Skip to main content
Log in

Topology optimization considering the Drucker–Prager criterion with a surrogate nonlinear elastic constitutive model

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

Abstract

We address material nonlinear topology optimization problems considering the Drucker–Prager strength criterion by means of a surrogate nonlinear elastic model. The nonlinear material model is based on a generalized J2 deformation theory of plasticity. From an algorithmic viewpoint, we consider the topology optimization problem subjected to prescribed energy, which leads to robust convergence in nonlinear problems. The objective function of the optimization problem consists of maximizing the strain energy of the system in equilibrium subjected to a volume constraint. The sensitivity analysis is quite effective and efficient in the sense that there is no extra adjoint equation. In addition, the nonlinear structural equilibrium problem is solved through direct minimization of the structural strain energy using Newton’s method with an inexact line search strategy. Four numerical examples demonstrate features of the proposed nonlinear topology optimization framework considering the Drucker–Prager strength criterion.

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.

Institutional subscriptions

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

Similar content being viewed by others

Notes

  1. For several inelastic constitutive models, the energy control has better convergence behavior than the load control method. For instance, Crisfield (1991) pointed out that load control is not preferable when a small addition to the load causes a relatively large additional displacement or when limit points are encountered. The energy control approach overcomes this difficulty in regions where the stress state tends to reach the strength limit.

References

  • Alberdi R, Khandelwal K (2017) Topology optimization of pressure dependent elastoplastic energy absorbing structures with material damage constraints. Finite Elem Anal Des 133:42–61

    Article  MathSciNet  Google Scholar 

  • Ascher UM, Greif C (2011) A first course in numerical methods. SIAM, Philadelphia

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

    Article  MathSciNet  Google Scholar 

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

    MATH  Google Scholar 

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

  • Bogomolny M, Amir O (2012) Conceptual design of reinforced concrete structures using topology optimization with elastoplastic material modeling. Int J Numer Methods Eng 90(13):1578–1597

    Article  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Boyd S, Vandenberghe L (2004) Convex optimization. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Chen WF, Han DJ (1988) Plasticity for structural engineers. Springer-Verlag, New York

    Book  Google Scholar 

  • Christensen PW, Klarbring A (2009) An introduction to structural optimization. Springer, Linköping

  • Crisfield MA (1991) Non-linear finite element analysis of solids and structures - volume 1: essentials. John Wiley & Sons

  • Davidson MW (2010) Robert Hooke: physics, architecture, astronomy, paleontology, biology. Lab Med 41(3):180–182

    Article  Google Scholar 

  • Drucker DC, Prager W (1952) Soil mechanics and plastic analysis or limit design. Q Appl Math 10(2):157–165

    Article  MathSciNet  Google Scholar 

  • Hencky H (1924) Zur Theorie plastischer Deformationen und der hierdurch. Z Angew Math 4(4):323–334

  • Hill R (1950) The mathematical theory of plasticity. Oxford University Press, Oxford

    MATH  Google Scholar 

  • Holmes DP (2019) Elasticity and stability of shape shifting structures. Curr Opin Colloid Interface Sci 40:118–137

  • Hooke R (1678) Lectures De Potentia Restitutiva, or of Spring. Explaining the Power of Springing Bodies. John Martyn, London

  • Kachanov LM (1971) Foundations of the theory of plasticity. North-Holland Publishing Company, Amsterdam

  • Klarbring A, Strömberg N (2012) A note on the min-max formulation of stiffness optimization including non-zero prescribed displacements. Struct Multidiscip Optim 45(1):147–149

    Article  MathSciNet  Google Scholar 

  • Lubarda VA (2000) Deformation theory of plasticity revisited. Proc Mont Acad Sci Arts 13:117–143

    Google Scholar 

  • Lubliner J (1990) Plasticity theory. Macmillan, New York

  • Luo Y, Kang Z (2012) Topology optimization of continuum structures with Drucker–Prager yield stress constraints. Comput Struct 90–91:65–75

    Article  Google Scholar 

  • Marsden JE, Hughes TJR (1983) Mathematical foundations of elasticity. Prentice-Hall, Upper Saddle River

    MATH  Google Scholar 

  • Niu F, Xu S, Cheng G (2011) A general formulation of structural topology optimization for maximizing structural stiffness. Struct Multidiscip Optim 43(4):561–572

    Article  Google Scholar 

  • Ramos Jr AS, Paulino GH (2016) Filtering structures out of ground structures - a discrete filtering tool for structural design optimization. Struct Multidiscip Optim 54(1):95–116

  • Rozvany GIN (2009) A critical review of established methods of structural topology optimization. Struct Multidiscip Optim 37(3):217–237

    Article  MathSciNet  Google Scholar 

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

    Article  Google Scholar 

  • Sanders ED, Ramos Jr AS, Paulino GH (2017) A maximum filter for the ground structure method: an optimization tool to harness multiple structural designs. Eng Struct 151:235–252

  • Sonato M, Piccolroaz A, Miszuris W, Mishuris G (2015) General transmission conditions for thin elasto-plastic pressure-dependent interphase between dissimilar materials. Int J Solids Struct 64–65:9–21

    Article  Google Scholar 

  • Souza Neto EA, Perić D, Owen DRJ (2008) Computational methods for plasticity: theory and applications. John Wiley & Sons, Chichester

  • Swan CC, Kosaka I (1997) Voigt–Reuss topology optimization for structures with nonlinear material behaviors. Int J Numer Methods Eng 40(20):3785–3814

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Tikhonov AN, Arsenin VY (1977) Solutions of ill posed problems. Wiley, New York

  • Timoshenko SP (1934) Theory of elasticity. McGraw Hill, New York

  • Univ. of Oxford (2014) Hooke Lecture. https://www.maths.ox.ac.uk/node/893. Accessed 16 July 2020

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

    Article  Google Scholar 

  • Zhao T, Ramos Jr AS, Paulino GH (2019) Material nonlinear topology optimization considering the von Mises criterion through an asymptotic approach: max strain energy and max load factor formulations. Int J Numer Methods Eng 118:804–828

Download references

Acknowledgments

This paper is dedicated to the memory of Robert Hooke (July 28, 1635 – March 3, 1703).

Funding

GHP and TZ acknowledge the financial support from the US National Science Foundation under project #1663244 and the endowment provided by the Raymond Allen Jones Chair at the Georgia Institute of Technology. ASR Jr. and ENL appreciate the financial support from the Brazilian National Council for Research and Development (CNPq). The information in this paper is the sole opinion of the authors and does not necessarily reflect the views of the sponsoring agencies.

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.

Replication of results

In Section 4, we provide the detailed parameters used for obtaining the results of the four numerical examples. In addition, we include the ABAQUS® user subroutine UMAT as supplementary material, which can be used to reproduce results presented in the paper.

Additional information

Responsible Editor: Helder C. Rodrigues

Dedicated to the memory of Robert Hooke (July 28, 1635 – March 3, 1703)

Publisher’s note

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

Electronic supplementary material

ESM 1

(FOR 15 kb).

Appendices

Appendix A. Illustration of the energy control approach

Let us consider a ground structure-based elastic formulation (Bendsøe and Sigmund 2003; Christensen and Klarbring 2009; Sanders et al. 2017), in which (43) can be written as:

$$ {\displaystyle \begin{array}{c}\underset{\boldsymbol{x}}{\max}\kern0.50em U\left(\boldsymbol{x},\boldsymbol{u}\left(\boldsymbol{x}\right)\right)\\ {}\mathrm{s}.\mathrm{t}.\begin{array}{c}\sum \limits_{e=1}^n{L}_e^T{x}_e-{V}_{\mathrm{max}}\le 0\\ {}{x}_e^{\mathrm{min}}\le {x}_e\le {x}_e^{\mathrm{max}},e=1,\dots, n\end{array}\\ {}\mathrm{with}\left\{\begin{array}{c}\boldsymbol{T}\left(\boldsymbol{x},\boldsymbol{u}\left(\boldsymbol{x}\right)\right)=\chi \left(\boldsymbol{x},\boldsymbol{u}\left(\boldsymbol{x}\right)\right){\boldsymbol{f}}_0\\ {}{\boldsymbol{f}}_0^T\boldsymbol{u}\left(\boldsymbol{x}\right)=2{C}_0\end{array}\right.\end{array}} $$
(56)

The vector x is a vector of design variables, with component xe being the cross-sectional area of truss member e—it is subjected to lower bound \( {x}_e^{\mathrm{min}} \) and upper bound \( {x}_e^{\mathrm{max}} \). In addition, n is the number of truss members in the ground structure, Le is the length of truss member e, Vmax is the upper bound on the total volume, and u(x) is the displacement vector. For illustrative purpose, we assume the particular case of linear elasticity. In the following, we solve a simple three-bar truss example to explain how to estimate a proper value for the prescribed energy C0 and demonstrate that C0 remains constant at each design iteration during the entire optimization process.

Fig. 18
figure 18

The three-bar truss example. a Design domain and boundary conditions. b, c, and d are the topologies at optimization iteration #1, #35, and #72, respectively. Blue bars are in tension and red bars are in compression

The three-bar example shown in Fig. 18(a) is made of a linear elastic material with the Young’s modulus E = 200 GPa. The structure has two degree of freedoms (dofs), and two reference forces are applied at each dof, respectively. The magnitudes of the two reference forces are f01 = 40 N and f02 = 80 N. The displacements at each of the two dofs are u1 and u2.

Let us assume that the designer suggests that the magnitudes of initial displacements at each of dofs are the same, i.e., u1 = u2 = 10−5 m. Based on this assumption, we estimate C0 as:

$$ {C}_0={C_0}_1+{C_0}_2=\frac{1}{2}{f_0}_1{u}_1+\frac{1}{2}{f_0}_2{u}_2=6\times {10}^{-4}\ \mathrm{N}\cdotp \mathrm{m} $$
(57)

This three-bar optimization problem converges with 72 iterations. As an example, we plot the topologies at optimization iteration #1, #35, and #72 in Fig. 18(b), (c), and (d), respectively. Table 11 shows the prescribed energy C0 at each optimization design iteration, which is composed of the energy C01 and C02 at each dof, respectively. As expected, C0 remains a constant in each optimization iteration. The data in Table 11 can be visualized in Fig. 19(a) and (b), which illustrate the energy control approach during the optimization process.

Table 11 Calculation of the prescribed energy at each optimization iteration
Fig. 19
figure 19

Displacement versus reaction force diagrams at (a) dof #1 and (b) dof #2, respectively. The shaded area represents prescribed energy at optimization iteration #0 (yellow), #1 (green), #35 (red), and #72 (blue), i.e., optimal design

Appendix B. Estimating the limit value of the prescribed energy \( {\boldsymbol{C}}_{\mathbf{0}} \)

Here, we provide a rational approach to estimate the limit value of the prescribed energy C0 for a design optimization problem given a fixed volume constraint. This approach includes two phases as follows:

  • Phase #1: Calculate an initial guess of the \( {C}_0^0 \) based on an approximated displacement vector u

$$ {C}_0^0=\frac{1}{2}{\boldsymbol{f}}_0^T\boldsymbol{u} $$
(58)
  • Phase #2: Estimate the limit value of the prescribed energy \( {\left({C}_0^i\right)}_{\mathrm{lim}} \) for the given optimized topology corresponding to the prescribed energy \( {C}_0^i \).

    • At step i (i = 0, 1, 2…), perform FEM analysis of the optimized topology obtained with \( {C}_0^i \), and plot the curve that represents the relationship between the prescribed energy C0 and the reaction load factor χ.

    • Select two points on the curve (i.e., C0 versus χ). One of the two points is \( \left[{C}_0^i,\kern0.5em {\chi}^i\right] \), and the other point \( \left[{\left({C}_0^i\right)}_k,\kern0.5em {\left({\chi}^i\right)}_k\right] \) is obtained iteratively such that:

$$ \frac{{\left({\chi}^i\right)}_k-{\left({\chi}^i\right)}_{k-1}}{{\left({C}_0^i\right)}_k-{\left({C}_0^i\right)}_{k-1}}\le \beta {K}_0 $$
(59)

where K0 is the slope of the curve as C0 is close to zero, and β is a small ratio (e.g., 4 × 10−2 as appropriate).

  • Based on the two selected points \( \left[{C}_0^i,\kern0.5em {\chi}^i\right] \) and \( \left[{\left({C}_0^i\right)}_k,\kern0.5em {\left({\chi}^i\right)}_k\right] \), build an asymptotic function with the functional format as follows:

$$ \chi \left({C}_0\right)=\frac{a{C}_0}{b+{C}_0} $$
(60)
  • Calculate the limit value of the prescribed energy \( {\left({C}_0^i\right)}_{\mathrm{lim}} \) at the current step i as follows:

$$ {\left({C}_0^i\right)}_{\mathrm{lim}}=b\ \left(\frac{1}{\sqrt{\alpha }}-1\right) $$
(61)

where α is a small ratio (e.g., 2 × 10−2 as appropriate).

  • Proceed to step i + 1. Stop, if the following criterion is satisfied

$$ \frac{\left|{\left({C}_0^{i+1}\right)}_{\mathrm{lim}}-{\left({C}_0^i\right)}_{\mathrm{lim}}\right|}{{\left({C}_0^i\right)}_{\mathrm{lim}}}\le 5\% $$
(62)
  • Then, \( {\left({C}_0^{i+1}\right)}_{\mathrm{lim}} \) is the estimated limit value of the prescribed energy for the given a fixed volume constraint.

For example, we investigate the limit value of the prescribed energy for the clamped design optimization problem in Section 4.2 using the approach mentioned above. In Fig. 20(a), the red curve represents the structural response of the optimized topology obtained with \( {C}_0^0=0.013\ \mathrm{MJ} \) (see Fig. 20(b)), and the black curve is an asymptotic approximation based on (60). From (61), we can obtain the limit prescribed energy for this topology (Fig. 20(b)) as \( {\left({C}_0^0\right)}_{\mathrm{lim}}=0.05\ \mathrm{MJ} \). With this value of the prescribed energy, the corresponding optimized topology is shown in Fig. 20(c). Similarly, we estimate the limit value of the prescribed energy for this topology as \( {\left({C}_0^1\right)}_{\mathrm{lim}}=0.13\ \mathrm{MJ} \). We repeat the procedure until (62) is satisfied, and then, the limit value of the prescribed energy for this clamped problem is obtained as \( {\left({C}_0^2\right)}_{\mathrm{lim}}=0.134\ \mathrm{MJ} \). As the prescribed energy increases, the corresponding topologies shown in Fig. 20(b), (c), and (d) are not changing. Instead, we note that the shape of those topologies is different.

Fig. 20
figure 20

(a) Structural response and its approximation for the optimized topology in (b) with \( {C}_0^0=0.013\ \mathrm{MJ} \). (c) Optimized topology with \( {C}_0^1=0.05\ \mathrm{MJ} \). (d) Optimized topology with \( {C}_0^2=0.13\ \mathrm{MJ} \)

Appendix C. Relationship between the increment of principal stress on the strength surface and the increment of principal strain

A reference principal strain tensor of the deformation at a material point can be written as:

$$ \left[{\varepsilon}^{\mathrm{ref}}\right]=\left[\begin{array}{ccc}{\varepsilon}_1^{\mathrm{ref}}& 0& 0\\ {}0& {\varepsilon}_2^{\mathrm{ref}}& 0\\ {}0& 0& {\varepsilon}_3^{\mathrm{ref}}\end{array}\right] $$
(63)

and the principal strain tensor controlled by a positive scaling factor ξ is denoted by:

$$ \left[\varepsilon \right]=\xi \left[{\varepsilon}^{\mathrm{ref}}\right] $$
(64)

Then, the first invariant of the principal strain tensor is:

$$ {J}_1=\xi {J}_1^{\mathrm{ref}} $$
(65)

By making reference to (1), we can calculate the principal stress components as:

$$ {\sigma}_i=\xi \left(\lambda\ {J}_1^{\mathrm{ref}}+2\ \mu\ {\varepsilon}_i^{\mathrm{ref}}\right),\kern0.5em i=1,2,3 $$
(66)

where λ and μ can be obtained from (4), (9), and (10) considering the Drucker–Prager criterion.

Next, by taking the derivatives of the principal stresses components with respect to the scaling factor ξ, we obtain that:

$$ \frac{d{\sigma}_i}{d\xi}=\left(\lambda +\xi \frac{d\lambda}{d\xi}\right){J}_1^{\mathrm{ref}}+2\ \left(\mu +\xi \frac{d\mu}{d\xi}\right){\varepsilon}_i^{\mathrm{ref}} $$
(67)

We then check that the terms in parentheses in (67) are independent of the scaling factor ξ, i.e.,

$$ \lambda +\xi \frac{d\lambda}{d\xi}=\frac{\left(1-6{\beta}^2\right){J}_1^{\mathrm{ref}}\sqrt{J_2^{\mathrm{ref}}}-\beta \left(6{J}_2^{\mathrm{ref}}-{J_1^{\mathrm{ref}}}^2\right)}{3{J}_1^{\mathrm{ref}}\sqrt{J_2^{\mathrm{ref}}}\left[6{\beta}^2\left(1+\upsilon \right)+1-2\upsilon \right]}E $$
(68)

and

$$ \mu +\xi \frac{d\mu}{d\xi}=\frac{6\upbeta \sqrt{J_2^{\mathrm{ref}}}-{J}_1^{\mathrm{ref}}}{2\sqrt{J_2^{\mathrm{ref}}}\left[6{\beta}^2\left(1+\upsilon \right)+1-2\upsilon \right]}\beta E $$
(69)

Therefore, we conclude that the increment of principal stress on the Drucker–Prager strength surface is constant for each reference strain tensor with respect to the increment of the principal strain.

Appendix D. Solving the nonlinear state equations: Newton’s method with line search

We solve (44) using Newton’s method with a backtracking line search strategy. We start with the Lagrangian function.

$$ \mathcal{L}\left(\boldsymbol{u},\chi \right)=U\left(\boldsymbol{u}\right)+\overset{\sim }{\chi}\left(2{C}_0-{\boldsymbol{f}}_0^T\boldsymbol{u}\ \right) $$
(70)

where \( \overset{\sim }{\chi } \) is the Lagrangian multiplier, which is also the reaction load factor in (43). According to the KKT optimality conditions, we readily obtain:

$$ \left\{\begin{array}{c}\frac{\partial \mathcal{L}}{\partial \boldsymbol{u}}\left({\boldsymbol{u}}^{\ast },\backslash {\mathrm{chi}}^{\ast}\right)=\mathbf{\nabla}U\left({\boldsymbol{u}}^{\ast}\right)-\backslash {\mathrm{chi}}^{\ast }{\boldsymbol{f}}_0=0\\ {}\frac{\partial \mathcal{L}}{\partial \chi}\left({\boldsymbol{u}}^{\ast },\backslash {\mathrm{chi}}^{\ast}\right)=2{C}_0-{\boldsymbol{f}}_0^T{\boldsymbol{u}}^{\ast }=0\end{array}\right. $$
(71)

At iteration k, we interpret the Newton step uk, and the associated multiplier χk + 1, as the solutions of a linearized approximation of the optimality conditions in (71). We substitute uk + uk for u and χk + 1 for χ, and replace the gradient by its linearized approximation near uk, to obtain the equations:

$$ \left\{\begin{array}{c}\mathbf{\nabla}U\left({\boldsymbol{u}}_k+\Delta {\boldsymbol{u}}_k\right)-{\chi}_{k+1}{\boldsymbol{f}}_0\approx \mathbf{\nabla}U\left({\boldsymbol{u}}_k\right)+\mathbf{\nabla}{\mathbf{\nabla}}^TU\left({\boldsymbol{u}}_k\right)\Delta {\boldsymbol{u}}_k-{\chi}_{k+1}{\boldsymbol{f}}_0=0\\ {}2{C}_0-{\boldsymbol{f}}_0^T\ \left({\boldsymbol{u}}_k+\Delta {\boldsymbol{u}}_k\right)=0\end{array}\right. $$
(72)

Since ∇∇TU(uk) = K(uk) and U(uk) = T(uk), then, (72) becomes:

$$ \left\{\begin{array}{c}\boldsymbol{T}\left({\boldsymbol{u}}_k\right)+\boldsymbol{K}\left({\boldsymbol{u}}_k\right)\Delta {\boldsymbol{u}}_k-{\chi}_{k+1}{\boldsymbol{f}}_0=0\\ {}2{C}_0-{\boldsymbol{f}}_0^T\ \left({\boldsymbol{u}}_k+\Delta {\boldsymbol{u}}_k\right)=0\end{array}\right. $$
(73)

Solving for uk using the first equation in the system (73), we obtain:

$$ \Delta {\boldsymbol{u}}_k={\boldsymbol{K}}^{-1}\left({\boldsymbol{u}}_k\right)\left[-\boldsymbol{T}\left({\boldsymbol{u}}_k\right)+{\chi}_{k+1}{\boldsymbol{f}}_0\right] $$
(74)

By means of the equality \( {\boldsymbol{f}}_0^T{\boldsymbol{u}}_k=2{C}_0 \), the second equation in the system (73) becomes:

$$ {\boldsymbol{f}}_0^T\ \Delta {\boldsymbol{u}}_k=0 $$
(75)

Substituting (74) into (75), and solving for χk + 1, we obtain:

$$ {\chi}_{k+1}=\frac{{\boldsymbol{f}}_0^T\Delta {\boldsymbol{u}}_k^{\prime }}{{\mathbf{f}}_0^T\Delta {\boldsymbol{u}}_k^{\prime \prime }}\kern0.5em ,\Delta {\boldsymbol{u}}_k^{\prime }={\boldsymbol{K}}^{-1}\left({\boldsymbol{u}}_k\right)\boldsymbol{T}\left({\boldsymbol{u}}_k\right)\kern0.5em ,\Delta {\boldsymbol{u}}_k^{\prime \prime }={\boldsymbol{K}}^{-1}\left({\boldsymbol{u}}_k\right){\boldsymbol{f}}_0 $$
(76)

By substituting the expression of χk + 1 in (76) into (74), we finally obtain the expression of the Newton step ∆uk as:

$$ \Delta {\boldsymbol{u}}_k=-\Delta {\boldsymbol{u}}_k^{\prime }+{\chi}_{k+1}\Delta {\boldsymbol{u}}_k^{\prime \prime } $$
(77)

For the sake of completeness, the detailed algorithm for Newton’s method, as employed in the present work, is provided in Table 12. The stiffness matrix might become singular near the limit state which can cause numerical difficulties. To prevent the possibility of a singular stiffness matrix, we add a Tikhonov regularization (Tikhonov and Arsenin 1977; Ramos Jr and Paulino 2016) parameter tTK into the tangent stiffness matrix as shown in lines 5 and 6 of Table 12. Through the testing of the numerical examples, we verify that the Tikhonov regularization technique is effective.

Table 12 Newton’s algorithm for solving nonlinear state equations

Appendix E. ABAQUS® UMAT subroutine (ESM)

The ABAQUS® user subroutine UMAT for the surrogate nonlinear elastic constitutive model considering the Drucker–Prager criterion is provided as ESM (Electronic Supplementary Material). A representative example of the supplementary material is presented here. For the Abaqus/CAE usage, please follow this sequence: Property module → Material Editor → General → User Material → Mechanical Constants → Input the user-defined material properties in the sequence of Young’s modulus, Poisson’s ratio, plastic Poisson’s ratio, and uniaxial strength stress. Run analysis with the present UMAT subroutine: Analysis → Edit job → General → User subroutine file.

figure a

Appendix F. Nomenclature

σ stress tensor

ɛ strain tensor

s deviatoric stress tensor

ɛd deviatoric strain tensor

I second-order identity tensor

σi principal stress components

εi principal strain components

λ, μ Lame’s parameters

ϕ1, ϕ2 functions representing the hardening behavior of the material

J1(ɛ) first invariant of the strain tensor

J2(ɛd) second invariant of the deviatoric strain tensor

J1(σ) first invariant of the stress tensor

J2(s) second invariant of the deviatoric stress tensor

σL linear elastic limit of the stress tensor

ɛL linear elastic limit of the strain tensor

σN nonlinear elastic limit of the stress tensor

ɛN nonlinear elastic limit of the strain tensor

E Young’s modulus

υ Poisson’s ratio

υp plastic Poisson’s ratio

σy uniaxial strength stress

β, kspositive constants for an elastic-perfectly-plastic material

φ strain energy density

φL linear elastic strain energy density

c Cohesion

ψ friction angle

η, ζ parameters defined based on the approximation to the Mohr–Coulomb criterion

σc material compressive strength

σt material tensile strength

U structural strain energy

C0 prescribed energy

f0 vector of given applied forces

u nodal displacement vector

ρ vector of element density variables

p constant penalization parameter

Vol Fracvolume fraction

R linear density filter radius

n number of elements discretizing the design domain

ve volume of element e

Vmax maximum material volume

T internal force vector

χ reaction load factor

\( \overset{\sim }{\chi } \)Lagrangian multiplier

KT tangent stiffness matrix

u Newton step

α step size by backtracking line search

J objective function

\( \mathcal{L} \)Lagrangian function

w magnitude of distributed load

x vector of the cross-sectional area for truss members

Le length of truss member e

κ a factor used in the inexact line search approach

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhao, T., Lages, E.N., Ramos, A.S. et al. Topology optimization considering the Drucker–Prager criterion with a surrogate nonlinear elastic constitutive model. Struct Multidisc Optim 62, 3205–3227 (2020). https://doi.org/10.1007/s00158-020-02671-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-020-02671-8

Keywords

Navigation