Open Access

Observability Inequalities for Parabolic Equations over Measurable Sets and Some Applications Related to the Bang-Bang Property for Control Problems

   | Dec 15, 2017

Cite

Introduction

This article serves as a review of observability inequalities from measurable sets for solutions to the heat equation. The purpose of trying to obtain the two observability inequalities that we will see and prove in this article, was that in control theory there is a very well known result, the Hilbert Uniqueness Method, that assures that the null controllability of an equation is equivalent to obtain an observability inequality for the adjoint equation. This result is attributed to J.L. Lion. In our previous research we were studying the null controllability of parabolic equations over measurable sets, so, for the Hilbert Uniqueness Method reason, we focused on proving the observability inequalities (Theorems 1 and 2) that we will see in this article.

In the next lines of the Introduction we will establish the type of problem we will work on, remember some apriori estimates for the parabolic equations and recall some previous results about this kind of work.

Then, in Section 2, we will establish and prove Theorem 1 and 2 which will give us two observability inequalities. We will continue, in Section 3, showing some applications of the obserbavility inequalities, the bang-bang property for the minimal time, optimal time and minimal norm control problems. In Section 4, we will establish some open problems related to observability inequalities and their applications to control theory. Finally, with Section 5, we will finish the article giving some details of a definition and a proof requiered in Section 3.

Let Ω be a bounded Lipschitz domain in ℝn and T be a fixed positive time. Consider the heat equation:

{tuΔu=0,inΩ×(0,1),u=0,onΩ×(0,T),u(0)=u0,inΩ,$$ \begin{equation}\label{1.1} \left\{\begin{array}{l} \partial_tu-\Delta u=0,\quad\mbox{in $\Omega\times(0,1)$,} \\[.8em]\quad u=0, \quad\mbox{on $\partial\Omega\times(0,T)$,} \\[1em] \displaystyle u(0)=u_0, \quad\mbox{in $\Omega$,} \end{array}\right. \end{equation} $$

with u0 in L2(Ω). The solution of (1) will be treated as either a function from [0, T] to L2(Ω) or a function of two variables x and t. Two important apriori estimates for the above equation are as follows:

u(T)L2(Ω)N(Ω,T,D)D|u(x,t)|dxdt,$$ \begin{equation} \|u(T)\|_{L^2(\Omega)}\leq N(\Omega,T,\mathscr{D})\int_{\mathscr{D}}|u(x,t)|\,dxdt, \end{equation} $$

for all u0 ∈ L2(Ω), where 𝒟 is a subset of Ω  ×  (0, T), and

u(T)L2(Ω)N(Ω,T,J)J|νu(x,t)|dσdt,$$ \begin{equation} \|u(T)\|_{L^2(\Omega)}\leq N(\Omega,T,\mathscr{J})\int_{\mathscr{J}}|\frac{\partial}{\partial\nu}u(x,t)|\,d\sigma dt, \end{equation} $$

for all u0 ∈ L2(Ω), where 𝒥 is a subset of ∂Ω  ×  (0, T). Such apriori estimates are called observability inequalities.

In the case that 𝒟 = ω× (0,T) and 𝒥 = Γ × (0,T) with ω and Γ accordingly open and nonempty subsets of Ω and ∂Ω, both inequalities (2) and (3) (where ∂Ω is smooth) were essentially first established, via the Lebeau-Robbiano spectral inequalities in [8]. These two estimates were set up to the linear parabolic equations (where ∂Ω is of class C2), based on the Carleman inequality provided in [7]. In the case when 𝒟 = ω × (0,T) and 𝒥 = Γ × (0,T) with ω and Γ accordingly subsets of positive measure and positive surface measure in Ω and ∂Ω, both inequalities (2) and (3) were built up in [1] with the help of a propagation of smallness estimate from measurable sets for real-analytic functions first established in [13]. For 𝒟 = ω × E, with ω and E accordingly an open subset of Ω and a subset of positive measure in (0, T), the inequality (2) (when ∂Ω is smooth) was proved in [14] with the aid of the Lebeau-Robbiano spectral inequality, and it was then verified for heat equations (when Ω is convex) with lower terms depending on the time variable, through a frequency function method in [11]. When 𝒟 = ω × E, with ω and E accordingly subsets of positive measure in Ω and (0, T), the estimate (2) (when ∂Ω is real-analytic) was obtained in [15].

In [2], we stablished the inequalities (2) and (3) when 𝒟 and 𝒥 were arbitrary subsets of positive measure and of positive surface measure in Ω  ×  (0, T) and ∂Ω  ×  (0, T) respectively. Such inequalities not only are mathematically interesting but also have important applications in the control theory of the heat equation, such as the bang-bang control, the time optimal control, the null controllability over a measurable set and so on.

We will see how we proved the two above-mentioned inequalities. We start assuming that the Lebeau-Robbiano spectral inequality stands on Ω. To introduce it, we write

0<λ1λ2λj$$ 0\lt \lambda_1\le\lambda_2\le\dots\leq\lambda_j\le\cdots $$

for the eigenvalues of −Δ with the zero Dirichlet boundary condition over ∂Ω, and {ej: j≥1 } for the set of L2(Ω)-normalized eigenfunctions, i.e.,

{Δej+λjej=0,inΩ,ej=0,onΩ.$$ \begin{equation} \left\{\begin{array}{l} \Delta e_j+\lambda_je_j=0,\quad\mbox{in $\Omega$,} \\[1em] \displaystyle e_j=0, \quad\mbox{on $\partial\Omega$.} \end{array}\right. \end{equation} $$

For λ > 0 we define

λf=λjλ(f,ej)ejandλf=λj>λ(f,ej)ej,$$ \mathscr E_\lambda f=\sum_{\lambda_j\le\lambda}( f,e_j)\,e_j\quad \mbox{and}\quad \mathscr E_\lambda^\perp f=\sum_{\lambda_j\gt\lambda}( f,e_j)\,e_j, $$

where

(f,ej)=Ωfejdx,whenfL2(Ω),j1.$$ ( f,e_j)=\int_{\Omega} f\, e_j\,dx,\ \mbox{when}\ f\in L^2(\Omega),\ j\ge 1. $$

Throughout this paper the following notations are used:

(f,g)=ΩfgdxandfL2(Ω)=(f,f)12.$$ (f,g)=\int_{\Omega}fg\,dx\quad \mbox{and}\ \|f\|_{L^2(\Omega)}=\left(f,f\right)^{\frac 12}. $$

ν is the unit exterior normal vector to Ω. dσ is surface measure on ∂Ω. BR(x0) stands for the ball centered at x0 in \ℝn of radius R, ΔR(x0) denotes BR(x0)∩∂Ω, BR=BR(0) and ΔRR(0). For measurable sets ω⊂ℝn and 𝒟 ⸦ ℝn × (0,T), |ω| and |𝒟| stand for the Lebesgue measures of the sets. For each measurable set 𝒥 in ∂Ω  ×  (0, T), 𝒥 in ∂ω × (0,T), |𝒥| denotes its surface measure on the lateral boundary of Ω  ×  ℝ. {etΔ : t≥0} is the semigroup generated by Δ with zero Dirichlet boundary condition over ∂Ω. Consequently, etΔf is the solution of equation (1) with the initial state f in L2(Ω). The Lebeau-Robbiano spectral inequality is as follows:

For each 0 < R ≤ 1, there isN=N (Ω, R), such that the inequality

λfL2(Ω)NeNλλfL2(BR(x0))$$ \begin{equation} \|\mathscr E_\lambda f\|_{L^2(\Omega)}\le Ne^{N\sqrt\lambda}\|\mathscr E_\lambda f\|_{L^2(B_R(x_0))} \end{equation} $$

holds, whenB4R(x0)⊂Ω, f ∈ L2(Ω) and λ > 0.

Observability inequalities

Our main results related to the observability inequalities are stated as follows, but, first, we will define the real-analyticity of the set Δ4R(q0).

Definition 1

Letq0 ∈ ∂Ω and 0 < R ≤ 1. We say that Δ4R(q0) is real-analytic with constants ρ and δ if for each q ∈ Δ4R(q0), there are a new rectangular coordinate system where q=0, and a real-analytic function ϕ:Bϱn1 $\phi: B'_{\varrho}\subset\mathbb{R}^{n-1} \rightarrow \mathbb{R}$ verifying

{ϕ(0)=0,|αϕ(x)||α|!δ|α|1,whenxBϱ,αn1,BϱΩ=Bϱ{(x,xn):xBϱ,xn>ϕ(x)},BϱΩ=Bϱ{(x,xn):xBϱ,xn=ϕ(x)}.$$ \begin{equation} \left\{\begin{array}{l} \phi(0')=0,\;\;|\partial^{\alpha}\phi(x')|\leq |\alpha|!\delta^{-|\alpha|-1}, \\[1em] \mbox{when}\;\;x'\in B'_{\varrho},\;\alpha\in\mathbb{N}^{n-1}, \\[1em] B_\varrho\cap\Omega=B_\varrho\cap\{(x',x_n):x'\in B'_{\varrho},\;\; x_n\gt\phi(x')\}, \\[1em] \displaystyle B_\varrho\cap\partial\Omega=B_\varrho\cap\{(x',x_n):x'\in B'_{\varrho},\;\; x_n=\phi(x')\}. \end{array}\right. \end{equation} $$

Here, Bρ denotes the open ball of radius ρ and with center at 0′ in ℝn−1.

In the next two theorems, we establish two observability inequalities for the heat equation over Ω  ×  (0, T). In Theorem 1, the observation is from a subset of positive measure in Ω  ×  (0, T), while in Theorem 2, the observation is from a subset of positive surface measure on ∂Ω  ×  (0, T).

Theorem 1

Suppose that a bounded domain Ω verifies the condition (5) andT > 0. Letx0 ∈ Ω andR ∈ (0, 1] be such thatB4R(x0)⊂Ω. Then, for each measurable set 𝒟 ⊂ BR(x0) × (0,T) with |𝒟| > 0, there is a positive constantB = B(Ω, T, R, 𝒟), such that

eTΔfL2(Ω)eBD|etΔf(x)|dxdt,$$ \begin{equation} \|e^{T\Delta}f\|_{L^2(\Omega)}\le e^{B}\int_{\mathscr D}|e^{t\Delta}f(x)|\,dxdt, \end{equation} $$

whenf ∈ L2(Ω).

Theorem 2

Suppose that a bounded Lipschitz domain Ω verifies the condition (5) andT > 0. Letq0 ∈ ∂Ω andR ∈ (0, 1] be such that Δ4R(q0) is real-analytic. Then, for each measurable set 𝒥 ⊂ ΔR(q0)× (0,T) with |𝒥| > 0, there is a positive constantB = B(Ω, T, R, 𝒥, such that

eTΔfL2(Ω)eBJ|νetΔf(x)|dσdt,$$ \begin{equation} \|e^{T\Delta}f\|_{L^2(\Omega)}\le e^{B}\int_{\mathscr J}|\frac{\partial}{\partial\nu}\, e^{t\Delta}f(x)|\,d\sigma dt, \end{equation} $$

whenf ∈ L2(Ω).

Next, we will see some results that will be necessary in the proof of the previous Theorem 1.

Lemma 3

LetBR(x0)⊂Ω and 𝒟 ⊂ BR(x0) × (0,T) be a subset of positive measure. Set

Dt={xΩ:(x,t)D},E={t(0,T):|Dt||D|/(2T)},t(0,T).$$ \begin{equation} \mathscr D_t=\{x\in \Omega : (x,t)\in\mathscr D\},\ E=\{t\in (0,T): |\mathscr D_t |\ge |\mathscr D|/(2T)\},\ t\in (0,T). \end{equation} $$

Then, 𝒟t ⊂ Ω is measurable for a.e. t ∈ (0, T), Eis measurable in (0, T), |E| ≥ |𝒟|/2|BR| and

χE(t)χDt(x)χD(x,t),inΩ×(0,T).$$ \begin{equation} \chi_E(t)\chi_{\mathscr D_t}(x)\le \chi_{\mathscr D}(x,t),\ \text{in}\ \Omega\times (0,T). \end{equation} $$

Proof

From Fubini’s theorem,

|D|=0T|Dt|dt=E|Dt|dt+[0,T]E|Dt|dt|BR||E|+|D|/2.$$ \begin{equation*} |\mathscr D|=\int_0^T|\mathscr D_t|\,dt=\int_E|\mathscr D_t|\,dt+\int_{[0,T]\setminus E}|\mathscr D_t|\,dt\le|B_R||E|+|\mathscr D|/2. \end{equation*} $$

Theorem 4

Letx0 ∈ Ω andR ∈ (0, 1] be such thatB4R(x0)⊂Ω. Let 𝒟⊂ BR(x0)× (0,T) be a measurable set with |𝒟| > 0. WriteE and 𝒟tfor the sets associated to 𝒟 in Lemma 3. Then, for eachη ∈ (0, 1), there areN = N(Ω,R, |𝒟|/(T|BR|),η) andθ = θ(Ω,R, |𝒟|/(T|BR|), η) withθ ∈ (0, 1), such that

et2ΔfL2(Ω)(NeN/(t2t1)t1t2χE(s)esΔfL1(Ds)ds)θet1ΔfL2(Ω)1θ,$$ \begin{equation} \|e^{t_2\Delta}f\|_{L^2(\Omega)} \le \left( N e^{N/(t_2-t_1)} \int_{t_1}^{t_2}\chi_E(s) \|e^{s\Delta}f\|_{L^1(\mathscr D_s)}\,ds \right)^{\theta}\|e^{t_1\Delta}f\|_{L^2(\Omega)}^{1-\theta}, \end{equation} $$

when 0 ≤ t1 < t2T, |E ∩ (t1, t2)| ≥ η(t2t1) andf ∈ L2(Ω). Moreover,

eN+1θt2t1et2ΔfL2(Ω)eN+1θq(t2t1)et1ΔfL2(Ω)Nt1t2χE(s)esΔfL1(Ds)ds,whenq(N+1θ)/(N+1).$$ \begin{equation} e^{-\frac{N+1-\theta}{t_2-t_1}}\|e^{t_2\Delta}f\|_{L^2(\Omega)}- e^{-\frac{N+1-\theta}{q\left(t_2-t_1\right)}}\|e^{t_1\Delta}f\|_{L^2(\Omega)} \\ \le N\int_{t_1}^{t_2}\chi_E(s) \|e^{s\Delta}f\|_{L^1(\mathscr D_s)}\,ds,\;\;\mbox{when}\;\;q\ge (N+1-\theta)/(N+1). \end{equation} $$

The reader can find the proof of the following Lemma 2 in either [10, pp. 256-257] or [11, Proposition 2.1].

Lemma 5

Let E be a subset of positive measure in (0, T). Letlbe a density point of E. Then, for each z > 1, there isl1=l1(z, E) in (l, T) such that, the sequence {lm} defined as

lm+1=l+zm(l1l),m=1,2,,$$ \begin{equation*} l_{m+1}= l+z^{-m}\left(l_1-l\right),\ m=1, 2,\cdots, \end{equation*} $$

verifies

|E(lm+1,lm)|13(lmlm+1),whenm1.$$ \begin{equation} |E\cap (l_{m+1}, l_m)|\ge \frac 13 \left(l_m-l_{m+1}\right),\ \text{when}\ m\ge 1. \end{equation} $$

Proof

[Theorem 1] LetE and 𝒟t be the sets associated to 𝒟t in Lemma 3 and l be a density point in E. For z > 1 to be fixed later, {lm} denotes the sequence associated to l and z in Lemma 5. Because (13) holds, we may apply Theorem 4, with η=1/3, t1=lm+1 and t2=lm, for each m ≥ 1, to get that there are N = N(Ω,R, |𝒟|/(T|BR|)) > 0 and θ = θ(Ω,R, |𝒟|/(T|B_R|)), with θ ∈ (0, 1), such that

eN+1θlmlm+1elmΔfL2(Ω)eN+1θq(lmlm+1)elm+1ΔfL2(Ω)Nlm+1lmχE(s)esΔfL1(Ds)ds,whenqN+1θN+1andm1.$$ \begin{equation} e^{-\frac{N+1-\theta}{l_m-l_{m+1}}}\|e^{l_m\Delta}f\|_{L^2(\Omega)}- e^{-\frac{N+1-\theta}{q\left(l_m-l_{m+1}\right)}}\|e^{l_{m+1}\Delta}f\|_{L^2(\Omega)} \\ \le N\int_{l_{m+1}}^{l_{m}}\chi_E(s) \|e^{s\Delta}f\|_{L^1(\mathscr D_s)}\,ds,\;\;\mbox{when}\;\;q\ge \frac{N+1-\theta}{N+1}\;\;\mbox{and}\;\;m\ge 1. \end{equation} $$

Setting z=1/q in (14) (which leads to 1<zN+1N+1θ) $1\lt z\le \frac{N+1}{N+1-\theta})$ and

γz(t)=eN+1θ(z1)(l1l)t,t>0,$$ \begin{equation*} \gamma_z(t)=e^{-\frac{N+1-\theta}{\left(z-1\right)\left(l_1-l\right)t}},\ t\gt0, \end{equation*} $$

recalling that

lmlm+1=zm(z1)(l1l),form1,$$ \begin{equation*} l_m-l_{m+1}=z^{-m}\left(z-1\right)\left(l_1-l\right),\ \text{for}\ m\ge 1, \end{equation*} $$

we have

γz(zm)elmΔfL2(Ω)γz(zm1)elm+1ΔfL2(Ω)Nlm+1lmχE(s)esΔfL1(Ds)ds,whenm1.$$ \begin{equation} \gamma_z(z^{-m})\|e^{l_m\Delta}f\|_{L^2(\Omega)}- \gamma_z(z^{-m-1})\|e^{l_{m+1}\Delta}f\|_{L^2(\Omega)} \\ \le N\int_{l_{m+1}}^{l_{m}}\chi_E(s) \|e^{s\Delta}f\|_{L^1(\mathscr D_s)}\,ds,\;\;\mbox{when}\;\;m\ge 1. \end{equation} $$

Choose now

z=12(1+N+1N+1θ).$$ \begin{equation*} z=\frac 12\left(1+\frac{N+1}{N+1-\theta}\right). \end{equation*} $$

The choice of z and Lemma 5 determines l1 in (l, T) and from (15)

γ(zm)elmΔfL2(Ω)γ(zm1)elm+1ΔfL2(Ω)Nlm+1lmχE(s)esΔfL1(Ds)ds,whenm1.$$ \begin{equation} \begin{split} \gamma(z^{-m})\|e^{l_m\Delta}f\|_{L^2(\Omega)}- \gamma(z^{-m-1})\|e^{l_{m+1}\Delta}f\|_{L^2(\Omega)} \\ \le N\int_{l_{m+1}}^{l_{m}}\chi_E(s) \|e^{s\Delta}f\|_{L^1(\mathscr D_s)}\,ds,\;\;\mbox{when}\;\;m\ge 1. \end{split} \end{equation} $$

with

γ(t)=eA/tandA=A(Ω,R,E,|D|/(T|BR|))=2(N+1θ)2θ(l1l).$$ \begin{equation*} \gamma(t)=e^{-A/t}\ \text{and}\ A=A(\Omega,R, E, |\mathscr D|/\left(T|B_R| \right))=\frac{2\left(N+1-\theta\right)^2}{\theta\left(l_1-l\right)}\, . \end{equation*} $$

Finally, because of

eTΔfL2(Ω)el1ΔfL2(Ω),supt0etΔfL2(Ω)<+,limt0+γ(t)=0,$$ \begin{equation*} \|e^{T\Delta}f\|_{L^2(\Omega)}\le \|e^{l_1\Delta}f\|_{L^2(\Omega)},\ \sup_{t\ge 0}\|e^{t\Delta}f\|_{L^2(\Omega)}\lt+\infty,\ \lim_{t\to 0+}\gamma(t)=0, \end{equation*} $$

and (10), the addition of the telescoping series in (16) gives

eTΔfL2(Ω)NezAD(Ω×[l,l1])|etΔf(x)|dxdt,forfL2(Ω),$$ \begin{equation*} \|e^{T\Delta}f\|_{L^2(\Omega)}\le Ne^{zA}\int_{\mathscr D\cap(\Omega\times [l,l_1])}|e^{t\Delta}f(x)|\,dxdt,\;\;\mbox{for}\;\; f\in L^2(\Omega), \end{equation*} $$

which proves (7) with B=zA+logN. □

Remark 1

The constant B in Theorem 1 depends on E because the choice of l1=l1(z, E) in Lemma 5 depends on the possible complex structure of the measurable set E (See the proof of Lemma 5 in [Proposition 2.1]). When𝒟 = ω × (0,T), one may takel=T/2, l1=T, z=2 and then,

B=A(Ω,R,|ω|/|BR|)/T.$$ \begin{equation*} B=A(\Omega, R, |\omega|/|B_R|)/T. \end{equation*} $$

Remark 2

The proof of Theorem 1 also implies the following observability estimate:

supm0suplm+1tlmezm+1AetΔfL2(Ω)ND(Ω×[l,l1])|etΔf(x)|dxdt,$$ \begin{equation*} \sup_{m\ge 0}\sup_{l_{m+1}\le t\le l_m}e^{-z^{m+1}A}\|e^{t\Delta}f\|_{L^2(\Omega)}\le N\int_{\mathscr D\cap (\Omega\times [l,l_1])}|e^{t\Delta}f(x)|\,dxdt, \end{equation*} $$

forf in L2(Ω), and with z, N and A as defined along the proof of Theorem 1. Here, l0=T.

Next, we will see some results that will be necessary in the proof of the previous Theorem 2.

Lemma 6

Let q0 ∈ ∂Ω and 𝒥 ⊂ ΔR(q0) × (0,T) be a subset with |𝒥| > 0. Set

Jt={xΩ:(x,t)J},E={t(0,T):|Jt||J|/(2T)},t(0,T).$$ \begin{equation*} \mathscr J_t=\{x\in \partial\Omega : (x,t)\in\mathscr J\},\ E=\{t\in (0,T): |\mathscr J_t|\ge |\mathscr J|/(2T)\},\ t\in (0,T). \end{equation*} $$

Then, 𝒥t ⊂ ΔR(q0) is measurable for a.e. t ∈ (0, T), Eis measurable in (0, T), |E| ≥ |𝒥|/(2|ΔR(q0)|) andχE(t)χ𝒥t(x) ≤ χ𝒥(x,t) over ∂Ω  ×  (0, T).

Proof

From Fubini’s theorem,

|J|=0T|Jt|dt=E|Jt|dt+[0,T]E|Jt|dt|R(x0)||E|+|J|/2.$$ \begin{equation*} |\mathscr J|=\int_0^T|\mathscr J_t|\,dt=\int_E|\mathscr J_t|\,dt+\int_{[0,T]\setminus E}|\mathscr J_t|\,dt\le |\triangle_R(x_0)||E|+|\mathscr J|/2. \end{equation*} $$

Theorem 7

Suppose that Ω verifies the condition (5). Assume thatq0 ∈ ∂Ω andR ∈ (0, 1] such that Δ4R(q0) is real-analytic. Let 𝒥 be a subset in ΔR(q0)  ×  (0, T) of positive surface measure on ∂Ω  ×  (0, T), E and 𝒥tbe the measurable sets associated to 𝒥 in Lemma 6. Then, for eachη ∈ (0, 1), there areN = N(Ω,R, |𝒥|/(TR(q0)|),η) andθ = θ(Ω,R, |𝒥|/(TR(q0)|),η) withθ ∈ (0, 1), such that the inequality

et2ΔfL2(Ω)(NeN/(t2t1)t1t2χE(t)νetΔfL1(Jt)dt)θet1ΔfL2(Ω)1θ,$$ \begin{equation} \|e^{t_2\Delta}f\|_{L^2(\Omega)} \le \left( N e^{N/(t_2-t_1)} \int_{t_1}^{t_2}\chi_E(t) \|\tfrac{\partial}{\partial\nu}e^{t\Delta}f\|_{L^1(\mathscr J_t)}\,dt \right)^{\theta}\|e^{t_1\Delta}f\|_{L^2(\Omega)}^{1-\theta}, \end{equation} $$

holds, when 0 ≤ t1 < t2Twitht2t1 < 1, |E ∩ (t1, t2)| ≥ η(t2t1) andf ∈ L2(Ω). Moreover,

eN+1θt2t1et2ΔfL2(Ω)eN+1θq(t2t1)et1ΔfL2(Ω)Nt1t2χE(t)νetΔfL1(Jt)dt,whenqN+1θN+1.$$ \begin{equation} e^{-\frac{N+1-\theta}{t_2-t_1}}\|e^{t_2\Delta}f\|_{L^2(\Omega)}- e^{-\frac{N+1-\theta}{q\left(t_2-t_1\right)}}\|e^{t_1\Delta}f\|_{L^2(\Omega)}\\ \le N\int_{t_1}^{t_2}\chi_E(t) \|\tfrac{\partial}{\partial\nu}\,e^{t\Delta}f\|_{L^1(\mathscr J_t)}\,dt,\;\;\mbox{when}\;\;q\ge \tfrac{N+1-\theta}{N+1}. \end{equation} $$

Proof

[Theorem 2] Let E and 𝒥t be the sets associated to 𝒥 in Lemma 6 and l be a density point in E. For z > 1 to be fixed later, {lm} denotes the sequence associated to l and z in Lemma 5. Because of (13) and from Theorem 7 with η=1/3, t1=lm+1 and t2=lm, with m≥1, there are N = N(Ω,R, |𝒥|/(TR(q0)|)) > 0 and θ = θ(Ω,R, |𝒥|/(TR(q0)|)), with θ ∈ (0, 1), such that

eN+1θlmlm+1elmΔfL2(Ω)eN+1θq(lmlm+1)elm+1ΔfL2(Ω)Nlm+1lmχE(s)νesΔfL1(Js)ds,whenqN+1θN+1andm1.$$ \begin{equation*} \begin{split} e^{-\frac{N+1-\theta}{l_m-l_{m+1}}}\|e^{l_m\Delta}f\|_{L^2(\Omega)}- e^{-\frac{N+1-\theta}{q\left(l_m-l_{m+1}\right)}}\|e^{l_{m+1}\Delta}f\|_{L^2(\Omega)}\\ \le N\int_{l_{m+1}}^{l_{m}}\chi_E(s) \|\tfrac{\partial}{\partial\nu}\,e^{s\Delta}f\|_{L^1(\mathscr J_s)}\,ds,\;\;\mbox{when}\;\;q\ge \frac{N+1-\theta}{N+1}\;\;\mbox{and}\;\;m\ge 1. \end{split} \end{equation*} $$

Let

z=12(1+N+1N+1θ).$$ \begin{equation*} z=\frac{1}{2}\left(1+\frac{N+1}{N+1-\theta}\right). \end{equation*} $$

Then, we can use the same arguments as those in the proof of Theorem 1 to verify Theorem 2. □

Remark 3

The proof of Theorem 2 also implies the following observability estimate:

supm0suplm+1tlmezm+1AetΔfL2(Ω)NJ(Ω×[l,l1])|νetΔf(x)|dσdt,$$ \begin{equation*}%{yuanyuan6} \sup_{m\ge 0}\sup_{l_{m+1}\le t\le l_m}e^{-z^{m+1}A}\|e^{t\Delta}f\|_{L^2(\Omega)}\le N\int_{\mathscr J\cap (\partial\Omega\times [l,l_1])}\left|\tfrac{\partial}{\partial\nu}\,e^{t\Delta}f(x)\right|\,d\sigma dt, \end{equation*} $$

forfinL2(Ω), withA=2(N + 1−θ)2/[θ(l1l)] and withz, Nandθas given along the proof of Theorem 2. Here, l0=T.

Remark 4

When 𝒥 = Γ × (0,T), Γ⊂ΔR(q0) is a measurable set, one may takel=T/2, l1=T, z=2 and the constant Bin Theorem 2 becomes

B=A(Ω,R,|Γ|/|R(q0)|)/T.$$ \begin{equation*} B=A(\Omega, R, |\Gamma |/|\triangle_R(q_0)|)/T. \end{equation*} $$

Applications of observability inequalities

We will now show some applications of the Theorems 1 and 2 in the control theory of the heat equation. Specifically, we will focus on the uniqueness and bang-bang properties of the minimal time, time optimal and minimal L-norm control problems.

In this section we assume that T > 0 and that Ω is a bounded Lipschitz domain verifying the condition (5).

First of all, we will show that Theorems 1 and 2 imply the null controllability with controls restricted over measurable subsets in Ω  ×  (0, T) and ∂Ω  ×  (0, T) respectively. Let 𝒟 be a measurable subset with positive measure in BR(x0)  ×  (0, T) with B4R(x0)⊂Ω. Let 𝒥 be a measurable subset with positive surface measure in ΔR(q0)  ×  (0, T), where q0 ∈ ∂Ω, R ∈ (0, 1] and Δ4R(q0) is real-analytic. Consider the following controlled heat equations:

{tuΔu=χDv,inΩ×(0,T],u=0,onΩ×[0,T],u(0)=u0,inΩ,$$ \begin{equation} \left\{\begin{array}{l} \partial_tu-\Delta u=\chi_{\mathscr{D}}v, \mbox{in}\ \Omega\times (0,T], \\[1em] u=0, \mbox{on}\ \partial\Omega\times [0,T], \\[1em] \displaystyle u(0)=u_0,\ \mbox{in}\ \Omega, \end{array}\right. \end{equation} $$

and

{tuΔu=0,inΩ×(0,T],1u=gχJ,onΩ×[0,T],u(0)=u0,inΩ,$$ \begin{equation} \left\{\begin{array}{l} \partial_tu-\Delta u=0,\ \mbox{in}\ \Omega\times (0,T], \\[1em] u=g\,\chi_{\mathscr J},\ \mbox{on}\ \partial\Omega\times [0,T], \\[1em] \displaystyle u(0)=u_0,\ \mbox{in}\ \Omega, \end{array}\right. \end{equation} $$

where u0 ∈ L2(Ω), v ∈ L(Ω  ×  (0, T)) and g ∈ L(∂Ω  ×  (0, T)) are controls. We say thatuis the solution to 20ifv ≡ u − etΔu0is the unique solution defined in [6, Theorem 3.2] to

{tvΔv=0,inΩ×(0,T),v=gχJ,onΩ×(0,T),v(0)=0,inΩ,$$ \begin{equation} \left\{\begin{array}{l} \partial_tv-\Delta v=0,\ \mbox{in}\ \Omega\times(0,T), \\[1em]v=g\chi_{\mathscr{J}},\ \mbox{on}\ \partial\Omega\times(0,T), \\[1em] \displaystyle v(0)=0,\ \mbox{in}\ \Omega, \end{array}\right. \end{equation} $$

withg in Lp(∂Ω  ×  (0, T)) for some 2 ≤ p ≤ ∞.

From now on, we always denote by u (· ; u0, v) and u (· ; u0, g) the solutions of equations (19) and (20) corresponding to v and g respectively.

Corollary 8

For eachu0 ∈ L2(Ω), there are bounded control functionsvandgwith

vL(Ω×(0,T))C1u0L2(Ω),$$ \|v\|_{L^\infty(\Omega\times (0,T))} \leq C_1\|u_0\|_{L^2(\Omega)}, $$

gL(Ω×(0,T))C2u0L2(Ω),$$ \begin{equation*} \displaystyle \|g\|_{L^\infty(\partial\Omega\times (0,T))}\leq C_2\|u_0\|_{L^2(\Omega)}, \end{equation*} $$

such thatu(T; u0, v)=0 andu(T; u0, g)=0. HereC1 = C(Ω, T, R, 𝒟) andC2 = C(Ω, T, R, 𝒥).

Proof

We only prove the boundary controllability. Let E be the measurable set associated to 𝒥 in Lemma 6. Write

J˜={(x,t):(x,Tt)J}andE˜={t:TtE}.$$ \begin{equation*} \widetilde{\mathscr{J}}=\left\{(x,t) : (x, T-t)\in {\mathscr{J}}\right\}\;\;\mbox{and}\;\;\widetilde{E}=\left\{t : T-t\in {E}\right\}. \end{equation*} $$

Let l > 0 be a density point of E˜ $\widetilde{E}$ (Hence, T − l is a density point of E). We choose z, l1 and the sequence {lm} as in the proof of Theorem 2 but with 𝒥 and E accordingly replaced by J˜ $\widetilde {\mathscr{J}}$ and E˜ $\widetilde{E}$ . It is clear that

0<l<<lm+1<lm<l1<l0=T,limm+lm=l.$$ \begin{equation*} 0\lt l\lt\dots\lt l_{m+1}\lt l_m\dots\lt l_1\lt l_0=T,\ \lim_{m\to +\infty}l_m=l. \end{equation*} $$

We set

=J(Ω×[Tl1,Tl])J.$$ \begin{equation*} \mathscr M=\mathscr J\cap\left(\partial\Omega\times [T-l_1,T-l]\right)\subset\mathscr J. \end{equation*} $$

It is clear that |𝓜| > 0. The proof of Theorem 2, the change of variables t=Tτ and Remark 3 show that the observability inequality

φ(0)L2(Ω)eB|φν(p,t)|dσdt,$$ \begin{equation} \|\varphi(0)\|_{L^2(\Omega)}\le e^B\int_{\mathscr M}|\tfrac{\partial\varphi}{\partial\nu}(p,t)|\,d\sigma dt, \end{equation} $$

holds, when φ is the unique solution in L([0,T],L2(Ω))L2([0,T],H01(Ω)) $L^\infty([0,T],L^2(\Omega))\cap L^2([0,T], H^1_0(\Omega))$ to

{tφ+Δφ=0,inΩ×[0,T),φ=0,onΩ×[0,T),φ(T)=φT,inΩ,$$ \begin{equation} \begin{cases} \partial_t\varphi+\Delta\varphi=0,\ &\text{in}\ \Omega\times [0,T),\\ \varphi=0,\ &\text{on}\ \partial\Omega\times [0,T),\\ \varphi(T)=\varphi_T,\ &\text{in}\ \partial\Omega, \end{cases} \end{equation} $$

for some φT in L2(Ω). Set

X={φν|:φ(t)=e(Tt)ΔφT,for0tT,for someφTL2(Ω)}.$$ \begin{equation*} X=\{\tfrac{\partial\varphi}{\partial\nu}|_{\mathscr M}: \varphi(t)=e^{(T-t)\Delta}\varphi_T,\ \text{for}\ 0\le t\le T,\ \text{for some}\ \varphi_T\in L^2(\Omega)\}. \end{equation*} $$

Since 𝓜 ⊂ ∂Ω ×[Tl1,Tl], X is a subspace of L1(𝓜) and from (22), the linear mapping Λ: X → ℝ, defined by

Λ(φν|)=(u0,φ(0)),$$ \begin{equation*} \Lambda(\tfrac{\partial\varphi}{\partial\nu}|_{\mathscr M})=(u_0,\varphi(0)), \end{equation*} $$

verifies

|Λ(φν|)|eBu0L2(Ω)|φν(p,t)|dσdt,whenφν|X.$$ \begin{equation*} \left|\Lambda(\tfrac{\partial\varphi}{\partial\nu}|_{\mathscr M})\right|\le e^B\|u_0\|_{L^2(\Omega)}\int_{\mathscr M}|\tfrac{\partial\varphi}{\partial\nu}(p,t)|\,d\sigma dt,\ \text{when}\ \tfrac{\partial\varphi}{\partial\nu}|_{\mathscr M}\in X. \end{equation*} $$

From the Hahn-Banach theorem, there is a linear extension T : L1(𝓜)→ ℝ of Λ, with

T(φν|)=(u0,φ(0)),whenφν|X,|T(f)|eBu0fL1(),for allfL1().$$ \begin{equation*} \begin{split}{left} T(\tfrac{\partial\varphi}{\partial\nu}|_{\mathscr M})=(u_0,\varphi(0)),\ \text{when}\ \tfrac{\partial\varphi}{\partial\nu}|_{\mathscr M}\in X,\\ |T(f)|\le e^B\|u_0\|\|f\|_{L^1(\mathscr M)},\ \text{for all}\ f\in L^1(\mathscr M). \end{split} \end{equation*} $$

Thus, T is in L1(𝓜)* = L(𝓜) and there is g in L(𝓜) verifying

T(f)=fgdσdt,for allfL1()andgL()eBu0.$$ \begin{equation*} T(f)=\int_{\mathscr M}fg\,d\sigma dt,\ \text{for all}\ f\in L^1(\mathscr M)\ \text{and}\ \|g\|_{L^\infty(\mathscr M)}\le e^B\|u_0\|. \end{equation*} $$

We extend g over ∂Ω  ×  (0, T) by setting it to be zero outside 𝓜 and denote the extended function by g again. Then it holds that u(T;u0, g)=0 provided that we know that

Ωu(T;u0,g)φTdx=Ωu0φ(0)dxgφνdσdt,for allφTL2(Ω).$$ \begin{equation} \int_{\Omega}u(T;u_0, g)\varphi_T\,dx=\int_{\Omega}u_0\varphi(0)\,dx-\int_{\mathscr M}g\,\tfrac{\partial\varphi}{\partial\nu}\,d\sigma dt,\ \text{for all}\ \varphi_T\in L^2(\Omega). \end{equation} $$

To prove (24), we first use the unique solvability for the problem

{tuΔu=0,inΩ×(0,T],u=γ,onΩ×[0,T],u(0)=0inΩ,$$ \begin{equation*} \begin{cases} \partial_tu-\Delta u=0,\ &\text{in}\ \Omega\times (0,T],\\ u=\gamma,\ &\text{on}\ \partial\Omega\times [0,T],\\ u(0)=0\ &\text{in}\ \Omega, \end{cases} \end{equation*} $$

with lateral Dirichlet data γ in Lp(∂Ω  ×  (0, T)), 2 ≤ p ≤ ∞, stablished in [6, Theorem 3.2] (See also [3, Theorems 8.1 and 8.3]). Then, because gχ𝓜 is bounded and supported in ∂Ω  ×  [T − l1, T − l]⊂∂Ω  ×  (2η, T − 2η) for some η > 0, the calculations leading to (24) can be justified via the regularization of gχ𝓜 and the approximation of Ω by smooth domains {Ωj;j≥1} as in [3, Lemma 2.2]. For the sake of completeness we provide the detailed proof of this identity in the Appendix in Section 5. □

Definition of the Minimal Time Control Problems and Main Results

In this section, we apply Theorems 1 and 2 to get the bang-bang property for the minimal time control problems usually called the first type of time optimal control problems; they are stated as follows. Let ω be a measurable subset with positive measure in BR(x0) and B4R(x0)⊂Ωm. Suppose that Δ4R(q0) is real-analytic for some q0 ∈ ∂Ω and R ∈ (0, 1] and let Γ be a measurable subset with positive surface measure of ΔR(x0). For each M > 0, we define the following control constraint set:

UM1={vmeasurable onΩ×R+:|v(x,t)|Mfor a.e.(x,t)Ω×R+}.UM1={vmeasurable onΩ×R+:|v(x,t)|Mfor a.e.(x,t)Ω×R+}.$$\eqalign{\mathscr{U}^1_M=\{v\;\;\text{measurable on}\;\; \Omega\times\mathbb{R}^+: \;\;|v(x,t)|\leq M\;\;\text{for a.e.}\;\;(x,t)\in\Omega\times\mathbb{R}^+\}.\\ \mathscr{U}^1_M=\{v\;\;\text{measurable on}\;\; \partial\Omega\times\mathbb{R}^+: \;\;|v(x,t)|\leq M\;\;\text{for a.e.}\;\;(x,t)\in \partial\Omega\times\mathbb{R}^+\}.}$$

Let u0 ∈ L2(Ω)∖{0}. Consider the minimal time control problems:

(TP)M1:TM1minvUM1t>0:etΔu0+0te(ts)Δ(χωv)ds=0$$ \begin{equation} (TP)^1_M:\;\; T^1_M\equiv \min\limits_{v\in\mathscr{U}^1_M} \left\{t\gt0:\;\;e^{t\Delta}u_0+\int_0^te^{(t-s)\Delta} ({\chi_{\omega}}v)\,ds=0\right\} \end{equation} $$

and

(TP)M2:TM2mingUM2t>0:u(x,t;g)=0for a.e.xΩ,$$ \begin{equation} (TP)^2_M:\;\; T^2_M\equiv \min\limits_{g\in\mathscr{U}^2_M} \left\{t\gt0:\;\;u(x,t;g)=0\;\;\mbox{for a.e.}\;\; x\in \Omega\right\}, \end{equation} $$

where u (·, · ;g) is the solution to

{tuΔu=0,inΩ×+,u=gχΓ,onΩ×+,u(0)=u0,inΩ.$$ \begin{equation} \begin{cases} \partial_t u- \Delta u=0,\ &\text{in}\ \Omega\times \mathscr {R}^+,\\ u=g\chi_{\Gamma},\ &\text{on}\ \partial\Omega\times\mathscr {R}^+,\\ u(0)=u_0,\ &\text{in}\ \Omega. \end{cases} \end{equation} $$

Any solution of (TP)Mi $(TP)_{M}^i$ , i=1, 2, is called a minimal time control to this problem. According to Theorem 1 and Theorem 3.3 in [12], problem (TP)M1 $(TP)^1_M$ has solutions. By Theorem 2, using the same arguments as those in the proof of Theorem 3.3 in [12], we can verify that there is gUM2 $g\in\mathscr{U}^2_M$ such that for some t > 0, u(x, t; g)=0 for a.e. x ∈ Ω.

Lemma 9

Problem(TP)M2 $(TP)^2_M$ has solutions.

Proof

Let {tn}n≥1, with tnTM2 $t_n\searrow T_M^2$ , and gnUM2 $g_n\in \mathscr{U}^2_M$ be such that u(x, tn; gn)=0 over Ω. Hence, on a subsequence,

gng*weakly star inL(Ω×(0,t1)).$$ \begin{equation} g_n\longrightarrow g^*\;\;\mbox{weakly star in}\;\; L^\infty(\partial\Omega\times(0,t_1)). \end{equation} $$

It suffices to show that

un(x,tn)u(x,tn;gn)u*(x,TM2)u(x,TM2;g*),for allxΩ.$$ \begin{equation} u_n(x,t_n)\equiv u(x,t_n; g_n)\longrightarrow u^*(x,T_M^2)\equiv u(x, T_M^2; g^*),\ \mbox{for all}\ x\in \Omega. \end{equation} $$

For this purpose, let G(x, y, t) be the Green’s function for Δ−∂t in Ω  ×  ℝ with zero lateral Dirichlet boundary condition. [6, Theorems 1.3 and 1.4] and [6, p. 643] show that for gUM2 $g\in \mathscr{U}^2_M$ and (x, t) ∈ Ω  ×  (0, T),

u(x,t;g)=etu00tΩGνq(x,q,ts)χΓ(q,s)g(q,s)dσqds$$ \begin{equation} u(x,t ;g)=e^{t\bigtriangleup}u_0- \int_{0}^{t}\int_{\partial\Omega}\tfrac{\partial G}{\partial\nu_q}(x,q,t-s)\,\chi_{\Gamma}(q, s)g(q,s)\,d\sigma_{q}ds \end{equation} $$

and

0TΩ|Gνq(x,q,τ)|2dσqdτ<+,whenxΩ,T>0.$$ \begin{equation} \int_0^T\int_{\partial\Omega}|\tfrac{\partial G}{\partial\nu_q}(x,q,\tau)|^2\,d\sigma_q d\tau\lt +\infty,\ \text{when}\ x\in\Omega,\ T\gt0. \end{equation} $$

Also, by standard interior parabolic regularity there is N=N(n,ε) with

|u(x,t;g)u(x,s;g)|N|ts|(gL(Ω×(0,T))+u0L2(Ω))$$ \begin{equation} |u(x,t;g)-u(x,s;g)|\le N|t-s|\left(\|g\|_{L^\infty(\partial\Omega\times (0,T))}+\|u_0\|_{L^2(\Omega)}\right) \end{equation} $$

when d(x, ∂Ω) > 3 $\sqrt{\varepsilon}$ and t > s ≥ ε. Now, when x ∈ Ω with d(x, ∂Ω) > 3 $\sqrt{\varepsilon}$ , it holds that

|un(x,tn)u*(x,TM2)||un(x,tn)un(x,TM2)|+|un(x,TM2)u*(x,TM2)|.$$ \begin{equation}|u_n(x,t_n)-u^*(x,T_M^2)|\leq |u_n(x,t_n)-u_n(x,T_M^2)|+ |u_n(x,T_M^2)-u^*(x,T_M^2)|. \end{equation} $$

This, along with (26), (28), (29) and (30) indicates that (27) holds for all x ∈ Ω with d(x, ∂Ω) > 3 $\sqrt{\varepsilon}$ . Since ε > 0 is arbitrary, (27) follows at once. □

Now, we can use the same methods as those in [14], as well as in Lemma 9, to get the following consequences of Theorems 1 and 2 respectively.

Corollary 10

Problem(TP)M1 $(TP)^1_M$ has the bang-bang property: any minimal time controlvsatisfies that |v(x, t)|=Mfor a.e. (x,t)ω×(0,TM1) $(x,t)\in \omega\times (0, T^1_M)$ . Consequently, this problem has a unique minimal time control.

Corollary 11

The problem(TP)M2 $(TP)^2_M$ has the bang-bang property: any minimal time boundary controlgsatisfies that |g(x, t)|=M for a.e. (x,t)Γ×(0,TM2) $(x,t)\in \Gamma\times (0, T^2_M)$ . Consequently, this problem has a unique minimal time control.

Definition of the Time Optimal Control Problems and Main Results

Next, we make use of Theorems 1 and 2 to study the bang-bang property for the time optimal control problems where the interest is on retarding the initial time of the action of a control with bounded L-norm. These problems are usually called the second type of time optimal control problems and are stated as follows: Let T > 0 and M > 0. Write ω and Γ for the sets given in Problems (TP)M1 $(TP)_M^1$ and (TP)M2 $(TP)_M^2$ respectively. Consider the controlled heat equations:

{tuΔu=χωχ(τ,T)v,inΩ×(0,T],u=0,onΩ×[0,T],u(0)=u0,inΩ$$ \begin{equation} \begin{cases} \partial_tu- \Delta u=\chi_{\omega}\chi_{(\tau,T)}v, &\text{in}\ \Omega\times (0,T],\\ u=0, &\text{on}\ \partial\Omega\times [0,T],\\ u(0)=u_0,\ &\text{in}\ \Omega \end{cases} \end{equation} $$

and

{tuΔu=0,inΩ×(0,T],u=χΓχ(τ,T)g,onΩ×[0,T],u(0)=u0,inΩ,$$ \begin{equation} \begin{cases} \partial_tu-\Delta u=0, &\text{in}\ \Omega\times (0,T],\\ u=\chi_{\Gamma}\chi_{(\tau,T)}g, &\text{on}\ \partial\Omega\times [0,T],\\ u(0)=u_0,\ &\text{in}\ \Omega, \end{cases} \end{equation} $$

where u0 ∈ L2(Ω). Write accordingly u (· ;χ(τ, T)v) and u (· ;χ(τ, T)g) for the solutions to equation (31) corresponding to χ(τ, T)v, and to equation (32) corresponding to χ(τ, T)g. Define the following control constraint sets:

UM,T1={vmeasurable onΩ×(0,T):|v(x,t)|Mfor a.e.(x,t)Ω×(0,T)}.UM,T2={gmeasurable onΩ×(0,T):|g(x,t)|Mfor a.e.(x,t)Ω×(0,T)}.$$ \begin{equation*}\nonumber \mathscr{U}^1_{M,T}=\{v\ \text{measurable on}\ \Omega \times(0,T): |v(x,t)|\leq M\ \text{for a.e.}\ (x,t)\in\Omega\times(0,T)\}.\\ \mathscr{U}^2_{M,T}=\{g\ \text{measurable on}\ \partial\Omega \times(0,T): |g(x,t)|\leq M\ \text{for a.e.}\ (x,t)\in\partial\Omega\times(0,T)\}. \end{equation*} $$

Consider the time optimal control problems:

(TP)M,T1:τM,T1supvUM,T1{τ[0,T):u(T;χ(τ,T)v)=0}$$ \begin{equation}(TP)^1_{M,T}:\ \tau^1_{M,T}\equiv \sup_{v\in\mathscr{U}^1_{M,T}} \left\{\tau\in[0,T):\;\;u(T; \chi_{(\tau,T)}v)=0\right\} \end{equation} $$

and

(TP)M,T2:τM,T2supgUM,T2{τ[0,T):u(T;χ(τ,T)g)=0}.$$ \begin{equation}(TP)^2_{M,T}:\;\; \tau^2_{M,T}\equiv \sup_{g\in\mathscr{U}^2_{M,T}} \left\{\tau\in[0,T):\;\;u(T; \chi_{(\tau,T)}g)=0\right\}. \end{equation} $$

Any solution of (TP)T,Mi $(TP)_{T,M}^i$ , i=1, 2, is called an optimal control to the corresponding problem.

Now, we can use the same arguments as those in the proof of Theorem 3.4 in [11] to get the following consequences of Theorem 1 and Theorem 2 respectively:

Corollary 12

Any optimal controlv* to Problem (TP)M,T1 $(TP)^1_{M,T}$ , if it exists, satisfies the bang-bang property: |v*(x, t)|=Mfor a.e. (x,t)ω×(τM,T1,T) $(x,t)\in \omega\times(\tau^1_{M,T}, T)$ .

Corollary 13

Any optimal controlg* to Problem(TP)M,T2 $(TP)^2_{M,T}$ , if it exists, satisfies the bang-bang property: |g*(x, t)|=Mfor a.e. (x,t)Γ×(τM,T2,T) $(x,t)\in \Gamma\times(\tau^2_{M,T}, T)$ .

Remark 5

By Theorem 1 (See also Remark 1) and the energy decay property for the heat equation, one can easily prove the following: for a fixed M > 0, there isvUM,T1 $v\in \mathscr{U}^1_{M,T}$ such thatu(T;χ(0,T)v)=0, whenTis large enough (such a controlvis called an admissible control); while for a fixedT > 0, the same holds whenMis large enough. The same can be said about Problem(TP)M,T2 $(TP)^2_{M,T}$ because of Theorem 2 (See also Remark 4). In the case where Problem(TP)M,T1 $(TP)^1_{M,T}$ has admissible controls, one can easily prove the existence of time optimal controls to this problem. In the case when Problem(TP)M,T2 $(TP)^2_{M,T}$ has admissible controls, one can make use of the similar method in the proof of Lemma 9 to verify the existence of time optimal controls for this problem.

Definition of the Minimal Norm Control Problems and Main Results

In this section, we apply Theorems 1 and 2 to get the bang-bang property for the minimal norm control problems; they are stated as follows. Let 𝒟 and 𝒥 be the subsets given at the beginning of this section. Let u0 ∈ L2(Ω), we define two control constraint sets as follows:

VD={vL(Ω×(0,T)):u(T;u0,v)=0}$$ \begin{equation*} \mathscr{V}_{\mathscr{D}}=\left\{v\in L^\infty (\Omega\times(0,T)): u(T; u_0, v)=0 \right\} \end{equation*} $$

and

VJ={gL(Ω×(0,T)):u(T;u0,g)=0}.$$ \begin{equation*}\nonumber \mathscr{V}_{\mathscr{J}}=\left\{g\in L^\infty (\partial\Omega\times(0,T)):\;u(T;u_0,g)=0 \right\}. \end{equation*} $$

Consider the minimal norm control problems:

(NP)D:MDmin{vL(Ω×(0,T)):vVD}$$ \begin{equation}(NP)_{\mathscr{D}}:\;\;\;\;M_{\mathscr{D}}\equiv \min\left\{\|v\|_{L^\infty(\Omega\times(0,T))}:\;\; v\in\mathscr{V}_{\mathscr{D}}\right\} \end{equation} $$

and

(NP)J:MJmin{gL(Ω×(0,T)):gVJ}.$$ \begin{equation}(NP)_{\mathscr{J}}:\;\;\;\;M_{\mathscr{J}}\equiv \min\left\{\|g\|_{L^\infty(\partial\Omega\times(0,T))}:\;\; g\in\mathscr{V}_{\mathscr{J}}\right\}. \end{equation} $$

Any solution of (NP)𝒟 (or NP)𝒥 is called a minimal norm control to this problem. According to Corollary 8, the sets VD $\mathscr{V}_{\mathscr{D}}$ and VJ $\mathscr{V}_{\mathscr{J}}$ are not empty. Since VD $\mathscr{V}_{\mathscr{D}}$ is not empty, it follows from the standard arguments that Problem (NP)𝒟 has solutions. Because VD $\mathscr{V}_{\mathscr{D}}$ is not empty, by using the similar arguments as those in the proof of Lemma 9, we can justify that Problem (NP)𝒥 has solutions.

We can use the same methods as those in [11] to get the following consequences of Theorem 1 and Theorem 2 respectively:

Corollary 14

Problem (NP)𝒟has the bang-bang property: any minimal norm controlvsatisfies that |v(x,t)| = M𝒟for a.e. (x,t) ∈ 𝒟. Consequently, this problem has a unique minimal norm control.

Corollary 15

The problem (NP)𝒥has the bang-bang property: any minimal norm boundary-controlgsatisfies that |g(x,t)| = M𝒥for a.e. (x,t) ∈ 𝒥. Consequently, this problem has a unique minimal norm control.

Open problems

In tis section we will establish the heat equation with similar conditions to what we studied before, but in this case we will require it to verify other type of boundary conditions instead of Dirichlet boundary conditions.

Let Ω be a bounded Lipschitz domain in ℝn and consider the following heat equation,

{tuΔu=0,inΩ×(0,1),νu=0,onΩ×(0,T),u(0)=u0,inΩ,$$ \begin{equation} \left\{\begin{array}{l} \partial_tu-\Delta u=0,\quad\mbox{in $\Omega\times(0,1)$,} \\\frac{\partial}{\partial\nu}u=0, \quad\mbox{on $\partial\Omega\times(0,T)$,} \\\displaystyle u(0)=u_0, \quad\mbox{in $\Omega$,} \end{array}\right. \end{equation} $$

with Neumann boundary condition and

{tuΔu=0,inΩ×(0,1),νu+αu=0,onΩ×(0,T),u(0)=u0,inΩ,$$ \begin{equation} \left\{\begin{array}{l} \partial_tu-\Delta u=0,\quad\mbox{in $\Omega\times(0,1)$,} \\\frac{\partial}{\partial\nu}u+\alpha u=0, \quad\mbox{on $\partial\Omega\times(0,T)$,} \\\displaystyle u(0)=u_0, \quad\mbox{in $\Omega$,} \end{array}\right. \end{equation} $$

with Robin boundary condition, where α ∈ ℝ and u0 in L2(Ω).

We proved two obserbavility inequalities (Theorems 1 and 2) for these kind of equations over measurable sets with Dirichlet boundary conditions, but if we change that condition to now use Neumann or Robin conditions, would we be able to prove some similar observability inequalities? And, if that’s the case, could we apply them to prove some bang-bang properties?

The idea of facing these questions is to spread our mathematical knowledge about this kind of problems and also to discover new interesting ways or limitations in the techniques we are used to working with. It could also be physically interesting because of the physical meaning of these new boundary conditions, as we will see now.

The Dirichlet boundary condition states that we have a constant temperature at the boundary. This can be considered as a model of an ideal cooler in a good contact having infinitely large thermal conductivity.

With the Neumann boundary condition case for the heat flow, we can say that we have a constant heat flux at the boundary or that it corresponds to a perfectly insulated boundary. If the flux is equal to zero, the boundary condition describes the ideal heat insulator with the heat diffusion. For the Laplace equation and drum modes, we could think this corresponds to allowing the boundary to flap up and down but not move otherwise.

Finally, the Robin boundary condition is the mathematical formulation of Newton’s law of cooling where the heat transfer coefficient α is utilized. The heat transfer coefficient is determined by details of the interface structure (sharpness, geometry) between two media. This law describes the boundary between metals and gas quite well and is good for the convective heat transfer.

Appendix

Here, we will give the definition of a Lipschitz domain and complete the proof of the equation (24) that appeared in the proof of Corollary 8.

Definition 2

Let Ω be a bounded domain in ℝn. øm is a Lipschitz domain (sometimes called strongly Lipschitz or Lipschitz graph domains) with constants m and ρ when for each point p on the boundary of Ω there is a rectangular coordinate system x=(x′, xn) and a Lipschitz function φ: ℝn−1 → ℝ verifying

ϕ(0)=0,|ϕ(x1)ϕ(x2)|m|x1x2|,for allx1,x2n1,$$ \begin{equation} \phi(0')=0,\quad |\phi(x_1')-\phi(x_2')|\le m|x_1'-x_2'|,\ \text{for all}\ x_1', x_2'\in\mathbb{R}^{n-1}, \end{equation} $$

p=(0′, 0) on this coordinate system and

Zm,ϱΩ={(x,xn):|x|<ϱ,ϕ(x)<xn<2mϱ},Zm,ϱΩ={(x,ϕ(x)):|x|<ϱ},$$ \begin{equation} \begin{split} Z_{m,\varrho}\cap\Omega=\{(x',x_n) : |x'|\lt \varrho,\ \phi(x')\lt x_n \lt 2m\varrho \},\\ Z_{m,\varrho}\cap\partial\Omega= \{(x',\phi(x')) : |x'|\lt \varrho\}, \end{split} \end{equation} $$

where Zm, ρ=Bρ  ×  (−2mρ, 2mρ).

Proof

(Proof of (24)] For each (p, τ) ∈ ∂Ω  ×  ℝ and fixed ξ > 0, we define

Γ(p)={xΩ:|xp|(1+ξ)d(x,Ω)},Γ(p,τ)={(x,t)Ω×(0,T):|xp|+|tτ|(1+ξ)d(x,Ω)}.$$ \begin{equation*} \Gamma(p)=\{x\in\Omega: |x-p|\le\left(1+\xi\right)d(x,\partial\Omega)\},\\ \Gamma(p,\tau)=\{(x,t)\in\Omega\times (0,T): |x-p|+\sqrt{|t-\tau|}\le\left(1+\xi\right)d(x,\partial\Omega)\}. \end{equation*} $$

The later are called respectively elliptic and parabolic non-tangential approach regions from the interior of Ω  ×  (0, T) to (p, τ). In particular,

Γ(p)×{τ}Γ(p,τ),for all(p,τ)Ω×(0,T).$$ \begin{equation*} \Gamma(p)\times\{\tau\}\subset\Gamma(p,\tau),\ \text{for all}\ (p,\tau)\in\partial\Omega\times (0,T). \end{equation*} $$

When u:Ω → ℝ or u:Ω× (0,T) → ℝ (or ℝn, define the elliptic and parabolic non-tangential maximal function of u in ∂Ω  ×  (0, T) as

u(p)=supxΓ(p)|u(x)|,u(p,τ)=sup(x,t)Γ(p,τ)|u(x,t)|,whenpΩandτ(0,T).$$ \begin{equation*} u^\ast(p)=\sup_{x\in\Gamma(p)}|u(x)|,\quad u^\sharp(p,\tau)=\sup_{(x,t)\in\Gamma(p,\tau)}|u(x,t)|,\ \text{when}\ p\in\partial\Omega\ \text{and}\ \tau\in (0,T). \end{equation*} $$

Let η > 0 be fixed such that [T − l1, T − l]⊂[2η, T − 2η], with l and l1 as defined in Corollary 8. Denote by u the solution to

{tuΔu=0,inΩ×(0,T),u=gχγ,onΩ×(0,T),u(0)=u0,inΩ.$$ \begin{equation*} \begin{cases} \partial_tu-\Delta u=0,\ &\text{in}\ \Omega\times(0,T),\\ u=g\chi_{\mathscr{M}}\equiv\gamma,\ &\text{on}\ \partial\Omega\times(0,T),\\ u(0)=u_0,\ &\text{in}\ \Omega. \end{cases} \end{equation*} $$

(See the beginning of Section 3 for the definition of the solution to this equation.)

Let γε in C01(Ω×(0,T)) $C^1_0(\partial\Omega\times(0,T))$ be a regularization of γ in ∂Ω  ×  [0, T] such that

γεL(Ω×[0,T])+εγεC1(Ω×[0,T])γL(Ω×[0,T]),supp(γε)Ω×[η,Tη]$$ \begin{equation*} \|\gamma^\varepsilon\|_{L^\infty(\partial\Omega\times[0,T])}+\varepsilon\,\|\gamma^\varepsilon\|_{C^1(\partial\Omega\times[0,T])} \leq \|\gamma\|_{L^\infty(\partial\Omega\times[0,T])},\\ \text{supp}(\gamma^\varepsilon)\subset\partial\Omega\times[\eta,T-\eta] \end{equation*} $$

and let vε be the solution to

{tvεΔvε=0,inΩ×(0,T),vε=γε,onΩ×(0,T),vε(0)=0,inΩ.$$ \begin{equation*} \begin{cases} \partial_tv^{\varepsilon}-\Delta v^{\varepsilon}=0,\ &\text{in}\ \Omega\times(0,T),\\ v^{\varepsilon}=\gamma^{\varepsilon},\ &\text{on}\ \partial\Omega\times(0,T),\\ v^{\varepsilon}(0)=0,\ &\text{in}\ \Omega. \end{cases} \end{equation*} $$

From [6, Theorem 3.2] [3, Theorem 6.1] or [4, Theorem 2.9]

vεL(Ω×[0,T])+ε(vε)L2(Ω×[0,T])γL(Ω×[0,T]),$$ \begin{equation} \|v^\varepsilon\|_{L^\infty(\partial\Omega\times[0,T])}+\epsilon \,\|\left(\nabla v^\epsilon\right)^\sharp\|_{L^2(\partial\Omega\times [0,T])}\le\|\gamma\|_{L^\infty(\partial\Omega\times[0,T])}, \end{equation} $$

and the limits

lim(x,t)(p,τ)(x,t)Γ(p,τ)vε(x,t)=vε(p,τ)$$ \begin{equation*} \lim_{\underset{(x,t)\in\Gamma(p,\tau)}{(x,t)\rightarrow (p,\tau)}}\nabla v^\epsilon(x,t)= \nabla v^\epsilon(p,\tau) \end{equation*} $$

exist and are finite for a.e. (p, τ) in ∂Ω  ×  (0, T). Also, vε ∈ C (Ω  ×  [0, T])∩ C(Ω  ×  [0, T]), vε = 0 for tη, and vε = 0 on ∂Ω  × (Tη, T]. Moreover, the Hölder regularity up to the boundary for bounded solutions to parabolic equations with zero local lateral Dirichlet data, shows that there are positive constants N = N(m, ρ, η) and α = α(m, ρ), with α ∈ (0, 1), such that

|vε(x1,t1)vε(x2,t2)|N[|x1x2|2+|t1t2|]α/2γL(Ω×[0,T]),$$ \begin{equation} \begin{split} &|v^\varepsilon(x_1,t_1)-v^\varepsilon(x_2,t_2)| \leq N\left[|x_1-x_2|^2+|t_1-t_2|\right]^{\alpha/2} \|\gamma\|_{L^\infty(\partial\Omega\times[0,T])}, \end{split} \end{equation} $$

when x1,x2Ω¯,Tη2t1,t2T $x_1,x_2\in\overline{\Omega},\;T-\frac\eta 2\leq t_1,t_2\leq T$ [9, Theorems 6.28 and 6.32].

Let φ(t) = e(TtφT, t ∈ (0, T), where φT is in L2(Ω). From the regularity of caloric functions [5, Theorem 1.7]

φC([0,T];L2(Ω))C(Ω×[0,T))C(Ω¯×[0,T))$$ \begin{equation} \varphi\in C([0,T];L^2(\Omega))\cap C^\infty(\Omega\times[0,T))\cap C(\overline{\Omega}\times[0,T)) \end{equation} $$

and from [6, Theorems 1.3 and 1.4] or the proof of (40) and (41) in this appendix, there are N = N(m, ρ) and ε = ε(m, ρ, n) > 0 such that

(φ)L(0,Tδ;L2+ε(Ω))Ne1/δφTL2(Ω),$$ \begin{equation} \|(\nabla \varphi)^\ast\| _{L^\infty(0,T-\delta;\, L^{2+\varepsilon}(\partial\Omega))} \leq Ne^{1/\delta}\,\|\varphi_T\|_{L^2(\Omega)}, \end{equation} $$

when 0 < δ < T and the limit

limxpxΓ(p)φ(x,τ)=φ(p,τ),$$ \begin{equation} \lim_{\underset{x\in\Gamma(p)}{x\to p}}\nabla\varphi(x,\tau)= \nabla\varphi(p,\tau), \end{equation} $$

exists and is finite for a.e. p ∈ ∂Ω and for all τ ∈ (0, T). Now, let Ωj⊂Ωj+1⊂Ω, j = 1, be a sequence of C-domains approximating Ω as in [3, Lemma 2.2]. Set, uε = vε+ etΔu0. By Green’s formula,

ddtΩjuε(t)φ(t)dx=Ωjuενjφφνjuεdσj.$$ \begin{equation*} \tfrac{d}{dt}\int_{\Omega_j}u^\varepsilon(t)\varphi(t)\,dx = \int_{\partial\Omega_j}\tfrac{\partial u^\varepsilon}{\partial\nu_j}\,\varphi-\tfrac{\partial \varphi}{\partial\nu_j}\, u^\varepsilon\,d\sigma_j. \end{equation*} $$

Integrating the above identity over [δ, Tδ] for a fixed δ(0,η2) $\delta\in(0,\frac{\eta}{2})$ , we get

Ωjuε(Tδ)φ(Tδ)dxΩjuε(δ)φ(δ)dx=Ωj×(δ,Tδ)uενjφφνjuεdσjdt.$$ \begin{equation} \begin{split} \int_{\Omega_j}u^\varepsilon(T-\delta)\varphi(T-\delta)\,dx -\int_{\Omega_j}u^\varepsilon(\delta)\varphi(\delta)\,dx \\ =\int_{\partial\Omega_j\times(\delta,T-\delta)}\tfrac{\partial u^\varepsilon}{\partial\nu_j}\,\varphi-\tfrac{\partial \varphi}{\partial\nu_j}\, u^\varepsilon\,d\sigma_jdt. \end{split} \end{equation} $$

Recall that uε(δ) = eδΔu0 and let j → +∞ in (42) with ε and δ being fixed. Then, (37), (39), (41) and the dominated convergence theorem show that

Ωuε(Tδ)φ(Tδ)dx=Ω(eδΔu0)φ(δ)dxΩ×(δ,Tδ)γεφνdσdt.$$ \begin{equation*} \int_{\Omega}u^\varepsilon(T-\delta)\varphi(T-\delta)\,dx =\int_{\Omega}(e^{\delta\Delta}u_0)\varphi(\delta)\,dx -\int_{\partial\Omega\times(\delta,T-\delta)}\gamma^\varepsilon\tfrac{\partial \varphi}{\partial\nu}\,d\sigma dt. \end{equation*} $$

Because γε is supported in [η, Tη], the later is the same as

Ωuε(Tδ)φ(Tδ)dx=Ω(eδΔu0)φ(δ)dxΩ×(η,Tη)γεφνdσdt,$$ \begin{equation} \int_{\Omega}u^\varepsilon(T-\delta)\varphi(T-\delta)\,dx =\int_{\Omega}(e^{\delta\Delta}u_0)\varphi(\delta)\,dx -\int_{\partial\Omega\times(\eta,T-\eta)}\gamma^\varepsilon\tfrac{\partial \varphi}{\partial\nu}\,d\sigma dt, \end{equation} $$

when 0 < δ < η/8. Next, from (38),

uε(Tδ)=vε(Tδ)+e(Tδ)Δu0=vε(T)+eTΔu0+O(δα/2),$$ \begin{equation*} u^\varepsilon(T-\delta)=v^\varepsilon(T-\delta)+e^{(T-\delta) \Delta}u_0=v^\varepsilon(T)+e^{T\Delta}u_0+O(\delta^{\alpha/2}), \end{equation*} $$

uniformly for xΩ¯ $x\in\overline{\Omega}$ , when 0 < δ < η/8. Hence, after letting δ → 0 in (43), we get

Ω(vε(T)+eTΔu0)φ(T)dx=Ωu0φ(0)dxΩ×(η,Tη)γεφνdσdt.$$ \begin{equation*} \int_{\Omega}(v^\varepsilon(T)+e^{T\Delta}u_0)\varphi(T)\,dx =\int_{\Omega}u_0\varphi(0)\,dx-\int_{\partial\Omega\times(\eta ,T-\eta)}\gamma^\varepsilon\tfrac{\partial\varphi}{\partial\nu}\, d\sigma dt. \end{equation*} $$

Also, from (37) and (38), vε converges uniformly over Ω¯×[Tη/2,T] $\overline{\Omega} \times[T-\eta/2,T]$ to some continuous function v˜ $\widetilde {v}$ as ε → 0. We claim that v˜=v $\widetilde{v}=v$ . If it is the case, we get after letting ε → 0 in the last equality, that

Ωu(T)φ(T)dx=Ωu0φ(0)dxΩ×(η,Tη)γφνdσdt,$$ \begin{equation*} \int_{\Omega}u(T)\varphi(T)\,dx =\int_{\Omega}u_0\varphi(0)\,dx-\int_{\partial\Omega\times(\eta ,T-\eta)}\gamma\tfrac{\partial\varphi}{\partial\nu}\, d\sigma dt, \end{equation*} $$

because γε(p, τ) → γ(p, τ) for a.e. (p, τ) ∈ ∂Ω  ×  (0, T), (39) and

supp(γε)supp(γ)Ω×[η,Tη].$$ \begin{equation*} \text{supp}(\gamma^\varepsilon)\cup \text{supp}(\gamma)\subset \partial\Omega\times[\eta,T-\eta]. \end{equation*} $$

Recalling that γ = gχ𝓜, we get

Ωu(T)φ(T)dx=Ωu0φ(0)dxΩ×(0,T)gχφνdσdt.$$ \begin{equation*} \int_{\Omega}u(T)\varphi(T)\,dx =\int_{\Omega}u_0\varphi(0)\,dx-\int_{\partial\Omega\times(0 ,T)}g\chi_{\mathscr{M}}\,\tfrac{\partial\varphi}{\partial\nu}\, d\sigma dt. \end{equation*} $$

Hence, (24) is proved.

To verify that v˜=v $\widetilde{v}=v$ over Ω¯×[0,T] $\overline{\Omega}\times[0,T]$ , observe that because vɛv is the unique solution to

{tuΔu=0,inΩ×(0,T),u=γεγ,onΩ×(0,T),u(0)=0,inΩ,$$ \begin{equation*} \begin{cases} \partial_tu-\Delta u =0,\ &\text{in}\ \Omega\times(0,T),\\ u=\gamma^{\epsilon}-\gamma,\ &\text{on}\ \partial\Omega\times(0,T),\\ u(0)=0,\ &\text{in}\ \Omega, \end{cases} \end{equation*} $$

whose parabolic non-tangential maximal function is in L2(∂Ω  ×  (0, T)) (See [6, Theorem 3.2]), it holds that

(vεv)L2(Ω×(0,T))NγεγL2(Ω×(0,T)).$$ \begin{equation} \|(v^\varepsilon-v)^\sharp\|_{L^2(\partial\Omega\times(0,T))} \leq N\|\gamma^\varepsilon-\gamma\|_{L^2(\partial\Omega\times(0,T))}. \end{equation} $$

For fixed p in ∂Ω, we may assume that p = (0′, 0) and that near p,

ΩZm,ϱ={(x,xn):ϕ(x)<xn<2mϱ,|x|ϱ},$$ \begin{equation*} \Omega\cap Z_{m,\varrho}=\{(x',x_n):\phi(x')\lt x_n\lt 2m\varrho,\; |x'|\leq \varrho\}, \end{equation*} $$

with φ as in (35) and (36). Then,

0TBρϕ(y)ϕ(y)+mρ|F(y,yn,t)|2dydyndtmρ0TBϱF(y,yn,t)2dydtmρΩ×(0,T)F(p,t)2dσdt,$$ \begin{equation*} \begin{split} \int_0^T\int_{B'_{\rho}}\int_{\phi(y')}^{\phi(y')+m\rho} |F(y',y_n,t)|^2\,dy'dy_ndt \\ \leq m\rho\int_0^T\int_{B'_\varrho}F^\sharp(y',y_n,t)^2\,dy'dt \leq m\rho\int_{\partial\Omega\times(0,T)}F^\sharp(p,t)^2\,d\sigma dt, \end{split} \end{equation*} $$

for all functions F. The above estimate, a covering argument and (44) show that

vεvL2(Ωmϱ×(0,T))NγεγL2(Ω×(0,T)),$$ \begin{equation} \|v^\varepsilon-v\|_{L^2(\Omega_{m\varrho}\times (0,T))}\le N\|\gamma^\varepsilon-\gamma\|_{L^2(\partial\Omega\times(0,T))}, \end{equation} $$

with Ωη = {x ∈ Ω: d(x, ∂Ω) ≤ η}. Recalling that vε = v = 0 for tη, the local boundedness properties of solutions to parabolic equations [9, Theorem 6.17] show that,

|(vεv)(x,τ)|(BR20(x)×[τR2202,τ]|vεv|2dyds)1/2,$$ \begin{equation*} |(v^\varepsilon-v)(x,\tau)|\leq \left(-\!\!\!\!\!\!\!\!\! \int_{B_{\frac{R}{20}}(x) \times[\tau-\frac{R^2}{20^2},\tau]}|v^\varepsilon-v|^2\,dyds \right)^{1/2}, \end{equation*} $$

when x ∈ ∂ΩR, 0 ≤ τT, and taking R<mϱ20 $R\lt\frac{m\varrho}{20}$ above, we find from (45) that

vεvL(ΩR×{0}ΩR×[0,T])NRγεγL2(Ω×(0,T)).$$ \begin{equation*} \|v^\varepsilon-v\|_{L^\infty(\Omega^R\times\{0\}\cup\partial\Omega ^R\times[0,T])} \leq N_R\left\|\gamma^\varepsilon-\gamma\right\|_{L^2(\partial \Omega\times(0,T))}. \end{equation*} $$

By the maximum principle and the above estimate

vεvL(ΩR×[0,T])NRγεγL2(Ω×(0,T))0,asε0,$$ \begin{equation*} \|v^\varepsilon-v\|_{L^\infty(\Omega^R\times[0,T])} \leq N_R\left\|\gamma^\varepsilon-\gamma\right\|_{L^2(\partial \Omega\times(0,T))}\longrightarrow 0,\;\;\text{as}\;\;\varepsilon\rightarrow 0, \end{equation*} $$

which shows that v˜=v $\widetilde{v}=v$ in Ω¯×[0,T] $\overline{\Omega}\times[0,T]$ . □

eISSN:
2444-8656
Language:
English
Publication timeframe:
2 times per year
Journal Subjects:
Life Sciences, other, Mathematics, Applied Mathematics, General Mathematics, Physics