Skip to main content
Log in

Topology optimization considering fracture mechanics behaviors at specified locations

  • RESEARCH PAPER
  • Published:
Structural and Multidisciplinary Optimization Aims and scope Submit manuscript

Abstract

As a typical form of material imperfection, cracks generally cannot be avoided and are critical for load bearing capability and integrity of engineering structures. This paper presents a topology optimization method for generating structural layouts that are insensitive/sensitive as required to initial cracks at specified locations. Based on the linear elastic fracture mechanics model (LEFM), the stress intensity of initial cracks in the structure is analyzed by using singularity finite elements positioned at the crack tip to describe the near-tip stress field. In the topology optimization formulation, the J integral, as a criterion for predicting crack opening under certain loading and boundary conditions, is introduced into the objective function to be minimized or maximized. In this context, the adjoint variable sensitivity analysis scheme is derived, which enables the optimization problem to be solved with a gradient-based algorithm. Numerical examples are given to demonstrate effectiveness of the proposed method on generating structures with desired overall stiffness and fracture strength property. This method provides an applicable framework incorporating linear fracture mechanics criteria into topology optimization for conceptual design of crack insensitive or easily detachable structures for particular applications.

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
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27
Fig. 28
Fig. 29
Fig. 30
Fig. 31

Similar content being viewed by others

References

  • Allaire G, Jouve F (2008) Minimum stress optimal design with the level set method. Eng Ana Bound Elem 32(11):909–918

    Article  MATH  Google Scholar 

  • Allaire G, Jouve F, Toader A-M (2004) Structural optimization using sensitivity analysis and a level-set method. J Comput Phys 194(1):363–393

    Article  MathSciNet  MATH  Google Scholar 

  • Amir O (2013) A topology optimization procedure for reinforced concrete structures. Comput Struct 114:46–58

    Article  Google Scholar 

  • Amir O, Sigmund O (2013) Reinforcement layout design for concrete structures based on continuum damage and truss topology optimization. Struct Multidiscip Optim 47(2):157–174

    Article  MathSciNet  MATH  Google Scholar 

  • Anderson TL, Anderson T (2005) Fracture mechanics: fundamentals and applications. CRC press

  • Banichuk N, Ragnedda F, Serra M (2006) Axisymmetric shell optimization under fracture mechanics and geometric constraint. Struct Multidiscip Optim 31(3):223–228

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  • Bendsøe MP, Díaz AR (1998) A method for treating damage related criteria in optimal topology design of continuum structures. Struct Optim 16(2–3):108–115

    Article  Google Scholar 

  • Bendsøe MP, Kikuchi N (1988) Generating optimal topologies in structural design using a homogenization method. Comput Methods Appl Mech Eng 71(2):197–224

    Article  MathSciNet  MATH  Google Scholar 

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

    MATH  Google Scholar 

  • Bruggi M (2008) On an alternative approach to stress constraints relaxation in topology optimization. Struct Multidiscip Optim 36(2):125–141

    Article  MathSciNet  MATH  Google Scholar 

  • Bruggi M, Duysinx P (2012) Topology optimization for minimum weight with compliance and stress constraints. Struct Multidiscip Optim 46(3):369–384

    Article  MathSciNet  MATH  Google Scholar 

  • Carlson A, Bowen AM, Huang Y, Nuzzo RG, Rogers JA (2012) Transfer printing techniques for materials assembly and micro/nanodevice fabrication. Adv Mater 24(39):5284–5318

    Article  Google Scholar 

  • Challis VJ, Roberts AP, Wilkins AH (2008) Fracture resistance via topology optimization. Struct Multidiscip Optim 36(3):263–271

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  • Davis M, Bond D (1999) Principles and practices of adhesive bonded structural joints and repairs. Int J Adhes Adhes 19(2):91–105

    Article  Google Scholar 

  • Deaton JD, Grandhi RV (2014) A survey of structural and multidisciplinary continuum topology optimization: post 2000. Struct Multidiscip Optim 49(1):1–38

    Article  MathSciNet  Google Scholar 

  • Desmorat B, Desmorat R (2008) Topology optimization in damage governed low cycle fatigue. Comptes Rendus Mecan 336(5):448–453

    Article  MATH  Google Scholar 

  • Duysinx P, Bendsøe MP (1998) Topology optimization of continuum structures with local stress constraints. Int J Numer Methods Eng 43(8):1453–1478

    Article  MathSciNet  MATH  Google Scholar 

  • Edke MS, Chang K-H (2011) Shape optimization for 2-D mixed-mode fracture using Extended FEM (XFEM) and Level Set Method (LSM). Struct Multidiscip Optim 44(2):165–181

    Article  Google Scholar 

  • Eschenauer H, Kobelev V (1992) Structural analysis and optimization modelling including fracture conditions. Int J Numer Methods Eng 34(3):873–888

    Article  MATH  Google Scholar 

  • Guo X, Zhang WS, Wang MY, Wei P (2011) Stress-related topology optimization via level set approach. Comput Methods Appl Mech Eng 200(47):3439–3452

    Article  MathSciNet  MATH  Google Scholar 

  • Henshell R, Shaw K (1975) Crack tip finite elements are unnecessary. Int J Numer Methods Eng 9(3):495–507

    Article  MATH  Google Scholar 

  • James KA, Waisman H (2013) Failure mitigation in optimal topology design using a coupled nonlinear continuum damage model. Com Methods Appl Mech Eng

  • Jansen M, Lombaert G, Schevenels M, Sigmund O (2014) Topology optimization of fail-safe structures using a simplified local damage model. Struct Multidis Optim 49(4):657–666

  • Le C, Norato J, Bruns T, Ha C, Tortorelli D (2010) Stress-based topology optimization for continua. Struct Multidiscip Optim 41(4):605–620

    Article  Google Scholar 

  • Lund E (1998) Shape optimization using Weibull statistics of brittle failure. Struct Optim 15(3–4):208–214

    Article  Google Scholar 

  • Luo Y, Kang Z (2012) Topology optimization of continuum structures with Drucker–Prager yield stress constraints. Comput Struct 90:65–75

    Article  Google Scholar 

  • Luo Y, Wang MY, Kang Z (2013) An enhanced aggregation method for topology optimization with local stress constraints. Comput Methods Appl Mech Eng 254:31–41

    Article  MathSciNet  MATH  Google Scholar 

  • Papila M, Haftka R (2003) Implementation of a crack propagation constraint within a structural optimization software. Struct Multidiscip Optim 25(5–6):327–338

    Article  Google Scholar 

  • Peng D, Jones R (2008) An approach based on biological algorithm for three-dimensional shape optimisation with fracture strength constrains. Comput Methods Appl Mech Eng 197(49):4383–4398

    Article  MATH  Google Scholar 

  • Potter K (2012) Resin transfer moulding. Springer Science & Business Media

  • Rice J (1968) A Path Independent Integral and the Approximate Analysis of Strain Concentration by Notches and Cracks. J Appl Mech 35:379

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Serra M (2000) Optimum beam design based on fatigue crack propagation. Struct Multidiscip Optim 19(2):159–163

    Article  Google Scholar 

  • Sigmund O (2001) A 99 line topology optimization code written in Matlab. Struct Multidiscip Optim 21(2):120–127

    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 

  • Thomas S, Mhaiskar M, Sethuraman R (2000) Stress intensity factors for circular hole and inclusion using finite element alternating method. Theor Appl Fract Mech 33(2):73–81

    Article  Google Scholar 

  • Wang MY, Wang X, Guo D (2003) A level set method for structural topology optimization. Comput Methods Appl Mech Eng 192(1):227–246

    Article  MathSciNet  MATH  Google Scholar 

  • Xie Y, Steven GP (1993) A simple evolutionary procedure for structural optimization. Comput Struct 49(5):885–896

    Article  Google Scholar 

Download references

Acknowledgements

The support of the Natural Science Foundation of China (U1508209, 11425207, 11302039) is gratefully acknowledged. The authors would like to thank Prof. Krister Svanberg for providing the source code of the MMA algorithm.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zhan Kang.

Appendix A

Appendix A

Details of derivation of discrete form of the J integral expression are as follows.

In a discrete form, the equilibrium equation reads

$$ \mathbf{K}\mathbf{d}=\mathbf{p}, $$
(A.1)

where d is the nodal displacement vector, p is the external force vector and K is the global stiffness matrix, which is expressed by

$$ \mathbf{K}={\displaystyle \sum_{i=1}^N{{\mathbf{G}}_i}^{\mathrm{T}}\left({\displaystyle {\int}_{\Omega_i}{{\mathbf{B}}_i}^{\mathrm{T}}{\mathbf{D}}_i{\mathbf{B}}_i\mathrm{d}\Omega}\right){\mathbf{G}}_i}. $$
(A.2)

After solving Equation A.1, the J integral can be calculated along a path composed entirely of element edges, as expressed in the matrix form

$$ J={\displaystyle \sum_{p=1}^M{\displaystyle {\int}_{\varGamma_p}\left(\frac{1}{2}{\boldsymbol{\upvarepsilon}}^{\mathrm{T}}\mathbf{D}\boldsymbol{\upvarepsilon } \mathrm{d}y-{\mathbf{T}}^{\mathrm{T}}\frac{\partial \mathbf{u}}{\partial x}\mathrm{d}s\right)}}, $$
(A.3)

where M is the total number of element edges which constitute the path of integral and Γ p is the p th element edge, T is the vector of traction on the contour and u is the displacement vector. For the element edges at which the strain and stress are discontinuous, the strain and traction are taken as their average values of two neighboring elements (for the pth element edge shown in Fig. 32, the two neighboring elements are EL  i p (inside the contour) and EL  o p (outside the contour)).

Fig. 32
figure 32

A schematic illustration of a finite element model of a crack

The first term of the J integral in Equation A.3 can be further written as

$$ \begin{array}{c}\hfill {J}_{{\mathrm{T}}_{\mathrm{I}}}=\frac{1}{8}{\displaystyle \sum_{p=1}^M{\displaystyle {\int}_{\varGamma_p}{\left({\mathbf{B}}_p^{\mathrm{i}}{\mathbf{d}}_p^{\mathrm{i}}+{\mathbf{B}}_p^{\mathrm{o}}{\mathbf{d}}_p^{\mathrm{o}}\right)}^{\mathrm{T}}{\mathbf{D}}_p\left({\mathbf{B}}_p^{\mathrm{i}}{\mathbf{d}}_p^{\mathrm{i}}+{\mathbf{B}}_p^{\mathrm{o}}{\mathbf{d}}_p^{\mathrm{o}}\right)\mathrm{d}y}}\hfill \\ {}\hfill ={\mathbf{d}}^{\mathrm{T}}\left({\mathbf{K}}_{\mathrm{I}}^{\mathrm{i}\mathrm{i}}+{\mathbf{K}}_{\mathrm{I}}^{\mathrm{o}\mathrm{o}}+{\mathbf{K}}_{\mathrm{I}}^{\mathrm{i}\mathrm{o}}\right)\mathbf{d}\hfill \\ {}\hfill ={\mathbf{d}}^{\mathrm{T}}{\mathbf{K}}_{\mathrm{I}}\mathbf{d}\hfill \end{array}, $$
(A.4)

where the superscripts “i” and “o” indicate the inner and outer side of the contour, respectively as shown in Fig. 32; B i p and B o p are the displacement–strain matrix of the two neighboring elements on the inner and outer side of the p th element edge, respectively. The expressions of K iiI , K ooI , K ioI are given as

$$ \left\{\begin{array}{l}{\mathbf{K}}_{\mathrm{I}}^{ii}=\frac{1}{8}{\displaystyle \sum_{p=1}^M{{\mathbf{G}}_p^{\mathrm{i}}}^{\mathrm{T}}\left({\displaystyle {\int}_{\varGamma_p}{{\mathbf{B}}_p^{\mathrm{i}}}^{\mathrm{T}}\mathbf{D}{\mathbf{B}}_p^{\mathrm{i}}\mathrm{d}y}\right){\mathbf{G}}_p^{\mathrm{i}}}\\ {}{\mathbf{K}}_{\mathrm{I}}^{\mathrm{o}\mathrm{o}}=\frac{1}{8}{\displaystyle \sum_{p=1}^M{{\mathbf{G}}_p^{\mathrm{o}}}^{\mathrm{T}}\left({\displaystyle {\int}_{\varGamma_p}{{\mathbf{B}}_p^{\mathrm{o}}}^{\mathrm{T}}\mathbf{D}{\mathbf{B}}_p^{\mathrm{o}}\mathrm{d}y}\right){\mathbf{G}}_p^{\mathrm{o}}}\\ {}{\mathbf{K}}_{\mathrm{I}}^{\mathrm{i}\mathrm{o}}=\frac{1}{4}{\displaystyle \sum_{p=1}^M{{\mathbf{G}}_p^{\mathrm{i}}}^{\mathrm{T}}\left({\displaystyle {\int}_{\varGamma_p}{{\mathbf{B}}_p^{\mathrm{i}}}^{\mathrm{T}}\mathbf{D}{\mathbf{B}}_p^{\mathrm{o}}\mathrm{d}y}\right){\mathbf{G}}_p^{\mathrm{o}}}\end{array}\right. $$
(A.5)

Similarly, the second term of the J integral is expressed as

$$ \begin{array}{c}\hfill {J}_{{\operatorname{T}}_{II}}=-\frac{1}{4}{\displaystyle \sum_{p=1}^M{\displaystyle {\int}_{\varGamma_p}{\left({\mathbf{B}}_p^{\mathrm{i}}{\mathbf{d}}_p^{\mathrm{i}}+{\mathbf{B}}_p^{\mathrm{o}}{\mathbf{d}}_p^{\mathrm{o}}\right)}^{\mathrm{T}}{\mathbf{D}}_p\mathbf{n}\left(\frac{\partial {\mathbf{N}}_p^{\mathrm{i}}}{\partial x}{\mathbf{d}}_p^{\mathrm{i}}+\frac{\partial {\mathbf{N}}_p^{\mathrm{o}}}{\partial x}{\mathbf{d}}_p^{\mathrm{o}}\right)\mathrm{d}s}}\hfill \\ {}\hfill =-{\mathbf{d}}^{\mathrm{T}}\left({\mathbf{K}}_{II}^{ii}+{\mathbf{K}}_{II}^{oo}+{\mathbf{K}}_{II}^{io}+{\mathbf{K}}_{II}^{oi}\right)\mathbf{d}\hfill \\ {}\hfill =-{\mathbf{d}}^{\mathrm{T}}{\mathbf{K}}_{\mathrm{II}}\mathbf{d}\hfill \end{array}, $$
(A.6)

where N is the shape function matrix; n is a matrix that consists of components of the direction vector of the integration contour; the expressions of n, K ii II , K oo II , K io II and K oi II are given as

$$ \mathbf{n}=\left[\begin{array}{l}{n}_1\kern0.5em 0\\ {}0\kern0.75em {n}_2\\ {}{n}_2\kern0.5em {n}_1\end{array}\right], $$
(A.7)
$$ \left\{\begin{array}{l}{\mathbf{K}}_{II}^{ii}=\frac{1}{4}{\displaystyle \sum_{p=1}^M{{\mathbf{G}}_p^{\mathrm{i}}}^{\mathrm{T}}\left({\displaystyle {\int}_{\varGamma_p}{{\mathbf{B}}_p^{\mathrm{i}}}^{\mathrm{T}}\mathbf{D}\mathbf{n}\frac{\partial {\mathbf{N}}_p^{\mathrm{i}}}{\partial x}\mathrm{d}s}\right){\mathbf{G}}_p^{\mathrm{i}}}\\ {}{\mathbf{K}}_{II}^{oo}=\frac{1}{4}{\displaystyle \sum_{p=1}^M{{\mathbf{G}}_p^{\mathrm{o}}}^{\mathrm{T}}\left({\displaystyle {\int}_{\varGamma_p}{{\mathbf{B}}_p^{\mathrm{o}}}^{\mathrm{T}}\mathbf{D}\mathbf{n}\frac{\partial {\mathbf{N}}_p^{\mathrm{o}}}{\partial x}\mathrm{d}s}\right){\mathbf{G}}_p^{\mathrm{o}}}\\ {}{\mathbf{K}}_{II}^{io}=\frac{1}{4}{\displaystyle \sum_{p=1}^M{{\mathbf{G}}_p^{\mathrm{i}}}^{\mathrm{T}}\left({\displaystyle {\int}_{\varGamma_p}{{\mathbf{B}}_p^{\mathrm{i}}}^{\mathrm{T}}\mathbf{D}\mathbf{n}\frac{\partial {\mathbf{N}}_p^{\mathrm{o}}}{\partial x}\mathrm{d}s}\right){\mathbf{G}}_p^{\mathrm{o}}}\\ {}{\mathbf{K}}_{II}^{oi}=\frac{1}{4}{\displaystyle \sum_{p=1}^M{{\mathbf{G}}_p^{\mathrm{o}}}^{\mathrm{T}}\left({\displaystyle {\int}_{\varGamma_p}{{\mathbf{B}}_p^{\mathrm{o}}}^{\mathrm{T}}\mathbf{D}\mathbf{n}\frac{\partial {\mathbf{N}}_p^{\mathrm{i}}}{\partial x}\mathrm{d}s}\right){\mathbf{G}}_p^{\mathrm{i}}}\end{array}\right. $$
(A.8)

Finally, the J integral is expressed in matrix form as

$$ J={J}_{{\operatorname{T}}_I}+{J}_{{\operatorname{T}}_{II}}={\mathbf{d}}^{\mathrm{T}}\left({\mathbf{K}}_{\mathrm{I}}-{\mathbf{K}}_{\mathrm{I}\mathrm{I}}\right)\mathbf{d}. $$
(A.9)

It is easy to prove that K I is symmetric and K II is asymmetric.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kang, Z., Liu, P. & Li, M. Topology optimization considering fracture mechanics behaviors at specified locations. Struct Multidisc Optim 55, 1847–1864 (2017). https://doi.org/10.1007/s00158-016-1623-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-016-1623-y

Keywords

Navigation