Skip to main content
Log in

Revisiting density-based topology optimization for fluid-structure-interaction problems

  • RESEARCH PAPER
  • Published:
Structural and Multidisciplinary Optimization Aims and scope Submit manuscript

Abstract

This study revisits the application of density-based topology optimization to fluid-structure-interaction problems. The Navier-Cauchy and Navier-Stokes equations are discretized using the finite element method and solved in a unified formulation. The physical modeling is limited to two dimensions, steady state, the influence of the structural deformations on the fluid flow is assumed negligible, and the structural and fluid properties are assumed constant. The optimization is based on adjoint sensitivity analysis and a robust formulation ensuring length-scale control and 0/1 designs. It is shown, that non-physical free-floating islands of solid elements can be removed by combining different objective functions in a weighted multi-objective formulation. The framework is tested for low and moderate Reynolds numbers on problems similar to previous works in the literature and two new flow mechanism problems. The optimized designs are consistent with respect to benchmark examples and the coupling between the fluid flow, the elastic structure and the optimization problem is clearly captured and illustrated in the optimized designs. The study reveals new features of topology optimization of FSI problems and may provide guidance for future research within the field.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22

Similar content being viewed by others

References

  • Alexandersen J, Aage N, Andreasen CS, Sigmund O (2014) Topology optimisation for natural convection problems. Int J Numer Meth Fluids, pp 699–721

  • Andreasen CS, Sigmund O (2011) Topology optimization of fluid-structure-interaction problems in poroelasticity. SMO 43:5

    MATH  Google Scholar 

  • Andreasen CS, Sigmund O (2013) Topology optimization of fluid–structure-interaction problems in poroelasticity. Comput Methods Appl Mech Eng 258(C):55–62

    Article  MathSciNet  MATH  Google Scholar 

  • Andreasen CS, Sigmund O, Gersborg-Hansen A (2009) Topology optimization of microfluidic mixers. Int J Numer Methods Fluids 61(5):498–513

    Article  MathSciNet  MATH  Google Scholar 

  • Bendsøe M, Kikuchi N (1988) Generating optimal topologies in structural design using a homogenization method. Comput Methods Appl Mech Eng, 1–28

  • Bendsøe M, Sigmund O (2003) Topology optimization - theory, methods and applications. Springer

  • Borrvall T, Petersson J (2003) Topology optimization of fluids in Stokes flow. Int J Numer Methods Fluids 41(1):77–107

    Article  MathSciNet  MATH  Google Scholar 

  • Brooks AN, Hughes T JR (1982) Streamline upwind/petrov-galerkin formulations for convection dominated flows with particular emphasis on the incompressible navier-stokes equations. Comput Methods Appl Mech Eng 32 (1):199–259

    Article  MathSciNet  MATH  Google Scholar 

  • Chen X (2016) Topology optimization of microfluidics — a review. Microchemi J 127:52–61

    Article  Google Scholar 

  • Christiansen RE, Lazarov BS, Jensen JS, Sigmund O (2015) Creating geometrically robust designs for highly sensitive problems using topology optimization: acoustc cavity design. Struct Multidiscip Optim 52(4):737–754

    Article  MathSciNet  Google Scholar 

  • Cook RD, Malkus DS, Plesha ME, Witt RJ (2002) Concepts and applications of finite element analysis, 4th edn. John Wiley & Sons Ltd

  • De Leon DM , Alexandersen J, Fonseca JS, Sigmund O (2015) Stress-constrained topology optimization for compliant mechanism design. Struct Multidiscip Optim 52(5):929–943

    Article  MathSciNet  Google Scholar 

  • Deng Y, Liu Z, Zhang P, Liu Y, Wu Y (2011) Topology optimization of unsteady incompressible Navier-Stokes flows. J Comput Phys 230(17):6688–6708

    Article  MathSciNet  MATH  Google Scholar 

  • Deng Y, Liu Z, Wu Y (2013) Topology optimization of steady and unsteady incompressible Navier-Stokes flows driven by body forces. Struct Multidiscip Optim 47(4):555–570

    Article  MathSciNet  MATH  Google Scholar 

  • Deuflhard P (2014) Newton methods for nonlinear problems. Affine Invar Adapt Algor, 1–437

  • Dowell EH, Hall KC (2001) Modeling of fluid-structure interaction. Annu Rev Fluid Mech 33(1):445–490

    Article  MATH  Google Scholar 

  • Dühring MB, Jensen JS, Sigmund O (2008) Acoustic design by topology optimization. J Sound Vib 317(3-5):557–575

    Article  Google Scholar 

  • Farhat C, Roux F-X (1991) A method of finite element tearing and interconnecting and its parallel solution algorithm. Int J Numer Methods Eng 32(6):1205–1227

    Article  MathSciNet  MATH  Google Scholar 

  • Farhat C, Lesoinnea M, Letallecb P (1998) Load and motion transfer algorithms for fluid / structure interaction problems with non-matching discrete interfaces: momentum and energy conservation, optimal discretization and application to aeroelasticity. Comput Methods Appl Mech Engrg 7825(97)

  • Gerbeau J-F, Vidrascu M (2003) A quasi-Newton algorithm based on a reduced model for fluid-structure interaction problems in blood flows. Research Report RR-4691

  • Gerbeau J-F, Vidrascu M, Frey P (2005) Fluid–structure interaction in blood flows on geometries based on medical imaging. Comput Struct 83(2–3):155–165

    Article  Google Scholar 

  • Gersborg-Hansen A, Haber RB, Sigmund O (2005) Topology optimization of channel flow problems. Struct Multidiscip Optim 30(3):181–192

    Article  MathSciNet  MATH  Google Scholar 

  • Hughes T JR, Franca LP, Balestra M (1986) A new finite element formulation for computational fluid dynamics V. Circumventing the babuška-brezzi condition: a stable petrov-galerkin formulation of the stokes problem accommodating equal-order interpolations. Comput Methods Appl Mech Eng 59(1):85–99

    Article  MATH  Google Scholar 

  • Jenkins N, Maute K (2015) Level set topology optimization of stationary fluid-structure interaction problems. Struct Multidiscip Optim, 1–17. Struct Multidiscip Optim 52(1):179–195

    Article  MathSciNet  Google Scholar 

  • Jenkins N, Maute K (2016) An immersed boundary approach for shape and topology optimization of stationary fluid-structure interaction problems. Struct Multidiscip Optim

  • Jensen JS, Sigmund O (2011) Topology optimization for nano-photonics. Laser Photon Rev 5(2):308–321

    Article  Google Scholar 

  • Kolaei A, Rakheja S, Richard MJ (2016) An efficient methodology for simulating roll dynamics of a tank vehicle coupled with transient fluid slosh. J Vibr Control

  • Kreissl S, Pingen G, Evgrafov A, Maute K (2010) Topology optimization of flexible micro-fluidic devices. Struct Multidiscip Optim 42(4):495–516

    Article  Google Scholar 

  • Kreissl S, Pingen G, Maute K (2011) Topology optimization for unsteady flow. Int J Numer Methods Eng

  • Liu S, Li Q, Chen W, Tong L, Cheng G (2015) An identification method for enclosed voids restriction in manufacturability design for additive manufacturing structures. Front Mech Eng 10(2):126– 137

    Article  Google Scholar 

  • Michaleris P, Vidal CA (1994) Tangent operators and design sensitivity formulations for transient non-linear coupled problems with applications to elastoplasticity. Int J Numer Methods Eng 37(1993):2471–2499

    Article  MATH  Google Scholar 

  • Nørgaard S, Sigmund O, Lazarov B (2016) Topology optimization of unsteady flow problems using the lattice Boltzmann method. J Comput Phys 307:291–307

    Article  MathSciNet  MATH  Google Scholar 

  • Okkels F, Bruus H (2007) Scaling behavior of optimally structured catalytic microfluidic reactors. Phys Rev E 75:016301

    Article  Google Scholar 

  • Pedersen CBW, Buhl T (2001) Topology synthesis of large-displacement compliant mechanisms. Int J Numer Meth Engng, 1–23

  • Picelli R, Vicente WM, Pavanello R (2015) Bi-directional evolutionary structural optimization for design-dependent fluid pressure loading problems. Eng Optim, 1–19

  • Picelli R, Vicente WM, Pavanello R (2017) Evolutionary topology optimization for structural compliance minimization considering design-dependent FSI loads. Fin Elem Anal Des 135:44–55. ISSN 0168874X

    Article  Google Scholar 

  • Schevenels M, Lazarov BS, Sigmund O (2011) Robust topology optimization accounting for spatially varying manufacturing errors. Comput Methods Appl Mech Eng 200(49):3613–3627

    Article  MATH  Google Scholar 

  • Shangguan W-B, Zhen-Hua L (2004) Experimental study and simulation of a hydraulic engine mount with fully coupled fluid–structure interaction finite element analysis model. Comput Struct 82(22):1751–1771

    Article  Google Scholar 

  • Sigmund O (1997) On the design of compliant mechanisms using topology optimization*. Mech Struct Mach 25(4):493–524

    Article  MathSciNet  Google Scholar 

  • Sigmund O (2009) Manufacturing tolerant topology optimization. Acta Mechanica Sinica/Lixue Xuebao 25(2):227–239

    Article  MATH  Google Scholar 

  • Sigmund O, Maute K (2013) Topology optimization approaches. Struct Multidiscip Optim 48(6):1031–1055

    Article  MathSciNet  Google Scholar 

  • Svanberg K (2006) The method of moving asymptotes - a new method for structural optimization. Struct Multidiscip Optim, 1–15

  • Tezduyar TE (1991) Stabilized finite element formulations for incompressible flow computations. Adv Appl Mech 28:1–44

    Article  MathSciNet  MATH  Google Scholar 

  • Tezduyar TE, Sathe S, Schwaab M, Conklin BS (2008) Arterial fluid mechanics modeling with the stabilized space – time fluid – structure interaction technique. Int J Numer Methods Fluids, 601–629

  • Vicente WM, Picelli R, Pavanello R, Xie YM (2015) Topology optimization of frequency responses of fluid–structure interaction systems. Finite Elem Anal Des 98(C):1–13. ISSN 0168874X

    Article  Google Scholar 

  • Wang F, Jensen JS, Sigmund O (2011a) Robust topology optimization of photonic crystal waveguides with tailored dispersion properties. J Opt Soc Amer B 28(3):387

    Article  Google Scholar 

  • Wang F, Lazarov BS, Sigmund O (2011b) On projection methods, convergence and robust formulations in topology optimization. Struct Multidiscip Optim 43(6):767–784

    Article  MATH  Google Scholar 

  • White FM, Corfield I (1991) Viscous fluid flow, vol 2. McGraw-Hill, New York

    Google Scholar 

  • Wu S, Wang Z (2014) A numerical simulation of fluid-structure interaction for refrigerator compressors suction and exhaust system performance analysis. In: International compressor engineering conference, pp 1–7

  • Yoon GH (2010) Topology optimization for stationary fluid–structure interaction problems using a new monolithic formulation. Int J Numer Methods Eng 82(5):591–616

    Article  MATH  Google Scholar 

  • Yoon GH (2014a) Compliant topology optimization for planar passive flap micro valve. J Nanosci Nanotechnol 14(10):7585–7591

    Article  Google Scholar 

  • Yoon GH (2014b) Stress-based topology optimization method for steady-state fluid structure interaction problems. Comput Methods Appl Mech Engrg 278:499–523

    Article  MathSciNet  Google Scholar 

  • Yoon GH, Sigmund O (2008) A monolithic approach for topology optimization of electrostatically actuated devices. Comput Methods Appl Mech Eng 197(45–48):4062–4075

    Article  MathSciNet  MATH  Google Scholar 

  • Yoon GH, Jensen JS, Sigmund O (2007) Topology optimization of acoustic–structure interaction problems using a mixed finite element formulation. Int J Numer Methods Eng 70(9):1049–1075

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors acknowledge the financial support received from the TopTen project sponsored by the Danish Council for Independent Research (DFF-4005-00320) and from the National Natural Science Foundation of China (Grant No. 51705311).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian Lundgaard.

Additional information

Responsible Editor: Hyunsun Alicia Kim

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 A.1 Details on the derivation of (3)

The Navier-Cauchy equations are given by:

$$\begin{array}{@{}rcl@{}} \frac{\partial{\sigma^{s}_{ij}}}{\partial{x_{j}}} + f_{i} &=& 0 \quad \text{in} \quad {{\Omega}_{S}} \\ \sigma^{s}_{ij} &=& C_{ijkl} \epsilon_{kl} \\ \epsilon^{s}_{kl} &= &\frac{1}{2} \left( \frac{\partial{d_{k}}}{\partial{x_{l}}} + \frac{\partial{d_{l}}}{\partial{x_{k}}} \right) \end{array} $$
(24)

The coupling between the fluid and the structure is given by (Farhat and Roux 1991; Yoon 2014a)

$$ \sigma^{s}_{ij} n_{j} = \sigma^{f}_{ij} n_{j} \quad \text{on} \quad {{\Gamma}_{SF}} $$
(25)

The weak form of (24) is given by:

$$ {\int}_{{{\Omega}_{S}}} {w^{h}_{i}} \frac{\partial{\sigma_{ij}^{s}}}{\partial{x_{j}}} \hspace{2pt}\text{d} V + {\int}_{{{\Omega}_{S}}} {w^{h}_{i}} f_{i} \hspace{2pt}\text{d} V = 0 $$
(26)

where \({w^{h}_{i}}\) is a suitable basis function. Integration by parts of higher dimensions on the first term of (26), yields:

$$\begin{array}{@{}rcl@{}} {\int}_{{{\Gamma}_{SF}}} {w^{h}_{i}} \sigma^{s}_{ij} n_{j} \hspace{2pt}\text{d} S - {\int}_{{{\Omega}_{S}}} \frac{\partial{{w^{h}_{i}}}}{\partial{x_{j}}} \sigma_{ij}^{s} \hspace{2pt}\text{d} V \\ + {\int}_{{{\Omega}_{S}}} {w^{h}_{i}} f_{i} \hspace{2pt}\text{d} V = 0 \end{array} $$
(27)

Shear stresses on the interface between the fluid and the structure are neglected for which reason (25) can be written as \(\sigma ^{s}_{ij} n_{j} = -p n_{i}\) on ΓSF, where p is the pressure on the interface surface. Equation (27) is now rewritten as:

$$\begin{array}{@{}rcl@{}} {\int}_{{{\Gamma}_{SF}}} {w^{h}_{i}} p^{h} n_{i} \hspace{2pt}\text{d} S - {\int}_{{{\Omega}_{S}}} \frac{\partial{{w^{h}_{i}}}}{\partial{x_{j}}} \sigma_{ij}^{s} \hspace{2pt}\text{d} V \\ + {\int}_{{{\Omega}_{S}}} {w^{h}_{i}} f_{i} \hspace{2pt}\text{d} V = 0 \end{array} $$
(28)

Integration by parts of higher dimensions on the first term of (28), yields:

$$\begin{array}{@{}rcl@{}} {\int}_{{{\Omega}_{S}}} \frac{\partial{{w^{h}_{i}}}}{\partial{x_{i}}} p^{h} \hspace{2pt}\text{d} V + {\int}_{{{\Omega}_{S}}} {w^{h}_{i}} \frac{\partial{p^{h}}}{\partial{x_{i}}} \hspace{2pt}\text{d} V \\ - {\int}_{{{\Omega}_{S}}} \frac{\partial{{w^{h}_{i}}}}{\partial{x_{j}}} \sigma_{ij}^{s} \hspace{2pt}\text{d} V + {\int}_{{{\Omega}_{S}}} {w^{h}_{i}} f_{i} \hspace{2pt}\text{d} V = 0 \end{array} $$
(29)

Equation (29) may now be rewritten from the segregated domains ΩS to a unified domain Ω, by introducing a design variable field 0 ≤ ρ ≤ 1 and the following interpolation function:

$$ C_{ijkl} = E(\rho) C^{0}_{ijkl} $$
(30)

Correct integration of the fluid pressure on the elastic structure is ensured by introducing the filter function Ψ(ρ):

$$ {\int}_{{\Omega}_{S}} \square \hspace{2pt}\text{d} V = {\int}_{{\Omega}} {\Psi}(\rho) \square \hspace{2pt}\text{d} V $$
(31)

Ψ(ρ) is a function which is unity for ρ = 1 and zero for ρ = 0. Inserting (30) and (31) into (29), we arrive at the following expression for the Navier-Cauchy equation defined in a unified domain Ω:

$$\begin{array}{@{}rcl@{}} {\int}_{{{\Omega}}} {\Psi}(\rho) \left( \frac{\partial{{w^{h}_{i}}}}{\partial{x_{i}}} p^{h} + {w^{h}_{i}} \frac{\partial{p^{h}}}{\partial{x_{i}}} \right) \hspace{2pt}\text{d} V \\ + {\int}_{{{\Omega}}} {w^{h}_{i}} f_{i} \hspace{2pt}\text{d} V = {\int}_{{{\Omega}}} E(\rho) \frac{\partial{{w^{h}_{i}}}}{\partial{x_{j}}} \sigma_{ij}^{s_{0}} \hspace{2pt}\text{d} V \end{array} $$
(32)

1.2 A.2 Benchmark examples

To demonstrate the features of the present framework, we have revisited the wall flow design problems presented in the works of Yoon (2010) and Picelli et al. (2017) and Jenkins and Maute (2016). These, what we call, benchmark designs problems are solved with the same physical parameters as in the respective papers but with our framework. The design solution, obtained with the framework presented in this study, to the design problem presented in Jenkins and Maute (2016) has been plotted in Fig. 23. The design solution shown in Jenkins and Maute (2016) and our design solution in Fig. 23 are by visual comparison quite similar. The small difference in the design solutions suggests that the internal pressure, the displacement dependency and /or the shear stress may have minor effects for this specific optimization problem. However, more studies and other optimization problems are required to fully understand the influence of the different modeling approaches and assumptions.

Fig. 23
figure 23

Design solution for the wall flow problem in (Jenkins and Maute 2016) with Re = 10

In Figs. 24 and 25, the Yoon (2010) and Picelli et al. (2017) benchmark design problems, solved by our framework, have been plotted for Re = 0.004 and Re = 12. The designs have been optimized for the full pressure coupling formulation in (3). It is unclear whether the design solutions in Yoon (2010) are solved for pressure-coupling term 1 or pressure-coupling term 1 and 2. In Yoon (2010), it is seen that an increased Reynolds number causes the wall support to move to the downstream part of the design domain. This tendency is conflicting with the tendencies observed in Picelli et al. (2017) and in Figs. 2425. In these design problems, it is observed that an increased Reynolds number causes the wall support to move to the upstream part of the design domain. As far as we are aware there does not exist any crosschecks between the Reynolds number and the optimized designs in the mentioned papers, for which reason it is challenging to assess how much significance we can attribute to the features of the optimized designs in Yoon (2010) and Picelli et al. (2017). However, please notice that our optimized designs in Figs. 2425 pass a crosscheck.

Fig. 24
figure 24

Design solutions for the wall flow problem stated in (Yoon 2010; Picelli et al. 2017) with Re = 0.004

Fig. 25
figure 25

Design solutions for the wall flow problem in Yoon (2010) and Picelli et al. (2017) with Re = 12

1.3 A.3 Details on the determination of interpolation functions

TO for FSI problems are highly non-linear, ill-posed and non-convex. Several model parameters influences the design processes and the design solutions, which require a significant amount of parameter tuning due to the deepness of the design space. Numerical experiments with the framework presented in this work, have suggested that the design process is highly dependent on the choice of interpolation functions, α(ρ),E(ρ) and Ψ(ρ) (abbreviated: H{α, E,Ψ}), and the choice of interpolation function parameters, pα, pE and pΨ (abbreviated: p{α, E,Ψ}). It is our experience that adequate choices of p{α, E,Ψ} and H{α, E,Ψ} are critical to obtain well-performing and 0/1 design solutions. As p{α, E,Ψ} and H{α, E,Ψ} are key to carry out successful optimization problems, we will in this section present a methodology which can be used to determine adequate p{α, E,Ψ} and H{α, E,Ψ} and hereby formulate well-posed optimization problems.

IMwEBR are based on shape sensitivities which may be better suited for TO for some multi-physics problems. IMwEBR have a well-defined boundary between the different types of physics, which ensures that the physics are resolved correctly throughout the entire design process, as no sub-domains are dependent on the quality of the interpolation functions of the intermediate design variables. In density-based methods the correctness of the physical modeling rely, among many other aspects, on the choice of p{α, E,Ψ} and H{α, E,Ψ}. The hypothesis is that topology sensitivities, f/ρd, may obtain the same well-behaving features as shape sensitivities, f/ρs, if some adequateH{α, E,Ψ} and p{α, E,Ψ} are chosen.

To determine adequate sets of H{α, E,Ψ} and p{α, E,Ψ}, we compare f/ρs and f/ρd for two different problems: (1) a purely elastic problem and (2) an FSI problem. The problem layouts and the boundary conditions have been sketched in Fig. 26. To carry out the study, we compare four different objective functions, fE, fC, fT and fP (abbreviated: f{E, C, T, P}):

  1. 1.

    Dissipated energy in the flow, fE, see (17).

  2. 2.

    Structural compliance fC, see (16)

  3. 3.

    The y-displacement of the tip of the beam in point {x, y} = {2, 0.45}:

    $$ f_{T} = \frac{1}{{\int}_{{{\Omega}_{T}}} \hspace{2pt}\text{d} V} {\int}_{{{\Omega}_{T}}} d_{y} \hspace{2pt}\text{d} V $$
    (33)
  4. 4.

    The pressure induced y-directional force on the beam:

    $$ f_{P} = {\int}_{{\Gamma}_{B}} p n_{j} \hspace{2pt}\text{d} S $$
    (34)
Fig. 26
figure 26

Problem layouts and boundary conditions used to compare the topology sensitivities and the shape sensitivities

The relationship between f{E, C, T, P}, ρS and ρD is determined with basis in a simple problem where a beam separates a channel into two regions of the same size. The problem is modeled in the unified framework. The beam separating the channel, ΩI, has fixed unity design variables and the tip of the beam is loaded with a force fy. With reference to Fig. 26, we consider two different problems: (a) An elastic problem where \(u_{x}^{*}= 0\) and fy = 10, and (b) an FSI problem where \(u_{x}^{*}= 1\) and fy = 0.

With reference to Fig. 26a, the topology sensitivities of various objective functions are computed by changing the design variables of the lowest line of elements of the vertical beam. With reference to Fig. 26b, the shape gradients of various objective functions are computed by changing the position of the nodes on the lower boundary of the beam. The position of the boundary is varied over the length of one element, entailing that f/ρd and f/ρs are comparable in material usage.

The relationship between f{C, P}, f{C, P}, ρd and ρs for the elastic problem have been compared in Fig. 27. The relationships are different but can be characterized by the following attribute: f{C, T}(ρs) and f{C, T}(ρd) are strictly monotonic entailing that f{C, T}/ρs and f{C, T}/ρd are strictly monotonic. Well-versed and crisp 0/1 designs and smooth optimization processes are obtained for a large number of TO for elastic problems, see e.g. (Bendsøe and Sigmund 2003). The hypothesis in this study is, that the well-posed properties of linear elastic problems are explained by the strictly monotonic features of the fC/ρd.

Fig. 27
figure 27

The relationship between fC, fP and ρd/ρs for various choices of H{α, E,Ψ} and p{α, E,Ψ}

We now point out attention to the FSI problem, where we investigate the monotonicity of f{E, C, T, P}(ρs) and f{E, C, T, P}(ρd). The relationships have been plotted in Fig. 28, and are characterized by:

  1. 1.

    f{E, C, T, P}(ρs) are strictly monotonic in all cases, which entail that f{E, C, T, P}/ρs are strictly monotonic in all cases.

  2. 2.

    f{E, C, T, P}(ρd) are strictly monotonic for some choices of p{α, E,Ψ} and H{α, E,Ψ}.

  3. 3.

    The relationship between f{E, C, T, P}(ρd) seem to be very sensitive with respect to the choice of pα, as a small change in pα may disrupts the monotonicity for all f{E, C, T, P}.

Fig. 28
figure 28

The relationship between fE, fC, fP, fT and ρd/ρs for various choices of H{α, E,Ψ} and p{α, E,Ψ}

To demonstrate the importance of the monotonicity of f{E, C, T, P}(ρ) we have included a numerical example where we compare two design optimized for two different pα. In Fig. 29, fD have been optimized for the flow obstacle problem (see Section 6.2) for pα = 0.5 ⋅ 10−6 and pα = 10−6. The design optimized with a strictly monotonic fD perform much better than the design optimized for non-monotonic f{D}. We notice that a small change in pα significantly influences the topology and the performance of the design solutions.

Fig. 29
figure 29

Optimized designs for the flow obstacle problem in Fig. 12 for two different interpolation function parameters

We suggest that the correlation between the monotonicity of the objective functions and the well-posedness of the design problems is likely to generalize to all multiphysics topology optimization problems.

1.4 A.4 Details on the Brinkman penalization parameter

Numerical experiments with the framework suggested that αmax should be chosen high (e.g. αmax = 109) to model the pressure field correctly. A low αmax, e.g. αmax = 105, provides a more well-posed optimization problem, however the pressure field is not modeled correctly. Modeling the pressure field incorrectly may provide unintuitive and physically meaningless optimized designs, as the coupling from the fluid to the structure is transfered through the pressure field. To demonstrate the relationship between the pressure field and the magnitude of αmax, we have plotted the pressure field along the line {x, y} = {x, 0.34} for the design shown in Fig. 30 in Fig. 31. The pressure fields for the COMSOL composite model and the unified model with αmax = 109 is closely correlated. However, for small αmax, a large difference between the composite model and the unified model is observed. As a final remark, the minor difference between the fields is caused by different finite element discretizations of the segregated and unified models.

Fig. 30
figure 30

Design used to evaluate the pressure

Fig. 31
figure 31

The pressure as function of x for various αmax

To demonstrate the occurrence of unintuitive optimized topologies for design problems with too low αmax, we consider two different design problems. The design in Fig. 32a has been optimized for αmax = 105 and the design in Fig. 32b has been optimized for αmax = 109. The designs provide superior performance for the model parameters under which the designs were optimized. However, the design optimizd for αmax = 105 does not performing well in a segregated FSI formulation. The too low αmax causes poor resolvement of the pressure field which causes unintuitive optimized designs. For comparison we have plotted the design optimized for αmax = 109 in Fig. 32b. This design performs well in a segregated model.

Fig. 32
figure 32

Design solutions for the flow obstacle problem in Fig. 12 for various αmax

The coupling between the structure and the fluid is carried out through the pressure field, for which reason adequate modeling of the pressure field is crucial in FSI problems. Previous work on topology optimization for fluid problems has used magnitudes of αmax which do not resolve the pressure field correctly. Non-intuitive designs may not have been observed in these studies because the pressure fields were not directly related to the objective functions or the constraints of the optimization problems.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lundgaard, C., Alexandersen, J., Zhou, M. et al. Revisiting density-based topology optimization for fluid-structure-interaction problems. Struct Multidisc Optim 58, 969–995 (2018). https://doi.org/10.1007/s00158-018-1940-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-018-1940-4

Keywords

Navigation