Skip to main content
Log in

Global optimality in minimum compliance topology optimization of frames and shells by moment-sum-of-squares hierarchy

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

This article has been updated

Abstract

The design of minimum-compliance bending-resistant structures with continuous cross-section parameters is a challenging task because of its inherent non-convexity. Our contribution develops a strategy that facilitates computing all guaranteed globally optimal solutions for frame and shell structures under multiple-load cases and self-weight. To this purpose, we exploit the fact that the stiffness matrix is usually a polynomial function of design variables, allowing us to build an equivalent non-linear semidefinite programming formulation over a semi-algebraic feasible set. This formulation is subsequently solved using the Lasserre moment-sum-of-squares hierarchy, generating a sequence of outer convex approximations that monotonically converges from below to the optimum of the original problem. Globally optimal solutions can subsequently be extracted using the Curto-Fialkow flat extension theorem. Furthermore, we show that a simple correction to the solutions of the relaxed problems establishes a feasible upper bound, thereby deriving a simple sufficient condition of global ε-optimality. When the original problem possesses a unique minimum, we show that this solution is found with a zero optimality gap in the limit. These theoretical findings are illustrated on several examples of topology optimization of frames and shells, for which we observe that the hierarchy converges in a finite (rather small) number of steps.

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

Change history

  • 09 February 2022

    The reference to equation 25 was updated in XML.

Notes

  1. For OC and MMA, we adopted the commonly used starting point of uniform mass distribution, i.e., \(a_{1}=a_{2}=0.2\sqrt {5}\). For fmincon, the default starting point was used.

  2. Let X,Y be real matrices of the same dimensions. Then, 〈X,Y〉 := Tr(XYT), where Tr is the trace operator.

  3. In particular, the transition from (17b) and (17e) to (21b) is reverse to how we approach Theorem 2 from Proposition 6.

  4. The lowest relaxation might not produce feasible upper bounds. For example, consider cII > 0,cIII = 0, and a one-element cantilever beam with one end fully clamped and the other carrying a moment load. Then, \(y_{a_{1}}=-1\), \(y_{{a_{1}^{2}}}=\infty \) with \(y_{c_{1}}=-1\) belongs to the set of optimal solutions to the first relaxation.

  5. For rank computation, we considered the eigenvalues with the absolute value smaller than 10− 8 to be singular.

  6. Fewer digits may prevent the solver from reaching all three global optima. Although an analytical formula for this specific \(\overline {V}\) can be derived, we omit it for the sake of brevity.

References

Download references

Acknowledgements

We thank Edita Dvořáková for providing us with her implementation of the Mitc4 shell elements (Dvořáková 2015). We have also gratefully appreciated useful comments and suggestions of two anonymous reviewers, who helped us improve the quality of this manuscript and brought our attention to Section 5.3 of Murota et al. (2010).

Funding

Marek Tyburec, Jan Zeman, and Martin Kružík received the support of the Czech Science Foundation project No. 19-26143X.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marek Tyburec.

Ethics declarations

Conflict of interest

The authors declare no competing interests.

Additional information

Responsible Editor: Shikui Chen

Replication of results

Source codes are available at Tyburec et al. (2020b).

Publisher’s note

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

Appendices

Appendix : 1: Relation to truss topology optimization

The problem formulation (21a) has already been known in the context of truss topology optimization (Vandenberghe and Boyd 1996), for which the constraints (21b) reduce to linear matrix inequalities (LMI). Consequently, the feasible set is convex, allowing for an efficient solution of (21a) by interior point methods, for example.

A natural question then arises: What happens when the rotational degrees of freedom are neglected, solving truss topology optimization problem instead of the frame one? To this goal, however, we must first satisfy the rather restrictive assumption that the truss ground structure is capable of carrying the loads fj(a), i.e.,

$$ \mathbf{f}_{j} (\mathbf{a}) \in \text{Im}\left( \mathbf{K}_{\mathrm{t},j}(\mathbf{a}) \right), \forall j \in \{1{\dots} n_{\text{lc}}\}, $$
(43)

where \(\mathbf {K}_{\mathrm {t},j} (\mathbf {a}) = \mathbf {K}_{j,0}+ {\sum }_{i=1}^{n_{\mathrm {e}}} \mathbf {K}_{j,i}^{(1)} a_{i}\), and that with all empty rows and columns removed Kt,j(a) is positive definite for all positive a. From the mechanical point of view, we require that no moment loads are imposed, a straight bar does not have to carry transverse loads, and the ground structure is well supported.

Suppose now that \(\mathbf {a}^{*}_{\mathrm {t}}\) are optimal cross-sections obtained from a solution to (21a) with the terms \(\mathbf {K}_{j,i}^{(2)}\) and \(\mathbf {K}_{j,i}^{(3)}\) neglected, and \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}^{*}_{\mathrm {t}}\) is the associated optimal objective function value, which can be computed from \(\mathbf {a}_{\mathrm {t}}^{*}\) as

$$ \boldsymbol{\omega}^{\mathrm{T}} \mathbf{c}^{*}_{\mathrm{t}} = \sum\limits_{j=1}^{n_{\text{lc}}}\left( \omega_{j} \left[\mathbf{f}_{j}(\mathbf{a}_{\mathrm{t}}^{*})\right]^{\mathrm{T}} \left[\mathbf{K}_{\mathrm{t},j}(\mathbf{a}_{\mathrm{t}}^{*}) \right]^{\dagger} \mathbf{f}_{j}(\mathbf{a}_{\mathrm{t}}^{*}) \right). $$
(44)

When the optimal cross-sections of a truss structure, \(\mathbf {a}^{*}_{\mathrm {t}}\), are reused in a frame structure, the resulting objective function value changes to

$$ \boldsymbol{\omega}^{\mathrm{T}} \mathbf{c}_{\mathrm{f}} = \sum\limits_{j=1}^{n_{\text{lc}}}\left( \omega_{j} \left[\mathbf{f}_{j}(\mathbf{a}_{\mathrm{t}}^{*})\right]^{\mathrm{T}} \left[\mathbf{K}_{\mathrm{t},j}(\mathbf{a}_{\mathrm{t}}^{*}) + \mathbf{K}_{\mathrm{b},j}(\mathbf{a}_{\mathrm{t}}^{*}) \right]^{\dagger} \mathbf{f}_{j}(\mathbf{a}_{\mathrm{t}}^{*}) \right) $$
(45)

with \(\mathbf {K}_{\mathrm {b},j}(\mathbf {a}) = {\sum }_{i=1}^{n_{\mathrm {e}}}\left (c_{\text {II}} \mathbf {K}_{j,i}^{(2)} a_{\mathrm {t},i}^{2} + c_{\text {III}}\mathbf {K}_{j,i}^{(3)} a_{\mathrm {t},i}^{3}\right )\).

To state a relation between \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}^{*}_{\mathrm {t}}\) and \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {f}}\) we recall a useful lemma:

Lemma 4

(Kovanic 1979) Let \(\mathbf {A} \in \mathbb {S}^{n}\) and \(\mathbf {B} \in \mathbb {R}^{n\times q}\). Then,

$$ \begin{array}{@{}rcl@{}} &&\left( \mathbf{A} + \mathbf{B}\mathbf{B}^{\mathrm{T}} \right)^{\dagger} = \mathbf{A}^{\dagger} \\ &&- \mathbf{A}^{\dagger} \mathbf{B} \left( \mathbf{I} + \mathbf{B}^{\mathrm{T}}\mathbf{A}^{\dagger} \mathbf{B}\right)^{-1} \mathbf{B}^{\mathrm{T}} \mathbf{A}^{\dagger} + \left( \mathbf{B}^{\dagger}_{\perp}\right)^{\mathrm{T}} \mathbf{B}^{\dagger}_{\perp} \end{array} $$
(46)

with \(\mathbf {B}_{\perp } = \left (\mathbf {I} - \mathbf {A}\mathbf {A}^{\dagger } \right ) \mathbf {B}\).

Using this lemma, we prove that \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {t}}\) provides an upper bound for \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {f}}\).

Lemma 5

Suppose that ωTc is the optimal objective function value of the frame structure design problem (21a) and (43) holds. Then, \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}^{*} \le \boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {f}} \le \boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}^{*}_{\mathrm {t}}\).

Proof

Because of \(\mathbf {f}_{j}(\mathbf {a}_{\mathrm {t}}^{*}) \in \text {Im}\left (\mathbf {K}_{\mathrm {t},j}(\mathbf {a}_{\mathrm {t}}^{*})\right )\), we clearly have \(\mathbf {f}_{j} (\mathbf {a}_{\mathrm {t}}^{*}) \in \text {Im}\left (\mathbf {K}_{\mathrm {t},j}(\mathbf {a}_{\mathrm {t}}^{*}) + \mathbf {K}_{\mathrm {b},j}(\mathbf {a}_{\mathrm {t}}^{*}) \right )\). Therefore, \(\mathbf {a}_{\mathrm {t}}^{*}\) is a feasible solution to the frame structure design problem (21a) and the associated objective function is bounded from below by the global optimum ωTc. Hence, \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}^{*} \le \boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {f}}\).

For the other inequality, we express (45) using Lemma 4. To this goal, let \(\mathbf {K}_{\mathrm {b},j}(\mathbf {a}^{*}_{\mathrm {t}}) = \mathbf {B}_{j} \mathbf {B}_{j}^{\mathrm {T}}\), where Bj is a real matrix because \(\mathbf {K}_{\mathrm {b},j}(\mathbf {a}^{*}_{\mathrm {t}}) \succeq 0\) by definition. Then, (45) can be written as

$$ \boldsymbol{\omega}^{\mathrm{T}} \mathbf{c}_{\mathrm{f}} = \boldsymbol{\omega}^{\mathrm{T}} \mathbf{c}_{\mathrm{t}}^{*} - \boldsymbol{\omega}^{\mathrm{T}} \mathbf{c}_{\mathrm{a}} + \boldsymbol{\omega}^{\mathrm{T}} \mathbf{c}_{\mathrm{b}}, $$
(47)

where

$$ \begin{array}{@{}rcl@{}} c_{\mathrm{a},j} &{=}& \left[\mathbf{f}_{j}(\mathbf{a}_{\mathrm{t}}^{*})\right]^{\mathrm{T}} \mathbf{A}_{j}^{\dagger} \mathbf{B}_{j} \left( \mathbf{I} {+} \mathbf{B}_{j}^{\mathrm{T}}\mathbf{A}_{j}^{\dagger} \mathbf{B}_{j}\right)^{-1} \mathbf{B}_{j}^{\mathrm{T}} \mathbf{A}_{j}^{\dagger} \mathbf{f}_{j}(\mathbf{a}_{\mathrm{t}}^{*}), \end{array} $$
(48a)
$$ \begin{array}{@{}rcl@{}} c_{\mathrm{b},j} &{=}& \left[\mathbf{f}_{j}(\mathbf{a}_{\mathrm{t}}^{*})\right]^{\mathrm{T}} \left( \mathbf{B}^{\dagger}_{\perp,j}\right)^{\mathrm{T}} \mathbf{B}^{\dagger}_{\perp,j} \mathbf{f}_{j}(\mathbf{a}_{\mathrm{t}}^{*}), \end{array} $$
(48b)

with \(\mathbf {A}_{j} = \mathbf {K}_{\mathrm {t},j}(\mathbf {a}_{\mathrm {t}}^{*})\). Clearly, (48a) is non-negative. For (48b), \(\left (\mathbf {B}^{\dagger }_{\perp }\right )^{\mathrm {T}} \mathbf {B}^{\dagger }_{\perp } \in \text {Ker}(\mathbf {A}_{j})\), so that \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {b}}\) vanishes. Hence, \(\boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {f}} = \boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {t}}^{*} - \boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {a}} \le \boldsymbol {\omega }^{\mathrm {T}} \mathbf {c}_{\mathrm {t}}^{*}\). □

Thus, when (43) holds true, the truss topology optimization produces an upper bound to the optimal objective of the frame structure topology optimization problem.

Appendix 2: Numerical performance of bound constraints

This section illustrates the effect of using different types of bound constraints, (24a) or (24b), on performance of the moment-sum-of-squares hierarchy.

We start by re-evaluating the problem in Section 4.3. Our numerical experiments in Table 6 reveal that for this specific problem the quadratic constraints (24b) are substantially tighter in terms of generated lower bounds, they require smaller relaxation degree to converge, and are computationally more efficient. Also notice that fewer constraints are needed when using (24b).

Table 6 Performance of the hierarchy on the problem with self-weight using different bound constraints. LB denotes lower bound, and PO stands for polynomial optimization. The bold lower bounds are certified global optima. In addition, nc stands for the number of semidefinite constraints of the size m and n is the number of variables. The entries ai denote the cross-section areas constructed from the first-order moments

For the problem in Section 4.4 discretized by Euler-Bernoulli frame elements, we observe that the difference in performance becomes less noticeable (see Table 7): the quadratic constraints lead to a better lower bound in the third relaxation only, but are still computationally more efficient because fewer constraints are needed.

Table 7 Performance of the hierarchy on the cantilever problem using different bound constraints. LB denotes lower bound, and PO stands for polynomial optimization. The bold lower bounds are certified global optima. In addition, nc stands for the number of semidefinite constraints of the size m and n is the number of variables. The entries ai denote the cross-section areas constructed from the first-order moments

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tyburec, M., Zeman, J., Kružík, M. et al. Global optimality in minimum compliance topology optimization of frames and shells by moment-sum-of-squares hierarchy. Struct Multidisc Optim 64, 1963–1981 (2021). https://doi.org/10.1007/s00158-021-02957-5

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-021-02957-5

Keywords

Navigation