• Aucun résultat trouvé

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

N/A
N/A
Protected

Academic year: 2022

Partager "Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php"

Copied!
25
0
0

Texte intégral

(1)

WHAT IS THE OPTIMAL SHAPE OF A FIN FOR ONE-DIMENSIONAL HEAT CONDUCTION?

GILLES MARCK, GR ´EGOIRE NADIN,AND YANNICK PRIVAT

Abstract. This article is concerned with the shape of small devices used to control the heat flowing between a solid and a fluid phase, usually called fins. The temperature along a fin in the stationary regime is modeled by a one-dimensional Sturm–Liouville equation whose coefficients strongly depend on its geometrical features. We are interested in the following issue: is there any optimal shape maximizing the heat flux at the inlet of the fin? Two relevant constraints are examined, by imposing either the volume or surface of the fin, and analytical nonexistence results are proved for both problems. Furthermore, using specific perturbations, we explicitly compute the optimal values and construct maximizing sequences. We show in particular that the optimal heat flux at the inlet is infinite in the first case and finite in the second one. Finally, we provide several extensions of these results for more general models of heat conduction, as well as several numerical illustrations.

Key words. heat conduction, calculus of variation, shape optimization, Sturm–Liouville equa- tion, volume constraint, surface constraint

AMS subject classifications.49J15, 49K15, 34E05 DOI.10.1137/130941377

1. Introduction. The increasing need for compact and efficient thermal systems leads to new challenges in the design of heat transfer devices. Across several engi- neering fields dealing with thermal management issues, a recurring problem concerns the optimal shape of several small elements, with the aim of locally increasing heat transfer. These systems, usually called fins, may adopt various shapes and designs, slightly extending the structure subjected to thermal loads.

For instance, fins are widely used to control the temperature of the heat exchang- ers located over computational processing units (CPU), since thermal overloads can have devastating effects on their physical integrity. Fins are also extensively used in in- dustrial manufacturing processes, mainly to evacuate the heat handled by the molded material, and must be designed to guarantee that the bulk temperature evolves inside a range specified by manufacturing constraints.

Therefore, efficient fin shapes are required to improve either cooling or warming processes, controlling the local temperature or increasing the heat transfer. Many engineering works have focused on modeling the direct problem in order to assess the efficiency of different fin shapes [3]. Nevertheless, very few theoretical results or mathematical proofs exist (see, for instance, [2, 10, 16]). The study proposed in the present article tackles this shape optimization problem, taking into account several relevant constraints on the admissible class of designs.

Engineering motivations behind fin optimization are rooted in a category of phys- ical problems related to efficient transport of a conservative flux. In particular, op- timizing the shape of a fin belongs to a larger class of problems aimed at reducing

Received by the editors October 16, 2013; accepted for publication (in revised form) April 14, 2014; published electronically August 14, 2014.

http://www.siam.org/journals/siap/74-4/94137.html

Universit´e Pierre et Marie Curie (Univ. Paris 6), CNRS UMR 7598, Laboratoire Jacques-Louis Lions, F-75005 Paris, France (gilles.marck@upmc.fr).

CNRS, Universit´e Pierre et Marie Curie (Univ. Paris 6), UMR 7598, Laboratoire Jacques-Louis Lions, F-75005 Paris, France (gregoire.nadin@upmc.fr, yannick.privat@upmc.fr). The third author was partially supported by ANR Project OPTIFORM.

1194

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(2)

x r

h,T Td

fluid flow

fin wall

hr,T dS

0

x φ(x) φ(x+dx)

φs(x) a(x+dx)

a(x)

0 x+dx

Ωa r

Fig. 1. Scheme of the axisymmetric fin.

thermal resistance during heat flows inside different media. Indeed, the main purpose of a fin is to thermally link a solid material subject to a heat flux with a fluid flow taking place around it. Its role consists of cooling down/warming up the solid phase by transferring heat to/from the fluid phase.

From the optimal design point of view, the main difficulties arise from the mul- tiple physical aspects of this problem, depicted in Figure 1. Indeed, the heat flowing from a solid domain through a fin to a fluid is only transported by conduction inside the structure, as far as its boundary. Then, according to the heat flux conservation property, it is fully transferred into the fluid phase. Hence, it is at the same time transported by the fluid motion and conducted between fluid elements, which is re- ferred to as aconducto-convection phenomenon (see [26]). As a consequence, the heat flowing from the fin to the fluid depends on the flow pattern around boundaries. This multiphysics optimal design problem with a Navier–Stokes/heat coupling system has been numerically investigated for some academic configurations in [22]. For the needs of simple and robust modeling, physical/engineering work often takes into consider- ation only the conduction phenomenon occurring inside the bulk material, using an average convective coefficient to stand for the heat transfer at the solid/fluid interface [3, 26]. This is the choice we make in what follows.

In this article, we will consider a fin subject to a steady-state thermal regime and attached to a device at constant temperature. A fin is considered as thermally efficient if it conveys the largest amount of heat while requiring the smallest volume of solid material. Two reasons motivate this claim: first, a fin is generally made of expensive components, because it requires high conductivity materials such as copper. Second, a fin is generally oriented orthogonally to the cross-flow direction, and we could expect that it is designed to generate the smallest possible perturbation in the fluid motion.

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(3)

Indeed, a smaller fin produces a smaller perturbation and requires less additional power to set the fluid into motion.

This feature leads us to the following optimal design problem: maximize the heat crossing the fin, by prescribing either its volume or its surface. From a mathematical point of view, the problems addressed within this article can be written as infinite- dimensional optimization problems subject to constraints of several natures: (i) global ones such as volume or surface, (ii) pointwise ones, and (iii) ordinary differential equation ones, since the cost functional depends on the solution of a Sturm–Liouville equation, whose coefficients are written as highly nonlinear functions of the unknown.

The idea of minimizing functionals depending on a Sturm–Liouville operator, such as eigenvalues or eigenfunctions, goes back at least to Krein in [18]; see also [14, 15, 20, 21] and [11] for a review on such problems. We also mention [12, 24], where two problems close to those addressed in this article have been solved in the context of mathematical biology (see Remark 4 for a comparison of these problems with ours).

The aim of the present paper is to investigate theoretically the issue of maximizing the heat flux at the inlet of a fin with respect to its shape, under surface or volume constraints. This analysis will be carried out using the simplified model (2.1) for the fin, for which we will prove that optimal shapes do not exist, but that one can construct a maximizing sequence in the volume constraint case and explicitly compute the optimal value of the heat flux. In the related work [4], the authors considered a simplified nonstationary model of fin, getting rid of the convective heat transfer from the lateral side of the structure (β = 0 in (2.1) below), and investigated a different criterion (the cooling rate of the fin), for which they completely solved the maximization problem using rearrangement techniques. We also mention [6, 7, 19, 23], in which more involved models are investigated exclusively in a numerical way.

The novelty of this article comes from not only the awareness of the convective heat transfer, making the optimal design problem harder to tackle, but also from the fact that we completely solve it theoretically. This is why the present study can be viewed, in some sense, as a first step toward understanding and deriving analytical results for more involved models.

This article is structured as follows: in section 2.1, a Sturm–Liouville model of temperature along the fin is derived using physical arguments. Section 2.2 is devoted to introducing the cost functionalF(a), which stands for the heat flux at the inlet of the fin, and providing continuity properties about it. The two main optimal design problems investigated are introduced in section 3: in a nutshell, the first consists of maximizing F(a) with a volume constraint, whereas the second aims to maximize F(a) with a lateral surface constraint. These problems are solved in section 3.2 and proved, respectively, in sections 3.3 and 3.4. More precisely, in each case we show a nonexistence result, provide the optimal value for this problem, and offer a way to construct maximizing sequences. Finally, all these results are extended to a more general setting in section 4. Several numerical illustrations of the results are also presented.

2. Modeling of an axisymmetric fin.

2.1. Sturm–Liouville model of the temperature conduction. Let us con- sider an axisymmetric fin Ωaof length >0 and radiusa(x) at abscissax, as displayed in Figure 1, defined in a Cartesian coordinate system by

Ωa ={(rcosθ, rsinθ, x)|r∈[0, a(x)), θ∈S1, x∈(0, )}.

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(4)

We will assume in what follows that the following hold:

(H1) a∈W1,∞(0, ), so that the surface element is defined almost everywhere.

(H2) There existsa0 >0 such that a(x)≥a0 for every x∈[0, ], so that the fin cannot collapse.

Figure 1 sums up the situation and the notation we will use throughout this article.

Some physical explanations about the derivation of the temperature model may be found in [3, 26]. Introduce the convective coefficient h > 0 modeling the heat transfer between the fin surface and the fluid flow, the convective coefficienthr>0 characterizing the heat transfer over the tip, andk >0,the thermal conductivity of the fin.

The inlet of the fin, as well as the fluid surrounding the fin, are assumed to be at a constant temperatures, denoted respectively byTd andT. Considering processes where the fin helps with cooling a thermal system, i.e., where the heat flows from its basis towards the fluid, we will assume that 0< T< Td.

Then, the temperatureT along the fin is solution of the following ordinary differ- ential equation (see [3, 26]):

(2.1)

(a2(x)T(x)) =βa(x)

1 +a(x)2(T(x)−T), x∈(0, ), T(0) =Td,

T() =−βr(T()−T),

where β = 2h/k is a positive real constant and βr = hr/k is a nonnegative real constant.

Remark 1. Let us comment on this model and the assumptions we have made on the physical parameterskandh.

From a rigorous point of view, the thermal conductivity k depends on the temperature T. However, if the operating temperature range in which the fin operates is small enough, it is reasonable to consider k as a constant parameter. Indeed, knowing thatT≤T(x)≤Td, as proved in Lemma 2.1, this hypothesis is valid if the variation of the thermal conductivity between k(T) and k(Td) is small. As a matter of fact, let’s consider a fin made of copper with working temperatures ranging from 0C to 100C: the order of magnitude of the copper thermal conductivity variation is about 2%. This assumption is widely used in the dedicated scientific/engineering literature [3, 19, 23, 26].

The same conclusion holds for the convective coefficienth, by assuming that it obeys Newton’s law, which means that it does not depend on the difference between T and the fin surface temperature. In the absence of radiative phenomena, this is a standard assumption for several fin models in the liter- ature [3, 7, 23, 26].

Since the function a is assumed to satisfy assumptions (H1) and (H2), system (2.1) has a unique solutionT ∈H1(0, ) by virtue of the Lax–Milgram lemma. The following lemma provides the monotonicity of this solution.

Lemma 2.1. Let a be a function satisfying assumptions (H1) and (H2). Then, the solutionT(·)of (2.1)is decreasing and satisfies

T≤T(x)≤Td

for everyx∈[0, ].

Proof. Let us use the standard change of variables for Sturm–Liouville problems (see, for example, [8]): y = x

0 dt

a(t)2. Setting 1 =

0 dt

a(t)2 and S(y) = T(x)−T

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(5)

for every x∈ [0, ], it follows from (2.1) that S is a solution of the boundary value problem

(2.2)

S(y) =βρ(y)S(y), y∈(0, 1), S(0) =Td−T,

S(1) =−βra()2S(1), where ρ(y) = a(x)3

1 +a(x)2 for every x [0, ]. Let us now prove that the function S(·) remains positive on [0, 1]. We denote byy0the largest zero of S(·) on [0, 1] whenever it exists. One has necessarily S(y0) = 0. Otherwise, the Cauchy–

Lipschitz theorem would immediately yield a contradiction. In particular, y0 =1. If S(y0) > 0, then S is positive on (y0, 1) since y0 is the largest zero of S, and thusS(·) is strictly convex and increasing on (y0, 1), which is incompatible with the boundary condition at 1. In the same way, if S(y0)< 0, S(·) is concave negative and decreasing on [y0, 1], and the same conclusion follows. As a result, the function S(·) is positive on [0, 1] and hence strictly convex according to (2.2). The function S(·) is thus decreasing according to the boundary condition at 1. The conclusion follows.

2.2. The cost functional F(a). The cost functionalF(a) stands for the heat flux at the inlet of the fin. It is defined by

(2.3) F(a) =−kπa(0)2T(0),

whereT(·) denotes the solution of (2.1). Integrating the main equation of (2.1) yields the new (integral) expression

(2.4) F(a) =kπβ

0 a(x)

1 +a(x)2(T(x)−T)dx+kπβra()2(T()−T). In the forthcoming analysis of the optimal design problems that we will investi- gate, we will need to use continuity properties of the cost functionalF. Let us write these properties precisely. Define the class of admissible designs,

Aa0,=

a∈W1,∞(0, ), a≥a0a.e. in (0, ) , and the product spaceAa0,, defined by

Aa0,=

(a, b), a∈ Aa0, andb=a

1 +a2 . Introduce the functionalFdefined on Aa0,by

(2.5) F(a, a

1 +a2) =F(a)

for every a ∈ Aa0,. Here and in what follows, the notation M(0, ) stands for the space of Radon measures on (0, ).

Definition 2.2. Let (an, bn)n∈Nbe a sequence of elements ofAa0,. We will say that (an, bn)n∈N τ-converges to(a, b)∈C0([0, ])× M(0, )if

(an)n∈N converges toa, locally uniformly in (0, ],

(bn)n∈Nconverges to b in the sense of measures.

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(6)

We endow Aa0, with the topology inherited from the τ-convergence. One has the following continuity result of the cost functionalF.

Proposition 2.3. Let (an, bn)n∈N be a sequence of elements of Aa0, which τ- converges to (a, b)∈C0([0, ])× M(0, ). Then, the sequence (F(an))n∈N converges toF(a, b)defined by

(2.6) F(a, b) =kπβb, T−TM(0,),C0([0,])+kπβra()2(T()−T),

whereT = ˜T+TdandT˜is the unique solution of the equation written under variational form: find T˜ H1(0, ) satisfying T˜(0) = 0 such that, for every test function ϕ H1(0, )satisfyingϕ(0) = 0, one has

0 a(x)2T˜(x)ϕ(x)dx+βb,( ˜T+Td−T)ϕM(0,),C0([0,]) +βra()2( ˜T() +Td−T)ϕ() = 0. (2.7)

Remark 2. It follows from Proposition 2.3 that the functionalF defined (with a slight abuse of notation) by (2.6) on the closure ofAa0,for the topology associated with theτ-convergence is a continuous extension of the functionF defined by (2.5).

Proof of Proposition 2.3. Consider a sequence (an, bn)n∈Nof elements ofAa0,as in the statement of Proposition 2.3. Denote by Tn the associated solution of (2.1).

Let us multiply the main equation of (2.1) by Tn−Tand then integrate by parts.

We obtain

a0min{β, a0}Tn−T2H1(0,)

0

an(x)2Tn(x)2+βbn(x)(Tn(x)−T)2 dx

=−βran()2(Tn()−T)2−an(0)2Tn(0)(Td−T). Integrating (2.1) on (0, ) yields

−βran()2(Tn()−T)−an(0)2Tn(0) =β

0 bn(x)(Tn(x)−T)dx, and, according to Lemma 2.1, it follows that

a0min{β, a0}Tn−T2H1(0,)≤βran2(Td−T)2+β(Td−T)2bn,1M(0,),C0([0,]). Since (an, bn)n∈Nτ-converges to (a, b), we deduce from the previous estimate that the sequence (Tn)n∈N is uniformly bounded inH1(0, ). Hence, using a Rellich theorem, there existsT ∈H1(0, ) such that, up to a subsequence, (Tn)n∈N converges toT, weakly inH1(0, ) and strongly in L2(0, ). Introduce for everyn Nthe function T˜n:=Tn−Td. The function ˜Tnis the unique solution of the system whose variational form can be written as: find ˜Tn ∈H1(0, ) such that ˜Tn(0) = 0,and for every test function ϕ∈H1(0, ) satisfyingϕ(0) = 0, one has

0

an(x)2T˜n(x)ϕ(x) +βbn(x)( ˜Tn(x) +Td−T)ϕ(x)

dx

=−βran()2( ˜Tn() +Td−T)ϕ().

The conclusion follows easily, passing to the limit into this variational formulation and noting that

F(an, bn)

=−an(0)2Tn(0) =β

0 bn(x)(Tn(x)−T)dx+βran()2(Tn()−T),

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(7)

since the injectionH1(0, )→C0([0, ]) is compact.

3. Optimal design problems for a simplified model of a fin.

3.1. The optimal design problems. In this section we will introduce and solve problems modeling the optimal shape of a fin for the Sturm–Liouville model (2.1). We will consider two kinds of constraints: a volume constraint and a lateral surface constraint. Let us define the volume and lateral surface functionals by

vol(a) =

0 a(x)2dx and surf(a) =

0 a(x)

1 +a(x)2dx.

We investigate the problem of maximizing the functionala→F(a) defined by (2.3) with a volume or a lateral surface constraint.1

Optimal design problem with volume or lateral surface constraint.

Let a0 and denote two positive real numbers. Fix V0 > a20 and S0 > a0. The optimal design problem with volume constraint is

(3.1) sup{F(a), a∈ Va0,,V0}, where

Va0,,V0 =

a∈W1,∞(0, ), a≥a0 a.e. in [0, ] and πvol(a)≤πV0 , and the optimal design problem with lateral surface constraint is

(3.2) sup{F(a), a∈ Sa0,,S0}, where

Sa0,,S0 =

a∈W1,∞(0, ), a≥a0 a.e. in [0, ] and 2πsurf(a)2πS0 . Remark 3. Physically, the lateral surface constraint can be justified by considering the fluid flowing around the fin. In realistic engineering configurations, the fluid is put into motion by external equipment, such as a pump or a fan. The pressure and kinetic energies provided by this system are dissipated all through the fluid flow, mainly because of the wall shear stress. In other words, one part of the dissipation is due to the fluid friction against the walls. As a consequence, limiting the surface area of the fin is a suitable way to reduce the operating cost of the whole thermal system.

In addition, limiting the fin surface may also help to reach shapes that are easier and cheaper to manufacture.

Remark 4. As underlined in the introduction, the optimal design problems ad- dressed here look similar to those addressed in [12, 24], which were devoted to the issue of understanding nerve fiber shapes by solving an optimization problem. In these works, a Sturm–Liouville operator whose coefficients depended on the shape of the nerve fiber was also introduced, and two criteria were considered: a kind of transfer function and the first eigenvalue of a self-adjoint operator. However, the techniques implemented in the present work are rather different and appear a bit more sophisti- cated. Indeed, the ideas of the proofs in [12, 24] were all based on the standard change of variables for Sturm–Liouville problems used in Lemma 2.1, which permitted one to consider auxiliary problems and thus get a lower bound for the optimal values.

1Recall that volume of Ωa=π

0a(x)2dx, and lateral surface of Ωa= 2π 0a(x)

1 +a(x)2dx.

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(8)

This trick was then used to construct minimizing sequences. Unfortunately, in the present work, we did not manage to adapt such techniques, and the arguments of the proofs rest upon the use of particular perturbations that are constructed in the spirit of rearrangement techniques (see Remark 5 and the proof of Theorem 3.3).

Before solving these shape optimization problems, let us make more precise the topological nature of the classesVa0,,V0andSa0,,S0inL(0, ). As will be highlighted in the proofs of Theorems 3.2 and 3.3, the classesVa0,,V0 andSa0,,S0 are not closed for the standard strong topology ofW1,∞(0, ). The following lemma investigates the L-boundedness of the elements of these classes.

Lemma 3.1. Let a0anddenote two positive real numbers, and letV0> a20 and S0> a0. Then,

the class Va0,,V0 is not a bounded set ofL(0, ),

the class Sa0,,S0 is a bounded set of L(0, ), and for everya∈ Sa0,,S0

a0≤a(x)

S02/2+ 4S0.

Proof. First consider the sequence of functions (an)n≥1defined by an(x) = n−2(Vn(n−a0−a220)

0)x ifx∈[0,2(V0−a20)/n], a0 ifx∈(2(V0−a20)/n, ]. By construction,

0 an(x)2dx=a20+(n−a20)(V0−a20) n < V0,

so that an ∈ Va0,,V0 for every n N. Since an C0([0, ]) and an(0) = n, it follows obviously thatVa0,,V0 is not bounded inL(0, ).

Second, considera∈ Sa0,,S0. One has S0

y

x

a(t)

1 +a(t)2dt≥ y

x

a(t)|a(t)|dt≥ y

x

a(t)a(t)dt =1

2|a(y)2−a(x)2| for every 0< x < y < . It follows that, for everyx∈[0, ],

(3.3) a(0)22S0≤a(x)2≤a(0)2+ 2S0. Using this inequality, one gets

S0

0 a(t)dt≥

a(0)22S0,

and thusa(0)2 ≤S02/2+ 2S0. The conclusion follows by combining this inequality with (3.3).

3.2. Main results: MaximizingF with a volume or a surface constraint.

In this section, we solve problems (3.1) and (3.2), which, according to the previous sections, model the optimal shape of a axisymmetric fin with either a volume or a surface constraint.

The following theorem highlights the ill-posed character of problem (3.1), where a volume constraint is imposed.

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(9)

Theorem 3.2. Leta0 and denote two positive real numbers, and letV0> a20. Problem (3.1)has no solution and

sup

a∈Va0,,V0

F(a) = +∞.

In the next theorem, we prove that imposing a lateral surface constraint on Ωa is not enough to get the existence of a solution for this problem. Nevertheless, in contrast to the previous case where a volume constraint was imposed, the value of the supremum ofF over the set Sa0,,S0 is now finite.

Theorem 3.3. Leta0 and denote two positive real numbers, and letS0> a0. Problem (3.2)has no solution, and

sup

a∈Sa0,,S0

F(a) =kπβ(Td−T)

a√3/20 γ

β + (S−a0)

, whereγ denotes the positive real number

(3.4) γ=

β a0sinh

β a0

+βrcosh β

a0 β

a0cosh β

a0

+βrsinh β

a0 .

Moreover, every sequence(an)n∈Nof elements ofSa0,,S0such that(an, an

1 +a2n)n∈N τ-converges to (a0, a0+ (S−a0)δ0), where δ0 denotes the Dirac measure at x= 0, is a maximizing sequence for problem (3.2).

Remark 5. In the proof of Theorem 3.3, the key idea for showing the nonexis- tence result and constructing maximizing sequences lies in applying a particular per- turbation of the functiona√

1 +a2, constructed in the vein of monotone or Schwarz rearrangements (but not itself a monotone or Schwarz rearrangement). We then ex- hibit an element a ∈ Aa0, belonging to the class of admissible designs, that makes the criterion increase.

Notice that it is standard in shape optimization to use rearrangements, such as the so-calledSteiner orSchwarz symmetrizations, to characterize the solutions of an optimal design problem (see, for instance, [17] for the definition and [11] for examples of use in shape optimization). Unfortunately, we did not manage to conclude the proof by applying standard rearrangement arguments, which explains why a more subtle analysis is performed to get the nonexistence result.

3.3. Proof of Theorem 3.2 (maximization of F with a volume con- straint). To prove this theorem, we will exhibit an explicit maximizing sequence (an)n∈N. The construction of the sequence (an)n∈N is based on the use of an inter- mediate sequence denoted (aS,m)m∈N, whereS denotes a positive real number. Fix S >0. The sequence (aS,m)m∈N is chosen to satisfy at the same time the following:

(aS,m)m∈N converges strongly toa0in L(0, ),

the sequence (bS,m)m∈N defined bybS,m=aS,m

1 +a2S,mfor everym∈N satisfies

surf(aS,m) =

0 bS,m(x)dx=S

for everym∈N and converges in the sense of measures toa0+ (S−a0)δ0, whereδ0 denotes the Dirac measure atx= 0.

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(10)

x= 0 m1 x= a0

Mm

x= 0 x=

a0

1 m

Fig. 2.Left: Graph ofbS,m=aS,m

1 +a2S,m. Right: Graph ofaS,m.

In particular, the sequence (aS,m, aS,m

1 +a2S,m)m∈N τ-converges to (a0, a0+ (S− a0)δ0) asmtends to +.

Let us now provide an example of such a sequence. A possible choice of function bS,mfor mlarge enough is given by

bS,m(x) =

a0+ (S−a0)m ifx∈[0,1/m], a0 ifx∈(1/m, ]. Solving the equation bS,m = aS,m

1 +a2S,m can be rewritten as aS,m|aS,m|

b2S,m−a2S,m = 1. This equation obviously has an infinite number of solutions. In particular, a family of solutions is obtained by making the functionaS,m oscillate an integer number of times in (0,1/m), each oscillation corresponding to two successive choices of the sign of aS,m on two intervals of the same length. The continuity of each functionaS,m

then determines that function in a unique way. As a consequence, it is possible to control the L-norm of aS,m by artfully choosing the number of oscillations, as we make precise below.

Introduce Mm = a0 + (S−a0)m for every m N. We now construct the sequence (aS,m)n∈N so that each function oscillatesm times on [0,1/m] (see Figure 2). More precisely, we defineaS,mby

(3.5)

aS,m(x) =

⎧⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

Mm2 (

Mm2 −a20−x)2 on 0,2m12

, aS,m(m12−x) on 1

2m2,m12

, aS,m

x−mi2

on i

m2,i+1m2

, i∈ {1, . . . , m−1},

a0 on 1

m, . Notice in particular that

aS,m−a0=aS,m

1 2m2

−a0= o 1

m2

as m→+∞.

The sequence (aS,m)n∈N then satisfies the expected properties above. Moreover, according to Proposition 2.3, one has

m→+∞lim F(aS,m) = lim

m→+∞F

aS,m, aS,m

1 +a2S,m

=F(a0, a0+ (S−a0)δ0).

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(11)

Notice that the solution T of (2.7) associated with the pair (a, b) = (a0, a0+ (S− a0)δ0) coincides with the solution of the same equation associated with the pair (a, b) = (a0, a0). Explicit computations thus lead to

T(x) =T+ (Td−T)

cosh β

a0x

−γsinh β

a0x

for everyx∈[0, ], whereγ is defined by (3.4). We then compute

m→+∞lim

F(aS,m)

=βa0+ (S−a0)δ0, T −TM(0,),C0([0,])+βra20(T()−T)

=β(Td−T)

a√3/20

β γ+ (S−a0)

.

Conclusion: Construction of a maximizing sequence (an)n∈N. Letn N such that n≥[a0] + 1. Consider the sequence (an,m)m∈N introduced previously. Since (an,m)m∈N converges strongly to a0 in L(0, ) and according to the previous con- vergence study, there existsp∈Nsuch that vol(an,p)≤V01n and

F(an,p) kπβ(Td−T)

a√3/20 γ

β + (n−a0)

1 n.

Let us denote by mn the first integer for which this property is verified, and byan

the function equal toa0 ifn≤[a0] and byan,mn otherwise. It follows not only that each elementan belongs toVa0,,V0, but also that

sup

a∈Va0,,V0

F(a)

kπβ F(an,mn)

kπβ (Td−T)

a√3/20 γ

β + (n−a0) 1 n

for everyn≥[a0] + 1. Lettingntend to +yields the conclusion of the theorem.

3.4. Proof of Theorem 3.3 (maximization of F with a lateral surface constraint). This proof is divided into several steps. Steps 1, 2, and 3 are devoted to showing that problem (3.2) has no solution, whereas Step 4 focuses on building maximizing sequences by introducing a kind of truncated optimal design problem. Let us argue by contradiction, assuming that problem (3.2) has a solution a∈ Sa0,,S0. Let us denote by b the function a√

1 +a2. The idea of the proof is to introduce an admissible perturbation aε of a in Sa0,,S0, suitably chosen to guarantee that F(aε)> F(a) for a givenε >0. We will proceed in several steps.

Step 1. Definition of the perturbationaε. Since the constant function equal toa0

is obviously not a solution of problem (3.2), there exists x0 (0, ) and an interval (x0−ε/2, x0+ε/2) on which a > a0. Introduce for such a choice ofx0 and ε the function

bε=b+c(χ[0,ε]−χ[x0−ε/2,x0+ε+2]),

wherecis a positive constant, chosen small enough to guarantee thatbε> a0almost everywhere in (0, ). We will use the following lemma to constructaε.

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(12)

x=ξ

x= ¯x x= ¯x+η

y=aη,1(x)

y=aη,2(x)

y=aε(x)

y=a(x)

x= 0 ε x0 x=

a0

y=a(x)

y=aε(x)

Fig. 3.Left: Zoom on one oscillation. Right: The perturbationaε.

Lemma 3.4. There exists a family(aε)ε>0 such thataε∈ Sa0,,S0 for everyε >0, aε

1 +a2ε =bε for almost everyε >0, and

aε−aL(0,)= O(ε2).

Proof. The proof consists of exhibiting an element aε ∈ Sa0,,S0 so that the statements of Lemma 3.4 are satisfied. First, we imposeaε =a on (ε, x0−ε/2) (x0+ε/2, ). Without loss of generality, let us now explain how to defineaε on [0, ε], the construction ofaεon [x0−ε/2, x0+ε/2] being similar. The construction method is close to that of the maximizing sequence presented in the proof of Theorem 3.2.

The idea is to impose oscillations onaεto control theLdistance betweenaandaε. We first explain how to create one oscillation on an interval [¯x,x¯ +η], where 0<x <¯ x¯+η < ε(see Figure 3). The functionaεis chosen so that

aε=aη,1 on (¯x, ξ) and aε=aη,2 on (ξ,x¯+η), where the functionaη,1 is a solution of the following Cauchy problem,

aη,1(x) =

bε(x)2−aη,1(x)2

aη,1(x) , x∈x,x¯+η), aη,1x) =ax);

the functionaη,2 is solution of the following Cauchy problem, aη,2(x) =−√

bε(x)2−aη,2(x)2

aη,2(x) , x∈x,x¯+η), aη,2x+η) =ax+η);

andξ∈x,x¯+η) is chosen so thataη,1(ξ) =aη,2(ξ). Notice that the functions aη,1

andaη,2 satisfy in particularaη,i

1 +a2η,i =bεfori∈ {1,2},aη,1 is increasing, and aη,2 is decreasing. Such a construction is possible, provided that the graphs of aη,1

and aη,2 intersect at a point whose abscissa belongs to (¯x,x¯+η). The intermediate value theorem yields that it is enough to show that

aη,1x+η)> ax+η) and aη,2x)> ax).

Without loss of generality, let us prove that the first assertion is true, the proof of the second one being similar. In fact, we will prove that aη,1 > a everywhere in

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(13)

x,x¯+η]. Let us argue by contradiction. Note thataη,1x)> ax) since ¯x < εand thusbεx)> bx). Denote byαthe first point of (¯x,x¯+η) such thataη,1(α) =a(α), provided that it exists. Sincebε> bon [¯x,x¯+η][0, ε], there holdsaη,1

1 +a2η,1>

a√

1 +a2 almost everywhere in any neighborhood ofα; thus there exists necessarily a neighborhoodOα ofαsuch that|aη,1(x)|>|a(x)| at every Lebesgue point ofaη,1 andainOα. It implies the existence ofν >0 such thataη,1(x)< a(x) on (α−ν, α), which is absurd.

It remains now to find the number of oscillation on [0, ε] and [x0−ε/2, x0+ε/2]

that guarantees thataεis as close to aas desired (in the sense of theLdistance).

Notice that, according to the previous definition ofaεon [¯x,x¯+η], one has a(x)2≤aε(x)22

x+η¯

¯ x

bε(x) +ax)2

2η(b+cL(0,)+ 2aL(0,)aL(0,)) +a(x)2 for everyx∈x,¯x+η]. Moreover, since a∈W1,∞(0, ),

0≤aε(x)−a(x)(b+cL(0,)+aL(0,)aL(0,))

a0 η

for everyx∈x,x¯+η]. It thus suffices to fix, for example,η=ε/kεwithkε= [1]+1, to ensure thataε−aL([¯x,¯x+η])= O(ε2).

Conclusion. We consider a regular subdivision of [0, ε] (resp., [x0−ε/2, x0+ε/2]) intokε intervals, and we use the process described previously to construct the graph ofaε by creating one oscillation on each of these intervals. This way, one getsaε aL([0,])= O(ε2). The lemma is then proved.

We now consider a family (aε)ε>0 chosen as in the statement of Lemma 3.4. In particular, and according to the construction ofaε made in the proof of Lemma 3.4, we will assume without loss of generality thataε() =a().

Forε >0, we denote byTεthe temperature associated with aε, i.e., the solution of (2.1) witha=aε.

Step2. Asymptotic development ofTεat the first order. Noticing that

0bε(x)dx=

0b(x)dx, (bε)ε>0 converges in the sense of measures to b, and (aε)ε>0 converges strongly to a in L(0, ) as ε 0, according to Lemma 3.4. Then, following the proof of Proposition 2.3, the family (Tε)ε>0 converges, up to a subsequence, weakly in H1(0, ) and strongly in L2(0, ), to T, the solution of (2.1) associated with the optimal radiusa.

Let us write an asymptotic development ofTεat the first order. For that purpose, introduce

Tε=Tε−T ε .

The functionTεis the solution of the following ordinary differential equation:

(3.6)

(a(x)2Tε(x)) =βb(x)Tε+βRεε(x)(Tε(x)−T)

aε(x)2−a(x)2 ε

Tε(x)

, x∈(0, ),

Tε(0) = 0,

Tε() =−βrTε(),

whereRε(x) =c(χ[0,ε]−χ[x0−ε/2,x0+ε+2]).

Downloaded 12/14/14 to 130.159.70.209. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

Références

Documents relatifs

Shu , Stability analysis and a priori error estimates to the third order explicit Runge-Kutta discontinuous Galerkin method for scalar conservation laws, SIAM J. Redistribution

Finally, we apply some spectral properties of M -tensors to study the positive definiteness of a class of multivariate forms associated with Z -tensors.. We propose an algorithm

In this paper we study the structure of finite dimensional estimation algebras with state dimension 3 and rank 2 arising from a nonlinear filtering system by using the theories of

In section 3, we use the sticky particle method to study weak solutions to the mCH equation for general Radon measure initial data.. Space-time BV estimates

Then, making use of this fractional calculus, we introduce the generalized definition of Caputo derivatives. The new definition is consistent with various definitions in the

Based on iteration (3), a decentralized algorithm is derived for the basis pursuit problem with distributed data to recover a sparse signal in section 3.. The algorithm

In the first set of testing, we present several numerical tests to show the sharpness of these upper bounds, and in the second set, we integrate our upper bound estimates into

For the direct adaptation of the alternating direction method of multipliers, we show that if the penalty parameter is chosen sufficiently large and the se- quence generated has a