265
Views
2
CrossRef citations to date
0
Altmetric
Articles

Shape optimization of unsteady natural convection fields

, &
Pages 691-707 | Received 12 Dec 2016, Accepted 25 May 2017, Published online: 08 Jun 2017

Abstract

This paper presents a numerical method for two shape optimization problems, namely, prescribing the temperature history distribution on sub-boundaries and maximizing the discharged heat on sub-boundaries of unsteady natural convection fields. The square error integral between the actual temperature distribution and the target temperature distribution on the sub-boundaries during a specified period of time was used as the objective functional for the prescribed temperature history distribution. The shape gradients of these shape determination problems were derived theoretically using the Lagrange multiplier method, adjoint variable method, and the material derivative formulae. Reshaping was performed by the traction method, which was proposed as an approach for solving shape optimization problems. Numerical programs for the shape determination problems are developed based on FreeFem++ in order to verify the proposed method.

1. Introduction

Research on the shape optimization of flow fields began with Pironneau [Citation1,Citation2], who proposed a method for updating shapes by treating displacements in the normal direction of the boundary as design variables in the distribution system and evaluating the shape gradient functions (sensitivities) by selecting the node coordinates on the boundary of a finite-element model as design variables. When the finite element method (FEM) was used for numerical analysis, the mesh within the region had to be updated after the boundary was updated [Citation3]. An unstable phenomenon involving the undulation of boundary shapes has been known to occur when shape gradient functions related to the displacement of such design variables were evaluated, and the resulting sensitivities were used to move nodes [Citation4]. Jameson [Citation5] proposed a numerical method for updating the boundary shapes using an elliptic equation defined on the boundary. This method replaced the actual sensitivities with smooth sensitivities in order to facilitate the use of the second-order differentiation conducted during the sensitivity evaluation in updating shapes in flow fields. Mohammadi et al. [Citation6] also proposed a similar method.

The traction method [Citation7,Citation8] was proposed for solving equations governing domain variables for linear elastic fields. Theoretical consideration for retaining the boundary smoothness following changes is possible with this method, and the undulating phenomenon did not occur [Citation9]. Furthermore, remeshing is not required for ordinary shape update analyses because the domain variation of all nodes within a region can be treated as design variables in the FEM while still retaining smoothness. This method enables sensitivity calculations of large-scale design variables for complex coupled fields based on the adjoint variable method [Citation10].

Shape determinations in natural convection fields have been studied by Momose et al. [Citation11], Park et al. [Citation12,Citation13], and Aounallah et al. [Citation14]. Momose et al. [Citation11] have proposed a shape sensitivity analysis method seeking to maximize the velocity in the sub-domain in natural convection fields; however, this method was limited to steady-state problems. Park et al. [Citation12,Citation13] used the adjoint variable method for shape identification problems in unsteady natural convection fields. They proposed to limit the design variables that express boundary shapes to the minimum number of variables required for the steady-state problem. Aounallah et al. [Citation14] in two-dimensions expressed the equations governing unsteady natural convection fields using flow functions and vorticity. They conducted an analysis using Bezier curves, which facilitate expressions of smooth boundary shapes with a minimal number of design variables, similar to Park et al. [Citation12,Citation13]. Topology shape optimization analyses have recently been conducted on heat convection fields by Yaji et al. [Citation15] and Alexandersen et al. [Citation16]. Yaji et al. [Citation15] optimized the topology of forced convection fields, while Alexandersen et al. [Citation16] used the adjoint variable method to optimize the topology of natural convection fields. However, in all these cases, the analysis was limited to steady-state problems.

This study considers two shape determination problems for unsteady natural convection fields. The first problem involves identification of shapes when prescribing the temperature distribution history on the sub-boundaries in natural convection fields. The problem of minimizing the integrated squared error between the actual temperature distribution and the target temperature distribution on sub-boundary is formulated in Section 4.1, and the shape gradient is derived using either the Lagrange multiplier method or the adjoint variable method and the material derivative function in Section 4.2. The second problem involves the determination of shapes for maximizing the amount of heat discharged on a sub-boundary. The problem is formulated, and the shape gradient density function is derived in Section 5 in a similar manner to that of the first problem. The traction method, Section 6, is applied to examples in Section 7 based on the shape gradient density functions derived for each problem.

2. Domain variation

Before formulating the shape optimization problem, a method of representing domain variation using the speed method will be discussed briefly. A more detailed explanation is presented in [Citation17].

Assume that a bounded domain Ω of Rd(d=2,3) with boundary Γ is variable. One approach to describing the domain variation is to use a one-parameter family of one-to-one mappings Ts(Ω):ΩΩs or its inverse Ts-1(Ωs):ΩsΩ, where s denotes the domain variation history. When a domain functional JΩ and a boundary functional JΓ of a distributed function ψ are considered, their derivatives J˙Ω and J˙Γ with respect to s at s=0 are given by the formulae of the material derivative:(1) JΩ=Ωψdx,J˙Ω=Ωψdx+Γψn·Vdx,JΓ=ΓψdΓ,J˙Γ=Γψ+nψ+ψκn·VdΓ,(1)

where n is an outward unit normal vector to the boundary, n(·)(·)·n, and κ denotes the quantity (d-1) times the average boundary curvature. The shape derivative ψ of the distributed function ψ indicates that the derivatives are fixed in spatial coordinates. The derivative V(Ωs) of Ts(Ω) with respect to s given by(2) V(Ωs)=TsΩs=TssTs-1(Ωs)(2)

is referred to as the velocity because of the analogy between s and time.

3. Governing equations for unsteady natural convection fields

Consider the unsteady natural convection field in the region Ω of Rd(d=2,3) in time interval [0, T]. Consider determining the flow velocity u(x,t)=(ui(x,t))i=1,d¯, pressure p(x,t) and temperature θ(x,t) in space xΩ at time t[0,T]. The Boussinesq-approximated Navier–Stokes equation, continuity equation, and energy equation are the governing equations for unsteady natural convection fields. They can be expressed as follows:(3) ρuit+ρujui,j=-p,i+μui,jj-eiρβg(θ-θ0),(x,t)Ω×[0,T],(3) (4) ui,i=0,(x,t)Ω×[0,T],(4) (5) θt+ujθ,j=αθ,jj,(x,t)Ω×[0,T],(5)

where the boundary Γ=Ω=ΓuΓσ=ΓθΓqΓh, see Figure .

Figure 1. Natural convection field.

Figure 1. Natural convection field.

The boundary conditions and initial conditions are:(6) ui(x,t)=u^i(x,t),t[0,T],xΓu,(6) (7) σi(x,t)=σ^i(x,t)=(-pδij+μui,j)nj=0,t[0,T],xΓσ,(7) (8) θ(x,t)=θ^(x,t),t[0,T],xΓθ,(8) (9) -αθ(x,t),jnj=q^(x,t),t[0,T],xΓq,(9) (10) -αθ(x,t),jnj=h^(θ(x,t)-θ^f),t[0,T],xΓh,(10) (11) ui(x,0)=uiini(x),xΩ,(11) (12) p(x,0)=pini(x),xΩ,(12) (13) θ(x,0)=θini(x),xΩ.(13)

Here, ρ represents the density, θ0 represents the base temperature, μ represents the dynamic viscosity, e1=e3=0,e2=-1, and g represents the acceleration of gravity. Moreover, β represents the coefficient of thermal expansion, α represents the thermal diffusivity, q^ and h^ represent the heat flux and the heat transfer coefficient normalized by heat capacity, θ^f represents the external temperature, uiini represents the initial flow velocity, pini represents the initial pressure, θini represents the initial temperature, δij represents the Kronecker delta. Finally, (·^) represents the known function on the boundary.

The weak forms of the respective governing Equations (Equation4)–(Equation6) can be expressed with adjoint flow velocity w(x,t)=(wi(x,t))i=1,d¯, adjoint pressure q(x,t) and adjoint temperature ξ(x,t) as follows:(14) 0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+c(w,p)+fV(w,θ)-l(w)}dt=0,wW,(14) (15) 0T{c(u,q)}dt=0,qQ,(15) (16) 0T{tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fqH(ξ)+fhH(θ,ξ)-fhfH(ξ)}dt=0,ξΞ.(16)

Furthermore, tV(u,t,w), aV(u,w), bV(v,u,w), c(wp), fV(w,θ), l(w), tH(θ,t,ξ), aH(θ,ξ), bH(u,θ,ξ), fq(ξ), fhH(θ,ξ), and fhfH(ξ) are defined as follows:(17) tV(u,t,w)=Ωρwiuitdx,aV(u,w)=Ωμwi,jui,jdx,bV(v,u,w)=Ωρwivjui,jdx,c(w,p)=-Ωwi,ipdx,fV(w,θ)=Ωwieiρβg(θ-θ0)dx,l(w)=Γσwiσ^idΓ,tH(θ,t,ξ)=Ωξθtdx,aH(θ,ξ)=Ωαξ,iθ,idx,bH(u,θ,ξ)=Ωξujθ,jdx,fqH(ξ)=Γqξq^dΓ,fhH(θ,ξ)=Γhh^ξθdΓ,fhfH(ξ)=Γhh^ξθ^fdΓ.(17)

Here, the superscripts V and H denote terms about the viscosity and heat, the subscripts q and h, hf denote terms about the heat flux and the heat transfer. Tensors described in this paper use the Einstein summation convention and differentiation (·),i=(·)/xi, while (·),t expresses the time derivative of the function.

The flow velocity u, its adjoint w, and the other variables are considered to be elements of the following functional spaces:(18) U={u(x,t)H1(Ω×[0,T])d|usatisfies(7) and(12)},(18) (19) Q={q(x,t)L2(Ω×[0,T])|Ωqdx=0if measure(Γσ)=0},(19) (20) Θ={θ(x,t)H1(Ω×[0,T])|θsatisfies(9)--(11),(14)},(20) (21) W={w(x,t)H1(Ω×[0,T])d|w(x,t)=0,t[0,T],xΓu,w(x,T)=0,xΩ},(21) (22) Ξ={ξ(x,t)H1(Ω×[0,T])|ξ(x,t)=0,t[0,T],xΓθ,ξ(x,T)=0,xΩ}.(22)

4. Prescribing temperature on a sub-boundary

4.1. Problem formulation

In this section, the problem of minimizing the square integration errors between the actual temperature θ|ΓD×[t1,t2] from time t=t1[0,T] to t=t2[0,T] and the target temperature θD|ΓD×[t1,t2] on sub-boundary ΓDΓ is formulated. We assume t1<t2. The domain variation of this natural convection field region Ω is denoted by Ts, and the domain Ω is assumed to vary to reach Ωs=Ts(Ω). For simplicity, we assume that the sub-boundaries ΓD, Γσ and Γh are invariable, that is Ts(ΓD)=ΓD, Ts(Γσ)=Γσ and Ts(Γh)=Γh for domain variation. The square integration error problem for temperature distribution from time t=t1 to t=t2 is formulated as follows:(23) GivenΩfindΩsthat minimizest1t2EΓD(θ)dtsubject to(15)--(17)andΩdxβVM,(23)

where βV is a coefficient related to the initial domain measure M, and(24) EΓD(θ)=ΓD(θ-θD)2dΓ.(24)

4.2. Shape gradient function

The square integration error problem can be rewritten as a retention problem without any restrictions from the Lagrange multiplier method. In this instance, the Lagrange function L(ui,p,θ,wi,q,ξ,Λ) is given as follows:(25) L=t1t2EΓD(θ)dt-0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+c(w,p)+fV(w,θ)-l(w)}dt-0T{c(u,q)}dt-0T{tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fqH(ξ)+fhH(θ,ξ)-fhfH(ξ)}dt+ΛΩdx-βVM.(25)

where wW, qQ, and ξΞ were introduced as Lagrange multiplier functions or the adjoint functions, with respect to the weak forms. The non-negative real constant number Λ is Lagrange multiplier with respect to the volume constraint. The derivative of L with respect to domain variation is derived using the velocity field (Equation3), as follows, [Citation7,Citation8],(26) L˙=-0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+c(w,p)+fV(w,θ)-l(w)}dt-0T{c(u,q)}dt-0T{tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fqH(ξ)+fhH(θ,ξ)-fhfH(ξ)}dt-0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+bV(u,u,w)+c(u,q)+bH(u,θ,ξ)}dt-0T{c(w,p)}dt-0T{fV(w,θ)+tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fhH(θ,ξ)}dt+t1t2EΓD(θ)dt+Λ˙Ωdx-βVM+lG(V).(26)

Here, (·) represents the derivative with respect to domain variation of the function fixed on the spatial coordinates, and(27) lG(V)=ΓuGn·VdΓ,(27)

where(28) G=0T-μwi,jui,j+wi,ip-θtξ-αξ,iθ,i-ξujθ,j-n(ξq^)-(ξq^)κdt+Λ,(28)

and ui,p,θ,wi,q,ξ, and Λ are determined by the following conditions:(29) 0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+c(w,p)+fV(w,θ)-l(w)+c(u,q)}dt=0,wW,qQ,(29) (30) 0T{tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fqH(ξ)+fhH(θ,ξ)-fhfH(ξ)}dt=0,ξΞ,(30) (31) 0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+bV(u,u,w)+c(u,q)+bH(u,θ,ξ)+c(w,p)}dt=0,uU,pQ,(31) (32) 0T{fV(w,θ)+tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fhH(θ,ξ)}dt-t1t2EΓD(θ)dt=0,θΘ,(32) (33) Λ0,ΩdxβVM,ΛΩdx-βVM=0.(33)

The derivative of the Lagrange function agrees with the derivative of the evaluation function, establishing the following relationship:(34) L˙|ui,p,θ,wi,q,ξ,Λ=E˙ΓD|ui,p,θ,wi,q,ξ,Λ=lG(V).(34)

Since Gn in Equation (Equation28) is a coefficient function of the velocity field V that provides minute variations in the domain, Gn is referred to as a sensitivity function or shape gradient function. Furthermore, the scalar function G is referred to as the shape gradient density function.

Equation (Equation30) is a weak form of the Navier–Stokes equation and the continuity equation. Equation (Equation31) is a weak form of the energy equation in the state equation. Equation (Equation32) is a weak form of the Navier–Stokes equation and continuous state equation for the adjoint problem, Equation (Equation33) is a weak form of the energy equation in the state equation for the adjoint problem, and (Equation34) is a constraint equation related to the Lagrange multiplier Λ. The traction method can be applied if the shape gradient function can be evaluated by analyzing ui,p,θ,wi,q,ξ, and Λ based on these equations.

Furthermore, considering the continuity equation for the adjoint equation, and assuming that the flow velocity on the design boundaries satisfies ui=0, the shape gradient density function G can be expressed from (Equation29) as(35) G=G0+G1,G0=0T-μwi,jui,j-θtξ-αξ,iθ,i-n(ξq^)-(ξq^)κdt,G1=Λ,(35)

where G0 is the shape gradient density function for the square integration errors and G1 is the shape gradient density function for the constraint conditions.

The same procedure can also be used to derive the adjoint equation and shape gradient function when prescribing the temperature θ|ΩD×[t1,t2] in sub-domain ΩD.

The adjoint equations for the shape identification problem can be expressed as follows:(36) 0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+bV(u,u,w)+c(u,q)+bH(u,θ,ξ)+c(w,p)}dt=0,uU,pQ,(36) (37) 0T{fV(w,θ)+tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fhH(θ,ξ)}dt-t1t2EΩD(θ)dt=0,θΘ,(37) (38) Λ0,ΩdxβVM,ΛΩdx-βVM=0.(38)

5. Maximizing heat discharge on a sub-boundary

5.1. Problem formulation

The problem of maximizing the discharged heat from time t=t1 to t=t2 on a sub-boundary ΓDΓh is formulated in the same way as in Section 4.1. For simplicity, we assume that the sub-boundaries Γσ, Γθ and Γh are invariable, that is Ts(Γσ)=Γσ, Ts(Γθ)=Γθ and Ts(Γh)=Γh for domain variation. The heat discharge maximization problem from time t=t1 to t=t2 is formulated as follows:(39) GivenΩfindΩsthat maximizest1t2JΓD(θ)dtsubject to(15)--(17)and(24),(39)

where(40) JΓD(θ)=ΓDΓhh^(θ-θ^f)dΓ.(40)

5.2. Shape gradient function

The Lagrange function L(ui,p,θ,wi,q,ξ,Λ) for this problem is given by the following formula:(41) L=-t1t2JΓD(θ)dt-0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+c(w,p)+fV(w,θ)-l(w)}dt-0T{c(u,q)}dt-0T{tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fqH(ξ)+fhH(θ,ξ)-fhfH(ξ)}dt+ΛΩdx-βVM.(41)

Assuming that the flow velocity satisfies ui=0 at the design boundary, the shape gradient density function for the problem is given by (Equation36).

The adjoint flow velocity wi, adjoint pressure q, adjoint temperature ξ, and Lagrange multiplier Λ related to the heat discharge maximization problem are determined by the following adjoint equations:(42) 0T{tV(u,t,w)+aV(u,w)+bV(u,u,w)+bV(u,u,w)+c(u,q)+bH(u,θ,ξ)+c(w,p)}dt=0,uU,pQ,(42) (43) 0T{fV(w,θ)+tH(θ,t,ξ)+aH(θ,ξ)+bH(u,θ,ξ)+fhH(θ,ξ)}dt+t1t2JΓD(θ)dt=0,θΘ,and(39).(43)

6. Solution

6.1. Traction method

The traction method has been proposed for solving the velocity field VD that indicates domain variation. Domain variation is based on the following governing equation [Citation7]:(44) aE(V,y)=-<Gn,y>,yD,(44)

and(45) D={V(H1(Ω))d|constraints of domain variation}.(45)

Here, aE(V,y) is a bilinear form indicating the strain energy on a linear elastic body, which is defined by the following equation with respect to the displacement vector function ui, vi:(46) aE(u,v)=ΩAijkluk,lvi,jdx,(46)

where Aijkl is a constant elastic tensor that is positive definite. Equation (Equation44) indicates that the velocity field V is analyzed as a displacement field when the negative shape gradient function -Gni is applied as an external force. The domain variation resulting from the traction method can be derived as a displacement field when the shape gradient function is applied to the pseudo-elasticity problem as an external force.

Furthermore, this analysis is a heat convection field analysis. For simplicity, the elastic tensor Aijkl is given by(47) Aijkl=δikδjl.(47)

Taking account of a bilinear Equation (Equation44) and G=G0+G1, the domain variation is expressed as V=V0+ΛV1. We obtain(48) aE(V0,y)=-<G0n,y>,yD.aE(ΛV1,y)=-<G1n,y>,yD.(48)

The Lagrange multiplier Λ is determined from(49) Γn·V0dΓ+ΛΓn·V1dΓ=0.(49)

Finally, because V0,V1,Λ are given, the domain xΩ is moved into x+VΩs.

6.2. Analysis procedure

The proposed shape determination analysis can be performed by repeating the following steps:

Step 1.

Give initial shape.

Step 2.

Analyze the flow velocity u(x,t), pressure p(x,t) and temperature θ(x,t), based on the state Equation (Equation30) and (Equation31) using initial condition.

Step 3.

Stop when the objective functional has converged.

Step 4.

Analyze the adjoint flow velocity w(x,t), adjoint pressure q(x,t) and adjoint temperature ξ(x,t), using the given flow velocity u(x,t) and temperature θ(x,t), based on the adjoint Equations (Equation32) and (Equation33) or Equations (Equation42) and (Equation43), using the conditions wi(x,T)=0 and ξ(x,T)=0 in (Equation22) and (Equation23).

Step 5.

Calculate the shape gradient density function G based on these results using Equation (Equation36).

Step 6.

Calculate the velocity field V based on the traction method using Equation (Equation44) to update the shape. Return to Step 2.

7. Numerical experiments

To verify the proposed method, numerical programs for the above two shape determination problems were developed based on FreeFem++. This section presents simple numerical experiments analyzed using the derived shape gradient function and the traction method to solve the problem of prescribing temperature history at a sub-boundary and the problem of maximizing heat discharge.

7.1. Prescribing temperature on a sub-boundary

An analysis model was considered similarly to the model presented by Park et al. [Citation13] and shown in Figure . In Figure , the target shape is represented by dotted lines, while the initial shape is represented by solid lines. The boundary conditions of the temperature field were assumed as follows. The lower surface boundary AB was assumed to be the high temperature base boundary Γθ and was set to θ^=1 K. Surfaces DA and BC on the left and right, respectively, were assumed to be low temperature base boundaries Γθ and were set to θ^=0 K. Surface DC at Γq was assumed to be a natural boundary, and the thermal insulation condition was set to q^=0. The temperature for the target shape on surface DC was assumed to be the target temperature θD(x,t) in Equation (Equation25). The boundary conditions for the flow field assumed that all surrounding boundaries were wall boundaries set to u1=u2=0.

Figure 2. Numerical model and boundary conditions for Park’s problem.

Figure 2. Numerical model and boundary conditions for Park’s problem.

The initial conditions were set to θini=0 and uiini=0 for the entire domain. Internally generated heat was ignored, and the pressure was considered to be unique in order to achieve an average value of 0. The time was set to t1=0 and t2=T, while time integration was performed from t=0 s to t=T=600 s with a time increment of Δt=0.8 s.

The prescribed temperature on boundary ΓD was considered for the upper surface DC. The design boundary Γdesign was considered to be the high temperature base boundary of lower surface AB. The remaining boundaries were set using constraints on the shape variation, specifically low temperature base boundaries DA and BC and thermal insulation boundary DC.

The temperature field θ, flow field velocity u and pressure p, adjoint temperature ξ, adjoint flow velocity w and adjoint pressure q, and shape updating analysis (velocity field V) were all performed using FreeFem++ [18,19]. A triangular secondary element was used for θ,u,ξ,w, and V, while a triangular primary element was used for pq.

Based on the FEM, an analysis was performed using the function ‘convect’ of FreeFem++ on the state equation in the time direction. The function ‘adaptmesh’ of FreeFem++ was used to perform remeshing for shape updates; ‘adaptmesh’ is a function that generates an adaptive mesh for improving solution accuracy of the state variables derived by finite-element analysis. Adaptmesh was used to improve the accuracy of the shape gradient density function G and velocity field V. The following physical values were used based on the book [Citation18]: density ρ = 1.0 kg/m3, dynamic viscosity μ=1.0×10-3 Pa s, Prandtl number Pr=7.01, thermal diffusivity α=μ/(ρPr), gravitational acceleration g=9.8 m/s2 downward, base temperature θ0=0 K, and coefficient of thermal expansion β=2.06×10-4 1/K. The dimensions of the shape were set to h=0.5 m, b=2 m and a=0.125 m.

The value of the objective functional at iteration k was denoted by J[k], and convergence was reached when(50) |J[k-1]-J[k]|J[k]<ε,(50)

where ε>0 is a small number.

For ε=10-5, a comparison of the initial, identified and target shapes is shown in Figure . The temperature θ, flow velocity u, and streamlines at the end time t=T=600 s for the respective shapes are shown in Figure . From Figure , it was confirmed that the temperature distribution, flow velocity distributions, and streamlines in the identified shape agreed with those in the target shape.

Figure 3. Shapes with finite-element meshes for Park’s problem.

Figure 3. Shapes with finite-element meshes for Park’s problem.

Figure 4. Temperature distributions, flow velocity distributions, and streamlines at time t=T=600 for Park’s problem.

Figure 4. Temperature distributions, flow velocity distributions, and streamlines at time t=T=600 for Park’s problem.

A comparison of the target shape (dotted lines) and the outline of the obtained shape (solid lines) is shown in Figure . The target and identifying shapes are practically identical. Figure shows the iterative history for the objective functional, normalized with initial value.

Figure 5. Comparison of target shape and identified shape for Park problem.

Figure 5. Comparison of target shape and identified shape for Park problem.

Figure 6. Numerical results: Iterative history for Park’s problem.

Figure 6. Numerical results: Iterative history for Park’s problem.

Figure 7. Numerical model and boundary condition for heat discharge maximization problem.

Figure 7. Numerical model and boundary condition for heat discharge maximization problem.

Figure 8. Mesh and distribution of temperature and streamlines at time t=T= 600 for initial shape for heat discharge maximization problem.

Figure 8. Mesh and distribution of temperature and streamlines at time t=T= 600 for initial shape for heat discharge maximization problem.

Figure 9. Numerical results: Mesh and distributions of temperature and streamlines at time t=T= 600 for optimum shape for heat discharge maximization problem.

Figure 9. Numerical results: Mesh and distributions of temperature and streamlines at time t=T= 600 for optimum shape for heat discharge maximization problem.

Figure 10. Iterative histories for heat discharge maximization problem.

Figure 10. Iterative histories for heat discharge maximization problem.

These results confirm that the objective functional, which is reduced to practically zero, and the identifying shapes, which are practically identical to the target shapes, can be obtained using the proposed method.

7.2. Maximizing heat discharge on a sub-boundary

A disc with five holes, as depicted in Figure , was used as an analysis model for the problem of the maximizing heat discharge for an unsteady natural convection field.

The boundary conditions considered hole A on the lower side to be the known temperature boundary Γθ and was set to θ^=1 K. Holes B, C, D and E were considered to be thermal insulation boundaries Γq and were set to q^=0 mK/s. The external boundaries considered h^=1 m/s and θ^f=0 K as heat transfer boundaries Γh. The boundary conditions for the flow field assumed that all surrounding boundaries were wall boundaries set to u1=u2= 0. The initial conditions of the entire domain were set to θini=0 and uiini=0. Internally generated heat was ignored, and the pressure was uniquely set to achieve an average of 0. The time was set to t1=0 and t2=T, and time integration was performed from t=0 s to t=T=600 s with Δt=0.8 s time increment. The four hole boundaries B, C, D, and E comprising thermal insulation boundary Γq were considered to be design boundaries Γdesign. Other boundaries were constrained with respect to domain variation in order to analyze the problem of maximizing the heat discharge at the external heat transfer boundary Γh.

The physical values were the same as those used in the example of Section 7.1, while the dimensions were set to d0=1.2 m, di=0.3 m, and a=0.7 m. The coefficient βV, which restrains the size of the domain, was varied for three values: in Case A, βV=1; in Case B, βV=0.9; and in Case C, βV=1.1. For ε=10-5, the obtained shape for optimization resulted in a sharpened shape along the horizontal direction for hole boundaries B and E, even when the FreeFem++ adaptmesh function was used after each shape update. It was therefore difficult to converge to the heat discharge value from the given initial guess. As a result, the value of ε was set to ε=10-3.

The mesh, temperature, and streamlines of the initial shape at end time t=T=600 s are shown in Figure , with the corresponding numerical results shown for Cases A, B, and C in Figure . The heat discharge, which is the objective functional being sought by shape updates, and domain size changes are shown in Figure . In Figure , the objective functional and the volume are normalized with initial value. The shape of hole boundaries B and E was updated in the horizontal direction in all cases based on the results shown in Figure . The temperature was elevated in the heat transfer boundary near hole boundaries B and E. As a result, the obtained shape had increasing heat discharge on the external heat transfer boundary Γh. A review of the Case B results reveals that the shape was modified for vertical extension on hole boundaries C and D, where the influence of the heat discharge is relatively low. This resulted in a shape that satisfies the restraining condition βV=0.9 to reduce the size of the domain. Similarly, the shape of boundaries C and D was vertically modified in Case C as well, resulting in a shape that had an increased domain size. The results shown in Figure confirm that, in all cases, the heat discharge increased while satisfying the constraining conditions of the domain size. An improvement of about 5% occurred with respect to the initial shape.

8. Conclusion

This paper proposed a numerical method for two problems: the problem of shape identification for prescribing the temperature history for a sub-boundary in unsteady natural convection fields, and the problem of shape optimization for maximizing heat discharge on a sub-boundary. These problems were formulated, and the shape gradient density functions were derived. The verification of the proposed method was accomplished using simple examples of two-dimensional problems by applying the traction method to the shape gradient distribution functions.

In addition, in this study, we did not consider noise in the input measurement and a corresponding stopping criterion for iterations. Because these are important matters about shape determination of inverse problems, we are going to examine them as future research themes.

Notes

No potential conflict of interest was reported by the authors.

References

  • Pironneau O. On optimum profiles in Stokes flow. J Fluid Mech. 1973;59(1):117–128.
  • Pironneau O. On optimum design in fluid mechanics. J Fluid Mech. 1974;64(1):97–110.
  • Pironneau O. Optimal shape design for elliptic systems. Berlin: Springer-Verlag; 1984.
  • Imam MH. Three-dimensional shape optimization. Int J Numer Methods Eng. 1982;18:661–673.
  • Jameson A. Optimum aerodynamic design using control theory. In: Hafez M, Oshima K, editors. Computational fluid dynamics review 1995. New York (NY): Wiley; 1995. p. 495–528.
  • Mohammadi B, Pironneau O. Applied shape optimization for fluids. New York(NY): Oxford University Press; 2001.
  • Azegami H. Solution to domain optimization problems. Trans Japan Soc Mech Eng Ser A. 1994;60:1479–1486. Japanese.
  • Azegami H, Shimoda M, Katamine E, et al. A Domain optimization technique for elliptic boundary value problems. In: Hernandez S, Brebbia CA, editors Computer aided optimum design of structures IV. Southampton: Structural optimization, Computational Mechanics Publications; 1995. p. 51–58.
  • Azegami H, Kaizu S, Shimoda M, et al. Irregularity of shape optimization problems and an improvement technique. In: Hernandez S, Brebbia CA, editors Computer aided optimum design of structures V. Southampton: Computational Mechanics Publications; 1997. p. 309–326.
  • Katamine E, Kiriyama Y, Azegami H. Multi objective shape optimization in forced heat-convection fields. Trans Japan Soc Mech Eng Ser B. 2013;89:2239–2253. Japanese.
  • Momose K, Kawano H, Kawahara G. An approach for shape optimization with sensitivity analysis in heat convection fields. Proc Thermal Eng Conf. 2009;:177–178. Japanese
  • Park HM, Ku JH. Shape identification for natural convection problems. Commun Numer Methods Eng. 2001;17:871–880.
  • Park HM, Shin HJ. Shape identification for natural convection problems using the adjoint variable method. J Comput Phys. 2003;186:198–221.
  • Aounallah M, Belkadi M, Adjlout L, et al. Numerical shape optimization of a confined cavity in natural convection regime. Comput Fluids. 2013;75:11–21.
  • Yaji K, Yamada T, Kubo S, et al. A topology optimization method for a coupled thermal-fluid problem using level set boundary expressions. Int J Heat Mass Transfer. 2015;81:878–888.
  • Alexandersen J, Aage N, Andreasen CS, et al. Topology optimization for natural convection problems. Int J Numer Methods Fluids. 2014;76(10):699–721.
  • Haug EJ, Choi KK, Komkov V. Design sensitivity analysis of structural systems. New York (NY): Academic Press; 1986.
  • Ootsuka K, Takaishi T. Finite element analysis using mathematical programming language FreeFem++. Tokyo: Kyoritsu; 2014. Japanese.
  • Hecht F. New development in FreeFem++. J Numer Math. 2012;20(3–4):251–265.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.