Skip to main content
Log in

Saturated poroelastic actuators generated by topology optimization

  • Research Paper
  • Published:
Structural and Multidisciplinary Optimization Aims and scope Submit manuscript

Abstract

In this paper the fluid-structure interaction problem of a saturated porous media is considered. The pressure coupling properties of porous saturated materials change with the microstructure and this is utilized in the design of an actuator using a topology optimized porous material. By maximizing the coupling of internal fluid pressure and elastic shear stresses a slab of the optimized porous material deflects/deforms when a pressure is imposed and an actuator is created. Several phenomenologically based constraints are imposed in order to get a stable force transmitting actuator.

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

Similar content being viewed by others

References

  • Auriault JL, Boutin C, Geindreau C (2009) Homogenization of coupled phenomena in heterogeneous media. ISTE Ltd and Wiley

  • Auriault JL, Sanchez-Palencia E (1977) Study of macroscopic behavior of a deformable saturated porous-medium. J Méc 16(4):575–603

    MathSciNet  MATH  Google Scholar 

  • Bensoussan A, Lions J, Papanicolaou G (1978) Asymptotic analysis for periodic structures. North-Holland, Amsterdam

    MATH  Google Scholar 

  • Biot MA (1941) General theory of three-dimensional consolidation. J Appl Phys 12(2):155–164

    Article  Google Scholar 

  • Biot MA (1955) Theory of elasticity and consolidation for a porous anisotropic solid. J Appl Phys 26(2):182–185

    Article  MathSciNet  MATH  Google Scholar 

  • Biot MA (1956) Theory of propagation of elastic waves in a fluid-saturated porous solid. I. Low-frequency range. J Acoust Soc Am 28(2):168–178

    Article  MathSciNet  Google Scholar 

  • 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 

  • Bourdin B (2001) Filters in topology optimization. Int J Numer Methods Eng 50:2143–2158

    Article  MathSciNet  MATH  Google Scholar 

  • Bruns T, Tortorelli D (2001) Topology optimization of non-linear elastic structures and compliant mechanisms. Comput Methods Appl Mech Eng 190:3443–3459

    Article  MATH  Google Scholar 

  • de Boer R, Ehlers W (1988) A historical review of the formulation of porous media theories. Acta Mech 74(1):1–8

    Article  Google Scholar 

  • Fillunger P (1913) Der auftrieb in talsperren. Österr. Wochenzeitschrift für den öffentl. Baudienst., vols I, II, III, pp 532–552, 522–556, 567–570

  • Guest JK, Prévost JH (2006) Optimizing multifunctional materials: design of microstructures for maximized stiffness and fluid permeability. Int J Solids Struct 43(22–23):7028–7047

    Article  MATH  Google Scholar 

  • Guest JK, Prévost JH (2007) Design of maximum permeability material structures. Comput Methods Appl Mech Eng 196(4–6):1006–1017

    Article  MATH  Google Scholar 

  • Hughes TJR, Franca LP (1987) A new finite element formulation for computational fluid dynamics: VII. The Stokes problem with various well-posed boundary conditions: symmetric formulations that converge for all velocity/pressure spaces. Comput Methods Appl Mech Eng 65(1):85–96

    Article  MathSciNet  MATH  Google Scholar 

  • Jones R (1999) Mechanics of composite materials, 2nd edn. Taylor & Francis, New York

    Google Scholar 

  • Kreissl S, Pingen G, Evgrafov A, Maute K (2010) Topology optimization of flexible micro-fluidic devices. Struct Multidiscipl Optim 42:495–516. doi:10.1007/s00158-010-0526-6

    Article  Google Scholar 

  • Lydzba D, Shao JF (2000) Study of poroelasticity material coefficients as response of microstructure. Mech Cohes-Frict Mater 5(2):149–171

    Article  Google Scholar 

  • Maute K, Allen M (2004) Conceptual design of aeroelastic structures by topology optimization. Struct Multidiscipl Optim 27:27–42. doi:10.1007/s00158-003-0362-z

    Article  Google Scholar 

  • Sanchez-Palencia E (1980) Non-homogeneous media and vibration theory. In: Lecture notes in physics, vol 127. Springer, Berlin

    Google Scholar 

  • Schanz M, Diebels S (2003) A comparative study of Biot’s theory and the linear theory of porous media for wave propagation problems. Acta Mech 161(3):213–235

    MATH  Google Scholar 

  • Sigmund O (1994) Materials with prescribed constitutive parameters: an inverse homogenization problem. Int J Solids Struct 31(17):2313–2329

    Article  MathSciNet  MATH  Google Scholar 

  • Sigmund O (1999) On the optimality of bone microstructure. In: Pedersen P, Bendsøe M (eds) Synthesis in bio solid mechanics. Kluwer, IUTAM, pp 221–234

  • Sigmund O (2000) A new class of extremal composites. J Mech Phys Solids 48(2):397–428

    Article  MathSciNet  MATH  Google Scholar 

  • Sigmund O, Torquato S (1996) Composites with extremal thermal expansion coefficients. Appl Phys Lett 69(21):3203–3205

    Article  Google Scholar 

  • Sigmund O, Torquato S (1997) Design of materials with extreme thermal expansion using a three-phase topology optimization method. J Mech Phys Solids 45(6):1037–1067

    Article  MathSciNet  Google Scholar 

  • Sigmund O, Torquato S, Aksay IA (1998) On the design of 1–3 piezocomposites using topology optimization. J Mater Res 13(4):1038–1048

    Article  Google Scholar 

  • Stolpe M, Svanberg K (2001) An alternative interpolation scheme for minimum compliance topology optimization. Struct Multidisc Optim 22(2):116–124

    Article  Google Scholar 

  • Svanberg K (1987) The method of moving asymptotes—a new method for structural optimization. Int J Numer Methods Eng 24(2):359–373

    Article  MathSciNet  MATH  Google Scholar 

  • Torquato S, Hyun S, Donev A (2002) Multifunctional composites: optimizing microstructures for simultaneous transport of heat and electricity. Phys Rev Lett 89(26):266601

    Article  Google Scholar 

  • von Terzaghi K (1923) Die Berechnung der Durchlässigkeit des Tones aus dem Verlauf der hydromechanischen Spannungserscheinungen. Sitzungsber Akad Wiss Wien, Math-Naturwiss Kl 132(3–4):125–138

    Google Scholar 

  • Xu S, Cheng G (2010) Optimum material design of minimum structural compliance under seepage constraint. Struct Multidisc Optim 41(4):575–587

    Article  MathSciNet  Google Scholar 

  • 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

    MATH  Google Scholar 

Download references

Acknowledgements

This research was conducted within the DCAMM Research School through a grant from the Danish Agency for Science, Technology and Innovation. The authors would like to thank the TopOpt research group (www.topopt.dtu.dk) for fruitful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Casper Schousboe Andreasen.

Appendix A: Adjoint sensitivity analysis

Appendix A: Adjoint sensitivity analysis

The Lagrangian \(\mathcal{L}\) is formed by adding the adjoint variable multiplied by the residual (= 0) to the objective function, Φ:

$$ \mathcal{L}=\Phi(\mathbf{u},\rho)+{\boldsymbol \lambda}^TR(\mathbf{u},\rho) $$
(33)

The derivative of the Lagrangian is obtained using the chain rule

$$ \begin{array}{rll} \frac{{\rm d} \mathcal{L}}{{\rm d} \rho}&=&\frac{\partial \mathcal{L}}{\partial \rho}+\frac{\partial \mathcal{L}}{\partial \mathbf{u}}\frac{{\rm d} \mathbf{u}}{{\rm d} \rho} \\ &=&\frac{\partial \Phi(\mathbf{u}, \rho)}{\partial \rho} +\frac{\partial \Phi(\mathbf{u}, \rho)}{\partial \mathbf{u}}\frac{{\rm d} \mathbf{u}}{{\rm d} \rho} \\ &&+{\boldsymbol \lambda}^T\frac{\partial R(\mathbf{u},\rho)}{\partial \rho} +{\boldsymbol \lambda}^T\frac{\partial R(\mathbf{u},\rho)}{\partial \mathbf{u}}\frac{{\rm d} \mathbf{u}}{{\rm d} \rho} \end{array} $$
(34)

rearranging and splitting yields: Find \({\boldsymbol \lambda}\) such that

$${\boldsymbol \lambda}^T\frac{\partial R(\mathbf{u}, \rho)}{\partial \mathbf{u}}\frac{{\rm d} \mathbf{u}}{{\rm d} \rho} +\frac{\partial \Phi(\mathbf{u}, \rho)}{\partial \mathbf{u}}\frac{{\rm d} \mathbf{u}}{{\rm d} \rho}=0 $$
(35)

and the sensitivities can be evaluated by computing

$$ \frac{{\rm d} \mathcal{L}}{{\rm d} \rho}=\frac{\partial \Phi(\mathbf{u}, \rho)}{\partial \rho}+{\boldsymbol \lambda}^T\frac{\partial R(\mathbf{u},\rho)}{\partial \rho} $$
(36)

This leads to the following adjoint problem for the objective function:

Find for all such that

$$ \int\! E_{ ijlm}\varepsilon_{ij}({\boldsymbol \lambda})\varepsilon_{lm}(\hat{{\boldsymbol \lambda}}){\rm d} \Omega \!=\!-\!\int\!\frac{1}{2}(\!-\!\delta_{ij}) E_{ijlm}\varepsilon_{lm} (\hat{{\boldsymbol \lambda}}){\rm d} \Omega $$
(37)

where the solution is used for the evaluation of the sensitivities

$$ \begin{array}{rll} \frac{{\rm d} \Phi}{{\rm d} \rho}&=& \int \bigg[\frac{1}{2}(1-\delta_{ij})E_{ijlm}'\varepsilon_{lm}(\mathbf{u})+\delta_{ij}\varepsilon_{ij}({\boldsymbol \lambda})\frac{E'}{E_0} \\ && \qquad -\,\varepsilon_{ij}({\boldsymbol \lambda})E_{ijlm}\varepsilon_{lm}(\mathbf{u})\bigg]{\rm d} \Omega \end{array} $$
(38)

1.1 A.1 Stiffness constraints

The stiffness constraints are related to the problem above as the equation system is the same except for the loading that changes according to the direction considered. The stiffness tensor entries are computed as

$$ E_{ijkh}=E_{ijkh}-E_{ijlm}\varepsilon_{lm}(\mathbf{u}^{kh})$$
(39)

and the sensitivities of the entries can be derived using the adjoint method.

$$ \begin{array}{rll} R_{ijkh}&=&\int_\Omega \varepsilon_{pq}({\boldsymbol \lambda}^{ij}) E_{pqkh}{\rm d} \Omega \\ &&-\int_\Omega \varepsilon_{pq}( {\boldsymbol \lambda}^{ij}) E_{pqlm} \varepsilon_{lm}(\mathbf{u}^{kh}) {\rm d} \Omega \\ \frac{\partial R_{ijkh}}{\partial \rho}&=&\int_\Omega \varepsilon_{pq}({\boldsymbol \lambda}^{ij}) E_{pqkh}'{\rm d} \Omega \\ && -\int_\Omega \varepsilon_{pq}( {\boldsymbol \lambda}^{ij}) E_{pqlm}' \varepsilon_{lm}(\mathbf{u}^{kh}) {\rm d} \Omega \\ \frac{\partial E_{ijkh}}{\partial \rho}&=&\int_\Omega E_{ijkh}'-E_{ijlm}'\varepsilon_{lm}(\mathbf{u}^{kh}) {\rm d} \Omega \\ \frac{\partial R_{ijkh}}{\partial \mathbf{u}}\frac{{\rm d} \mathbf{u}}{{\rm d} \rho} &=&-\int_\Omega \varepsilon_{pq}({\boldsymbol \lambda}^{ij}) E_{pqlm} \varepsilon_{lm}'(\mathbf{u}^{kh}) {\rm d} \Omega \\ \frac{\partial E_{ijkh}}{\partial \mathbf{u}}\frac{{\rm d} \mathbf{u}}{{\rm d} \rho} &=&- \int_\Omega E_{ijlm} \varepsilon_{lm}'(\mathbf{u}^{kh}) {\rm d} \Omega \end{array} $$

which yields the following adjoint problem

$$ -\!\int_\Omega \varepsilon_{pq}({\boldsymbol \lambda}^{ij}) E_{pqlm} \varepsilon_{lm}'(\mathbf{u}^{kh}) {\rm d} \Omega \!-\! \int_\Omega E_{ijlm} \varepsilon_{lm}'(\mathbf{u}^{kh}) {\rm d} \Omega\! =\! 0 $$
(40)
$$ \Rightarrow\! \int_\Omega \varepsilon_{pq}({\boldsymbol \lambda}^{ij}) E_{pqlm} \varepsilon_{lm}(\hat{{\boldsymbol \lambda}}) {\rm d} \Omega \!=\!-\!\int_\Omega E_{ijlm}\varepsilon_{lm}(\hat{{\boldsymbol \lambda}}) {\rm d} \Omega $$
(41)

which due to the symmetry of E is equivalent to the original problem but with opposing sign and hence the sensitivities can be computed from the original solution as

$$ \begin{array}{rll} \frac{{\rm d} E_{ijkh}}{{\rm d} \rho}&=&\int_{\Omega}\biggl[E_{ijkh}'-E_{ijlm}'\varepsilon_{lm}(\mathbf{u}^{kh}) \\ && \qquad +\,\varepsilon_{pq}(\mathbf{u}^{ij})E_{pqlm}'\varepsilon_{lm}(\mathbf{u}^{kh}) \\ && \qquad -\,E_{pqkh}'\varepsilon_{pq}(\mathbf{u}^{ij})\biggl]{\rm d} \Omega \end{array} $$
(42)

1.2 A.2 Permeability constraints

For each constraint based on the normal permeability (i = k),

$$ K_{ik}=\frac{1}{|\Omega|}\int_\Omega \emph{v}_i^k {\rm d} \Omega $$
(43)

the sensitivity needs to be computed. The derivation is as follows. Forming the residual

$$ \begin{array}{rll} R(\hat{\mathbf{v}},\mathbf{v}^k,\rho) &= &-\int_\Omega \hat{\emph{v}}_{i,j}{\emph{v}}_{i,j}^k {\rm d}\Omega - \int_\Omega\zeta(\rho){\emph{v}}_i^k\hat{\emph{v}}_{i} {\rm d}\Omega \\ && +\, \int_\Omega\hat{\emph{v}}_{i,i}p {\rm d}\Omega+ \int_\Omega \hat{\emph{v}}_i\delta_{ik}{\rm d}\Omega - \int_\Omega \hat{p}{\emph{v}}_{i,i}^k{\rm d}\Omega \end{array} $$
(44)

Then forming the derivative of the Lagrangian

$$ \begin{array}{rll} \frac{{\rm d} \mathcal{L}}{{\rm d} \rho}&=&\frac{\partial K(\mathbf{v})}{\partial \mathbf{v}}\frac{{\rm d} \mathbf{v}}{{\rm d} \rho}+\frac{\partial R({\boldsymbol \lambda},\mathbf{v},\rho)}{\partial \rho} \\ &&+\frac{\partial R({\boldsymbol \lambda},\mathbf{v},\rho)}{\partial \mathbf{v}}\frac{{\rm d} \mathbf{v}}{{\rm d} \rho} +\frac{\partial R({\boldsymbol \lambda},\mathbf{v},\rho)}{\partial p}\frac{{\rm d} p}{{\rm d} \rho} \end{array} $$
(45)

which leads to the following adjoint problem

Find and \(p^k\in\mathcal{Q}\) such that and \(\hat{p}\in \mathcal{Q}\)

$$ \begin{array}{rll} -\int_\Omega \lambda^k_{i,j}\hat{\lambda}_{i,j} {\rm d}\Omega &-& \int_\Omega\zeta(\rho)\hat{\lambda}_i\lambda^k_{i} {\rm d}\Omega - \int_\Omega\lambda^k_{i,i}p {\rm d}\Omega \\ &+& \,\int_\Omega \lambda^k_i{\rm d}\Omega + \int_\Omega p\hat{\lambda}_{i,i}{\rm d}\Omega = 0 \end{array} $$
(46)

which is seen to be equivalent to the state problem when i = k and \(\lambda_i = \emph{v}_i\). Hence the sensitivities can be computed by evaluating

$$ \frac{{\rm d} K_{ik}(\mathbf{v})}{{\rm d} \rho}=-\frac{{\rm d} \zeta(\rho)}{{\rm d} \rho}\emph{v}_i^k\emph{v}_i^k $$
(47)

but only for i = k.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Andreasen, C.S., Sigmund, O. Saturated poroelastic actuators generated by topology optimization. Struct Multidisc Optim 43, 693–706 (2011). https://doi.org/10.1007/s00158-010-0597-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-010-0597-4

Keywords

Navigation