Skip to main content
Log in

Comparison of approximate shape gradients

  • Published:
BIT Numerical Mathematics Aims and scope Submit manuscript

Abstract

Shape gradients of PDE constrained shape functionals can be stated in two equivalent ways. Both rely on the solutions of two boundary value problems (BVPs), but one involves integrating their traces on the boundary of the domain, while the other evaluates integrals in the volume. Usually, the two BVPs can only be solved approximately, for instance, by finite element methods. However, when used with finite element solutions, the equivalence of the two formulas breaks down. By means of a comprehensive convergence analysis, we establish that the volume based expression for the shape gradient generally offers better accuracy in a finite element setting. The results are confirmed by several numerical experiments.

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

Similar content being viewed by others

Notes

  1. For simplicity, we assume that the operator \(\mathcal{L}\) is self-adjoint.

  2. Note that \(p_h\) is not a proper Ritz–Galerkin solution of (2.7), because the right-hand side is perturbed.

  3. We write \(C\) for generic constants, whose value may differ between different occurrences. They may depend only on \(\varOmega \), shape-regularity and quasi-uniformity of the meshes.

  4. For the sake of readability, we use the same notation for scalar and vectorial Sobolev norms.

  5. Many bounds in this proof rely on duality techniques, which introduce so-called adjoint BVPs. For the sake of readability, we abuse the notation and we always denote by \(w\) the solutions of these BVPs.

  6. In experiments 1 and 4 we consider domains with curved boundaries. In this case the refined mesh is always adjusted to fit the boundary.

  7. The experiments are performed in MATLAB and are based on the library LehrFEM developed at ETHZ.

References

  1. Adams, R.A., Fournier, J.J.F.: Sobolev Spaces, 2nd edn. Elsevier/Academic Press, Amsterdam (2003)

    MATH  Google Scholar 

  2. Allaire, G.: Conception optimale de structures. Springer, New York (2007)

  3. Allaire, G., de Gournay, F., Jouve, F., Toader, A.M.: Structural optimization using topological and shape sensitivity via a level set method. Control Cybernet. 34(1), 59–80 (2005)

    MATH  MathSciNet  Google Scholar 

  4. Becker, R., Rannacher, R.: An optimal control approach to a posteriori error estimation in finite element methods. Acta Numer. 10, 1–102 (2001)

    Article  MATH  MathSciNet  Google Scholar 

  5. Berggren, M.: A unified discrete-continuous sensitivity analysis method for shape optimization. In: Applied and Numerical Partial Differential Equations, pp. 25–39. Springer, New York (2010)

  6. Braess, D.: Finite Elements. Theory, Fast Solvers, and Applications in Elasticity Theory, 3rd edn. Cambridge University Press, Cambridge (2007)

  7. Brenner, S.C., Scott, L.R.: The Mathematical Theory of Finite Element Methods, 3rd edn. Springer, New York (2008)

  8. Bucur, D., Buttazzo, G.: Variational Methods in Shape Optimization Problems. Birkhäuser Boston Inc., Basel (2005)

  9. Céa, J.: Conception optimale ou identification de formes: calcul rapide de la dérivée directionnelle de la fonction coût. RAIRO Modél. Math. Anal. Numér. 20(3), 371–402 (1986)

    MATH  MathSciNet  Google Scholar 

  10. Delfour, M.C., Zolésio, J.P.: Shapes and Geometries. Metrics, Analysis, Differential Calculus, and Optimization. 2nd edn. Society for Industrial and Applied Mathematics (SIAM) (2011)

  11. Eppler, K.: Boundary integral representations of second derivatives in shape optimization. Discuss. Math. Differ. Incl. Control Optim. 20(1), 63–78 (2000)

    Article  MATH  MathSciNet  Google Scholar 

  12. Eppler, K.: Second derivatives and sufficient optimality conditions for shape functionals. Control Cybernet. 29(2), 485–511 (2000)

    MATH  MathSciNet  Google Scholar 

  13. Eppler, K., Harbrecht, H.: Coupling of FEM and BEM in shape optimization. Numer. Math. 104(1), 47–68 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  14. Firl, M., Wüchner, R., Bletzinger, K.U.: Regularization of shape optimization problems using FE-based parametrization. Struct. Multidiscip. Optim. 47(4), 507–521 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  15. Gunzburger, M.D.: Perspectives in flow control and optimization. Society for Industrial and Applied Mathematics (SIAM) (2003)

  16. Harbrecht, H.: On output functionals of boundary value problems on stochastic domains. Math. Methods Appl. Sci. 33(1), 91–102 (2010)

    MATH  MathSciNet  Google Scholar 

  17. Harbrecht, H., Tausch, J.: On the numerical solution of a shape optimization problem for the heat equation. SIAM J. Sci. Comput. 35(1), A104–A121 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  18. Haslinger, J., Mäkinen, R.A.E.: Introduction to Shape Optimization. Theory, Approximation, and Computation. Society for Industrial and Applied Mathematics (SIAM) (2003)

  19. Laporte, E., Le Tallec, P.: Numerical Methods in Sensitivity Analysis and Shape Optimization. Birkhäuser Boston Inc., Basel (2003)

  20. Le, C., Bruns, T., Tortorelli, D.: A gradient-based, parameter-free approach to shape optimization. Comput. Methods Appl. Mech. Eng. 200(9–12), 985–996 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  21. McFee, S., Webb, J., Lowther, D.: A tunable volume integration formulation for force calculation in finite-element based computational magnetostatics. IEEE Trans. Magnetics 24(1), 439–442 (1988)

    Article  Google Scholar 

  22. Monk, P.: Finite Element Methods for Maxwell’s Equations. Clarendon Press (2003)

  23. Monk, P., Süli, E.: The adaptive computation of far-field patterns by a posteriori error estimation of linear functionals. SIAM J. Numer. Anal. 36(1), 251–274 (1999)

    Article  MathSciNet  Google Scholar 

  24. Murat, F., Simon, J.: Etude de problèmes d’optimal design. In: Optimization Techniques Modeling and Optimization in the Service of Man Part 2, pp. 54–62. Springer, Berlin (1976)

  25. Murat, F., Simon, J.: Sur le contrôle par un domaine géométrique. Internal Report No 76 015, Laboratoire d’Analyse Numérique de l’Université Paris 6 (1976)

  26. Pironneau, O.: Optimal Shape Design for Elliptic Systems. Springer, Berlin (1984)

  27. Simon, J.: Differentiation with respect to the domain in boundary value problems. Numer. Funct. Anal. Optim. 2(7–8), 649–687 (1981)

  28. Sokołowski, J., Zolésio, J.P.: Introduction to Shape Optimization. Shape Sensitivity Analysis. Springer, New York (1992)

    Book  MATH  Google Scholar 

  29. Udawalpola, R., Wadbro, E., Berggren, M.: Optimization of a variable mouth acoustic horn. Int. J. Numer. Methods Eng. 85(5), 591–606 (2011)

    Article  MATH  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. Paganini.

Additional information

Communicated by R. Winther.

The work of A. Paganini and S. Sargheini was partly supported by ETH Grant CH1-02 11-1.

Appendix

Appendix

Closely following [10, Ch. 10, Sect. 6], we give a detailed derivation of Formulas (2.9)–(2.13). Let \(u\) be the weak solution in \(H^{1}({\varOmega })\) of the following state problem:

$$\begin{aligned} \left\{ \begin{array}{l@{\quad }l} -\Delta u + u = f &{} \text {in } \varOmega \,,\\ u = g &{} \text {on } \partial \varOmega \,. \end{array}\right. \end{aligned}$$
(5.1)

It is assumed that the Dirichlet problem (5.1) is \(H^2\)-regular, so that its solution \(u\) is at least in \(H^2(\varOmega )\) for \(f\in L^{2}({\varOmega })\). We consider the shape functional

$$\begin{aligned} \mathcal{J}(\varOmega ) = \int _{\varOmega } j(u) \, d{\mathbf {x}}\,, \end{aligned}$$

and we introduce the Lagrangian

$$\begin{aligned} \fancyscript{L} (\varOmega ,v,q,\lambda ) :\,=\int _\varOmega j(v) +(\Delta v - v + f)q\, d{\mathbf {x}}+ \int _{\partial \varOmega } \lambda (g - v)\, dS\,, \end{aligned}$$
(5.2)

where the functions \(v\), \(q\) and \(\lambda \) are in \(H^2(\mathbb {R}^d)\). Performing integration by parts, the Lagrangian can be rewritten as

$$\begin{aligned}&\fancyscript{L} (\varOmega ,v,q,\lambda ) = \int _\varOmega j(v) - \nabla v \cdot \nabla q - v\,q + f\,q\, d{\mathbf {x}}+ \int _{\partial \varOmega } \frac{\partial {v}}{\partial {\mathbf {n}}}\,q + \lambda (g - v)\, dS\,,\\&\quad = \int _\varOmega j(v) +(\Delta q - q)v + f\,q\, d{\mathbf {x}}+ \int _{\partial \varOmega } \frac{\partial {v}}{\partial {\mathbf {n}}}\,q - \frac{\partial {q}}{\partial {\mathbf {n}}}\,v + \lambda (g - v)\, dS\,. \end{aligned}$$

The saddle point of \(\fancyscript{L}(\varOmega ,\cdot \,,\cdot \,,\cdot )\) is characterized by

$$\begin{aligned} \left\langle \frac{\partial \fancyscript{L}(\varOmega ,v,q,\lambda )}{\partial v}, \phi \right\rangle _{\varOmega } \!\!= \left\langle \frac{\partial \fancyscript{L}(\varOmega ,v,q,\lambda )}{\partial q}, \phi \right\rangle _{\varOmega } \!\!= \left\langle \frac{\partial \fancyscript{L}(\varOmega ,v,q,\lambda )}{\partial \lambda }, \phi \right\rangle _{\partial \varOmega } \!\!= 0 \end{aligned}$$

for all \(\phi \in H^2(\mathbb {R}^d)\), which, by density, leads to

$$\begin{aligned}&\left\{ \begin{array}{l@{\quad }l} -\Delta v + v = f&{} \text {in } \varOmega \,,\\ v = g&{} \text {on } \partial \varOmega \,, \end{array}\right. \end{aligned}$$
(5.3a)
$$\begin{aligned}&\left\{ \begin{array}{l@{\quad }l} -\Delta q + q = j'(v) &{}\text {in } \varOmega \,,\\ q = 0 &{}\text {on } \partial \varOmega \,, \end{array}\right. \end{aligned}$$
(5.3b)
$$\begin{aligned}&\quad \lambda = -\frac{\partial {q}}{\partial {\mathbf {n}}}\quad \text {on } \partial \varOmega \,, \end{aligned}$$
(5.3c)

weakly in \(H^1(\mathbb {R}^d)\). Thus, for \(\varOmega \) fixed,

$$\begin{aligned} \mathcal{J}(\varOmega ) = \inf _{v \in H^2(\mathbb {R}^d)} \sup _{q,\lambda \in H^2(\mathbb {R}^d)} \fancyscript{L}(\varOmega ,v,q,\lambda )\,, \end{aligned}$$
(5.4)

because

$$\begin{aligned} \mathcal{J}(\varOmega ) = \fancyscript{L}(\varOmega ,u,q,\lambda ) \quad \forall \, q,\lambda \in H^2(\mathbb {R}^d)\,. \end{aligned}$$

Recall that the material derivative of a generic function \(f\) with respect to the deformation \(T_\mathcal{V}\) is defined as

$$\begin{aligned} \dot{f} :\,=\lim _{s\searrow 0} \frac{f\circ T_{s\cdot \mathcal{V}} - f}{s}\,. \end{aligned}$$

Note that, if \(f\) is independent of \(\varOmega \), \(\dot{f}\in H^1(\mathbb {R}^d)\) for \(f\in H^2(\mathbb {R}^d)\).

To compute the Eulerian derivative of \(\mathcal{J}(\varOmega )\), the Correa–Seeger theorem can be applied on the right-hand side of (5.4) [10, Ch. 10, Sect. 6.3], so that a formula for \(d\mathcal{J}(\varOmega )\) can be obtained by evaluating the Eulerian derivative of the Lagrangian (5.2) in its saddle point. For \(T_\mathcal{V}({\mathbf {x}}):\,={\mathbf {x}}+ \mathcal{V}({\mathbf {x}})\), the Eulerian derivative of (5.2) reads

$$\begin{aligned}&\lim _{s\searrow 0} \frac{\fancyscript{L}(T_{s\cdot \mathcal{V}}(\varOmega ),v,q,\lambda ) -\fancyscript{L}(\varOmega ,v,q,\lambda )}{s} \\&\quad = \int _\varOmega \left( j'(v) \dot{v} - \nabla \dot{v}\cdot \nabla q - \nabla v\cdot \nabla \dot{q} + \nabla v \cdot ({\mathbf {D}}\mathcal{V}+{\mathbf {D}}\mathcal{V}^T) \nabla q \right. \\&\left. \qquad - \dot{v}\,q - v\,\dot{q} + \dot{f}q + f\dot{q} + \mathrm{div }(\mathcal{V}) \left( j(v) -\nabla v \cdot \nabla q - v\, q + fq \right) \right) \, d{\mathbf {x}}\\&\qquad +\int _{\partial \varOmega } \dot{\frac{\partial {v}}{\partial {\mathbf {n}}}}q + \frac{\partial {v}}{\partial {\mathbf {n}}}\dot{q} +\lambda (\dot{g}-\dot{v}) + \dot{\lambda }(g-v) + \mathrm{div }_\Gamma (\mathcal{V}) \left( \frac{\partial {v}}{\partial {\mathbf {n}}}q + \lambda (g-v) \right) \,dS \\&\quad = \int _\varOmega j'(v) \dot{v} + \Delta q \, \dot{v} - q\, \dot{v} \, d{\mathbf {x}}+ \int _\varOmega \Delta v \, \dot{q} -v \, \dot{q} + f \, \dot{q} \, d{\mathbf {x}}\\&\qquad + \int _{\partial \varOmega } \dot{\frac{\partial {v}}{\partial {\mathbf {n}}}} q + \dot{\lambda }(g-v) + \mathrm{div }_\Gamma (\mathcal{V})\left( \frac{\partial {v}}{\partial {\mathbf {n}}} q + \lambda (g-v) \right) \, dS \\&\qquad + \int _\varOmega \nabla v \cdot ({\mathbf {D}}\mathcal{V}+{\mathbf {D}}\mathcal{V}^T) \nabla q + \dot{f}q + \mathrm{div }(\mathcal{V}) \left( j(v) -\nabla v \cdot \nabla q - v\, q + fq \right) \, d{\mathbf {x}}\\&\qquad + \int _{\partial \varOmega } \lambda (\dot{g}-\dot{v})-\frac{\partial {q}}{\partial {\mathbf {n}}}\dot{v} \, dS\,. \end{aligned}$$

So, in the saddle point defined by (5.3), we have

$$\begin{aligned}&\lim _{s\searrow 0} \frac{\fancyscript{L}(T_{s\cdot \mathcal{V}}(\varOmega ),v,q,\lambda ) -\fancyscript{L}(\varOmega ,v,q,\lambda )}{s}= \\&\quad =\int _\varOmega \nabla v \cdot ({\mathbf {D}}\mathcal{V}+{\mathbf {D}}\mathcal{V}^T) \nabla q + \dot{f}q + \mathrm{div }(\mathcal{V}) \left( j(v) -\nabla v \cdot \nabla q - v\, q + fq \right) \, d{\mathbf {x}}\\&\qquad + \int _{\partial \varOmega } -\frac{\partial {q}}{\partial {\mathbf {n}}}\dot{g} \, dS \\&\quad =\int _\varOmega \left( \nabla v \cdot ({\mathbf {D}}\mathcal{V}+{\mathbf {D}}\mathcal{V}^T) \nabla q + \dot{f}q + (j'(v) - q) \dot{g} - \nabla q \cdot \nabla \dot{g}\right. \\&\qquad \left. + \mathrm{div }(\mathcal{V}) \left( j(v) -\nabla v \cdot \nabla q - v\, q + fq \right) \right) \, d{\mathbf {x}}\,, \end{aligned}$$

which, after an additional integration by parts on the term \(\dot{f}q = \nabla f\cdot \mathcal{V}q\), corresponds to Formula (2.9). Formula (2.10) is obtained performing additional integrations by parts and exploiting the vector calculus identity

$$\begin{aligned} \mathcal{V}\cdot \nabla \left( \nabla v\cdot \nabla q\right) + \nabla v \cdot \left( {\mathbf {D}}\mathcal{V}+{\mathbf {D}}\mathcal{V}^T\right) \nabla q = \nabla \left( \mathcal{V}\cdot \nabla v\right) \cdot \nabla q +\nabla v\cdot \nabla \left( \mathcal{V}\cdot \nabla v \right) . \end{aligned}$$

We refer to [5, Sect. 6] for a detailed derivation. Alternatively, (2.10) can be derived with the so-called “fast derivation” method of Céa, which, formally, does not rely on the concept of material derivative, cf. [9] and [2, Ch. 6.4.3].

Similarly, Formula (2.11) can be derived considering the Lagrangian

$$\begin{aligned} \nonumber \mathcal{L}(\varOmega ,v,q)&:= \int _\varOmega j(v) +(\Delta v - v + f)q \,d{\mathbf {x}}+\int _{\partial \varOmega } gq - \frac{\partial {v}}{\partial {\mathbf {n}}}q\,dS\,,\nonumber \\&=\int _\varOmega j(v) - \nabla v\cdot \nabla q - vq + fq\,d{\mathbf {x}}+ \int _{\partial \varOmega } gq\,dS\,,\nonumber \\&=\int _\varOmega j(v) + (\Delta q - q)v + fq \,d{\mathbf {x}}+\int _{\partial \varOmega } gq-\frac{\partial {q}}{\partial {\mathbf {n}}}v\,dS\,. \end{aligned}$$
(5.5)

Its saddle point is characterized by

$$\begin{aligned}&\left\{ \begin{array}{l@{\quad }l} -\Delta v + v = f &{} \text {in } \varOmega \,,\\ \frac{\partial {v}}{\partial {\mathbf {n}}}=g &{} \text {on } \partial \varOmega \,, \end{array}\right. \end{aligned}$$
(5.6a)
$$\begin{aligned}&\left\{ \begin{array}{l@{\quad }l} -\Delta q + q = j'(v) &{} \text {in } \varOmega \,,\\ \frac{\partial {q}}{\partial {\mathbf {n}}} = 0 &{} \text {on } \partial \varOmega \,. \end{array}\right. \end{aligned}$$
(5.6b)

Thus, the Eulerian derivative of (5.5) in (5.6) reads

$$\begin{aligned}&\lim _{s\searrow 0} \frac{\fancyscript{L}(T_{s\cdot \mathcal{V}}(\varOmega ),v,q,\lambda ) -\fancyscript{L}(\varOmega ,v,q,\lambda )}{s}= \\&\quad = \int _\varOmega \nabla v \cdot ({\mathbf {D}}\mathcal{V}+{\mathbf {D}}\mathcal{V}^T) \nabla q + \dot{f}q + \mathrm{div }(\mathcal{V}) \left( j(v) -\nabla v \cdot \nabla q - v\, q + fq \right) \, d{\mathbf {x}}\\&\qquad + \int _{\partial \varOmega } \dot{g} q + \mathrm{div }_\Gamma (\mathcal{V})\left( gq \right) \, dS\,. \end{aligned}$$

In this case, the term \(\mathrm{div }_\Gamma (\mathcal{V})\) does not vanish, and to recover Formula (2.12) it is necessary to perform an integration by parts on the boundary, from which stems the mean curvature term. For piecewise smooth boundaries, this step has to be performed carefully, because, as in Remark 2.3, additional contributions of corner points appear.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hiptmair, R., Paganini, A. & Sargheini, S. Comparison of approximate shape gradients. Bit Numer Math 55, 459–485 (2015). https://doi.org/10.1007/s10543-014-0515-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10543-014-0515-z

Keywords

Mathematics Subject Classification

Navigation