We discuss a control constrained boundary optimal control problem for the Boussinesq-type system arising in the study of the dynamics of an arterial network. We suppose that the control object is described by an initial-boundary value problem for $ 1D $ system of pseudo-parabolic nonlinear equations with an unbounded coefficient in the principle part and the Robin-type of boundary conditions. The main question we study in this part of the paper is about the existence of optimal solutions and first-order optimality conditions.
Citation: Ciro D'Apice, Olha P. Kupenko, Rosanna Manzo. On boundary optimal control problem for an arterial system: First-order optimality conditions[J]. Networks and Heterogeneous Media, 2018, 13(4): 585-607. doi: 10.3934/nhm.2018027
[1] | Ciro D'Apice, Olha P. Kupenko, Rosanna Manzo . On boundary optimal control problem for an arterial system: First-order optimality conditions. Networks and Heterogeneous Media, 2018, 13(4): 585-607. doi: 10.3934/nhm.2018027 |
[2] | Alberto Bressan, Yunho Hong . Optimal control problems on stratified domains. Networks and Heterogeneous Media, 2007, 2(2): 313-331. doi: 10.3934/nhm.2007.2.313 |
[3] | Oliver Kolb, Simone Göttlich, Paola Goatin . Capacity drop and traffic control for a second order traffic model. Networks and Heterogeneous Media, 2017, 12(4): 663-681. doi: 10.3934/nhm.2017027 |
[4] | Rong Huang, Zhifeng Weng . A numerical method based on barycentric interpolation collocation for nonlinear convection-diffusion optimal control problems. Networks and Heterogeneous Media, 2023, 18(2): 562-580. doi: 10.3934/nhm.2023024 |
[5] | Dilip Sarkar, Shridhar Kumar, Pratibhamoy Das, Higinio Ramos . Higher-order convergence analysis for interior and boundary layers in a semi-linear reaction-diffusion system networked by a $ k $-star graph with non-smooth source terms. Networks and Heterogeneous Media, 2024, 19(3): 1085-1115. doi: 10.3934/nhm.2024048 |
[6] | Ling Zhang, Xuewen Tan, Jia Li, Fan Yang . Dynamic analysis and optimal control of leptospirosis based on Caputo fractional derivative. Networks and Heterogeneous Media, 2024, 19(3): 1262-1285. doi: 10.3934/nhm.2024054 |
[7] | Didier Georges . Infinite-dimensional nonlinear predictive control design for open-channel hydraulic systems. Networks and Heterogeneous Media, 2009, 4(2): 267-285. doi: 10.3934/nhm.2009.4.267 |
[8] | Huanhuan Li, Meiling Ding, Xianbing Luo, Shuwen Xiang . Convergence analysis of finite element approximations for a nonlinear second order hyperbolic optimal control problems. Networks and Heterogeneous Media, 2024, 19(2): 842-866. doi: 10.3934/nhm.2024038 |
[9] | Tong Yan . The numerical solutions for the nonhomogeneous Burgers' equation with the generalized Hopf-Cole transformation. Networks and Heterogeneous Media, 2023, 18(1): 359-379. doi: 10.3934/nhm.2023014 |
[10] | Vincent Renault, Michèle Thieullen, Emmanuel Trélat . Optimal control of infinite-dimensional piecewise deterministic Markov processes and application to the control of neuronal dynamics via Optogenetics. Networks and Heterogeneous Media, 2017, 12(3): 417-459. doi: 10.3934/nhm.2017019 |
We discuss a control constrained boundary optimal control problem for the Boussinesq-type system arising in the study of the dynamics of an arterial network. We suppose that the control object is described by an initial-boundary value problem for $ 1D $ system of pseudo-parabolic nonlinear equations with an unbounded coefficient in the principle part and the Robin-type of boundary conditions. The main question we study in this part of the paper is about the existence of optimal solutions and first-order optimality conditions.
The main goal of this paper is to study one class of optimal control problems (OCPs) for a viscous Boussinesq system arising in the study of the dynamics of cardiovascular networks. We consider the boundary control problem for a
$ Minimize J(g,h,η,u):=12∫ΩαΩ(u(T)−uΩ)2 dx+ν2∫T0∫Ω(ηxx)2 dxdt+12∫T0|∫ΩαQ(η(t)+r0uxt(t)−ηQ) dx|2 dt+12∫T0(βg|g|2+βh|h|2) dt $ | (1) |
subject to the constraints
$ \left\{ ηt+ηxu+ηux+12r0ux−νηxx=0 in Q,[u−(δux)x]t+12(u2)x+μηx=f in Q, \right. $ | (2) |
$ \left\{ η(0,⋅)=η0 in Ω,u(0,⋅)−(δ(⋅)ux(0,⋅))x=u0 in Ω, \right. $ | (3) |
$ \left\{ η(⋅,0)=η(⋅,L)=η∗ in (0,T),δ(0)˙ux(⋅,0)+σ0u(⋅,0)=g, in (0,T),δ(L)˙ux(⋅,L)+σ1u(⋅,L)=h, in (0,T),δ(L)ux(0,L)=δ(0)ux(0,0)=0 \right. $ | (4) |
and
$ \label{2.6} (g,h)\in G_{ad}\times H_{ad}\subset L^2(0,T)\times L^2(0,T). $ | (5) |
Here,
Optimal control problem (1)-(5) comes from the fluid dynamic models of blood flows in arterial systems. It is well known that the cardiovascular system consists of a pump that propels a viscous liquid (the blood) through a network of flexible tubes. The heart is one key component in the complex control mechanism of maintaining pressure in the vascular system. The aorta is the main artery originating from the left ventricle and then bifurcates to other arteries, and it is identified by several segments (ascending, thoracic, abdominal). The functionality of the aorta, considered as a single segment, is worth exploring from a modeling perspective, in particular in relationship to the presence of the aortic valve.
In the first part of our investigation (see [5]) we make use of the standard viscous hyperbolic system (see [2,21]) which models cross-section area
$ \frac{\partial S}{\partial t}+\frac{\partial (Su)}{\partial x}-\nu \frac{{{\partial }^{2}}S}{\partial {{x}^{2}}} = 0, $ | (6) |
$ \frac{\partial u}{\partial t}+u\frac{\partial u}{\partial x}+\frac{1}{\rho }\frac{\partial P}{\partial x} = f, $ | (7) |
where
$ \label{3.F2} \eta = r-r_{0} = \frac{1}{\sqrt{\pi} }(\sqrt{S}-\sqrt{S_{0}})\simeq \frac{S-S_{0}}{2\sqrt{\pi S_{0}}}, $ | (8) |
where
The fluid structure interaction is modeled using inertial forces, which gives the pressure law
$ P=Pext+βr20η+ρωh∂2η∂t2. $ | (9) |
Here,
This leads to the following Boussinesq system:
$ \label{3.F4} \left\{ ηt+ηxu+ηux+12r0ux−νηxx=0,ut+uux+2Ehρr20ηx+ρωhρηxtt=f, \right. $ |
where
As for the OCP that is related with the arterial system, we are interested in finding the optimal heart rate (HR) which leads to the minimization of the following cost functional
$ J=∫t0+Tpulset0|Pavg(t)−Pref|2dt=∫t0+Tpulset0|1L∫L0P(x,t)dx−Pref|2dt. $ | (10) |
The systolic period is taken to be consistently one quarter of
It is easy to note that relations (8)-(9) lead to the following representation for the cost functional (10)
$ \label{3.F9} J = \int_{t_0}^{t_0+T_{pulse}} \left| \frac{1}{L}\int_0^L P(x,t)\, dx-P_{ref}\right|^2\, dt\\ = \frac{1}{L^2} \int_{t_0}^{t_0+T_{pulse}} \left| \int_0^L \left(P_{ext}(t)+\frac{2Eh}{r_{0}^{2}}\eta(t,x) + \rho _{\omega }h\,\eta_{tt}(t,x) -L\,P_{ref}\right)\, dx\right|^2\, dt. $ | (11) |
Since
The research in the field of the cardiovascular system is very active (see, for instance the literature describing the dynamics of the vascular network coupled with a heart model, [2,9,10,12,15,16,17,18,19,20,21]). However, there seems to be no studies that focus on both aspects at the same time: a detailed description of the four chambers of the heart and on the spatial dynamics in the aorta. Some numerical aspects of optimizing the dynamics of the pressure and flow in the aorta as well as the heart rate variability, taking into account the elasticity of the aorta together with an aortic valve model at the inflow and a peripheral resistance model at the outflow, based on the discontinuous Galerkin method and a two-step time integration scheme of Adam-Bashfort, were recently treated in [3] for the Boussinesq system like (2). More broadly, theory and applications of optimization and control in spatial networks, basing on the different types of conservation laws have been extensively developed in literature, have been successfully applied to telecommunications, transportation or supply networks ([6,7]).
From mathematical point of view, the characteristic feature of the Boussinesq system (2) is the fact that it involves a pseudo-parabolic operator with unbounded coefficient in its principle part. In the first part of this paper it was shown that for any pair of boundary controls
Let
$ \|u{{\|}_{{{L}^{2}}(\Omega ,\delta\ dx)}} = {{\left( \int_{\Omega }{{{u}^{2}}}\delta\ dx \right)}^{1/2}} < +\infty . $ |
We set
$ (\varphi,\psi)_{V_0} = \left(\varphi^\prime,\psi^\prime\right)_H\ \forall\,\varphi, \psi\in V_0 $ |
and
$ (\varphi,\psi)_{V} = (\varphi,\psi)_H+\left(\varphi^\prime,\psi^\prime\right)_H\ \forall\, \varphi, \psi\in V, $ |
respectively.
We also make use of the weighted Sobolev space
$ \|u\|_{V_\delta} = \left(\int_\Omega \left(u^2+\delta(u^\prime)^2\right)\,dx\right)^{1/2} $ |
is finite. Note that due to the following estimate,
$ ‖u‖2V:=∫Ω(u2+(u′)2)dx≤max{1,δ−10}∫Ω(u2+δ(u′)2)dx=max{1,δ−10}‖u‖2Vδ. $ | (12) |
Recall that
Let us recall some well-known inequalities, that will be useful in the sequel (see [5]).
●
● (Friedrich's Inequality) For any
$ \label{4.F} \|u\|_H\le L\|u_x\|_H = L\|u\|_{V_0}. $ | (13) |
By
$ \|u\|_{L^2(0,T;V_0)}: = \left(\int_0^T \|u(t)\|^2_{V_0}\,dt \right)^{1/2} < +\infty. $ |
By analogy we can define the spaces
$ \int_0^T\varphi(t)\left < \dot{u}(t),v\right > _{V^\ast;V}\,dt = -\int_0^T\dot{\varphi}(t)\left < u(t),v\right > _{V^\ast;V}\,dt,\ \ \ \ \forall\,v\in V, $ |
where
We also make use of the following Hilbert spaces
$ W0(0,T)={u∈L2(0,T;V0): ˙u∈L2(0,T;V∗0)},Wδ(0,T)={u∈L2(0,T;Vδ): ˙u∈L2(0,T;V∗δ)}, $ |
supplied with their common inner product, see [8,p. 473], for instance.
Remark 1. The following result is fundamental (see [8]): Let
(ⅰ)
$ \max\limits_{1\le t\le T} \|u(t)\|_H\le C_E\left(\|u\|_{L^2(0,T;V)}+\|\dot{u}\|_{L^2(0,T;V^\ast)}\right); $ |
(ⅱ) if
$ \label{1.3} \int_s^t\left(\left < \dot{u}(\gamma),v(\gamma)\right > _{V^\ast;V}+\left < u(\gamma),\dot{v}(\gamma)\right > _{V^\ast;V}\right)\,d\gamma = \left(u(t),v(t)\right)_H-\left(u(s),v(s)\right)_H $ | (14) |
for all
The similar assertions are valid for the Hilbert triplet
Let
$ f\in {{L}^{\infty }}(0,T;H),\ \ \mu \in {{L}^{\infty }}(0,T;V),\ \ {{\sigma }_{0}}\in {{L}^{\infty }}(0,T),\ \ {{\sigma }_{1}}\in {{L}^{\infty }}(0,T), $ | (15) |
$ {{\alpha }_{\Omega }}\in {{L}^{\infty }}(\Omega ),\ \ {{\alpha }_{Q}}\in {{L}^{\infty }}(Q),\ \ {{u}_{\Omega }}\in {{L}^{2}}(\Omega ),\ \ {{\eta }_{Q}}\in {{L}^{2}}(0,T;H), $ | (16) |
$ {{u}_{0}}\in {{V}_{\delta }},\ \ {{\eta }_{0}}\in H_{0}^{1}(\Omega ),\ \ {{r}_{0}}\in {{H}^{1}}(\Omega ), $ | (17) |
be given distributions. In particular,
We assume that the sets of admissible boundary controls
$ Gad={g∈L2(0,T): g0≤g≤g1 a.e. in (0,T)},Had={h∈L2(0,T): h0≤h≤h1 a.e. in (0,T)}, $ | (18) |
where
The optimal control problem we consider in this paper is to minimize the discrepancy between the given distributions
Definition 3.1. We say that, for given boundary controls
$ \eta (t) = w(t)+{{\eta }^{*}},\ \ \ \ w(\cdot )\in {{W}_{0}}(0,T),\ \ \ \ u(\cdot )\in {{W}_{\delta }}(0,T), $ | (19) |
$ \delta (L){{u}_{x}}(0,L) = 0,\ \ \ \ \ \ \ \ \delta (0){{u}_{x}}(0,0) = 0, $ | (20) |
$ {{\left( w(0),\chi \right)}_{H}} = {{\left( {{\eta }_{0}}-{{\eta }^{*}},\chi \right)}_{H}}\ \ \ \ \ \ \ \text{for all }\chi \in H, $ | (21) |
$ {{\left( u(0)-{{(\delta {{u}_{x}}(0))}_{x}},\chi \right)}_{{{V}_{\delta }}}} = {{\left( {{u}_{0}},\chi \right)}_{{{V}_{\delta }}}}\ \ \ \ \ \ \ \text{for all }\chi \in {{V}_{\delta }}, $ | (22) |
and the following relations
$ ⟨˙w(t),φ⟩V∗0;V0+((w(t)u(t))x,φ)H+ν(wx(t),φx)H +12(r0ux(t)+2η∗ux(t),φ)H=0, $ | (23) |
$ ⟨˙u(t),ψ⟩V∗δ;Vδ+∫Ωδ˙ux(t)ψxdx+(u(t)ux(t),ψ)H+(μ(t)wx(t),ψ)H +σ1(t)u(t,L)ψ(L)−σ0(t)u(t,0)ψ(0) =(f(t),ψ)H+h(t)ψ(L)−g(t)ψ(0) $ | (24) |
hold true for all
Remark 2. Let us mention that if we multiply the left- and right-hand sides of equations (23)-(24) by function
$ \int_{0}^{T}{{{\left\langle {{A}_{1}}(w(t),u(t)),\varphi (t) \right\rangle }_{V_{0}^{*};{{V}_{0}}}}dt = 0,\ \ \ \ \ \ \forall \varphi (\cdot )\in {{L}^{2}}(0,T;{{V}_{0}}),} $ | (25) |
$ \int_{0}^{T}{{{\left\langle {{A}_{2}}(w(t),u(t)),\psi (t) \right\rangle }_{V_{\delta }^{*};{{V}_{\delta }}}}dt = 0,\ \ \ \ \ \ \forall \psi (\cdot )\in {{L}^{2}}(0,T;{{V}_{\delta }}),} $ | (26) |
where
$ {{A}_{1}}(w,u) = \frac{\partial w}{\partial t}-\nu {{w}_{xx}}+{{w}_{x}}u+w{{u}_{x}}+\frac{1}{2}{{r}_{0}}{{u}_{x}}+{{\eta }^{*}}{{u}_{x}}\in V_{0}^{*}, $ | (27) |
$ {{A}_{2}}(w,u) = \left[ ∂∂t(u−(δux)x)+12(u2)x+μwx−fδ(0)˙ux(⋅,0)+σ0u(⋅,0)−gδ(L)˙ux(⋅,L)+σ1u(⋅,L)−h \right]\in V_{\delta }^{*}. $ | (28) |
Lemma 3.2 ([5]). Assume that the conditions (15)-(17) hold true. Let
$ (η(⋅),u(⋅))∈(W0(0,T)+η∗)×Wδ(0,T),w∈L∞(0,T;H)∩L2(0,T;H2(Ω)∩V0),˙w∈L2(0,T;H), u∈W1,∞(0,T;Vδ) $ | (29) |
and there exists a constant
$ \|w\|_{{{L}^{2}}(0,T;{{H}^{2}}(\Omega ))}^{2}+\|w\|_{{{L}^{\infty }}(0,T;H)}^{2}+\|\dot{w}\|_{{{L}^{2}}(0,T;H)}^{2}\le {{D}_{*}}, $ | (30) |
$ \|u\|_{{{L}^{\infty }}(0,T;{{V}_{\delta }})}^{2}+\|\dot{u}\|_{{{L}^{\infty }}(0,T;{{V}_{\delta }})}^{2}\le {{D}_{*}}. $ | (31) |
We also define the feasible set to the problem (1)-(5), (18) as follows:
$ \Xi = \left\{ (g,h,\eta ,u)\ \left| g∈Gad, h∈Had,η(t)=w(t)+η∗, w∈W0(0,T), u∈Wδ(0,T),(w(t),u(t)) satisfies relations (19)-(24)for all φ∈V0, ψ∈Vδ, and a.e. t∈[0,T],J(g,h,η,u)<+∞. \right. \right\} $ | (32) |
We say that a tuple
$ J\left(g^0,h^0,\eta^0,u^0\right) = \inf\limits_{(g,h,\eta,u)\in\Xi}J(g,h,\eta,u). $ |
In [5] it was shown that
While proving these hypotheses, the authors in [5] obtained a series of useful estimates for the weak solutions to initial-boundary value problem (2)-(4).
Lemma 3.3. [5,Lemmas 6.3 and 6.5 along with Remark 6.5] Let
$ \label{5.8} \|w(t)\|_H^2+\|u(t)\|^2_{V_\delta}\le C_1,\ \ \ \|\dot{w}(t)\|_{V_0^\ast}\le C_2,\ \ \ \|\dot{u}(t)\|_{V_\delta}\le C_3. $ | (33) |
In the context of solvability of OCP (18)-(5), the regularity of the solutions of the corresponding initial-boundary value problem (2)-(4) plays a crucial role.
Theorem 3.4 ([5]). The set of feasible solutions
Now we proceed with the result concerning existence of optimal solutions to OCP (1)-(5), (18).
Theorem 3.5. For each
$ f∈L∞(0,T;L2(Ω)), μ∈L∞(0,T;V), σ0∈L∞(0,T), σ1∈L∞(0,T),αΩ∈L∞(Ω), αQ∈R+, uΩ∈L2(Ω), ηQ∈W(0,T;H),u0∈Vδ, η0∈V0, r0∈H1(Ω), δ∈L1(Ω) $ |
the optimal control problem (1)-(5), (18) admits at least one solution
Proof. We apply for the proof the direct method of the calculus of variations. Let us take
$ {{\Xi }_{\lambda }} = \left\{ (g,h,\eta ,u)\in \Xi \ \ :\ \ J(g,h,\eta ,u)\le \lambda \right\}\ne \varnothing . $ |
Since the cost functional (1) is bounded below on
$ ‖ηxx‖2L2(0,T;L2(Ω))=‖wxx‖2L2(0,T;L2(Ω))≤‖w‖2L2(0,T;H2(Ω))≤D∗,‖uxt‖2L2(0,T;H)≤max{1,δ−10}‖˙u‖2L∞(0,T;Vδ)≤D∗. $ |
Therefore, within a subsequence, still denoted by the same index, we can suppose that
$ gn⇀g0 in L2(0,T), hn⇀h0 in L2(0,T),un→u0 strongly in L2(0,T;H),un∗⇀u0 weakly-∗ in L∞(0,T;Vδ),˙un⇀v weakly in L2(0,T;Vδ) and weakly-∗ in L∞(0,T;Vδ), $ |
where
$ \|{{u}_{n}}(t)\|_{{{V}_{\delta }}}^{2}\le {{C}_{1}}\ \ \ \ \text{for}\ \text{all}\ n\in \mathbb{N}\ \ \text{and}\ \ \text{for}\ \ \text{all}\ \ t\in [0,T], $ |
whence, passing to a subsequence, if necessary, we obtain
$ un(T,⋅)⇀u0(T,⋅) in Vδ,un(T,⋅)→u0(T,⋅) strongly in H $ |
due to the continuity of embedding
$ ηn(t,x)⇀η0(t,x) in V0, ˙u(t,x)⇀˙u0(t,x) in Vδ for a.e. t∈[0,T],(ηn(t,x)+r0(x)un xt(t,x)−ηQ)⇀(η0(t,x)+r0(x)u0xt(t,x)-ηQ)) in L1(Ω)for a.e. t∈[0,T],∫ΩaQ(ηn(t,x)+r0(x)un xt(t,x)−ηQ)dx→→∫ΩaQ(η0(t,x)+r0(x)un xt(t,x)−ηQ))dx for a.e. t∈[0,T],limn→∞∫T0(∫ΩaQ(ηn(t,x)+r0(x)un xt(t,x)−ηQ)dx)2 dt = ∫T0(∫ΩaQ(η0(t,x)+r0(x)un xt(t,x)−ηQ)))2 dt, $ |
we have
This section aims to prove a range of auxiliary results that will be used in the sequel. Throughout this section the tuple
The following proposition aims to prove rather technical result, however it is useful for substantiation of the first-order optimality conditions to the initial OCP (1)-(5).
Proposition 1. Let
$ u0[u0xxη0+2u0xη0x+η0xxu0]−(αQ)2∫Ω(η0−ηQ)dx∈L2(0,T;V∗),η0[u0xxη0+2u0xη0x+η0xxu0]∈L2(0,T;V∗). $ |
Proof. To begin with, let us prove that
$\eta^0\big[u^0_{xx}\eta^0+2u^0_x \eta^0_x+\eta^0_{xx}u^0\big]\in L^2(0,T;V^\ast).$ |
Obviously, in order to show that
$ u^0\big[u^0_{xx}\eta^0+2u^0_x \eta^0_x+\eta^0_{xx}u^0\big]- (\alpha_Q)^2\int_\Omega \left(\eta^0-\eta_Q\right)\,dx\in L^2(0,T;V^\ast) $ |
it would be enough to apply the similar arguments. Since
$ \left\|u^0_{xx}\eta^0+2u^0_x \eta^0_x+\eta^0_{xx}u^0\right\|_{V^\ast}\le \widetilde{C}\text{ for a.a. }t\in[0,T]. $ |
It should be noticed that as far as
$ u_{x}^{0}\in {{L}^{2}}(\Omega ;\delta \ dx) \hookrightarrow {{L}^{2}}(\Omega )\ \ \ \ \text{for}\ \text{a}.\text{a}.\ t\in [0;T], $ |
then
Also the fact that
$‖u0xx(t)η0(t)+2u0x(t)η0x(t)+η0xx(t)u0(t)‖V∗=sup‖v‖V≤1⟨u0xx(t)η0(t)+2u0x(t)η0x(t)+η0xx(t)u0(t),v⟩V∗;V=∫Ωu0xx(t)η0(t)vdx+2∫Ωu0x(t)η0x(t)vdx+∫Ωη0xx(t)u0(t)vdx≤‖η0(t)‖C(¯Ω)‖v‖V‖u0xx(t)‖V∗+‖η0x(t)‖L∞(Ω)‖u0x(t)‖H‖v‖H+‖u0‖C(¯Ω)‖ηxx(t)‖H‖v‖H≤‖v‖V×(‖η0(t)‖C(¯Ω)‖u0xx‖V∗+‖η0x(t)‖L∞(Ω)‖u0x(t)‖L2(Ω)+‖u0‖C(¯Ω)‖ηxx(t)‖L2(Ω))⏟C(t).$ |
It is clear that if only
$ \|u^0_{xx}\|_{V^\ast} = \left\|\frac{1}{\delta}\left((\delta u_x^0)_x-\delta_x u^0_x\right)\right\|_{V^\ast} \le \frac{1}{\delta_0}\left(\|(\delta u_x^0)_x\|_{V^\ast}+\|\delta_x u^0_x\|_{V^\ast}\right) $ | (34) |
and
$ ‖C(t)‖2L2(0;T)≤2δ20‖η0‖2C(0,T;H)∫T0(‖(δu0x)x‖2V∗+‖δxu0x‖2V∗)dt+2max{L,L−1}δ0∫T0‖η0x‖2V‖u0‖2Vδdt+‖u0‖2C(0,T;H)∫T0‖η0xx‖2Hdt≤2δ20‖η0‖2C(0,T;H)∫T0(‖(δu0x)x‖2V∗+‖δxu0x‖2V∗)dt+2max{L,L−1}δ0‖u0‖2W1,∞(0,T;Vδ)‖η0‖2L2(0,T;H2)+‖u0‖2C(0,T;H)‖η0‖2L2(0,T;H2). $ | (35) |
Let us show that the integrals
$∫T0‖δxu0x(t)‖2V∗dt=∫T0(sup‖v‖V≤1∫Ω|δx||u0x(t)||v|dx)2dt≤∫T0(sup‖v‖V≤1‖v‖C(¯Ω)‖δ‖V‖u(t)‖V)2dt≤c2(E)δ0‖v‖2V‖δ‖2V‖u‖2L2(0,T;Vδ)≤c2(E)Tδ0‖δ‖2V‖u‖2L∞(0,T;Vδ).$ |
Now, to estimate the second integral, we make use of the equation (2)
$∫T0‖(δu0x)x‖2V∗dt=∫T0(sup‖v‖V≤1∫Ω|(δu0x)xv|dx)2dt=∫T0(sup‖v‖V≤1∫Ω|[∫t0(f(s)−u0(s)u0x(s)−μ(s)η0x(s))ds+u0(t)+u0+(δ(u0)x)x]v|dx)2dt≤∫T02(sup‖v‖V≤1∫Ω|∫t0(f(s)v−u0(s)u0x(s)v−μ(s)η0x(s)v)ds|dx)2dt+∫T02(sup‖v‖V≤1∫Ω|(u0(t)+u0+(δ(u0)x)x)v|dx)2dt≤∫T02(sup‖v‖V≤1∫Ω∫T0|f(s)v−u0(s)u0x(s)v−μ(s)η0x(s)v)|dsdx)2dt+∫T02(sup‖v‖V≤1[‖u0(t)‖V‖v‖V+‖u0‖V‖v‖V+‖(δ(u0)x)x‖V∗‖v‖V])2dt≤∫T02(sup‖v‖V≤1∫T0∫Ω[|f(s)v|+|u0(s)u0x(s)v|+|μ(s)η0x(s)v|]dxds)2dt+∫T06([‖u0(t)‖2V+‖u0‖2V+‖(δ(u0)x)x‖2V∗])2dt≤∫T02(sup‖v‖V≤1∫T0(‖f(t)‖H‖v‖V+‖u0(t)‖C(¯Ω)‖u0(t)‖V‖v‖V+‖μ(t)‖H‖η0(t)‖V‖v‖C(¯Ω))ds)2dt+6Tδ0‖u0‖2L∞(0,T;Vδ)+6T‖u0‖2V+6T‖(δ(u0)x)x‖2V∗≤6T[T‖f‖2L2(0,T;H)+(c(E))2max{1,δ−10}T‖u0‖4L∞(0,T;Vδ)+(c(E))2‖μ‖2L2(0,T;H)‖η0‖2L2(0,T;V)]+6Tδ0‖u0‖2L∞(0,T;Vδ)+6T‖u0‖2V+6T‖(δ(u0)x)x‖2V∗<+∞.$ |
It is worth to mention here that, in fact,
$ \int_\Omega (\delta (u_0)_x)^2\,dx\le \|\delta\|_{C(\overline{\Omega})}\int_\Omega \delta((u_0)_x)^2\,dx\le c(E)\|\delta\|_V\|u_0\|_{V_\delta}. $ |
It remains to note that the property
Let us consider two operators
$ A, B:L^2(0,T;V_0)\times L^2(0,T;V_\delta)\to \left[L^2(0,T;V_0^\ast)\right]^2\times \left[L^2(0,T)\right]^2, $ |
defined on the set of vector functions
$ (A{\bf{p}})(t): = A(t){\bf{p}}(t) = \left( p(t)q(t)−(δqx(t))xγ1[δqx(t)]−γ2[δqx(t)] \right), $ | (36) |
$ (B{\bf{p}})(t): = B(t){\bf{p}}(t) = \left( u0px(t)+νpxx(t)+(μq)x(t)(η0+12r0)px(t)+12(r0)xp(t)+u0qx(t)−(σ1(t)+γ1[u0])γ1[q(t)](σ0(t)+γ2[u0])γ2[q(t)] \right). $ | (37) |
Here, we use the fact that
Lemma 4.1. The operator
for some constant
$\|A(t){\bf{v}}\|_{\widetilde{V}^\ast}\le C\|{\bf{v}}\|_{\widetilde{V}}+g(t),\ \ \ for \ a.e.\ \ \ t\in[0,T],\ \forall\,{\bf{v}}\in \widetilde{V};$ |
it is strictly monotone uniformly with respect to
$ ⟨A(t)v1−A(t)v2,v1−v2⟩˜V∗;˜V≥‖v11−v12‖2H+m‖v21−v22‖2Vδ,∀v1,v2∈˜V and for a.e. t∈[0,T]. $ |
Moreover, the operator
$ \|B{\bf{v}}_1-B{\bf{v}}_2\|_{L^2(0,T;\widetilde{V}^\ast)}\le L\|{\bf{v}}_1-{\bf{v}}_2\|_{L^2(0,T;\widetilde{V})}, \ for \ all\ {\bf{v}}_1,\,{\bf{v}}_2\in L^2(0,T;\widetilde{V}). $ |
Proof. Since the radial continuity of operator
$‖A(t)v‖˜V∗=sup‖z‖˜V≤1|⟨A(t)v,z⟩˜V∗;˜V|=sup‖z‖V0+‖y‖Vδ≤1|∫Ω(vz+wy)dx−∫Ω(δwx)xydx+δ(L)wx(L)y(L)−δ(0)wx(0)y(0)|=sup‖z‖˜V≤1|∫Ω(vz+wy)dx+∫Ωδwxyxdx|≤sup‖z‖˜V≤1(‖v‖H‖z‖H+‖w‖H‖y‖H+‖w‖Vδ‖y‖Vδ)≤2(‖v‖V0+‖y‖Vδ)=2‖v‖˜V.$ |
As for the monotonicity property, for every
$⟨A(t)p1−A(t)p2,p1−p2⟩˜V∗;˜V=∫Ω(p1−p2)2dx+∫Ω(q1−q2)2dx−∫Ω[(δ(q1)x)x−(δ(q2)x)x](q1−q2)dx+[δ(L)(q1(⋅,L))x−δ(L)(q2(⋅,L))x](q1(⋅,L)−q2(⋅,L))−[δ(0)(q1(⋅,0))x−δ(0)(q2(⋅,0))x](q1(⋅,0)−q2(⋅,0))=‖p1−p2‖H+‖q1−q2‖H+‖q1−q2‖2L2(Ω,δdx).$ |
It remains to show the Lipschitz continuity of operator
$‖Bv−Bw‖L2(0,T;˜V∗)=sup‖z‖˜V≤1|⟨Bv−Bw,z⟩˜V∗;˜V|=∫T0[|(u0(t)(v1x(t)−w1x(t)),z1(t))H|+ν|(v1x(t)−w1x(t),z1x(t))H|+|(μx(v2(t)−w2(t)),z1(t))H|+|(μ(v2x(t)−w2x(t)),z1(t))H|+12|((r0+2η0)(v1x(t)−w1x(t)),z2(t))H|+12|((r0)x(v1(t)−w1(t)),z2(t))H|+|(u0(t)(v2x(t)−w2x(t)),z2(t))H|+|(σ1(t)+u0(t,L))(v2(t,L)−w2(t,L))z2(t,L)|+|(σ0(t)+u0(t,0))(v2(t,0)−w2(t,0))z2(t,0)|]dt≤‖u0‖C(Q)‖v1−w1‖L2(0,T;V0)‖z1‖L2(0,T;V0)+ν‖v1−w1‖L2(0,T;V0)‖z1‖L2(0,T;V0)+∫T0(2‖z‖C(¯Ω)δ−1/20‖μ‖V‖v2−w2‖Vδ+12(‖r0+2η0‖H+‖r0‖V)‖v1−w1‖V‖z2‖C(¯Ω))dt+‖u0‖C(Q)δ−10‖v2−w2‖Vδ‖z2‖Vδ+∫T0(|σ1(t)|+|σ0(t)|+2‖u0(t)‖C(¯Ω))‖v2(t)−w2(t)‖C(¯Ω)dt.$ |
Taking into account the continuous embedding
$ \|v\|_{C(\overline{\Omega})}\le c(E)\|v\|_V\le c(E)\delta_0^{-1/2}\|v\|_{V_\delta}, $ |
we finally have
$ \|B{\bf{v}}-B{\bf{w}}{{\|}_{{{L}^{2}}(0,T;{{\widetilde{V}}^{*}})}}\le L\|{\bf{v}}-{\bf{w}}{{\|}_{{{L}^{2}}(0,T;\widetilde{V})}}, $ |
where
$C1=‖u0‖C(Q)+ν+c(E)(‖r0‖V+‖η0‖C(0,T;H)),C2=2c(E)δ−10‖μ‖L∞(0,T;V)+‖u0‖C(Q)δ−10+c(E)(‖σ1‖L2(0,T)+‖σ2‖L2(0,T)+2‖u0‖C(Q)).$ |
This concludes the proof.
Lemma 4.2. Operator
$A:L^2(0,T;V_0)\times L^2(0,T;V_\delta)\to \left[L^2(0,T;V_0^\ast)\right]^2\times \left[L^2(0,T)\right]^2,$ |
which is defined by (36), is radially continuous, strictly monotone and there exists an inverse Lipschitz-continuous operator
$ A^{-1}:\left[L^2(0,T;V_0^\ast)\right]^2\times \left[L^2(0,T)\right]^2\to L^2(0,T;V_0)\times L^2(0,T;V_\delta) $ |
such that
$ (A−1f)(t)=A−1(t)f(t) for a.e. t∈[0,T]and for all f∈[L2(0,T;V∗0)]2×[L2(0,T)]2, $ |
where
$A(t):V_0\times V_\delta\to \left[V_0^\ast\right]^2\times\mathbb{R}\times\mathbb{R}.$ |
Proof. It is easy to see that the action of operator
$A(t)p(t)=(A1(t)p(t)A2(t)q(t)),A1:L2(0,T;V0)→L2(0,T;V∗0),A2:L2(0,T;Vδ)→L2(0,T;V∗0)×L2(0,T)×L2(0,T),$ |
where
$ A_1(t)p(t) = p(t) \text{ and }A_2(t) q(t) = \left(q(t)−(δqx(t))xγ1[δqx(t)]−γ2[δqx(t)]\right).$ |
It is easy to see, that
$\langle (A_2 q_1)(t)-(A_2q_2)(t),q_1(t)-q_2(t)\rangle_{V_\delta^\ast;V_\delta} = \|q_1-q_2\|_{V_\delta}.$ |
Moreover,
$A_2^{-1}:L^2(0,T;V_0^\ast)\times L^2(0,T)\times L^2(0,T)\to L^2(0,T;V_\delta)$ |
such that
$ (A_2^{-1}f)(t) = A_2^{-1}(t)f(t)\ \text{for a.e. }\ t\in [0,T]\ \text{ and }\ \forall\,f\in \left[L^2(0,T;V_0^\ast)\right]\times \left[L^2(0,T)\right]^2, $ |
where
Before proceeding further, we make use of the following result concerning the solvability of Cauchy problems for pseudoparabolic equations (for the proof we refer to [11,Theorem 2.4]).
Theorem 4.3. For operators
$A, B:L^2(0,T;V_0)\times L^2(0,T;V_\delta)\to \left[L^2(0,T;V_0^\ast)\right]^2\times \left[L^2(0,T)\right]^2$ |
defined in (36), (37), and for any
$ F\in \left[L^2(0,T;V_0^\ast)\right]^2\times \left[L^2(0,T)\right]^2\ \ \ and\ \ \ \ b\in V_0^\ast\times V_\delta^\ast, $ |
the Cauchy problem
$ (A(t)p)′t+B(t)p=F(t),A(T)p(T)=b $ |
admits a unique solution.
In this section we focus on the derivation of the first-order optimality conditions for optimization problem (1)-(5). The Lagrange functional
$L:(W0(0,T)∩L2(0,T;H2(Ω)∩V0))×W1,∞(0,T;Vδ)×L2(0,T)×L2(0,T)×R×(W0(0,T)∩L2(0,T;H2(Ω)∩V0))×W1,∞(0,T;Vδ)→R,$ |
associated to problem (1)-(5) (see also Remark 2) is defined by
$L(w,u,g,h,λ,p,q)=λJ(g,h,w,u)−∫T0[⟨A1(w,u),p⟩V∗0;V0+⟨A2(w,u),q⟩V∗δ;Vδ]dt=λJ(g,h,w,u)−∫T0[⟨˙w,p⟩V∗0;V0−ν⟨wxx,p⟩V∗0;V0+((wu)x,p)H+12((r0+2η∗)ux,p)H]dt−∫T0[⟨˙u−(δ˙ux)x,q⟩V∗δ;Vδ+12((u2)x,q)H+(μwx,q)H−(f,q)H]dt−∫T0[(δ(L)˙ux(t,L)+σ1(t)u(t,L)−h)q(t,L)−(δ(0)˙ux(t,0)+σ0(t)u(t,0)−g)q(t,0)]dt=λJ(g,h,w,u)−∫T0[⟨˙w,p⟩V∗0;V0−ν⟨wxx,p⟩V∗0;V0+((wu)x,p)H+12((r0+2η∗)ux,p)H]dt−∫T0[⟨˙u,q⟩V∗δ;Vδ+∫Ωδ˙uxqxdx+12((u2)x,q)H+(μwx,q)H−(f,q)H]dt−∫T0[σ1(t)u(t,L)q(t,L)−h(t)q(t,L)−σ0(t)u(t,0)q(t,0)+g(t)q(t,0)]dt.$ |
Let us shift the correspondent derivatives from
$L(w,u,g,h,λ,p,q)=λJ(g,h,w,u)+∫T0[⟨w,˙p⟩V∗0;V0+ν⟨w,pxx⟩V∗0;V0+(wu,px)H+12(u,((r0+2η∗)p)x)H]dt−∫Ωp(T)w(T)dx+∫Ωp(0)w(0)dx+∫T0[⟨u,˙q⟩V∗δ;Vδ+∫Ωδux˙qxdx+12(u2,qx)H+(w,(μq)x)H+(f,q)H]dt−⟨u(T,⋅),q(T,⋅)⟩V∗δ;Vδ−∫Ωδux(T)qx(T)dx+⟨u(0,⋅),q(0,⋅)⟩V∗δ;Vδ+∫Ωδux(0)qx(0)dx−∫T0[σ1(t)u(t,L)q(t,L)−h(t)q(t,L)−σ0(t)u(t,0)q(t,0)+g(t)q(t,0)]dt=λJ(g,h,w,u)+∫T0[⟨w,˙p⟩V∗0;V0+ν⟨w,pxx⟩V∗0;V0+(wu,px)H+12(u,((r0+2η∗)p)x)H]dt−∫Ωp(T)w(T)dx+∫Ωp(0)w(0)dx+∫T0[⟨u,˙q⟩V∗δ;Vδ+∫Ωδux˙qxdx+12(u2,qx)H+(w,(μq)x)H+(f,q)H]dt−⟨u(T,⋅),q(T,⋅)−(δqx(T,⋅))x⟩V∗δ;Vδ−δ(L)u(T,L)qx(T,L)+δ(0)u(T,0)qx(T,0)+⟨u(0,⋅)−(δux(0,⋅))x,q(0,⋅)⟩V∗δ;Vδ−∫T0[σ1(t)u(t,L)q(t,L)−h(t)q(t,L)−σ0(t)u(t,0)q(t,0)+g(t)q(t,0)]dt−12∫T0(u2(t,L)q(t,L)−u2(t,0)q(t,0))dt=λJ(g,h,w,u)+∫T0[⟨w,˙p⟩V∗0;V0+ν⟨w,pxx⟩V∗0;V0+(wu,px)H+12(u,((r0+2η∗)p)x)H]dt−∫Ωp(T)w(T)dx+∫Ωp(0)w(0)dx+∫T0[⟨u,˙q−(δ˙qx)x⟩V∗δ;Vδ+12(u2,qx)H+(w,(μq)x)H+(f,q)H]dt−⟨u(T,⋅),q(T,⋅)−(δqx(T,⋅))x⟩V∗δ;Vδ−∫T0[(σ1(t)q(t,L)−(δ(L)˙qx(t,L))u(t,L)−h(t)q(t,L)]dt−∫T0[σ0(t)(q(t,0)−(δ(0)˙qx(t,0))u(t,0)−g(t)q(t,0)]dt−12∫T0(u2(t,L)q(t,L)−u2(t,0)q(t,0))dt.$ |
For each fixed
$ (w,u,g,h)\in \Big(W_0(0,T)\cap L^2(0,T;H^2(\Omega)\cap V_0)\Big)\times W^{1,\infty}(0,T;V_\delta)\times L^2(0,T)\times L^2(0,T). $ |
Notice that, for a fixed
Further we make use of the following relation
Also, to simplify the deduction and in order to avoid the demanding of the increased smoothness on solutions of the initial Boussinesq system (2)-(5), we use (see [4] and [5]) elastic model for the hydrodynamic pressure
$ P(t,x) = P_{ext}+\frac{\beta }{r_{0}^{2}}\eta $ |
instead of the inertial one
$ P=Pext+βr20η+ρωh∂2η∂t2=Pext+βr20η−12ρωhr0uxt. $ | (38) |
Indeed, if we suppose the wall thickness
As a result, the cost functional
$ J(g,h,w,u)=12∫ΩαΩ(u(T)−uΩ)2dx+12∫T0∫Ω((w(t)u(t))x+ux(t)η∗)2dxdt+12∫T0|∫ΩαQ(w(t)+η∗−ηQ)dx|2dt+12∫T0(βg|g|2+βh|h|2)dt. $ | (39) |
In order to formulate the conjugate system for an optimal solution
$ z\in W_0(0,T)\cap L^2(0,T;H^2(\Omega)\cap V_0)\ \ \ \ \text{and}\ \ \ \ v\in W^{1,\infty}(0,T;V_\delta)\times L^2(0,T). $ |
With that in mind we emphasize the following point. Since the elements
$ w+z\in W_0(0,T)\cap L^2(0,T;H^2(\Omega)\cap V_0)\ \ \ \ \text{and}\ \ \ \ u+v\in W^{1,\infty}(0,T;V_\delta)\times L^2(0,T) $ |
are some admissible solutions to OCP (39), (2)-(5), it follows that the increments
$ \left\{ z(0,⋅)=0 in Ω,v(0,⋅)−(δ(⋅)vx(0,⋅))x=0 in Ω, \right. $ | (40) |
$ \left\{ z(⋅,0)=z(⋅,L)=0 in (0,T),δ(0)˙vx(⋅,0)+σ0v(⋅,0)=0, in (0,T),δ(L)˙vx(⋅,L)+σ1v(⋅,L)=0, in (0,T),δ(L)vx(0,L)=δ(0)vx(0,0)=0. \right. $ | (41) |
Taking into account the definition of the Fréchet derivative of nonlinear mappings, we get
$ J(g,h,w+z,u) = J(g,h,w,u)+{{J}_{w}}z+{{R}_{0}}(w,z), $ |
where
$ R0(w,z)=12∫T0∫Ω((zu)x)2dxdt+∫T0|∫ΩaQz(t)|2dt, $ | (42) |
and
$Jwz=J(g,h,w+z,u)−J(g,h,w,u)−R0(w,z)=12∫T0∫Ω(((w(t)+z(t))u(t))x+ux(t)η∗)2dxdt−12∫T0∫Ω((w(t)u(t))x+ux(t)η∗)2dxdt+12∫T0|∫ΩαQ(w(t)+z(t)+η∗−ηQ)dx|2dt−12∫T0|∫ΩαQ(w(t)+η∗−ηQ)dx|2dt=∫T0∫Ω((w(t)u(t))x+ux(t)η∗)((z(t)u(t))x)dxdt+∫T0(∫ΩαQ(w(t)+η∗−ηQ)dx)(∫ΩαQz(t)dx)dt=∫T0∫Ω(wxu+uxw+uxη∗)(uxz+zxu)dxdt+α2Q∫T0∫Ω(∫Ω(w(t)+η∗−ηQ)dx)z(t)dxdt=∫T0∫Ω[(wxuxu+(ux)2w+(ux)2η∗)−(wxu2+uxuw+uxuη∗)x]z(t)dxdt+α2Q∫T0∫Ω(∫Ω(w(t)+η∗−ηQ)dx)z(t)dxdt.$ |
It is obviously follows from (42) that
$ \frac{|R_0(w,x)|}{\|z\|_{L^2(0,T;H^2(\Omega)\cap V_0)}}\rightarrow 0\ \ \ \ \text{as}\ \ \ \ \|z\|_{L^2(0,T;H^2(\Omega)\cap V_0)}\to 0. $ |
Hence, after some transformations, we have
$ Jwz=∫T0∫Ω(−u[uxx(w+η∗)+2uxwx+wxxu]+α2Q∫Ω(w(t)+η∗−ηQ)dx)z(t)dxdt. $ | (43) |
Treating similarly to the other derivative, we obtain
$ J(g,h,w,u+v) = J(g,h,w,u)+J_u v+\widetilde{\mathcal{R}}_0(u,v), $ |
where the remainder
$ ˜R0(u,v)=12∫ΩaΩv2(T)dx+12∫T0∫Ω((wv)x+vxη∗)2dxdt,|˜R0(u,v)|/‖v‖W1,∞(0,T;Vδ)→0 as ‖v‖W1,∞(0,T;Vδ)→0, $ | (44) |
and
$ Juv=J(g,h,w,u+v)−J(g,h,w,u)−˜R0(u,v)=12∫ΩαΩ(u(T)+v(T)−uΩ)2dx−12∫ΩαΩ(u(T)−uΩ)2dx+12∫T0∫Ω((w(t)(u(t)+v(t)))x+(ux(t)+vx(t))η∗)2dxdt−12∫T0∫Ω((w(t)u(t))x+ux(t)η∗)2dxdt=∫ΩαΩ(u(T)−uΩ)v(T)dx−∫T0∫Ω(w+η∗)[uxx(w+η∗)+2uxwx+wxxu]dxdt+∫T0η∗((w0(t,L)u0(t,L))x+u0xη∗)v(t,L)dt−∫T0η∗((w0(t,0)u0(t,0))x+u0x(t,0)η∗)v(t,0)dt. $ | (45) |
We are now in a position to identify the Fréchet derivatives
$Lwz=λJwz+∫T0[⟨z,˙p⟩V∗0;V0+ν⟨z,pxx⟩V∗0;V0+(zu,px)H+(z,(μq)x)H]dt−⟨z(T),p(T)⟩V∗0;V0$ |
and
$Luv=λJuv+∫T0[(wv,px)H+12(v,((r0+2η∗)p)x)H]dt+∫T0[⟨v,˙q−δ(˙qx)x⟩V∗δ;Vδ+(uv,qx)H]dt−⟨v(T,⋅),q(T,⋅)−(δqx(T,⋅))x⟩V∗δ;Vδ−∫T0[(σ1(t)q(t,L)−δ(L)˙qx(t,L))v(t,L)−(σ0(t)q(t,0)−δ(0)˙qx(t,0))v(t,0)]dt−∫T0(u(t,L)v(t,L)q(t,L)−u(t,0)v(t,0)q(t,0))dt−δ(L)v(T,L)qx(T,L)+δ(0)v(T,0)qx(T,0).$ |
As for the Fréchet derivatives
$Lgk(t)=L(w,u,g+k,h,p,q)−L(w,u,g,h,p,q)−R(g,k)=∫T0βgg(t)k(t)dt−∫T0k(t)q(t,0)dt−R(g,k),Lhl(t)=L(w,u,g,h+l,p,q)−L(w,u,g,h,p,q)−R(h,l)=∫T0βhh(t)l(t)dt+∫T0l(t)q(t,L)dt−R2(h,l),$ |
where
$ R1(g,k)=12∫T0βgk2(t)dt, R2(h,l)=12∫T0βhl2(t)dt,|R1(g,k)|/‖k‖L2(0,T)→0 as ‖k‖L2(0,T)→0, and |R2(h,l)|/‖l‖L2(0,T)→0 as ‖l‖L2(0,T)→0. $ |
Taking into account the calculations given above, we arrive at the following representation of the first-order optimality conditions for OCP (2)-(5), (39).
Theorem 5.1. Let
$ (p,q)\in \Big[W_0(0,T)\cap L^2(0,T;H^2(\Omega)\cap V_0)\Big]\times W^{1,\infty}(0,T;V_\delta) $ |
such that the following system
$ ∫T0[⟨˙w0(t),φ⟩V∗0;V0+((w0(t)u0(t))x,φ)H+ν(w0x(t),φx)H+12(r0u0x(t)+2η∗u0x(t),φ)H]dt=0, $ | (46) |
$ ∫T0[⟨˙u0(t),ψ⟩V∗δ;Vδ+∫Ωδ˙u0x(t)ψxdx+(u0(t)u0x(t),ψ)H+(μ(t)w0x(t),ψ)H+σ1(t)u0(t,L)ψ(L)−σ0(t)u0(t,0)ψ(0)]dt=∫T0[(f(t),ψ)H+h0(t)ψ(L)−g0(t)ψ(0)]dt, $ | (47) |
$ ∫T0[⟨˙p(t),φ(t)⟩V∗0;V0+ν⟨pxx(t),φ(t)⟩V∗0;V0+(px(t)u0(t),φ(t))H+((μq(t))x,φ(t))H]dt−(p(T),φ(T))H=∫T0∫Ω(u0[u0xxη0+2u0xη0x+η0xxu0])φ(t)dxdt−∫T0∫Ω(α2Q∫Ω(η0(t)−ηQ(t))dx)φ(t)dxdt, $ | (48) |
$ ∫T0[⟨˙q(t)−(δ˙qx(t))x,ψ(t)⟩V∗δ;Vδ+(qx(t)u0(t),ψ(t))H]dt+∫T0[(px(t)η0(t),ψ(t)))H+12((r0p(t))x,ψ(t))H]dt−∫T0[(σ1(t)+u0(t,L))q(t,L)−δ(L)˙qx(t,L)]ψ(t,L)dt+∫T0[(σ0(t)+u0(t,0))q(t,0)−δ(0)˙qx(t,0)]ψ(t,0)dt−⟨v(T,⋅),q(T,⋅)−(δqx(T,⋅))x⟩V∗δ;Vδ−δ(L)qx(T,L)ψ(T,L)+δ(0)qx(T,0))ψ(T,0)=∫T0∫Ωη0[u0xx(t)η0(t))+2u0x(t)η0x(t)+η0xx(t)u0(t)]ψ(t)dxdt−∫ΩaΩ(u0(T)−uΩ)ψ(T)dx−∫T0η∗(η0x(t,L)u0(t,L)+η∗u0x(t,L))ψ(t,L)dt+∫T0η∗(η0x(t,0)u0(t,0)+η∗u0x(t,0))ψ(t,0)dt, $ | (49) |
$ ∫T0(βgg0(t)−q(t,0))(g(t)−g0(t))dt≥0, ∀g∈Gad, $ | (50) |
$ ∫T0(βhh0(t)+q(t,L))(h(t)−h0(t))dt≥0 ∀h∈Had, $ | (51) |
$ η0(t)=w0(t)+η∗, $ | (52) |
$ δ(L)u0x(0,L)=0, δ(0)u0x(0,0)=0, δ(L)qx(T,L)=0, δ(0)qx(T,0)=0, $ | (53) |
$ w0(0)=η00−η∗, p(T)=0, p(⋅,0)=p(⋅,L)=0, $ | (54) |
$ u0(0)−(δu0x(0))x=u0, q(T)−(δqx(T))x=λaΩ(u0(T)−uΩ) $ | (55) |
holds true for all
$\varphi\in W_0(0,T)\cap L^2(0,T;H^2(\Omega)\cap V_0),\ \psi\in W^{1,\infty}(0,T;V_\delta),\ \varphi\in V_0, \ \psi\in V_\delta,$ |
and a.e.
Proof. Since the derived optimality conditions (46)-(55) are the direct consequence of the Lagrange principle, we focus on the solvability of the variational problems (48)-(49) for the adjoint variables
$ pt+νpxx+pxu0+(μq)x=λu0[u0xxη0+2u0xη0x+η0xxu0]−λ(αQ)2∫Ω(η0−ηQ)dx, $ | (56) |
$ [q−(δqx)x]t+qxu0+pxη0+12(r0p)x=λη0[u0xxη0+2u0xη0x+η0xxu0], $ | (57) |
$ δ(L)˙qx(⋅,L)−(σ1+u0(⋅,L))q(⋅,L)=−λη∗(η0x(⋅,L)u0(⋅,L)+u0x(⋅,L)η∗), $ | (58) |
$ δ(0)˙qx(⋅,0)−(σ0+u0(⋅,0))q(⋅,0)=−λη∗(η0x(⋅,0)u0(⋅,0)+u0x(⋅,0)η∗), $ | (59) |
$ q(T)−(δqx(T))x=λaΩ(u0(T)−uΩ), $ | (60) |
$ δ(L)qx(T,L)=δ(0)qx(T,0)=0, $ | (61) |
$ p(T)=0, p(⋅,0)=p(⋅,L)=0. $ | (62) |
In the operator presentation, the system (56)-(62) takes the form (see [11]):
$ (A(t)p)′t+B(t)p=F(t), A(T)p(T)=b, $ |
where the operators
$A(t), B(t): L^2(0,T;V_0)\times L^2(0,T;V_\delta)\to \left[L^2(0,T;V_0^\ast)\right]^2\times \left[L^2(0,T)\right]^2$ |
are defined in (36)-(37), and
$b=(0,λaΩ(u0(T)−uΩ),0,0)∈V∗0×V∗0×R×R,F(t)=(f1,f2,ϕ1,ϕ2)t∈[L2(0,T;V∗0)]2×[L2(0,T)]2,f1(t)=λu0[u0xxη0+2u0xη0x+η0xxu0]−λ(αQ)2∫Ω(η0−ηQ)dx,f2(t)=λη0[u0xxη0+2u0xη0x+η0xxu0],ϕ1(t)=−λη∗(η0x(t,L)u0(t,L)+u0x(t,L)η∗),ϕ2(t)=λη∗(η0x(t,0)u0(t,0)+u0x(t,0)η∗).$ |
As a result, the existence of a unique pair
$ F\in \left[L^2(0,T;V_0^\ast)\right]^2\times \left[L^2(0,T)\right]^2\ \text{ and }\ {\bf{b}}\in V_0^\ast\times V_0^\ast\times\mathbb{R}\times\mathbb{R}, $ |
the Lagrange multiplier
$\mathcal{L} = \mathcal{L}(w,u,g,h,\lambda,p,q)$ |
can be taken equal to
[1] | R. A. Adams, Sobolev Spaces, Academic Press, New York, 1975. |
[2] | J. Alastruey, Propagation in the Cardiovascular System: Development, Validation and Clinical Applications, Ph.D thesis, Imperial College London, 2006. |
[3] |
Boundary control for an arterial system. J. of Fluid Flow, Heat and Mass Transfer (2016) 3: 25-33. ![]() |
[4] |
Flow optimization of the vascular networks. Mathematical Biosciences and Engineering (2017) 14: 607-624. ![]() |
[5] |
On boundary optimal control problem for an arterial system: Existence of feasible solutions. Journal of Evolution Equations (2018) 1-42. ![]() |
[6] |
On relaxation of state constrained optimal control problem for a PDE-ODE model of supply chains. Networks and Heterogeneous Media (2014) 9: 501-518. ![]() |
[7] |
On optimization of a highly re-entrant production system. Networks and Heterogeneous Media (2016) 11: 415-445. ![]() |
[8] |
R. Dautray and J.-L. Lions,
Mathematical Analysis and Numerical Mathods for Science and Technology, Vol. 5: Evolutional Problems I, Springer-Verlag, Berlin, 1992. doi: 10.1007/978-3-642-58090-1
![]() |
[9] |
Numerical modeling of 1D arterial networks coupled with a lumped parameters, description of the heart. Comput. Methods Biomech. Biomed. Eng. (2006) 9: 273-288. ![]() |
[10] | L. Formaggia, A. Quarteroni and A. Veneziani, Cardiovascular Mathematics: Modeling and Simulation of the Circulatory System, Springer Verlag, Berlin, 2010. |
[11] | H. Gajewski, K. Gröger and K. Zacharias, Nichtlineare Operatorgleichungen und Operatordifferentialgleichungen, Akademie-Verlag, Berlin, 1974. |
[12] |
F. C. Hoppensteadt and C. Peskin,
Modeling and Simulation in Medicine and the Life Sciences, Springer-Verlag, New York, 2002. doi: 10.1007/978-0-387-21571-6
![]() |
[13] | M. O. Korpusov and A. G. Sveshnikov, Nonlinear Functional Analysis and Mathematical Modelling in Physics: Methods of Nonlinear Operators, KRASAND, Moskov, 2011 (in Russian). |
[14] | A. Kufner, Weighted Sobolev Spaces, Wiley & Sons, New York, 1985. |
[15] |
F. Liang, D. Guan and J. Alastruey, Determinant factors for arterial hemodynamics in hypertension: Theoretical insights from a computational model-based study,
ASME Journal of Biomechanical Engineering, 140 (2018), 031006. doi: 10.1115/1.4038430
![]() |
[16] |
Asymptotic nonlinear and dispersive pulsatile flow in elastic vessels with cylindrical symmetry. Computers & Mathematics with Applications (2018) 75: 4022-4027. ![]() |
[17] |
Blood pressure and blood flow variation during postural change from sitting to standing: model development and validation. J. Appl. Physiol (2005) 99: 1523-1537. ![]() |
[18] |
Numerical modeling of the pressure wave propagation in the arterial flow. International Journal for Numerical Methods in Fluids (2003) 43: 651-671. ![]() |
[19] |
The cardiovascular system: Mathematical modelling, numerical algorithms and clinical applications. Acta Numerica (2017) 16: 365-590. ![]() |
[20] |
Validation of a one-dimensional model of the systemic arterial tree. Am J Physiol Heart Circ Physiol. (2009) 297: H208-H222. ![]() |
[21] |
Computational modeling of 1D blood flow with variable mechanical properties and its application to the simulation of wave propagation in the human arterial system. Internat. J. for Numerical Methods in Fluids (2003) 43: 673-700. ![]() |
1. | Maria Pia D'Arienzo, Luigi Rarità, 2020, 2293, 0094-243X, 420041, 10.1063/5.0026464 | |
2. | Maria Pia D'Arienzo, Luigi Rarità, 2020, 2293, 0094-243X, 420042, 10.1063/5.0026462 | |
3. | Евгений Сергеевич Барановский, Evgenii Sergeevich Baranovskii, Задача оптимального стартового управления для двумерных уравнений Буссинеска, 2022, 86, 1607-0046, 3, 10.4213/im9099 |