Skip to main content
Log in

Self-consistent Clustering Analysis-Based Moving Morphable Component (SMMC) Method for Multiscale Topology Optimization

  • Published:
Acta Mechanica Solida Sinica Aims and scope Submit manuscript

Abstract

Current multiscale topology optimization restricts the solution space by enforcing the use of a few repetitive microstructures that are predetermined, and thus lack the ability for structural concerns like buckling strength, robustness, and multi-functionality. Therefore, in this paper, a new multiscale concurrent topology optimization design, referred to as the self-consistent analysis-based moving morphable component (SMMC) method, is proposed. Compared with the conventional moving morphable component method, the proposed method seeks to optimize both material and structure simultaneously by explicitly designing both macrostructure and representative volume element (RVE)-level microstructures. Numerical examples with transducer design requirements are provided to demonstrate the superiority of the SMMC method in comparison to traditional methods. The proposed method has broad impact in areas of integrated industrial manufacturing design: to solve for the optimized macro and microstructures under the objective function and constraints, to calculate the structural response efficiently using a reduced-order model: self-consistent analysis, and to link the SMMC method to manufacturing (industrial manufacturing or additive manufacturing) based on the design requirements and application areas.

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

Similar content being viewed by others

Availability of Data

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

References

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

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  3. Zhou M, Rozvany G. The COC algorithm, part II: topological, geometrical and generalized shape optimization. Comput Methods Appl Mech Eng. 1991;89(1–3):309–36.

    Article  Google Scholar 

  4. Allaire G, Jouve F, Toader A-M. Structural optimization using sensitivity analysis and a level-set method. J Comput Phys. 2004;194(1):363–93.

    Article  MathSciNet  MATH  Google Scholar 

  5. Wang MY, Wang X, Guo D. A level set method for structural topology optimization. Comput Methods Appl Mech Eng. 2003;192(1–2):227–46.

    Article  MathSciNet  MATH  Google Scholar 

  6. Xie YM, Steven GP. A simple evolutionary procedure for structural optimization. Comput Struct. 1993;49(5):885–96.

    Article  Google Scholar 

  7. Li H, Knapik S, Li Y, Park C, Guo J, Mojumder S, et al. Convolution hierarchical deep-learning neural network tensor decomposition (C-HiDeNN-TD) for high-resolution topology optimization. Comput Mech. 2023:1–20.

  8. Liu L, Yan J, Cheng G. Optimum structure with homogeneous optimum truss-like material. Comput Struct. 2008;86(13–14):1417–25.

    Article  Google Scholar 

  9. Deng J, Yan J, Cheng G. Multi-objective concurrent topology optimization of thermoelastic structures composed of homogeneous porous material. Struct Multidiscip Optim. 2013;47:583–97.

    Article  MathSciNet  MATH  Google Scholar 

  10. Niu B, Yan J, Cheng G. Optimum structure with homogeneous optimum cellular material for maximum fundamental frequency. Struct Multidiscip Optim. 2009;39:115–32.

    Article  Google Scholar 

  11. Yan J, Cheng G-d, Liu L. A uniform optimum material based model for concurrent optimization of thermoelastic structures and materials. Int J Simul Multidiscip Design Optim. 2008;2(4):259–66.

    Article  Google Scholar 

  12. Yan X, Huang X, Zha Y, Xie Y. Concurrent topology optimization of structures and their composite microstructures. Comput Struct. 2014;133:103–10.

    Article  Google Scholar 

  13. Camanho P, Bessa M, Catalanotti G, Vogler M, Rolfes R. Modeling the inelastic deformation and fracture of polymer composites–Part II: smeared crack model. Mech Mater. 2013;59:36–49.

    Article  Google Scholar 

  14. de Souza Neto EA, Peric D, Owen DR. Computational methods for plasticity: theory and applications. New York: Wiley; 2011.

    Google Scholar 

  15. Gurson AL. Continuum theory of ductile rupture by void nucleation and growth: part I—yield criteria and flow rules for porous ductile media; 1977.

  16. Sigmund O, Maute K. Topology optimization approaches: a comparative review. Struct Multidiscip Optim. 2013;48(6):1031–55.

    Article  MathSciNet  Google Scholar 

  17. Michel J-C, Suquet P. Computational analysis of nonlinear composite structures using the nonuniform transformation field analysis. Comput Methods Appl Mech Eng. 2004;193(48–51):5477–502.

    Article  MathSciNet  MATH  Google Scholar 

  18. Oskay C, Fish J. Eigendeformation-based reduced order homogenization for failure analysis of heterogeneous materials. Comput Methods Appl Mech Eng. 2007;196(7):1216–43.

    Article  MathSciNet  MATH  Google Scholar 

  19. Roussette S, Michel J-C, Suquet P. Nonuniform transformation field analysis of elastic–viscoplastic composites. Compos Sci Technol. 2009;69(1):22–7.

    Article  Google Scholar 

  20. Yvonnet J, He Q-C. The reduced model multiscale method (R3M) for the non-linear homogenization of hyperelastic media at finite strains. J Comput Phys. 2007;223(1):341–68.

    Article  MathSciNet  MATH  Google Scholar 

  21. Krysl P, Lall S, Marsden JE. Dimensional model reduction in non-linear finite element dynamics of solids and structures. Int J Numer Meth Eng. 2001;51(4):479–504.

    Article  MathSciNet  MATH  Google Scholar 

  22. Guo X, Zhang W, Zhong W. Doing topology optimization explicitly and geometrically—a new moving morphable components based framework. J Appl Mech. 2014;81(8):081009.

    Article  Google Scholar 

  23. Zhang W, Yuan J, Zhang J, Guo X. A new topology optimization approach based on moving morphable components (MMC) and the ersatz material model. Struct Multidiscip Optim. 2016;53:1243–60.

    Article  MathSciNet  Google Scholar 

  24. Sigmund O. Morphology-based black and white filters for topology optimization. Struct Multidiscip Optim. 2007;33:401–24.

    Article  Google Scholar 

  25. Zhang W, Li D, Yuan J, Song J, Guo X. A new three-dimensional topology optimization method based on moving morphable components (MMCs). Comput Mech. 2017;59:647–65.

    Article  MathSciNet  MATH  Google Scholar 

  26. Guo X, Zhou J, Zhang W, Du Z, Liu C, Liu Y. Self-supporting structure design in additive manufacturing through explicit topology optimization. Comput Methods Appl Mech Eng. 2017;323:27–63.

    Article  MathSciNet  MATH  Google Scholar 

  27. Sun J, Tian Q, Hu H, Pedersen NL. Topology optimization of a flexible multibody system with variable-length bodies described by ALE–ANCF. Nonlinear Dyn. 2018;93:413–41.

    Article  MATH  Google Scholar 

  28. Liu Z, Bessa M, Liu WK. Self-consistent clustering analysis: an efficient multi-scale scheme for inelastic heterogeneous materials. Comput Methods Appl Mech Eng. 2016;306:319–41.

    Article  MathSciNet  MATH  Google Scholar 

  29. Liu Z, Fleming M, Liu WK. Microstructural material database for self-consistent clustering analysis of elastoplastic strain softening materials. Comput Methods Appl Mech Eng. 2018;330:547–77.

    Article  MathSciNet  MATH  Google Scholar 

  30. Liu Z, Kafka OL, Yu C, Liu WK. Data-driven self-consistent clustering analysis of heterogeneous materials with crystal plasticity. Advances in computational plasticity: a book in honour of D Roger J Owen. 2018:221–42.

  31. Li Y, Wang Y, Ma R, Hao P. Improved reliability-based design optimization of non-uniformly stiffened spherical dome. Struct Multidiscip Optim. 2019;60:375–92.

    Article  Google Scholar 

  32. Huang O, Saha S, Guo J, Liu WK. An introduction to kernel and operator learning methods for homogenization by self-consistent clustering analysis. Comput Mech. 2023:1–25.

  33. Heinrich C, Aldridge M, Wineman A, Kieffer J, Waas AM, Shahwan K. The influence of the representative volume element (RVE) size on the homogenized response of cured fiber composites. Modell Simul Mater Sci Eng. 2012;20(7): 075007.

    Article  Google Scholar 

Download references

Acknowledgements

We are grateful to all those we have had the pleasure of working with on this manuscript. Special thanks are extended to Dr. Jiaying Gao from Northwestern University for the enlightening discussions we had.

Funding

This research received no specific grant from any funding agency in the public, commercial, or not-for-profit sectors.

Author information

Authors and Affiliations

Authors

Contributions

YL proposed the idea of SMMC, did analyses of the model and was a major contributor in writing the manuscript. JG and HL ran the numerical examples and gave some analysis. HC co-contributed to proposing the idea of the method, instructed the other authors, and also helped revise the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Huihan Chen.

Ethics declarations

Conflict of interests

The authors declare that they have no competing interests.

Appendices

Appendix A: Self-consistent Clustering Analysis (SCA)

In order to predict the mechanical behaviors of microstructures, finite element and fast Fourier transform (FFT) methods are often used to analyze the mechanical response of representative volume elements (RVE). However, the huge computational cost associated with solving the microscale model at each integration point limits the application of these methods. To address this issue, Liu et al. proposed the self-consistent clustering analysis (SCA) based on a reduced-order RVE model [28]. Compared to the original DNS, SCA offers a great improvement in computational efficiency by significantly reducing the degrees of freedom through compressing the microstructural data into clusters in the offline stage. Furthermore, material nonlinear constitutive laws (such as plasticity and damage) can be obtained with the reduced-order model in the online stage. The basic idea behind SCA is as follows.

Offline Stage

In offline stage, the relationship between the elastic microscopic strain \({{\varvec{\varepsilon}}}^{\text{micro}}({\varvec{x}})\) and the homogeneous elastic macroscopic strain \({{\varvec{\varepsilon}}}^{\text{macro}}\) at each material point can be obtained using the strain concentration tensor A(x), that is

$${{\varvec{\varepsilon}}}^{\text{micro}}({\varvec{x}})={\varvec{A}}({\varvec{x}}):{{\varvec{\varepsilon}}}^{\text{macro}}$$
(A1)

where \({{\varvec{\varepsilon}}}^{\text{macro}}\) is the elastic macroscopic strain on the RVE with the boundary conditions, \({{\varvec{\varepsilon}}}^{\text{micro}}({\varvec{x}})\) is the elastic local strain at point x in the microscale domain Ω. Based on the strain concentration tensor Am(x), the voxels are grouped into a few clusters using k-means clustering.

After clustering, the interaction tensors \({{\varvec{D}}}^{IJ}\), representing the influence of stress in the jth cluster on the strain in the ith cluster, can be computed. They can be written as an integral of Green’s function in an RVE domain Ω with periodic boundary conditions:

$${{\varvec{D}}}^{IJ}=\frac{1}{{c}^{I}|\Omega |}{\int }_{\Omega }{\int }_{\Omega }{\chi }^{I}({\varvec{x}}){\chi }^{J}\left({{\varvec{x}}}^{\boldsymbol{^{\prime}}}\right){\boldsymbol{\Phi }}^{0}\left({\varvec{x}},{{\varvec{x}}}^{\boldsymbol{^{\prime}}}\right)\text{d}{{\varvec{x}}}^{\boldsymbol{^{\prime}}}\text{d}{\varvec{x}}$$
(A2)

where \({\boldsymbol{\Phi }}^{0}\left({\varvec{x}},{{\varvec{x}}}{\prime}\right)\) is the fourth-order periodic Green’s function associated with an isotropic linear elastic reference material with stiffness tensor C0. \(c^{{\varvec{I}}}\) is the volume fraction of the ith cluster, and \(|\Omega |\) is the volume of the RVE domain. \({\chi }^{I}\)=1 if x is in the ith cluster. Moreover, \({\boldsymbol{\Phi }}^{0}\left({\varvec{x}},{{\varvec{x}}}{\prime}\right)\) has a simple form in the Fourier space as

$${\widehat{{\varvec{\phi}}}}^{0}\left({\varvec{\xi}}\right)=\frac{1}{4{\mu }^{0}}{\widehat{{\varvec{\phi}}}}^{1}\left({\varvec{\xi}}\right)+\frac{{\lambda }^{0}+{\mu }^{0}}{{\mu }^{0}({\lambda }^{0}+2{\mu }^{0})}{\widehat{{\varvec{\phi}}}}^{2}\left({\varvec{\xi}}\right)$$
(A3)

with

$${\widehat{\phi }}_{ijkl}^{1}\left({\varvec{\xi}}\right)=\frac{1}{{\left|{\varvec{\xi}}\right|}^{2}}({\delta }_{ik}{\xi }_{j}{\xi }_{l}+{\delta }_{il}{\xi }_{j}{\xi }_{k}+{\delta }_{jl}{\xi }_{i}{\xi }_{k}+{\delta }_{jk}{\xi }_{i}{\xi }_{l})$$
(A4)
$${\widehat{\phi }}_{ijkl}^{2}\left({\varvec{\xi}}\right)=-\frac{{\xi }_{i}{\xi }_{j}{\xi }_{k}{\xi }_{l}}{|\varvec{\xi} {|}^{4}}$$
(A5)

where \({\lambda }^{0}\) and \({\mu }^{0}\) are Lamé constants of the reference material, \({\varvec{\xi}}\) is the Fourier space corresponding to x in real space, and \({\delta }_{ij}\) is the Kronecker delta function.

Online Stage

The equilibrium condition in the RVE can be rewritten as a continuous Lippmann–Schwinger integral equation by introducing a homogeneous isotropic linear elastic reference material [28] as follows

$$\begin{aligned}\begin{aligned}\Delta {\varvec{\varepsilon}}_{m}({\varvec{x}})+&\int {\boldsymbol{\Phi}}^{0}\left({\varvec{x}},{\varvec{x}}{\prime}\right):\left[\Delta \sigma_{m}\left({\varvec{x}}{\prime}\right)-{\varvec{C}}^{0}\right.\\&\left.:\Delta {\varvec{\varepsilon}}_{m}\left({\varvec{x}}{\prime}\right)\right]{\textrm{d}}{\varvec{x}}{\prime}-\Delta {\varvec{\varepsilon}}^{0}=\varvec{0}\end{aligned}\end{aligned}$$
(A6)

where \(\Delta {{\varvec{\varepsilon}}}^{0}\) is the far-field strain increment controlling the evolution of the local strain. The far-field strain is uniform in the RVE. The stiffness tensor \({{\varvec{C}}}^{0}\) of the isotropic linear elastic reference material can be described using Lamé parameters \({\lambda }^{0}\) and \({\mu }^{0}\). By averaging the incremental integral equation and considering the effect of macro-strain constraints or macro-stress constraints, Eq. (A6) can be rewritten. For more details, please refer to [28]. Following the clustering idea in the offline stage, the discretized integral equation of the ith cluster can be derived as

$$\begin{aligned}&\Delta \varvec{\varepsilon }_{m}^{I}+\sum_{I=1}^{k}{{\varvec{D}}}^{IJ}:\left[\Delta \varvec{\sigma }_{m}^{J}-{{\varvec{C}}}^{0}:\Delta \varvec{\varepsilon }_{m}^{J}\right]\\ & \quad -\Delta \varvec{\varepsilon }^{0}=\mathbf{0} \end{aligned}$$
(A7)

where \(\Delta {{\varvec{\varepsilon}}}_{m}^{I}\) and \(\Delta {{\varvec{\sigma}}}_{m}^{J}\) are the microscopic strain and stress increments in the jth cluster, respectively. \({{\varvec{D}}}^{IJ}\) is the interaction tensor computed from the offline stage. After discretization, the far-field strain increment, which is equal to the average strain in the RVE, can be expressed as

$$ \Delta \varvec{\varepsilon} ^{0} = \sum\limits_{{I = 1}}^{k} {c^{I} \Delta \varvec{\varepsilon} _{m}^{I} } $$
(A8)

For the macro-strain constraint, the discrete form can be written as

$$ \mathop \sum \limits_{{I = 1}}^{k} c^{I} \Delta {\varvec{\varepsilon }}_{m}^{I} = \Delta {\varvec{\varepsilon }}_{M} \,{\text{or}}\,\Delta {\varvec{\varepsilon} }^{0} = \Delta {\varvec{\varepsilon }}_{M} $$
(A9)

where \(\Delta {{\varvec{\varepsilon}}}_{M}\) is the macroscopic strain increment. Similarly, the discrete form of macro-stress constraint is

$$ \sum\limits_{{I = 1}}^{k} {c^{I} \Delta \varvec{\sigma} _{m}^{I} } = \Delta \varvec{\sigma} _{M} $$
(A10)

where \(\Delta {{\varvec{\sigma}}}_{M}\) is the macroscopic strain increment. For the continuous Lippmann–Schwinger equation in Eq. (A6), the stiffness tensor of the reference material \({{\varvec{C}}}^{0}\) does not influence the solution. However, once the equation is transformed into a discrete form, the equilibrium condition is not strictly satisfied at each point in the RVE, and the solution depends on \({{\varvec{C}}}^{0}\). Therefore, the effect of \({{\varvec{C}}}^{0}\) on the reduced-order system needs to be considered.

To build a \({{\varvec{C}}}^{0}\)-based iteration, a self-consistent scheme is used in the online stage [28]. The stiffness tensor of the reference material \({{\varvec{C}}}^{0}\) is approximately the same as the macroscale tangent stiffness tensor \(\bar{{\varvec{C}}}\).

Appendix B: Moving Morphable Component (MMC) Method

The framework of SMMC is based on the idea of moving morphable component (MMC) method proposed by Zhang et al. [23] This section gives a short review of the MMC method.

In the MMC method, the morphable structural components, which can move, deform, and overlap freely, are used as the primary building blocks for topology optimization. The geometries/topologies of structural components are described using a topology description function (TDF) with a set of explicit parameters. In the MMC framework, topology optimization is achieved by optimizing the shapes, lengths, widths, orientations, and layout of the components. For the ith component in the structure, the TDF can be expressed as [22]

$$\chi^{i} \left( \varvec{x} \right) = 1 - \left[ {\left( {\frac{{\left( {x - x_{{i0}} } \right)\cos \theta _{i} + \left( {y - y_{{i0}} } \right)\sin \theta _{i} }}{{L_{i} }}} \right)^{m} + \left( {\frac{{ - \left( {x - x_{{i0}} } \right)\sin \theta _{i} + \left( {y - y_{{i0}} } \right)\cos \theta _{i} }}{{t_{i} }}} \right)^{m} } \right] $$
(B1)

where \(\left({x}_{i0},{y}_{i0}\right)\) represents the coordinates of the center of the component, \({L}_{i}\), \({\theta }_{i}\), and \({t}_{i}\) denote the half-length, inclined angle, and variable thickness of the component, respectively, as shown in Fig. 2. \(m\) is a large even integer number (we take p = 6). With the use of \({\chi }^{i}({\varvec{x}})\), the region \({\Omega }^{i}\) occupied by the ith component can be expressed as

$$\left\{\begin{array}{l}{\chi }^{i}\left(x\right)>0,\mathrm{ if} x\in {\Omega }^{i}\\ {\chi }^{i}\left(x\right)=0, \mathrm{if} x\in \partial {\Omega }^{i}\\ {\chi }^{i}\left(x\right)<0, \mathrm{if} x\in D\backslash \left({\Omega }^{i}\cup \partial {\Omega }^{i}\right)\end{array}\right.$$
(B2)

Then, the corresponding structural topology can be described as

$$\left\{\begin{array}{l}{\chi }^{s}\left(x\right)>0, \mathrm{if} x\in {\Omega }^{i}\\ {\chi }^{s}\left(x\right)=0, \mathrm{if} x\in \partial {\Omega }^{i}\\ {\chi }^{s}\left(x\right)<0, \mathrm{if} x\in D\backslash \left({\Omega }^{i}\cup \partial {\Omega }^{i}\right)\end{array}\right.$$
(B3)

where \({\Omega }^{s}={\cup }_{i=1}^{n}{\Omega }^{i}\) is the region occupied by the solid structural components, and \({\chi }^{s}\) is the TDF of the whole structure, which can be formulated as [8]

$${\chi }^{s}({\varvec{x}})={\max}({\chi }^{1},{\chi }^{2},...,{\chi }^{n})$$
(B4)

where n is the total number of components in the structure.

Based on the MMC topology optimization method, the topology optimization problem can be formulated as [22, 23]:

$$ \begin{aligned} & {\text{Find}}\,\varvec{d} = \left( {\left( {\varvec{d}^{1} } \right)^{{ \top }} , \ldots ,\left( {\varvec{d}^{\varvec{i}} } \right)^{{ \top }} , \ldots ,\left( {\varvec{d}^{\varvec{n}} } \right)^{{ \top }} } \right)^{{ \top }} ,\varvec{b} = \left( {\left( {\varvec{b}^{1} } \right)^{{ \top }} , \ldots ,\left( {\varvec{b}^{\varvec{i}} } \right)^{{ \top }} , \ldots ,\left( {\varvec{b}^{\varvec{n}} } \right)^{{ \top }} } \right)^{{ \top }} ,\varvec{~u}\left( {\varvec{x}} \right) \\ & {\text{Minimize}}\,C = \mathop \sum \limits_{{i = 1}}^{n} \mathop \int \limits_{{\Omega ^{i} \backslash \;\left( { \cup _{{1\; \le j < i}} \;\left( {\;\Omega ^{i} \cap \Omega ^{j} } \right)} \right)\;}} \varvec{f}^{\varvec{i}} \cdot \varvec{u}{\text{d}}V + \mathop \int \limits_{{\Gamma _{{\mathfrak{t}}} }} \varvec{t} \cdot \varvec{u}{\text{d}}S \\ & {\text{s}}{\text{.t}}{\text{.}} \\ & g_{j} \left( \varvec{d} \right) \le 0,j = 1,...,m \\ & \varvec{d} \subset {\mathcal{U}}_{d} \\ \end{aligned}$$
(B5)

where \({\varvec{d}}_{i}\) = (\({x}_{i0},{y}_{i0}\),\({L}_{i}\), \({t}_{1i}\), \({t}_{2i}\), \({t}_{3i}\),\({\theta }_{i}\)) (\(i=1,\dots ,n\)) denotes the vector of design variables for the ith component. \({g}_{j}\)(\(j=1,\dots ,m\)) is the constraint function/functional, and \({\mathcal{U}}_{d}\) represents the admissible set of d.

For the compliance optimization problem, Eq. (B5) can be rewritten as:

$$ \begin{gathered} {\text{Find}}\,\varvec{d} = \left( {\left( {\varvec{d}^{1} } \right)^{ \top } , \ldots ,\left( {\varvec{d}^{i} } \right)^{ \top } , \ldots ,\left( {\varvec{d}^{n} } \right)^{ \top } } \right)^{ \top } ,\varvec{u}\left( {\varvec{x}} \right) \hfill \\ {\text{Minimize}}\,C = \mathop \sum \limits_{{i = 1}}^{n} \mathop \int \limits_{{\Omega ^{i} \backslash \;\left( { \cup _{{1\; \le j < i}} \;\left( {\;\Omega ^{i} \cap \Omega ^{j} } \right)} \right)\;}} \varvec{f}^{i} \cdot \varvec{u}{\text{d}}V + \mathop \int \limits_{{\Gamma _{{\mathfrak{t}}} }} \varvec{t} \cdot \varvec{u}{\text{d}}S \hfill \\ {\text{s}}.{\text{t}}. \hfill \\ \mathop \sum \limits_{{i = 1}}^{n} \mathop \int \limits_{{\Omega ^{i} \backslash \;\left( { \cup _{{1\; \le j < i}} \;\left( {\;\Omega ^{i} \cap \Omega ^{j} } \right)} \right)}} \varvec{E}^{\varvec{i}} :\varvec{\varepsilon }\left( \varvec{u} \right):\varvec{\varepsilon }\left( \varvec{v} \right){\text{d}}V \hfill \\ \quad = \mathop \sum \limits_{{\varvec{i} = 1}}^{\varvec{n}} \mathop \int \limits_{{\Omega ^{i} \backslash \;\left( { \cup _{{1\; \le j < i}} \;\left( {\;\Omega ^{i} \cap \Omega ^{j} } \right)} \right)}} \varvec{f} \cdot \varvec{v}{\text{d}}V+\mathop \int \limits_{{\Gamma _{t} }} \varvec{t} \cdot \varvec{v}{\text{d}}S,\varvec{~}\forall \varvec{v} \in {\mathbf{\mathcal{U}}}_{{{\text{ad}}}} \hfill \\ \quad V\left( \varvec{d} \right) \le \bar{V} \hfill \\ \quad \varvec{d} \subset {\mathcal{U}}_{d} \hfill \\ \quad \varvec{u} = {\bar{\varvec{u}}},\,{\text{on}}\,\Gamma _{u} \hfill \\ \end{gathered} $$
(B6)

In Eq. (B6), \(\bar{{{\varvec{u}}}}\) is the prescribed displacement on the Dirichlet boundary \({\Gamma }_{u}\). The symbol \({\varvec{\varepsilon}}\) is the second-order linear strain tensor. \({\varvec{f}}^{i}\) is the body force density in \({\Omega }^{i}\), while t is the surface traction on the Neumann boundary \({\Gamma }_{t}\). \({\varvec{u}}\) and \({\varvec{v}}\) are the displacement field and the corresponding test function defined on \(\Omega ={\cup }_{i=1}^{n}{\Omega }^{i}\). The fourth-order isotropic elasticity tensor of the ith component’s material is expressed as \({{\varvec{E}}}^{{\varvec{i}}}={E}^{i}/(1+{\nu }_{i})\left[{\varvec{I}}+{\nu }^{i}/\left(1-2{\nu }^{i}\right){\varvec{\delta}}\otimes{\varvec{\delta}}\right]\) (\({\varvec{I}}\) and \({\varvec{\delta}}\) are the fourth- and second-order identity tensors, respectively). The upper bound of the available volume of solid material is defined as\(\bar{V}\). \({{\varvec{E}}}^{i}\) is Young’s modulus, and \({\nu }^{i}\) is Poisson’s ratio. When \({\varvec{x}}\in {\Omega }^{i}\backslash \hspace{0.33em}({\cup }_{1\hspace{0.33em}\le j<i}\hspace{0.33em}(\hspace{0.33em}{\Omega }^{i}\cap {\Omega }^{j}))\), the assumption is made that \({\varvec{E}}({\varvec{x}})={{\varvec{E}}}^{{\varvec{i}}}\) and\({\varvec{f}}={{\varvec{f}}}^{{\varvec{i}}}\). Besides, it is assumed that \({{\varvec{E}}}^{{\varvec{i}}}=\cdots ={{\varvec{E}}}^{{\varvec{n}}}={\varvec{E}}\) and \({\nu }^{i}=\cdots ={\nu }^{n}=\nu \). To simplify the problem, \(\bar{{{\varvec{u}}}}=0\) is set in the following discussion.

If only single-phase material is taken into account, Eq. (B6) can be transferred into the following equations.

$$ \begin{aligned} & {\text{Find}}\,{\varvec{d}} = \left( {\left( {{\varvec{d}}^{1} } \right)^{{ \top }} , \ldots ,\left( {{\varvec{d}}^{{\varvec{i}}} } \right)^{{ \top }} , \ldots ,\left( {{\varvec{d}}^{{\varvec{n}}} } \right)^{{ \top }} } \right)^{{ \top }} ,\\ & {\varvec{b}} = \left( {\left( {{\varvec{b}}^{1} } \right)^{{ \top }} , \ldots ,\left( {{\varvec{b}}^{{\varvec{i}}} } \right)^{{ \top }} , \ldots ,\left( {{\varvec{b}}^{{\varvec{n}}} } \right)^{{ \top }} } \right)^{{ \top }} ,\varvec{ u}\left( {{\varvec{x}}} \right) \\ & {\text{Minimize}}\,C = \mathop \int \limits_{{\Omega }} H\left( {\phi^{s} \left( {{\varvec{x}};{\varvec{d}}} \right)} \right){\varvec{f}} \cdot {\varvec{u}}{\text{d}V} + \mathop \int \limits_{{\Gamma_{{{t}}} }} {\varvec{t}} \cdot {\varvec{u}}{\text{d}S} \\ & {\text{s}}.{\text{t}}. \\ & \mathop \int \limits_{\Omega } \left( {H\left( {\phi^{s} \left( {{\varvec{x}};{\varvec{d}}} \right)} \right)} \right)^{q} {\varvec{E}}:{\varvec{\varepsilon}}\left( {\varvec{u}} \right):{\varvec{\varepsilon}}\left( {\varvec{v}} \right){\text{d}V} \\ & \quad = \mathop \int \limits_{\Omega } H\left( {\phi^{s} \left( {{\varvec{x}};{\varvec{d}}} \right)} \right){\varvec{f}} \cdot {\varvec{v}}{\text{d}{V + }}\mathop \int \limits_{{\Gamma_{t} }} {\varvec{t}} \cdot {\varvec{v}}{\text{d}S},\varvec{ }\forall {\varvec{v}} \in {{\mathcal{\bf U}}}_{{{\text{ad}}}} \\ & \varvec{ }\mathop \int \limits_{\Omega } H\left( {\phi^{s} \left( {{\varvec{x}};{\varvec{d}}} \right)} \right){\text{d}V} \le \bar{V} \\ & {\varvec{d}} \subset {\mathcal{U}}_{d} \\ & {\varvec{u}} = \bar{\varvec{u}},\,{\text{on}}\,\Gamma_{u} \\ \end{aligned} $$
(B7)

where \(H=H(x)\) is the Heaviside function, and we take q as an integer lager than 1.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, Y., Guo, J., Li, H. et al. Self-consistent Clustering Analysis-Based Moving Morphable Component (SMMC) Method for Multiscale Topology Optimization. Acta Mech. Solida Sin. 36, 884–898 (2023). https://doi.org/10.1007/s10338-023-00433-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10338-023-00433-9

Keywords

Navigation