Skip to main content
Log in

On compliance and buckling objective functions in topology optimization of snap-through problems

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

Abstract

This paper deals with topology optimization of static geometrically nonlinear structures experiencing snap-through behaviour. Different compliance and buckling criterion functions are studied and applied for topology optimization of a point loaded curved beam problem with the aim of maximizing the snap-through buckling load. The response of the optimized structures obtained using the considered objective functions are evaluated and compared. Due to the intrinsic nonlinear nature of the problem, the load level at which the objective function is evaluated has a tremendous effect on the resulting optimized design. A well-known issue in buckling topology optimization is artificial buckling modes in low density regions. The typical remedy applied for linear buckling does not have a natural extension to nonlinear problems, and we propose an alternative approach. Some possible negative implications of using symmetry to reduce the model size are highlighted and it is demonstrated how an initial symmetric buckling response may change to an asymmetric buckling response during the optimization process. This problem may partly be avoided by not exploiting symmetry, however special requirements are needed of the analysis method and optimization formulation. We apply a nonlinear path tracing algorithm capable of detecting different types of stability points and an optimization formulation that handles possible mode switching. This is an extension into the topology optimization realm of a method developed, and used for, fiber angle optimization in laminated composite structures. We finally discuss and pinpoint some of the issues related to buckling topology optimization that remains unsolved and demands further research.

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

Similar content being viewed by others

References

  • Bendsøe MP (1989) Optimal shape design as a material distribution problem. Struct Multidisc Optim 1(4):193–202

    Article  Google Scholar 

  • Bendsøe MP, Kikuchi N (1988) Generating optimal topologies in structural design using a homogenization method. Comput Methods Appl Mech Eng 71:197–224

    Article  Google Scholar 

  • Bendsøe MP, Sigmund O (1999) Material interpolation schemes in topology optimization. Arch Appl Mech 69:635–654

    Article  Google Scholar 

  • Bendsøe MP, Sigmund O (2003) Topology optimization – theory, methods and applications, 2nd edn. Springer, Berlin Heidelberg, New York, ISBN: 3-540-42992-1

    Google Scholar 

  • Bendsøe MP, Olhoff N, Taylor JE (1983) A variational formulation for multicriteria structural optimization. J Struct Mech 11:523–544

    Article  Google Scholar 

  • Borri C, Hufendiek HW (1985) Geometrically nonlinear behaviour of space beam structures. J Struct Mech 13:1–26

    Article  Google Scholar 

  • Brendel B, Ramm E (1980) Linear and nonlinear stability analysis of cylindrical shells. Comput Struct 12:549–558

    Article  MATH  Google Scholar 

  • Bruns TE, Sigmund O (2004) Toward the topology design of mechanisms that exhibit snap-through behavior. Comput Methods Appl Mech Eng 193:3973–4000

    Article  MathSciNet  MATH  Google Scholar 

  • Bruns TE, Sigmund O, Tortorelli DA (2002) Numerical methods for the topology optimization of structures that exhibit snap-through. Int J Numer Methods Eng 55:1215–1237

    Article  MATH  Google Scholar 

  • Buhl T, Pedersen CBW, Sigmund O (2000) Stiffness design of geometrically nonlinear structures using topology optimization. Struct Multidisc Optim 19:93–104

    Article  Google Scholar 

  • Courant R, Hilbert D (1953) Methods of mathematical physics, vol 1. Interscience Publishers, New York

    Google Scholar 

  • Crisfield MA (1981) A fast incremental/iterative solution procedure that handles “snap-through”. Comput Struct 13:55–62

    Article  MATH  Google Scholar 

  • Du J, Olhoff N (2007) Topological design of freely vibrating continuum structures for maximum values of simple and multiple eigenfrequencies and frequency gaps. Struct Multidisc Optim 34:91–110. doi:10.1007/s00158-007-0101-y

    Article  MathSciNet  Google Scholar 

  • Hinton E (ed) (1992) NAFEMS Introduction to Nonlinear Finite Element Analysis. Bell and Bain Ltd, Glasgow, ISBN 1-874376-00-X

    Google Scholar 

  • Kemmler R, Lipka A, Ramm E (2005) Large deformations and stability in topology optimization. Struct Multidisc Optim 30:459–476

    Article  MathSciNet  MATH  Google Scholar 

  • Khot NS, Venkayya VB, Berke L (1976) Optimal structural design with stability constraints. Int J Numer Methods Eng 10:1097–1114

    Article  MATH  Google Scholar 

  • Lee SJ, Hinton E (2000) Dangers inherited in shells optimized with linear assumptions. Comput Struct 78:473–486

    Article  Google Scholar 

  • Lindgaard E, Lund E (2010) Nonlinear buckling optimization of composite structures. Compt Methods Appl Mech Eng 199(37–40):2319–2330. doi:10.1016/j.cma.2010.02.005

    Article  MathSciNet  MATH  Google Scholar 

  • Lindgaard E, Lund E (2011) A unified approach to nonlinear buckling optimization of composite structures. Comput Struct 89(3–4):357–370. doi:10.1016/j.compstruc.2010.11.008

    Article  Google Scholar 

  • Lindgaard E, Lund E, Rasmussen K (2010) Nonlinear buckling optimization of composite structures considering “worst” shape imperfections. Int J Solids Struct 47(22–23):3186–3202. doi:10.1016/j.ijsolstr.2010.07.020

    Article  MATH  Google Scholar 

  • Lund E, Stegmann J (2005) On structural optimization of composite shell structures using a discrete constitutive parametrization. Wind Energy 8:109–124. doi:10.1002/we.132

    Article  Google Scholar 

  • Manickarajaha D, Xiea YM, Steven GP (2000) Optimisation of columns and frames against buckling. Comput Struct 75:45–54

    Article  Google Scholar 

  • Min S, Kikuchi N (1997) Optimal reinforcement design of structures under the buckling load using the homogenization design method. Struct Eng Mech 5:565–576

    Google Scholar 

  • Neves MM, Rodrigues H, Guedes JM (1995) Generalized topology design of structures with a buckling load criterion. Struct Optim 10:71–78

    Article  Google Scholar 

  • Neves MM, Sigmund O, Bendsøe MP (2002) Topology optimization of periodic microstructures with a penalization of highly localized buckling modes. Int J Numer Meth Eng 54:809–834

    Article  MATH  Google Scholar 

  • Olhoff N (1989) Multicriterion structural optimization via bound formulation and mathematical programming. Struct Optim 1:11–17

    Article  Google Scholar 

  • Olhoff N, Rasmussen SH (1977) On single and bimodal optimum buckling loads of clamped columns. Int J Solids Struct 13:605–614

    Article  MATH  Google Scholar 

  • Rozvany GIN, Zhou M, Birker T (1992) Generalized shape optimization without homogenization. Struct Multidisc Optim 4(3–4):250–252

    Google Scholar 

  • Sekimoto T, Noguchi H (2001) Homologous topology optimization in large displacement and buckling problems. JSME Int J 44:616–622

    Article  Google Scholar 

  • Seyranian AP, Lund E, Olhoff N (1994) Multiple eigenvalues in structural optimization problems. Struct Optim 8:207–227

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Taylor JE, Bendsøe MP (1984) An interpretation of min-max structural design problems including a method for relaxing constraints. Int J Solids Struct 20(4):301–314

    Article  MATH  Google Scholar 

  • Wilson EL, Itoh T (1983) An eigensolution strategy for large systems. Comput Struct 16:259–265

    Article  MATH  Google Scholar 

  • Wittrick WH (1962) Rates of change of eigenvalues, with reference to buckling and vibration problems. J R Aeronaut Soc 66:590–591

    Google Scholar 

Download references

Acknowledgments

The authors gratefully acknowledge the support from the Danish Center for Scientific Computing (DCSC) for the hybrid Linux Cluster “Fyrkat” at Aalborg University, Denmark.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Esben Lindgaard.

Appendix A: Design sensitivity analysis

Appendix A: Design sensitivity analysis

1.1 A.1 Sensitivity of linear displacements

The displacement sensitivities \(\frac{d\mathbf{D}}{d\rho_e}\) are computed by direct differentiation of the static equilibrium equation, (1), w.r.t. a design variable ρ e , e = 1, ..., N e .

$$ \mathbf{K_0} \frac{d\mathbf{D}}{d\rho_e}=-\frac{d\mathbf{K_0}}{d\rho_e}\mathbf{D} + \frac{d\mathbf{R}}{d\rho_e}, \quad e=1,\ldots,N_e $$
(20)

The displacement sensitivity \(\frac{d\mathbf{D}}{d\rho_e}\) can be evaluated by backsubstitution of the factored global initial stiffness matrix in (20). The initial stiffness matrix has already been factored when solving the static problem in (1) and can here be reused, whereby only the new terms on the right hand side of (20), called the pseudo load vector, need to be calculated. Note that the force vector derivative, \(\frac{d\mathbf{R}}{d\rho_e}\), is zero for design independent loads. The global initial stiffness matrix derivative \(\frac{d\mathbf{K_0}}{d\rho_e}\) only involve the derivative of the current local element stiffness matrix \(\frac{d\mathbf{k}_{0}}{d\rho_e}\) which is determined by central difference approximations at the element level. The element stiffness derivative could easily be evaluated analytically, however for ease of programming a semi-analytical approach has been used.

1.2 A.2 Sensitivity of nonlinear displacements

The sensitivities of nonlinear displacements are computed by considering the residual or force unbalance equation at a converged load step n,

$$ \mathbf{Q}^n(\mathbf{D}^n(\boldsymbol \rho),\boldsymbol \rho)=\mathbf{F}^n-\mathbf{R}^n=\boldsymbol 0 $$
(21)

where \(\mathbf{Q}^n(\mathbf{D}^n(\boldsymbol \rho),\boldsymbol \rho)\) is the so-called residual or force unbalance, F n is the global internal force vector, and R n is the global applied load vector. Taking the total derivative of this equilibrium equation with respect to any of the design variables ρ e ,e = 1, ..., N e , we obtain

$$ \frac{d\mathbf{Q}^n}{d\rho_e} = \frac{\partial\mathbf{Q}^n}{\partial\rho_e} +\frac{\partial\mathbf{Q}^n}{\partial\mathbf{D}^n}\frac{d\mathbf{D}^n}{d\rho_e}=\boldsymbol 0 $$
(22)
$$ \text{where \quad} \frac{\partial\mathbf{Q}^n}{\partial\mathbf{D}^n} = \frac{\partial\mathbf{F}^n}{\partial\mathbf{D}^n}-\frac{\partial\mathbf{R}^n}{\partial\mathbf{D}^n} $$
(23)
$$ \text{and \quad} \frac{\partial\mathbf{Q}^n}{\partial\rho_e} = \frac{\partial\mathbf{F}^n}{\partial\rho_e}-\frac{\partial\mathbf{R}^n}{\partial\rho_e} $$
(24)

We note that (23) reduces to the tangent stiffness matrix. Since it is assumed that the current load is independent of deformation, \(\frac{\partial\mathbf{R}^n}{\partial\mathbf{D}^n}=\boldsymbol 0\), we obtain

$$ \frac{\partial\mathbf{F}^n}{\partial\mathbf{D}^n}= \mathbf{K}_{\mathbf{T}}^n $$
(25)

By inserting the tangent stiffness and (24) into (22), we obtain the displacement sensitivities \(\frac{d\mathbf{D}^n}{d\rho_e}\) as

$$ \mathbf{K}_{\mathbf{T}}^n\frac{d\mathbf{D}^n}{d\rho_e}=\frac{\partial\mathbf{R}^n}{\partial\rho_e}-\frac{\partial\mathbf{F}^n}{\partial\rho_e} $$
(26)

The partial derivative of the load vector, \(\frac{\partial\mathbf{R}^n}{\partial\rho_e}\), can explicitly be expressed by two terms by taking the partial derivative of (6)

$$ \frac{\partial\mathbf{R}^n}{\partial\rho_e}=\gamma^n \frac{\partial\mathbf{R}}{\partial\rho_e} + \frac{\partial\gamma^n}{\partial\rho_e} \mathbf{R} $$
(27)

For design independent loads \(\frac{\partial\mathbf{R}}{\partial\rho_e}=\boldsymbol 0\) and for a fixed load level \(\frac{\partial\gamma^n}{\partial\rho_e}=0\). The pseudo load vector, i.e. the right hand side of (26), is determined at the element level by central difference approximations and assembled to global vector derivatives. Again, the partial derivative only involves the element which is associated with the current design variable.

1.3 A.3 Linear compliance

The design sensitivity of linear compliance is obtained by the adjoint approach, see e.g. Bendsøe and Sigmund (2003; Lund and Stegmann (2005). The sensitivity with respect to any design variable ρ e , e = 1, ..., N e is

$$ \frac{dC_L}{d\rho_e} = - \mathbf{D}^\text{T} \, \frac{d\mathbf{K_0}}{d\rho_e} \, \mathbf{D} $$
(28)

The global initial stiffness matrix derivatives \(\frac{d\mathbf{K_0}}{d\rho_e}\) are determined semi-analytically at the element level by central difference approximations and assembled to global matrix derivatives.

1.4 A.4 Nonlinear end compliance

The design sensitivity of nonlinear end compliance at a converged load step n with respect to any design variable, ρ e , e = 1,...,N e , is obtained by the adjoint approach, see e.g. Bendsøe and Sigmund (2003)

$$ \frac{dC_{GNL}}{d\rho_e}= \boldsymbol{\lambda}^\text{T} \frac{\partial\mathbf{Q}^n}{\partial\rho_e} = \boldsymbol{\lambda}^\text{T} \left( \frac{\partial\mathbf{F}^n}{\partial\rho_e}-\frac{\partial\mathbf{R}^n}{\partial\rho_e} \right) $$
(29)

Assuming the end load fixed and independent of design changes we have that \(\frac{\partial\mathbf{R}^n}{\partial\rho_e}=\boldsymbol 0\). The adjoint vector \(\boldsymbol{\lambda}\), which is not to be confused with the eigenvector, is obtained as the solution to the adjoint equation

$$ \mathbf{K}_{\mathbf{T}}^n \, \boldsymbol{\lambda}=-\mathbf{R}^n $$
(30)

The partial derivatives in the right hand side of (29) are determined at the element level by central difference approximations and assembled to global vector derivatives.

1.5 A.5 Linear buckling

The linear buckling load factor sensitivities are determined by

$$ \frac{d\lambda_j}{d\rho_e}={\phi}_j^T \left( \frac{d\mathbf{K_0}}{d\rho_e} + \lambda_j \frac{d\mathbf{K}_{\boldsymbol \sigma}}{d\rho_e} \right) {\phi}_j $$
(31)

where the eigenvalue problem in (2) has been differentiated with respect to any design variable, ρ e , e = 1, ..., N e , assuming that λ j is simple, see e.g. Courant and Hilbert (1953); Wittrick (1962). The global matrix derivatives of K 0 and \(\mathbf{K}_{\boldsymbol \sigma}\) are determined semi-analytically at the element level by central difference approximations and assembled to global matrix derivatives. The stress stiffness matrix is an implicit function of the displacement field, i.e. \(\mathbf{K}_{\boldsymbol \sigma}(\mathbf{D}(\boldsymbol{\rho}),\boldsymbol{\rho})\), and thus depends on all elements within the model. Both displacement field and design variables need to be perturbed in the element central difference approximation. The displacement field is perturbed via the calculated displacement sensitivities in (20) such that \(\Delta \mathbf{D} \approx \frac{d\mathbf{D}}{d\rho_e} \Delta \rho_e\).

1.6 A.6 Nonlinear buckling

The nonlinear buckling load factor sensitivities at load step n are determined by

$$ \frac{d\lambda_j}{d\rho_e}={\phi}_j^T \left( \frac{d\mathbf{K_0}}{d\rho_e} + \frac{d\mathbf{K}_{\mathbf{L}}^n}{d\rho_e} + \lambda_j \frac{d\mathbf{K}_{\boldsymbol \sigma}^n}{d\rho_e} \right) {\phi}_j $$
(32)

and

$$ \frac{d\gamma_j^c}{d\rho_e}=\frac{d\lambda_j}{d\rho_e} \, \gamma^n $$
(33)

where the eigenvalue problem in (11) has been differentiated with respect to any design variable, ρ e , e = 1, ..., N e , assuming that λ j is simple, see Lindgaard and Lund (2010). It is assumed that the final load level is fixed and that the nonlinear buckling load has been determined at load step n by evaluation of (10) and (11). The global matrix derivatives of K 0 , \(\mathbf{K}_{L}^n\), and \(\mathbf{K}_{\boldsymbol \sigma}^n\) are determined in the same manner as for the linear buckling load sensitivities, i.e. semi-analytical central difference approximations at the element level and assembly to global matrix derivatives. The displacement field is perturbed via the calculated sensitivities of the nonlinear displacements in (26) such that \(\Delta \mathbf{D}^n \approx \frac{d\mathbf{D}^n}{d\rho_e} \Delta \rho_e\).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Lindgaard, E., Dahl, J. On compliance and buckling objective functions in topology optimization of snap-through problems. Struct Multidisc Optim 47, 409–421 (2013). https://doi.org/10.1007/s00158-012-0832-2

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-012-0832-2

Keywords

Navigation