Skip to main content

Advertisement

Log in

Mixed-integer second-order cone optimization for composite discrete ply-angle and thickness topology optimization problems

  • Research Article
  • Published:
Optimization and Engineering Aims and scope Submit manuscript

Abstract

Discrete variable topology optimization problems are usually solved by using solid isotropic material with penalization (SIMP), genetic algorithms (GA), or mixed-integer nonlinear optimization (MINLO). In this paper, we propose formulating discrete ply-angle and thickness topology optimization problems as a mixed-integer second-order cone optimization (MISOCO) problem. Unlike SIMP and GA methods, MISOCO efficiently finds the problem’s globally optimal solution. Furthermore, in contrast with existing MISOCO formulations of discrete ply-angle optimization problems, our reformulations allow the structure to change topology, consider the more realistic Tsai–Wu stress yield criteria constraint, and eliminate checkerboard patterns using simple linear constraints. We address two types of discrete ply-angle and thickness problems: a structural mass minimization problem and a compliance optimization problem where the objective is to maximize the structural stiffness. For each element, one first chooses if the element is present or not in the structure. One can then choose the element’s ply-angle and thickness from a finite set of possibilities for the former case. The discrete design space for ply-angle and thickness is a result of manufacturing limitations. To improve the problem’s MISOCO solution approach, we develop valid inequality constraints to tighten the continuous relaxation of the MISOCO reformulation. We compare the performance of various MISOCO solvers: Gurobi, CPLEX, and MOSEK to solve the MISOCO reformulation. We also use BARON to solve the original MINLO formulations of the problems. Our results show that solving the MISOCO problem’s formulation using MOSEK is the most efficient solution approach.

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

Notes

  1. https://SichengHe@bitbucket.org/SichengHe/misoco_composite.git.

References

  • Achtziger W, Stolpe M (2009) Global optimization of truss topology with discrete bar areas—part II: implementation and numerical results. Comput Optim Appl 44(2):315–341

    MathSciNet  MATH  Google Scholar 

  • Albanesi A, Bre F, Fachinotti V, Gebhardt C (2018) Simultaneous ply-order, ply-number and ply-drop optimization of laminate wind turbine blades using the inverse finite element method. Compos Struct 184:894–903

    Google Scholar 

  • Andersen ED, Andersen KD (2000) The MOSEK interior point optimizer for linear programming: an implementation of the homogeneous algorithm. In: Frenk H, Roos K, Terlaky T, Zhang S (eds) Applied optimization. Springer, Berlin, pp 197–232

    Google Scholar 

  • Andersen ED, Roos K, Terlaky T (2003) On implementing a primal-dual interior-point method for conic quadratic optimization. Math Program 95(2):249–277

    MathSciNet  MATH  Google Scholar 

  • Belotti P, Góez JC, Pólik I, Ralphs TK, Terlaky T (2013) On families of quadratic surfaces having fixed intersections with two hyperplanes. Discrete Appl Math 161(16–17):2778–2793

    MathSciNet  MATH  Google Scholar 

  • Belotti P, Góez JC, Pólik I, Ralphs TK, Terlaky T (2017) A complete characterization of disjunctive conic cuts for mixed integer second order cone optimization. Discrete Optim 24:3–31

    MathSciNet  MATH  Google Scholar 

  • Bendsøe MP, Sigmund O (1999) Material interpolation schemes in topology optimization. Arch Appl Mech (Ingenieur Archiv) 69(9–10):635–654

    MATH  Google Scholar 

  • Bourdin B (2001) Filters in topology optimization. Int J Numer Methods Eng 50(9):2143–2158

    MathSciNet  MATH  Google Scholar 

  • Brampton CJ, Wu KC, Kim HA (2015) New optimization method for steered fiber composites using the level set method. Struct Multidiscip Optim 52(3):493–505

    MathSciNet  Google Scholar 

  • Brooks TR, Martins JRRA (2018) On manufacturing constraints for tow-steered composite design optimization. Compos Struct 204:548–559

    Google Scholar 

  • Bruyneel M (2011) SFP—a new parameterization based on shape functions for optimal material selection: application to conventional composite plies. Struct Multidiscip Optim 43(1):17–27

    Google Scholar 

  • Cheng G (1995) Some aspects of truss topology optimization. Struct Optim 10(3–4):173–179

    Google Scholar 

  • Cheng GD, Guo X (1997) \(\epsilon\)-relaxed approach in structural topology optimization. Struct Optim 13(4):258–266

    Google Scholar 

  • Díaz A, Sigmund O (1995) Checkerboard patterns in layout optimization. Structural Optimization 10(1):40–45

    Google Scholar 

  • Ghiasi H, Fayazbakhsh K, Pasini D, Lessard L (2010) Optimum stacking sequence design of composite materials part II: variable stiffness design. Compos Struct 93(1):1–13

    Google Scholar 

  • Gillet A, Francescato P, Saffre P (2010) Single- and multi-objective optimization of composite structures: the influence of design variables. J Compos Mater 44(4):457–480

    Google Scholar 

  • Grossmann IE, Voudouris V, Ghattas O (1992) Mixed-integer linear programming reformulations for some nonlinear discrete design optimization problems. In: Floudas CA, Pardalos PM (eds) Recent advances in global optimization. Princeton University Press, Princeton, pp 478–512

    Google Scholar 

  • Gurobi Optimization, LLC (2018) Gurobi optimizer reference manual

  • Haftka RT (1985) Simultaneous analysis and design. AIAA J 23(7):1099–1103

    MathSciNet  MATH  Google Scholar 

  • Haftka RT, Walsh JL (1992) Stacking-sequence optimization for buckling of laminated plates by integer programming. AIAA J 30(3):814–819

    Google Scholar 

  • Hahn HT, Tsai SW (1980) Introduction to composite materials. Technomic Publishing Co., Inc., New York

    Google Scholar 

  • Huang D, Friedmann PP (2016) An integrated aerothermoelastic analysis framework for predicting the response of composite panels. In: 15th dynamics specialists conference. American Institute of Aeronautics and Astronautics

  • IBM ILOG (2018) IBM ILOG CPLEX v12.8.0 user’s manual for CPLEX

  • Kennedy GJ (2016) A full-space barrier method for stress-constrained discrete material design optimization. Struct Multidiscip Optim 54(3):619–639

    MathSciNet  Google Scholar 

  • Kennedy GJ, Martins JRRA (2013) A laminate parametrization technique for discrete ply-angle problems with manufacturing constraints. Struct Multidiscip Optim 48(2):379–393

    MathSciNet  Google Scholar 

  • Kim JS, Kim CG, Hong CS (1999) Optimum design of composite structures with ply drop using genetic algorithm and expert system shell. Compos Struct 46(2):171–187

    Google Scholar 

  • Lobo MS, Vandenberghe L, Boyd S, Lebret H (1998) Applications of second-order cone programming. Linear Algebra Appl 284(1–3):193–228

    MathSciNet  MATH  Google Scholar 

  • Marmaras K (2014) Optimal design of composite structures under manufacturing constraints. PhD thesis, Technical University of Denmark

  • Martins JRRA, Lambe AB (2013) Multidisciplinary design optimization: a survey of architectures. AIAA J 51(9):2049–2075

    Google Scholar 

  • McCormick GP (1976) Computability of global solutions to factorable nonconvex programs: part I—convex underestimating problems. Math Program 10(1):147–175

    MATH  Google Scholar 

  • Mela K (2014) Resolving issues with member buckling in truss topology optimization using a mixed variable approach. Struct Multidiscip Optim 50(6):1037–1049

    MathSciNet  Google Scholar 

  • Munoz E (2010) Global optimization for discrete topology design problems by generalized Benders’s decomposition. PhD thesis, Technical University of Denmark

  • Nikbakt S, Kamarian S, Shakeri M (2018) A review on optimization of composite structures part I: laminated composites. Compos Struct 195:158–185

    Google Scholar 

  • Petersen CC (1971) A note on transforming the product of variables to linear form in linear programs. Working paper, Purdue University

  • Rasmussen MH, Stolpe M (2008) Global optimization of discrete truss topology design problems using a parallel cut-and-branch method. Comput Struct 86(13–14):1527–1538

    Google Scholar 

  • Riche RL, Haftka RT (1993) Optimization of laminate stacking sequence for buckling load maximization by genetic algorithm. AIAA J 31(5):951–956

    MATH  Google Scholar 

  • Sahinidis NV (2017) BARON 17.8.9: global optimization of mixed-integer nonlinear programs. User’s manual

  • Shahabsafa M, Mohammad-Nezhad A, Terlaky T, Zuluaga L, He S, Hwang JT, Martins JRRA (2018) A novel approach to discrete truss design problems using mixed integer neighborhood search. Struct Multidiscip Optim 58(6):2411–2429

    MathSciNet  Google Scholar 

  • Shahabsafa M, Fakhimi R, Lei W, He S, Martins JRRA, Terlaky T, Zuluaga LF (2020) Truss topology design and sizing optimization with guaranteed kinematic stability. Struct Multidiscip Optim. https://doi.org/10.1007/s00158-020-02698-x

    Article  Google Scholar 

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

    Google Scholar 

  • Sigmund O, Petersson J (1998) Numerical instabilities in topology optimization: a survey on procedures dealing with checkerboards, mesh-dependencies and local minima. Struct Optim 16(1):68–75

    Google Scholar 

  • Sørensen R, Lund E (2015) Thickness filters for gradient based multi-material and thickness optimization of laminated composite structures. Struct Multidiscip Optim 52(2):227–250

    Google Scholar 

  • Sørensen SN, Stolpe M (2015) Global blending optimization of laminated composites with discrete material candidate selection and thickness variation. Struct Multidiscip Optim 52(1):137–155

    MathSciNet  Google Scholar 

  • Stegmann J, Lund E (2005) Discrete material optimization of general composite shell structures. Int J Numer Methods Eng 62(14):2009–2027

    MATH  Google Scholar 

  • Stolpe M (2007) On the reformulation of topology optimization problems as linear or convex quadratic mixed 0–1 programs. Optim Eng 8(2):163–192

    MathSciNet  MATH  Google Scholar 

  • Stolpe M (2014) Truss topology optimization with discrete design variables by outer approximation. J Glob Optim 61(1):139–163

    MathSciNet  MATH  Google Scholar 

  • Stolpe M, Stidsen T (2007) A hierarchical method for discrete structural topology design problems with local stress and displacement constraints. Int J Numer Methods Eng 69(5):1060–1084

    MathSciNet  MATH  Google Scholar 

  • Stolpe M, Svanberg K (2003) Modelling topology optimization problems as linear mixed 0–1 programs. Int J Numer Methods Eng 57(5):723–739

    MathSciNet  MATH  Google Scholar 

  • Sved G, Ginos Z (1968) Structural optimization under multiple loading. Int J Mech Sci 10(10):803–805

    Google Scholar 

  • Tawarmalani M, Sahinidis NV (2005) A polyhedral branch-and-cut approach to global optimization. Math Program 103(2):225–249

    MathSciNet  MATH  Google Scholar 

  • Tsai SW, Wu EM (1971) A general theory of strength for anisotropic materials. J Compos Mater 5(1):58–80

    Google Scholar 

  • Zhou Y, Nomura T, Saitou K (2018) Multi-component topology and material orientation design of composite structures (MTO-C). Comput Methods Appl Mech Eng 342:438–457

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This research was supported by Air Force Office of Scientific Research Grant No. FA9550-15-1-0222. We thank Prof. Sahinidis for providing us a license for the BARON software.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sicheng He.

Additional information

Publisher's Note

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

Appendix

Appendix

1.1 Constitutive matrix calculation

1.1.1 In-plane constitutive relation

In this section, we present the constitutive matrix \({\mathbf {Q}}_i\) derivation procedures. In a frame that the x axis is aligned with the fiber direction, we have the following constitutive relationships:

$$\begin{aligned} \begin{bmatrix} \sigma _1\\ \sigma _2\\ \tau _{12} \end{bmatrix} = {\mathbf {Q}} \begin{bmatrix} \epsilon _1\\ \epsilon _2\\ \epsilon _{12} \end{bmatrix}, \end{aligned}$$

where \({\mathbf {Q}}\) is given as

$$\begin{aligned} {\mathbf {Q}}= \begin{bmatrix} Q_{11} &{} \quad Q_{12} &{} \quad 0\\ Q_{21} &{} \quad Q_{22} &{} \quad 0\\ 0 &{} \quad 0 &{} \quad 2Q_{66}\\ \end{bmatrix}, \end{aligned}$$

and

$$\begin{aligned} Q_{11}& = \frac{E_1}{1-\nu _{21}\nu _{12}}\\ Q_{12}& = \frac{\nu _{12} E_2}{1-\nu _{21}\nu _{12}}=Q_{21}\\ Q_{22}& = \frac{E_2}{1-\nu _{21}\nu _{12}}\\ Q_{66}& = G_{12}. \end{aligned}$$

For more general cases where the x axis and the fiber direction are with an angle of \(\theta\), using the laws of tensor transformation relations, we have:

$$\begin{aligned} \begin{bmatrix} \sigma _x\\ \sigma _y\\ \tau _{xy} \end{bmatrix} = {\mathbf {T}}^{-1}(\theta ) \begin{bmatrix} \sigma _1\\ \sigma _2\\ \tau _{12} \end{bmatrix}, \end{aligned}$$

where

$$\begin{aligned} {\mathbf {T}}^{-1}(\theta )= \begin{bmatrix} \cos ^2\theta &{} \quad \sin ^2\theta &{} \quad -2\cos \theta \sin \theta \\ \sin ^2\theta &{} \quad \cos ^2\theta &{} \quad 2\cos \theta \sin \theta \\ \cos \theta \sin \theta &{} \quad -\cos \theta \sin \theta &{} \quad \cos ^2\theta - \sin ^2\theta \end{bmatrix}, \end{aligned}$$

and

$$\begin{aligned} \begin{bmatrix} \epsilon _1\\ \epsilon _2\\ \epsilon _{12} \end{bmatrix} = {\mathbf {T}}(\theta ) \begin{bmatrix} \epsilon _x\\ \epsilon _y\\ \epsilon _{xy} \end{bmatrix}. \end{aligned}$$

Thus, the general constitutive law is given as:

$$\begin{aligned} \begin{bmatrix} \sigma _x\\ \sigma _y\\ \tau _{xy} \end{bmatrix} = {\mathbf {T}}^{-1}(\theta ) {\mathbf {Q}} {\mathbf {T}}(\theta ) \begin{bmatrix} \epsilon _x\\ \epsilon _y\\ \epsilon _{xy} \end{bmatrix}. \end{aligned}$$

For the jth choice with the angle as \(\theta _j\), we define:

$$\begin{aligned} {\mathbf {Q}}_j={\mathbf {T}}^{-1}(\theta _j) {\mathbf {Q}} {\mathbf {T}}(\theta _j). \end{aligned}$$

We can give an explicit expression for \({\mathbf {Q}}_j\) by expanding the previous equation:

$$\begin{aligned} Q_{j, 11}& = Q_{11}\cos ^4\theta _j + Q_{22}\sin ^4\theta _j+2(Q_{12}+2Q_{66})\sin ^2\theta _j\cos ^2\theta _j\\ Q_{j, 12}& = (Q_{11}+Q_{22}-4Q_{66})\cos ^2\theta _j\sin ^2\theta _j+Q_{12}(\sin ^4\theta _j+\cos ^4\theta _j)\\ Q_{j, 22}& = Q_{11}\sin ^4\theta _j + Q_{22}\cos ^4\theta _j+2(Q_{12}+2Q_{66})\sin ^2\theta _j\cos ^2\theta _j\\ Q_{j, 16}& = (Q_{11}-Q_{12}-2Q_{66})\cos ^3\theta _j\sin \theta _j-(Q_{22}-Q_{12}-2Q_{66})\cos \theta _j\sin ^3\theta _j\\ Q_{j, 26}& = (Q_{11}-Q_{12}-2Q_{66})\cos \theta _j\sin ^3\theta _j-(Q_{22}-Q_{12}-2Q_{66})\cos ^3\theta _j\sin \theta _j\\ Q_{j, 66}& = (Q_{11}+Q_{22}-2Q_{12}-2Q_{66})\cos ^2\theta _j\sin ^2\theta _j+Q_{66}(\sin ^4\theta _j+\cos ^4\theta _j).\\ \end{aligned}$$

1.1.2 Out-of-plane constitutive relation

The out-of-plane constitutive relationship is given as

$$\begin{aligned} \begin{bmatrix} \sigma _{yz}\\ \sigma _{xz} \end{bmatrix} = {\mathbf {C}}_j \begin{bmatrix} \psi _y + \frac{\partial w}{\partial y}\\ \psi _x + \frac{\partial w}{\partial x}, \end{bmatrix}, \end{aligned}$$

where

$$\begin{aligned} {\mathbf {C}}_j= \begin{bmatrix} C_{j, 44} &{} \quad C_{j, 45}\\ C_{j, 45} &{} \quad C_{j, 55} \end{bmatrix}, \end{aligned}$$

and

$$\begin{aligned} C_{j, 44}& = C_{44}\cos ^2\theta _j + C_{55}\sin ^2\theta _j\\ C_{j, 45}& = (C_{55} - C_{44})\cos \theta _j\sin \theta _j, \\ C_{j, 55}& = C_{44}\sin ^2\theta _j + C_{55}\cos ^2\theta _j. \end{aligned}$$

We also know that \(C_{44}=G_{23}\) and \(C_{55}=G_{12}\).

1.2 Strain operator

We calculate the strain of the upper surface center for each element that is later used in Tsai–Wu failure criterion. The in-plane strain is given as:

$$\begin{aligned} \begin{aligned} \pmb {\epsilon }&= \pmb {\epsilon }_l + z{\chi }, \\ \pmb {\epsilon }_l&={\mathbf {L}}_1{\mathbf {u}}, \\ {\chi }&={\mathbf {L}}_2{\mathbf {u}}, \end{aligned} \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} {\mathbf {L}}_1&=\left[ \begin{array}{c|ccccc|c} &{} \quad N_{i,x} &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad \\ \dots &{} \quad 0 &{} \quad N_{i,y} &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad \dots \\ &{} \quad N_{i,y} &{} \quad N_{i,x} &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad \\ \end{array}\right] ,\\ {\mathbf {L}}_2&=\left[ \begin{array}{c|ccccc|c} &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad N_{i,x} &{} \quad 0 &{} \quad \\ \dots &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad N_{i,y} &{} \quad \dots \\ &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad N_{i,y} &{} \quad N_{i,x} &{} \quad \\ \end{array}\right] .\\ \end{aligned} \end{aligned}$$

For the shape function \(N_i\), we have

$$\begin{aligned} \begin{aligned} N_{i+3j-3}&= f_i(\eta ) f_j(\xi ), \,\, i,j=1,2,3,\\ f_1(x)&= \frac{1}{2}x(x-1),\\ f_2(x)&= 1-x^2,\\ f_3(x)&= \frac{1}{2}(1+x)x. \end{aligned} \end{aligned}$$

To evaluate the strain of the center point over the top surface, we simply set \((\eta , \xi , z)=(0, 0, h/2)\). Finally,

$$\begin{aligned} \pmb {\epsilon }_i={\mathbf {B}}({\mathbf {x}}_i){\mathbf {u}}_i =\left( {\mathbf {L}}_1(0)+\frac{h}{2}{\mathbf {L}}_2(0)\right) {\mathbf {u}}_i. \end{aligned}$$

1.3 Local stiffness matrix calculation

In this section, we present the derivation of the local stiffness matrix \({\mathbf {K}}_{ij}\) derivation, which is based on classic lamination theory (CLT). We begin the derivation by giving the expression for matrices \({\mathbf {A}}, {\mathbf {B}}, {\mathbf {D}}\) representing the extensional stiffness matrix, the coupling stiffness matrix and the bending stiffness matrix, respectively:

$$\begin{aligned} \begin{bmatrix} {\mathbf {A}}&\quad {\mathbf {B}}&\quad {\mathbf {D}} \end{bmatrix} = \int _{-h_i/2}^{h_i/2} \begin{bmatrix} 1&\quad z&\quad z^2 \end{bmatrix} {\mathbf {Q}}_j dz. \end{aligned}$$

The traverse shear stiffness matrix is given as:

$$\begin{aligned} {\mathbf {S}}=\kappa \int _{-h_i/2}^{h_i/2}{\mathbf {C}}_j dz, \end{aligned}$$

where \(\kappa =\frac{5}{6}\). Notice that to be consistent with the assumption made in this paper here we assume that there is only one layer of material.

We now present the stiffness matrix expression:

$$\begin{aligned} \begin{aligned} {\mathbf {K}}_{ij}&= {\mathbf {K}}_{ij,L_1} + {\mathbf {K}}_{ij,L_2}\\&=\int _{A_i} \left( L_1^\intercal {\mathbf {A}} L_1 +L_1^\intercal {\mathbf {B}}L_2 + L_2^\intercal {\mathbf {B}}L_1 + L_2^\intercal {\mathbf {D}} L_2 \right) dA_i + \int _{A_i} \left( L_3^\intercal {\mathbf {S}}L_3\right) dA_i \end{aligned} \end{aligned}$$

Then we apply the Gaussian quadrature for numerical integration. For \(K_{ij,L_1}\), we apply \(3\times 3\) integration points; while for \(K_{ij,L_2}\), we apply \(2\times 2\) integration points. This helps to avoid “shear locking”.

1.4 A brief overview of SOC

A SOC constraint can be expressed in the following format (Lobo et al. 1998)

$$\begin{aligned} \Vert {\mathbf {A}} {\mathbf {x}} + {\mathbf {b}} \Vert \le {\mathbf {c}}^\top {\mathbf {x}} + d, \end{aligned}$$
(27)

where \(\Vert {\cdot }\Vert\) denotes the Euclidean norm, \({\mathbf {x}}\in {{\mathbb {R}}}^n\) is the variable vector, \({\mathbf {A}} \in {{\mathbb {R}}}^{m \times n}\), \({\mathbf {b}}\in {{\mathbb {R}}}^{m}\), \({\mathbf {c}}\in {{\mathbb {R}}}^n\), and \(d\in {{\mathbb {R}}}\).

A SOC constraint appears naturally when the feasible set of a convex quadratic function is considered. In particular, let \({\mathbf {Q}}\) be PSD, then

$$\begin{aligned} \{(y, {\mathbf {x}})\in {{\mathbb {R}}}\times {{\mathbb {R}}}^n: y^2 \ge {\mathbf {x}}^\intercal {\mathbf {Q}} {\mathbf {x}}\}, \end{aligned}$$
(28)

is equal to the following set

$$\begin{aligned} \{(y, {\mathbf {x}})\in {{\mathbb {R}}}\times {{\mathbb {R}}}^n: \Vert {\mathbf {Q}}^{1/2}{\mathbf {x}}\Vert \le y\}. \end{aligned}$$
(29)

Many more examples, including truss topology optimization problems, showing how SOC constraints arise in practice are discussed by Lobo et al. (1998).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

He, S., Shahabsafa, M., Lei, W. et al. Mixed-integer second-order cone optimization for composite discrete ply-angle and thickness topology optimization problems. Optim Eng 22, 1589–1624 (2021). https://doi.org/10.1007/s11081-020-09573-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11081-020-09573-0

Keywords

Navigation