941
Views
5
CrossRef citations to date
0
Altmetric
Research Article

A Port-Hamiltonian formulation of linear thermoelasticity and its mixed finite element discretization

ORCID Icon, ORCID Icon, &
Pages 643-661 | Received 08 Feb 2021, Accepted 02 Apr 2021, Published online: 05 May 2021

Abstract

A port-Hamiltonian formulation for general linear coupled thermoelasticity and for the thermoelastic bending of thin structures is presented. The construction exploits the intrinsic modularity of port-Hamiltonian systems to obtain a formulation of linear thermoelasticity as an interconnection of the elastodynamics and heat equations. The derived model can be readily discretized by using mixed finite elements. The discretization is structure-preserving, since the main features of the system are retained at a discrete level. The proposed model and discretization strategy are validated against a benchmark problem of thermoelasticity, the Danilovskaya problem.

1. Introduction

Thermoelasticity is the study of elastic bodies undergoing thermal excitation. It is a clear example of a multiphysics phenomenon since the heat transfer and elastic vibrations within the body mutually interact. Over the last twenty years, distributed port-Hamiltonian (pH) systems have attracted the attention of different research communities [Citation1]. An important peculiarity of port-Hamiltonian systems (pHs) is that they are naturally modular [Citation2]. This feature is particularly appealing in the case of multiphysics phenomena like thermoelasticity, since each physical domain can be modeled independently from the others and subsequently interconnected to the rest in a physically motivated way.

Flexible structures have been largely investigated into the pH framework as well as the heat equation (consult for instance [Citation3] for the Timoshenko beam, [Citation4] for the Euler-Bernoulli beam, [Citation5, Citation6] for thick and thin plates and [Citation7, Citation8] for the heat equation). More complicated models arising from fluid dynamics have also been considered [Citation9–12]. The development of new models within the pH framework has been accompanied with an increased interest in numerical discretization methods, capable of retaining the main features of the distributed system in its finite-dimensional counterpart. Recently, it has become evident that there is a strict link between discretization of port-Hamiltonian systems and mixed finite elements [Citation13]. An example of this connection is given in [Citation14], where a velocity-stress formulation for the wave dynamics is shown to be Hamiltonian and its mixed discretization preserves such a structure.

Two main contributions are presented in this article. First, a linear model of thermoelasticity is obtained in the pH formalism. Each physics is described separately and the final system is obtained considering a power-preserving interconnection of the heat equation and linear elastodynamics formulated as port-Hamiltonian systems. The construction applies to both general linear thermoelasticity and bending of thin structures. For the latter case, the elastic vibrations take place in a reduced domain (uni-dimensional for beams and bi-dimensional for plates), whereas the thermal diffusion occurs in the three-dimensional domain. This generalizes models in which the heat diffusion is reduced to the same domain of the elastic vibrations (cf. [Citation15] for plates and [Citation16] for beams). The second contribution is a mixed finite elements discretization method which is structure-preserving. Two different mixed formulations are presented. One allows incorporating Neumann boundary conditions directly into the weak form as natural conditions. The other incorporates Dirichlet conditions as natural boundary conditions. The proposed discretization is then applied to the Danilovskaya problem, assessing the validity of both the model and the associated discretization.

The paper is organized as follows. In Section 2 linear thermoelasticity is constructed as the interconnection of the heat equation and linear elastodynamics. First, the heat equation is formulated as a pH system. Then, the same procedure is applied to the elastodynamics. This methodology is then applied to the thermoelastic bending of thin structures, i.e. beams and plates in Section 3. The discretization strategy is discussed in §4. By careful application of the integration by parts, two discretizations, sharing the same structure and properties of the infinite-dimensional system, are obtained. In Section 5, the proposed model and discretization are tested using the Danilovskaya problem. This problem has been frequently used, since an analytical solution in the Laplace domain is known.

2. Port-Hamiltonian linear coupled thermoelasticity

In this section, the classical model of thermoelasticity is reformulated in a pH fashion by interconnecting the heat equation and the linear elastodynamics problem both seen as pHs. The construction makes use of the intrinsic modularity of pHs [Citation2]. It is shown that the interconnection preserves a quadratic functional that plays the role of a fictitious energy. The resulting system is dissipative with respect to this functional.

2.1. Classical thermoelasticity

Consider a bounded connected set ΩRd, d{1,2,3}. The classical equations for linear fully-coupled thermoelasticity for an isotropic thermoelastic material are [Citation17, Citation18] (1) ρ2ut2=Div(ΣET),ρcϵTt=div(jQ)Cβ:εt,ΣET=ΣE+ΣT,ΣE=D(ε),ΣT=Cβθ,θ:=(TT0)/T0,ε=Grad (u),jQ=k grad T.(1) where ρ, cϵ, k, T0 are the mass density, the specific heat density at constant strain, the thermal conductivity and the reference temperature. The vector field uRd is the displacement, the scalar field T is the temperature, εRsymd×d is the infinitesimal strain tensor, ΣERsymd×d is the symmetric stress tensor contribution due to mechanical deformation, ΣTRsymd×d the symmetric stress tensor contribution due to a thermal field, and jQRd is the heat flux. Tensor D is the stiffness tensor. For an isotropic homogeneous material, it takes the form (2) D(·)=2μ(·)+λ Tr (·)Id×d,(2) where coefficients λ, μ are the Lamé parameters. Coefficient k is the thermal conductivity. The coupling term is expressed as: (3) Cβ:=T0β(2μ+3λ)Id×d,(3) where β is the thermal expansion coefficient. The operator Div is the divergence of a tensor field Div Σ(x,t)=(i=1dΣijxi)1jd.

The linearized strain tensor (also called infinitesimal strain tensor) is the symmetric gradient of the displacement (4) ε:=Grad u,   where   Grad u=12[u+(u)].(4)

Operator grad is the gradient of a scalar field, while div is the divergence of a vector field. The notation A:B=Tr (AB)=i,jAijBij denotes the tensor contraction. The reader may consult [Citation19, Chapter 1] or [Citation20, Chapter 8] for a detailed derivation on these equations.

Given a partition of the boundary Ω=ΓDEΓNE=ΓDTΓNT for the elastic and thermal domain, the general boundary conditions read (see ) (5) Dirichletb.c.:Neumannb.c.:u known on ΓDE×(0,+),ΣET·n known on ΓNE×(0,+),T known on ΓDT×(0,+),jQ·n known on ΓNT×(0,+),(5) where n is the outgoing normal vector at the boundary. Note that there are 4 different cases of boundary conditions all together, since at each point of the boundary both a vectorial b.c. on the elastic part and a scalar b.c. on the thermal part must be taken into account. In the following sections, the pHs formulation of the heat equation and elastodynamics are given. Then, an equivalent coupled system is constructed by interconnecting these two systems in a structured manner.

Figure 1. Boundary conditions for the thermoelastic problem, with 4 cases.

Figure 1. Boundary conditions for the thermoelastic problem, with 4 cases.

2.2. The heat equation as a port-Hamiltonian descriptor system

Consider the heat equation in a bounded connected set ΩRd, d{1,2,3}, describing the evolution of the temperature field T(x,t) (6) ρcϵTt=div (k grad T)+uT,xΩ,(6) where ρ, cϵ, k have the same meaning as in Equation(1) and uT is a distributed heat source. This model can be put in pH form by means of a canonical interconnection structure. To model the Fourier law, an algebraic relationship has to be incorporated to obtain a pH system (cf. [Citation8, Chapter 2]). Here, in the same manner, a differential-algebraic formulation is exploited.

Let T0 be a constant reference temperature (the introduction of this variable is instrumental for coupled thermoelasticity). The functional HT=12ΩρcϵT0(TT0T0)2dΩ has the physical dimension of an energy and represents a Lyapunov functional of this system. Even though it does not represent the internal energy, which is classically used in thermodynamics, it has some important and useful properties. Select as energy variable αT:=ρcϵ(TT0).

The corresponding co-energy is eT:=δHTδαT=αTρcϵT0=TT0T0=:θ.

Introducing the heat flux jQ:=k grad T as additional variable, the heat equation (Equation6) is equivalently reformulated as (7) [1000]t(αTjQ)=[0divgrad(T0k)1](eTjQ)+[10]uT,yT=[10](eTjQ),(7) where yT represents the distributed output, which is power-conjugated to the distributed heat source input uT. In matrix notation, it is obtained (8) ETtαT=(JTRT)eT+BT uT,yT=BT* eT(8) where αT=(αT,jQ),eT=(eT,jQ) and ET=[1000],JT=[0divgrad0],RT=[000(T0k)1],BT=[10].

This system is an example of pH descriptor system (cf. [Citation21, Citation22] for the finite-dimensional case). The Hamiltonian reads (9) HT=12ΩαT2ρϵεT0dΩ.(9)

The power rate is then deduced (10) ḢT=ΩδHTδαTαTtdΩ=ΩeTtαTdΩ,=ΩeT·ET tαTdΩ,=ΩeT·{(JTRT)eT+BTuT}dΩ,=ΩuT yTdΩΩ(eT div jQ+jQ· grad eT+||jQ||2kT0)dΩ,=ΩuT yTdΩΩ||jQ||2kT0dΩΩeTjQ·ndS,ΩuT yTdΩΩeTjQ·ndS.(10)

This choice of Hamiltonian allows retrieving the classical boundary conditions (i.e. temperature, or inward heat flux) and leads to a dissipative system. Other formulations, based on the entropy or the internal energy as Hamiltonian functionals, are possible for the heat equation [Citation23, Citation24]. These provide either an accrescent or a lossless system. Unfortunately these formulations are non linear and their discretization is a difficult task [Citation25].

2.3. Port-Hamiltonian linear elastodynamics

Consider the linearized equation of elastodynamics [Citation20, Chapter 4] (11) ρ2ut2(x,t) Div (D Grad u)=f,xΩ,(11) where u and D have the same meaning as in (1). The term f represents an external force. To derive a pH formulation, the total energy, that includes the kinetic and deformation energy, is used (12) HE=12Ω{ρ||tu||2+Σ:ε}dΩ.(12)

Recall that ε=Grad u and Σ=Dε.

The energy variables are then the linear momentum and the deformation field αv=ρv,Aε=ε, where v:=tu. The Hamiltonian can be rewritten as a quadratic functional in the energy variables (13) HE=12Ω{1ρ||αv||2+(DAε):Aε}dΩ.(13)

The co-energy variables are given by (14) ev:=δHδαv=v,Eε:=δHδAε=Σ.(14)

The tensor-valued co-energy variable Eε is obtained by taking the variational derivative with respect to a tensor (cf. [Citation26, Chapter 3] and [Citation5]).

The equivalent port-Hamiltonian reformulation of system Equation(11) is then given by (cf. [Citation26, Chapter 3]) (15) t(αvAε)=[0DivGrad0](evEε)+[Id×d0]uE,yE=[Id×d0](evEε),(15) where the distributed input uE:=f plays the role of the previously introduced forcing. The energy rate verifies the following (16) ḢE=Ω{ev·tαv+Eε:tAε}dΩ,=ΩuE·yEdΩ+Ωev·(Eε·n)dS.(16)

2.4. Thermoelasticity as two coupled Port-Hamiltonian systems

Given the pHDAE formulation of the heat equation (Equation7) and the pH formulation of elasticity Equation(15), linear thermoelasticity can be expressed as a coupled port-Hamiltonian system by considering the following interconnection (17) uE=Div (Cβ yT),uT=Cβ:Grad (yE).(17)

The interconnection is power preserving, since it can be compactly written as (18) uE=Aβ(yT),uT=Aβ*(yE),   where   Aβ(·)=Div (Cβ ·),(18)

where Aβ* denotes the formal adjoint (cf. ). The assertion is justified by the following proposition.

Figure 2. Schematic interconnection block diagram for linear thermoelasticity.

Figure 2. Schematic interconnection block diagram for linear thermoelasticity.

Proposition 1.

Let C0(Ω),C0(Ω,Rd) be the space of smooth scalar functions and vector-valued functions with compact support in Ω. Given yTC0(Ω),yEC0(Ω,Rd), the coupling operator (19) Aβ:C0(Ω)C0(Ω,Rd),yTDiv (CβyT)(19) has formal adjoint (20) Aβ*:C0(Ω,Rd)C0(Ω)yE+Cβ: Grad (yE)(20)

Proof.

It is necessary to show (21) yE, AβyTL2(Ω,Rd)=Aβ*yE, yTL2(Ω),(21) where for uE,yEC0(Ω),uT,yTC0(Ω) (22) uE, yEL2(Ω,Rd)=ΩEuE·yEdΩ,uT, yTL2(Ω)=ΩTuTyTdΩ.(22)

By applying the integration by parts, the proof is readily obtained (23) yE, AβyTL2(Ω,Rd)=ΩyE·Div (CβyT)dΩ,=Ω Grad (yE):CβyTdΩ,=ΩAβ*(yE)yTdΩ,=Aβ*yE, yTL2(Ω).(23)

If the compact support assumption is removed, it is obtained (24) uT,yTL2(Ω)+uE,yEL2(Ω,R3)=Ω{(Cβ: Grad ev)eT+Div (CβeT)·ev}dΩ,=Ω div (eT Cβ·ev)dΩ,=Ω(eT Cβ·n)·evdS.(24)

Using the expression of yT,yE, considering that T0 is constant and applying Schwarz theorem for smooth function, the inputs are equal to uE=Div (ΣT),uT=Cβ: Grad (v)=Cβ:εt.

The coupled thermoelastic problem can now be written as (25) [1000010000100000]t(αvAεαTjQ)=[0Div Aβ0Grad000Aβ*00div00 grad (T0k)1](evEεeTjQ),(25) with total energy given by H=HE+HT. The power balance for each subsystem is given by (26) ḢE=ΩuE·yEdΩ+Ωev·(Eε·n)dS,(26) (27) ḢTΩuTyTdΩΩθjQ·ndS,(27)

The overall power balance is easily computed considering EquationEqs. (26), Equation(27) and Equation(24). (28) Ḣ=ḢE+ḢTΩ{[EεeTCβ]·n}·evdSΩeTjQ·ndS.(28)

This result is the same as the one stated in [Citation18, p. 332]. From the power balance the classical boundary conditions are retrieved. This allows defining appropriate boundary operators for the thermoelastic problem (29) u=[γ0ΓDE0000γΓNEγΓNE(Cβ ·)000γ0ΓDT0000γΓNT]B(evEεeTjQ),y=[0γΓDEγΓDE(Cβ ·)0γ0ΓNE000000γΓDT00γ0ΓNT0]C(evEεeTjQ).(29)

The notation γ0Γ*° (with *={D,N} and °={E,T}) indicates the Dirichlet trace over the set Γ*°, namely γ0Γ*°a=a|Γ*°, and γΓ*° indicates the normal trace along Γ*°, namely γΓ*°a=a·n|Γ*°.

System (25) together with (29) is a pH system with collocated boundary control and observation. Indeed, it shows that the classical thermoelastic problem can be modeled as two coupled subsystems, demonstrating the modularity of the pH paradigm.

Remark 1.

Notice that the boundary operators in EquationEq. (29) contain a coupling between the thermal and mechanical variables. This is due to the fact that the coupling operator Aβ is of differential nature; otherwise, the coupling would only appear in the domain, but not on the boundary.

3. Thermoelastic Port-Hamiltonian bending

In this section, the thermoelastic bending of thin beam and plate structures is described as a pH system. Starting from classical coupled thermoelastic models, suitable pH formulations are obtained. These couple a mechanical system defined on a reduced domain (uni-dimensional for beams, bi-dimensional for plates) to the thermal diffusion defined in the three-dimensional space.

Here, instead of detailing each physics (thermal and elastic) and the interconnection between the two, the pH system is derived from the coupled equations. It is shown that the final pH system possesses the same structure as EquationEq. (25).

3.1. Thermoelastic Euler-Bernoulli beam

The model for the linear thermoelastic vibrations of an isotropic thin rod is detailed in [Citation27, Citation28]. The domain of the beam is uni-dimensional ΩE={0,L}, while the thermal domain is three-dimensional ΩT={0,L}×S, where S is the set representing the beam cross section. For simplicity, the set S is assumed to be constant along the x-axis. The ruling equations are (30) ρA2wt2=EI4wx4βET02x2Szθdydz,x{0,L}=ΩE,ρcϵ,BT0θt=div (kT0 grad θ)+βT0Ez3wx2t,(x,y,z)ΩE×S=ΩT,(30) where w(x, t) is the vertical displacement of the beam, I=Sz2dydz the second moment of area, E the Young modulus and A the area of the surface S. The constant cϵ,B is due to the thermoelastic coupling (cf. [Citation27, Citation28] for a detailed explanation). The other terms have the same meaning as in §2. The normalized temperature θ(x,y,z,t) depends on all spatial coordinates. For simplicity, the physical parameters are assumed to be constant.

The coupling operator is defined as (31) Aβ,B(yT):=βET0xx(SzyTdydz).(31)

To unveil an interconnection that is power preserving with respect to a certain function, the formal adjoint of the coupling operator is needed.

Proposition 2.

Let C0(ΩT),C0(ΩE) be the space of smooth functions with compact support defined in ΩT and ΩE, respectively. Given yTC0(ΩT),yEC0(ΩE) the formal adjoint of the coupling operator is (32) Aβ,B*(yE)=βET0z xxyE.(32)

Proof.

The formal adjoint is defined by the relation (33) yE, Aβ,ByTL2(ΩE)=Aβ,B*yE, yTL2(ΩT),(33) where for uE,yEC0(ΩE),uT,yTC0(ΩT) (34) uE, yEL2(ΩE)=ΩEuEyEdx,uT, yTL2(ΩT)=ΩTuTyTdxdydz.(34)

Using Def. Equation(31) and integration by parts, one finds (35) yE, Aβ,ByTL2(ΩE)=ΩEyEAβ,ByTdx,=ΩEyEβET0xx(SzyTdydz)dx,=ΩE(xxyE)βET0(SzyTdydz)dx,(35)

Since ΩT=ΩE×S and thanks to Fubini theorem, it is found (36) ΩExx(yE)βET0(Sz yTdydz)dx=ΩES(xxyE)βET0z yTdxdydz,=ΩT(xxyE)βET0z yTdxdydz,=Aβ,B*yE, yTL2(ΩT).(36)

This concludes the proof. □

Using EquationEqs. (31) and Equation(32), system Equation(30) is rewritten as (37) ρA2wt2=EI4wx4+Aβ,Bθ,ρcϵ,BT0θt=div (kT0 grad θ)Aβ,B*wt.(37)

Consider the Hamiltonian functional (38) H=HE+HT=12ΩE{ρA(wt)2+EI(2wx2)2}dx+12ΩTρcϵ,BT0θ2dxdydz.(38)

The energy variables are chosen as follows (39) αw=ρAtw,ακ=xx2w,αT=ρcϵ,B T0 θ.(39)

The corresponding co-energy variables read (40) ew:=δHδαw=tw,eκ:=δHδακ=EIxx2w,eT:=δHδαT=θ.(40)

System Equation(37) can now be rewritten as (41) [1000010000100000]t(αwακαTjQ)=[0xx2Aβ,B0xx2000Aβ,B*00div00 grad (kT0)1](eweκeTjQ),(41)

This system is the equivalent of Equation(25) for the bending of beams. Hence, following the same reasoning, it can be obtained starting from each subsystem in pH form by means of an appropriate interconnection.

3.2. Thermoelastic Kirchhoff plate

For the bending of thin plate, several models have been proposed [Citation27, Citation29–31]. Here, the Chadwick model [Citation27] is considered. The thin plate occupies the open connected set ΩE×{h2,h2}, where h is the plate thickness. The system of equations describes the midplane vertical displacement and the evolution of the temperature in the 3 D domain (42) ρh2wt2=div Div2D(DbHess2Dw)βT0E1νΔ2D(h/2h/2zθdz),    (x,y)ΩE,ρcϵ,PT0θt=div3D(kT0 grad 3Dθ)+βT0Ez1νΔ2D(wt),(x,y,z)ΩE×{h2,h2}=ΩT,(42) where w(x,y,t) is the vertical deflection, E is the Young modulus, ν the Poisson modulus and cϵ,P a constant (depending on the heat capacity at constant strain and other coupling parameters, cf. [Citation27]). Symbols Δ2D=xx2+yy2 stands for the two-dimensional Laplacian. The notation Hess denotes the Hessian operator. This operator can be decomposed as Hess=Grad° grad  [Citation6]. The subscript 2D, 3D refers to the spatial dependency of the operators. Tensor Db is the bending stiffness tensor, defined by (43) Db(·):=Eh312(1ν2)[(1ν)(·)+ν Tr (·)I2×2].(43)

The coupling operator is here defined as (44) Aβ,P(yT):=βT0E1νΔ2D(h/2h/2zyTdz).(44)

Analogously to the case of the Euler-Bernoulli beam, its formal adjoint is sought for.

Proposition 3.

Let C0(ΩT),C0(ΩE) be the space of smooth functions with compact support defined in ΩT and ΩE respectively. Given yTC0(ΩT),yEC0(ΩE) the formal adjoint of the coupling operator is (45) Aβ,B*(yE)=βT0Ez1νΔ2DyE.(45)

Proof.

The proof is completely identical to that of Prop. 2. □

System Equation(42) is rewritten as (46) ρh2wt2=div Div2D(Db Hess2Dw)+Aβ,Pθ,ρcϵ,PT0θt= div3D(kT0 grad 3Dθ)Aβ,P*(wt),(46)

The Hamiltonian functional equals (47) H=HE+HT=12ΩE{ρh(wt)2+(DbHess2Dw): Hess2Dw}dxdy+12ΩTρcϵ,PT0θ2dxdydz.(47)

The energy and co-energy variables are (48) αw=ρhtw,ew=tw,Aκ=Hess2Dw,Eκ=Db Hess2Dw,αT=ρcϵ,PT0θ,eT=θ.(48)

System Equation(46) is rewritten as (49) [1000010000100000]t(αwAκαTjQ)=[0div Div2DAβ,P0Hess2D000Aβ,P*00div3D00 grad 3D(kT0)1](ewEκeTjQ),(49) where div Div2D and Hess2D are formally adjoint operators [Citation6]. The final system reproduces the same structured coupling already observed for (25) and (41) before.

Remark 2.

The choice can be made to reduce the thermoelastic bending to two problems defined on the same spatial domain (cf. [Citation16] for beams in 1D, and [Citation15] for plates in 2D) by introducing the following approximation of the temperature field (50) θ(x,y,z)=θ0+z θ1,(50) where θ0=θ0(x),θ1=θ1(x) for beams, and θ0=θ0(x,y),θ1=θ1(x,y) for plates. This introduces a strong simplification, since the thermal phenomena typically occur in the whole three-dimensional space, and not only in 1D or 2D as this approach implies.

Remark 3

(Lagnese [Citation29] and Nowacki [Citation32] models in pH form). The models by Lagnese and Nowacki consider the thermal evolution equation in the variable Θ(x,y,t)=1Ih/2h/2zθ(x,y,z,t)dz,whereI=h312, corresponding to the first moment of the temperature. In their formulation, a linear term appears in the evolution equation for the temperature. This term arises from the second derivative with respect to z in the Laplacian h/2h/2z2z2θ(x,y,z,t)dz=(zzθθ)|h/2h/2=θ|h/2h/2.

The term zθ|h/2h/2 is zero because of an assumed zero flux condition on the plate faces (cf. [Citation31]). For the second term, Lagnese and Nowacki assume that θ(x,y,z,t) is linear in z [Citation31]. This means that θ(x,y,z,t)zΘ(x,y,t).

Then, it holds θ(x,y,z,t)|h/2h/2=hΘ(x,y,t), so that h/2h/2z2z2θ(x,y,z,t)dz=hΘ(x,y,t).

This obviously introduces an inconsistency, as the term in the integral should be zero. However, it allows to retrieve the damping due the term zzθ(x,y,z,t) in the reduced model. After this clarification, it is possible to state the port-Hamiltonian realization of the Nowacki and Lagnese model [1000010000100000]t(αwAκαTjQ)=[0div Div2DAβ,PL0Hess2D000Aβ,PL*0T0khdiv2D00 grad 2D(kT0I)1](ewEκeTjQ), where the underlying variables are defined as αw=ρhtw,ew=tw,Aκ=Hess2Dw,Eκ=Db Hess2Dw,αT=ρcϵ,PT0IΘ,eT=Θ.

Here the coupling operators Aβ,PL,Aβ,PL* are given by Aβ,PL(·)=βT0EI1νΔ2D(·),Aβ,PL*(·)=βT0EI1νΔ2D(·).

4. Mixed finite element discretization

The numerical discretization is illustrated considering the linear thermoelasticity system (25). By using the expression of the coupling operator (3), and using a pure co-energy formulation, system (25) takes the form (51) [ρ0000D10000ρcϵT000000]t(evEεeTjQ)=[0DivCβ grad 0Grad000Cβ div00div00 grad (T0k)1](evEεeTjQ),(51) where Cβ=T0β(2μ+3λ). To obtain a suitable mixed formulation, the system is first put into weak form by considering the test functions vvC(Ω,Rd),VεC(Ω,Rsymd×d),vTC(Ω),vQC(Ω,Rd): (52a) vv, ρtevΩ=vv,  DivEεΩvv, Cβ grad eTΩ,(52a) (52b) Vε, D1tEεΩ=Vε,  Grad evΩ,(52b) (52c) vT, ρcϵT0teTΩ=vT, Cβ div evΩvT,  div jQΩ,(52c) (52d) 0=vQ,  grad eTΩvQ, (T0k)1jQΩ.(52d)

The notation ·, ·Ω indicates a suitable L2 inner product over the domain, depending on the nature (scalar, vectorial or tensorial) of the considered variables. Two different mixed formulations can be obtained, depending on which lines undergo the integration by parts.

4.1. First mixed formulation

The first mixed formulation is obtained by integrating by parts the Div and grad operators in line Equation(52a) and the second div operator in Equation(52c). The following system is then obtained (53) vv, ρtevΩ=Grad vv, EεΩ+div vv, CβeTΩ+vv, (EεCβeT)·nΩ,Vε, D1tEεΩ=Vε,  Grad evΩ,vT, ρcϵT0teTΩ=vT, Cβ div evΩ+grad vT, jQΩ+vT, jQ·nΩ,0=vQ,  grad eTΩvQ, (T0k)1jQΩ.(53) where ·, ·Ω indicates a suitable L2 inner product over the boundary. In this formulation Neumann boundary conditions are natural ones. Introducing a Galerkin approximation (54) vv=i=1nvϕvivvi,Vε=i=1nεΦεivεi,vT=i=1nTϕTivTi,vQ=i=1nQϕQivQi,ev=i=1nvϕvievi,Eε=i=1nεΦεieεi,eT=i=1nTϕTieTi,jQ=i=1nQϕQijQi,(54) the following system is obtained (55) [Mρ0000M@@D10000MρcϵT000000](e.ve.εe.Tj.Q)=[0DGradDAβ*0DGrad000DAβ*00Dgrad00DgradRQ](eveεeTjQ),(55) where, for simplicity, homogeneous boundary conditions have been assumed: (EεCβeT)·n|Ω=0,jQ·n|Ω=0.

The mass matrices are defined as follows Mρij=ϕvi, ρϕvjΩ,i,j{1,nv},M@@D1ij=Φεi, D1ΦεjΩ,i,j{1,nε},MρcϵT0ij=ϕTi, ρcϵT0ϕvjΩ,i,j{1,nT},

The matrices DGrad,Dgrad are given by DGradij=Φεi, Grad ϕvjΩ,i{1,nε},j{1,nv},Dgradij=ϕQi, gradϕTjΩ,i{1,nQ},j{1,nT},

The coupling matrix DAβ* arises from the discretization of the coupling operator Aβ* DAβ*ij=ϕTi, Aβ*ϕvjΩ=ϕTi, Cβdiv ϕvjΩ,i{1,nT},j{1,nv}.

The dissipation matrix reads RQij=ϕQi, (T0k)1ϕQjΩ,i,j{1,nQ}.

Suitable mixed finite elements for elastodynamics and heat equations that prove compatible with this discretization are detailed in [Citation33, Citation34], respectively.

4.2. Second mixed formulation

The second mixed formulation is obtained by integrating by parts the Grad operator in line (52 b), the first div operator in (52c) and the grad operator in (52d). The following system is then obtained (56) vv, ρtevΩ=vv,  Div EεΩvv, Cβ grad eTΩ,Vε, D1tEεΩ=Div Vε, evΩ+Vε·n, evΩ,vT, ρcϵT0teTΩ=grad vT, CβevΩvT,  div jQΩvTCβ·n, evΩ,0=div vQ, eTΩvQ, (T0k)1jQΩvQ·n, eTΩ.(56) where ·, ·Ω indicates a suitable L2 inner product over the boundary. In this formulation Dirichlet boundary conditions are natural: ev|Ω:=Velocity,eT=TT0T0Ω:=DimensionlessTemperature.

Introducing the Galerkin approximation (54), the following system is obtained (57) [Mρ0000MD10000MρcϵT000000](e.ve.εe.Tj.Q)=[0DDivDAβ0DDiv000DAβ00Ddiv00DdivRQ](eveεeTjQ).(57)

The matrices DDiv,Ddiv,DAβ are given by (58) DDivij=ϕvi,  Div ΦεjΩ,i{1,nv},j{1,nε},Ddivij=ϕTi,  div ϕQjΩ,i{1,nT},j{1,nQ},DAβij=ϕvi, AβϕTjΩ=ϕvi, Cβ grad ϕTjΩ,i{1,nv},j{1,nT}.(58)

For this discretization, stable mixed elements for elastodynamics can be found in [Citation35], and for the heat equation in [Citation36].

5. Validation of the model: The Danilovskaya problem

In this section the pH discretization of the Danilovskaya problem [Citation37] is performed. For this problem an analytical solution in the Laplace domain is available [Citation38]. First the pH formulation is illustrated, second the discretization strategy is briefly discussed. Numerical results are then presented.

5.1. The Danilovskaya problem

The Danilovskaya problem is a one-dimensional thermoelastic model in the infinite half-space x0. The initial conditions for this problem are all null. The system is excited by a sudden thermal heating at x = 0. Furthermore, the variables vanish at . Consequently, the following boundary conditions apply T(0,t)=T1H(t),limxT(x,t)=0,σET(0,t)=0,limxu(x,t)=0, where H(t) is the Heaviside function. Since the effect of the elastic vibration on the thermal field is weak, a dimensionless constant cδ is usually introduced to strengthen the coupling from the mechanical to the thermal domain [Citation39]. This dimensionless constant reads (59) cδ=δρcϵ(2μ+λ)β2(3λ+2μ)2T0,(59) where δ{0,1} is a variable for switching on and off the strong coupling from the mechanical to the thermal domain. The problem can now be recast as a pH system in co-energy variables (60) [ρ0000(2μ+λ)10000ρcϵT000000]t(eveεeTjQ)=[0xAβ0x000cδAβ*00x00x(T0k)1](eveεeTjQ),(60) where Aβ(·):=x(Cβ ·) (cf. EquationEq. (18)). Notice that the coupling parameter cδ breaks the Hamiltonian structure. The boundary conditions in the pH variables read (61) eT(0,t)=T1T0T0H(t),(eεCβev)(0,t)=0,(61) (62) limxeT(x,t)=0,    limxev(x,t)=0.(62)

Remark 4

(Boundary conditions for the numerical simulation). In the numerical simulation, the vanishing conditions at (62) are replaced by Neumann conditions at the extremity of the simulation domain Ω={0,L} [Citation39] (63) (eεCβev)(L,t)=0,jQ(L,t)=0.(63)

5.2. Discretization of the thermoelastic system

The first mixed formulation, detailed in §4.1, is employed here. This choice leads to the following weak form for the numerical domain Ω={0,L} (64) vv, ρtevΩ=xvv, eεΩ+Aβ*vv, eTΩ+γ0vv, γn(eεCβeT)Ω,vε, (2μ+λ)1teεΩ=+vε, xevΩ,vT, ρcϵT0teTΩ=vT, cδAβ*evΩ+xvT, jQΩγ0vT, γnjQΩ,0=vj, xeTΩvj, (T0k)1ejΩ,(64)

where vv,vε,vT,vj are the test functions. For this discretization the boundary condition eT(0,t)=T1T0T0H(t),

is imposed strongly as an essential condition. The other boundary terms disappear because of Equation(63). The following system is obtained (65) [Mρ0000M(2μ+λ)10000MρcϵT000000](e.ve.εe.Te.Q)=[0DGradDAβ*0DGrad000cδDAβ*00Dgrad00DgradRQ](eveεeTeQ).(65)

5.3. Numerical results

To assess the validity of the solution, the numerical results are compared with the analytical solution in the Laplace domain. The dimensionless displacement field û and temperature θ are introduced û=(λ+2μ)CxCβu,T̂=TT0T0.

In the parameters for the simulation are reported. The constant Cx, Cv are the characteristic length and velocity of the problem [Citation39]. The dimensionless constant L̂,t̂end are the dimensionless length and time of the problem. The solution in the Laplace domain for the dimensionless variable is given by [Citation38] (66) T̂(s)=1s(C12C22)[(C12s2)exp(C1x̂)(C22s2)exp(C2x̂)],û(s)=1s(C12C22)[C1exp(C1x̂)C2exp(C2x̂)],(66) where x̂=x/Cx is the dimensionless space variable and C1,C2 are given by (67) C1(s)=[s2[(1+δ+s)+[(1+δ+s)24s]12]]12,C2(s)=[s2[(1+δ+s)[(1+δ+s)24s]12]]12.(67)

Table 1. Settings and parameters for the thermoelastic problem.

For the semi-discretization in space, Continous Galerkin elements of order 1 (CG1) are employed for ev, eT, while Discontinous Galerkin of order 0 (DG0) are used for eε,jQ. This choice is in accordance with the Finite Elements constructed in [Citation33, Citation34]. Given the differential-algebraic nature of the problem, an implicit method is required to perform the time integration. For this reason, the Crank-Nicholson scheme is used. The matrices are constructed using the Firedrake finite element library [Citation40].

In the analytical and numerical displacement and temperature at x̂=1 are compared for weak δ = 0 and strong coupling δ = 1. The inverse of the Laplace transform is computed using the de Hoog method [Citation41] (available through the invertlaplace function of the mpmath Python library). The displacement is retrieved from the velocity field using the trapezoidal rule. The numerical solution matches the analytical one, thus assessing the validity of the model (60) and its discretization (65). In and the numerical solutions for the dimensionless displacement and temperature are reported for weak δ = 0 and strong coupling δ = 1.

Figure 3. Dimensionless displacement and temperature at x̂=1.

Figure 3. Dimensionless displacement and temperature at x̂=1.

Figure 4. Displacement solution for the Danilovskaya problem.

Figure 4. Displacement solution for the Danilovskaya problem.

Figure 5. Temperature solution for the Danilovskaya problem.

Figure 5. Temperature solution for the Danilovskaya problem.

6. Conclusion

It has been shown that classical linear thermoelastic problems are equivalent to two coupled port-Hamiltonian systems. This is especially interesting for the simulation of thermoelastic phenomena: each subsystem can be discretized separately and then coupled to the other using the discretized coupling operator. This allows to easily track how the energy flows between the two physics. Two different discretization has been proposed, depending on which kind of boundary conditions are to be weakly enforced. The best strategy is of course problem dependent. This new model of thermoelasticity may be of interest for control theorists and practitioners, given the increasing importance of port-Hamiltonian systems in control theory.

Finally, this contribution also discusses the results of discretization on a model problem only in the uni-dimensional case, where all the differential operators reduce to the same. An important point that deserves additional attention is the construction of stable mixed finite elements for the general three-dimensional problem. Reliable numerical models can then be employed for generating model-based control actions. Important future developments may include the reformulation of thermoelastic linear shells as well as non-linear thermoelasticity within the pH framework.

Acknowledgments

This work is supported by the project ANR-16-CE92-0028, entitled Interconnected Infinite-Dimensional systems for Heterogeneous Media, INFIDHEM, financed by the French National Research Agency (ANR) and the Deutsche Forschungsgemeinschaft (DFG). Further information is available at https://websites.isae-supaero.fr/infidhem/the-project.

References

  • R. Rashad, F. Califano, A. J. van der Schaft, and S. Stramigioli, “Twenty years of distributed port-Hamiltonian systems: A literature review,” IMA J. Math. Control Inform., vol. 37, no. 4, pp. 1400–1422, 2020. DOI: 10.1093/imamci/dnaa018.
  • M. Kurula, H. Zwart, A. J. van der Schaft, and J. Behrndt, “Dirac structures and their composition on Hilbert spaces,” J. Math. Anal. Appl., vol. 372, no. 2, pp. 402–422, 2010. DOI: 10.1016/j.jmaa.2010.07.004.
  • A. Macchelli and C. Melchiorri, “Modeling and control of the Timoshenko beam. The distributed port Hamiltonian approach,” SIAM J. Control Optim., vol. 43, no. 2, pp. 743–767, 2004. DOI: 10.1137/S0363012903429530.
  • S. Aoues, F. L. Cardoso-Ribeiro, D. Matignon, and D. Alazard, “Modeling and control of a rotating flexible spacecraft: A port-Hamiltonian approach,” IEEE Trans. Contr. Syst. Technol., vol. 27, no. 1, pp. 355–362, 2019. DOI: 10.1109/TCST.2017.2771244.
  • A. Brugnoli, D. Alazard, V. Pommier-Budinger, and D. Matignon, “Port-Hamiltonian formulation and symplectic discretization of plate models. Part I: Mindlin model for thick plates,” Appl. Math. Modell., vol. 75, pp. 940–960, Nov. 2019. DOI: 10.1016/j.apm.2019.04.035.
  • A. Brugnoli, D. Alazard, V. Pommier-Budinger, and D. Matignon, “Port-Hamiltonian formulation and symplectic discretization of plate models. Part II: Kirchhoff model for thin plates,” Appl. Math. Modell., vol. 75, 961–981, Nov. 2019. DOI: 10.1016/j.apm.2019.04.036.
  • P. Kotyczka, “Structured discretization of the heat equation: Numerical properties and preservation of flatness,” in 23rd MTNS, Hong Kong, Jul. 16–20, 2018.
  • P. Kotyczka, Numerical Methods for Distributed Parameter Port-Hamiltonian Systems, Technical University of Munich: TUM University Press, 2019.
  • F. L. Cardoso-Ribeiro, D. Matignon, and V. Pommier-Budinger, “Port-Hamiltonian model of two-dimensional shallow water equations in moving containers,” IMA J. Math. Control Inform., vol. 37, no. 4, pp. 1348–1366, Jul. 2020. DOI: 10.1093/imamci/dnaa016.
  • R. Rashad, F. Califano, F. P. Schuller, and S. Stramigioli, “Port-Hamiltonian modeling of ideal fluid flow: Part I. Foundations and kinetic energy,” Journal of Geometry and Physics, pp. 104201, 2021a. DOI: 10.1016/j.geomphys. 2021.104201.
  • R. Rashad, F. Califano, F. P. Schuller, and S. Stramigioli, “Port-Hamiltonian modeling of ideal fluid flow: Part II. Compressible and incompressible flow,” Journal of Geometry and Physics, pp. 104199, 2021b. DOI: 10.1016/j.geomphys.2021.104199.
  • R. Altmann and P. Schulze, “A Port-Hamiltonian formulation of the Navier–Stokes equations for reactive flows,” Syst. Control Lett., vol. 100, pp. 51–55, 2017. DOI: 10.1016/j.sysconle.2016.12.005.
  • F. L. Cardoso-Ribeiro, D. Matignon, and L. Lefèvre, “A partitioned finite element method for power-preserving discretization of open systems of conservation laws,” IMA J. Math. Control Inform., vol. 12, 2020. DOI: 10.1093/imamci/dnaa038.
  • R. C. Kirby and T. T. Kieu, “Symplectic-mixed finite element approximation of linear acoustic wave equations,” Numer. Math., vol. 130, no. 2, pp. 257–291, Jun. 2015. DOI: 10.1007/s00211-014-0667-4.
  • G. Avalos and I. Lasiecka, “Boundary controllability of thermoelastic plates via the free boundary conditions,” SIAM J. Control Optim., vol. 38, no. 2, pp. 337–383, 2000. DOI: 10.1137/S0363012998339836.
  • S. W. Hansen and B. Y. Zhang, “Boundary control of a linear thermoelastic beam,” J. Math. Anal. Appl., vol. 210, no. 1, pp. 182–205, 1997. DOI: 10.1006/jmaa.1997.5437.
  • M. A. Biot, “Thermoelasticity and irreversible thermodynamics,” J. Appl. Phys., vol. 27, no. 3, pp. 240–253, 1956. DOI: 10.1063/1.1722351.
  • D. E. Carlson, “Linear thermoelasticity,” in Linear Theories of Elasticity and Thermoelasticity: Linear and Nonlinear Theories of Rods, Plates, and Shells, C. Truesdell, Ed. Berlin, Heidelberg: Springer, 1973, pp. 297–345. DOI: 10.1007/978-3-662-39776-3_.2.
  • R. B. Hetnarski and M. R. Eslami, Thermal Stresses: Advanced Theory and Applications, vol. 158. Springer Netherlands, Dordrecht: Springer, 2009.
  • R. Abeyaratne, Lecture Notes on the Mechanics of Elastic Solids. Volume II: Continuum Mechanics, 1st ed., Cambridge, MA: American Society of Mechanical Engineers, 2012.
  • C. Beattie, V. Mehrmann, H. Xu, and H. Zwart, “Linear port-Hamiltonian descriptor systems,” Math. Control Signals Syst., vol. 30, no. 4, p. 17, 2018. DOI: 10.1007/s00498-018-0223-3.
  • V. Mehrmann and R. Morandin, “Structure-preserving discretization for port-Hamiltonian descriptor systems,” in 2019 IEEE 58th Conference on Decision and Control (CDC), Nice (France), 2019, pp. 6863–6868. DOI: 10.1109/CDC40024.2019.
  • V. Duindam, A. Macchelli, S. Stramigioli, and H. Bruyninckx, Modeling and Control of Complex Physical Systems, Springer-Verlag Berlin Heidelberg, 2009.
  • A. Serhani, G. Haine, and D. Matignon, “Anisotropic heterogeneous n-D heat equation with boundary control and observation: I. Modeling as port-Hamiltonian system,” IFAC-PapersOnLine, vol. 52, no. 7, pp. 51–56, 2019. 3rd IFAC Workshop on Thermodynamic Foundations for a Mathematical Systems Theory TFMST 2019. DOI: 10.1016/j.ifacol.2019.07.009.
  • A. Serhani, G. Haine, and D. Matignon, “Anisotropic heterogeneous n-D heat equation with boundary control and observation: II. Structure-preserving discretization,” IFAC-PapersOnLine, vol. 52, no. 7, pp. 57–62, 2019. 3rd IFAC Workshop on Thermodynamic Foundations for a Mathematical Systems Theory TFMST 2019. DOI: 10.1016/j.ifacol.2019.07.010.
  • A. Brugnoli, “A Port-Hamiltonian formulation of flexible structures. Modelling and structure-preserving finite element discretization,” Ph.D. thesis, Université de Toulouse, ISAE-SUPAERO, France, 2020.
  • P. Chadwick, “On the propagation of thermoelastic disturbances in thin plates and rods,” J. Mech. Phys. Solids, vol. 10, no. 2, pp. 99–109, 1962. DOI: 10.1016/0022-5096(62)90013-3.
  • R. Lifshitz and M. L. Roukes, “Thermoelastic damping in micro- and nanomechanical systems,” Phys. Rev. B, vol. 61, no. 8, pp. 5600–5609, 2000. DOI: 10.1103/PhysRevB.61.5600.
  • J. E. Lagnese, Boundary Stabilization of Thin Plates, Society for Industrial and Applied Mathematics, University City, Philadelphia, 1989. DOI: 10.1137/1.9781611970821.
  • J. G. Simmonds, “ Major simplifications in a current linear model for the motion of a thermoelastic plate,” Q. Appl. Math., vol. 57, no. 4, pp. 673–679,1999. DOI: 10.1090/qam/1724299.
  • A. N. Norris, “Dynamics of thermoelastic thin plates: A comparison of four theories,” J. Therm. Stress., vol. 29, no. 2, pp. 169–195, 2006. DOI: 10.1080/01495730500257482.
  • W. Nowacki, Thermoelasticity, Oxford, UK: Pergamon Press, 1962.
  • G. Cohen and S. Fauqueux, “Mixed spectral finite elements for the linear elasticity system in unbounded domains,” SIAM J. Sci. Comput., vol. 26, no. 3, pp. 864–884, 2005. DOI: 10.1137/S1064827502407457.
  • Z. Weng, X. Feng, and P. Huang, “A new mixed finite element method based on the Crank–Nicolson scheme for the parabolic problems,” Appl. Math. Modell., vol. 36, no. 10, pp. 5068–5079, 2012. DOI: 10.1016/j.apm.2011.12.044.
  • D. Arnold and J. Lee, “Mixed methods for elastodynamics with weak symmetry,” SIAM J. Numer. Anal., vol. 52, no. 6, pp. 2743–2769, 2014. DOI: 10.1137/13095032X.
  • S. Memon, N. Nataraj, and A. K. Pani, “An a posteriori error analysis of mixed finite element Galerkin approximations to second order linear parabolic problems,” SIAM J. Numer. Anal., vol. 50, no. 3, pp. 1367–1393, 2012. DOI: 10.1137/100782760.
  • V. I. Danilovskaya, “Thermal stresses in an elastic half-space due to sudden heating of its boundary,” Prikladnaya Matematika i Mech., vol. 14, no. 316–318, 1950. (in Russian)
  • M. Balla, “Analytical study of the thermal shock problem of a half-space with various thermoelastic models,” Acta Mech., vol. 89, no. 1–4, 73–92, Mar. 1991. DOI: 10.1007/BF01171248.
  • E. Rabizadeh, A. Saboor Bagherzadeh, and T. Rabczuk, “Goal-oriented error estimation and adaptive mesh refinement in dynamic coupled thermoelasticity,” Comput. Struct., vol. 173, pp. 187–211, 2016. DOI: 10.1016/j.compstruc.2016.05.024.
  • F. Rathgeber, D. A. Ham, L. Mitchell, M. Lange, F. Luporini, A. T. T. McRae, G. T. Bercea, G. R. Markall, and P. H. J. Kelly, “Firedrake: Automating the finite element method by composing abstractions,” ACM Trans. Math. Softw., vol. 43, no. 3, pp. 1–27, 2017. DOI: 10.1145/2998441.
  • F. R. de Hoog, J. H. Knight, and A. N. Stokes, “An improved method for numerical inversion of Laplace transforms,” SIAM J. Sci. Stat. Comput., vol. 3, no. 3, pp. 357–366, 1982. DOI: 10.1137/0903022.