This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the regularity of optimal potentials in control problems governed by elliptic equations

Abstract

In this paper we consider optimal control problems where the control variable is a potential and the state equation is an elliptic partial differential equation of a Schrödinger type, governed by the Laplace operator. The cost functional involves the solution of the state equation and a penalization term for the control variable. While the existence of an optimal solution simply follows by the direct methods of the calculus of variations, the regularity of the optimal potential is a difficult question and under the general assumptions we consider, no better regularity than the BVBV one can be expected. This happens in particular for the cases in which a bang-bang solution occurs, where optimal potentials are characteristic functions of a domain. We prove the BVBV regularity of optimal solutions through a regularity result for PDEs. Some numerical simulations show the behavior of optimal potentials in some particular cases.

G. Buttazzo,   J. Casado-Díaz††,   F. Maestre††

Dipartimento di Matematica, Università di Pisa,

Largo B. Pontecorvo, 5

56127 Pisa, ITALY

†† Dpto. de Ecuaciones Diferenciales y Análisis Numérico,

Facultad de Matemáticas, C. Tarfia s/n

41012 Sevilla, SPAIN

Keywords: optimal potentials, BV regularity, bang-bang property, shape optimization, control problems.

2020 Mathematics Subject Classification: 49Q10, 49J45, 35B65, 35R05, 49K20.

1 Introduction

The starting point of our research is an optimal control problem of the form

minΩ[j(x,u)+ψ(m)]𝑑x,\min\int_{\Omega}\big{[}j(x,u)+\psi(m)\big{]}\,dx, (1.1)

governed by the state equation

{Δu+mu=fin ΩuH01(Ω).\begin{cases}-\Delta u+mu=f\quad\text{in }\Omega\\ u\in H^{1}_{0}(\Omega).\end{cases} (1.2)

Here Ω\Omega is a bounded open subset of N\mathbb{R}^{N}, the control variable mm is assumed to be nonnegative, fL2(Ω)f\in L^{2}(\Omega), and j,ψj,\psi are suitable integrands. We assume that ψ\psi has a superlinear growth, which automatically implies that the control variables are in L1(Ω)L^{1}(\Omega). Notice that, when j(x,u)=f(x)uj(x,u)=f(x)u, the problem can be written in the variational form

min{2(m)+Ψ(m):mL1(Ω),m0},\min\Big{\{}-2\mathscr{E}(m)+\Psi(m)\ :\ m\in L^{1}(\Omega),\ m\geq 0\Big{\}},

where

Ψ(m)=Ωψ(m)𝑑x(m)=min{Ω[12|u|2+12mu2f(x)u]𝑑x:uH01(Ω)}.\begin{split}&\Psi(m)=\int_{\Omega}\psi(m)\,dx\\ &\mathscr{E}(m)=\min\left\{\int_{\Omega}\Big{[}\frac{1}{2}|\nabla u|^{2}+\frac{1}{2}mu^{2}-f(x)u\Big{]}\,dx\ :\ u\in H^{1}_{0}(\Omega)\right\}.\end{split}

In this case it is possible to see that the control variable mm can be eliminated, obtaining the auxiliary variational problem

min{Ω[|u|2+ψ(u2)2f(x)u]𝑑x:uH01(Ω)},\min\left\{\int_{\Omega}\Big{[}|\nabla u|^{2}+\psi^{*}(u^{2})-2f(x)u\Big{]}\,dx\ :\ u\in H^{1}_{0}(\Omega)\right\}, (1.3)

where ψ\psi^{*} denotes the Fenchel-Moreau conjugate of the function ψ\psi. Setting g(s)=s(ψ)(s2)g(s)=s(\psi^{*})^{\prime}(s^{2}), the unique solution u^\hat{u} of (1.3) can be obtained through the PDE

{Δu+g(u)=fin ΩuH01(Ω),\begin{cases}-\Delta u+g(u)=f\quad\text{in }\Omega\\ u\in H^{1}_{0}(\Omega),\end{cases} (1.4)

and the optimal control m^\hat{m} can be then recovered as

m^=(ψ)(u^2).\hat{m}=(\psi^{*})^{\prime}(\hat{u}^{2}).

For general integrands j(x,u)j(x,u) the elimination procedure above is not possible, and the necessary conditions of optimality involve an adjoint state variable and the corresponding adjoint PDE. Nevertheless, we can show that the optimal control problem (1.1), (1.2) admits a solution (u^,m^)(\hat{u},\hat{m}). The main goal of the present paper is to show that, under suitable assumptions on the data, the optimal control m^\hat{m} has additional regularity properties. In particular, we show that m^BV(Ω)\hat{m}\in BV(\Omega).

We stress that, under the general assumptions we consider on the function ψ\psi, higher regularity properties on m^\hat{m} do not hold. Indeed, when

ψ(s)={sif s[α,β](with 0α<β)+otherwise,\psi(s)=\begin{cases}s&\text{if }s\in[\alpha,\beta]\qquad\text{(with $0\leq\alpha<\beta$)}\\ +\infty&\text{otherwise,}\end{cases}

the optimal control m^\hat{m} is of bang-bang type, that is

m^=α+(βα)1E\hat{m}=\alpha+(\beta-\alpha)1_{E}

for a suitable set EE, which then turns out to be a set with finite perimeter.

Problems of this form arise for instance in some biological models, governed by logistic diffusive equations, where one aims to control the size of a total population or the optimal location of resources, see for instance [20] and [21].

The proof of the BVBV regularity above is obtained through a careful analysis of a nonlinear elliptic PDE of the form (1.4). In general, for the cases arising from optimal control problems, the right-hand side ff has a low summability and does not belong to the space H1(Ω)H^{-1}(\Omega); the definition of solution is then more involved and has to be given as in the theory of renormalized solutions (see for instance [3], [17]).

In Section 2 we list the notation that is used along the paper, in Section 3 we study the semilinear problems (1.4) with the nonlinearity gg possibly discontinuous and the right-hand side ff having a low summability. Our goal is to show that, when fBV(Ω)f\in BV(\Omega), the solution uu is such that g(u)BV(Ω)g(u)\in BV(\Omega). Section 4 deals with the application of the result above to the optimal control problem (1.1), (1.2). In Section 5 we consider some relevant examples with various particular choices of the data. Finally in Section 6 we provide some numerical simulations which show the bang-bang behavior of optimal solutions in some cases, as well as the continuous behavior in some other ones.

2 Notation

In this section we fix the notation that we use in the rest of the paper.

  • For ss\in\mathbb{R}, we denote by sgn(s)\operatorname{sgn}(s) the sign of ss.

  • For a measurable set ANA\subset\mathbb{R}^{N}, we denote by |A||A| its Lebesgue measure.

  • For a function g:g:\mathbb{R}\to\mathbb{R}, and ss\in\mathbb{R}, we denote by g(s)g_{-}(s), and g+(s)g_{+}(s) the limits on the left and on the right in ss, respectively (if they exist).

  • For k>0k>0 we define the functions Tk,SkW1,()T_{k},S_{k}\in W^{1,\infty}(\mathbb{R}) by

    Tk(s)={sif |s|<kksgn(s)if |s|k,Sk(s)={1if |s|k2|s|kif k<|s|<2k0if |s|2k.T_{k}(s)=\begin{cases}s&\hbox{if }|s|<k\\ k\,\operatorname{sgn}(s)&\hbox{if }|s|\geq k,\end{cases}\qquad S_{k}(s)=\begin{cases}1&\hbox{if }|s|\leq k\\ \displaystyle 2-{|s|\over k}&\hbox{if }k<|s|<2k\\ 0&\hbox{if }|s|\geq 2k.\end{cases}
  • We denote by (Ω){\cal M}(\Omega) the space of Borel measures in Ω\Omega with finite total mass.

  • If ψ:{}\psi:\mathbb{R}\to\mathbb{R}\cup\{\infty\} is a lower semicontinuous proper convex function, we denote by dom(ψ)\operatorname{dom}(\psi) the domain of ψ\psi, defined by

    dom(ψ)={s:ψ(s)<}.\operatorname{dom}(\psi)=\big{\{}s\in\mathbb{R}\ :\ \psi(s)<\infty\big{\}}.

    It is a non-empty interval of \mathbb{R}, which we assume not reduced to a point. For sdom(ψ)s\in\operatorname{dom}(\psi), we denote by ψ(s)\partial\psi(s) the subdifferential of ψ\psi at ss, defined by

    ψ(s)={τ:τ(ts)ψ(t)ψ(s) for every tdom(ψ)}.\partial\psi(s)=\big{\{}\tau\in\mathbb{R}\ :\ \tau(t-s)\leq\psi(t)-\psi(s)\ \text{ for every }t\in\operatorname{dom}(\psi)\big{\}}. (2.1)

    Taking into account that the map

    (s,t)ψ(t)ψ(s)tsfor s<t(s,t)\mapsto{\psi(t)-\psi(s)\over t-s}\qquad\text{for }s<t (2.2)

    is non-decreasing in both variables ss and tt, we have that for every sdom(ψ)s\in\operatorname{dom}(\psi) the two limits

    {d+ψ(s):=limε0ψ(s+ε)ψ(s)ε(,]dψ(s):=limε0ψ(s)ψ(sε)ε[,)\begin{cases}\displaystyle d_{+}\psi(s):=\lim_{\varepsilon\searrow 0}{\psi(s+\varepsilon)-\psi(s)\over\varepsilon}\in(-\infty,\infty]\\ \displaystyle d_{-}\psi(s):=\lim_{\varepsilon\searrow 0}{\psi(s)-\psi(s-\varepsilon)\over\varepsilon}\in[-\infty,\infty)\end{cases}

    exist and

    ψ(s)=[dψ(s),d+ψ(s)].\partial\psi(s)=\big{[}d_{-}\psi(s),d_{+}\psi(s)\big{]}. (2.3)

    We also recall that every convex function ψ\psi is locally Lipschitz in the interior of its domain.

  • We denote by ψ\psi^{*} the Fenchel-Moreau conjugate of ψ\psi, defined by

    ψ(t):=supsdom(ψ){tsψ(s)}t.\psi^{*}(t):=\sup_{s\in\operatorname{dom}(\psi)}\big{\{}ts-\psi(s)\big{\}}\qquad\forall\,t\in\mathbb{R}.

    We recall that ψ\psi^{*} is also a convex lower semicontinuous function and that

    (ψ)=ψ.(\psi^{*})^{*}=\psi.

    Moreover, we have

    tψ(s)sψ(t)ts=ψ(s)+ψ(t).t\in\partial\psi(s)\iff s\in\partial\psi^{*}(t)\iff ts=\psi(s)+\psi^{*}(t). (2.4)

In the following we consider the integral functionals

J(u)=Ωj(x,u)𝑑xdefined for uH01(Ω);Ψ(m)=Ωψ(m)𝑑xdefined for mL1(Ω),m0;F(u,m)=J(u)+Ψ(m).\begin{split}&J(u)=\int_{\Omega}j(x,u)\,dx\qquad\text{defined for }u\in H^{1}_{0}(\Omega);\\ &\Psi(m)=\int_{\Omega}\psi(m)\,dx\qquad\text{defined for }m\in L^{1}(\Omega),\ m\geq 0;\\ &F(u,m)=J(u)+\Psi(m).\end{split}

The function ψ:[0,){}\psi:[0,\infty)\to\mathbb{R}\cup\{\infty\} is assumed lower semicontinuous proper and convex, and such that

limsψ(s)s=;\lim_{s\to\infty}\frac{\psi(s)}{s}=\infty; (2.5)

in this way the functional Ψ\Psi is well defined on L1(Ω)L^{1}(\Omega) and Ψ(m)<\Psi(m)<\infty implies that mL1(Ω)m\in L^{1}(\Omega). It is easy to see that, with the conditions above, the function ψ\psi is bounded from below; hence up to the addition of a constant, that does not modify our optimization problem, we may assume ψ\psi non-negative.

Concerning the integrand j(x,s)j(x,s) we assume measurability in xx, lower semicontinuity in ss and the bound

a(x)c|s|2j(x,s)a(x)-c|s|^{2}\leq j(x,s) (2.6)

for suitable c0c\geq 0 and aL1(Ω)a\in L^{1}(\Omega).

The optimization problem we consider is then

min{F(u,m):Δu+mu=f,uH01(Ω),mL1(Ω)},\min\big{\{}F(u,m)\ :\ -\Delta u+mu=f,\ u\in H^{1}_{0}(\Omega),\ m\in L^{1}(\Omega)\big{\}},

where fH1(Ω)f\in H^{-1}(\Omega) is prescribed. In Section 4 we show that the optimal control problem above admits an optimal pair (u^,m^)(\hat{u},\hat{m}). Under some additional assumptions, we obtain the corresponding necessary conditions of optimality and we study the regularity properties of (u^,m^)(\hat{u},\hat{m}).

3 Semilinear problems with a discontinuous term

In the present section we are interested in the existence and regularity properties of the solutions of a semi-linear problem of the form

{Δu+g(u)=fin Ωu=0on Ω,\begin{cases}-\Delta u+g(u)=f&\text{in }\Omega\\ u=0&\text{on }\partial\Omega,\end{cases} (3.1)

where the function gg is non-decreasing and not necessarily continuous. Moreover, we assume only low integrability on the right-hand side ff. In particular, we do not assume fH1(Ω)f\in H^{-1}(\Omega), which leads us to work with renormalized or entropy solutions (see for instance [3], [4], [16]). In this sense, we introduce the following definition of solution for problem (3.1).

Definition 3.1.

Assume fL1(Ω)f\in L^{1}(\Omega). We say that a pair (u,w)(u,w) is a solution of (3.1) if it satisfies

{uH01(Ω)if N=1uW01,p(Ω)p<NN1,Tk(u)H01(Ω)k>0if N2,\displaystyle\begin{cases}u\in H^{1}_{0}(\Omega)&\hbox{if }N=1\\ \displaystyle u\in W^{1,p}_{0}(\Omega)\ \forall\,p<{N\over N-1},\quad T_{k}(u)\in H^{1}_{0}(\Omega)\ \forall\,k>0&\hbox{if }N\geq 2,\end{cases} (3.2)
wL1(Ω),g(u)wg+(u) a.e. in Ω,\displaystyle w\in L^{1}(\Omega),\quad g_{-}(u)\leq w\leq g_{+}(u)\ \hbox{ a.e. in }\Omega, (3.3)
{Ωuvdx+Ωwv𝑑x=Ωfv𝑑xvH01(Ω)L(Ω) such that k>0 with v=0 a.e. in {|u|>k}.\displaystyle\begin{cases}\displaystyle\int_{\Omega}\nabla u\cdot\nabla v\,dx+\int_{\Omega}wv\,dx=\int_{\Omega}fv\,dx\\ \forall\,v\in H^{1}_{0}(\Omega)\cap L^{\infty}(\Omega)\hbox{ such that }\exists k>0\ \hbox{ with }\nabla v=0\hbox{ a.e. in }\{|u|>k\}.\end{cases} (3.4)

The existence and uniqueness of solutions for problem (3.1) is given by the following theorem.

Theorem 3.2.

Let ΩN\Omega\subset\mathbb{R}^{N} be a bounded open set, and g:g:\mathbb{R}\to\mathbb{R} a non-decreasing function. Then, for every fL1(Ω)f\in L^{1}(\Omega), there exists a unique solution (u,w)(u,w) of (3.1) in the sense of Definition 3.1. Moreover, it satisfies

wg+(0)L1(Ω)fg+(0)L1(Ω),\displaystyle\|w-g_{+}(0)\|_{L^{1}(\Omega)}\leq\|f-g_{+}(0)\|_{L^{1}(\Omega)}, (3.5)
{uH01(Ω)CfL1(Ω)if N=1,Tk(u)H01(Ω)2Ω|fg+(0)||Tk(u)|𝑑xk>0,uW01,p(Ω)Cfg+(0)L1(Ω)p<NN1if N2.\displaystyle\begin{cases}\displaystyle\|u\|_{H^{1}_{0}(\Omega)}\leq C\|f\|_{L^{1}(\Omega)}&\hbox{if }N=1,\\ \displaystyle\|T_{k}(u)\|^{2}_{H^{1}_{0}(\Omega)}\leq\int_{\Omega}|f-g_{+}(0)||T_{k}(u)|\,dx\quad\forall\,k>0,\\ \displaystyle\|u\|_{W^{1,p}_{0}(\Omega)}\leq C\|f-g_{+}(0)\|_{L^{1}(\Omega)}\quad\forall\,p<{N\over N-1}&\hbox{if }N\geq 2.\end{cases} (3.6)

In these estimates the constant CC only depends on |Ω||\Omega|, pp, and NN.

Remark 3.3.

Assuming in Theorem 3.2, ff in Lq(Ω)L^{q}(\Omega), with q>1q>1, the classical estimates for renormalized solutions (see for instance [3], [4], [17]) combined with Stampacchia’s estimates (see [25]) also prove that the solution (u,w)(u,w) of (3.1) satisfies

{uW01,NqNq(Ω)if 1<q2NN+2uH01(Ω)LNqN2q(Ω)if 2NN+2<q<N2uH01(Ω)L(Ω)if N2<q.\begin{cases}\displaystyle u\in W_{0}^{1,{Nq\over N-q}}(\Omega)&\displaystyle\hbox{if }1<q\leq{2N\over N+2}\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle u\in H^{1}_{0}(\Omega)\cap L^{Nq\over N-2q}(\Omega)&\displaystyle\hbox{if }{2N\over N+2}<q<{N\over 2}\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle u\in H^{1}_{0}(\Omega)\cap L^{\infty}(\Omega)&\displaystyle\hbox{if }{N\over 2}<q.\end{cases} (3.7)

In particular, assuming fL2NN+2(Ω)f\in L^{2N\over N+2}(\Omega), we have that the solution uu in Theorem 3.2 is in H01(Ω)H^{1}_{0}(\Omega). In this case, it can be defined in a simpler way as the unique solution of the strictly convex minimum problem

minvH01(Ω)Ω(12|v|2+G(v)fv)𝑑x,\min_{v\in H^{1}_{0}(\Omega)}\int_{\Omega}\left({1\over 2}|\nabla v|^{2}+G(v)-fv\right)dx,

with G:G:\mathbb{R}\to\mathbb{R} defined by

G(s)=0sg(r)𝑑rs.G(s)=\int_{0}^{s}g(r)\,dr\qquad\forall\,s\in\mathbb{R}.
Remark 3.4.

Theorem 3.2 and (3.7) can be extended with the same proofs to the case where the operator Δu-\Delta u is replaced by div(A(x)u)-\mbox{\rm div}\,(A(x)\nabla u) with AA a matrix function in L(Ω)N×NL^{\infty}(\Omega)^{N\times N} satisfying the ellipticity condition

c>0 such that A(x)ξξc|ξ|2ξN, a.e. xΩ.\exists\,c>0\hbox{ such that }\ A(x)\xi\cdot\xi\geq c|\xi|^{2}\quad\forall\,\xi\in\mathbb{R}^{N},\hbox{ a.e. }x\in\Omega.

Moreover, the equation

div(A(x)u)+w=f-\mbox{\rm div}\,(A(x)\nabla u)+w=f (3.8)

is satisfied in the sense of distributions for Ω\Omega. Indeed, in the case of the Laplacian operator, since Δ-\Delta is an isomorphism from W01,p(Ω)W^{1,p}_{0}(\Omega) into W1,p(Ω)W^{-1,p}(\Omega), for 1<p<1<p<\infty, it is known that (3.4) is equivalent to require that (u,w)(u,w) satisfies (3.8) in the distributions sense in Ω\Omega. Thus, Theorem 3.2 shows that for every fL1(Ω)f\in L^{1}(\Omega), there exists a unique solution (u,w)W01,p(Ω)L1(Ω)(u,w)\in W_{0}^{1,p}(\Omega)\cap L^{1}(\Omega), 1<p<N/(N1)1<p<N/(N-1), in the distributions sense, of

{Δu+w=fin Ωu=0on Ωg(u)wg+(u)a.e. in Ω.\begin{cases}-\Delta u+w=f&\text{in }\Omega\\ u=0&\hbox{on }\partial\Omega\\ g_{-}(u)\leq w\leq g_{+}(u)&\text{a.e. in }\Omega.\end{cases} (3.9)

However, if we replace Δu-\Delta u by div(Au)-\mbox{\rm div}\,(A\nabla u) with AA as above, this is no longer true. We refer to [24] for a classical counter-example to the uniqueness of the distributional solution.

The following result proves the continuous dependence with respect to the right-hand side and a maximum principle.

Theorem 3.5.

Let ΩN\Omega\subset\mathbb{R}^{N} be a bounded open set and g:g:\mathbb{R}\to\mathbb{R} a non-decreasing function. For f1,f2L1(Ω)f_{1},f_{2}\in L^{1}(\Omega) we take (u1,w1)(u_{1},w_{1}), (u2,w2)(u_{2},w_{2}) solutions of (3.1) with f=f1f=f_{1} and f=f2f=f_{2} respectively. Then, we have

w1w2L1(Ω)f1f2L1(Ω);\displaystyle\|w_{1}-w_{2}\|_{L^{1}(\Omega)}\leq\|f_{1}-f_{2}\|_{L^{1}(\Omega)}; (3.10)
Tk(u1u2)H01(Ω)kf1f2L1(Ω)k>0;\displaystyle\|T_{k}(u_{1}-u_{2})\|_{H^{1}_{0}(\Omega)}\leq k\|f_{1}-f_{2}\|_{L^{1}(\Omega)}\qquad\forall\,k>0; (3.11)
{u1u2H01(Ω)Cf1f2L1(Ω)if N=1u1u2W01,p(Ω)Cf1f2L1(Ω)p(1,NN1)if N2.\displaystyle\begin{cases}\|u_{1}-u_{2}\|_{H^{1}_{0}(\Omega)}\leq C\|f_{1}-f_{2}\|_{L^{1}(\Omega)}&\hbox{if }N=1\\ \displaystyle\|u_{1}-u_{2}\|_{W^{1,p}_{0}(\Omega)}\leq C\|f_{1}-f_{2}\|_{L^{1}(\Omega)}\quad\forall\,p\in\Big{(}1,{N\over N-1}\Big{)}&\hbox{if }N\geq 2.\end{cases} (3.12)

The constant C>0C>0 in this last inequality only depends on |Ω||\Omega| if N=1N=1. For N2N\geq 2 it only depends on NN, pp and |Ω||\Omega|. In addition,

f1f2 a.e. in Ωu1u2 a.e. in Ω.f_{1}\leq f_{2}\hbox{ a.e. in }\Omega\ \Longrightarrow\ u_{1}\leq u_{2}\hbox{ a.e. in }\Omega.

Our main result in this section proves that the function ww in Theorem 3.2 is in BV(Ω)BV(\Omega) when ff is in BV(Ω)BV(\Omega). It will be used in Theorem 4.7 to deduce some regularity results for the solution of a control problem governed by an elliptic equation, where the control variable corresponds to the coefficients of the zero order’s term.

Theorem 3.6.

Let ΩN\Omega\subset\mathbb{R}^{N} be a bounded open set of class C1,1C^{1,1}, and g:g:\mathbb{R}\to\mathbb{R} a non-decreasing function. Then, for fBV(Ω),f\in BV(\Omega), the solution (u,w)(u,w) of (3.9) is in W2,NN1(Ω)×BV(Ω)W^{2,{N\over N-1}}(\Omega)\times BV(\Omega). Moreover, there exists C>0C>0 depending only on Ω\Omega such that

uW2,NN1(Ω)C(fBV(Ω)+|g+(0)|),\displaystyle\|u\|_{W^{2,{N\over N-1}}(\Omega)}\leq C\big{(}\|f\|_{BV(\Omega)}+|g_{+}(0)|\big{)}, (3.13)
w(Ω)NC(fBV(Ω)+|g+(0)|).\displaystyle\|\nabla w\|_{{\cal M}(\Omega)^{N}}\leq C\big{(}\|f\|_{BV(\Omega)}+|g_{+}(0)|). (3.14)

Let us now prove Theorems 3.2, 3.5 and 3.6.

Proof of Theorem 3.2.

When gg is a continuous function, the result is well known from the theory of renormalized solutions for elliptic PDE (see for instance [3], [4], [17]). We recall how the corresponding estimates are obtained.

Taking Tk(u)T_{k}(u), with k>0k>0 as test function in (3.4) we have

{|u|<k}|u|2𝑑x+Ω(g(u)g(0))Tk(u)𝑑x=Ω(fg(0))Tk(u)𝑑x.\int_{\{|u|<k\}}\hskip-6.0pt|\nabla u|^{2}\,dx+\int_{\Omega}\big{(}g(u)-g(0)\big{)}T_{k}(u)\,dx=\int_{\Omega}(f-g(0))T_{k}(u)\,dx. (3.15)

Thanks to the fact that gg is non-decreasing, we have (g(u)g(0))Tk(u)0(g(u)-g(0))T_{k}(u)\geq 0 a.e. in Ω\Omega and this gives the second estimate in (3.6). From this inequality, using the argument in Theorem 1 of [4], we conclude that uu satisfies (3.6).

Dividing by kk in (3.15) and taking the limit as k0k\to 0 we get

limk01k{|u|<k}|u|2𝑑x+Ω|g(u)g(0)|𝑑x=Ω(fg(0))sgn(u)𝑑x.\lim_{k\to 0}{1\over k}\int_{\{|u|<k\}}\hskip-6.0pt|\nabla u|^{2}dx+\int_{\Omega}|g(u)-g(0)|dx=\int_{\Omega}(f-g(0))\operatorname{sgn}(u)\,dx.

This proves (3.5) with w=g(u)w=g(u).

Let us now prove the existence of solution for (3.9) in the case of gg just non-decreasing. For this purpose we replace gg by gn=gρng_{n}=g\ast\rho_{n}, nn\in\mathbb{N}, with ρn\rho_{n} a sequence of mollifiers functions defined as

ρn(s)=nρ(ns)s,\rho_{n}(s)=n\rho\big{(}ns)\qquad\forall\,s\in\mathbb{R},

with

ρC(),ρ0 in ,support(ρ)(1,0),ρ(s)𝑑s=1.\rho\in C^{\infty}(\mathbb{R}),\quad\rho\geq 0\ \hbox{ in }\mathbb{R},\quad{\rm support}(\rho)\subset(-1,0),\quad\int_{\mathbb{R}}\rho(s)ds=1.

Then, gng_{n} satisfies

gnC(),gn is non-decreasing,g(s)gn(s)g(s+1n)s.g_{n}\in C^{\infty}(\mathbb{R}),\quad g_{n}\hbox{ is non-decreasing},\quad g(s)\leq g_{n}(s)\leq g\Big{(}s+{1\over n}\Big{)}\quad\forall\,s\in\mathbb{R}. (3.16)

Taking unu_{n} the solution of (3.1) for g=gng=g_{n}, and using estimates (3.5) and (3.6), we deduce the existence of a subsequence of nn, still denoted by nn, and a function uu such that

unu in W01,p(Ω)p<NN1(p=2 if N=1),\displaystyle u_{n}\rightharpoonup u\ \hbox{ in }W^{1,p}_{0}(\Omega)\quad\forall\,p<{N\over N-1}\quad(p=2\hbox{ if }N=1), (3.17)
Tk(un)Tk(u) in H01(Ω)k>0,\displaystyle T_{k}(u_{n})\rightharpoonup T_{k}(u)\ \hbox{ in }H^{1}_{0}(\Omega)\quad\forall\,k>0, (3.18)
Ω|gn(un)gn(0)|𝑑xΩ|fgn(0)|𝑑x.\displaystyle\int_{\Omega}|g_{n}(u_{n})-g_{n}(0)|\,dx\leq\int_{\Omega}|f-g_{n}(0)|\,dx. (3.19)

From (3.15) we also have

limn({|un|<k}|un|2𝑑x+Ω(gn(un)gn(0))Tk(un)𝑑x)=Ω(fg+(0))Tk(u)𝑑x,\lim_{n\to\infty}\left(\int_{\{|u_{n}|<k\}}\hskip-6.0pt|\nabla u_{n}|^{2}\,dx+\int_{\Omega}\big{(}g_{n}(u_{n})-g_{n}(0)\big{)}T_{k}(u_{n})\,dx\right)=\int_{\Omega}(f-g_{+}(0))T_{k}(u)\,dx,

for every k>0k>0. Dividing by kk and taking the limit as kk\to\infty, gives

limklimn(1k{|un|<k}|un|2𝑑x+{|un|>k}|gn(un)|𝑑x)=0.\lim_{k\to\infty}\lim_{n\to\infty}\left({1\over k}\int_{\{|u_{n}|<k\}}\hskip-7.0pt|\nabla u_{n}|^{2}dx+\int_{\{|u_{n}|>k\}}\hskip-5.0pt\big{|}g_{n}(u_{n})|dx\right)=0. (3.20)

Let us prove that gn(un)g_{n}(u_{n}) is compact in the weak topology of L1(Ω)L^{1}(\Omega). By (3.16), (3.19) and the Dunford-Pettis theorem, it is enough to prove that gn(un)g_{n}(u_{n}) is equi-integrable, i.e. that for every ε>0\varepsilon>0, there exists δ>0\delta>0 such that

E|gn(un)|𝑑xεEΩ, measurable, with |E|<δ.\int_{E}|g_{n}(u_{n})|dx\leq\varepsilon\qquad\forall\,E\subset\Omega,\hbox{ measurable, with }|E|<\delta.

For such ε\varepsilon, thanks to (3.20), there exist k,m>0k,m>0 such that

{|un|>k}|gn(un))|dx<ε2nm.\int_{\{|u_{n}|>k\}}\hskip-5.0pt\big{|}g_{n}(u_{n}))|dx<{\varepsilon\over 2}\quad\forall\,n\geq m.

Choosing then

δ1<ε2sup[k,k+1]|g(s)|,\delta_{1}<{\varepsilon\over 2\sup_{[-k,k+1]}|g(s)|},

and taking into account (3.16), we deduce that for every EΩE\subset\Omega, measurable with |E|<δ|E|<\delta, we have

E|gn(un)|𝑑xEsup[k,k+1]|g(s)|dx+{|un|>k}|gn(un)|𝑑x<εnm.\int_{E}|g_{n}(u_{n})|dx\leq\int_{E}\,\sup_{[-k,k+1]}|g(s)|dx+\int_{\{|u_{n}|>k\}}\hskip-8.0pt|g_{n}(u_{n})|dx<\varepsilon\quad\forall\,n\geq m.

On the other hand, since every finite subset of functions in L1(Ω)L^{1}(\Omega) is equi-integrable, there exists δ2>0\delta_{2}>0 such that

E|gn(un)|𝑑x<εn<m.\int_{E}|g_{n}(u_{n})|dx<\varepsilon\quad\forall\,n<m.

Thus, taking δ=min{δ1,δ2}\delta=\min\{\delta_{1},\delta_{2}\} we deduce the equi-integrability of gn(un)g_{n}(u_{n}). Extracting a subsequence if necessary, we then deduce the existence of wL1(Ω)w\in L^{1}(\Omega) such that

gn(un)w in L1(Ω).g_{n}(u_{n})\rightharpoonup w\ \hbox{ in }L^{1}(\Omega). (3.21)

From (3.16), (3.17) and the Rellich-Kondrachov compactness theorem, we also have

g(u)wg+(u) a.e. in Ω.g_{-}(u)\leq w\leq g_{+}(u)\quad\hbox{ a.e. in }\Omega.

Let us prove that (u,w)(u,w) satisfies (3.4). We take vH01(Ω)L(Ω)v\in H^{1}_{0}(\Omega)\cap L^{\infty}(\Omega) such that there exists k>0k>0 with v=0\nabla v=0 a.e. in {|u|>k}.\{|u|>k\}. For m>0m>0, we use Sm(un)vS_{m}(u_{n})v as test function in the equation for unu_{n}. We get

1m{m<|un|<2m}|un|2sgn(un)v𝑑x+ΩunvSm(un)𝑑x+Ωgn(un)Sm(un)v𝑑x=ΩfSm(un)v𝑑x.\begin{split}&\displaystyle-{1\over m}\int_{\{m<|u_{n}|<2m\}}\hskip-12.0pt|\nabla u_{n}|^{2}\operatorname{sgn}(u_{n})v\,dx+\int_{\Omega}\nabla u_{n}\cdot\nabla v\,S_{m}(u_{n})\,dx\\ &\qquad\qquad\displaystyle+\int_{\Omega}g_{n}(u_{n})S_{m}(u_{n})v\,dx=\int_{\Omega}fS_{m}(u_{n})v\,dx.\end{split}

Taking into account (3.18), (3.21) and the fact that Sm(un)vS_{m}(u_{n})v is bounded in L(Ω)L^{\infty}(\Omega) and converges in measure to Sm(u)vS_{m}(u)v, we can pass to the limit as nn\to\infty in this inequality to get

|ΩuvSm(u)𝑑x+ΩwSm(u)v𝑑xΩfSm(u)v𝑑x|vL(Ω)mlim supn{m<|un|<2m}|un|2𝑑x.\begin{array}[]{l}\displaystyle\left|\int_{\Omega}\nabla u\cdot\nabla v\,S_{m}(u)\,dx+\int_{\Omega}wS_{m}(u)v\,dx-\int_{\Omega}fS_{m}(u)v\,dx\right|\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle\leq{\|v\|_{L^{\infty}(\Omega)}\over m}\limsup_{n\to\infty}\int_{\{m<|u_{n}|<2m\}}\hskip-18.0pt|\nabla u_{n}|^{2}dx.\end{array}

By (3.20), Tk(u)H01(Ω)T_{k}(u)\in H^{1}_{0}(\Omega) and v=0\nabla v=0 a.e. in {|u|>k}\{|u|>k\}, we can now pass to the limit as mm\to\infty to deduce that (3.4) holds.

The uniqueness of solutions for (3.9) follows from Theorem 3.5. ∎

Proof of Theorem 3.5.

Let us assume N2N\geq 2. The case N=1N=1 is simpler taking into account the continuous imbedding of L1(Ω)L^{1}(\Omega) into H1(Ω)H^{-1}(\Omega).

For m,k>0m,k>0, we take Sm(u1)Sm(u2)Tk(u1u2)+S_{m}(u_{1})S_{m}(u_{2})T_{k}(u_{1}-u_{2})^{+} as test function in the difference of the equations satisfied by (u1,w1)(u_{1},w_{1}) and (u2,w2)(u_{2},w_{2}). This gives

{0<u1u2<k}|(u1u2)|2Sm(u1)Sm(u2)𝑑x+Ω(w1w2)Tk(u1u2)+Sm(u1)Sm(u2)𝑑x1m{m<|u2|<2m}Tk(u1u2)+sgn(u2)Sm(u1)(u1u2)u2dx1m{m<|u1|<2m}Tk(u1u2)+sgn(u1)Sm(u2)(u1u2)u1dx=Ω(f1f2)Tk(u1u2)+Sm(u1)Sm(u2)𝑑x.\begin{array}[]{l}\displaystyle\int_{\{0<u_{1}-u_{2}<k\}}\hskip-12.0pt|\nabla(u_{1}-u_{2})|^{2}S_{m}(u_{1})S_{m}(u_{2})dx\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle+\int_{\Omega}(w_{1}-w_{2})T_{k}(u_{1}-u_{2})^{+}S_{m}(u_{1})S_{m}(u_{2})\,dx\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle-{1\over m}\int_{\{m<|u_{2}|<2m\}}\hskip-16.0ptT_{k}(u_{1}-u_{2})^{+}\operatorname{sgn}(u_{2})S_{m}(u_{1})\nabla(u_{1}-u_{2})\cdot\nabla u_{2}\,dx\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle-{1\over m}\int_{\{m<|u_{1}|<2m\}}\hskip-16.0ptT_{k}(u_{1}-u_{2})^{+}\operatorname{sgn}(u_{1})S_{m}(u_{2})\nabla(u_{1}-u_{2})\cdot\nabla u_{1}\,dx\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle=\int_{\Omega}(f_{1}-f_{2})\,T_{k}(u_{1}-u_{2})^{+}S_{m}(u_{1})S_{m}(u_{2})\,dx.\end{array} (3.22)

Here we use the second estimate in (3.6) for u=u1u=u_{1}, u=u2u=u_{2}, with k=mk=m, which dividing by mm and taking the limit as mm\to\infty, gives

limm1m{|u1|<2m}|u1|2𝑑x=1mlimm{|u2|<2m}|u2|2𝑑x=0.\lim_{m\to\infty}{1\over m}\int_{\{|u_{1}|<2m\}}\hskip-12.0pt|\nabla u_{1}|^{2}dx={1\over m}\lim_{m\to\infty}\int_{\{|u_{2}|<2m\}}\hskip-12.0pt|\nabla u_{2}|^{2}dx=0.

This allows us to pass to the limit in (3.22), as mm\to\infty, to deduce

{0<u1u2<k}|(u1u2)|2𝑑x+Ω(w1w2)Tk(u1u2)+𝑑x=Ω(f1f2)Tk(u1u2)+𝑑x,k>0.\begin{array}[]{l}\displaystyle\int_{\{0<u_{1}-u_{2}<k\}}\hskip-16.0pt|\nabla(u_{1}-u_{2})|^{2}dx+\int_{\Omega}(w_{1}-w_{2})T_{k}(u_{1}-u_{2})^{+}dx\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle=\int_{\Omega}(f_{1}-f_{2})T_{k}(u_{1}-u_{2})^{+}dx,\qquad\forall\,k>0.\end{array} (3.23)

Now, we observe that the conditions

g(u1)w1g+(u1),g(u2)w2g+(u2) a.e in Ω,g_{-}(u_{1})\leq w_{1}\leq g_{+}(u_{1}),\ \ g_{-}(u_{2})\leq w_{2}\leq g_{+}(u_{2})\qquad\hbox{ a.e in }\Omega,

imply

(w1w2)Tk(u1u2)+0a.e. in Ω.(w_{1}-w_{2})T_{k}(u_{1}-u_{2})^{+}\geq 0\qquad\hbox{a.e. in }\Omega.

Therefore (3.23) proves

{0<u1u2<k}|(u1u2)|2𝑑xk{0<u1u2<k}|f1f2|𝑑x.\int_{\{0<u_{1}-u_{2}<k\}}\hskip-16.0pt|\nabla(u_{1}-u_{2})|^{2}\,dx\leq k\int_{\{0<u_{1}-u_{2}<k\}}\hskip-16.0pt|f_{1}-f_{2}|\,dx.

Adding the analogous inequality with u1,u2u_{1},u_{2} replaced by each other, we deduce (3.11). This inequality also implies (3.12) (see [4]).

Dividing by kk in (3.23) and passing to the limit as k0k\to 0, we get

limk01k{0<u1u2<k}|(u1u2)|2𝑑x+{u2<u1}|w1w2|𝑑x={u2<u1}(f1f2)𝑑x.\lim_{k\to 0}{1\over k}\int_{\{0<u_{1}-u_{2}<k\}}\hskip-16.0pt|\nabla(u_{1}-u_{2})|^{2}\,dx+\int_{\{u_{2}<u_{1}\}}\hskip-12.0pt|w_{1}-w_{2}|\,dx=\int_{\{u_{2}<u_{1}\}}\hskip-12.0pt(f_{1}-f_{2})\,dx.

Using the analogous equality with u1,u2u_{1},u_{2} replaced by each other we conclude (3.10).

If f1f2f_{1}\leq f_{2} a.e. in Ω\Omega, then (3.23) proves

{0<u1u2<k}|(u1u2)|2𝑑x=0k>0,\int_{\{0<u_{1}-u_{2}<k\}}\hskip-16.0pt|\nabla(u_{1}-u_{2})|^{2}\,dx=0\qquad\forall\,k>0,

and then that u1u2u_{1}\leq u_{2} a.e. in Ω\Omega. ∎

Proof of Theorem 3.6.

Let us first assume gg in W1,()W^{1,\infty}(\mathbb{R}), fW1,1(Ω)L2(Ω)f\in W^{1,1}(\Omega)\cap L^{2}(\Omega). Then uu belongs to H2(Ω)H^{2}(\Omega). Taking into account the boundary condition u=0u=0 on Ω\partial\Omega, and ΩC1,1\Omega\in C^{1,1}, we can use equation (4.19) in the proof of Lemma 4.3 in [16] to prove that the second derivatives of uu satisfy

{Δiu+g(u)iu=ifin Ω,1iNu=|u|sνon ΩD2uνν+g(0)=f+huon Ω,\begin{cases}-\Delta\partial_{i}u+g^{\prime}(u)\partial_{i}u=\partial_{i}f&\hbox{in }\Omega,\quad 1\leq i\leq N\\ \nabla u=|\nabla u|s\nu&\hbox{on }\partial\Omega\\ -D^{2}u\,\nu\cdot\nu+g(0)=f+h\cdot\nabla u&\hbox{on }\partial\Omega,\end{cases} (3.24)

where ν\nu denotes the unitary outside normal to Ω\Omega, h,sh,s satisfy

hL(Ω)N,sL(Ω),s{1,1}, a.e. in Ω.h\in L^{\infty}(\partial\Omega)^{N},\quad s\in L^{\infty}(\partial\Omega),\ s\in\{-1,1\},\hbox{ a.e. in }\Omega.

and they only depend on Ω\Omega.

For ε>0\varepsilon>0, we multiply the first equation in (3.24) by

iu|u|+εH1(Ω)L(Ω).{\partial_{i}u\over|\nabla u|+\varepsilon}\in H^{1}(\Omega)\cap L^{\infty}(\Omega).

Integrating by parts, adding in ii, and using the boundary conditions in (3.24), we get

Ω(|D2u|2|u|+ε|D2uu|2|u|(|u|+ε)2)𝑑x+Ωg(u)|u|2|u|+ε𝑑x=Ωs(f+hug(0))|u||u|+ε𝑑s(x)+Ωfu|u|+ε𝑑x.\begin{array}[]{l}\displaystyle\int_{\Omega}\Big{(}{|D^{2}u|^{2}\over|\nabla u|+\varepsilon}-{|D^{2}u\nabla u|^{2}\over|\nabla u|(|\nabla u|+\varepsilon)^{2}}\Big{)}\,dx+\int_{\Omega}g^{\prime}(u){|\nabla u|^{2}\over|\nabla u|+\varepsilon}\,dx\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle=-\int_{\partial\Omega}s\big{(}f+h\cdot\nabla u-g(0)\big{)}{|\nabla u|\over|\nabla u|+\varepsilon}\,ds(x)+\int_{\Omega}{\nabla f\cdot\nabla u\over|\nabla u|+\varepsilon}\,dx.\end{array}

Using here that

|D2u|2|D2uu|2|u|(|u|+ε)|D2u|2|D2uu|2|u|2=|D2u(Iuu|u|2)|20,|D^{2}u|^{2}-{|D^{2}u\nabla u|^{2}\over|\nabla u|(|\nabla u|+\varepsilon)}\geq|D^{2}u|^{2}-{|D^{2}u\nabla u|^{2}\over|\nabla u|^{2}}=\Big{|}D^{2}u\Big{(}I-{\nabla u\otimes\nabla u\over|\nabla u|^{2}}\Big{)}\Big{|}^{2}\geq 0,

a.e. in Ω\Omega, we can pass to the limit as ε0\varepsilon\to 0 to deduce

Ω1|u|(|D2u|2|D2uu|2|u|2)𝑑x+Ωg(u)|u|𝑑x=Ωs(f+hug(0))𝑑σ(x)+Ωfu|u|𝑑x.\begin{array}[]{l}\displaystyle\int_{\Omega}{1\over|\nabla u|}\Big{(}|D^{2}u|^{2}-{|D^{2}u\nabla u|^{2}\over|\nabla u|^{2}}\Big{)}\,dx+\int_{\Omega}g^{\prime}(u)|\nabla u|\,dx\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\displaystyle=-\int_{\partial\Omega}s\big{(}f+h\cdot\nabla u-g(0)\big{)}\,d\sigma(x)+\int_{\Omega}{\nabla f\cdot\nabla u\over|\nabla u|}\,dx.\end{array} (3.25)

In the first term of the right-hand side, we can apply the trace theorem for functions in W1,1(Ω)W^{1,1}(\Omega) which proves the existence of CC depending only on Ω\Omega such that

fL1(Ω)CfL1(Ω)N.\|f\|_{L^{1}(\partial\Omega)}\leq C\|\nabla f\|_{L^{1}(\Omega)^{N}}. (3.26)

Using also that the imbedding of W1,1(Ω)W^{1,1}(\Omega) into Lp(Ω)L^{p}(\Omega) is compact for 1<p<NN11<p<{N\over N-1}, we deduce that for every δ>0\delta>0, there exists C>0C>0 depending on Ω\Omega, pp and δ\delta, such that

ΔuLp(Ω)δ(g(u)L1(Ω)N+fL1(Ω)N)+C(g(u)L1Ω)+fL1(Ω)).\|\Delta u\|_{L^{p}(\Omega)}\leq\delta\big{(}\|\nabla g(u)\|_{L^{1}(\Omega)^{N}}+\|\nabla f\|_{L^{1}(\Omega)^{N}}\big{)}+C\big{(}\|g(u)\|_{L^{1}\Omega)}+\|f\|_{L^{1}(\Omega)}\big{)}.

Applying then that (Δ)1(-\Delta)^{-1} is continuous from Lp(Ω)L^{p}(\Omega) into W01,p(Ω)W2,p(Ω)W^{1,p}_{0}(\Omega)\cap W^{2,p}(\Omega), and the trace theorem for Sobolev spaces, we conclude that, for another constant C>0C>0, we have

uL1(Ω)Nδ(g(u)L1(Ω)N+fL1(Ω)N)+C(g(u)L1Ω)+fL1(Ω)),\|\nabla u\|_{L^{1}(\partial\Omega)^{N}}\leq\delta\big{(}\|\nabla g(u)\|_{L^{1}(\Omega)^{N}}+\|\nabla f\|_{L^{1}(\Omega)^{N}}\big{)}+C\big{(}\|g(u)\|_{L^{1}\Omega)}+\|f\|_{L^{1}(\Omega)}\big{)},

which combined with (3.5), with w=g(u)w=g(u), proves

uL1(Ω)Nδg(u)L1(Ω)N+C(fW1,1(Ω)+|g(0)|),\|\nabla u\|_{L^{1}(\partial\Omega)^{N}}\leq\delta\|\nabla g(u)\|_{L^{1}(\Omega)^{N}}+C\big{(}\|f\|_{W^{1,1}(\Omega)}+|g(0)|), (3.27)

with CC depending on Ω\Omega and δ\delta.

Choosing δ\delta such that δhL(Ω)<1\delta\|h\|_{L^{\infty}(\partial\Omega)}<1, we can use (3.26) and (3.27) in (3.25), to conclude the existence of C>0C>0 depending only on Ω\Omega such that (3.14) holds. From this estimate, the continuous imbedding of W1,1(Ω)W^{1,1}(\Omega) into LNN1(Ω)L^{N\over N-1}(\Omega), and uu solution of (3.9), with w=g(u)w=g(u), we also have that uu satisfies (3.13). This proves the result for gW1,()g\in W^{1,\infty}(\mathbb{R}), fW1,1(Ω)L2(Ω)f\in W^{1,1}(\Omega)\cap L^{2}(\Omega).

The case where gg is just an increasing function follows by approximating gg by a sequence of smooth functions gng_{n} as in the proof of Theorem 3.2.

The case where ff is just in BV(Ω)BV(\Omega) follows by replacing ff by a sequence fnf_{n} in W1,1(Ω)L2(Ω)W^{1,1}(\Omega)\cap L^{2}(\Omega), such that fnf_{n} converges to ff in L1(Ω)L^{1}(\Omega) and fnL1(Ω)N\|\nabla f_{n}\|_{L^{1}(\Omega)^{N}} converges to f(Ω)N\|\nabla f\|_{{\cal M}(\Omega)^{N}}. ∎

4 Applications to optimal potentials problems

In this section, we are interested in the study of an optimal control problem for an elliptic equation of a Schrödinger type, where the control variable is the potential. Namely, we consider the problem

minΩ(j(x,u)+ψ(m))𝑑x{Δu+mu=f in ΩuH01(Ω),mL1(Ω),m0 a.e. in Ω,\begin{split}&\displaystyle\min\int_{\Omega}\big{(}j(x,u)+\psi(m)\big{)}\,dx\\ &\begin{cases}-\Delta u+m\,u=f\ \hbox{ in }\Omega\\ u\in H^{1}_{0}(\Omega),\quad m\in L^{1}(\Omega),\quad m\geq 0\ \hbox{ a.e. in }\Omega,\end{cases}\end{split} (4.1)

where ψ\psi is a lower semicontinuous convex function. The constraint m0m\geq 0 in Ω\Omega is introduced in the cost functional taking

ψ(s)=+s<0.\psi(s)=+\infty\qquad\forall\,s<0.

Problems of this type intervene in several shape optimization problems, where mm is a Borel measure of capacitary type, not Radon in general (see for instance [2], [5], [6], [7], [16], [9], [10], [12], [13], [14] and [22]). Similarly to these papers, the solution uu of (4.1) must be understood in the variational sense

{uH01(Ω)Lm2(Ω)Ω(uv+muv)𝑑x=f,vvH01(Ω)Lm2(Ω).\begin{cases}u\in H^{1}_{0}(\Omega)\cap L^{2}_{m}(\Omega)\\ \displaystyle\int_{\Omega}\big{(}\nabla u\cdot\nabla v+muv\big{)}\,dx=\langle f,v\rangle\quad\forall\,v\in H^{1}_{0}(\Omega)\cap L^{2}_{m}(\Omega).\end{cases}

In the present paper, we are interested in obtaining some regularity results for the optimal controls m^\hat{m}, which in our case are integrable functions. In particular, we show that in several cases the optimal controls m^\hat{m} are of bang-bang type, and then discontinuous. However, we show that, under some suitable assumptions, m^\hat{m} is a BVBV function and then that the interfaces have finite perimeter.

Our first result proves the existence of solution for (4.1). We refer to Theorem 2.19 in [16] for a related result in a more general setting.

Theorem 4.1.

Let ΩN\Omega\subset\mathbb{R}^{N} be a bounded open set, j:Ω×j:\Omega\times\mathbb{R}\to\mathbb{R} measurable in the first component and lower semicontinuous in the second one, satisfying (2.6), and ψ:[0,]\psi:\mathbb{R}\to[0,\infty] a convex lower semicontinuous function with dom(ψ)[0,)\operatorname{dom}(\psi)\subset[0,\infty), such that (2.5) holds. Then, for every fH1(Ω)f\in H^{-1}(\Omega), problem (4.1) has a least one solution m^L1(Ω)\hat{m}\in L^{1}(\Omega).

The optimality conditions for (4.1) are given by Theorem 4.2 below (see also Theorem 4.1 in [16]). Since our aim in the present work is to present some regularity conditions for the solutions of problem (4.1), let us assume that the right-hand side ff in the state equation satisfies

fW1,r(Ω),with {r2if N=1r>Nif N2.f\in W^{-1,r}(\Omega),\quad\text{with }\begin{cases}r\geq 2&\text{if }N=1\\ r>N&\text{if }N\geq 2.\end{cases} (4.2)

By Stampacchia’s estimates (see for instance [25]), this implies that there exists M>0M>0, such that for every mL1(Ω)m\in L^{1}(\Omega), m0m\geq 0 a.e. in Ω\Omega, the solution uu of the state equation in (4.1) is in L(Ω)L^{\infty}(\Omega), with uL(Ω)M\|u\|_{L^{\infty}(\Omega)}\leq M. In particular, this means that the value of j(x,s)j(x,s) for |s|M|s|\geq M is not important and then we can replace in Theorem 4.1 condition (2.6) by

inf{|s|M}j(x,s)L1(Ω)M>0.\inf_{\{|s|\leq M\}}j(x,s)\in L^{1}(\Omega)\qquad\forall\,M>0.

Further we will assume j(x,.)C1()j(x,.)\in C^{1}(\mathbb{R}) and satisfying

j(.,0)L1(Ω),max|s|M|sj(,s)|Lr2(Ω)M>0.j(.,0)\in L^{1}(\Omega),\qquad\max_{|s|\leq M}|\partial_{s}j(\cdot,s)|\in L^{r\over 2}(\Omega)\quad\forall\,M>0. (4.3)

In these conditions, the following result holds.

Theorem 4.2.

Assume that in Theorem 4.1, the right-hand side ff and the function jj satisfy (4.2) and (4.3) respectively, and define h:h:\mathbb{R}\to\mathbb{R} by

h(τ)=max{sdom(ψ):τψ(s)}.h(\tau)=\max\big{\{}s\in\operatorname{dom}(\psi)\ :\ \tau\in\partial\psi(s)\big{\}}. (4.4)

Then, if m^\hat{m} is a solution of (4.1), u^\hat{u} is the corresponding state function, solution of

{Δu^+m^u^=fin Ωu^=0on Ω,\begin{cases}-\Delta\hat{u}+\hat{m}\,\hat{u}=f&\text{in }\Omega\\ \hat{u}=0&\text{on }\partial\Omega,\end{cases} (4.5)

and z^\hat{z} is the adjoint state, solution of

{Δz^+m^z^=sj(x,u^)in Ωz^=0on Ω,\begin{cases}-\Delta\hat{z}+\hat{m}\,\hat{z}=\partial_{s}j(x,\hat{u})&\text{in }\Omega\\ \hat{z}=0&\text{on }\partial\Omega,\end{cases} (4.6)

we have

m^L(Ω),u^z^ψ(m^),h(u^z^)m^h(u^z^), a.e. in Ω.\hat{m}\in L^{\infty}(\Omega),\quad\hat{u}\hat{z}\in\partial\psi(\hat{m}),\ \ h_{-}(\hat{u}\hat{z})\leq\hat{m}\leq h(\hat{u}\hat{z}),\quad\hbox{ a.e. in }\Omega. (4.7)
Remark 4.3.

Taking into account that dom(ψ)[0,)\operatorname{dom}(\psi)\subset[0,\infty), ψ\psi lower semicontinuous and (2.5) we deduce that

limsdψ(s)=,\lim_{s\to\infty}d_{-}\psi(s)=\infty,

and that, taking

α:=infdom(ψ)0,\alpha:=\inf\operatorname{dom}(\psi)\geq 0, (4.8)

one of the following conditions hold

limsαψ(s)=,limsαd+ψ(s)= or αdom(ψ),dψ(α)=.\lim_{s\searrow\alpha}\psi(s)=\infty,\ \lim_{s\searrow\alpha}d_{+}\psi(s)=-\infty\quad\hbox{ or }\quad\alpha\in\operatorname{dom}(\psi),\ d_{-}\psi(\alpha)=-\infty.

Therefore, for every τ\tau\in\mathbb{R}, there exists sdom(ψ)s\in\operatorname{dom}(\psi) such that τψ(s)\tau\in\partial\psi(s).

Remark 4.4.

By (2.3) and (2.4), we have that hh in Theorem 4.2) is also given by

h(τ)=d+ψ(τ)τ.h(\tau)=d_{+}\psi^{\ast}(\tau)\qquad\forall\,\tau\in\mathbb{R}.

It is always a non-decreasing function, continuous on the right. Moreover, it also satisfies

h(τ)=max{sdom(ψ):τdψ(s)}.h(\tau)=\max\big{\{}s\in\operatorname{dom}(\psi):\ \tau\geq d_{-}\psi(s)\big{\}}. (4.9)

From (4.7) and the regularity results for elliptic equations, we deduce that the optimal measure m^\hat{m} is very regular if hh is and the functions jj and ff are very regular too. In this sense, the following proposition provides some necessary and sufficient conditions to have hh continuous and locally Lipschitz-continuous respectively.

Proposition 4.5.

The function hh defined by (4.4) satisfies

hC0()ψ is strictly convex. h\in C^{0}(\mathbb{R})\Longleftrightarrow\psi\ \hbox{ is strictly convex. } (4.10)
hLip()0<infs1,s2dom(ψ)s1<s2dψ(s2)d+ψ(s1)s2s1.h\in{\rm Lip}(\mathbb{R})\iff 0<\inf_{s_{1},s_{2}\in\operatorname{dom}(\psi)\atop s_{1}<s_{2}}{d_{-}\psi(s_{2})-d_{+}\psi(s_{1})\over s_{2}-s_{1}}. (4.11)

Further assumptions on ψ\psi than those in Proposition 4.5 provide more regularity for hh, but it is interesting to note that

d+ψ(α)>ψ(α)=(,d+ψ(α)]h(s)=α,s(,d+ψ(α)),d_{+}\psi(\alpha)>-\infty\Longrightarrow\partial\psi(\alpha)=\big{(}-\infty,d_{+}\psi(\alpha)\big{]}\Longrightarrow h(s)=\alpha,\ \forall\,s\in(-\infty,d_{+}\psi(\alpha)\big{)},

with α\alpha defined by (4.8). Thus, hh cannot be an analytic function if d+ψ(α)>.d_{+}\psi(\alpha)>-\infty. Even more, we have the following result.

Proposition 4.6.

Assume that α\alpha defined by (4.8) is such that d+ψ(α)>d_{+}\psi(\alpha)>-\infty. Then

h(d+ψ(α))limsαd+ψ(s)d+ψ(α)sα=.\exists\,h^{\prime}(d_{+}\psi(\alpha))\iff\lim_{s\searrow\alpha}{d_{+}\psi(s)-d_{+}\psi(\alpha)\over s-\alpha}=\infty. (4.12)

From Proposition 4.5, we have that hh (and then m^\hat{m}) is not continuous if ψ\psi is not strictly convex. Moreover, Proposition 4.6 shows that even if hh is continuous, it is not derivable in general. Theorem 4.7 below provides a sufficient condition to get m^u^z^\hat{m}\hat{u}\hat{z} in BV(Ω)BV(\Omega), and then shows that the discontinuity surfaces of m^\hat{m} have finite perimeter.

Theorem 4.7.

In addition to the conditions in Theorem 4.1 we assume ΩC1,1\Omega\in C^{1,1},

g(τ):=h(τ)ττ,g(\tau):=h(\tau)\tau\qquad\forall\,\tau\in\mathbb{R}, (4.13)

non-decreasing in τ\tau, and

max|s|M|xsj(.,s)|Lq(Ω),max|s|M|ss2j(.,s)|L1(Ω),M>0,\max_{|s|\leq M}|\nabla_{x}\partial_{s}j(.,s)|\in L^{q}(\Omega),\quad\max_{|s|\leq M}|\partial^{2}_{ss}j(.,s)|\in L^{1}(\Omega),\quad\forall\,M>0, (4.14)

with

q2NN+1 if 1N2,q>N2 if N3.q\geq{2N\over N+1}\ \hbox{ if }1\leq N\leq 2,\qquad q>{N\over 2}\ \hbox{ if }N\geq 3. (4.15)

Then, for every fBV(Ω)Lq(Ω)f\in BV(\Omega)\cap L^{q}(\Omega) and every solution m^\hat{m} of (4.1), we have

u^,z^W2,q(Ω),m^u^z^BV(Ω),\hat{u},\hat{z}\in W^{2,q}(\Omega),\qquad\hat{m}\hat{u}\hat{z}\in BV(\Omega),

with u^\hat{u}, z^\hat{z} the solutions of (4.5) and (4.6) respectively.

Remark 4.8.

In the assumptions of Theorem 4.7, the functions u^\hat{u}, z^\hat{z} are continuous, and thus, the set E:={u^z^=0}E:=\{\hat{u}\hat{z}=0\} is a closed subset of Ω¯\overline{\Omega} (which contains the boundary). The fact that m^u^z^\hat{m}\hat{u}\hat{z} belongs to BV(Ω)BV(\Omega), proves then that m^\hat{m} belongs to BVloc(ΩE)BV_{loc}(\Omega\setminus E).

Remark 4.9.

Since hh is non-decreasing, a sufficient condition to have gg non-decreasing is to assume d+ψ(α)0d_{+}\psi(\alpha)\geq 0, with α\alpha defined by (4.8).

Proof of Theorem 4.1.

In order to prove the existence of solution, we apply the direct method of the calculus of variations. We take mnL1(Ω)m_{n}\in L^{1}(\Omega), mn0m_{n}\geq 0 a.e. in Ω\Omega, such that the solution unu_{n} of the state equation in (4.1) satisfies

limnΩ(j(x,un)+ψ(mn))𝑑x=,\lim_{n\to\infty}\int_{\Omega}\big{(}j(x,u_{n})+\psi(m_{n})\big{)}dx={\cal I},

where we have denoted by {\cal I} the infimum of (4.1). In particular

lim supnΩψ(mn)𝑑x<,\limsup_{n\to\infty}\int_{\Omega}\psi(m_{n})\,dx<\infty,

which, taking into account (2.5), implies that mnm_{n} is compact in the weak topology of L1(Ω)L^{1}(\Omega). Moreover, fH1(Ω)f\in H^{-1}(\Omega) implies that unu_{n} is bounded in H01(Ω)H^{1}_{0}(\Omega). Therefore, extracting a subsequence if necessary, there exist m^L1(Ω)\hat{m}\in L^{1}(\Omega), and u^H01(Ω)\hat{u}\in H^{1}_{0}(\Omega) such that

mnm^ in L1(Ω),unu^ in H01(Ω).m_{n}\rightharpoonup\hat{m}\ \hbox{ in }L^{1}(\Omega),\qquad u_{n}\rightharpoonup\hat{u}\ \hbox{ in }H^{1}_{0}(\Omega). (4.16)

From these convergences, the Rellich-Kondrachov compactness theorem, the lower semicontinuity of ψ\psi, and Fatou’s Lemma, we deduce

=limnΩ(j(x,un)+ψ(mn))𝑑x=lim infnΩ(j(x,un)a+c|un|2+ψ(mn))𝑑xΩ(a+c|u|2)𝑑xΩ(j(x,u^)+ψ(m^))𝑑x.\begin{split}{\cal I}&=\lim_{n\to\infty}\int_{\Omega}\big{(}j(x,u_{n})+\psi(m_{n})\big{)}\,dx\\ &=\liminf_{n\to\infty}\int_{\Omega}\big{(}j(x,u_{n})-a+c|u_{n}|^{2}+\psi(m_{n})\big{)}\,dx-\int_{\Omega}\big{(}-a+c|u|^{2})\,dx\\ &\geq\int_{\Omega}\big{(}j(x,\hat{u})+\psi(\hat{m})\big{)}\,dx.\end{split}

Proving then that u^\hat{u} is the solution of (4.5), we will deduce that m^\hat{m} is a solution of (4.1). For this purpose, given vH01(Ω)L(Ω)v\in H^{1}_{0}(\Omega)\cap L^{\infty}(\Omega), and l>0l>0, we take Sl(un)vS_{l}(u_{n})v as test function in the equation satisfied by unu_{n}. This gives

1lΩ|un|2vsgn(un)𝑑x+ΩunvSl(un)𝑑x+ΩmnunSl(un)v𝑑x=f,Sl(un)v.-{1\over l}\int_{\Omega}|\nabla u_{n}|^{2}v\operatorname{sgn}(u_{n})\,dx+\int_{\Omega}\nabla u_{n}\cdot\nabla v\,S_{l}(u_{n})\,dx+\int_{\Omega}m_{n}u_{n}S_{l}(u_{n})v\,dx=\langle f,S_{l}(u_{n})v\rangle.

By (4.16), the Rellich-Kondrachov compactness theorem, and Sl(un)vS_{l}(u_{n})v bounded in L(Ω)L^{\infty}(\Omega), we can pass to the limit in nn, in the three last terms in this equality. In the first term, we can use that unu_{n} is bounded in H01(Ω)H^{1}_{0}(\Omega) and that vv belongs to L(Ω)L^{\infty}(\Omega). Thus, there exists C>0C>0 independent of mm such that

|Ω(u^v+m^u^v)Sl(u^)𝑑xf,Sl(u^)v|Cl.\left|\int_{\Omega}\big{(}\nabla\hat{u}\cdot\nabla v+\hat{m}\hat{u}v)S_{l}(\hat{u})\,dx-\langle f,S_{l}(\hat{u})v\rangle\right|\leq{C\over l}.

Using that Sl(u^)vS_{l}(\hat{u})v converges strongly to vv in H01(Ω)H^{1}_{0}(\Omega) as ll\to\infty, and the Lebesgue dominated convergence theorem, we can pass to the limit as ll\to\infty in this equality to get

Ω(u^v+m^u^v)𝑑x=f,v,vH01(Ω)L(Ω).\int_{\Omega}\big{(}\nabla\hat{u}\cdot\nabla v+\hat{m}\hat{u}v)\,dx=\langle f,v\rangle,\quad\forall\,v\in H^{1}_{0}(\Omega)\cap L^{\infty}(\Omega).

If vv is just in H01(Ω)Lm2(Ω)H^{1}_{0}(\Omega)\cap L^{2}_{m}(\Omega), we prove that this equality also holds true just replacing vv by Tk(v)T_{k}(v) and then passing to the limit as kk\to\infty. ∎

Proof of Theorem 4.2.

Let m^\hat{m} be a solution of (4.1) and define u^\hat{u}, z^\hat{z} as the solutions of (4.5) and (4.6) respectively. Since ff satisfies (4.2), Stampacchia’s estimates (see [25]) show that u^\hat{u} belongs to H01(Ω)L(Ω)H^{1}_{0}(\Omega)\cap L^{\infty}(\Omega). By (4.3) we then have that z^\hat{z} solution of (4.6) is well defined and belongs to H01(Ω)Lm^2(Ω)H^{1}_{0}(\Omega)\cap L^{2}_{\hat{m}}(\Omega).

Now, for another non-negative function mL1(Ω)m\in L^{1}(\Omega) such that ψ(m)\psi(m) is integrable, k>0k>0, and ε(0,1]\varepsilon\in(0,1], we define mεm_{\varepsilon} as

mε=m^+εTk(mm^),m_{\varepsilon}=\hat{m}+\varepsilon\,T_{k}(m-\hat{m}),

and uεu_{\varepsilon} as the solution of

{Δuε+mεuε=fin Ωuε=0on Ω.\begin{cases}-\Delta u_{\varepsilon}+m_{\varepsilon}u_{\varepsilon}=f&\hbox{in }\Omega\\ u_{\varepsilon}=0&\hbox{on }\partial\Omega.\end{cases}

Taking into account that

mε={(1ε)m^+εmif |mm^|k(1εk|mm^|)m^+εk|mm^|mif |mm^|>k,m_{\varepsilon}=\begin{cases}(1-\varepsilon)\hat{m}+\varepsilon m&\hbox{if }|m-\hat{m}|\leq k\\ \displaystyle\Big{(}1-{\varepsilon k\over|m-\hat{m}|}\Big{)}\hat{m}+{\varepsilon k\over|m-\hat{m}|}m&\hbox{if }|m-\hat{m}|>k,\end{cases}

that m^\hat{m} is a solution of (4.1), and the convexity of ψ\psi, we deduce

Ωj(x,uε)+{|mm^|k}((1ε)ψ(m^)+εψ(m))𝑑x+{|mm^|>k}((1εk|mm^|)ψ(m^)+εk|mm^|ψ(m))𝑑xΩ(j(x,uε)+ψ(mε))𝑑xΩ(j(x,u^)+ψ(m^))𝑑x,\begin{split}&\int_{\Omega}j(x,u_{\varepsilon})+\int_{\{|m-\hat{m}|\leq k\}}\hskip-8.0pt\big{(}(1-\varepsilon)\psi(\hat{m})+\varepsilon\,\psi(m)\big{)}\,dx\\ &\qquad+\int_{\{|m-\hat{m}|>k\}}\hskip-8.0pt\Big{(}\Big{(}1-{\varepsilon k\over|m-\hat{m}|}\Big{)}\psi(\hat{m})+{\varepsilon k\over|m-\hat{m}|}\psi(m)\Big{)}\,dx\\ &\qquad\geq\int_{\Omega}\big{(}j(x,u_{\varepsilon})+\psi(m_{\varepsilon})\big{)}\,dx\geq\int_{\Omega}\big{(}j(x,\hat{u})+\psi(\hat{m})\big{)}dx,\end{split}

for every ε>0\varepsilon>0. Using then that

uεuεu in H01(Ω),{u_{\varepsilon}-u\over\varepsilon}\rightarrow u^{\prime}\ \hbox{ in }H^{1}_{0}(\Omega),

with uH01(Ω)L(Ω)u^{\prime}\in H^{1}_{0}(\Omega)\cap L^{\infty}(\Omega) the solution of

{Δu+m^u+Tk(mm^)u^=0in Ωu=0on Ω,\begin{cases}-\Delta u^{\prime}+\hat{m}u^{\prime}+T_{k}(m-\hat{m})\hat{u}=0&\hbox{in }\Omega\\ u^{\prime}=0&\hbox{on }\partial\Omega,\end{cases}

we get

Ωsj(x,u^)udx+{|mm^|k}(ψ(m)ψ(m^))𝑑x+{|mm^|>k}k(ψ(m)ψ(m^))|mm^|𝑑x0.\begin{split}\int_{\Omega}\partial_{s}j(x,\hat{u})u^{\prime}dx+\int_{\{|m-\hat{m}|\leq k\}}\hskip-10.0pt\big{(}\psi(m)-\psi(\hat{m})\big{)}\,dx\\ \qquad+\int_{\{|m-\hat{m}|>k\}}\hskip-10.0pt{k\big{(}\psi(m)-\psi(\hat{m})\big{)}\over|m-\hat{m}|}\,dx\geq 0.\end{split} (4.17)

On the other hand, taking uu^{\prime} as test function in (4.6), and z^\hat{z} as test function in (4.5) we deduce

Ωsj(x,u)udx=Ω(z^u+m^z^u)𝑑x=ΩTk(mm^)u^z^𝑑x.\int_{\Omega}\partial_{s}j(x,u)u^{\prime}\,dx=\int_{\Omega}\big{(}\nabla\hat{z}\cdot\nabla u^{\prime}+\hat{m}\hat{z}u^{\prime}\big{)}\,dx=-\int_{\Omega}T_{k}(m-\hat{m})\hat{u}\hat{z}\,dx.

Therefore (4.17) provides

ΩTk(mm^)u^z^𝑑x+{|mm^|k}(ψ(m)ψ(m^))𝑑x+{|mm^|>k}k(ψ(m)ψ(m^))|mm^|𝑑x0.-\int_{\Omega}T_{k}(m-\hat{m})\hat{u}\hat{z}\,dx+\int_{\{|m-\hat{m}|\leq k\}}\hskip-10.0pt\big{(}\psi(m)-\psi(\hat{m})\big{)}\,dx+\int_{\{|m-\hat{m}|>k\}}\hskip-10.0pt{k\big{(}\psi(m)-\psi(\hat{m})\big{)}\over|m-\hat{m}|}\,dx\geq 0.

Using here that mm, m^\hat{m}, ψ(m)\psi(m) and ψ(m^)\psi(\hat{m}) belong to L1(Ω)L^{1}(\Omega) and that the second assertion in (4.3), z^\hat{z} solution of (4.6) and Stampacchia’s estimates imply that z^\hat{z} is in L(Ω)L^{\infty}(\Omega), we can take the limit as kk\to\infty to deduce

Ω(ψ(m)mu^z^)𝑑xΩ(ψ(m^)m^u^z^)𝑑x,mL1(Ω),m0 a.e. in Ω,Ωψ(m)𝑑x<.\begin{split}&\int_{\Omega}\big{(}\psi(m)-m\hat{u}\hat{z}\big{)}\,dx\geq\int_{\Omega}\big{(}\psi(\hat{m})-\hat{m}\hat{u}\hat{z}\big{)}\,dx,\\ &\forall\,m\in L^{1}(\Omega),\ m\geq 0\hbox{ a.e. in }\Omega,\ \int_{\Omega}\psi(m)\,dx<\infty.\end{split}

This implies that m^\hat{m} satisfies

m^dom(ψ),ψ(m^)m^u^z^=minsdom(ψ){ψ(s)su^z^},a.e. in Ω,\hat{m}\in\operatorname{dom}(\psi),\quad\psi(\hat{m})-\hat{m}\hat{u}\hat{z}=\min_{s\in\operatorname{dom}(\psi)}\big{\{}\psi(s)-s\hat{u}\hat{z}\big{\}},\quad\hbox{a.e. in }\Omega,

or equivalently,

m^u^z^=ψ(m^)+ψ(u^z^).\hat{m}\hat{u}\hat{z}=\psi(\hat{m})+\psi^{\ast}(\hat{u}\hat{z}).

By (2.4) this is also equivalent to m^ψ(u^z^)\hat{m}\in\partial\psi^{\ast}(\hat{u}\hat{z}), and also to u^z^ψ(m^).\hat{u}\hat{z}\in\partial\psi(\hat{m}). From (2.3) applied to ψ\psi^{\ast} and Remark 4.4 we then deduce the third assertion in (4.7). Combined with u^\hat{u} and z^\hat{z} in L(Ω)L^{\infty}(\Omega), this also implies that m^\hat{m} is in L(Ω)L^{\infty}(\Omega). ∎

Proof of Proposition 4.5.

Let us prove (4.10). If ψ\psi is strictly convex, the quotient function defined by (2.2) restricted to the set

{(s,t)dom(ψ)×dom(ψ):s<t},\big{\{}(s,t)\in\operatorname{dom}(\psi)\times\operatorname{dom}(\psi)\ :\ s<t\big{\}},

is strictly increasing in ss and tt. This proves

d+ψ(s1)<dψ(s2)s1,s2dom(ψ),s1<s2.d_{+}\psi(s_{1})<d_{-}\psi(s_{2})\qquad\forall\,s_{1},s_{2}\in\operatorname{dom}(\psi),\ s_{1}<s_{2}.

By Remark 4.3, this implies that for every τ\tau\in\mathbb{R}, there exists a unique sdom(ψ)s\in\operatorname{dom}(\psi) such that τψ(s)\tau\in\partial\psi(s). By definition (4.4) of hh we deduce that h(τ)h(\tau) agrees with such ss. Now, we observe that the lower semicontinuity of ψ\psi and definition (2.1) of ψ\partial\psi imply the following continuity property for ψ\partial\psi:

sn,sdom(ψ),τnψ(sn),τ,sns,τnττψ(s).s_{n},s\in\operatorname{dom}(\psi),\ \tau_{n}\in\partial\psi(s_{n}),\ \tau\in\mathbb{R},\quad s_{n}\to s,\ \tau_{n}\to\tau\Longrightarrow\tau\in\partial\psi(s).

Thanks to the uniqueness of ss proved above, this can also be read as

sn=h(τn),sns,τnτs=h(τ),s_{n}=h(\tau_{n}),\ s_{n}\to s,\ \tau_{n}\to\tau\Longrightarrow s=h(\tau),

and then proves the continuity of hh.

For the reciprocal, we argue by contradiction. If ψ\psi is not strictly convex, then, there exists an interval [c,d]dom(ψ)[c,d]\subset\operatorname{dom}(\psi), with c<dc<d such that ψ\psi is an affine function with a certain slope λ\lambda\in\mathbb{R} in this interval. Moreover c0c\geq 0 can be chosen as

c=minsdom(ψ)d+ψ(s)=λ.c=\min_{s\in\operatorname{dom}(\psi)}d_{+}\psi(s)=\lambda.

If c=αc=\alpha, defined by (4.8), then ψ(c)=(,λ]\partial\psi(c)=(-\infty,\lambda] and thus, definition (4.4) of hh implies

h(τ)=cτ<λ,h(λ)d.h(\tau)=c\qquad\forall\,\tau<\lambda,\ h(\lambda)\geq d.

If c>αc>\alpha, then d+ψ(s)<λd_{+}\psi(s)<\lambda for every s<cs<c, and thus

h(τ)<cτ<λ,h(λ)d.h(\tau)<c\qquad\forall\,\tau<\lambda,\ h(\lambda)\geq d.

In both cases, we conclude that hh is not continuous at τ=λ\tau=\lambda.

In order to prove (4.11) we first observe that the right-hand side implies ψ\psi strictly convex. Since hLiploc()h\in{\rm Lip}_{\rm loc}(\mathbb{R}) implies hh continuous, we conclude that the left-hand side in (4.11) also implies ψ\psi strictly convex. Therefore it is enough to prove the result for ψ\psi strictly convex. As we saw at the beginning of the proof, this implies that for every τ\tau\in\mathbb{R} there exists a unique sdom(ψ)s\in\operatorname{dom}(\psi) such that τψ(s)\tau\in\partial\psi(s), and this ss satisfies h(τ)=s.h(\tau)=s. Then, taking into account (2.3), we deduce the existence of L>0L>0 such that

h(τ2)h(τ1)L(τ2τ1),h(\tau_{2})-h(\tau_{1})\leq L(\tau_{2}-\tau_{1}),

for every τ1τ2\tau_{1}\leq\tau_{2}, is equivalent to

s2s1L(dψ(s1)d+ψ(s2)),s_{2}-s_{1}\leq L\big{(}d_{-}\psi(s_{1})-d_{+}\psi(s_{2})\big{)},

for every s1,s2dom(ψ)s_{1},s_{2}\in\operatorname{dom}(\psi), and then, that (4.11) holds. ∎

Proof of Proposition 4.6.

Since d+ψ(α)>d_{+}\psi(\alpha)>-\infty, we have ψ(α)=(,d+ψ(α)]\partial\psi(\alpha)=(-\infty,d_{+}\psi(\alpha)], and then h(s)=αh(s)=\alpha for every s<d+ψ(α)s<d_{+}\psi(\alpha). Therefore, if hh is derivable at d+(ψ(α))d_{+}(\psi(\alpha)), we must have

h(d+ψ(α))=α,limε0h(d+ψ(α)+ε)αε=0.h(d_{+}\psi(\alpha))=\alpha,\qquad\lim_{\varepsilon\searrow 0}{h(d_{+}\psi(\alpha)+\varepsilon)-\alpha\over\varepsilon}=0.

Thus, for every ρ>0\rho>0, there exists δ>0\delta>0 such that 0<ε<δ0<\varepsilon<\delta implies

h(d+ψ(α)+ε)<α+ερ,h(d_{+}\psi(\alpha)+\varepsilon)<\alpha+\varepsilon\rho,

which by (4.9) can also be read as

sdom(ψ),dψ(s)d+ψ(α)+εs<α+ερ,s\in\operatorname{dom}(\psi),\ d_{-}\psi(s)\leq d_{+}\psi(\alpha)+\varepsilon\Longrightarrow s<\alpha+\varepsilon\rho,

and then that

sα+ερdψ(s)>d+ψ(α)+ε.s\geq\alpha+\varepsilon\rho\Rightarrow d_{-}\psi(s)>d_{+}\psi(\alpha)+\varepsilon.

Taking s=α+ερs=\alpha+\varepsilon\rho and letting ε0\varepsilon\to 0, this gives

limε0dψ(α+ερ)d+ψ(α)ερ>1ρρ>0,\lim_{\varepsilon\searrow 0}{d_{-}\psi(\alpha+\varepsilon\rho)-d_{+}\psi(\alpha)\over\varepsilon\rho}>{1\over\rho}\qquad\forall\,\rho>0,

and then

limε0dψ(α+ε)d+ψ(α)ε=.\lim_{\varepsilon\searrow 0}{d_{-}\psi(\alpha+\varepsilon)-d_{+}\psi(\alpha)\over\varepsilon}=\infty.

The proof of the reciprocal follows by a similar argument. ∎

Proof of Theorem 4.7.

Taking into account that m^\hat{m} belongs to L(Ω)L^{\infty}(\Omega), that Ω\Omega is C1,1C^{1,1}, fLq(Ω)f\in L^{q}(\Omega), the first assertion in (4.14) and u^\hat{u}, z^\hat{z} solutions of (4.5) and (4.6) respectively, we can apply the regularity results for elliptic equations (see e.g. [23], chapter 7) to deduce that u^\hat{u}, z^\hat{z} belong to W2,q(Ω)W^{2,q}(\Omega).

Now, we use that u^z^\hat{u}\hat{z} satisfies

Δ(u^z^)+m^u^z^=2u^z^+fz^+sj(x,u^)u^in Ω.-\Delta(\hat{u}\hat{z})+\hat{m}\hat{u}\hat{z}=-2\nabla\hat{u}\cdot\nabla\hat{z}+f\hat{z}+\partial_{s}j(x,\hat{u})\hat{u}\qquad\hbox{in }\Omega.

where thanks to (4.14), f^BV(Ω)Lq(Ω)\hat{f}\in BV(\Omega)\cap L^{q}(\Omega) and u^,z^W2,q(Ω)\hat{u},\hat{z}\in W^{2,q}(\Omega), we have that the right-hand side of this equation belongs to BV(Ω)BV(\Omega). Using then that (4.7) and definition (4.13) of gg imply

g(u^z^)m^u^z^g(u^z^)a.e. in Ω,g_{-}(\hat{u}\hat{z})\leq\hat{m}\hat{u}\hat{z}\leq g_{-}(\hat{u}\hat{z})\qquad\hbox{a.e. in }\Omega,

and that we are assuming gg non-decreasing in \mathbb{R}, we can apply Theorem 3.6 to deduce that m^u^z^\hat{m}\hat{u}\hat{z} belongs to BV(Ω)BV(\Omega). ∎

5 Some examples

In the present section we illustrate the results obtained in the previous one, by applying them to some classical examples. In Section 6 we will also perform some numerical computations relative to these examples.

First example. We consider the case where we are looking for a non-negative function mm such that the solution uu of the state equation in (4.1) minimizes

J(u):=Ωj(x,u)𝑑x,J(u):=\int_{\Omega}j(x,u)\,dx, (5.1)

and it is such that the norm of mm in some space Lp(Ω)L^{p}(\Omega), 1<p<1<p<\infty is not too large. This can be modeled by (4.1), with ψ\psi given by

ψ(s)=1(,0)+ksp1[0,).\psi(s)=\infty 1_{(-\infty,0)}+ks^{p}1_{[0,\infty)}.

with kk a positive to parameter to choose. Then,

ψ(s)={(,0]if s=0,{kpsp1}if s>0,h(τ)=(τkp)1p11(0,).\partial\psi(s)=\begin{cases}(-\infty,0]&\hbox{if }s=0,\\ \{kp\,s^{p-1}\}&\hbox{if }s>0,\end{cases}\qquad h(\tau)=\Big{(}{\tau\over kp}\Big{)}^{1\over p-1}1_{(0,\infty)}.

By Theorem 4.1 and by the fact that ψ\psi is strictly convex, we have that hh is continuous. It is in C1()C^{1}(\mathbb{R}) when p<2p<2, i.e. when condition (4.12) holds, and it is locally Lipschitz continuous if p2p\leq 2, (and then condition (4.11) holds in bounded subsets of dom(ψ)\operatorname{dom}(\psi)).

From (4.7), we deduce that, taking jj and ff regular enough, every solution m^\hat{m} of (4.1) satisfies

m^=(u^z^kp)1p11{u^z^>0}a.e. in Ω,\hat{m}=\Big{(}{\hat{u}\hat{z}\over kp}\Big{)}^{1\over p-1}1_{\{\hat{u}\hat{z}>0\}}\qquad\hbox{a.e. in }\Omega, (5.2)

with u^\hat{u} and z^\hat{z} the solutions of (4.5) and (4.6) respectively. In particular, this means that u^\hat{u}, z^\hat{z} solve the nonlinear system

{Δu^+(u^z^kp)1p11{u^z^>0}u^=fin Ω,Δz^+(u^z^kp)1p11{u^z^>0}z^=sj(x,u^)in Ω,u^=z^=0on Ω.\begin{cases}\displaystyle-\Delta\hat{u}+\Big{(}{\hat{u}\hat{z}\over kp}\Big{)}^{1\over p-1}1_{\{\hat{u}\hat{z}>0\}}\hat{u}=f&\hbox{in }\Omega,\\ \displaystyle-\Delta\hat{z}+\Big{(}{\hat{u}\hat{z}\over kp}\Big{)}^{1\over p-1}1_{\{\hat{u}\hat{z}>0\}}\hat{z}=\partial_{s}j(x,\hat{u})&\hbox{in }\Omega,\\ \displaystyle\hat{u}=\hat{z}=0&\hbox{on }\partial\Omega.\end{cases} (5.3)

By Theorem 4.7 we obtain that

m^u^z^=(u^z^)p(kp)1p1 1{u^z^>0}\hat{m}\hat{u}\hat{z}={\big{(}\hat{u}\hat{z}\big{)}^{p^{\prime}}\over(kp)^{1\over p-1}}\,1_{\{\hat{u}\hat{z}>0\}}

belongs to BV(Ω)BV(\Omega) (and even to W1,1(Ω)W^{1,1}(\Omega)). However, in order to have m^\hat{m} in W1,1(Ω)W^{1,1}(\Omega), we need (u^z^)2pp1(u^z^)1{u^z^}(\hat{u}\hat{z})^{2-p\over p-1}\nabla(\hat{u}\hat{z})1_{\{\hat{u}\hat{z}\}} to be in L1(Ω)NL^{1}(\Omega)^{N}, which is not clear for p>2p>2, i.e. when hh is not locally Lipschitz.

Second example. We now consider the case where we want to minimize functional (5.1), with uu the solution of the state equation in (4.1), under the constraint m[α,β]m\in[\alpha,\beta], with 0α<β0\leq\alpha<\beta. The problem corresponds to (4.1), with ψ\psi given by

ψ(s)=1(,α)(β,).\psi(s)=\infty 1_{(-\infty,\alpha)\cup(\beta,\infty)}. (5.4)

Thus, dom(ψ)=[α,β]\operatorname{dom}(\psi)=[\alpha,\beta], and

ψ(s)={(,0]if s=α{0}if α<s<β[0,)if s=β,h(τ)={αif τ<0βif τ0,\partial\psi(s)=\begin{cases}(-\infty,0]&\hbox{if }s=\alpha\\ \{0\}&\hbox{if }\alpha<s<\beta\\ [0,\infty)&\hbox{if }s=\beta,\end{cases}\qquad h(\tau)=\begin{cases}\alpha&\hbox{if }\tau<0\\ \beta&\hbox{if }\tau\geq 0,\end{cases}

As expected, since ψ\psi is not strictly convex, we get by (4.10) that hh is not continuous. Condition (4.7) reads in this case as

{m^=αa.e. in {u^z^<0}m^=βa.e. in {u^z^>0}m^[α,β]a.e. in {u^z^=0}.\begin{cases}\hat{m}=\alpha&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}<0\}\\ \hat{m}=\beta&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}>0\}\\ \hat{m}\in[\alpha,\beta]&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}=0\}.\end{cases} (5.5)

Therefore, the value of m^\hat{m} is not determined on the set where u^z^\hat{u}\hat{z} vanishes. When this set has zero measure, assertion (5.5) shows that m^\hat{m} is a bang-bang control, which only takes the values α\alpha and β\beta. Taking into account that g(s)=h(s)sg(s)=h(s)s is non-decreasing, we deduce from Theorem 4.7 that m^u^z^\hat{m}\hat{u}\hat{z} is in BV(Ω)BV(\Omega).

As a simple case where we can assure that m^\hat{m} is a bang-bang control, we take

j(x,s)=γ(x)sa.e. xΩ,s,j(x,s)=\gamma(x)s\qquad\hbox{a.e. }x\in\Omega,\ \forall\,s\in\mathbb{R}, (5.6)

Assuming γ,fLq(Ω)\gamma,f\in L^{q}(\Omega), with qq satisfying (4.15), we have that u^,z^\hat{u},\hat{z} are in W2,q(Ω)W^{2,q}(\Omega). Therefore, (4.5) and (4.6) imply

f=0 a.e. in {u^=0},γ=0 a.e. in {z^=0}.f=0\ \hbox{ a.e. in }\{\hat{u}=0\},\qquad\gamma=0\ \hbox{ a.e. in }\{\hat{z}=0\}.

Thus, assuming |{f=0}|=|{γ=0}|=0,|\{f=0\}|=|\{\gamma=0\}|=0, we conclude that the set {u^z^=0}\{\hat{u}\hat{z}=0\} has zero measure.

It is also simple to give a counterexample where the set {u^z^=0}\{\hat{u}\hat{z}=0\} has positive measure and the control m^\hat{m} is not a bang-bang control. Just take m~C0(Ω¯;[α,β])\tilde{m}\in C^{0}(\overline{\Omega};[\alpha,\beta]), not constant, and u~\tilde{u} the solution of (4.5), with m^\hat{m} replaced by m~\tilde{m}. Defining

j(x,s)=|su~(x)|2a.e. xΩ,s,j(x,s)=|s-\tilde{u}(x)|^{2}\qquad\hbox{a.e. }x\in\Omega,\ \forall\,s\in\mathbb{R},

we deduce that problem (4.1) has the unique solution m^=m~\hat{m}=\tilde{m}, for which the functional vanishes. Observe that in this case sj(x,u^)=0\partial_{s}j(x,\hat{u})=0 a.e. in Ω\Omega, and then z^\hat{z} is the null function. Therefore the set {u^z^=0}\{\hat{u}\hat{z}=0\} is the whole set Ω\Omega, and condition (5.5) does not provide any information about m^\hat{m}.

Third example. We consider a mixture of the first and second examples. Now, the goal is to minimize (5.1), with uu the solution of the state equation in (4.1) and m[α,β]m\in[\alpha,\beta], such that its norm in Lp(Ω)L^{p}(\Omega) is not too great, with 1p<1\leq p<\infty. The problem corresponds to take in (4.1)

ψ(s)=1(,α)(β,)+ksp1[α,β],\psi(s)=\infty 1_{(-\infty,\alpha)\cup(\beta,\infty)}+ks^{p}1_{[\alpha,\beta]},

with k>0k>0 a positive constant to choose.

In the strictly convex case p>1p>1, we have

ψ(s)={(,kpαp1]if s=α{kpsp1}if α<s<β[kpβp1,)if s=β,h(τ)={αif τ<kpαp1(τkp)1p1if kpαp1τ<kpβp1βif τkpβp1.\partial\psi(s)\hskip-2.0pt=\hskip-2.0pt\begin{cases}(-\infty,kp\alpha^{p-1}]&\hbox{if }s=\alpha\\ \{kps^{p-1}\}&\hbox{if }\alpha<s<\beta\\ [kp\beta^{p-1},\infty)&\hbox{if }s=\beta,\end{cases}\qquad h(\tau)\hskip-2.0pt=\hskip-2.0pt\begin{cases}\alpha&\hbox{if }\tau<kp\alpha^{p-1}\\ \displaystyle\Big{(}{\tau\over kp}\Big{)}^{1\over p-1}&\hbox{if }kp\alpha^{p-1}\leq\tau<kp\beta^{p-1}\\ \beta&\hbox{if }\tau\geq kp\beta^{p-1}.\end{cases}

As in the first example, the strict convexity of ψ\psi provides a function hh which is continuous. Therefore, the optimal controls are continuous and even, they are in some Sobolev space if p2p\leq 2 (assuming jj and ff smooth enough).

In the case p=1p=1, we have

ψ(s)={(,k]if s=α{k}if α<s<β[k,)if s=β,h(τ)={αif τ<kβif τk.\partial\psi(s)=\begin{cases}(-\infty,k]&\hbox{if }s=\alpha\\ \{k\}&\hbox{if }\alpha<s<\beta\\ [k,\infty)&\hbox{if }s=\beta,\end{cases}\qquad h(\tau)=\begin{cases}\alpha&\hbox{if }\tau<k\\ \beta&\hbox{if }\tau\geq k.\end{cases}

As in the second example, (4.7) provides

{m^=αa.e. in {u^z^<k}m^=βa.e. in {u^z^>k}m^[α,β]a.e. in {u^z^=k},\begin{cases}\hat{m}=\alpha&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}<k\}\\ \hat{m}=\beta&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}>k\}\\ \hat{m}\in[\alpha,\beta]&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}=k\},\end{cases}

and then the optimal controls are bang-bang controls if the set {u^z^=k}\{\hat{u}\hat{z}=k\} has null measure.

Since h(τ)τh(\tau)\tau is still a non-decreasing, Theorem 4.7 proves that m^u^z^\hat{m}\hat{u}\hat{z} is in BV(Ω)BV(\Omega) for every optimal control m^\hat{m}. Using also that m^=α\hat{m}=\alpha in a neighborhood of the closed set {u^z^=0}\{\hat{u}\hat{z}=0\} we deduce that in this case m^\hat{m} is also in BV(Ω)BV(\Omega).

As a particular case we can take

j(x,s)=f(x)s.j(x,s)=f(x)s. (5.7)

This is a classical problem corresponding to the minimization of the energy. In this case control problem (4.1) can also be written in the simplest form

maxαmβminuH01(Ω){Ω(|u|2+mu22fu)𝑑xkΩm𝑑x}.\max_{\alpha\leq m\leq\beta}\min_{u\in H^{1}_{0}(\Omega)}\left\{\int_{\Omega}\Big{(}|\nabla u|^{2}+mu^{2}-2fu\Big{)}dx-k\int_{\Omega}m\,dx\right\}. (5.8)

For k=0k=0, it is clear that the solution corresponds to m^=β\hat{m}=\beta. Thus, the interesting case corresponds (as we assumed above) to k>0k>0. This means that we just want to spend a limited amount of the optimal potential β\beta (for example, because it is more expensive).

From (5.7) we get z^=u^\hat{z}=\hat{u}, and then (4.7) gives

{m^=αa.e. in {|u^|2<k}m^=βa.e. in {|u^|2>k}.\begin{cases}\hat{m}=\alpha&\hbox{a.e. in }\{|\hat{u}|^{2}<k\}\\ \hat{m}=\beta&\hbox{a.e. in }\{|\hat{u}|^{2}>k\}.\end{cases} (5.9)

Taking fLq(Ω)f\in L^{q}(\Omega), with qq satisfying (4.15), we have u^\hat{u} in W2,q(Ω)W^{2,q}(\Omega), and then (4.5) provides

km^=f a.e. in {u^=k},km^=f a.e. in {u^=k}.\sqrt{k}\hat{m}=f\ \hbox{ a.e. in }\big{\{}\hat{u}=\sqrt{k}\big{\}},\qquad-\sqrt{k}\hat{m}=f\ \hbox{ a.e. in }\big{\{}\hat{u}=-\sqrt{k}\big{\}}.

Therefore, a sufficient condition to have m^\hat{m} a bang-bang control is to assume that the set {f[kβ,kα][kα,kβ]}\{f\in[-\sqrt{k}\beta,-\sqrt{k}\alpha]\cup[\sqrt{k}\alpha,\sqrt{k}\beta]\} has null measure. This holds in particular if α\alpha is positive, ff belongs to L(Ω)L^{\infty}(\Omega) and kk is large enough.

Fourth example. Related to the case p=1p=1 in the third example, let us take

ψ(s)=1(,α)(β,)ks1[α,β],\psi(s)=\infty 1_{(-\infty,\alpha)\cup(\beta,\infty)}-ks1_{[\alpha,\beta]},

with 0α<β0\leq\alpha<\beta, k>0k>0. In this case we are interested in controls mm which take their values in [α,β][\alpha,\beta] and its integral is large. Similarly to the third example, we have

ψ(s)={(,k]if s=α{k}if α<s<β[k,)if s=β,h(τ)={αif τ<kβif τk,\partial\psi(s)=\begin{cases}(-\infty,-k]&\hbox{if }s=\alpha\\ \{-k\}&\hbox{if }\alpha<s<\beta\\ [-k,\infty)&\hbox{if }s=\beta,\end{cases}\qquad h(\tau)=\begin{cases}\alpha&\hbox{if }\tau<-k\\ \beta&\hbox{if }\tau\geq-k,\end{cases}

Theorem 4.1 proves that the optimal controls satisfy

{m^=αa.e. in {u^z^<k}m^=βa.e. in {u^z^>k}m^[α,β]a.e. in {u^z^=k},\begin{cases}\hat{m}=\alpha&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}<-k\}\\ \hat{m}=\beta&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}>-k\}\\ \hat{m}\in[\alpha,\beta]&\hbox{a.e. in }\big{\{}\hat{u}\hat{z}=-k\},\end{cases}

and then they are not continuous in general. Moreover, in this case the function

g(τ)=h(τ)τ={ατif τ<kβτif τk,g(\tau)=h(\tau)\tau=\begin{cases}\alpha\tau&\hbox{if }\tau<-k\\ \beta\tau&\hbox{if }\tau\geq-k,\end{cases}

decreases at τ=k\tau=-k. Thus, this is an example where Theorem 4.7 does not apply and therefore, we do not know if m^u^z^\hat{m}\hat{u}\hat{z} is in BV(Ω)BV(\Omega).

A classical example corresponds to the maximization of the energy i.e. (compare with (5.7))

F(x,s)=f(x)s.F(x,s)=-f(x)s. (5.10)

Similarly to (5.8), it is known that the problem can also be written as

minαmβminuH01(Ω){Ω(|u|2+mu22fu)𝑑xkΩm𝑑x}.\min_{\alpha\leq m\leq\beta}\min_{u\in H^{1}_{0}(\Omega)}\left\{\int_{\Omega}\Big{(}|\nabla u|^{2}+mu^{2}-2fu\Big{)}dx-k\int_{\Omega}m\,dx\right\}. (5.11)

For k=0k=0 the solution is given by m^=α\hat{m}=\alpha and then the interesting case is k>0k>0, as assumed above. Now, the function z^\hat{z} is equal to u^-\hat{u}, and therefore m^\hat{m} satisfies (compare with (5.9)

{m^=αa.e. in {k<|u^|2}m^=βa.e. in {|u^|2<k}.\begin{cases}\hat{m}=\alpha&\hbox{a.e. in }\big{\{}k<|\hat{u}|^{2}\}\\ \hat{m}=\beta&\hbox{a.e. in }\big{\{}|\hat{u}|^{2}<k\}.\end{cases}

Thus, a suffcient condition to assure that the optimal controls only take the values α\alpha and β\beta is also to asume that the set {f[kβ,kα][kα,kβ]}\{f\in[-\sqrt{k}\beta,-\sqrt{k}\alpha]\cup[\sqrt{k}\alpha,\sqrt{k}\beta]\} has null measure.

6 Some numerical simulations

In this section, we illustrate the results of the previous ones through the numerical resolution, in the 2D case, of problem (4.1) for the first three examples in Section 5. For the first example we will consider p=2p=2 and for the third one p=1p=1. We apply a gradient descent method with projection. It depends of the function ψ\psi associated to the volume constraint of the potential. The corresponding algorithm is related to Theorem 4.2 providing the optimality conditions to (4.1). We refer to [1], [15] for similar algorithms in optimal design problems. It reads as follows.

  • Initialization: choose an admisible function m0L1(Ω)m_{0}\in L^{1}(\Omega), such that Ψ(m0)<\Psi(m_{0})<\infty.

  • For j0j\geq 0, iterate until stop condition as follows.

    • Compute uj,zju_{j},z_{j} solution of (4.5), (4.6), for m^=mj\hat{m}=m_{j}.

    • Compute m~j\tilde{m}_{j} descent direction associated to uju_{j} and zjz_{j}, as:

      m~j={ujzj2kmjujzj2kmjL2(Ω)in the first example,sgn(ujzj)in the second example,sgn(kujzj)in the third example.\tilde{m}_{j}=\begin{cases}\displaystyle{u_{j}z_{j}-2km_{j}\over\|u_{j}z_{j}-2km_{j}\|_{L^{2}(\Omega)}}&\hbox{in the first example,}\\ \operatorname{sgn}(u_{j}z_{j})&\hbox{in the second example,}\\ \operatorname{sgn}(k-u_{j}z_{j})&\hbox{in the third example.}\end{cases}
    • Update the function mjm_{j}:

      mj+1=Pψ(mj+ϵjm~j)m_{j+1}=P_{\psi}(m_{j}+\epsilon_{j}\tilde{m}_{j})

      where PψP_{\psi} is the projection operator from \mathbb{R} into the domain of ψ\psi, i.e.

      Pψ(s)={s+in the first example,min{β,max{s,α}}in the second and third examples.P_{\psi}(s)=\begin{cases}s^{+}&\hbox{in the first example,}\\ \min\{\beta,\max\{s,\alpha\}\}&\hbox{in the second and third examples.}\end{cases}
  • Stop if |I(mj)I(mj1)||I(m0)|<tol\frac{|I(m_{j})-I(m_{j-1)}|}{|I(m_{0})|}<tol, for tol>0tol>0 small.

In all the simulations we have chosen Ω\Omega as the ball BB of center zero and radius one in dimension two. For the first example, we have chosen a simple case where the solution is radial. Thus, we just solve the corresponding one-dimensional problem. The implementation for this example has been carried out using Matlab R2022a. The second and third examples have been implemented using the free software FreeFemm++ v 4.4-3 (see [19] and http://www.freefem.org/). The corresponding results are as follows.

First example. For s0s_{0}\in\mathbb{R}, we take:

j(x,s)=12|ss0|2,ψ(s)=1(,0)+k|s|2,f=1.j(x,s)=\frac{1}{2}\big{|}s-s_{0}\big{|}^{2},\qquad\psi(s)=\infty 1_{(-\infty,0)}+k|s|^{2},\qquad f=1.

Taking mm as the null function, the solution of the state equation in (4.1) is

u(x)=1|x|24.u(x)={1-|x|^{2}\over 4}.

Thus, the interesting case corresponds to s0(0,1/4)s_{0}\in(0,1/4). When k=0k=0 assumption (2.5) is not satisfied, and the optimal control is not necessarily given by a function m^L1(Ω)\hat{m}\in L^{1}(\Omega) (we refer to [6], [8], [9], [10], [11], [12], [13] for some other existence results on optimal potentials). Indeed, it can be proved that the optimal control m^\hat{m} is given by the Radon measure

m^=1s01{|x|<a}dx+4s01+a2(12loga)4s0loga1{|x|=a}dσ,\hat{m}={1\over s_{0}}1_{\{|x|<a\}}dx+{4s_{0}-1+a^{2}(1-2\log a)\over 4s_{0}\log a}1_{\{|x|=a\}}d\sigma,

where dσd\sigma is the 1-dimensional measure and aa is characterized by

0<a<1,4s01+a2<2a2loga,0<a<1,\quad 4s_{0}-1+a^{2}<2a^{2}\log a,
logaa1r(4s01+r2)logrdr=(4s01+a2)a1rlog2rdr.\log a\int_{a}^{1}r(4s_{0}-1+r^{2})\log r\,dr=(4s_{0}-1+a^{2}\Big{)}\int_{a}^{1}r\log^{2}r\,dr.

The corresponding optimal control u^\hat{u} is given by

u^(x)={s0if |x|<a1|x|24+(4s01+a2)log|x|4logaif a|x|1.\hat{u}(x)=\begin{cases}s_{0}&\hbox{if }|x|<a\\ \displaystyle{1-|x|^{2}\over 4}+{(4s_{0}-1+a^{2})\log|x|\over 4\log a}&\hbox{if }a\leq|x|\leq 1.\end{cases}

In Figure 1 we represent the optimal state function u^\hat{u} for s0=0.1s_{0}=0.1 (then a0.2825a\sim 0.2825)

In Figure 2 we represent on the top the optimal control and optimal state function obtained by solving the above problem (corresponding to k=0k=0) numerically. In the middle and the bottom we also represent the optimal control and optimal state function, taking k=107k=10^{-7} and k=105k=10^{-5} respectively. Since the solutions are radial functions, we have applied the algorithm to the corresponding one-dimensional problem. Observe that in this radial representation the singular part of the optimal measure m^\hat{m} is given by two Dirac masses at r=ar=-a and r=ar=a. In the numerical computation this provides the two corners which appear in the figure.

Refer to caption
Figure 1: Example 1: Optimal state u^\hat{u} for k=0k=0 (analytic solution).
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: First example: Optimal control m^\hat{m} (left), and optimal state u^\hat{u} (right). Top: k=0k=0. Middle: k=105k=10^{-5}. Bottom: k=107k=10^{-7}.

Second example. As a particular case of the second example in the previous section, we take

j(x,s)=g(x)s,ψ(s)=1(,α)(β,+),j(x,s)=g(x)s,\qquad\psi(s)=\infty 1_{(-\infty,\alpha)\cup(\beta,+\infty),}

with

g(x1,x2)=x12x22.g(x_{1},x_{2})=x_{1}^{2}-x_{2}^{2}.

The right-hand side in the state equation in (4.1) is given by

f(x1,x2)=10(x12+x22)sin(13arctan|x2x1|)1{|x1|>1010}.f(x_{1},x_{2})=10(x_{1}^{2}+x_{2}^{2})\sin\Big{(}13\arctan\Big{|}\frac{x_{2}}{x_{1}}\Big{|}\Big{)}1_{\{|x_{1}|>10^{-10}\}}.

Since the function ff changes sign several times, we expect that the corresponding optimal state function u^\hat{u}, solution of (4.5), changes sign several times too. The function gg, which is the right-hand side of the adjoint state function z^\hat{z}, solution of (4.6), also changes sign four times. Taking into account (5.5), this produces a bang-bang control with α\alpha and β\beta being exchanged in several regions. For our numerical experiments we consider α=0\alpha=0, and we take three different values for β\beta: 11, 10210^{2} and 10410^{4}. The corresponding results are shown in Figure 3. We observe that as β\beta grows up, the optimal control m^\hat{m} is concentrated on smaller sets. We expect that for β\beta tending to infinity, m^\hat{m} goes to a singular measure.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: Second example: right-hand side ff: (top left). Optimal control m^\hat{m} for β=1\beta=1 (top right), β=102\beta=10^{2} (down left) and β=104\beta=10^{4} (down right).

Third Example. In this case we take

j(x,s)=s,ψ(s)=1(,α)(β,+)+ks1[α,β],j(x,s)=s,\qquad\psi(s)=\infty 1_{(-\infty,\alpha)\cup(\beta,+\infty)}+ks1_{[\alpha,\beta]},

and

f(x1,x2)=1ω1(x1,x2)+1ω2(x1,x2)+1ω3(x1,x2)+1ω4(x1,x2).f(x_{1},x_{2})=1_{\omega_{1}}(x_{1},x_{2})+1_{\omega_{2}}(x_{1},x_{2})+1_{\omega_{3}}(x_{1},x_{2})+1_{\omega_{4}}(x_{1},x_{2}).

where ωi=B(ci,r)\omega_{i}=B(c_{i},r) (i=1,,4i=1,\dots,4) are small balls of radius r=3/10r=\sqrt{3}/10 and centers c1=(0,0.5)c_{1}=(0,0.5), c2=(0,0.5)c_{2}=(0,-0.5), c3=(0.5,0)c_{3}=(0.5,0) and c4=(0.5,0)c_{4}=(-0.5,0). Taking α=0\alpha=0, β=1\beta=1, we have carried out three numerical experiments corresponding to k=0.00175k=0.00175, k=0.0014k=0.0014 and k=0.001k=0.001 respectively. The results are given in Figure 4. We get a bang-bang optimal potential for the three experiments. Moreover, as expected, the norm in L1(B)L^{1}(B) of m^\hat{m} increases when kk decreases, taking the values 0.1727660.172766, 0.5319710.531971 and 0.874948,0.874948, respectively.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Right-hand side ff (top left). Optimal control m^\hat{m} for k=0.00175k=0.00175 (top right), k=0.0014k=0.0014 (down left) and k=0.001k=0.001 (down right).

Acknowledgments. The work of GB is part of the project 2017TEXA3H “Gradient flows, Optimal Transport and Metric Measure Structures” funded by the Italian Ministry of Research and University. GB is member of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM).

The work of JCD and FM is a part of the FEDER project PID2020-116809GB-I00 of the Ministerio de Ciencia e Innovación of the government of Spain and the project UAL2020-FQM-B2046 of the Consejería de Transformación Económica, Industria, Conocimiento y Universidades of the regional government of Andalusia, Spain.

References

  • [1] G. Allaire. Shape optimization by the homogenization method. Appl. Math. Sci. 146, Springer-Verlag, New York 2002.
  • [2] J.C. Bellido, G. Buttazzo, B. Velichkov. Worst-case shape optimization for the Dirichlet energy. Nonlinear Anal., 153 (2017), 117–129.
  • [3] P. Bénilan, L. Boccardo, T. Gallouët, R. Gariepy, M. Pierre, J.L. Vázquez. An L1L^{1}-theory of existence and uniqueness of solutions on nonlinear elliptic equations. Ann. Scuola Norm. Sup. Pisa Cl. Sci., 22 (2) (1995), 241–273.
  • [4] L. Boccardo, T. Gallouët. Non-linear elliptic and parabolic equations involving measure data. J. Funct. Anal., 87 (1989), 149–169.
  • [5] D. Bucur. Minimization of the kk-th eigenvalue of the Dirichlet Laplacian. Arch. Ration. Mech. Anal., 206 (3) (2012), 1073–1083.
  • [6] D. Bucur, G. Buttazzo. Variational Methods in Shape Optimization Problems. Birkhäuser, Basel 2005.
  • [7] D. Bucur, G. Buttazzo, B. Velichkov. Spectral optimization problems for potentials and measures. SIAM J. Math. Anal., 46 (4) (2014), 2956–2986.
  • [8] G. Buttazzo, J. Casado-Díaz, F. Maestre. On the existence of optimal potentials on unbounded domains. SIAM J. Math. Anal., 53 (2021), 1088–1121.
  • [9] G. Buttazzo, G. Dal Maso Shape optimization for Dirichlet problems: relaxed solutions and optimality conditions. Appl. Math. Optim., 23 (1991), 17–49.
  • [10] G. Buttazzo, G. Dal Maso. An existence result for a class of shape optimization problems. Arch. Ration. Mech. Anal., 122 (1993), 183–195.
  • [11] G. Buttazzo, A. Gerolin, B. Ruffini, B. Velichkov. Optimal potentials for Schördinger operators. J. Éc. Polytech. Math., 1 (2014), 71–100.
  • [12] G. Buttazzo, F. Maestre, B. Velichkov. Optimal potentials for problems with changing sign data. J. Optim. Theory Appl., 178 (3) (2018), 743–762.
  • [13] G. Buttazzo, B. Velichkov. A shape optimal control problem with changing sign data. SIAM J. Math. Anal., 50 (3) (2018), 2608–2627.
  • [14] E.A. Carlen, R.L. Frank, E.H. Lieb. Stability Estimates for the Lowest Eigenvalue of a Schrödinger Operator. Geom. Funct. Anal., 24 (1) (2014), 63–84.
  • [15] J. Casado-Díaz. Optimal design of multi-phase materials with a cost functional that depends nonlinearly on the gradient. SpringerBriefs in Mathematics, Springer, Cham 2022.
  • [16] G. Buttazzo, J. Casado-Díaz, F. Maestre. On the existence of optimal potentials on unbounded domains. SIAM J. Math. Anal., 53 (2021), 1088–1121.
  • [17] G. Dal Maso, F. Murat, L. Orsina, A. Prignet. Renormalized solutions of elliptic equations with general measure data. Ann. Scuola Norm. Sup. Pisa Sci., 28 (1999), 741–808.
  • [18] E. Giusti. Minimal surfaces and function of bounded variation. Birkhäuser, Boston 1984.
  • [19] F. Hecht New development in FreeFem++. J. Numer. Math., 20 (2012), 251–265.
  • [20] I. Mazari, G. Nadin, Y. Privat. Optimisation of the total population size for logistic diffusive equations: bang-bang property and fragmentation rate. Comm. Partial Differential Equations, 47 (4) (2022), 797–828.
  • [21] I. Mazari, G. Nadin, Y. Privat. Optimal location of resources maximizing the total population size in logistic models. J. Math. Pures Appl., 134 (2020), 1–35.
  • [22] D. Mazzoleni, A. Pratelli. Existence of minimizers for spectral problems. J. Math. Pures Appl., 100 (3) (2013), 433–453.
  • [23] D. Gilbarg, N. Trudinger. Elliptic partial differential equations of second order. Springer, Berlin 1998.
  • [24] J. Serrin. Pathological solution of elliptic differential equations. Ann. Scuola Norm. Sup. Pisa Cl. Sci., 18 (1964), 385–387.
  • [25] G. Stampacchia. Le problème de Dirichlet pour les équations elliptiques du second ordre à coefficients discontinus. Ann. Inst. Fourier, 15 (1965), 189–258.