Home Null controllability for a degenerate population model in divergence form via Carleman estimates
Article Open Access

Null controllability for a degenerate population model in divergence form via Carleman estimates

  • Genni Fragnelli EMAIL logo
Published/Copyright: October 15, 2019

Abstract

In this paper we consider a degenerate population equation in divergence form depending on time, on age and on space and we prove a related null controllability result via Carleman estimates.

MSC 2010: 35K65; 92D25; 93B05; 93B07

1 Introduction

We consider the following population model in divergence form describing the dynamics of a single species:

yt+ya(k(x)yx)x+μ(t,a,x)y=f(t,a,x)χωin Q,y(t,a,1)=y(t,a,0)=0on QT,A,y(0,a,x)=y0(a,x)in QA,1,y(t,0,x)=0Aβ(a,x)y(t,a,x)dain QT,1. (1.1)

Here Q := (0, T) × (0, A) × (0, 1), QT,A := (0, T) × (0, A), QA,1: = (0, A) × (0, 1) and QT,1 := (0, T) × (0, 1). Moreover, y(t, a, x) is the distribution of certain individuals at location x ∈ (0, 1), at time t ∈ (0, T), where T is fixed, and of age a ∈ (0, A). A is the maximal age of life, while β and μ are the natural fertility and the natural death rates, respectively. Thus, the formula 0A βy da denotes the distribution of newborn individuals at time t and location x. In the model χω is the characteristic function of the control region ω ⊂ (0, 1); the function k is the dispersion coefficient and we assume that it depends on the space variable x and degenerates at the boundary of the state space. We say that the function k is

Definition 1.1

Weakly degenerate (WD) if kW1,1([0, 1]),

k>0 in (0,1) and k(0)=k(1)=0,

and there exist M1, M2 ∈ (0, 1) such that xk′(x) ≤ M1k(x) and (x – 1)k′(x) ≤ M2k(x) for a.e. x ∈ [0, 1].

or

Definition 1.2

Strongly degenerate (SD) if kW1,∞([0, 1]),

k>0 in (0,1) and k(0)=k(1)=0,

and there exist M1, M2 ∈ [1, 2) such that xk′(x) ≤ M1k(x) and (x – 1)k′(x) ≤ M2k(x) for a.e. x ∈ [0, 1].

For example, as k one can consider k(x) = xα(1 – x)β, α, β > 0. Clearly, we say that k is weakly or strongly degenerate only at 0 if Definition 1.1 is satisfied only at 0, i.e. kW1,1([0, 1]), k > 0 in(0, 1], k(0) = 0 and, there exists M1 ∈ (0, 1) or M1 ∈ [1, 2) such that xk′(x) ≤ M1k(x) for a.e. x ∈ [0, 1]. Analogously at 1.

In the last centuries, population models have been widely investigated by many authors from many points of view (see, for example, [5], [9], [14], [20]). From the general theory for the Lotka-McKendrick system, it is known that the asymptotic behavior of the solution depends on the so called net reproduction rate R0: if R0 > 1, the solution is exponentially growing; if R0 < 1, the solution is exponentially decaying; if R0 = 1, the solution tends to the steady state solution. Clearly, if R0 > 1 and the system represents the distribution of a damaging insect population or of a pest population, it is very worrying. For example, in 2017 B. Zhong, C. Lv, W. Qin show that the net reproduction rate for the Tirathaba rufivena (which causes a lot of damages for the crop, for example, of fruits and flowers) depends on the temperature: it is 10.40 if the temperature is 28 C and it is 4.13 is the temperature is 20 C (see, for example, [25]); in 2011 S. S. Win, R. Muhamad, Z. A. M. Ahmad, N. A. Adam show that the net reproduction rate for the Nilaparvata lugens (which caused a lot of damages for the rice crop throughout South and South-East Asia since the early 1970’s) was about 10 (see, for example, [24]). For this reason, recently great attention is given to null controllability. For example in [21], where (1.1) models an insect growth, the control corresponds to a removal of individuals by using pesticides.

There are a lot of papers that deal with null controllability for (1.1) when the dispersion coefficient k is a constant or a strictly positive function (see, for example, [3]). If y is independent of a and k degenerates at the boundary or at an interior point of the domain we refer, for example, to [2], [15] and to [17], [18], [19] if μ is singular at the same point of k. To our best knowledge, [1] is the first paper where y depends on t, a and x and the dispersion coefficient k can degenerate. In particular, the authors assume that k degenerates at the boundary (for example k(x) = xα, being x ∈ (0, 1) and α > 0). Using Carleman estimates for the adjoint problem, the authors prove null controllability for (1.1) under the condition TA. However, this assumption is not realistic when A is too large. To overcome this problem in [10], the authors used Carleman estimates and a fixed point method via the Leray - Schauder Theorem. However, in [10] the authors consider a dispersion coefficient that can degenerate only at a point of the boundary and they use the fixed point technique in which the birth rate β must be in C2(Q) specially in the proof of [10, Proposition 4.2]. In the recent paper [13], we studied null controllability for (1.1) in non divergence form and with a diffusion coefficient degenerating at a one point of the boundary domain or in an interior point. Observe that, in the case of a boundary degeneracy, we cannot derive the null controllability for (1.1) by the one of the problem in non divergence form or vice versa, see [7]. For this reason here we study the null controllability for (1.1) assuming that k degenerates at the boundary of the domain and T < A completing [1]. We underline that here, contrary to [10] and [13], we assume also that k can degenerate at both points of the boundary domain (see Theorem 4.8) and β is only a continuous function. On the other hand, while in [10] the authors used Carleman estimates, a generalization of the Leray - Schauder fixed point Theorem and the multi-valued theory, here we use only Carleman estimates, some results of [13] and a technique based on cut off functions, making the proof slimmer and easier to read. Moreover, the technique that we use to prove Theorem 4.8 can be applied also to the problem in non divergence form considered in [13], generalizing [13, Theorem 4.8]. Finally, in the proof of the last theorem we make precise a calculation of [13, Theorem 4.8] which was not accurate. Observe that in this paper, as in [13], we do not consider the positivity of the solution, even if it is clearly interesting. This problem is related to the minimum time issue, i.e. given T cannot be arbitrarily small, but this study is still a work in progress, see [23] for related results in non degenerate cases.

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

2 Well posedness results

On the rates μ and β we assume:

Hypothesis 2.1

The functions μ and β are such that

βC(Q¯A,1) and β0 in QA,1,μC(Q¯) and μ0 in Q. (2.1)

To prove well posedness of (1.1), we introduce, as in [2], the following Hilbert spaces

Hk1:={uL2(0,1)u absolutely continuous in [0,1],kuxL2(0,1) and u(1)=u(0)=0}

and

Hk2:={uHk1(0,1)|kuxH1(0,1)}.

We have, as in [2] or [16], that the operator

A0u:=(kux)x,D(A0):=Hk2(0,1)

is self–adjoint, nonpositive and generates a strongly continuous semigroup on the space L2(0, 1).

Now, setting Aau:=ua, we have that

Au:=AauA0u,

for

uD(A)=uL2(0,A;D(A0)):uaL2(0,A;Hk1(0,1)),u(0,x)=0Aβ(a,x)u(a,x)da,

generates a strongly continuous semigroup on L2(QA,1) := L2(0, A; L2(0, 1)) (see also [4]). Moreover, the operator B(t) defined as

B(t)u:=μ(t,a,x)u,

for uD(𝓐), can be seen as a bounded perturbation of 𝓐 (see, for example, [2]); thus also (𝓐 + B(t), D(𝓐)) generates a strongly continuous semigroup.

Setting L2(Q) := L2(0, T; L2(QA,1)), the following well posedness result holds:

Theorem 2.1

Assume that k is weakly or strongly degenerate at 0 and/or at 1. For all fL2(Q) and y0L2(QA,1), the system (1.1) admits a unique solution

yU:=C([0,T];L2(QA,1)))L2(0,T;H1(0,A;Hk1(0,1)))

and

supt[0,T]y(t)L2(QA,1)2+0T0AkyxL2(0,1)2dadtCy0L2(QA,1)2+CfL2(Q)2, (2.2)

where C is a positive constant independent of k, y0 and f.

In addition, if f ≡ 0, then yC1([0, T]; L2(QA,1)).

For the existence of the solution and the regularity of it we refer, for example, to [11] and [22]. On the other hand, we postpone the proof of (2.2) to the Appendix.

3 Carleman estimates

From the general theory, it is known that null controllability for a linear parabolic system is, roughly speaking, equivalent to the observability for the associated homogeneous adjoint problem (see, for example, [12]). Thus, the key point is to prove such an inequality. A usual strategy in showing the observability inequality is to prove that certain global Carleman estimates hold true for the adjoint operator. Hence, this section is devoted to obtain global Carleman estimates for the operator which is the adjoint of the given one in both the weakly and the strongly degenerate cases. In particular, we consider the following adjoint system associated to (1.1):

zt+za+(k(x)zx)xμ(t,a,x)z=f,(t,a,x)Q,z(t,a,0)=z(t,a,1)=0,(t,a)QT,A,z(t,A,x)=0,(t,x)QT,1. (3.1)

3.0.0.1 Carleman inequalities when the degeneracy is at 0

In this subsection we will consider the case when k(0) = 0 and we assume that μ satisfies (2.1). On the other hand, on k we make additional assumptions:

Hypothesis 3.1

The function kC0[0, 1] ⋂ C1(0, 1] is such that k(0) = 0, k > 0 on (0, 1] and there exists M1 ∈ (0, 2) such that xk′(x) ≤ M1k(x) for all x ∈ [0, 1]. Moreover, if M1 ≥ 1 one has to require that there exists θ ∈ (0, M1], such that the function xk(x)xθ is nondecreasing near 0.

We remark that the assumption xa′ ≤ M1a, with M1 < 2, is essential in all our results and it is the same made, for example, in [2]. It implies that 1a L1(0, 1) and, in particular, if K < 1, then 1a L1(0, 1). Thus, the case M1 ≥ 2 is excluded. Summing up, we will confine our analysis to the case of 1a L1(0, 1). This is, however, the interesting case from the viewpoint of null controllability. In fact, if 1a L1(0, 1) and y is independent of a, then (1.1) fails to be null controllable on the whole interval [0, 1], and regional null controllability is the only property that can be expected, see [8].

Moreover, the assumption “∃θ such that the function xk(x)xθ is nondecreasing near 0” is just technical and it is equivalent to the following one: “∃θ such that θaxa′ near 0”. Clearly, the prototype is k(x) = xα, α ∈ (0, 2).

Now, let us introduce the weight function

φ(t,a,x):=Θ(t,a)(p(x)2pL(0,1)), (3.2)

where Θ is as in (4.7) and p(x):=0xyk(y)dy. Observe that φ(t, a, x) < 0 for all (t, x) ∈ Q and φ(t, a, x) → – ∞ as t → 0+, T or a → 0+. The following estimate holds:

Theorem 3.1

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

V:=L2(QT,A;Hk2(0,1))H1(0,T;H1(0,A;Hk1(0,1)))

satisfies, for all ss0,

QsΘkvx2+s3Θ3x2kv2e2sφdxdadtCQf2e2sφdxdadt+sC0T0AΘ(t,a)[kvx2e2sφ](t,a,1)dadt.

Clearly the previous Carleman estimate holds for every function v that satisfies (3.1) in (0, T) × (0, A) × (B, C) as long as (0, 1) is substituted by (B, C) and k satisfies Hypothesis 3.1 in (B, C).

3.0.0.2 Proof of Theorem 3.1

As a first step assume that μ ≡ 0.

In order to prove Theorem 3.1, we define, for s > 0, the function

w(t,a,x):=esφ(t,a,x)v(t,a,x)

where v is the solution of (3.1) in 𝓥; observe that, since v ∈ 𝓥, w ∈ 𝓥. Clearly, one has that w satisfies

(esφw)t+(esφw)a+(k(esφw)x)x=f(t,a,x),(t,x)Q,w(0,a,x)=w(T,a,x)=0,(a,x)QA,1,w(t,A,x)=w(t,0,x)=0,(t,x)QT,1,w(t,a,0)=w(t,a,1)=0,(t,a)QT,A. (3.3)

Defining Lw := wt + wa + (kwx)x and Lsw := eL(ew), the equation of (3.3) can be recast as follows

Lsw=Ls+w+Lsw=esφf,

where

Ls+w:=(kwx)xs(φt+φa)w+s2kφx2w,Lsw:=wt+wa2skφxwxs(kφx)xw.

As usual, we compute the inner product <Ls+w,Lsw>L2(Q) whose first expression is given in the following lemma

Lemma 3.1

Assume Hypothesis 3.1. The following identity holds

<Ls+w,Lsw>L2(Q)=s2Q(φtt+φaa)w2dxdadt+sQk(x)(k(x)φx)xxwwxdxdadt2s2Qkφxφtxw2dxdadt2s2Qkφxφxaw2dxdadt+sQ(2k2φxx+kkφx)wx2dxdadt+s3Q(2kφxx+kφx)kφx2w2dxdadt+sQφatw2dxdadt.{D.T.} (3.4)
{B.T.}QT,A[kwxwt]01dadt+QT,A[kwxwa]01dadts2QA,1φaw20Tdxda.+QT,A[sφx(k(x)wx)2+s2k(x)φtφxw2s3k2φx3w2]01dadt+QT,A[sk(x)(k(x)φx)xwwx]01dadt+s2QT,A[kφxφaw2]01dadt12QT,1[kwx2]0Adxdt+12QT,1[(s2kφx2s(φt+φa))w2]0Adxdt.

Proof

It results, integrating by parts,

<Ls+w,Lsw>L2(Q)=I1+I2+I3+I4,

where

I1=Q(kwx)x(wt2skφxwxs(kφx)xw)dxdadt,I2=Q(sφtw+s2kφx2w)(wt2skφxwxs(kφx)xw)dxdadt,I3=Q((kwx)xs(φt+φa)w+s2kφx2w)wadxdadt

and I4=sQφaw(wt2skφxwxs(kφx)xw)dxdadt.

By [2, Lemma 3.1], we get

I1+I2:=s2Qφttw2dxdadt+sQk(x)(k(x)φx)xxwwxdxdadt2s2Qk(x)φxφtxw2dxdadt+sQ(2k2φxx+k(x)kφx)wx2dxdadt+s3Q(2k(x)φxx+kφx)k(x)φx2w2dxdadt+QT,A[k(x)wxwt]x=0x=1dadt+QT,A[sφx(k(x)wx)2+s2k(x)φtφxw2s3k2φx3w2]x=0x=1dadt+QT,A[sk(x)(k(x)φx)xwwx]x=0x=1dadt. (3.5)

Next, we compute I3 and I4. Integrating by parts, we have

I3=120T01[kwx2]0Adxdt+0T0A[kwxwa]01dadt+120T01[(s2kφx2s(φt+φa))w2]0Adxdt+s2Qφaaw2dxdadt+s2Qφtaw2dxdadts2Qkφxφxaw2dxdadt. (3.6)

On the other hand

I4=s2Qφatw2dxdadts2Qkφxφaxw2dxdadts20A01φaw20Tdxda+s20T0A[kφxφaw2]01dadt. (3.7)

Adding (3.5) - (3.7), (3.4) follows immediately.□

As a consequence of the definition of φ, one has the next estimate:

Lemma 3.2

Assume Hypothesis 3.1. There exist two strictly positive constants C and s0 such that, for all ss0, all solutions w of (3.3) satisfy the following estimate

sCQΘkwx2dxdadt+s3CQΘ3x2kw2dxdadt{D.T.}.

Proof

The distributed terms of <Ls+w,Lsw>L2(Q) take the form

{D.T.}=s2Q(Θtt+Θaa)(p2pL(0,1))w2dxdadt2s2QΘΘtx2kw2dxdadt2s2QΘΘax2kw2dxdadt+sQΘ(2kkx)wx2dxdadt+s3QΘ3(2kkx)x2k2w2dxdadt+sQΘta(p2pL(0,1))w2dxdadt. (3.8)

Now, observe that there exists c > 0 such that

ΘμcΘν if 0<μ<ν|ΘΘt|cΘ3,|ΘΘa|cΘ3,|Θaa|cΘ32,|Θtt|cΘ32 and |Θta|cΘ32. (3.9)

Hence, proceeding as in the proof of [2, Lemma 3.5], one can deduce

(3.8)s2Q(Θtt+Θaa)(p2pL(0,1))w2dxdadts3C4QΘ3x2kw2dxdadts3C4QΘ3x2kw2dxdadt+CsQΘkwx2dxdadt+s3CQΘ3x2kw2dxdadt+sQΘta(p2pL(0,1))w2dxdadt. (3.10)

Now, it results

sQΘtt(p2pL(0,1))C2sQΘ3/2b(x)w2dxdadt+sC2QΘ3/2w2dxdadt,

where b(x)=0xyk(y)dy. As in [2, Lemma 3.5], one can estimate the last two terms in the following way

s2QΘ3/2b(x)w2dxdadtC16s3QΘ3x2k(x)w2dxdadt,

for s large enough and

s2QΘ3/2w2dxdadtC4sQΘk(x)wx2dxdadt+C16s3QΘ3x2k(x)w2dxdadt.

Hence

sQΘttψ(x)w2dxdadtC4s01Θk(x)wx2dxdadt+C8s3QΘ3x2k(x)w2dxdadt.

The same estimate holds also for

QΘaa(p2pL(0,1))dxdadt and QΘta(p2pL(0,1))dxdadt.

Using the above estimates in (3.10) the thesis follows immediately for s0 large enough.□

The next lemma holds.

Lemma 3.3

Assume Hypothesis 3.1. The boundary terms in (3.4) become

{B.T.}=QT,A[sΘxkwx2]01dadt. (3.11)

Proof

Using the definition of φ, [2, Lemma 3.6], the boundary conditions of w and proceeding as in [13, Lemma 3.2], the boundary terms of <Ls+w,Lsw>L2(Q) become

{B.T.}=QT,A[kwxwa]01dadts2QA,1φaw20TdxdaQT,A[sΘxkwx2]01dadt+s2QT,A[kφxφaw2]01dadt12QT,1[kwx2]0Adxdt+12QT,1[(s2kφx2s(φt+φa))w2]0Adxdt=QT,A[sΘxkwx2]01dadt.

As a consequence of Lemmas 3.3 and 3.2, we have

Proposition 3.1

Assume Hypothesis 3.1. There exist two strictly positive constants C and s0 such that, for all ss0, all solutions w of (3.3) in 𝓥 satisfy

sCQΘkwx2dxdadt+s3CQΘ3x2kw2dxdadtCQf2e2sφdxdadt+QT,A[sΘkwx2](t,a,1)dadt.

Recalling the definition of w, we have v = ew and vx = (wxxw)e. Thus, Theorem 3.1 follows immediately by Proposition 3.1 when μ ≡ 0.

Now, we assume that μ ≢ 0.

To complete the proof of Theorem 3.1 we consider the function f = f + μv. Hence, there are two strictly positive constants C and s0 such that, for all ss0, the following inequality holds

QsΘkvx2+s3Θ3x2kv2e2sφdxdadtCQf¯2e2sφdxdadt+sC0T0AΘ(t,a)[kvx2e2sφ](t,a,1)dadt. (3.12)

On the other hand, we have

Qf¯2e2sφdxdadt2(Q|f|2e2sφdxdadt+Q|μ|2|v|2e2sφdxdadt). (3.13)

Now, if M1 < 1, applying the Hardy-Poincaré proved in [2, Proposition 2.1] to the function ν := ev, we obtain

Q|μ|2|v|2e2sφdxdadtμ2Qν2dxdadtCQν2k(x)x2dxdadtCQk(x)νx2dxdadtCQk(x)e2sφvx2dxdadt+Cs2QΘ2e2sφx2kv2dxdadt. (3.14)

If M1 ≥ 1, using the Young’s inequality to the function ν := ev, we have

Q|μ|2|v|2e2sφdxdadtμ2Qν2dxdadtCQk1/3x2/3ν23/4x2kν21/4dxdadtCQk1/3x2/3ν2dxdadt+CQx2kν2dxdadt. (3.15)

Now, consider the function y(x) = (k(x)x4)1/3. Clearly, y(x)=k(x)x2k(x)2/3Ck(x)andk1/3x2/3=y(x)x2. Moreover, using Hypothesis 3.1, one has that the function y(x)xq=k(x)xθ13, where q:=4+ϑ3 ∈ (1, 2), is nondecreasing near 0. The Hardy-Poincaré inequality (see [2, Proposition 2.1.]) implies

Qk1/3x2/3ν2dxdadt=Qyx2ν2dxdadtCQy(νx)2dxdadtCQk(x)νx2dxdadtCQk(x)e2sφvx2dxdadt+Cs2QΘ2e2sφx2kv2dxdadt. (3.16)

In any case, by (3.14), (3.15) and (3.16),

Q|μ|2|v|2e2sφdxdadtCQk(x)e2sφvx2dxdadt+Cs2QΘ2e2sφx2kv2dxdadt.

Using this last inequality in (3.13), it follows

Q|f¯|2~e2sφdxdadt2Q|f|2~e2sφdxdadt+CQk(x)e2sφvx2dxdadt+Cs2QΘ2e2sφx2kv2dxdadtCQ|f|2~e2sφdxdadt+CQΘk(x)e2sφvx2dxdadt+Cs2QΘ3e2sφx2kv2dxdadt. (3.17)

Substituting in (3.12), one can conclude

QsΘkvx2+s3Θ3x2kv2e2sφdxdadtC(Q|f|2e2sφkdxdadt+s0T0AΘ(t,a)[kvx2e2sφ](t,a,1)dadt),

for all s large enough.

3.0.0.3 Carleman inequalities when the degeneracy is at 1

In this subsection we will consider the case when k(1) = 0. Again μ satisfies (2.1) and on k we make the following assumption:

Hypothesis 3.2

The function kC0[0, 1] ⋂ C1[0, 1) is such that k(1) = 0, k > 0 on [0, 1) and there exists M2 ∈ (0, 2) such that (x – 1) k′(x) ≤ M2k(x) for all x ∈ [0, 1]. Moreover, if M2 ≤ 1 one has to require that there exists θ ∈ (0, M2], such that the function xk(x)|1x|θ is nonincreasing near 1.

For Hypothesis 3.2 we can make the same considerations made for Hypothesis 3.1.

As in the previous subsection, let us introduce the weight function

φ¯(t,a,x):=Θ(t,a)(p¯(x)2p¯L(0,1)), (3.18)

where Θ is as in (4.7) and p¯(x):=0xy1k(y)dy. As before, φ̄(t, a, x) < 0 for all (t, x) ∈ Q and φ̄(t, a, x) → –∞ as t → 0+, T or a → 0+. The following estimate holds:

Theorem 3.2

Assume that Hypothesis 3.2 is satisfied. Then, there exist two strictly positive constants C and s0 such that every solution v of (3.1) in 𝓥 satisfies, for all ss0,

QskΘvx2+s3Θ3(x1)2kv2e2sφ¯dxdadtCQf2e2sφ¯dxdadt+sC0T0AΘ(t,a)[(1x)vx2e2sφ¯](t,a,0)dadt.

The previous Carleman estimate holds for every function v that satisfies (3.1) in (0, T) × (0, A) × (B, C) as long as (0, 1) is substituted by (B, C) and k satisfies Hypothesis 3.2 in (B, C).

The proof of Theorem 3.2 is analogous to one of Theorem 3.1 so we omit it. However, we underline that in the proof of Theorem 3.1 we use [2, Lemma 3.1] which is proved only if k degenerates at 0; actually we observe that the proof of [2, Lemma 3.1] does not depend on the degeneracy point; hence, it holds also if k(1) = 0. Instead, Lemma 3.3, if k(1) = 0, becomes

QT,A[sΘ(x1)kwx2]01dadt=QT,A[sΘkwx2](t,a,0)dadt;

thus, Proposition 3.1 can be rewritten in the following way

sCQΘkwx2dxdadt+s3CQΘ3(x1)2kw2dxdadtCQf2e2sφ¯dxdadt+QT,A[sΘkwx2](t,a,0)dadt.

If μ ≢ 0, we can proceed as in the proof of Theorem 3.1. However, while in that case we use the Hardy-Poincaré inequality proved in [2, Proposition 2.1] which holds only if k(0) = 0, in this case we have to use the following inequality whose proof we postpone to the Appendix.

Proposition 3.2

(Hardy-Poincaré inequalities). Assume that k : [0, 1] ⟶ ℝ+ is in 𝓒([0, 1]), k(1) = 0, k > 0 on [0, 1)

Case (i):

Hypothesis (HP1): Assume that k is such that there exists θ ∈ (0, 1) such that the function

xk(x)(1x)θisnondecreasinginneighbourhoodofx=1.

Then, there is a constant C > 0 such that for any function w, locally absolutely continuous on [0, 1), continuous at 1 and satisfying

w(1)=0,and01k(x)|w(x)|2dx<+,

the following inequality holds

01k(x)(1x)2w2(x)dxC01k(x)|w(x)|2dx. (3.19)

If hypothesis (HP1) is replaced by

Hypothesis (HP1)’: Assume that k is such that there exists θ ∈ (0, 1) such that the function

xk(x)(1x)θisnondecreasingin[0,1),

Then, for any function w, locally absolutely continuous on [0, 1), continuous at 1 and satisfying

w(1)=0,and01k(x)|w(x)|2dx<+,

the inequality (3.19) holds with the explicit constant C=4(1θ)2.

Case (ii):

Hypothesis (HP2): Assume that k is such that there exists θ ∈ (1, 2) such that the function

xk(x)(1x)θisnonincreasinginaneighbourhoodofx=1,

Then, there is a constant C > 0 such that for any function w, locally absolutely continuous on [0, 1) satisfying

w(0)=0,and01k(x)|w(x)|2dx<+,

the inequality (3.19) holds.

If hypothesis (HP2) is replaced by

Hypothesis (HP2)’: Assume that k is such that there exists θ ∈ (1, 2) such that the function

xk(x)(1x)θisnonincreasingin[0,1).

Then, for any function w, locally absolutely continuous on [0, 1) satisfying

w(0)=0,and01k(x)|w(x)|2dx<+,

the inequality (3.19) holds with the explicit constant C=4(1θ)2.

4 Observability and controllability

In this section we will prove, as a consequence of the Carleman estimates established in Section 3, observability inequalities for the associated adjoint problem of (1.1). From now on, we assume that the control set ω is such that

ω=(α,ρ)(0,1). (4.1)

Moreover, on k and β we assume the following assumptions:

Hypothesis 4.1

The function kC0[0, 1]⋂ C1(0, 1) is such that k(0) = 0 = k(1), k > 0 on (0, 1) and there exist M1, M2 ∈ (0, 2) such that xk′(x) ≤ M1k(x) and (x – 1)k′(x) ≤ M2k(x) for all x ∈ [0, 1]. Moreover, one has to require that:

  1. if M1 ≥ 1, there exists θ ∈ (0, M1], such that the function xk(x)xθ is nondecreasing near 0;

  2. if M2 ≤ 1, there exists y ∈ (0, M2], such that the function xk(x)|1x|y is nonincreasing near 1.

Hypothesis 4.2

Assume T < A and suppose that there exists āT such that

β(a,x)=0for all(a,x)[0,a¯]×[0,1]. (4.2)

Observe that Hypothesis 4.2 is the biological meaningful one. Indeed, ā is the minimal age in which the female of the population become fertile, thus it is natural that before ā there are no newborns. For other comments on Hypothesis 4.2 we refer to [12].

Under the previous hypotheses, the following observability inequality holds:

Proposition 4.1

Suppose that Hypotheses 3.1 or 3.2 or 4.1 and 4.2 hold. Then, for every δ ∈ (T, A), there exists a strictly positive constant C = C(δ) such that every solution v ∈ 𝓤 of

vt+va+(k(x)vx)xμ(t,a,x)v+β(a,x)v(t,0,x)=0,(t,x,a)Q,v(t,a,0)=v(t,a,1)=0,(t,a)QT,A,v(T,a,x)=vT(a,x)L2(QA,1),(a,x)QA,1v(t,A,x)=0,(t,x)QT,1, (4.3)

satisfies

0A01v2(Ta¯,a,x)dxdaC0δ01vT2(a,x)dxda+0T0Aωv2dxdadt. (4.4)

Here vT(a, x) is such that vT(A, x) = 0 in (0, 1).

Remark 1

  1. If T = ā, the observability inequality given in the previous proposition is the corresponding of [1, Proposition 3.1], where the authors proved it under different assumptions and with TA.

  2. Moreover, as in [13], observe that in (4.4) the presence of the integral 0δ01vT2(a,x)dxda is related to the presence of the term β(a, x) v(t, 0, x) in the equation of (4.3). In fact, estimating such a term using the method of characteristic lines, we obtain the previous integral. Obviously, if vT(a, x) = 0 a.e. in (0, δ) × (0, 1), we obtain the classical observability inequality.

Before proving Proposition 4.1 we will give some results that will be very helpful. As a first step we introduce the following class of functions

W:={vsolution of (4.3)|vTD(A2)},

where D(𝓐2) = {uD(𝓐) | 𝓐uD(𝓐)} is densely defined in D(𝓐) (see, for example, [6, Lemma 7.2]) and hence in L2(QA,1). Obviously,

W=C1([0,T];D(A))V:=L2(QT,A;Hk2(0,1))H1(0,T;H1(0,A;Hk1(0,1)))U.

Proposition 4.2

(Caccioppoli’s inequality). Let ωand ω two open subintervals of (0, 1) such that ω′ ⊂⊂ ω ⊂⊂ (0, 1). Let ψ(t, a, x) := Θ(t, a)Ψ(x), where Θ is defined in (4.7) and ΨC1(0, 1) is a strictly negative function. Then, there exist two strictly positive constants C and s0 such that, for all ss0,

0T0Aωvx2e2sψdxdadtC0T0Aωv2dxdadt+Qf2e2sψdxdadt, (4.5)

for every solution v of (3.1).

The proof of the previous proposition is similar to the one given in [13], but we repeat it in the Appendix for the reader’s convenience.

Moreover, the following non degenerate inequality proved in [13] is crucial:

Theorem 4.1

[see [13, Theorem 3.2]] Let z ∈ 𝓩 be the solution of (3.1), where fL2(Q), kC1([0, 1]) is a strictly positive function and

Z:=L2(QT,A;H2(0,1)H01(0,1))H1(0,T;H1(0,A;H01(0,1))).

Then, there exist two strictly positive constants C and s0, such that, for any ss0, z satisfies the estimate

Q(s3ϕ3z2+sϕzx2)e2sΦdxdadtC(Qf2e2sΦdxdadtsκ0T0Ake2sΦϕ(zx)2x=0x=1dadt), (4.6)

where the functions ϕ and Φ are defined as follows

ϕ(t,a,x)=Θ(t,a)eκσ(x),Θ(t,a)=1t4(Tt)4a4,Φ(a,t,x)=Θ(t,a)Ψ(x),Ψ(x)=eκσ(x)e2κσ, (4.7)

(t, a, x) ∈ Q, κ > 0 and σ(x):=dx11k(t)dt, where 𝔡 = ∥k′∥L(0,1).

Remark 2

The previous Theorem still holds under the weaker assumption kW1,∞(0, 1) without any additional assumption.

On the other hand, if we require kW1,1(0, 1) then we have to add the following hypothesis: there exist two functions 𝔤 ∈ L1(0, 1), 𝔥 ∈ W1,∞(0, 1) and two strictly positive constants 𝔤0, 𝔥0 such that 𝔤(x) ≥ 𝔤0 and

k(x)2k(x)x1g(t)dt+h0+k(x)g(x)=h(x)fora.e.x[0,1].

In this case, i.e. if kW1,1(0, 1), the function Ψ in (4.7) becomes

Ψ(x):=r0x1k(t)t1g(s)dsdt+0xh0k(t)dtc, (4.8)

where r and 𝔠 are suitable strictly positive functions. For other comments on Theorem 4.1 we refer to [13].

With the aid of Theorems 3.1, 3.2, 4.1 and Proposition 4.2, we can now show ω-local Carleman estimates for (3.1).

Theorem 4.2

Assume Hypothesis 3.1. Then, there exist two strictly positive constants C and s0 such that every solution v of (3.1) in 𝓥 satisfies, for all ss0,

QsΘkvx2+s3Θ3x2kv2e2sφdxdadtCQf2dxdadt+0T0Aωv2dxdadt.

Proof

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

0ξ(x)1, for all x[0,1],ξ(x)=1,x[0,(2α+ρ)/3],ξ(x)=0,x[(α+2ρ)/3,1].

We define w(t, a, x) := ξ(x)v(t, a, x) where v ∈ 𝓥 satisfies (3.1). Then w satisfies

wt+wa+(kwx)xμw=ξf+(kξxv)x+ξxkvx=:h,(t,a,x)Q,w(t,a,0)=w(t,a,1)=0,(t,a)QT,A.

Thus, applying Theorem 3.1 and Proposition 4.2,

0T0A02α+ρ3sΘkvx2+s3Θ3x2kv2e2sφdxdadt=0T0A02α+ρ3sΘkwx2+s3Θ3x2kw2e2sφdxdadtQsΘkwx2+s3Θ3x2kw2e2sφdxdadtCQh2e2sφdxdadtCQf2e2sφdxdadt+0T0Aωv2dxdadt+0T0Aωvx2e2sφdxdadtCQf2e2sφdxdadt+0T0Aωv2dxdadt, (4.9)

where ω:=2α+ρ3,α+2ρ3.

Now, consider z = ηv, where η = 1 – ξ and take ∈ (0, α). Then z satisfies

zt+za+(kzx)xμz=ηf+(kηxv)x+ηxkvx=:h,(t,a,x)QT,A×(α¯,1)=:Q¯,z(t,a,α¯)=z(t,a,1)=0,(t,a)QT,A. (4.10)

Clearly the equation satisfied by z is not degenerate, thus applying Theorem 4.1 and Proposition 4.2, one has

Q¯(s3ϕ3z2+sϕzx2)e2sΦdxdadtCQ¯h2e2sΦdxdadtCQ¯f2e2sΦdxdadt+0T0Aω(v2+vx2)e2sΦdxdadtCQf2e2sΦdxdadt+0T0Aωv2dxdadt.

Hence

0T0Aα+2ρ31(s3ϕ3v2+sϕvx2)e2sΦdxdadt=0T0Aα+2ρ31(s3ϕ3z2+sϕzx2)e2sΦdxdadtCQf2e2sΦdxdadt+0T0Aωv2dxdadt,

for a strictly positive constant C. Proceeding, for example, as in [16] one can prove the existence of ς > 0, such that, for all (t, a, x) ∈ [0, T] × [0, A] × [, 1], we have

e2sφςe2sΦ,x2k(x)e2sφςe2sΦ. (4.11)

Thus, for a strictly positive constant C,

0T0Aα+2ρ31sΘkvx2+s3Θ3x2kv2e2sφdxdadtC0T0Aα+2ρ31(s3ϕ3v2+sϕvx2)e2sΦdxdadtCQf2e2sΦdxdadt+0T0Aωv2dxdadt. (4.12)

Now, consider α̃ ∈ (α, (2α + ρ)/3), ρ̃ ∈ ((α + 2ρ)/3, ρ) and a smooth function τ : [0, 1] → ℝ such that

0τ(x)1, for all x[0,1],τ(x)=1,x[(2α+ρ)/3,(α+2ρ)/3],τ(x)=0,x[0,α~][ρ~,1],

and define ζ(t, a, x) := τ(x)v(t, a, x). Clearly, ζ satisfies (4.10) with h := τf + (xv)x + τxkvx. Observe that in this case τx, τxx ≢ 0 in ω¯:=α~,2α+ρ3α+2ρ3,ρ~. As before, by Theorem 4.1, Proposition 4.2 and (4.11), we have

0T0A2α+ρ3α+2ρ3sΘkvx2+s3Θ3x2kv2e2sφdxdadtC0T0A2α+ρ3α+2ρ3(s3ϕ3v2+sϕvx2)e2sΦdxdadt=C0T0A2α+ρ3α+2ρ3(s3ϕ3ζ2+sϕζx2)e2sΦdxdadtCQf2e2sΦdxdadt+0T0Aωv2dxdadt. (4.13)

Adding (4.9), (4.12) and (4.13), the thesis follows.□

Proceeding as before one can prove

Theorem 4.3

Assume Hypothesis 3.2. Then, there exist two strictly positive constants C and s0 such that every solution v of (3.1) in 𝓥 satisfies, for all ss0,

QsΘkvx2+s3Θ3(1x)2kv2e2sφdxdadtCQf2dxdadt+0T0Aωv2dxdadt.

Remark 3

Observe that the results of Theorems 4.2 and 4.3 still hold true if we substitute the domain (0, T) × (0, A) with a general domain (T1, T2) × (y, A), provided that μ and β satisfy the required assumptions. In this case, in place of the function Θ defined in (4.7), we have to consider the weight function

Θ~(t,a):=1(tT1)4(T2t)4(ay)4. (4.14)

Using the previous local Carleman estimates one can prove the next observability inequalities.

Theorem 4.4

Assume Hypotheses 3.1 or 3.2 and 4.2 with T > ā. Then, for every δ ∈ (0, A), there exists a strictly positive constant C = C(δ) such that every solution v of (4.3) in 𝓥 satisfies

0A01v2(Ta¯,a,x)dxdaC0T0δ01v2(t,a,x)dxdadt+C0a¯01vT2(a,x)dxda+0T0Aωv2dxdadt.

Moreover, if vT(a, x) = 0 for all (a, x) ∈ (0, ā) × (0, 1), one has

0A01v2(Ta¯,a,x)dxdaC0T0δ01v2(t,a,x)dxdadt+0T0Aωv2dxdadt.

Proof

As in [13] and using the method of characteristic lines, one can prove the following implicit formula for v solution of (4.3):

S(Tt)vT(T+at,), (4.15)

if t + a and

v(t,a,)=S(Tt)vT(T+at,)+aT+atS(sa)β(s,)v(s+ta,0,)ds,Γ=a¯aAS(sa)β(s,)v(s+ta,0,)ds,Γ=ΓA,T, (4.16)

otherwise. Here (S(t))t≥0 is the semigroup generated by the operator 𝓐0μId for all uD(𝓐0) (Id is the identity operator), ΓA,T := Aa + t and

Γ:=min{a¯,ΓA,T}. (4.17)

In particular, it results

v(t,0,):=S(Tt)vT(Tt,), (4.18)

if tTā.

Proceeding as in [13, Theorem 4.4], with suitable changes, one has that there exists a positive constant C such that:

QA,1v2(T~,a,x)dxdaCTa¯2Ta¯4QA,1v2(t,a,x)dxdadt. (4.19)

However, here we make all the calculations in order to make precise some steps in [13, Theorem 4.4], where there is a misprint. Indeed in (4.19) the integration is made on Ta¯2,Ta¯4 in place of T4,3T4 as in [13]. The integration on T4,3T4 is correct if T = ā as in the next Corollary or in [13, Corollary 4.1].

Indeed, define, for ς > 0, the function w = eςtv, where v solves (4.3). Then w satisfies

wt+wa+(k(x)wx)x(μ(t,a,x)+ς)w=β(a,x)w(t,0,x),(t,x,a)Q~,w(t,a,0)=w(t,a,1)=0,(t,a)Q~T,A,w(T,a,x)=eςTvT(a,x),(a,x)QA,1,w(t,A,x)=0,(t,x)Q~T,1, (4.20)

where := (, T) × QA,1, T,A := (, T) × (0, A) and T,1 := (, T) × (0, 1). Multiplying the equation of (4.20) by –w and integrating by parts on Qt := (, t) × (0, A) × (0, 1), it results

12QA,1w2(t,a,x)dxda+eςT~2QA,1v2(T~,a,x)dxda+12T~t01w2(τ,0,x)dxdτ+ςQtw2(τ,a,x)dxdadτQtβw(τ,0,x)wdxdadτβL(Q)1ϵQtw2dxdadτ+ϵAβL(Q)T~t01w2(τ,0,x)dxdτ, (4.21)

for ϵ > 0. Choosing ϵ=12βL(Q)Aandς=βL(Q)ϵ, we have

QA,1v2(T~,a,x)dxdaCQA,1w2(t,a,x)dxdaCQA,1v2(t,a,x)dxda.

Then, integrating over Ta¯2,Ta¯4, we have (4.19).

Now, take δ ∈ (0, A). By (4.19), we have

QA,1v2(T~,a,x)dxdaCTa¯2Ta¯40δ+δA01v2(t,a,x)dxdadt. (4.22)

Consider the term Ta¯2Ta¯4δA01v2(t,a,x)dxdadt. If Hypothesis 3.1 holds, proceeding as in [2, Lemma 3.2], one has

01v2dxC01kvx2dx+01x2kv2dx, (4.23)

for a strictly positive constant C. Hence,

Ta¯2Ta¯4δA01v2(t,a,x)dxdadtCTa¯2Ta¯4δA01Θ~kvx2e2sφ~dxdadt+CTa¯2Ta¯4δA01Θ~3x2kv2e2sφ~dxdadt,

where Θ̃ is defined in (4.14) with T1 := Tā, T2 := T, y = 0 and φ̃ is the function associated to Θ̃ according to (3.2). Analogously, if Hypothesis 3.2 holds, then we obtain

Ta¯2Ta¯4δA01v2(t,a,x)dxdadtCTa¯2Ta¯4δA01Θ~kvx2e2sφ¯~dxdadt+CTa¯2Ta¯4δA01Θ~3(1x)2kv2e2sφ¯~dxdadt,

where φ̄̃ is the function associated to Θ̃ according to (3.18). Thus, by Theorem 4.2 or 4.3 applied to := (Tā, T) × (0, A) × (0, 1) and Remark 3,

Ta¯2Ta¯4δA01v2(t,a,x)dxdadtCQ¯f2dxdadt+0T0Aωv2dxdadt,

where, in this case, f(t, a, x) := –β(a, x)v(t, 0, x). Thus

Ta¯2Ta¯4δA01v2(t,a,x)dxdadtCβL(Q)2Q¯v2(t,0,x)dxdadt+0T0Aωv2dxdadt, (4.24)

for a strictly positive constant C. By (4.18), (4.24) and proceeding as in [13], we have

Ta¯2Ta¯4δA01v2(t,a,x)dxdadtCβL(Q)2Qa¯,1vT2(a,x)dxda+0T0Aωv2dxdadt, (4.25)

for a strictly positive constant C. By (4.22) and (4.25), it results

QA,1v2(T~,a,x)dxdaC0T0δ01v2(t,a,x)dxdadt+CQa¯,1vT2(a,x)dxda+0T0Aωv2dxdadt. (2.26)

Corollary 4.1

Assume ā = T, Hypotheses 3.1 or 3.2 and 4.2. Then, for every δ ∈ (0, A), there exists a strictly positive constant C = C(δ) such that every solution v of (4.3) in 𝓥 satisfies

0A01v2(0,a,x)dxdaC0T0δ01v2(t,a,x)dxdadt+C0a¯01vT2(a,x)dxda+0T0Aωv2dxdadt.

Moreover, if vT(a, x) = 0 for all (a, x) ∈ (0, ā) × (0, 1), one has

0A01v2(0,a,x)dxdaC0T0δ01v2(t,a,x)dxdadt+0T0Aωv2dxdadt.

Actually, proceeding as in [13] with suitable changes, we can improve the previous results in the following way:

Theorem 4.5

Assume Hypotheses 3.1 or 3.2 and 4.2. Then, for every δ ∈ (T, A), there exists a strictly positive constant C = C(δ) such that every solution v of (4.3) in 𝓥 satisfies

0A01v2(Ta¯,a,x)dxdaC0δ01vT2(a,x)dxda+0T0Aωv2dxdadt.

By Theorem 4.5 and using a density argument, one can deduce Proposition 4.1. As a consequence one can prove, as in [13], the following null controllability results:

Theorem 4.6

Assume Hypotheses 3.1 or 3.2 or 4.1 and 4.2. Then, given T > 0 and y0L2(QA,1), for every δ ∈ (T, A), there exists a control fδL2() such that the solution yδ ∈ 𝓤 of

yt+ya(k(x)yx)x+μ(t,a,x)y=fδ(t,x,a)χωinQ~,y(t,a,1)=y(t,a,0)=0onQ~T,A,y(T~,a,x)=y0(a,x)inQA,1,y(t,0,x)=0Aβ(a,x)y(t,a,x)dainQ~T,1, (4.27)

satisfies

yδ(T,a,x)=0a.e.(a,x)(δ,A)×(0,1).

Moreover, there exists C = C(δ) > 0 such that

fδL2(Q~)Cy0L2(QA,1). (4.28)

Here, we recall, = (, T) × (0, A) × (0, 1), T,A = (, T) × (0, A) and T,1 = (, T) × (0, 1).

Observe that if T = ā, Theorem 4.6 is exactly the null controllability result that we expect. Indeed, in this case (4.27) coincide with (1.1). On the other hand, if T > ā, the null controllability for (1.1) is given in the next theorem and it is based on the previous result:

Theorem 4.7

Assume Hypotheses 3.1 or 3.2 and 4.2. Then, given T ∈ (0, A) and y0L2(QA,1), for every δ ∈ (T, A), there exists a control fδL2(Q) such that the solution yδ of (1.1) satisfies

yδ(T,a,x)=0a.e.(a,x)(δ,A)×(0,1).

Moreover, there exists C = C(δ) > 0 such that

fδL2(Q)Cy0L2(QA,1). (4.29)

Proof

The proof is similar to the one of [13, Theorem 4.8]. However, here we make all the calculations in order to make precise some steps in [13, Theorem 4.8], where there is a misprint and a term was missing.

As a first step, set := Tā ∈ (0, T). By Theorem 2.1, there exists a unique solution u of

ut+ua(k(x)ux)x+μ(t,a,x)u=0in (0,T~)×(0,A)×(0,1),u(t,a,1)=u(t,a,0)=0on (0,T~)×(0,A),u(0,a,x)=y0(a,x)in (0,A)×(0,1),u(t,0,x)=0Aβ(a,x)u(t,a,x)dain (0,T~)×(0,1). (4.30)

Set 0(a, x) := u(, a, x); clearly 0L2(QA,1). Now, consider

wt+wa(k(x)wx)x+μ(t,a,x)w=h(t,x,a)χωin Q~,w(t,a,1)=w(t,a,0)=0on Q~T,A,w(T~,a,x)=y~0(a,x)in QA,1,w(t,0,x)=0Aβ(a,x)w(t,a,x)dain Q~T,1. (4.31)

Again, by Theorem 2.1, there exists a unique solution wδ of (4.31) and, by the previous Theorem, there exists a control hδL2() such that

wδ(T,a,x)=0 a.e. (a,x)(δ,A)×(0,1)

and

hδL2(Q~)Cy~0L2(QA,1),

for a positive constant C.

Now, define yδ and fδ by

yδ:=u,in[0,T~],wδ,in[T~,T]andfδ:=0,in[0,T~],hδ,in[T~,T].

Then yδ satisfies (1.1) and fδL2(Q) is such that

yδ(T,a,x)=0 a.e. (a,x)(δ,A)×(0,1).

Indeed yδ(T, a, x) = wδ(T, a, x) = 0 a.e. (a, x) ∈ (δ, A) × (0, 1).

Now, we prove (4.29). As a first step, as in [13], we multiply the equation of (4.30) by u. Then, integrating over QA,1, we obtain:

12ddt0A01u2dxda+1201u2(t,A,x)dx+0A01kux2dxda+0A01μu2dxda=1201u2(t,0,x)dx.

Hence, using the fact that u(t, 0, x) = 0A β(a, x) u(t, a, x)da, we have

12ddt0A01u2dxda12010Aβ(a,x)u(t,a,x)da2dxC20A01u2dxda.

Setting F(t):=u(t)L2(QA,1)2 and multiplying the previous inequality by eCt, it results

ddteCtF(t)0.

Integrating over (0, t), for all t ∈ [0, T], we obtain

0A01u2(t,a,x)dxdaeCT0A01u2(0,a,x)dxda=eCT0A01y02(a,x)dxda.

In particular,

0A01u2(T~,a,x)dxdaeCT0A01y02(a,x)dxda.

Thus,

fδL2(Q)2=T~T0A01hδ2dxdadtCy~0L2(QA,1)2=C0A01u2(T~,a,x)dxdaC0A01y02(a,x)dxda, (4.32)

for a strictly positive constant C. Hence, (4.29) follows.□

As a consequence of the previous theorem, we obtain the null controllability property if the coefficient k degenerates at 0 and at 1 at the same time.

Theorem 4.8

Assume Hypotheses 4.1 and 4.2. Then, given T ∈ (0, A) and y0L2(QA,1), for every δ ∈ (T, A), there exists a control fδL2(Q) such that the solution yδ of (1.1) satisfies

yδ(T,a,x)=0a.e.(a,x)(δ,A)×(0,1).

Moreover, there exists C = C(δ) > 0 such that

fδL2(Q)Cy0L2(QA,1). (4.33)

Proof

Fix y0L2(QA,1) and consider the two problems

(P1)yt+ya(k(x)yx)x+μ(t,a,x)y=f(t,a,x)χωin (0,T)×(0,A)×(0,β¯),y(t,a,β¯)=y(t,a,0)=0on QT,A,y(0,a,x)=y0(a,x)in (0,A)×(0,β¯),y(t,0,x)=0Aβ(a,x)y(t,a,x)dain (0,T)×(0,β¯), (4.34)

and

(P2)yt+ya(k(x)yx)x+μ(t,a,x)y=f(t,a,x)χωin (0,T)×(0,A)×(α¯,1),y(t,a,1)=y(t,a,α¯)=0on QT,A,y(0,a,x)=y0(a,x)in (0,A)×(α¯,1),y(t,0,x)=0Aβ(a,x)y(t,a,x)dain (0,T)×(α¯,1), (4.35)

where ∈ (0, α) and β̄ ∈ (β, 1). Thus, by Theorem 4.7, there exist two controls h1,δ and h2,δ such that the solutions u1,δ and u2,δ of (P1) and (P2), associated to h1,δ and h2,δ, respectively, satisfy

u1,δ(T,a,x)=0a.e. (a,x)(δ,A)×(0,β¯),

and

u2,δ(T,a,x)=0a.e. (a,x)(δ,A)×(α¯,1).

Moreover, there exists C > 0 such that

0T0A0β¯h1,δ2dxdadtCy0L2(QA,1)2

and

0T0Aα¯1h2,δ2dxdadtCy0L2(QA,1)2.

Denote with u1 and h1 (respectively u2 and h2) the trivial extensions of u1,δ and h1,δ (respectively u2,δ and h2,δ) to [β̄, 1] (respectively [0, ]), so that all functions are defined in the interval [0, 1]. Clearly, they depends always on δ and

hiL2(Q)Cy0L2(QA,1),i=1,2. (4.36)

Now, let u3 be the solution of

yt+ya(k(x)yx)x+μ(t,a,x)y=0in (0,T)×(0,A)×(0,1),y(t,a,1)=y(t,a,0)=0on QT,A,y(0,a,x)=y0(a,x)in (0,A)×(0,1),y(t,0,x)=0Aβ(a,x)y(t,a,x)dain (0,T)×(0,1), (4.37)

and consider the three smooth cut off functions ξ, η, ϕ : [0, 1] → ℝ defined as

0ξ(x)1, for all x[0,1],ξ(x)=1,x[0,(2α+ρ)/3],ξ(x)=0,x[(α+2ρ)/3,1],0η(x)1, for all x[0,1],η(x)=0,x[0,(2α+ρ)/3],η(x)=1,x[(α+2ρ)/3,1]

and ϕ := 1 – ξη. Finally, take

y(t,a,x)=ξu1+ηu2+F(t)ϕu3,

where F(t):=TtT.

It is easy to verify that y(t, a, 0) = y(t, a, 1) = 0, y(0, a, x) = y0(a, x) (since F(0) = 1) and y(t, 0, x) = 0A β(a, x)y(t, a, x)da. Moreover,

y(T,a,x)=0a.e. (a,x)(δ,A)×(0,1)

and y satisfies the equation of (1.1) with

fδ=ξh1χω+ηh2χω1Tϕu3F(t)kϕu3,xF(t)(kϕu3)xkξu1,x(kξu1)xkηu2,x(kηu2)x.

Obviously, the support of fδ is contained in ω and, since kC1(ω), the terms (u3)x, (kξu1)x and (u2)x are L2(0, 1) (recall that ϕ′(x) = ξ′(x) = η′(x) = 0 for all x ∈ (0, 1) ∖ ω); thus fδL2(Q) as required. As in [19], estimate (4.33) follows by the definition of fδ, (4.36) and (2.2) for ui, i = 1, 2, 3.□

Observe that the previous result can be proved also for the problem in non divergence form considered in [13].

A Appendix

A.1 Proof of (2.2)

Multiplying the equation of (1.1) by y and integrating over (0, A) × (0, 1), we obtain

12ddty(t)L2(QA,1)2+1201y2(t,A,x)dx1201y2(t,0,x)dx+0A01kyx2dxda=0A01μy2dxda+0Aωfydxda.

Hence, using the initial condition y(t, 0, x) = 0A β(a, x) y(t, a, x)da, the assumptions on β and μ and the inequality of Jensen, one has

12ddty(t)L2(QA,1)2+1201y2(t,A,x)dx+0A01kyx2dxdaC20A01y2(t,a,x)dxda+120A01f2dxda, (A.1)

where C is a positive constant. Since 01y2(t,A,x)dxand0A01kyx2dxda are positive, we deduce

ddty(t)L2(QA,1)2Cy(t)L2(QA,1)2+f(t)L2(QA,1)2.

Setting F(t):=y(t)L2(QA,1)2 and multiplying the previous inequality by eCt, one has

ddt(eCtF(t))eCtf(t)L2(QA,1)2. (A.2)

Integrating (A.2) over (0, t), for all t ∈ [0, T] it follows

eCtF(t)F(0)+0teCτf(τ)L2(QA,1)2dτ.

Hence, for all t ∈ [0, T],

F(t)eCTF(0)+0Tf(τ)L2(QA,1)2dτ

and

supt[0,T]y(t)L2(QA,1)2Cy0L2(QA,1)2+fL2(Q)2dτ. (A.3)

Therefore, by (A.1), it follows

12ddty(t)L2(QA,1)2+0A01kyx2dxdaC20A01y2(t,a,x)dxda+120A01f2dxda.

Integrating over (0, T), we have

12y(T)L2(QA,1)2+0T0A01kyx2dxdadt12y0L2(QA,1)2+C20T0A01y2(t,a,x)dxdadt+120T0A01f2dxdadt.

Hence, by (A.3),

0T0AkyxL2(0,1)2dadty0L2(QA,1)2+C0Ty(t)L2(QA,1)2dt+fL2(Q)2Cy0L2(QA,1)2+fL2(Q)2dτ (A.4)

and (2.2) follows by (A.3) and (A.4).

A.2 Proof of Proposition 3.2

We consider case (i), Hypothesis (HP1). Fix β ∈ (θ, 1) arbitrarily for the moment. Since w(1) = 0, we have

01k(x)(1x)2w2(x)dx=01k(x)(1x)2(x1(1y)β/2w(y)(1y)β/2dy)2dx.

This implies

01k(x)(1x)2w2(x)dx01k(x)(1x)2(x1(1y)β|w(y)|2dyx1(1y)βdy)dx.

Hence, we have

01k(x)(1x)2w2(x)dx11β01k(x)(1x)1+β(x1(1y)β|w(y)|2dy)dx.

By the Theorem of Fubini, it follows

01k(x)(1x)2w2(x)dx11β01(1y)β|w(y)|2(0yk(x)(1x)1+βdx)dy. (A.5)

Now, divide the right hand side of (A.5) into three parts, i.e.

01(1y)β|w(y)|2(0yk(x)(1x)1+βdx)dy=Lϵ+Mϵ+Nϵ,

where

Lϵ=01ϵ(1y)β|w(y)|2(0yk(x)(1x)1+βdx)dy,Mϵ=1ϵ1(1y)β|w(y)|2(01ϵk(x)(1x)1+βdx)dy,

and

Nϵ=1ϵ1(1y)β|w(y)|2(1ϵyk(x)(1x)1+βdx)dy.

Thanks to our hypothesis, there exists ϵ > 0 such that the function

xk(x)(1x)θ is nondecreasing on [1ϵ,1);

thus, for Nϵ, we have

Nϵ=1ϵ1(1y)β|w(y)|2(1ϵyk(x)(1x)1+βdx)dy1ϵ1(1y)βθ|w(y)|2k(y)1ϵy(1x)θβ1dxdy1(βθ)1ϵ1k(y)|w(y)|2dy. (A.6)

For Mϵ, we have

Mϵsup[0,1ϵ]k1ϵ1(1y)βk(y)k(y)|w(y)|2(01ϵ(1x)(1+β)dx)dyϵθββk(1ϵ)sup[0,1ϵ]k1ϵ1(1y)βθk(y)|w(y)|2C1ϵ1k(x)|w(x)|2dx. (A.7)

Proceeding in a similar way, we obtain

LϵC01ϵk(x)|w(x)|2dx. (A.8)

Using (A.6), (A.7) and (A.8) in (A.5), we obtain

01k(x)(1x)2w2(x)dxC01k(x)|w(x)|2dx, (A.9)

where the constant C depends on a, ϵ, θ and β. If one assumes that Hypothesis (HP1)’ holds, that is

xk(x)(1x)θ is nondecreasing on [0,1),

then, one can take ϵ = 1 in the above computations, so that Lϵ = Mϵ = 0. Using then (A.6) in (A.5), with ϵ = 1, one obtains

01k(x)(1x)2w2(x)dx1(1β)(βθ)01k(x)|w(x)|2dx.

We then remark that this last estimate is optimal for β=θ+12, which gives the desired result.

We now consider case (ii), Hypothesis (HP2). Fix β ∈ (1, θ) arbitrarily for the moment. As before

01k(x)(1x)2w2(x)dx=01k(x)(1x)2(0x(1y)β/2w(y)(1y)β/2dy)2dx;

so that

01k(x)(1x)2w2(x)dx01k(x)(1x)2(0x(1y)β|w(y)|2dy0x(1y)βdy)dx.

It follows that

01k(x)(1x)2w2(x)dx1β101k(x)(1x)1+β(0x(1y)β|w(y)|2dy)dx.

Applying the Theorem of Fubini, we have

01k(x)(1x)2w2(x)dx1β101(1y)β|w(y)|2(y1k(x)(1x)1+βdx)dy. (A.10)

As before, we rewrite

01(1y)β|w(y)|2(y1k(x)(1x)1+βdx)dy=Iϵ+Jϵ+Kϵ,

where

Iϵ=01ϵ(1y)β|w(y)|2(y1ϵk(x)(1x)1+βdx)dy,Jϵ=01ϵ(1y)β|w(y)|2(1ϵ1k(x)(1x)1+βdx)dy

and

Kϵ=1ϵ1(1y)β|w(y)|2(y1k(x)(1x)1+βdx)dy.

Thanks to our hypothesis, there exists ϵ > 0 such that the function

xk(x)(1x)θ is nonincreasing on [1ϵ,1),

thus Kϵ can be estimated in the following way:

Kϵ=1ϵ1(1y)β|w(y)|2(y1k(x)(1x)1+βdx)dy1ϵ1(1y)β|w(y)|2k(y)(1y)θy1(1x)θβ1dx1(θβ)1ϵ1k(x)|w(x)|2dx. (A.11)

For Jϵ, we can proceed in a similar way, obtaining

Jϵ=01ϵ(1y)β|w(y)|2(1ϵ1k(x)(1x)1+βdx)dyϵβk(1ϵ)inf[0,1ϵ]k01ϵk(y)|w(y)|2dy. (A.12)

For Iϵ, we have

Iϵ=01ϵ(1y)β|w(y)|2k(y)k(y)(y1ϵk(x)(1x)1+βdx)dysup[0,1ϵ]kβinf[0,1ϵ]k01ϵk(y)|w(y)|2(1y)βϵβdyC01ϵk(x)|w(x)|2dx. (A.13)

Using (A.11), (A.12) and (A.13) in (A.10), we obtain

01k(x)(1x)2w2(x)dxC01k(x)|w(x)|2dx, (A.14)

where the constant C depends on a, ϵ, θ and β. If one assumes that hypothesis (HP2)’ holds, that is

xk(x)(1x)θ is noninreasing on [0,1),

then, one can take ϵ = 1 in the above computations, so that Jϵ = Iϵ = 0. Using then (2.1) in (A.10), with ϵ = 1, one obtains

01k(x)(1x)2w2(x)dx1(β1)(θβ)01k(x)|w(x)|2dx.

We then remark that this last estimate is optimal for β=θ+12, which gives the desired result.

A.3 Proof of Proposition 4.2

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

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

Then, integrating by parts one has

0=0Tddt0A01(ξesψ)2v2dxdadt=Q2sψt(ξesψ)2v2+2(ξesψ)2v(va(kvx)x+μv+f)dxdadt=2sQψt(ξesψ)2v2dxdadt+2sQψa(ξesψ)2v2dxdadt+2Qξ2e2sψxkvvxdxdadt+2Q(ξ2e2sψk)vx2dxdadt+2Qξ2e2sψμv2dxdadt+2Qξ2e2sψfvdxdadt.

Hence, using Young’s inequality

2Qξ2e2sψkvx2dxdadt=2sQψtξesψ2v2dxdadt2sQψa(ξesψ)2v2dxdadt2Qξ2e2sψxkvvxdxdadt2Qξ2e2sψμv2dxdadt2Qξ2e2sψfvdxdadt2sQψt(ξesψ)2v2dxdadt2sQψa(ξesψ)2v2dxdadt+Qk(ξ2e2sφ)xξesφv2dxdadt+Qξ2e2sφkvx2dxdadt.+(2μL(Q)+1)Qξ2v2dxdadt+Qξ2e2sψf2dxdadt.

Thus,

infω{k}0T0Aωe2sψvx2dxdadtsupω×(0,T){4kξesψx22s(ψt+ψa)(ξesψ)2}+2μL(Q)+10T0Aωv2dxdadt+Qf2e2sψdxdadt.

Acknowledgments

The author is a member of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM) and she is supported by the FFABR “Fondo per il finanziamento delle attività base di ricerca” 2017, by the INdAM- GNAMPA Project 2019 ”Controllabilità di PDE in modelli fisici e in scienze della vita” and by PRIN 2017-2019 Qualitative and quantitative aspects of nonlinear PDEs.

References

[1] B. Ainseba, Y. Echarroudi, L. Maniar Null controllability of population dynamics with degenerate diffusion, Differential Integral Equations (2013), 1397–1410.10.57262/die/1378327432Search in Google Scholar

[2] F. Alabau-Boussouira, P. Cannarsa, G. Fragnelli, 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

[3] S. Aniţa, Analysis and control of age-dependent population dynamics, Mathematical Modelling: Theory and Applications 11 (2000), Kluwer Academic Publishers, Dordrecht.10.1007/978-94-015-9436-3Search in Google Scholar

[4] V. Barbu, M. Iannelli, M. Martcheva, On the controllability of the Lotka-McKendrick model of population dynamics, J. Math. Anal. Appl. 253 (2001), 142—165.10.1006/jmaa.2000.7075Search in Google Scholar

[5] R. Borges, A. Calsina, S. Cuadrado, O. Diekmann, Delay equation formulation of a cyclin-structured cell population model, J. Evol. Equ. 14 (2014), 841–862.10.1007/s00028-014-0241-7Search in Google Scholar

[6] H. Brezis, Functional Analysis, Sobolev Spaces and Partial Differential Equations, Springer Science+Business Media, LLC 2011.10.1007/978-0-387-70914-7Search in Google Scholar

[7] P. Cannarsa, G. Fragnelli, D. Rocchetti, 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

[8] P. Cannarsa, P. Martinez, J. Vancostenoble, Persistent regional controllability for a class of degenerate parabolic equations, Commun. Pure Appl. Anal. 3 (2004), 607–635.10.3934/cpaa.2004.3.607Search in Google Scholar

[9] O. Diekmann, Ph. Getto, Boundedness, global existence and continuous dependence for nonlinear dynamical systems describing physiologically structured populations, J. Differential Equations 215 (2005), 268–319.10.1016/j.jde.2004.10.025Search in Google Scholar

[10] Y. Echarroudi, L. Maniar, Null controllability of a model in population dynamics, Electron. J. Differential Equations 2014 (2014), 1–20.Search in Google Scholar

[11] K.J. Engel, R. Nagel, One-Parameter Semigroups for Linear Evolution Equations, Springer-Verlag, New York, Berlin, Heidelberg, 1999.Search in Google Scholar

[12] E. Fernández-Cara, S. Guerrero, Global Carleman inequalities for parabolic systems and applications to controllability, SIAM J. Control Optim. 45, 1395-1446.s10.1137/S0363012904439696Search in Google Scholar

[13] G. Fragnelli, Carleman estimates and null controllability for a degenerate population model, Journal de Mathématiques Pures et Appliqués 115 (2018), 74—126.10.1016/j.matpur.2018.01.003Search in Google Scholar

[14] G. Fragnelli, P. Martinez, J. Vancostenoble, Qualitative properties of a population dynamics system describing pregnancy, Math. Models Methods Appl. Sci. 15 (2005), 507–554.10.1142/S0218202505000455Search in Google Scholar

[15] G. Fragnelli, D. Mugnai, Carleman estimates and observability inequalities for parabolic equations with interior degeneracy, Advances in Nonlinear Analysis 2 (2013), 339–378.10.1515/anona-2013-0015Search in Google Scholar

[16] G. Fragnelli, D. Mugnai, Carleman estimates, observability inequalities and null controllability for interior degenerate non smooth parabolic equations, Mem. Amer. Math. Soc. 242 (2016), v+84 pp. Corrigendum, to appear.10.1090/memo/1146Search in Google Scholar

[17] G. Fragnelli, D. Mugnai, Carleman estimates for singular parabolic equations with interior degeneracy and non smooth coefficients, Adv. Nonlinear Anal. 6 (2017), 61–84.10.1515/anona-2015-0163Search in Google Scholar

[18] G. Fragnelli, D. Mugnai, Controllability of strongly degenerate parabolic problems with strongly singular potentials, Electron. J. Qual. Theory Differ. Equ. 50 (2018), 1–11.10.14232/ejqtde.2018.1.50Search in Google Scholar

[19] G. Fragnelli, D. Mugnai, Controllability of degenerate and singular parabolic problems: the double strong case with Neumann boundary conditions, Opuscula Math. 39 (2019), 207—225.10.7494/OpMath.2019.39.2.207Search in Google Scholar

[20] G. Fragnelli, L. Tonetto, A population equation with diffusion, J. Math. Anal. Appl. 289 (2004), 90–99.10.1016/j.jmaa.2003.08.047Search in Google Scholar

[21] Y. He, B. Ainseba, Exact null controllability of the Lobesia botrana model with diffusion, J. Math. Anal. Appl. 409 (2014), 530–543.10.1016/j.jmaa.2013.07.020Search in Google Scholar

[22] J.L. Lions, E. Magenes, Non-homogeneous boundary value problems and applications. Vol. I. Translated from the French by P. Kenneth. Die Grundlehren der mathematischen Wissenschaften, Band 181. Springer-Verlag, New York-Heidelberg (1972).Search in Google Scholar

[23] D. Maity, M. Tucsnak and E. Zuazua, Controllability and positivity constraints in population dynamics with age structuring and diffusion, Journal de Mathématiques Pures et Appliqués. In press, 10.1016/j.matpur.2018.12.006.10.1016/j.matpur.2018.12.006Search in Google Scholar

[24] S. S. Win, R. Muhamad, Z. A. M. Ahmad, N. A. Adam, Population Fluctuations of Brown Plant Hopper Nilaparvata lugens Stal. and White Backed Plant Hopper Sogatella furcifera Horvath on Rica, Jounal of Entomology 8 (2011), 183–190.10.3923/je.2011.183.190Search in Google Scholar

[25] B. Zhong, C. Lv and W. Qin, Effect of Temperature on the Population Growth of Tirathaba rufivena (Lepidoptera: Pyralidae) on Areca catechu (Arecaceae), Florida Entomologist 100 (2017), 578–582.10.1653/024.100.0314Search in Google Scholar

Received: 2018-12-29
Accepted: 2019-05-16
Published Online: 2019-10-15

© 2019 Genni Fragnelli, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Articles in the same Issue

  1. Frontmatter
  2. On the moving plane method for boundary blow-up solutions to semilinear elliptic equations
  3. Regularity of solutions of the parabolic normalized p-Laplace equation
  4. Cahn–Hilliard equation on the boundary with bulk condition of Allen–Cahn type
  5. Blow-up solutions for fully nonlinear equations: Existence, asymptotic estimates and uniqueness
  6. Radon measure-valued solutions of first order scalar conservation laws
  7. Ground state solutions for a semilinear elliptic problem with critical-subcritical growth
  8. Generalized solutions of variational problems and applications
  9. Existence and non-existence results for Kirchhoff-type problems with convolution nonlinearity
  10. Nonlinear Sherman-type inequalities
  11. Global regularity for systems with p-structure depending on the symmetric gradient
  12. Homogenization of a net of periodic critically scaled boundary obstacles related to reverse osmosis “nano-composite” membranes
  13. Noncoercive resonant (p,2)-equations with concave terms
  14. Evolutionary quasi-variational and variational inequalities with constraints on the derivatives
  15. Sharp estimates on the first Dirichlet eigenvalue of nonlinear elliptic operators via maximum principle
  16. Localization and multiplicity in the homogenization of nonlinear problems
  17. Remarks on a nonlinear nonlocal operator in Orlicz spaces
  18. A Picone identity for variable exponent operators and applications
  19. On the weakly degenerate Allen-Cahn equation
  20. Continuity results for parametric nonlinear singular Dirichlet problems
  21. Construction of type I blowup solutions for a higher order semilinear parabolic equation
  22. Singularly perturbed Choquard equations with nonlinearity satisfying Berestycki-Lions assumptions
  23. Comparison results for nonlinear divergence structure elliptic PDE’s
  24. Constant sign and nodal solutions for parametric (p, 2)-equations
  25. Monotonicity formulas for coupled elliptic gradient systems with applications
  26. Berestycki-Lions conditions on ground state solutions for a Nonlinear Schrödinger equation with variable potentials
  27. A class of semipositone p-Laplacian problems with a critical growth reaction term
  28. The role of superlinear damping in the construction of solutions to drift-diffusion problems with initial data in L1
  29. Reconstruction of Tesla micro-valve using topological sensitivity analysis
  30. Lewy-Stampacchia’s inequality for a pseudomonotone parabolic problem
  31. Global well-posedness of nonlinear wave equation with weak and strong damping terms and logarithmic source term
  32. Regularity Criteria for Navier-Stokes Equations with Slip Boundary Conditions on Non-flat Boundaries via Two Velocity Components
  33. Homoclinics for singular strong force Lagrangian systems
  34. A constructive method for convex solutions of a class of nonlinear Black-Scholes equations
  35. On a class of nonlocal nonlinear Schrödinger equations with potential well
  36. Superlinear Schrödinger–Kirchhoff type problems involving the fractional p–Laplacian and critical exponent
  37. Regularity for minimizers for functionals of double phase with variable exponents
  38. Boundary blow-up solutions to the Monge-Ampère equation: Sharp conditions and asymptotic behavior
  39. Homogenisation with error estimates of attractors for damped semi-linear anisotropic wave equations
  40. A-priori bounds for quasilinear problems in critical dimension
  41. Critical growth elliptic problems involving Hardy-Littlewood-Sobolev critical exponent in non-contractible domains
  42. On the Sobolev space of functions with derivative of logarithmic order
  43. On a logarithmic Hartree equation
  44. Critical elliptic systems involving multiple strongly–coupled Hardy–type terms
  45. Sharp conditions of global existence for nonlinear Schrödinger equation with a harmonic potential
  46. Existence for (p, q) critical systems in the Heisenberg group
  47. Periodic traveling fronts for partially degenerate reaction-diffusion systems with bistable and time-periodic nonlinearity
  48. Some hemivariational inequalities in the Euclidean space
  49. Existence of standing waves for quasi-linear Schrödinger equations on Tn
  50. Periodic solutions for second order differential equations with indefinite singularities
  51. On the Hölder continuity for a class of vectorial problems
  52. Bifurcations of nontrivial solutions of a cubic Helmholtz system
  53. On the exact multiplicity of stable ground states of non-Lipschitz semilinear elliptic equations for some classes of starshaped sets
  54. Sign-changing multi-bump solutions for the Chern-Simons-Schrödinger equations in ℝ2
  55. Positive solutions for diffusive Logistic equation with refuge
  56. Null controllability for a degenerate population model in divergence form via Carleman estimates
  57. Eigenvalues for a class of singular problems involving p(x)-Biharmonic operator and q(x)-Hardy potential
  58. On the convergence analysis of a time dependent elliptic equation with discontinuous coefficients
  59. Multiplicity and concentration results for magnetic relativistic Schrödinger equations
  60. Solvability of an infinite system of nonlinear integral equations of Volterra-Hammerstein type
  61. The superposition operator in the space of functions continuous and converging at infinity on the real half-axis
  62. Estimates by gap potentials of free homotopy decompositions of critical Sobolev maps
  63. Pseudo almost periodic solutions for a class of differential equation with delays depending on state
  64. Normalized multi-bump solutions for saturable Schrödinger equations
  65. Some inequalities and superposition operator in the space of regulated functions
  66. Area Integral Characterization of Hardy space H1L related to Degenerate Schrödinger Operators
  67. Bifurcation of time-periodic solutions for the incompressible flow of nematic liquid crystals in three dimension
  68. Morrey estimates for a class of elliptic equations with drift term
  69. A singularity as a break point for the multiplicity of solutions to quasilinear elliptic problems
  70. Global and non global solutions for a class of coupled parabolic systems
  71. On the analysis of a geometrically selective turbulence model
  72. Multiplicity of positive solutions for quasilinear elliptic equations involving critical nonlinearity
  73. Lack of smoothing for bounded solutions of a semilinear parabolic equation
  74. Gradient estimates for the fundamental solution of Lévy type operator
  75. π/4-tangentiality of solutions for one-dimensional Minkowski-curvature problems
  76. On the existence and multiplicity of solutions to fractional Lane-Emden elliptic systems involving measures
  77. Anisotropic problems with unbalanced growth
  78. On a fractional thin film equation
  79. Minimum action solutions of nonhomogeneous Schrödinger equations
  80. Global existence and blow-up of weak solutions for a class of fractional p-Laplacian evolution equations
  81. Optimal rearrangement problem and normalized obstacle problem in the fractional setting
  82. A few problems connected with invariant measures of Markov maps - verification of some claims and opinions that circulate in the literature
Downloaded on 1.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2020-0034/html
Scroll to top button