Loading [MathJax]/extensions/TeX/boldsymbol.js
Research article Special Issues

Some comparison results and a partial bang-bang property for two-phases problems in balls

  • In this paper, we present two type of contributions to the study of two-phases problems. In such problems, the main focus is on optimising a diffusion function a under L and L1 constraints, this function a appearing in a diffusive term of the form (a) in the model, in order to maximise a certain criterion. We provide a parabolic Talenti inequality and a partial bang-bang property in radial geometries for a general class of elliptic optimisation problems: namely, if a radial solution exists, then it must saturate, at almost every point, the L constraints defining the admissible class. This is done using an oscillatory method.

    Citation: Idriss Mazari. Some comparison results and a partial bang-bang property for two-phases problems in balls[J]. Mathematics in Engineering, 2023, 5(1): 1-23. doi: 10.3934/mine.2023010

    Related Papers:

    [1] Filippo Gazzola, Elsa M. Marchini . The moon lander optimal control problem revisited. Mathematics in Engineering, 2021, 3(5): 1-14. doi: 10.3934/mine.2021040
    [2] Aleksandr Dzhugan, Fausto Ferrari . Domain variation solutions for degenerate two phase free boundary problems. Mathematics in Engineering, 2021, 3(6): 1-29. doi: 10.3934/mine.2021043
    [3] Valentina Volpini, Lorenzo Bardella . Asymptotic analysis of compression sensing in ionic polymer metal composites: The role of interphase regions with variable properties. Mathematics in Engineering, 2021, 3(2): 1-31. doi: 10.3934/mine.2021014
    [4] Hyeonbae Kang, Shigeru Sakaguchi . A symmetry theorem in two-phase heat conductors. Mathematics in Engineering, 2023, 5(3): 1-7. doi: 10.3934/mine.2023061
    [5] Lucio Boccardo, Giuseppa Rita Cirmi . Regularizing effect in some Mingione’s double phase problems with very singular data. Mathematics in Engineering, 2023, 5(3): 1-15. doi: 10.3934/mine.2023069
    [6] Stefano Biagi, Serena Dipierro, Enrico Valdinoci, Eugenio Vecchi . A Hong-Krahn-Szegö inequality for mixed local and nonlocal operators. Mathematics in Engineering, 2023, 5(1): 1-25. doi: 10.3934/mine.2023014
    [7] Guido De Philippis, Filip Rindler . Fine properties of functions of bounded deformation-an approach via linear PDEs. Mathematics in Engineering, 2020, 2(3): 386-422. doi: 10.3934/mine.2020018
    [8] Sandro Salsa, Francesco Tulone, Gianmaria Verzini . Existence of viscosity solutions to two-phase problems for fully nonlinear equations with distributed sources. Mathematics in Engineering, 2019, 1(1): 147-173. doi: 10.3934/Mine.2018.1.147
    [9] Cristiana De Filippis . Optimal gradient estimates for multi-phase integrals. Mathematics in Engineering, 2022, 4(5): 1-36. doi: 10.3934/mine.2022043
    [10] Filippo Gazzola, Gianmarco Sperone . Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040
  • In this paper, we present two type of contributions to the study of two-phases problems. In such problems, the main focus is on optimising a diffusion function a under L and L1 constraints, this function a appearing in a diffusive term of the form (a) in the model, in order to maximise a certain criterion. We provide a parabolic Talenti inequality and a partial bang-bang property in radial geometries for a general class of elliptic optimisation problems: namely, if a radial solution exists, then it must saturate, at almost every point, the L constraints defining the admissible class. This is done using an oscillatory method.



    Scope of the paper  In this paper, we aim at investigating several properties for a natural shape optimisation problem that arises in heterogeneous heat conduction: what is the optimal way to design the properties of a material in order to optimise its performance? This question has received a lot of attention from the mathematical community over the last decades [1,10,11,12,13,14,15,16,22,30] and our goal in this paper is to offer some complementary qualitative results. Mathematically, these problems are often dubbed two-phase problems and write, in their most general form, as follows: considering that the piece consists of a basic material, with conductivity α>0, we try to find the best location ω for the inclusion of another material having conductivity β>α. The resulting diffusive part of the equation under consideration writes

    ((α+(βα)1ω))). (1.1)

    This diffusive part is supplemented with a source term, and can be considered in elliptic or parabolic models. We study some aspects of both cases in the present paper in the case of radial geometries. To state the generic type of question we are interested in, we write down the typical equation in the elliptic case: for a ball Ω, a source term fL2(Ω) (the influence of which is also discussed) and an inclusion ωΩ, let uω be the unique solution of

    {((α+(βα)1ω)uω)=finΩ,uω=0onΩ. (1.2)

    We consider a volume constraint, enforced by a parameter V1(0;Vol(Ω)), and we investigate the problem

    supωΩ,Vol(ω)=V1J(ω):=Ωj(uω),

    for a certain non-linearity j. More specifically, we consider the set

    M(Ω):={aL(Ω):a=α+(βα)1ωforsomemeasurableωΩ,Vol(ω)=V1}, (1.3)

    also called the set of bang-bang functions, as well as its natural compactification for the weak L topology,

    A(Ω):={aL(Ω),αaβ,Ωa=V0:=αVol(Ω)+(βα)V1}.

    For aA(Ω), we define ua,f as the solution of (1.2) with α+(βα)1ω replaced with a. We will be interested in two formulations: the initial (unrelaxed) one

    supaM(Ω)Ωj(ua)

    as well as the relaxed one

    supaA(Ω)Ωj(ua).

    It should be noted that we will also for some results have to optimise with respect to the source term f, but that the main difficulty usually lies in handling the term a. The two formulations of the problem have their interest, as it may be interesting to see when the two coincides. In other words, is a solution to the second problem a solution of the first one? Let us already underline several basic facts: first, as is customary in this type of optimisation problems (we detail the references later on and for the moment refer to [31]) we do not expect existence of solutions in all geometry, and the proper type of relaxation should rather be of the H-convergence type. Nevertheless, we offer some results about these two problems. Second, the type of problems we are considering are not energetic (in the sense that the criterion we aim at optimising can not a priori be derived from the natural energy associated with the PDE constraint). This leads to several difficulties, most notably in handling the adjoint of the optimisation problem and in the ensuing loss of natural convexity or concavity of the functional to optimise. Third, we distinguish between two types of results: the first type correspond to Talenti inequalities, where we rearrange both the coefficient a and the source term f. In the elliptic case, this follows from results of [6], and our contribution here is the application of these methods to the parabolic case. A second type of result, given in Theorem II, deals with a possible identification of the two formulations (i.e., if a solution to the relaxed problem exists then it is a solution of the unrelaxed one) in radial geometries, and we do not need for this second type of results to rearrange the source term f. This result is the main contribution of this article.

    Informal statement of the results  Our goal is thus threefold. For the sake of presentation we indicate to which case (i.e., optimisation with respect to a, f or both) each item corresponds. We write ua,f for the solution of the equation with diffusion a and source term f. We will need a comparison inequality provided in [6] and that we recall in Theorem A.

    1) Existence and partial characterisation in radial geometries (optimisation with respect to a and f) This matter of existence and/or characterisation of optimal a in the case of radial geometries is the topic of the two first results. The Talenti inequality from [6] leads to a comparison principle, but leaves open the question of the existence of optimal shapes: if Ω is a centred ball, is there a radially symmetric solution a to the optimisation problem in Ω of the form

    a=α+(βα)1ω

    for some measurable subset ωΩ? We prove in Theorem I that it is the case when the function j is convex and we also allow ourselves to rearrange the source term f. We use the ideas contained in [15] to do so and prove this theorem for the sake of completeness; we highlight that the main contribution here is to prove that the methods of [15] work for non-energetic functionals.

    2) Weak bang-bang property under monotonicity assumption in radial geometries (optimisation with respect to a) The second result of the "elliptic problem" part, is the main result of this paper, Theorem II. In it, we give a weak bang-bang property that does not require convexity assumptions on the function j (and so no clear convexity on J). Namely, we prove that if, in a centred ball Ω, a solution a to the optimisation problem exists, and if j is increasing, then this solution has to be of bang-bang type. It is notable that, in this theorem, we do not require the term f to be rearranged as well and that we can handle non-energetic problem. This is proved by introducing, for two-phase problems, an oscillatory method reminiscent of the ideas of [25].

    3) Comparison results for parabolic models (optimisation with respect to time-dependent a and f) We provide, in Theorem III, a parabolic Talenti inequality. The proof is an adaptation of a result of [29], combined with the methods of [6].

    Plan of the paper  This paper is organised as follows: in Section 1.2 we present the models, the optimisation problems and give some elements about the Schwarz rearrangement. In Section 1.3 we state our main results. Section 1.4 contains the bibliographical references. The rest of the paper is devoted to the proofs of the main results. Finally, in the Conclusion, we state several open problems that we deem interesting.

    Henceforth, Ω is a centred ball in IRd, and V0(0;Vol(Ω)) is a fixed parameter that serves as a volume constraint. As explained in the first paragraph, we are interested in both elliptic and parabolic models. This leads us to define two admissible classes: the first one, used for elliptic problems, is

    A(Ω):={aL(Ω):αaβa.e.inΩ,Ωa=V0} (1.4)

    while the second, defined for a certain time horizon T>0, is

    A(Ω;T):={aL((0;T)×Ω):αaβa.e.in(0;T)×Ω,fora.e.t(0;T),Ωa(t,)=V0} (1.5)

    The set of admissible sources, on which we also place a volume constraint modelled via a constant F0(0;Vol(Ω)), is

    F(Ω):={fL(Ω):0f1a.e.inΩ,Ωf=F0}. (1.6)

    Similarly, we define, in the parabolic case,

    F(Ω;T):={fL((0;T)×Ω):0f1a.e.in(0;T)×Ω,fora.e.t(0;T)Ωf(t,)=F0}. (1.7)

    Main equation in the elliptic case  In the elliptic case, the main equation reads as follows: for any aA(Ω) and any fF(Ω), uell,a,f is the unique solution of

    {(auell,a,f)=finΩ,uell,a,f=0inΩ. (1.8)

    The solution uell,a,f is the unique minimiser in W1,20(Ω) of the energy functional

    Ea,f:W1,20(Ω)u12Ωa|u|2Ωfu. (1.9)

    Remark 1. Although for the classes A(Ω) and A(Ω;T) the lower bounds 0<αa ensure coercivity of the associated energy, it may be asked whether the non-negativity constraint on the sources can be relaxed. It may be difficult, as we need in our proofs the following crucial fact: when Ω is the ball, when f and a are radially symmetric functions of A(Ω) and F(Ω) respectively, the solution ua,f is radially non-increasing in Ω. This may not be the case, for instance when f<0 close to the center of the ball. Thus we choose simplicity and assume f0 almost everywhere.

    In the elliptic case, the goal is to solve the following problem: let jC1(IR) be a given non-linearity, then the problem is

    supaA(Ω),fF(Ω){Jell(a,f):=Ωj(uell,a,f)}Pell,j

    In [6], a comparison result that we will use later on is proved; we recall it in Theorem A. This comparison result states roughly speaking, that if j is increasing, there exist two radially symmetric functions ˜a and f such that Jell(a,f)Jell(˜a,f), with f still admissible; ˜a, however, may violate some constraints. Here, our main contribution is Theorem I, in which we prove it is possible to choose a radially symmetric ˜a that satisfies the constraints if we assume that j is convex and C2. This is done by adapting the methods of [15].

    Second, in Theorem II, we are interested in the following alternative formulation: fF(Ω) being fixed, solve

    supaA(Ω){Jell(a):=Ωj(uell,a,f)}.Pell,j,f

    We prove, using an oscillatory technique that, if a solution a exists and if j is increasing, then we must have aM(Ω). We underline that this result does not require rearranging f.

    Main equation in the parabolic case  In the parabolic case, the main equation reads as follows: for any aA(Ω;T), any fF(Ω;T), uparab,a,f is the unique solution of

    {uparab,a,ft(auparab,a,f)=fin(0;T)×Ω,uparab,a,f=0onΩ,uparab,a,f(0,)=0inΩ. (1.10)

    The parabolic optimisation problem assumes the following form: for two given non-linearities j1 and j2 in C1(IR) we consider the optimisation problem

    supaA(Ω;T),fF(Ω;T){Jparab(a,f):=(0;T)×Ωj1(uparab,a,f)+Ωj2(uparab,a,f(T))}.Pparab,j1,j2

    The main result is Theorem III, in which a parabolic isoperimetric inequality (with respect to the coefficient a) is obtained: namely, it is better to have radially symmetric a and f.

    In this section we recall the key points about the Schwarz rearrangement, which will be used constantly throughout this paper, and about the rearrangement of Alvino and Trombetti [6,7] that is crucial in dealing with two-phase isoperimetric problems.

    Schwarz rearrangement: definitions, properties and order relations  We refer to section 1.4 for further references, for instance for parabolic isoperimetric inequalities and for the time being we recall the basic definitions of the Schwarz rearrangement. We refer to [19,20,23] for a thorough introduction.

    Definition 2 (Schwarz rearrangement of sets). For a given bounded connected open set Ω0, the Schwarz rearrangement Ω0 of Ω0 is the unique centred ball BΩ0=B(0;RΩ0) such that

    Vol(BΩ0)=Vol(Ω0). (1.11)

    For rearrangements of functions, we use the distribution function: for any p[1;+), for any function uLp(Ω), u0, its distribution function is

    μu:IR+tVol({u>t}). (1.12)

    Definition 3 (Schwarz rearrangement of a function). For any function uLp(Ω0),u0, its Schwarz rearrangement is the unique radially symmetric function uLp(Ω0) having the same distribution function as u. u# stands for the one-dimensional function such that u=u#(cd||d) where cd:=Vol(B(0;1)).

    As a consequence of the equimeasurability of the function and of its rearrangement* there holds:

    * Two functions are called equimeasurable if they have the same distribution functions.

    p[1;+),uLp(Ω0),u0,Ωup=Ω0(u)p. (1.13)

    Two results are particularly important in the study of the Schwarz rearrangement:

    Hardy-Littlewood inequality: for any two non-negative functions f,gL2(Ω),

    Ω0fgΩ0fg. (1.14)

    PólyaSzegö inequality: for any p[1;+), for any uW1,p0(Ω),u0,

    uW1,p0(Ω0)andΩ0|u|pΩ0|u|p. (1.15)

    Finally, we will rely on an ordering of the set of functions.

    Definition 4. Let RΩ0>0 be the radius of the ball Ω0. For any two non-negative functions f,gL1(Ω0) we write

    fg

    if

    r[0;RΩ0],B(0;r)fB(0;r)g. (1.16)

    This ordering [17] provides the natural framework for comparison theorems in elliptic and parabolic equations [3,4,28,29,36,37,38]. The following property is proved in [4,Proposition 2]: for any non-decreasing convex function F such that F(0)=0, for any two non-negative functions f,gL1(Ω),

    fgF(f)F(g). (1.17)

    We now pass to the definition of rearrangement sets:

    Definition 5 (Rearrangement sets). For any non-negative function fL1(Ω0) we define

    CΩ0(f):={φL1(Ω0),φ=f} (1.18)

    and

    KΩ0(f):={φL1(Ω0),φ0a.e.,φf,Ω0φ=Ω0f}. (1.19)

    The following result can be found in [8,27,34]: for any non-negative fL1(Ω), KΩ(f) is a weakly compact, convex set; its extreme points are the elements of CΩ(f).

    The Alvino-Trombetti rearrangement: definition and property  The Alvino-Trombetti rearrangement is very useful when handling two-phase problems, and was introduced in [6,7] to establish some comparison principles for some elliptic equations with a diffusion matrix. The goal is the following: let uW1,20(Ω)L(Ω) be a non-negative function and let aA(Ω). We want to prove that there exists ˜a that is radially symmetric, such that ˜aId is uniformly elliptic and such that

    Ωa|u|2Ω˜a|u|2. (1.20)

    One defines ˜a as the unique radially symmetric function such that

    Fora.e.t(0;uL),{ut}1˜a={ut}1a. (1.21)

    It can be checked [15] that

    ˜a1KΩ((a)1).

    Remark 6. In particular, if all the level-sets of u have Lebesgue measure zero and the gradient of u does not vanish on these level-sets, this definition rewrites as

    Fora.e.t(0;uL),{u=t}1˜a|u|={u=t}1a|u|. (1.22)

    This fact follows from the co-area formula, which states in particular that

    {uc}1a=c{u=t}1a|u|.

    [7,Lemma 1.2] or [15,Proposition 4.9] assert that: for any u0,uW1,20(Ω), for any aA(Ω), ˜a being defined by (1.21), there holds

    Ωa|u|2Ω˜a|u|2. (1.23)

    Talenti inequalities for the relaxed problem  Let us start by recalling an application of the Alvino-Trombetti rearrangement to Talenti-like inequalities. Talenti inequalities originated in the seminal [36] and have, since then, been widely studied [2,3,4,8,9,28,29,37,38]. Roughly speaking, they amount to comparing, using the relation , the solution u of an elliptic problem with the solution u of a "symmetrised" elliptic equation. This first result [6] is the stepping stone to our main theorem and holds for the relaxed version of the problem:

    Theorem A ([6], Comparison results, optimisation w.r.t. a and f). Let Ω be a centred ball. For any aA(Ω) and any fF(Ω), ˜a being defined by (1.21), there holds

    uell,a,fuell,˜a,f. (1.24)

    As a consequence, for any increasing function j,

    Ωj(uell,a,f)Ωj(uell,˜a,f).

    Thus, it seems quite interesting to investigate whether the optimisation problem (Pell,j) has a radial solution. This would seem natural given the equation above. However, the Alvino-Trombetti rearrangement only provides us with a rearranged coefficient ˜a such that the inverse (˜a)1KΩ(a1). This last set is however different from A(Ω). The same problem arises when considering bang-bang functions a. We nonetheless obtain existence properties for the unrelaxed problem.

    Theorem I (Existence and bang-bang property in radial geometry for convex integrand, optimisation w.r.t. a and f). Assume j is a convex C2 function. Let R>0. Let Ω=B(0;R). The optimisation problem

    supaM(Ω),fF(Ω)Ωj(uell,a,f)

    has a solution (¯a,¯f)M(Ω)×F(Ω).

    The proof of this theorem is inspired by the proof of existence of optimal profiles for eigenvalue problems in [15].

    Finally, the last result for elliptic problems deals with a bang-bang property when optimising only with respect to a: is it true that, if we just assume that j is increasing, if a solution ¯a of (Pell,j,f) exists, then it is bang-bang? We can only partially answer this question, in the next theorem. It is the main result of our paper.

    Theorem II (Weak bang-bang property for increasing cost functions, optimisation w.r.t radially symmetric a). Assume j is an increasing function such that j>0 on IR+. Let R>0. Let Ω=B(0;R) and fF(Ω). Then, if the optimisation problem

    supaA(Ω),aradiallysymmetricΩj(uell,a,f)Pell,j,a

    has a solution ¯a, there holds

    ¯aM(Ω)or,inotherwords,¯a=α+(βα)1ωforsomemeasurableωΩ.

    The proof of this theorem is based on the development of an oscillatory method recently introduced in [25].

    In this second part, we state our main result devoted to the parabolic optimisation problem (Pparab,j1,j2). The proof of the parabolic isoperimetric inequality is done by adapting the proofs of Theorem A and of [29,Theorem 2.1]. For the sake of clarity, for a function u of two variables u=u(t,x), the notation u(t,) stands for the Schwarz rearrangement of u(t,) with respect to the space variable x.

    Theorem III (Comparison results, optimisation w.r.t. a and f). Let Ω=B(0;R).Let aM(Ω;T) and fF(Ω;T). Then there exists a radially symmetric function ˜a defined on (0;T)×Ω such that α˜aβ almost everywhere and such that, for almost every t(0;T) and every r(Ω;R) there holds

    B(0;r)uparab,a,fB(0;r)uparab,˜a,f.

    In particular, if j1 and j2 are convex increasing functions there holds

    Jparab(a,f)Jparab(˜a,f).

    Let us now offer some comments about this result, and about the method of proof.

    Remark 7 (Comments on Theorem III). 1) The first thing that has to be noted is that, exactly as in the elliptic case, although the new weight ¯a satisfies the correct upper and lower bounds α¯β, there is a priori no guarantee that ¯aM(Ω). Some other arguments would then be needed in order to conclude as to the integral constraint. It is not clear at this stage how one may go about this question.

    2) The second remark has to do with the method of proof that is employed. The two main available approaches in the context of parabolic equations are, on the one hand, dealing with the parabolic problem directly, as is done in [29] and as we do here, and on the other hand by time-discretisation of the evolution problem, as in [3]. We believe the second of these approaches may prove more delicate. To see why, let us recall the main steps of the proof of [3]: the authors, which in particular try to prove a comparison result for rearrangement of the source f in the parabolic equation

    utΔu=f,

    approximate this equation by the discretisation with time step NIN

    N(uk+1,Nuk,N)Δuk+1,N=fk,N

    where fk,N=N(k+1)/Nk/Nf. On each of these discretised problem, they use an elliptic Talenti inequality, yielding a comparison with the solution of the same system with fk,N as a right-hand side. This Schwarz symmetrisation operation is independent of the time-step (in the sense that the definition of f does not depend on the time step N). In our case however, since the Alvino-Trombetti rearrangement depends on the function uparab,a,f evaluated at the time t, this would translate, at the discretised level, as a rearrangement that would depend on both indexes k and N. This may lead to potential difficulties in passing to the limit.

    Let us underline that this type of parabolic comparison results can be very useful when dealing with parabolic eigenvalue optimisation problems, as is done for instance in [32,Theorem 3.9].

    In this paper, we offer contributions that may be viewed from several point of views, each of which stemming from very rich domains in mathematical analysis.

    Two-phase spectral optimisation problems  Two-phase optimisation problems have a rich history, and are deeply linked to homogenisation phenomenas. We refer, for instance, to [1,31] for a presentation of this rich theory, and we underline that one of the striking features of these problems is that there is often a lack of existence results. These results are typically obtained by proving that should an optimiser exist, then an overdetermined problem that can only solved in radial geometries should have a solution. This is done by using Serrin type theorems [35], and this phenomenon occurs in dimension d2. A typical and famous example of such problems is the optimisation of the first Dirichlet eigenvalue of the operator (a) under the constraint that aA(Ω). To the best of our knowledge, the proof of non-existence of an optimal aA(Ω) when Ω is not a ball was only recently completed in a series of papers by Casado-Diaz [10,11,12]. However, these negative results in the case of non-radial geometries do not allow to conclude as for the existence and/or characterisation of optimisers in radially symmetric domains. In this case, the same spectral optimisation problem being under consideration, the first proof of existence can be found in [15], using the Alvino-Trombetti rearrangement. We borrow from their ideas in the proof of Theorem I (and we highlight the fact that we do not consider here energetic problems). To underline the complexity of this spectral optimisation problem, let us also mention [22], in which it is shown that, in the ball, the qualitative features of the optimiser a strongly depend on the volume constraint. We also refer to [13,26] for the study of the spectral optimisation of operators with respect to a weight aA(Ω) that appears both in the principal symbol (a) and as a potential.

    Elliptic and parabolic Talenti inequalities  Talenti inequalities, which originate in the seminal [36] have been the subject of an intense research activity. For parabolic equations, the study of such inequalities started, as far as we are aware, in the works of Bandle [9], Vazquez [38] and were later deeply analysed by Alvino, Trombetti and Lions [2,3] on the one hand, and by Mossino and Rakotoson on the other [29]. We would like to mention that we have recently obtained a quantitative parabolic isoperimetric inequality for the source term in [24]. Alvino, Nitsch and Trombetti have recently proved an elliptic Talenti inequality under Robin boundary conditions, using a very fine analysis of the Robin problem [5]. This Robin Talenti inequality was then used in, for instance, [21,33].

    Proof of Theorem I. For the first part of the theorem, we consider the case Ω=B(0;R) where R>0 is a fixed constant. We work with functions aM(Ω). In other words, there exists ωΩ measurable such that

    a=α+(βα)1ω,

    and we aim at solving

    supaM(Ω),fF(Ω)J(f,a)=Ωj(uell,a,f),

    under the assumption that j is convex on IR+.

    Let us first note that for any aM(Ω) we have a=α+(βα)1B where B=B(0;r) satisfies βVol(B)+Vol(ΩB)=V0. As a consequence, for any a1,a2M(Ω),

    K((a1)1)=K((a2)1)

    where K() is the rearrangement class defined in definition 5. For the sake of notational convenience, we define

    K:=K(a1)whereaisanyelementofM(Ω).

    By Theorem A, for any aA(Ω) there exists a radially symmetric ˜a such that

    ˜a1K,J(˜a,¯f)J(a,f)

    where ¯f is simply the Schwarz rearrangement of f. By convexity of the functional with respect to f, ¯f is a bang-bang function. We henceforth consider it fixed and focus on optimisation with respect to a.

    The problem with this reformulation is that there is a priori no guarantee that ˜aM(Ω), and it is in general false. To overcome this difficulty, we now focus on a slightly simplified version of our problem:

    suparadiallysymmetrics.t.˜aKΩj(ua)whereuasolves{(aua)=¯finΩ,ua=0onΩ. (2.1)

    Another refomulation, a priori encompassing a larger class, is

    sup˜μradiallysymmetrics.t.μK(H(μ):=Ωj(vμ))wherevμsolves{(1μvμ)=¯finΩ,vμ=0onΩ. (2.2)

    We now proceed in several steps, following the ideas of [15]:

    1) Existence of solutions to (2.2): we first prove, in lemma 1, that there exists a solution ¯μ to (2.2). This is done via the direct method in the calculus of variations.

    2) The bang-bang property for ¯μ: we then prove, in Lemma 9, that any solution ¯μ of (2.2) is a bang-bang function. In other words, there exists a measurable subset ωΩ such that

    ¯μ=1α+(βα)1ω.

    This is done via a convexity argument. As a consequence, 1¯μM(Ω), hence concluding the proof.

    Existence of solutions to (2.2)   The main result of this paragraph is the following lemma:

    Lemma 8. There exists a solution ¯μ of the variational problem (2.2).

    Proof of Lemma 8. We first note the following fact: if a sequence of radially symmetric functions {μk}KIN weakly converges in L to μ (which is an element of K by the closedness of K for the weak convergence, see [27]) then, up to a subsequence,

    H(μk)kH(μ).

    Obtaining this result boils down to proving that, for any sequence of radially symmetric functions {μk}KIN weakly converges in L to μ there holds

    (μkkμ)(vμkkvμa.e.uptoasubsequence).

    Indeed, it then simply suffices to use the dominated convergence theorem to obtain the required result. Let us then prove that for any sequence of radially symmetric functions {μk}KIN weakly converging in L to μ,

    vμkkvμinL2(Ω). (2.3)

    However, since we are working with radially symmetric functions, this follows from explicit integration in radial coordinates of

    {(1μvμk)=¯finΩ,vμk=0onΩ

    which gives, for any kIN (and with a slight abuse of notation),

    rd1vμk(r)=μk(r)r0sd1f(s)ds.

    Thus, from the Rellich-Kondrachov embedding

    vμkkvμ{weaklyinW1,20(Ω),stronglyinL2(Ω).

    It suffices to extract a subsequence that is converging almost everywhere.

    We turn back to the proof of the lemma: let {μk}kIN be a maximising sequence for (2.2). Since the set K is weakly compact, and since for any kIN μk is radially symmetric, there exists a radially symmetric μK such that, up to a subsequence,

    μkkμweaklyinL.

    Hence, up to a subsequence,

    H(μk)kH(μ)

    so that μ is a solution of (2.2).

    The bang-bang property for ¯μ   We now present the key point of the proof of Theorem I, the bang-bang property.

    Lemma 9. Any solution ¯μ of (2.2) is of bang-bang type: there exists ωΩ such that

    μ=1α+(βα)1ω.

    Proof of Lemma 9. We argue by contradiction and assume that there exists a solution ¯μ of (2.2) that is not of bang-bang type. We will reach a conclusion using a second order information on the functional H, namely, by using the first and second order Gâteaux-derivative of the functional H. Let us first observe that it is standard [18] to see that the map Kμvμ is Gâteaux-differentiable. Furthermore, for a given μK and an admissible perturbation h at μ (i.e., such that μ+thK for t>0 small enough) the first order Gâteau-derivative of vμ in the direction h is the unique solution ˙vμ(h) of

    {(1μ˙vμ(h))+(hμ2vμ)=0inΩ,˙vμ(h)=0onΩ, (2.4)

    and the first Gâteaux-derivative of H at μ in the direction h is given by

    ˙H(μ)[h]=Ωj(vμ)˙vμ(h). (2.5)

    This leads to introducing the adjoint state pμ as the unique solution of

    {(1μpμ)=j(vμ)inΩ,pμ=0onΩ, (2.6)

    Remark 10. It should be noted that by explicit integration of the equation on vμ in radial coordinate, vμL(Ω); as a consequence, j(vμ) is an L function, so that pμ is well-defined.

    Multiplying, on the one hand (2.4) by pμ, on the other hand (2.6) by vμ, and integrating by parts gives

    ˙H(μ)[h]=Ω1μpμ,˙vμ(h)=Ωhμ2vμ,pμ=Ωhvμμ,pμμ. (2.7)

    We now compute the second order derivative of the criterion in a similar manner: the second order Gâteaux derivative of vμ at μ in the direction h is zero. In other words, denoting by ¨vμ this second order derivative, we have

    ¨vμ=0.

    Indeed, this follows from the explicit computation of vμ as

    rd1vμk(r)=μk(r)r0sd1f(s)ds.

    Thus, it appears that μvμ is linear. As a consequence we have that the second order Gâteaux derivative of H at μ in the direction h, henceforth abbreviated as ¨H(μ), is given by

    ¨H(μ)=Ωj(vμ)(˙vμ)2+Ωj(vμ)¨vμ=Ωj(vμ)(˙vμ)2. (2.8)

    Hence, if j is convex, so is H. Thus any solution ¯μ of (2.2) is an extreme point of K. In other words

    ¯μC(a1)whereaisanyelementofM(Ω).

    It follows that 1¯μM(Ω).

    Conclusion of the proof   As noted at the beginning of the proof, for any aM(Ω) there exists ˜a such that ˜a1K and such that J(˜a,¯f)J(a,f). Since

    H(1˜a)=J(˜a,¯f)

    it follows that

    J(˜a,¯f)H(¯μ)where¯μisthesolutionof(2.2)givenbylemma8.

    From proposition 2 ¯μ is a bang-bang function. As a consequence, 1¯μ:=¯a is an element of M(Ω). Thus

    J(a,f)J(˜a,¯f)H(¯μ)=J(¯a,¯f).

    Thus ¯a is a solution of the initial optimisation problem.

    The proof of the theorem is now complete.

    Proof of Theorem II. Throughout this proof we assume that we are given a radially symmetric solution ¯a of the optimisation problem

    supaA(Ω),aradiallysymmetricΩj(uell,a,f)

    and we want to prove that ¯aM(Ω). To reach the desired conclusion we argue by contradiction and we assume that ¯aM(Ω). We emphasise once again that this proof does not require rearranging the source term f. Since f is assumed to be fixed, we write J(a) for J(a,f)=Ωj(uell,a,f) and ua for uell,a,f.

    Let us single out the following result, that follows from direct integration in radial coordinates of (1.10):

    Lemma 11. For any radially symmetric aA(Ω) and fF(Ω), uaW1,(Ω), ua is radial and we furthermore have, with a slight abuse of notation, for a.e. r(0;R),

    ua(r)=1a(r)rd1r0ξd1f(ξ)dξ. (3.1)

    In particular, ua is a non-positive function and, for any ε>0, sup[ε;R]ua<0. It is strictly decreasing if f>0 in a neighbourhood of 0.

    We now compute the Gateaux derivatives of both the maps auell,a,f and of aJ(a) (we note that the fact that both maps are Gateaux differentiable follow from standard arguments). We note that, to compute them, it is not necessary to assume that the coefficients a and f are radiallly symmetric.

    The first-order Gateaux derivative of uell,a,f at a in an admissible direction h (i.e., such that a+thA(Ω) for t>0 small enough), denoted by ˙ua, is the unique solution to

    {(a˙ua)=(hua)inΩ,˙ua=0onΩ. (3.2)

    The Gateaux derivative of J at a in the direction h is given by

    ˙J(a)[h]=Ωj(ua)˙ua. (3.3)

    This leads to introducing the adjoint state pa as the unique solution to

    {(apa)=j(ua)inΩ,pa=0onΩ. (3.4)

    Multiplying (3.4) by ˙ua and (3.2) by pa and integrating by parts gives

    ˙J(a)[h]=Ωj(ua)˙ua=Ωapa,˙ua=Ωhua,pa. (3.5)

    In the same way, the second order Gateaux derivative of ua at a in the direction h, denoted by ¨ua, is the unique solution to

    {(a¨ua)=2(h˙ua)inΩ,¨ua=0onΩ, (3.6)

    and the second order Gateaux derivative of J at a in the direction h is given by

    ¨J(a)[h,h]=Ωj(ua)(˙ua)2+Ωj(ua)¨ua. (3.7)

    However, multiplying (3.6) by pa, integrating by parts and using the weak formulation of (3.4) yields

    Ωj(ua)¨ua=Ωapa,¨ua=2Ωh˙ua,pa. (3.8)

    Plugging (3.8) in (3.7) gives

    ¨J(a)[h,h]=Ωj(ua)(˙ua)22Ωh˙ua,pa. (3.9)

    We now use the radial symmetry assumption: since h,a and f are radially symmetric, and since ua(0)=˙ua(0)=0, (3.2) implies, in radial coordinates, as

    a˙ua=hua. (3.10)

    Furthermore, we have the following lemma:

    Lemma 12. If j>0 on IR+ then pa is a radially symmetric decreasing function:

    pa<0in(0;R).

    Proof of Lemma 12. The fact that pa is decreasing simply follows from, first, the strong maximum principle which implies that

    ua>0in[0;R)

    and, second, from explicit integration of the equation on pa in radial coordinates, which gives

    pa(r)=1a(r)rd1r0j(ua)rd1dr<0forr>0.

    The radiality of pa implies

    ˙ua,pa=˙uapa.

    As a consequence of (3.10), we have, for a constant Md>0

    Ωh˙ua,˙pa=MdR0rd1h(r)˙ua(r)pa(r)dr=MdR0a(r)(˙ua)2paua(r)dr=MdR0a(r)(˙ua)2apaaua(r)dr.

    Let us first define

    φ:r(0;R]apaaua(r).

    We observe that φ>0 in (0;R] and that, as r0,

    liminfr0φ(r)liminfr0j(ua(0))f(r)>0. (3.11)

    If f>0 in a neighbourhood of 0 then we can extend φ by j(ua(0))f(0)>0 in 0. If on the other hand f=0 in a neighbourhood of 0, then φ+ when r0. In any case, there exists a constant A>0 such that

    φA02>0in[0;R].

    We define the function

    Φ:Ωxφ(|x|),

    and we thus have

    Ωh˙ua,˙pa=2MdR0rd1h(r)˙ua(r)pa(r)dr=2MdR0a(r)(˙ua)2paua(r)dr=2ΩΦa|˙ua|2A0Ωa|˙ua|2.

    Finally, as jC2(IR+) and uaL(Ω) there exists a constant B>0 such that

    j(ua)BinΩ. (3.12)

    We end up with the following estimate on ¨J(a)[h,h]:

    ¨J(a)[h,h]=Ωj(ua)(˙ua)22Ωh˙ua,pa (3.13)
    A0Ωa|˙ua|2BΩ˙u2a. (3.14)

    Remark 13. It should be noted that, at this level, we recover the convexity of the functional J if we assume that j0. Indeed, in that case we can take B=0.

    Let us now turn back to the core of the proof: we have a maximiser ¯aM(Ω). Let

    ˜ω:={α<a<β}.

    Since ¯aM(Ω),

    Vol(˜ω)>0.

    Furthermore, for any hL(˜ω), extended by 0 outside of ˜ω and such that ˜ωh=0, we have (since both h and h are admissible perturbations at ¯a)

    ˙J(a)[h]=0. (3.15)

    To reach a contradiction, it suffices to build hL(˜ω){0} such that

    \begin{equation} \boldsymbol{\int_{{{{\Omega}}}}h 1_{{{\tilde{\omega}}}} = 0}\;{\rm{ and }}\; \boldsymbol{\ddot{\mathcal J}(a)[h1_{{\tilde{\omega}}}, h1_{{\tilde{\omega}}}] > 0.} \end{equation} (3.16)

    Actually, by approximation it suffices to build h\in L^2({{\tilde{\omega}}}) such that (3.16) holds. For the sake of notational simplicity, for any h\in L^2({{\tilde{\omega}}}) , we identify h with h1_{{\tilde{\omega}}}\in L^2({{{\Omega}}}) . To obtain the existence of such a perturbation we single out the estimate

    \begin{equation} \ddot{\mathcal J}(a)[h, h] \geqslant A_0\int_{{{{\Omega}}}}a |{{\nabla}} \dot u_a|^2-B\int_{{{{\Omega}}}}\dot u_a^2. \end{equation} (3.17)

    We introduce the sequence \{\sigma_k\, , \psi_k\}_{k\in {\rm{I\kern-0.21emN}}\; ^*} of eigenvalues of the operator -\nabla\cdot(a{{\nabla}}) . We pick a non-decreasing of the eigenvalue sequence:

    0 < \sigma_0 \leqslant \sigma_1 \leqslant \dots \leqslant \sigma_k\underset{k\to \infty}\rightarrow +\infty.

    The eigenequations are given by

    \begin{equation} \forall k\in {\rm{I\kern-0.21emN}}\;^*\, , \begin{cases}-\nabla \cdot(a{{\nabla}} \psi_k) = \sigma_k\psi_k&{\rm{ in }}\;{{{\Omega}}}\, , \\ \psi_k = 0&{\rm{ on }}\;\partial {{{\Omega}}}\, , \\ \int_{{{{\Omega}}}} \psi_k^2 = 1.\end{cases} \end{equation} (3.18)

    For any admissible perturbation h at a , we decompose \dot u_a in this basis as

    \begin{equation} \dot u_a = \sum\limits_{k = 1}^\infty\alpha_k(h) \psi_k, \end{equation} (3.19)

    where the coefficients \left\{\alpha_k(h)\right\}_{k\in {\rm{I\kern-0.21emN}}\; ^*} are determined by equation (3.2). If we assume that, for an integer K large enough, we have

    \begin{equation} \forall k \leqslant K\, , \alpha_k(h) = 0\, , \sum\limits_{k = K}^\infty\alpha_k(h)^2 > 0 \end{equation} (3.20)

    then we obtain, by expanding the right hand-side of (3.17),

    \begin{equation} \boldsymbol{\ddot{\mathcal J}(a)[h1_{{\tilde{\omega}}}, h1_{{\tilde{\omega}}}] \geqslant A\sum\limits_{k = K}^\infty\sigma_k {\alpha_k(h)^2}-B\sum\limits_{k = K}^\infty{\alpha_k(h)^2} \geqslant \left(A\sigma_K-{B}{}\right)\sum\limits_{k = K}^\infty \frac{\alpha_k(h)^2}{\sigma_k} > 0.} \end{equation} (3.21)

    As a consequence it remains to construct a perturbation h such that

    \begin{equation} \nabla\cdot (h{{\nabla}} u_a) = \sum\limits_{k \geqslant K}\eta_k \psi_k. \end{equation} (3.22)

    We need however to be careful, since \nabla \cdot \left(h{{\nabla}} u_a\right) merely lies in W^{-1, 2}({{{\Omega}}}) . To overcome this difficulty, we define, for any h\in L^2({{\tilde{\omega}}}) , the coefficient

    \begin{equation} \eta_k(h): = \int_{{{{\Omega}}}} h \langle {{\nabla}} u_a\, , {{\nabla}} \psi_k\rangle. \end{equation} (3.23)

    Since \nabla \cdot \left(h{{\nabla}} u_a\right)\in W^{-1, 2}({{{\Omega}}}) , each of this quantities is well-defined. Furthermore, setting, for any k\in {\rm{I\kern-0.21emN}}\; ,

    \begin{equation} \alpha_k(h): = \int_{{{{\Omega}}}} \psi_k \dot u_a \end{equation} (3.24)

    we have

    \begin{equation} \alpha_k(h): = \int_{{{{\Omega}}}} \psi_k \dot u_a = \frac1{\sigma_k}\int_{{{{\Omega}}}} a\langle {{\nabla}} \psi_k\, , {{\nabla}} \dot u_a\rangle = \frac{\eta_k(h)}{\sigma_k}. \end{equation} (3.25)

    Since \dot u_a\in W^{1, 2}_0({{{\Omega}}}) we have, in W^{1, 2}_0({{{\Omega}}}) , the decomposition

    \begin{equation} \dot u_a = \sum\limits_{k = 1}^\infty \frac{\eta_k(h)}{\sigma_k}\psi_k. \end{equation} (3.26)

    As a consequence, to ensure a decomposition of the form (3.19) it suffices to find, for K large enough, an h\in L^2({{\tilde{\omega}}}) such that

    h is radially symmetric,

    \Vert h\Vert_{L^2({{\tilde{\omega}}})} = 1\, ,

    ● For any k = 1, \dots, K-1 , \eta_k(h) = 0 ,

    ● There holds \int_{{{{\Omega}}}}h1_{{\tilde{\omega}}} = \int_{{\tilde{\omega}}} h = 0 .

    We define L^2_{{\operatorname{rad}}}(\tilde \omega) as the space of radially symmetric functions in L^2(\tilde \omega) . Let us first note that for any k\in \{1, \dots, K\} the linear maps \eta_k: = h\mapsto \eta_k(h) are continuous on L^2_{{\operatorname{rad}}}(\tilde \omega) . This continuity property is a consequence of the radial symmetry assumption on the coefficients, which from Lemma 11 implies that {{\nabla}} u_a\in L^\infty({{{\Omega}}}) . Indeed, we can then simply write

    \begin{align*} \left| \eta_k(h)\right|& \leqslant \Vert {{\nabla}} u_a\Vert_{L^\infty({{{\Omega}}})} \int_{{{{\Omega}}}} \vert h\vert \cdot\vert {{\nabla}} \psi_k\vert \\& \leqslant \Vert {{\nabla}} u_a\Vert_{L^\infty({{{\Omega}}})} \Vert {{\nabla}} \psi_k\Vert_{L^2({{{\Omega}}})}\Vert h\Vert_{L^2({{{\Omega}}})} \\& = \Vert {{\nabla}} u_a\Vert_{L^\infty({{{\Omega}}})} \Vert {{\nabla}} \psi_k\Vert_{L^2({{{\Omega}}})}\Vert h\Vert_{L^2({{\tilde{\omega}}})}, \end{align*}

    whence the continuity.

    Defining R_0\in L^2_{{\operatorname{rad}}}({{\tilde{\omega}}})' as

    R_0(h): = \int_{{\tilde{\omega}}} h,

    which is obviously continuous on L^2({{\tilde{\omega}}}) we are hence looking for h_k such that

    \begin{equation} \Vert h_K\Vert_{L^2({{\tilde{\omega}}})}\, , h_K\in \ker(R_0)\cap \left(\bigcap\limits_{k = 1}^{K-1} \ker(\eta_k)\right). \end{equation} (3.27)

    However, L^2_{{\operatorname{rad}}}({{\tilde{\omega}}}) is an infinite dimensional Hilbert space, \ker(R_0) and \bigcap_{k = 1}^{K-1} \ker(\eta_k) are closed subspaces of finite co-dimension, hence \ker(R_0)\cap \left(\bigcap_{k = 1}^{K-1} \ker(\eta_k)\right) has finite co-dimension. In particular it is non empty, so there exists h_K such that (3.27) holds. The conclusion follows.

    Proof of Theorem III. For the proof of the parabolic Talenti inequalities we follow the main ideas of [29,Theorem 2.1] and of [7]. Since the proof is very similar we mostly present the main steps. To alleviate notations, we simply write u_{a, f} for {{u_{{{\operatorname{parab}}}, a, f}}} . For a fixed a\in \mathcal M({{\Omega}}; T) , we define, for almost every t\in (0;T) , \tilde a(t, \cdot) as the Alvino-Trombetti rearrangement of a(t, \cdot) with respect to {{u_{a, f}}}(t, \cdot) . In other words, for almost every t\in (0;T) and almost every s\in (0;\Vert {{u_{a, f}}}(t, \cdot)\Vert_{L^\infty}) ,

    \int_{\{u_{a, f}^*(t, \cdot) \geqslant s\}}\frac1{\tilde a (t, \cdot)} = \int_{\{u_{a, f}(t, \cdot) \geqslant s\}}\frac1{a(t, \cdot)}.

    From [7,Proof of Lemma 1.2] we have, with S_d = d\mathrm{Vol}({{\mathbb B}}(0, 1))^{\frac1d} ,

    S_d^2 \mu_{{u_{a, f}}}(t, s)^{2-\frac2d} \leqslant\left(-\frac{d}{ds}\int_{\{u_{a, f}(t, \cdot) > s\}}\frac1{a}\right)\left(-\frac{d}{ds}\int_{\{u_{a, f}(t, \cdot) > s\}} a|{{\nabla}} {{u_{a, f}}}|^2\right).

    Let \overline a is the one-dimensional counterpart of \tilde a (i.e., \tilde a = \overline a(c_d|\cdot|^d) ). Then, as in [7], there holds, almost everywhere,

    -\frac{d}{ds}\int_{\{u_{a, f}(t, \cdot) > s\}}\frac1{a} = -\frac{\frac{\partial \mu_{{u_{a, f}}}}{\partial s}}{\overline a}(t, s).

    On the other hand the same arguments as in [3,Proof of Theorem 1] (see also [36]) show that, almost everywhere

    -\frac{d}{ds}\int_{\{u_{a, f}(t, \cdot) > s\}}a|{{\nabla}} u_{a, f}|^2 = \int_{\{u_{a, f}(t, \cdot) > s\}}\left(f-\frac{\partial u_{a, f}}{\partial t}\right).

    We can hence conclude that

    \begin{equation} S_d^2\overline a \mu_{{u_{a, f}}}(t)^{2-\frac2d} \leqslant -\frac{\partial \mu_{{u_{a, f}}}}{\partial s}(t, s)\int_{\{{{u_{a, f}}} > t\}}\left(f-\frac{\partial {{u_{a, f}}}}{\partial t}\right). \end{equation} (4.1)

    We now rewrite

    \int_{\{{{u_{a, f}}}(t, \cdot) > s\}}\frac{\partial {{u_{a, f}}}}{\partial t} = \int_0^{\mu_{{u_{a, f}}}(t, s)}\frac{\partial u^\#_{a, f}}{\partial t}.

    Introducing as in [29] the function k defined as

    k(t, \xi): = \int_0^\xi u^\#(t, \cdot)

    we hence obtain

    \begin{equation} \int_{\{{{u_{a, f}}}(t, \cdot) > s\}}\frac{\partial {{u_{a, f}}}}{\partial t} = \frac{\partial k}{\partial t}(t, \mu(t, s)). \end{equation} (4.2)

    By the Hardy-Littlewood inequality we have

    \begin{equation} \int_{\{{{u_{a, f}}} > t\}}f \leqslant \int_0^{\mu_{{{u_{a, f}}}}(t, s)}f^\#. \end{equation} (4.3)

    Combining these estimates we are left with

    \begin{equation} 1 \leqslant -S_d^{-2}\overline a^{-1} \mu_{{u_{a, f}}}(t, s)^{-2+\frac2d} \frac{\partial \mu_{{u_{a, f}}}}{\partial s}(t, s)\left( \int_0^{\mu_{{{u_{a, f}}}}(t, s)}f^\#-\frac{\partial k}{\partial t}(t, \mu(t, s))\right) \end{equation} (4.4)

    which, after integration, gives

    \begin{equation} 0 \leqslant -\frac{\partial^2 k}{\partial \xi^2} \leqslant S_d^{-2}\overline a^{-1} \xi^{-2+\frac2d} \left( \int_0^{\xi}f^\#-\frac{\partial k}{\partial t}(t, \xi)\right). \end{equation} (4.5)

    We denote by k_* the function obtained by replacing {{u_{a, f}}} by u_{\tilde a, f^*} in the definition of k . Since all the previous inequalities become equalities in this case it follows that the function K: = k-k^* satisfies

    \begin{equation} \begin{cases} \frac{\partial k}{\partial t}-S_d^2\overline a\xi^{2-\frac{2}d}\frac{\partial^2 K}{\partial^2 \xi^2} \leqslant 0&{\rm{ in }}\;(0;{{\operatorname{Vol}}}({{\Omega}}))\times (0;T) \\ K(0, \cdot) = 0\, , \\ K(t, 0) = 0 = \frac{\partial K}{\partial \xi}(t, {{\operatorname{Vol}}}({{\Omega}})). \end{cases} \end{equation} (4.6)

    From the maximum principle, we have K \leqslant 0 , so that the conclusion follows. If j_1 and j_2 are convex non-decreasing functions, the second conclusion of the theorem follows from [4,Proposition 2].

    In this paper, we have undertaken the study of certain non-energetic two-phase optimisation problems. Of course, our results are partial, and we now present some open problems that we think are worth investigating.

    Open problem I: rearrangements for the time-independent case   The first crucial question has to do with the parabolic problem. Indeed, since the Alvino-Trombetti rearrangement we use is defined differently for every time t , the question of time-independent a remains completely open, and we believe it may be fruitful to investigate in the future.

    Open problem II: possible relaxations of the problem, bang-bang property for the parabolic optimisation problem   The second problem has to do with the conclusion of Theorem I. A more satisfying conclusion that we could not reach would have been a weak bang-bang property, namely that, a profile a\in \mathcal A({{\Omega}}) being given, there exists \tilde a\in \mathfrak M({{\Omega}}) that improves the criterion. Usually, this type of property is obtained using the convexity or concavity of the functional. However, here, what we proved in Theorem II was that the second-order derivative of the functional is positive on an infinite dimensional subspace of the space of admissible perturbations. It is unclear whether this weaker information may be sufficient.

    Open problem III: Robin boundary conditions   Finally, let us note that, following the recent progresses in the study of Robin Talenti inequalities [5], it may be very interesting to try and understand which type of rearrangement of the weight a may be suitable to obtain Talenti inequalities for two-phases problems under Robin boundary conditions.

    I. Mazari was partially supported by the Austrian Science Fund (FWF) through the grant I4052-N32. This work was also partially funded by the French ANR Project ANR-18-CE40-0013 - SHAPO on Shape Optimization and by the Project "Analysis and simulation of optimal shapes - Application to lifesciences" of the Paris City Hall. The author would like to warmly thank the referee for his or her suggestions.

    The author declares no conflict of interest.



    [1] G. Allaire, Shape optimization by the homogenization method, New York: Springer, 2002. http://dx.doi.org/10.1007/978-1-4684-9286-6
    [2] A. Alvino, P. Lions, G. Trombetti, A remark on comparison results via symmetrization, P. Roy. Soc. Edinb. A, 102 (1986), 37–48. http://dx.doi.org/10.1017/S0308210500014475 doi: 10.1017/S0308210500014475
    [3] A. Alvino, P.-L. Lions, G. Trombetti, Comparison results for elliptic and parabolic equations via Schwarz symmetrization, Ann. Inst. H. Poincaré Anal. Non Linéaire, 7 (1990), 37–65. http://dx.doi.org/10.1016/S0294-1449(16)30303-1 doi: 10.1016/S0294-1449(16)30303-1
    [4] A. Alvino, P.-L. Lions, G. Trombetti, Comparison results for elliptic and parabolic equations via symmetrization: a new approach, Differ. Integral Equ., 4 (1991), 25–50.
    [5] A. Alvino, C. Nitsch, C. Trombetti, A Talenti comparison result for solutions to elliptic problems with Robin boundary conditions, 2019, arXiv: 1909.11950.
    [6] A. Alvino, G. Trombetti, Sulle migliori costanti di maggiorazione per una classedi equazioni ellittiche degeneri, Ricerche Mat., 27 (1978), 413–428.
    [7] A. Alvino, G. Trombetti, A lower bound for the first eigenvalue of an elliptic operator, J. Math. Anal. Appl., 94 (1983), 328–337. http://dx.doi.org/10.1016/0022-247X(83)90066-5 doi: 10.1016/0022-247X(83)90066-5
    [8] A. Alvino, G. Trombetti, P. Lions, On optimization problems with prescribed rearrangements, Nonlinear Anal. Theor., 13 (1989), 185–220. http://dx.doi.org/10.1016/0362-546X(89)90043-6 doi: 10.1016/0362-546X(89)90043-6
    [9] C. Bandle, Isoperimetric inequalities and applications, Pitman Publishing, 1980.
    [10] J. Casado-Díaz, Smoothness properties for the optimal mixture of two isotropic materials: The compliance and eigenvalue problems, SIAM J. Control Optim., 53 (2015), 2319–2349. http://dx.doi.org/10.1137/140971087 doi: 10.1137/140971087
    [11] J. Casado-Díaz, Some smoothness results for the optimal design of a two-composite material which minimizes the energy, Calc. Var., 53 (2015), 649–673. http://dx.doi.org/10.1007/s00526-014-0762-5 doi: 10.1007/s00526-014-0762-5
    [12] J. Casado-Díaz, A characterization result for the existence of a two-phase material minimizing the first eigenvalue, Ann. Inst. H. Poincaré Anal. Non Linéaire, 34 (2016), 1215–1226. http://dx.doi.org/10.1016/j.anihpc.2016.09.006 doi: 10.1016/j.anihpc.2016.09.006
    [13] F. Caubet, T. Deheuvels, Y. Privat, Optimal location of resources for biased movement of species: the 1D case, SIAM J. Appl. Math., 77 (2017), 1876–1903. http://dx.doi.org/10.1137/17M1124255 doi: 10.1137/17M1124255
    [14] C. Conca, A. Laurain, R. Mahadevan, Minimization of the ground state for two phase conductors in low contrast regime, SIAM J. Appl. Math., 72 (2012), 1238–1259. http://dx.doi.org/10.1137/110847822 doi: 10.1137/110847822
    [15] C. Conca, R. Mahadevan, L. Sanz, An extremal eigenvalue problem for a two-phase conductor in a ball, Appl. Math. Optim., 60 (2009), 173–184. http://dx.doi.org/10.1007/s00245-008-9061-x doi: 10.1007/s00245-008-9061-x
    [16] M. Dambrine, D. Kateb, On the shape sensitivity of the first Dirichlet eigenvalue for two-phase problems, Appl. Math. Optim., 63 (2010), 45–74. http://dx.doi.org/10.1007/s00245-010-9111-z doi: 10.1007/s00245-010-9111-z
    [17] G. H. Hardy, Note on a theorem of Hilbert, Math. Z., 6 (1920), 314–317. http://dx.doi.org/10.1007/BF01199965 doi: 10.1007/BF01199965
    [18] A. Henrot, M. Pierre, Shape variation and optimization: a geometrical analysis, Switzerland: European Mathematical Society Publishing House, 2018.
    [19] B. Kawohl, Rearrangements and convexity of level sets in PDE, Berlin, Heidelberg: Springer, 1985. http://dx.doi.org/10.1007/BFb0075060
    [20] S. Kesavan, Symmetrization and applications, World Scientific, 2006. http://dx.doi.org/10.1142/6071
    [21] J. J. Langford, P. McDonald, Extremizing temperature functions of rods with Robin boundary conditions, 2021, arXiv: 2101.09600.
    [22] A. Laurain, Global minimizer of the ground state for two phase conductors in low contrast regime, ESAIM: COCV, 20 (2014), 362–388. http://dx.doi.org/10.1051/cocv/2013067 doi: 10.1051/cocv/2013067
    [23] E. Lieb, M. Loss, Analysis, Providence, Rhode Island: American Mathematical Society, 2001.
    [24] I. Mazari, Quantitative estimates for parabolic optimal control problems under l^\infty and l^1 constraints in the ball: Quantifying parabolic isoperimetric inequalities, Nonlinear Anal., 215 (2022), 112649. http://dx.doi.org/10.1016/j.na.2021.112649 doi: 10.1016/j.na.2021.112649
    [25] I. Mazari, G. Nadin, Y. Privat, Optimisation of the total population size for logistic diffusive equations: bang-bang property and fragmentation rate, Commun. Part. Diff. Eq., in press. http://dx.doi.org/10.1080/03605302.2021.2007533
    [26] I. Mazari, G. Nadin, Y. Privat, Shape optimization of a weighted two-phase Dirichlet eigenvalue, Arch. Rational Mech. Anal., 243 (2022), 95–137. http://dx.doi.org/10.1007/s00205-021-01726-4 doi: 10.1007/s00205-021-01726-4
    [27] L. Migliaccio, Sur une condition de Hardy, Littlewood, Polya, C. R. Acad. Sci. Paris Sér. I Math., 297 (1983), 25–28.
    [28] J. Mossino, Inégalités isopérimétriques et applications en physique, Paris: Hermann, 1984.
    [29] J. Mossino, J. M. Rakotoson, Isoperimetric inequalities in parabolic equations, Ann. Scuola Norm. Sci. Ser. 4, 13 (1986), 51–73.
    [30] F. Murat, Contre-exemples pour divers problèmes où le contrôle intervient dans les coefficients, Annali di Matematica, 112 (1977), 49–68. http://dx.doi.org/10.1007/BF02413475 doi: 10.1007/BF02413475
    [31] F. Murat, L. Tartar, Calculus of variations and homogenization, In: Topics in the mathematical modelling of composite materials, Boston, MA: Birkhäuser, 1997,139–173. http://dx.doi.org/10.1007/978-1-4612-2032-9_6
    [32] G. Nadin, The principal eigenvalue of a space–time periodic parabolic operator, Annali di Matematica, 188 (2009), 269–295. http://dx.doi.org/10.1007/s10231-008-0075-4 doi: 10.1007/s10231-008-0075-4
    [33] F. D. Pietra, C. Nitsch, R. Scala, C. Trombetti, An optimization problem in thermal insulation with Robin boundary conditions, Commun. Part. Diff. Eq., 46 (2021), 2288–2304. http://dx.doi.org/10.1080/03605302.2021.1931885 doi: 10.1080/03605302.2021.1931885
    [34] J. V. Ryff, Majorized functions and measures, Proc. Acad. Sci. Amsterdam, 71 (1968), 431–437.
    [35] J. Serrin, A symmetry problem in potential theory, Arch. Rational Mech. Anal., 43 (1971), 304–318. http://dx.doi.org/10.1007/BF00250468 doi: 10.1007/BF00250468
    [36] G. Talenti, Elliptic equations and rearrangements, Ann. Scuola Norm. Sci. Ser. 4, 3 (1976), 697–718.
    [37] G. Trombetti, J. L. Vazquez, A symmetrization result for elliptic equations with lower-order terms, Annales de la Faculté des sciences de Toulouse: Mathématiques Ser. 5, 7 (1985), 137–150.
    [38] J. L. Vazquez, Symétrisation pour u_t = {\Delta}\varphi(u) et applications, C. R. Acad. Sci. Paris Sér. I Math., 295 (1982), 71–74.
  • This article has been cited by:

    1. Idriss Mazari, A note on the rearrangement of functions in time and on the parabolic Talenti inequality, 2022, 68, 0430-3202, 137, 10.1007/s11565-022-00392-y
    2. Dario Mazzoleni, Benedetta Pellacci, Calculus of variations and nonlinear analysis: advances and applications, 2023, 5, 2640-3501, 1, 10.3934/mine.2023059
    3. Idriss Mazari-Fouquer, Optimising the carrying capacity in logistic diffusive models: Some qualitative results, 2024, 393, 00220396, 238, 10.1016/j.jde.2024.02.007
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2059) PDF downloads(186) Cited by(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog