Skip to main content
Log in

Topology optimization with local stress constraints: a stress aggregation-free approach

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

Abstract

This paper presents a consistent topology optimization formulation for mass minimization with local stress constraints by means of the augmented Lagrangian method. To solve problems with a large number of constraints in an effective way, we modify both the penalty and objective function terms of the augmented Lagrangian function. The modification of the penalty term leads to consistent solutions under mesh refinement and that of the objective function term drives the mass minimization towards black and white solutions. In addition, we introduce a piecewise vanishing constraint, which leads to results that outperform those obtained using relaxed stress constraints. Although maintaining the local nature of stress requires a large number of stress constraints, the formulation presented here requires only one adjoint vector, which results in an efficient sensitivity evaluation. Several 2D and 3D topology optimization problems, each with a large number of local stress constraints, are provided.

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

Similar content being viewed by others

Notes

  1. A variation of the vanishing constraints is used in the present study to solve stress-constrained topology optimization problems.

  2. According to Cheng and Guo (1997), the restriction on the variable lower bound to be ε2 is not necessary. They demonstrated that, in order to guarantee convergence, the lower bound on the design variables has to be a higher order term smaller than ε as \(\varepsilon \rightarrow 0.\)

  3. However, the stress-constrained problem is not well-behaved because, given the degenerate nature of the constraints, the Lagrange multiplier set associated with a stationary point is unbounded. For optimization problems of this type, Izmailov et al. (2012) and Andreani et al. (2012) showed that the AL method exhibits global convergence properties, which suggests that this method is a viable alternative to solve stress-constrained topology optimization problems.

  4. Notice the modular structure developed for the stress-constrained topology optimization problem. Due to this feature, different constitutive behaviors can be incorporated in the present computational mechanics framework.

  5. As an example, suppose that for a given AL step we have g = − 0.5, λ = 1, and μ = 1. For this combination of constraint values and AL parameters, we have that, \(h=\max \limits (g,-~\lambda /\mu )=\max \limits (-~0.5, -~1)=-~0.5\), which yields \(P=\lambda h+\frac {1}{2}\mu h^{2}=-~3/8<0\), where P is the penalization term of the AL function.

  6. The piecewise constraint given by (17) is C1(z) because, for \(\sigma _{j}^{\text {v}}/\sigma _{\lim } >1\), \(g_{j}(\textbf {z})={\rho ^{p}_{j}}(\sigma _{j}^{\text {v}}/\sigma _{\lim }-1)^{2}\), which is the finite composition of \(C^{\infty }\) functions in this domain, and for \(\sigma _{j}^{\text {v}}/\sigma _{\lim } <1\), gj(z) = 0, which is also \(C^{\infty }\). Moreover, when \(\sigma _{j}^{\text {v}}/\sigma _{\lim }=1\), both the value of \({\rho ^{p}_{j}}(\sigma _{j}^{\text {v}}/\sigma _{\lim }-1)^{2}\) and its first derivative with respect to z are equal to zero, which is the same value of gj(z) and its derivative with respect to z, when \(\sigma _{j}^{\text {v}}/\sigma _{\lim }<1\).

  7. Because the interphase penalization F(z) in (19) is only used when we apply continuation on the filter radius, we have decided not to include it in the current derivation.

  8. Stagnation is reached when the average change in the design variables between two consecutive iterations is smaller than a given tolerance, i.e., when Change < tol (cf. Algorithm 2) and the constraints are yet not satisfied.

  9. When two consecutive filter matrices have similar topology, it indicates that the material distribution between two consecutive iterations has not changed significantly. Alternatively, we could use a criterion based on \(\left \|\textbf {P}_{i+1}-\textbf {P}_{i} \right \|\) to stop the filter reduction, but this means storing both Pi and Pi+ 1, which requires a substantial amount of RAM memory.

  10. The function defined in Eq. (36) is not unique, and other functions such as ρ(1 − ρ) can be used for the same purpose. We choose the function in Eq. (36) because its slight asymmetry with respect to ρ = 0.5 tends to favor designs with lower weight.

  11. In addition to all parameters shown in Table 1, the MMA parameters (Svanberg 1987) used in all examples for the minimization of the AL sub-problems are asyinit = 0.2, asyinc = 1.2, asydec = 0.7, move = 0.1.

  12. The stress shown in this example, as well as those shown in subsequent examples, is the stress measure \(\widetilde {\sigma }_{e}^{\text {v}}\) (18) normalized with respect to the stress limit, \(\sigma _{\lim }\).

  13. The isosurfaces as well as the STL files are obtained using the MATLAB-based graphical tool TOPslicer (Zegard and Paulino 2016).

  14. The computational costs reported in this section are based on the topology optimization results obtained in a computer with an i7-4930k CPU at 3.40 GHz and 64 GB of RAM and a NVIDIA GEFORCE GTX 1080 Ti GPU running on a 64-bit operating system.

  15. On the other hand, an inconsistent formulation refers to the case in which stresses are not treated locally, i.e., they are either aggregated or clustered.

  16. The interested reader is referred to Bertsekas (1996, p. 161) for complete proof of the differentiability of the AL function used with (42).

References

  • Achtziger W, Kanzow C (2008) Mathematical programs with vanishing constraints: optimality conditions and constraint qualifications. Math Program 114(1):69–99

    MathSciNet  MATH  Google Scholar 

  • Achtziger W, Hoheisel T, Kanzow C (2013) A smoothing-regularization approach to mathematical programs with vanishing constraints. Comput Optim Appl 55(3):733–767

    MathSciNet  MATH  Google Scholar 

  • Amstutz S, Novotny AA (2010) Topological optimization of structures subject to von Mises stress constraints. Struct Multidiscip Optim 41(3):407–420

    MathSciNet  MATH  Google Scholar 

  • Andreani R, Haeser G, Schuverdt M, Silva P (2012) A relaxed constant positive linear dependence constraint qualification and applications. Math Program 135(1-2):255–273

    MathSciNet  MATH  Google Scholar 

  • Bell E. T. (1986) Men of Mathematics. Simon and Schuster

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

    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

    MathSciNet  MATH  Google Scholar 

  • Bendsøe MP (1995) Optimization of structural topology, shape, and material. Springer, Berlin

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

    MATH  Google Scholar 

  • Bendsøe MP, Sigmund O (2003) Topology optimization: theory, methods and applications. Springer, Berlin

  • Bertsekas DP (1996) Constrained optimization and Lagrange multiplier methods (ptimization and Neural Computation Series), 1 ed. Athena Scientific

  • Bertsekas DP (1999) Nonlinear programming, 2nd ed. Athena Scientific

  • Borrvall T, Petersson J (2001) Topology optimization using regularized intermediate density control. Comput Methods Appl Mech Eng 190(37-38):4911–4928

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

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

    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

    MathSciNet  MATH  Google Scholar 

  • Cauchy AL (1827) De la pression ou tension dans un corps solide. Exercices Math 2:42–56

    Google Scholar 

  • Cheng G, Jiang Z (1992) Study on topology optimization with stress constraints. Eng Optim 20(2):129–148

    Google Scholar 

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

    Google Scholar 

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

    Google Scholar 

  • Christensen P, Klarbring A (2008) An introduction to structural optimization. Solid Mechanics and Its Applications. Springer, Netherlands

  • da Silva GA, Beck AT, Sigmund O (2019) Stress-constrained topology optimization considering uniform manufacturing uncertainties. Comput Methods Appl Mech Eng 344:512–537

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Duysinx P, Sigmund O (1998b) New developments in handling stress constraints in optimal material distribution. In: Proceedings of the 7th AIAA/USAF/NASA/ISSMO Symposium on Multidisciplinary Analysis and Optimization, vol 1, pp 1501–1509

  • Emmendoerfer Jr, H, Fancello EA (2014) A level set approach for topology optimization with local stress constraints. Int J Numer Methods Eng 99(2):129–156

    MathSciNet  MATH  Google Scholar 

  • Emmendoerfer Jr, H, Fancello EA (2016) Topology optimization with local stress constraint based on level set evolution via reaction-diffusion. Comput Methods Appl Mech Eng 305:62–88

    MathSciNet  MATH  Google Scholar 

  • Ermoliev YM, Kryazhimskii AV, Ruszczyński A (1997) Constraint aggregation principle in convex optimization. Math Program 76(3):353–372

    MathSciNet  MATH  Google Scholar 

  • Gibson I, Rosen D, Stucker B (2015) Additive manufacturing technologies: 3D printing, rapid prototyping, and direct digital manufacturing, vol 2. Springer, Berlin

  • Giraldo-Londoño O, Paulino GH (2020) PolyStress: a Matlab implementation for local stress-constrained topology optimization considering material nonlinearity. Struct Multidiscip Optim. https://doi.org/10.1007/s00158-020-02664-7

  • Guest JK, Prévost JH, Belytschko T (2004) Achieving minimum length scale in topology optimization using nodal design variables and projection functions. Int J Numer Methods Eng 61(2):238–254

    MathSciNet  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(47-48):3439–3452

    MathSciNet  MATH  Google Scholar 

  • Gurtin ME (1981) An introduction to continuum mechanics, vol 158. Academic Press

  • Hoheisel T, Kanzow C (2008) Stationary conditions for mathematical programs with vanishing constraints using weak constraint qualifications. J Math Anal Appl 337(1):292–310

    MathSciNet  MATH  Google Scholar 

  • Holmberg E, Torstenfelt B, Klarbring A (2013) Stress constrained topology optimization. Struct Multidiscip Optim 48(1):33–47

    MathSciNet  MATH  Google Scholar 

  • Izmailov AF, Solodov MV, Uskov EI (2012) Global convergence of augmented lagrangian methods applied to optimization problems with degenerate constraints, including problems with complementarity constraints. SIAM J Optim 22(4):1579–1606

    MathSciNet  MATH  Google Scholar 

  • James K, Lee E, Martins J (2012) Stress-based topology optimization using an isoparametric level set method. Finite Elem Anal Des 58:20–30

    Google Scholar 

  • Kirsch U (1989) Optimal topologies in truss structures. Comput Methods Appl Mech Eng 72 (1):15–28

    MATH  Google Scholar 

  • Kirsch U, Taye S (1986) On optimal topology of grillage structures. Eng Comput 1:229–243

    Google Scholar 

  • Kirsch U (1990) On singular topologies in optimum structural design. Struct Optim 2(3):133–142

    Google Scholar 

  • Kiyono C, Vatanabe S, Silva E, Reddy J (2016) A new multi-p-norm formulation approach for stress-based topology optimization design. Compos Struct 156:10 –19. 70th Anniversary of Professor J. N. Reddy

    Google Scholar 

  • Kreisselmeier G, Steinhauser R (1979) Systematic control design by optimizing a vector performance index. IFAC Proc Vol 12(7):113–117. IFAC Symposium on computer Aided Design of Control Systems, Zurich, Switzerland

    MATH  Google Scholar 

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

    Google Scholar 

  • Lee E, James KA, Martins JRRA (2012) Stress-constrained topology optimization with design-dependent loading. Struct Multidiscip Optim 46(5):647–661

    MathSciNet  MATH  Google Scholar 

  • Lian H, Christiansen AN, Tortorelly DA, Sigmund O (2017) Combined shape and topology optimization for minimization of maximal von Mises stress. Struct Multidiscip Optim 55(5):1541–1557

    MathSciNet  Google Scholar 

  • Love AEH (1892) A treatise on the mathematical theory of elasticity, Vol 1. Cambridge University Press, Cambridge

  • Malvern LE (1969) Introduction to the mechanics of a continuous Mmedium. Prentice-Hall

  • Navarrina F, Muiños I, Colominas I, Casteleiro M (2005) Topology optimization of structures: a minimum weight approach with stress constraints. Adv Eng Softw 36(9):599–606

    MATH  Google Scholar 

  • Nocedal J, Wright SJ (2006) Numerical optimization, 2nd ed. Springer, Berlin

  • Paris J, Navarrina F, Colominas I, Casteleiro M (2009) Topology optimization of continuum structures with local and global stress constraints. Struct Multidiscip Optim 39(4):419–437

    MathSciNet  MATH  Google Scholar 

  • Paris J, Navarrina F, Colominas I, Casteleiro M (2010) Block aggregation of stress constraints in topology optimization of structures. Adv Eng Softw 41(3):433–441

    MATH  Google Scholar 

  • Park YK (1995) Extensions of optimal layout design using the homogenization method. Ph.D thesis, University of Michigan, Ann Arbor

  • Pereira JT, Fancello EA, Barcellos CS (2004) Topology optimization of continuum structures with material failure constraints. Struct Multidiscip Optim 26(1-2):50–66

    MathSciNet  MATH  Google Scholar 

  • Petersson J (2001) On continuity of the design-to-state mappings for trusses with variable topology. Int J Eng Sci 39(10):1119– 1141

    MathSciNet  MATH  Google Scholar 

  • Rozvany GIN, Birker T (1994) On singular topologies in exact layout optimization. Struct Optim 8(4):228–235

    Google Scholar 

  • Rozvany GIN (2001) On design-dependent constraints and singular topologies. Struct Multidiscip Optim 21(2):164–172

    MathSciNet  Google Scholar 

  • Sharma A, Maute K (2018) Stress-based topology optimization using spatial gradient stabilized XFEM. Struct Multidiscip Optim 57(1):17–38

    MathSciNet  Google Scholar 

  • Sigmund O (2007) Morphology-based black and white filters for topology optimization. Struct Multidiscip Optim 33(4-5):401–424

    Google Scholar 

  • Sigmund O (2009) Manufacturing tolerant topology optimization. Acta Mech Sinica 25(2):227–239

    MATH  Google Scholar 

  • Stolpe M, Svanberg K (2001) On the trajectories of the epsilon-relaxation approach for stress-constrained truss topology optimization. Struct Multidiscip Optim 21(2):140–151

    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

    MathSciNet  MATH  Google Scholar 

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

    Google Scholar 

  • Talischi C, Paulino GH, Pereira A, Menezes IFM (2012a) PolyMesher: a general-purpose mesh generator for polygonal elements written in Matlab. Struct Multidiscip Optim 45(3):309–328

    MathSciNet  MATH  Google Scholar 

  • Talischi C, Paulino GH, Pereira A, Menezes IFM (2012b) PolyTop: a Matlab implementation of a general topology optimization framework using unstructured polygonal finite element meshes. Struct Multidiscip Optim 45(3):329–357

    MathSciNet  MATH  Google Scholar 

  • Talischi C, Paulino GH (2013) An operator splitting algorithm for Tikhonov-regularized topology optimization. Comput Methods Appl Mech Eng 253:599–608

    MathSciNet  MATH  Google Scholar 

  • Timoshenko S, Goodier JN (1951) Theory of elasticity, 2 ed. McGraw-Hill

  • Verbart A, Langelaar M, van Keulen F (2016) Damage approach: a new method for topology optimization with local stress constraints. Struct Multidiscip Optim 53(5):1081–1098

    MathSciNet  Google Scholar 

  • Verbart A, Langelaar M, Keulen F. v. (2017) A unified aggregation and relaxation approach for stress-constrained topology optimization. Struct Multidiscip Optim 55(2):663–679

    MathSciNet  Google Scholar 

  • Wang MY, Wang S (2005) Bilateral filtering for structural topology optimization. Int J Numer Methods Eng 63(13):1911–1938

    MathSciNet  MATH  Google Scholar 

  • Wang F, Lazarov BS, Sigmund O (2011) On projection methods, convergence and robust formulations in topology optimization. Struct Multidiscip Optim 43(6):767–784

    MATH  Google Scholar 

  • Wong KV, Hernandez A (2012) A review of additive manufacturing. ISRN Mechanical Engineering 2012, pp 1–10

  • Xia Q, Shi T, Liu S, Wang MY (2012) A level set solution to the stress-based structural shape and topology optimization. Comput Struct 90:55–64

    Google Scholar 

  • Xu S, Cai Y, Cheng G (2010) Volume preserving nonlinear density filter based on Heaviside functions. Struct Multidiscip Optim 41(4):495–505

    MathSciNet  MATH  Google Scholar 

  • Yang RJ, Chen CJ (1996) Stress-based topology optimization. Struct Optim 12(2):98–105

    Google Scholar 

  • Zegard T, Paulino GH (2016) Bridging topology optimization and additive manufacturing. Struct Multidiscip Optim 53(1):175–192

    Google Scholar 

Download references

Acknowledgments

This paper is dedicated to the memory of Augustin-Louis Cauchy (August 21, 1789 - May 23, 1857). We thank Mrs. Emily D. Sanders for useful suggestions that contributed to improve the manuscript.

Funding

F.V.S., O.G.-L., and G.H.P. acknowledge the financial support from the US National Science Foundation under grant #1663244 and the endowment provided by the Raymond Allen Jones Chair at the Georgia Institute of Technology. F.V.S. and I.F.M.M. acknowledge the support provided by Brazilian agencies CNPQ and FAPERJ, and Tecgraf/PUC-Rio (Group of Technology in Computer Graphics), Rio de Janeiro, Brazil.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Glaucio H. Paulino.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Responsible Editor: Kurt Maute

Dedicated to the memory of Augustin-Louis Cauchy (August 21, 1789 – May 23, 1857)

Publisher’s note

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

Replication of results

The paper includes details of the numerical implementation and all input parameters for the numerical examples are provided to facilitate replication of the results. To help with the dissemination of the original ideas presented here, we follow-up with the publication of an educational paper and corresponding Matlab code for stress constrained topology optimization in the SMO Journal (Giraldo-Londoño and Paulino 2020).

Appendices

Appendix A: Augmented Lagrangian method for inequality constraints

The procedure described in Section 4 is designed to solve optimization problems with equality constraints. For the sake of completeness, and to provide the reader interested in applying the AL method to problems with inequality constraints, we present an extension of the AL method for inequality constraints. As usual, we handle inequality constraints by introducing slack variables sj,j = 1,…,Nc to the optimization problem (Nocedal and Wright 2006). Consider the following optimization problem with inequality constraints:

$$ \begin{array}{ll} \underset{\textbf{z}\in\mathbb{R}^{n}}{\min} ~~&f(\textbf{z})\\ ~~~\text{s.t.} ~~&g_{j}(\textbf{z}) \leq 0~~\forall j=1,\ldots,N_{c}\\ &\textbf{L}\leq\textbf{z}\leq \textbf{U}, \end{array} $$
(37)

where L and U define the lower and upper bounds of the design variables, respectively. Introducing slack variables, constraints gj(z) ≤ 0 are rewritten as follows:

$$ h_{j}(\textbf{z})=g_{j}(\textbf{z})+s_{j}=0,~s_{j}\ge0,~j=1,\ldots,N_{c}. $$
(38)

Consequently, the approximate sub-problem that needs to be solved at the k th step of the AL method is as follows:

$$ \begin{array}{ll} \underset{\textbf{z},~\textbf{s}}{\min} ~~&J^{(k)}(\textbf{z},\textbf{s})=f(\textbf{z})+\sum\limits_{j=1}^{N_{c}}\left[\lambda_{j}^{(k)}\!\left( g_{j}(\textbf{z})+s_{j}\right)+\frac{\mu^{(k)}}{2}\left( g_{j}(\textbf{z})+s_{j}\right)^{2}\right]\\ ~~~\text{s.t.} ~~&\textbf{L}\leq\textbf{z}\leq \textbf{U}\\ &s_{j}\ge 0~~\forall j=1,\ldots,N_{c}. \end{array} $$
(39)

The minimization of J(k)(z,s) with respect to the slack variables is obtained explicitly for any fixed z by solving the optimization problem as follows:

$$ \begin{array}{ll} \underset{s_{j}}{\min} ~~&\left[\lambda_{j}^{(k)}\left( g_{j}(\textbf{z})+s_{j}\right)+\frac{\mu^{(k)}}{2}\left( g_{j}(\textbf{z})+s_{j}\right)^{2}\right]\\ ~~~\text{s.t.} ~~&s_{j}\ge 0. \end{array} $$
(40)

The optimization statement (40) is defined in terms of the slack variable, sj, associated with constraint gj. As a result, its solution can be found in closed form using the stationary conditions of the Lagrangian of Eq. (40), which leads to the following:

$$ s_{j}=\max\left[0,-\left( \frac{\lambda_{j}^{(k)}}{\mu^{(k)}}+g_{j}(\textbf{z})\right)\right]~~\forall j=1,\ldots,N_{c}. $$
(41)

Substituting (41) into (38) leads to the following:

$$ h_{j}(\textbf{z})=\max\left[g_{j}(\textbf{z}),-\frac{\lambda_{j}^{(k)}}{\mu^{(k)}}\right]~~\forall j=1,\ldots,N_{c}. $$
(42)

Using (42), the inequality constraints gj(z) ≤ 0 of (37) can be replaced by equality constraints, allowing the problem to be solved using the procedure described for solving the equality-constrained problem (6). As inferred from (42), the slack variables do not need to be computed explicitly, facilitating the implementation of the AL method with inequality constraints. One must recall that the Lagrange multiplier estimators, \(\lambda _{j}^{(k)}\), and the penalty factor, μ(k), remain constant for each AL sub-problem, and thus the AL function is continuously differentiable (with respect to the design variables) at each AL step. Despite the presence of the maximum function, the AL function used with (42) is differentiable even at the points in which \(g_{j}(\textbf {z})=-\lambda _{j}^{(k)}/\mu ^{(k)}\).Footnote 16 If (17) is substituted into (42), it follows that gj(z) = hj(z), which is the case in our implementation of the AL-based method. However, if one were to use a different stress constraint definition that can take negative values, then (42) (in its explicit form) would be necessary in the implementation.

Appendix B: Apparent “local” von Mises stress vs. stress measure

Here, we clarify the difference between the apparent “local” von Mises stress, σv, and the stress measure, \(\widetilde {\sigma }^{\text {v}}\), described in Section 5.2. Using a stress measure of the form \(\widetilde {\sigma }^{\text {v}}=\rho ^{\alpha }\sigma ^{\text {v}}\) is a typical procedure in the context of stress-based topology optimization (e.g., Bruggi and Duysinx 2012; Lee et al. 2012), because the apparent “local” von Mises stress is large in regions of low density. The high stresses in regions of low density can be seen in Fig. 14a (only regions with density above 0.05 are shown), which displays the normalized apparent “local” von Mises stress for the 16,380 mesh solution with linear filter of the L-bracket example in Table 3. We observe that the largest apparent local stress occurs at the boundaries between solid and void, in which the density is small due to the filter operator. In contrast, Fig. 14b shows the normalized stress measure of the same solution, in which the stress measure strictly satisfies the stress limit of 1. By adopting \(\widetilde {\sigma }^{\text {v}}=\rho ^{0.5}\sigma ^{\text {v}}\), we reduce the effect of the vaguely defined stress in the regions of low density.

Fig. 14
figure 14

Normalized apparent “local” von Mises stress a and normalized stress measure b for the 16,380 mesh solution with linear filter of the L-bracket example in Table 3. The stress is only displayed for the regions with density above 0.05. Because both stresses are normalized, the value of 1 represents the stress limit

Appendix C: Comparison of the piecewise vanishing constraint with ε-relaxation

We compare the optimization results obtained using our piecewise vanishing constraint with those obtained using the ε-relaxed constraint (Cheng and Guo 1997). The ε-relaxation approach used here is the one proposed by Cheng and Guo (1997) and Petersson (2001) in which we start with a value of ε = 0.01, and we set the lower bound of the design variable to ε2. The value of ε is divided by two every time that we restart the AL parameters to gradually decrease the value of ε reaching a final value of ε = 0.000625. We also update the lower bound of the design variable so that it is always equal to ε2.

Figure 15 displays the results that we obtain using each of these constraints. As compared to the designs obtained using the piecewise vanishing constraint, those obtained using the ε-relaxed constraint have a significantly larger volume and contain regions of pure solid material (which appear sub-optimal) near the bottom left corner of the bracket. Based on the information displayed on Fig. 15, the results obtained using the piecewise vanishing constraint outperform those obtained using ε-relaxation constraint in terms of both optimized volume and quality of results.

Fig. 15
figure 15

Optimized topologies for an L-bracket obtained using the piecewise vanishing constraint (top) and using the ε-relaxed constraint (bottom)

Appendix D: Effect of weight factors γe

The introduction of weight factors γe is one of the main characteristics of the present formulation. Here, we investigate the effects of using the weight factors in the optimization results obtained for the L-bracket. Figure 16 presents the results that we obtain using γe, as well as, those not using γe. We observe that, as compared to the results obtained with γe, those obtained without γe have a higher final volume and have fewer beam-like elements. The factor γe helps the optimizer to overcome unfavorable local optima by increasing the relevance of the objective function in elements which have low stress value. If γe is not considered, the optimizer is likely to get trapped in a bad local optima with high volume, preventing it to achieve structures with more slender elements, as we expect for this benchmark problem.

Fig. 16
figure 16

Optimized topologies for an L-bracket obtained when weight factors γe are considered (top) and when they are not considered (bottom)

Appendix E: Effect of scale factor η

The scale factor, η, is introduced to normalize the values of the AL penalization parameters, μ(k) and \(\lambda _{j}^{(k)}\), such that we eliminate the need for adjusting the numerical values of μ(1) and \(\lambda _{j}^{(1)}\) for problems with different number of constraints. We demonstrate the effectiveness of the parameter η through the numerical results shown in Fig. 17. These results correspond to the optimized topologies obtained for the L-bracket when the parameter η is either used or not. For the case when the parameter η is not used, we re-calibrate the values of μ(1) and \(\lambda _{j}^{(1)}\) for a mesh size of 16,380 elements, and these values are used to obtain the optimized topologies for the other mesh sizes. As the mesh is refined (i.e., as the number of constraints increases), the results obtained when η is not used are clearly worse than those obtained when η is used. That is because when η is not considered, the magnitude of the penalty parameter of the AL function (19) increases as the number of constraints increases. Without the η parameter to normalize this effect, the optimizer becomes trapped in local optima with higher volume.

Fig. 17
figure 17

Optimized topologies for an L-bracket obtained when scale factor η is considered (top) and when it is not considered (bottom)

Appendix F: Effect of the number of AL parameter restarts

We investigate the effect of restarting the AL parameters, \(\lambda _{j}^{(k)}\) and μ(k), as well as the weight factors, \(\gamma _{e}^{(k)}\), on the final solution (cf. Section 7.1). As illustrated by Fig. 18, restarting these parameters helps us achieve a solution with overall lower volume. The results presented in the figure correspond to the final solutions of the L-bracket problem with 16,380 elements that we obtained using the filter reduction approach for various numbers of restarts.

Fig. 18
figure 18

Effect of the number of AL parameter restarts on the optimized topologies obtained for an L-bracket meshed with 16,380 elements: a no restart; b one restart; c two restarts; d three restarts; and e four restarts. As the number of restarts increases, the small-scale artifacts are removed and the optimized volume fraction becomes smaller

As shown by the results in Fig. 18, the topology obtained with no restart contains several small-scale artifacts, which cannot be removed without increasing the stress in the structure. As a result, the solution gets trapped in an unfavorable local optimum with higher volume. After the first restart, the small-scale features disappear and the final volume becomes smaller. If we keep restarting the AL parameters and weight factors when the optimization stagnates, the total volume that we are able to reach keeps decreasing because the optimizer is able to find local optima with lower volume.

Appendix G: Effect of stress limit \(\sigma _{\lim }\)

We analyze the effect of the stress limit, \(\sigma _{\lim }\), in the optimization results obtained for the L-bracket with 16,380 elements. The optimization results obtained using our AL-based framework with the filter reduction scheme are depicted in Fig. 19. The results demonstrate that increasing the stress limit leads to designs with lower volume fraction and more slender members, yet all these designs are topologically similar. A more significant change of topology is observed when \(\sigma _{\lim }\ge 80\) Pa, in which the vertical member of the left edge of the L-bracket begins to tilt. The results also demonstrate that the optimized volume decreases monotonically as we increase the stress limit. Although not shown in the figure, our numerical results also indicate that the minimum stress limit possible for this optimization problem is around 37 Pa. That is because the final volume obtained from the optimization results increase significantly as the stress limit approaches this value.

Fig. 19
figure 19

Optimized volume of the L-bracket meshed with 16,380 elements as a function of the stress limit

Appendix H: Nomenclature

α :

Parameter used to update penalty parameter μ(k)

λ (k) :

Vector of approximated Lagrange multipliers at the k th iteration of the AL method

ξ :

Adjoint vector used for sensitivity evaluation

η :

Scale factor used in the AL function (19)

γL, γU:

Lower and upper bounds of weight factors γe, respectively

γ e :

Weight factor for element e used in the modified AL function (19)

μ (k) :

Penalty parameter at the k th iteration of the AL method

ν :

Poisson’s ratio of solid material

\(\sigma _{e}^{\text {v}}\) :

von Mises stress at the centroid of element e

\(\sigma _{\lim }\) :

Stress limit for a given material

𝜖 :

Ersatz stiffness

ε :

Relaxation parameter in ε-relaxed approach

ρ :

Vector of filtered densities

\(\tilde {\rho }_{e}\) :

Volume fraction of element e, defined using a smooth Heaviside projection function

Δ:

Infinitesimal quantity

E 0 :

Young’s modulus of solid material

J(k)(z):

Augmented Lagrangian function at iteration k

N c :

Number of stress constraints

N e :

Number of elements in a finite element mesh

β:

Mass penalization parameter used in the smooth Heaviside projection function

K :

Global stiffness matrix

P :

Filter matrix

f :

Global force vector

k e :

Element stiffness matrix

s :

Vector of slack variables

u :

Global displacement vector

z :

Vector of design variables

ai, bi:

Parameters used to define the evolution of weight factors γe (i = 1, 2)

f :

Objective function

gj(z):

j th stress constraint

h :

Equality constraint

hj(z):

j th modified stress constraint used in the AL method with inequality constraints

m(z):

Mass of the structure

p :

Stiffness penalization parameter

q :

Relaxation parameter in qp-relaxed approach

r :

Filter radius

s :

Exponent of the polynomial filter

v e :

Volume of element e for density equal one

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Senhora, F.V., Giraldo-Londoño, O., Menezes, I.F.M. et al. Topology optimization with local stress constraints: a stress aggregation-free approach. Struct Multidisc Optim 62, 1639–1668 (2020). https://doi.org/10.1007/s00158-020-02573-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-020-02573-9

Keywords

Navigation