You are on page 1of 31

Structural and Multidisciplinary Optimization (2023) 66:195

https://doi.org/10.1007/s00158-023-03648-z

RESEARCH PAPER

Thermodynamically consistent concurrent material and structure


optimization of elastoplastic multiphase hierarchical systems
Tarun Gangwar1,2   · Dominik Schillinger2

Received: 28 April 2023 / Revised: 14 July 2023 / Accepted: 27 July 2023


© The Author(s) 2023

Abstract
The concept of concurrent material and structure optimization aims at alleviating the computational discovery of optimum
microstructure configurations in multiphase hierarchical systems, whose macroscale behavior is governed by their micro-
structure composition that can evolve over multiple length scales from a few micrometers to centimeters. It is based on the
split of the multiscale optimization problem into two nested sub-problems, one at the macroscale (structure) and the other at
the microscales (material). In this paper, we establish a novel formulation of concurrent material and structure optimization
for multiphase hierarchical systems with elastoplastic constituents at the material scales. Exploiting the thermomechanical
foundations of elastoplasticity, we reformulate the material optimization problem based on the maximum plastic dissipa-
tion principle such that it assumes the format of an elastoplastic constitutive law and can be efficiently solved via modified
return mapping algorithms. We integrate continuum micromechanics based estimates of the stiffness and the yield criterion
into the formulation, which opens the door to a computationally feasible treatment of the material optimization problem. To
demonstrate the accuracy and robustness of our framework, we define new benchmark tests with several material scales that,
for the first time, become computationally feasible. We argue that our formulation naturally extends to multiscale optimiza-
tion under further path-dependent effects such as viscoplasticity or multiscale fracture and damage.

Keywords  Multiphase topology optimization · Concurrent design · Continuum micromechanics · Homogenization ·


Elastoplasticity · Path-dependent optimization

1 Introduction et al. 2014; Fratzl and Weinkamer 2007; Ritchie et al. 2009;
Bhushan 2009; Egan et al. 2015). In other words, biologi-
Multiphase hierarchical systems apply the concept of micro- cal materials adapt their form (or shape/structure) against
heterogeneity repetitively across a hierarchy of well-sepa- the dynamic external environment and improve the micro-
rated length scales: composite microstructures at a smaller structure architecture, fulfilling the local needs imposed by
scale form the base constituents for new microstructures at physiological, phylogenetic, and reproductive constraints
the next larger scale. This principle constitutes the backbone (Wolff 1986; Gibson 2012; Gao et al. 2003). A rational
of virtually all biological materials, enabling them to com- understanding of microstructure interdependencies across
bine various functional properties at different length scales hierarchical scales on the macroscale properties helps pave
with favorable mechanical properties at the macroscale the way forward to many engineering applications involving
through evolutionary mechanisms (Wegst et al. 2015; Zheng biological materials such as the genetic tailoring of crops
(Brulé et al. 2016; McCann et al. 2014), bone remodeling
(Rodrigues et al. 1999; Blanchard et al. 2016), and the fab-
Responsible Editor: Jun Wu
rication of bioinspired engineering materials (Wegst et al.
* Tarun Gangwar 2015; Holstov et al. 2015).
tarun.gangwar@ce.iitr.ac.in Multiscale modeling of hierarchical materials in con-
junction with structural optimization methods constitutes
1
Department of Civil Engineering, Indian Institute a promising pathway to elucidate optimum microstructure
of Technology Roorkee, Roorkee, India
configurations in multiphase hierarchical systems. In this
2
Institute for Mechanics, Computational Mechanics Group, context, recently developed concurrent multiscale analysis
Technical University of Darmstadt, Darmstadt, Germany

13
Vol.:(0123456789)
195   Page 2 of 31 T. Gangwar, D. Schillinger

and topology optimization methods (Xia and Breitkopf Wu et al. (2021) that extensively covers existing approaches
2014, 2015; Rodrigues et  al. 2002; Coelho et  al. 2008; for designing hierarchical structures for linear end-compli-
Nakshatrala et al. 2013; Da et al. 2017; Zhang et al. 2018) ance minimization problems. The variational structure of the
naturally fit to the dual optimization of structure (form) and displacement-based formulation of end-compliance optimi-
material (microstructure architecture). The idea is to decom- zation corresponds to a saddle point problem with respect
pose the multiscale problem into two nested sub-problems, to the admissible set for design variables and the space of
one at the macroscale (structure) and the other at the micro- kinematically admissible displacements (Lipton 1994). With
scale (material). At each macroscale material point, the pointwise definitions of material design variables, the sad-
microscale sub-problem provides a locally optimal mate- dle point property enables a natural decomposition into the
rial response and can be interpreted as the reformulation material and structure optimization subproblems (Jog et al.
of a material constitutive law for the macroscale structure 1994). The equivalent variational structure for combined
optimization problem. The material optimization problem is analysis and optimization of path-dependent problems that
typically solved within an FE2 type computational homog- consider the complete deformation process is non-trivial and
enization framework (Feyel and Chaboche 2000; Fish 2013) yet to be investigated, which is also one of the conclusions
at each macroscale Gauss point. Other approaches not based in the current review article (Wu et al. 2021).
on scale separation or periodicity such as Nguyen and Schil- The second critical roadblock is the computational cost
linger (2019) also seem possible. We note that throughout for multiscale analysis through computational homogeniza-
the article, we will use the term multiphase hierarchical tion that for multiphase hierarchical systems grows exponen-
system for the combined representation of the multiphase tially with each scale characterization (Yuan and Fish 2009;
hierarchical material and the macrostructure domain that Le et al. 2015; Liu et al. 2016; Bessa et al. 2017). Adding
habitats it. the topology optimization at the structure level results in
The base constituents in multiphase hierarchical systems even higher computational cost, since it requires solving
often exhibit elastoplastic material properties resulting in many multiscale problems for different realizations of the
a path-dependent macroscale mechanical response and structure topology during a typical optimization algorithm.
dissipation-driven self-adapting mechanisms. In the case This drawback limits existing approaches to small two-scales
of path-dependent problems, efficient methods for the com- problems, even in the simplest case of hierarchical materials
putational optimization of multiphase hierarchical systems with linear elastic constituents. Continuum micromechanics
are still in its infancy (Da 2019). In the context of fracture provides a rigorous framework to derive analytical estimates
resistance and damage, recent contributions have proposed of macroscale elastoplastic properties (Zaoui 2002; Suquet
structure optimization methods optimizing the inclusion 2014; Morin et al. 2017) and has been successfully applied
characteristics in matrix-inclusion type multiphase materi- to describe many multiphase hierarchical systems such as
als for path-dependent objective functions (Xia et al. 2018; plant, wood, bone, and cementitious materials (Gangwar
Li et al. 2021; Kato 2010; Kato and Ramm 2013; Hilchen- and Schillinger 2019; Gangwar et al. 2021; Hofstetter et al.
bach and Ramm 2015). The corresponding optimization 2005; Hellmich et al. 2004; Fritsch et al. 2009; Pichler and
formulations, however, remain in the format of a monoscale Hellmich 2011). In the context of concurrent material and
design, where all the morphological design parameters of the structure optimization, we recently integrated continuum
material are represented at the structure scale. In this case, micromechanics based homogenized estimates in end-com-
the structure scale discretization is dictated by the smallest pliance optimization problems, which rendered our frame-
length scale of constituents, making these methods com- work computationally tractable for multiphase hierarchical
putationally prohibitive for multiphase hierarchical systems systems with several material length scales (Gangwar and
with several well-separated length scales. To the best of our Schillinger 2021).
knowledge, no work has been reported so far in the literature In this article, we establish, for the first time, a thermo-
on a decomposed concurrent material and structure optimi- dynamically consistent formulation of concurrent material
zation formulation for path-dependent problems involving and structure optimization, including suitable sub-problem
elastoplastic multiphase hierarchical systems. formulations for multiphase hierarchical systems with elas-
The first major roadblock in the context of elastoplastic toplastic constituents at the material scales. The structure
behavior across hierarchical scales is the non-trivial prob- optimization problem addresses the macroscale density dis-
lem decomposition into material and structure optimization tribution, while the material optimization problem at each
subproblems. The current state of concurrent material and material point seeks the optimized macroscale response with
structure optimization methods focuses only on end-com- respect to microscale (material) design variables. In par-
pliance type optimization problems with an overall linear ticular, we reformulate the material optimization problem
elastic response at both the material and structure levels. based on the maximum plastic dissipation principle such
Interested readers are referred to a recent review article by that it assumes the format of an elastoplastic constitutive law

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 3 of 31  195

that can be efficiently solved via modified return mapping x is denoted as 𝜌(x) . The domain Ω is subjected to a traction
algorithms. To focus on the key concepts of the decomposed ̄t(t) at the Neumann boundary ΓN and the prescribed displace-
formulation, we make a few assumptions on the macroscale ments ū E (t) at the Dirichlet boundary ΓD with a body force
behavior of the inelastic hierarchical materials including iso- b(x, t) , where t ∈ [0, T] . Then, the macroscale displacement
thermal processes, the existence of associative flow rule, and field u(x,
̄ t) at a material point x and at time t ∈ [0, T] is a
rate-independent ideal elastoplastic response. mapping ū ∶ Ω × [0, T] → ℝ3 . We define the corresponding
We express the homogenized stiffness and yield cri- velocity and strain fields at (x, t) ∈ Ω × [0, T] as
terion as a function of material design variables within a ( 𝜕 u(x,
𝜕 u(x,
̄ t) ̄ t) )
continuum micromechanics framework, which enables the v(x, t) ∶= and E(x, t) ∶= sym , (1)
computationally tractable treatment of our optimization 𝜕t 𝜕x
formulation. In particular, we focus on a quadratic stress where sym(◻) represents the symmetric part of a second-
average micromechanical approach for estimating homog- order tensor.
enized yield criterion, which has been successfully used in With the kinematically admissible velocity field v(x, t) and
modeling the limit strength of a broad range of hierarchical the macroscale stress field 𝚺(x, t) , the mechanical work iden-
materials such as metal-matrix composite, cement-mortar, tity is
wood, and crop stems (Suquet 1997; Pichler and Hellmich
2011; Hofstetter et al. 2008; Gangwar et al. 2021). d
T(v) + Pint (𝚺, v) = Pext (v) ∀t ∈ [0, T], (2)
Our article is organized as follows: In Sect. 2, we briefly dt
review the relevant thermomechanical principles of elasto- where
plasticity along with multiscaling concepts in continuum
micromechanics, which form the basis of our further devel-
2 ∫Ω
1
kinetic energy T(v) = 𝜌|v|2 dΩ,
opments. In Sect. 3, we formulate the path-dependent stiff-
ness maximization problem, decomposing material and
∫Ω
𝜕v
stress power Pint (𝚺, v) = 𝚺∶ dΩ,
structure optimization sub-problems for elastoplastic mul- 𝜕x
tiphase hierarchical systems. We then describe its discretiza- (3)
∫Ω
tion within the framework of the finite element method. In external power Pext (v) = b(x, t) ⋅ v dΩ +
Sect. 4, we develop an algorithmic procedure for the mate-
rial optimization problem based on the maximum plastic dis- ∫ ΓN
̄t(t) ⋅ v ds.
sipation principle. In Sect. 5, we consolidate all our devel-
opments in an algorithmic framework and provide pertinent This identity directly comes from the application of the prin-
implementation details. Finally, we verify our framework ciple of virtual power (PVP) with a specific choice of the
with benchmark problems in Sect. 6. velocity field as the test function (Simo and Hughes 2006).
The PVP is a fundamental exposition and can be applied for
various applications ranging from, for instance, modeling
2 A brief review of fundamental concepts DNA macromolecules to elastic foundations (Kalliauer et al.
2020; Höller et al. 2019). Please refer to Germain for sys-
2.1 Thermomechanical formulation temic application of PVP to derive fundamental equations
of elastoplasticity in continuum mechanics (Germain 1973).

We start by briefly reviewing the basic principles of elasto- Remark 1  We emphasize that we chose (◻) ̄ notation for the
plasticity from a thermodynamics viewpoint, including the macroscale displacement ū  , and the given boundary condi-
mechanical work identity, the notion of plastic dissipation tions ū E and ̄t . Later in this article, the displacement solution
from the second law of thermodynamics, and the derivation and boundary conditions drive the optimization algorithm,
of the constitutive equations for associative plasticity reflect- and, therefore, this choice directly relates with the intro-
ing on the principle of maximum plastic dissipation. We duced notations for the sought solutions for the optimized
primarily follow the exposition of Simo and Hughes (2006). multiscale configurations.

2.1.1 The mechanical work identity 2.1.2 Constitutive relations from the second law


of thermodynamics
We consider a macroscale initial boundary value problem
defined on a domain Ω and restrict our attention to a time The constitutive relations between the macroscale stress
interval [0, T] . The position of a material point in the domain 𝚺(x, t) and the macroscale displacement field u(x,
̄ t) (through
Ω is denoted by x . The macroscale density at a material point

13
195   Page 4 of 31 T. Gangwar, D. Schillinger

the macroscale strains E(x, t) ) close the global governing energy function with respect to the elastic part of the strain
equations stated in (2) and (3). The second law of thermo- tensor. The second equation (8) represents the irreversible
dynamics governs the form of these constitutive relations, nature of an elastoplastic process implying that the dissipa-
and we summarize the important results in the following. tion energy is always non-negative. This relation constrains
First, we decompose the macroscale strain tensor E into an the possible stress states a material can undergo and indi-
elastic and plastic part assuming small strains, denoted by cates that the stress depends on the rate of the plastic part
Ee and Ep , as of the strain tensor.

E = Ee + Ep . (4)
2.1.3 The principle of maximum plastic dissipation
We introduce the notion of internal potential energy and
dissipation within the context of elastoplasticity. We define The principle of maximum plastic dissipation is a corner-
the internal energy of the system as stone of the mathematical formulation of associative plas-
ticity. In the following, we derive the material constitutive
∫Ω
Vint = Ψ(Ee ) dΩ, (5) equations for perfect plasticity from the viewpoint of this
principle. We first assume a yield criterion 𝔉(𝝉) , with 𝝉 ∈ 𝕊
where Ψ(Ee ) is the Helmholtz free energy density defined in denoting any possible stress state. Its zero isosurface is the
terms of the stored elastic energy function W and the con- usually convex yield surface that encloses the space of
tributions from hardening effects. In this presentation, we admissible stresses

𝔼Σ ∶= 𝝉 ∈ 𝕊 | 𝔉(𝝉) ≤ 0 .
consider the case of perfect plasticity, which implies that the { }
contribution from hardening is zero and Ψ = W . (9)
Next, we look at the difference between the stress power
Pint (𝚺, v) and the rate of change of the internal energy Vint , For a given plastic strain Ep ∈ 𝕊 , we define the plastic dis-
which we denote by Dmech . Assuming isothermal conditions, sipation Dp at a material point for perfect plasticity as
the Clausius–Duhem version of the second law of thermo- p p
Dp [𝝉; Ė ] ∶= 𝝉 ∶ Ė . (10)
dynamics (Truesdell and Noll 2004; Tadmor et al. 2012)
follows as where 𝝉 ∈ 𝔼Σ now denotes an admissible stress state.
In the local form, the principle of maximum plastic dissi-
V ≥ 0 ∀t ∈ [0, T].
d
Dmech ∶= Pint (𝚺, v) − (6) pation states that, for a given plastic strain Ep ∈ 𝕊 , the plas-
dt int
tic dissipation Dp attains its maximum for the actual stress
We identify Dmech as the total instantaneous mechanical dis- tensor 𝚺 among all possible stresses 𝝉 ∈ 𝔼Σ . Mathematically,
sipation in the domain Ω at time t ∈ [0, T] , which is always the principle is
non-negative. p
{
p
}
Inserting the definitions of Pint (𝚺, v) and Vint from (3) and Dp [𝚺; Ė ] = max Dp [𝝉; Ė ] . (11)
𝝉∈𝔼
(5) and using the strain decomposition (4), we arrive at
Σ

[( ) ] The classical formulation of associative plasticity (flow rule,


∶ E + 𝚺 ∶ E dΩ ≥ 0, (7)
𝜕Ψ(Ee )
�Ω
D mech
= 𝚺− ̇ e ̇ p loading/unloading conditions) directly follows from this
𝜕Ee principle. To this end, we first transform the maximization
principle into a minimization problem by changing the sign
where (◻) ̇ denotes the material time derivative of a quantity.
of the objective function. Next, we transform the constraint
The rate of elastic and plastic part of the macroscale strain
minimization problem into an unconstrained problem by
tensor lie in the space of second-order symmetric tensors
e p e p introducing the cone of Lagrange multipliers
𝕊 , that is Ė , Ė ∈ 𝕊 . Any permissible value of Ė , Ė ∈ 𝕊
𝕂 ∶= 𝛿 ∈ L2 (Ω) | 𝛿 ≥ 0 ,
{ }
will lead to a total strain rate, thanks to the additive decom-
(12)
position Ė = Ė e + Ė p ∈ 𝕊 , which can describe a plausible
kinematic process. The principle of thermodynamic deter-
where L2 denotes the space of all square integrable
minism requires that (7) remains valid for any kinematic
functions. The corresponding Lagrangian functional
process, which implies
Lp ∶ 𝕊 × 𝕂 × 𝕊 → ℝ is then
and 𝚺 ∶ Ė p ≥ 0.
𝜕Ψ(Ee )
𝚺= (8) p p
Lp (𝝉, 𝛿; Ė ) ∶= −𝝉 ∶ Ė + 𝛿 𝔉(𝝉). (13)
𝜕Ee
The first equation is a typical elastic constitutive relation, The solution to (11) is then given by a point (𝚺, 𝛾) ∈ 𝕊 × 𝕂
where the stress is defined as the derivative of the free satisfying the Karush-Kuhn-Tucker optimality conditions for
the Lagrangian functional (13). The conditions are

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 5 of 31  195

∑ ∑
) || 𝜕𝔉(𝝉) ||
p
𝜕 Lp (𝝉, 𝛾; Ė p E= 𝜙r 𝜺r and 𝚺= 𝜙r 𝝈 r . (16)
| = −Ė + 𝛾 | = 0,
𝜕𝝉 | 𝜕𝝉 ||𝚺 (14)
r r
|𝚺,𝛾
𝛾 ≥ 0, 𝔉(𝚺) ≤ 0, and 𝛾 𝔉(𝚺) = 0. A link between the macroscale strain E and the average
microscopic strain 𝜺r of phase r is established with a fourth
The first equation in (14) is the associated flow rule, often order concentration tensor 𝔸r as
also called the normality of the flow rule. The second and
third equations in (14) are the classical loading/unloading 𝜺r = 𝔸r ∶ E. (17)
conditions. The only requirement for these equations to hold A comparison of the macroscale constitutive relation
uniquely is the convexity of the elastic range 𝔼Σ . A suffi- 𝚺 = ℂ ∶ E with (16) and (17) yields the homogenized stiff-
cient condition for this requirement is the convexity of the ness ℂ in terms of the volume fraction, stiffness, and con-
yield criterion function 𝔉(𝝉) . We will exploit these aspects centration tensor of constituent phases as
later on for devising solution strategies for our optimization

framework. ℂ= 𝜙rr ∶ 𝔸r . (18)
r

It is clear from (18) that the estimation of the concentration


2.2 Multiscaling concepts in continuum
tensors 𝔸 entails the homogenized stiffness ℂ . The simplest
micromechanics
choice for 𝔸 is to assume a uniform strain state through-
out the RVE, that is 𝔸 = 𝕀 , where 𝕀 is a fourth-order sym-
Continuum micromechanics forms a rigorous foundation for
metric unit tensor. This choice leads to the well-established
the (semi-)analytical estimation of homogenized stiffness and
Voigt mixture rule for homogenized stiffness. However, the
strength properties of materials with hierarchical microstruc-
Voigt rule does not consider any other statistical informa-
tures. Here, we state the key results that are relevant in the
tion beyond the volume fraction of phases. We note that the
context of this article. For a detailed review, interested readers
Voigt rule is often applied for “homogenization/interpola-
are referred to the presentations given in Zaoui (2002), Suquet
tion” between a solid material and voids in conjunction with
(1997).
relaxation to ill-defined 0-1 type problems in topology opti-
mization (Bendsøe and Sigmund 1999; Allaire and Aubry
1999).
2.2.1 Estimation of homogenized elastic properties
The estimation of the concentration tensor 𝔸r based on
Eshelby’s matrix-inclusion solutions can incorporate the
The goal of continuum micromechanics is to estimate the
volume fraction, the shape of phases, and their interac-
homogenized response of a representative volume element
tion with each other. Eshelby’s matrix-inclusion problem
(RVE) filled with microheterogeneous material. For the exist-
relates strains in an ellipsoidal inclusion perfectly bonded
ence of such an RVE, a minimal requirement is that the charac-
with the surrounded homogeneous infinite elastic matrix
teristic length, d, of the considered inhomogeneities and defor-
to the applied homogeneous strains at infinity. Following
mation mechanisms is much smaller than the size, l, of the
Zaoui (2002), the estimation of 𝔸r from the matrix-inclu-
RVE. Moreover, l must be much smaller than the characteristic
sion problem entails the homogenized stiffness expression
length scale of the variation in the loading on the macroscale
as
structure, L. Therefore, a proper scale separation implies
∑ [∑
d ≪ l ≪ L. (15) ℂ= 𝜙r ∶ [𝕀 + ℙ0r ∶(𝕔r − ℂ0 )]−1 ∶ 𝜙s [𝕀+
(19)
r s
In each phase r of the RVE, the average microscopic stress ]−1
𝝈 r , the average microscopic strain 𝜺r , and the phase stiffness ℙ0s ∶ (𝕔s − ℂ )] 0 −1
.
𝕔r are linked as: 𝝈 r = 𝕔r ∶ 𝜺r  . The kinematic compatibil-
ity for the homogeneous strain boundary conditions for the Here, the Hill tensor ℙ0r characterizes the morphology of the
RVE relates the macroscale strain tensor E with the volume inclusion phase r and its interaction with the surrounding
average of microscopic strains 𝜺r . Similarly, the equilibrated reference matrix with stiffness tensor ℂ0 . The Hill tensor ℙ0r
microscopic stresses 𝝈 r and the macroscale stress tensor 𝚺 depends on the morphology, that is, the shape and orienta-
fulfill the volume average relation following the homogene- tion of the inclusion phase as well as the stiffness tensor of
ous stress boundary conditions. With 𝜙r as the volume frac-
tion of the phase r, these relations are

13
195   Page 6 of 31 T. Gangwar, D. Schillinger


the reference matrix. The analytical expressions for ℙ0r can √
≤ 0.
𝜕ℂ 𝜙̄ w 𝜎wY
be found in (Laws 1977, 1985; Masson 2008). With (19), 𝔉(𝝉) = 𝝉 ∶ [ℂ]−1 ∶ ∶ [ℂ]−1 ∶ 𝝉 −
the homogenized stiffness of the RVE can be expressed as a 𝜕𝜇w 3 𝜇w
(22)
function of constituent phase characteristics.
We √ emphasize t hat (22) is of t he for m of
2.2.2 Estimation of homogenized elastic limit strength 𝔉 = 𝝉 ∶ 𝕄 ∶ 𝝉 − R ≤ 0 that represents the general quad-
ratic form of classical rate-independent plasticity models. In
A macroscale RVE reaches the elastic limit state when any the context of this article, it implies that the elastic domain
one of the constituents in the RVE yields. Let us focus on defined by (22) satisfies two critical geometric properties.
the weakest constituent phase, denoted by index r = w . We These properties are (1) the convexity of the elastic domain
assume that its elastic limit behavior is described by the yield and (2) the degree-one homogeneity of the yield criterion.
criterion These properties are very important for developing solution

𝖋w (𝝈 ∗w ) ≤ 0,
algorithms for our optimization formulation, and we will
(20) recall them later in subsequent sections.
where 𝝈 ∗w is the effective stress measure in the weak phase
w. Moreover, we assume that it is the only constituent that
exhibits inelastic behavior. The effective stress or “stress
3 A framework for path‑dependent
peaks” in phase w can then be estimated with the second-
concurrent material and structure
order moment of the stress field in this phase, which is the
optimization
quadratic stress average over the phase volume Vw Suquet
In this section, we formulate a concurrent material and struc-
(1997), expressed as
ture optimization method maximizing the path-dependent
� �1∕2 stiffness for elastoplastic multiphase hierarchical systems.
Vw ∫Vw 2
1 1
𝝈 ∗w = ⟨𝝈 ∶ 𝝈⟩1∕2
w =
𝝈 ∶ 𝝈 dV . (21) We then focus on the finite element discretization of this
formulation.
Next we assume that 𝖋w is a scalar deviatoric stress-based
yield criterion such as the von Mises criterion with known 3.1 Thermodynamically consistent formulation
yield strength 𝜎wY  , bulk modulus 𝜇w and effective volume
function 𝜙̄ w for the weak phase w. Following (Suquet 1997), We start by looking at a representative problem illustrated in
with the admissible stress 𝝉 and homogenized stiffness ℂ , Fig. 1. We assume a fixed reference domain Ω subjected to a
the weak phase criterion 𝖋w translates to the macroscopic traction ̄t(t) at the Neumann boundary ΓN and the prescribed
yield criterion 𝔉 as displacements ū E (t) at the Dirichlet boundary ΓD with a body
force b(x, t) , where t ∈ [0, T] .

Fig. 1  Sketch of a representa-
tive elastoplastic multiphase
hierarchical system

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 7 of 31  195

3.1.1 Global optimization problem with micromechanical Aad and Ead define the set of admissible design variables
design variables at the macro- and microscales, respectively, with possible
design constraints. The second part of this equation is the
We introduce the definition of macroscale density 𝜌(x) and continuous version of the objective function proposed in
microstructure characterization m(x, t) . We assume that the Fritzen et al. (2016). On the one hand, the velocity field
macroscale density 𝜌(x) is fixed with respect to time, while v(x, t) and, thus, the displacement field u(x, ̄ t) implicitly
m(x, t) is a function of loading history representing a local depend on 𝜌(x) and m(x, t) . On the other hand, the first part
adaption of microstructure with time. The set m(x, t) contains of (25) is an explicit expression in terms of the design varia-
the geometric and mechanical characterization of phases that bles 𝜌(x) and m(x, t) with known macroscale strains E and Ep
span multiple well-separated microscales, consisting of vol- that arise from the solution of the global equilibrium equa-
ume fraction, material properties, shape, and orientation of the tions. The first part of (25) is the basis for the development
different phases in the hierarchical system. The homogenized and mathematical analysis of our optimization formulation.
material constitutive relations defined by the plastic dissipation Later on in this article, we will come back to the second part
Dp and the Helmholtz free energy Ψ in Sect. 2.1 depend on the of (25) to motivate sensitivity calculations.
macroscale density 𝜌(x) and the microstructure characterization
field m(x, t) . Therefore, the design vector is [𝜌(x), m(x, t)]T .
A typical objective is to maximize the structural stiff- 3.1.2 Definition of the sets of admissible design variables
ness for the path-dependent nonlinear structure designs. It
translates as the maximization of the total mechanical work The admissible set Aad seeks a limit on the total material
expended in the course of a deformation process (Fritzen mass Mreq available for design. Mathematically, it can be
et al. 2016). Assuming a quasi-static case with no inertial defined as
effects, the total mechanical work fw in the considered time {
interval [0, T] follows directly from the mechanical work Aad = 𝜌(x) | 𝜌(x) = [𝜌min , 𝜌max ],
identity (2) and the definition of stress power (6) as (26)
𝜌(x)dΩ ≤ Mreq , x ∈ Ω ,
}
T [ ] T
�Ω
∫0 ∫0 ext
d
fw = Dmech + Vint dt = P (v)dt. (23)
dt
where 𝜌min and 𝜌max are the bounds on the macroscale mate-
Utilizing the definitions of D  , Vint and Pext (v) from
mech rial density 𝜌.
Sect. 2.1, we arrive at Without loss of generality, the definition of the admis-
T[ { } ] sible set Ead is illustrated via the representative multiphase
∫0 ∫Ω
fw =
p ̇ e ) dΩ dt
Dp [𝚺; Ė ] + Ψ(E hierarchical system shown in Fig. 1. We observe a well-
(24) separated three-scale hierarchical system with three base
T[ ]
constituent materials denoted as Material A, B, and C
∫0 ∫ Ω ∫ ΓN
= b(x, t) ⋅ v dΩ + ̄
t(t) ⋅ v ds dt.
with densities 𝜌A , 𝜌B , and 𝜌C  , respectively. Material A and
B are linear elastic, and Material C exhibits a perfectly
We note that the (pseudo-)time t represents the loading elastoplastic response. At a material point P, the volume
history. fraction of Material B and C at the lowermost scale are 𝛾B
Augmenting Dp and Ψ with 𝜌(x) and m(x, t) , we set up our and 𝛾C such that 𝛾B + 𝛾C = 1 . Material B forms spherical
optimization problem using the introduced definition of total inclusions in the matrix of Material C at this scale. The
mechanical work fw from (24) as homogenized material from this scale forms the matrix M
T { that hosts the inclusions of Material A with the orienta-
∫0 ∫Ω
p
max fw = max Dp [𝜌(x), m(x, t), 𝚺;Ė ] tion 𝜃A and elongation 𝜁A at the mesoscale. The density of
𝜌(x) ∈ Aad 𝜌(x) ∈ Aad the matrix M is simply 𝜌M = (𝛾B 𝜌B + 𝛾C 𝜌C ) . The volume
m(x, t) ∈ Ead m(x, t) ∈ Ead fraction of Material A and matrix M are 𝜙A and 𝜙M with
}
̇
+ Ψ[𝜌(x), m(x, t);Ee ] dΩ dt 𝜙A + 𝜙M = 1 . The microstructure characterization field set
m(x, t) is thus {𝜙A (x, t), 𝜃A (x, t), 𝜁A (x, t), 𝛾C (x, t)} . We note
T [
that this set can be arbitrarily extended if necessary.
∫ ∫Ω
= max b(x, t) ⋅ v dΩ
𝜌(x) ∈ Aad 0 With these definitions, the microscale design admissi-
m(x, t) ∈ Ead ble set Ead for the representative multiphase hierarchical
] system shown in Fig. 1 follows as
∫ΓN
+ ̄t(t) ⋅ v ds dt.

(25)

13
195   Page 8 of 31 T. Gangwar, D. Schillinger

{ T {

𝜌(x) ∈ Aad ∫0 ∫Ω
Ead = m(x, t) | 𝜌(x) = 𝜌A 𝜙A (x, t) + 𝜌M (x, t)(1− max
p
̄ t), 𝚺; Ė ]+
Dp [𝜌(x), m(x,
𝜙A (x, t)),
≤ 1,
}
0 < 𝜙min max ̇
Ψ[𝜌(x), ̄ t); Ee ] dΩ dt
m(x,
A < 𝜙A (x, t) < 𝜙A
𝜌M (x, t) = 𝜌B (1 − 𝜸 C (x, t)) + 𝜌C 𝛾C (x, t), (27) T [ ]
0 < 𝛾Cmin < 𝛾C (x, t) < 𝛾Cmax ≤ 1,
𝜌(x) ∈ Aad ∫0 ∫Ω ∫ ΓN
= max b(x, t) ⋅ v dΩ + ̄t(t) ⋅ v ds dt.

𝜃A (x, t) ∈ [−𝜋∕2, 𝜋∕2], (30)


}
𝜻 A (x, t) ∈ [1, 𝜁 max
], ∀(x, t) ∈ Ω × [0, T] . We also explicitly write the equivalent statement in terms of
the total external work done in the deformation process from
Here, the volume fraction of Material A is bounded by 𝜙min (25), which we will exploit in the subsequent sections for
A
and 𝜙max  
, and the volume fraction of Material C is bounded discretization and sensitivity calculations. In this statement,
A
by 𝛾Cmin and 𝛾Cmax at their respected scales. Furthermore, the ̄ t) optimizes the following sub-problem or “material”
m(x,
elongation ratio of the inclusions of Material A is bounded optimization problem:
by 𝜁 max . We note that the bounds are constant and do not {
p
}
depend on the loading history or the material position in the max ̇
Dp [m(x, t), 𝚺;Ė ] + Ψ[m(x, t);Ee ]
m(x,t)∈Ead (𝜌(x))
domain. We emphasize again that the multiscale configu- { }
p ̇
max{𝝉 ∶ Ė } + Ψ[m(x, t);Ee ] (31)
ration of Fig. 1 is used for illustration, but the underlying = max
m(x,t)∈Ead (𝜌(x)) 𝝉∈𝔼Σ
representation is easily generalized to cover any other mul-
∀(x, t) ∈ Ω × [0, T].
tiphase hierarchical system.
The macroscale density 𝜌(x) dictates the construction of the
3.1.3 Decomposition into material and structure admissible space Ead , and, therefore, we take it out from the
optimization problems definitions of Dp and Ψ in (31) and consider it in Ead . In the
second line, we rewrite the definition of the plastic dissipa-
The definition of the admissible set Ead at a material point x tion Dp with the help of the principle of maximum plastic
only depends on the macroscale density 𝝆(x) of this point. It dissipation discussed in Sect. 2.1.3.
implies that the definition of Ead is pointwise, and we can thus The constitutive relations in (31) defined through the
rewrite the statement (25) as Helmholtz free energy Ψ and the plastic dissipation Dp
T { remain to be discussed. For linearized elasticity, the stored
m(x,t)∈Ead (𝜌(x)) ∫0 ∫Ω
elastic energy function W takes a quadratic form in the elas-
p
max max Dp [𝜌(x), m(x, t), 𝚺; Ė ]
𝜌(x) ∈ Aad tic part of the strain tensor Ee = E − Ep . The homogenized
} elasticity tensor ℂ is a function of the microstructure char-
̇
+ Ψ[𝜌(x), m(x, t);Ee ] dΩ dt.
acterization field set m(x, t) ∈ Ead (𝜌(x)) . For the perfect
(28) plasticity case, that is Ψ = W  , the quadratic form follows as
We also assume that the macroscale density 𝝆(x) is fixed
1
with respect to the loading history. Therefore, we are Ψ[m(x, t); Ee ] = (E − Ep ) ∶ ℂ(m(x, t)) ∶ (E − Ep ), (32)
2
allowed to swap the integral and maximization operations
in (28), hence The elastic constitutive equation from (8) entails the follow-
T { ing stress–strain relationship:

∫ ∫
p
max max Dp [𝜌(x), m(x, t), 𝚺;Ė ] 𝜕Ψ(Ee )
𝜌(x) ∈ Aad 0 Ω m(x,t)∈Ead (𝜌(x))
𝚺= = ℂ(m(x, t)) ∶ (E − Ep ). (33)
} 𝜕Ee
̇
+ Ψ[𝜌(x), m(x, t); Ee ] dΩ dt.
Similarly, the elastoplastic material constitutive equations
(29) through the maximum plastic dissipation principle stated in
The statement (29) allows us to decompose the optimiza- (11) is augmented to include 𝜌(x) and m(x, t) as
tion problem into two sub-problems. The outer “structure” { }
p p
optimization problem is Dp [m(x, t), 𝚺; Ė ] = max 𝝉 ∶ Ė , (34)
𝝉∈𝔼 Σ

where the definition of the elastic closure 𝔼Σ in terms of 𝜌(x)


and m(x,
̄ t) is

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 9 of 31  195

𝔼Σ ∶= 𝝉 ∈ 𝕊 | m(x, t) ∈ Ead (𝜌(x)), 𝔉(𝝉, m(x, t)) ≤ 0 .


{ }
consistent with the standard finite element discretization of
the initial boundary value problem introduced in Simo and
(35) Hughes (2006), to represent the introduced quantities in the
The homogenized stiffness ℂ(m(x, t)) and the homogenized global equilibrium equations.
yield criterion 𝔉(𝝉, m(x, t)) can be estimated as a function
of microstructure variables for instance via the continuum 3.2.1 Model definitions in the discrete setting
micromechanics principles outlined in Sect. 2.2.
We start by dividing the time interval [0, T] into nload parti-
Remark 2  Please note that, in this article, we assume perfect tions and split the domain Ω into Ne finite elements:
plasticity for the definitions of the Helmholtz free energy Ψ
nload −1 Ne
and the plastic dissipation Dp appearing in the decomposed ⋃ ⋃
[0, T] = [tn , tn+1 ] and Ω = Ωj . (36)
optimization formulation stated in (30) and (31). For mod- n=0 j=1
eling plasticity with hardening within the decomposed for-
mulation, the Helmholtz free energy Ψ in (32) and the plastic Here, Ωj is the domain of element j, and each element is
dissipation Dp in (35) can be augmented to include harden- equipped with Ngp Gauss quadrature points. We focus on
ing contributions by utilizing continuum micromechanics- a typical (quasi-)time interval [tn , tn+1 ] with known equili-
based homogenization schemes, such as outlined in Fritsch brated state at time step tn . In the context of this work, we
et al. (2009); Morin et al. (2017) for the example of bones. choose standard nodal finite elements with Lagrange basis
Throughout this article, however, we keep the assumption functions that can be assembled to approximate the mac-
of perfectly associated plasticity to focus on the foundations roscale displacements and strains at load increment (n + 1)
of the decomposed formulation and the general ideas for its over the complete domain:
computational treatment.
u ≈ Nun+1 and E ≈ Bun+1 (37)
3.1.4 Interpretation as an inelastic constitutive law where N is the (assembled) displacement interpolation oper-
ator and B is the (assembled) strain–displacement opera-
The combination of (30) and (31) constitutes the concur- tor (Hughes 2000). In the sense of the standard Galerkin
rent material and structure optimization formulation. The method, we use the same finite element basis functions for
maximization problem in (30) seeks the optimal material the representation of the solution and the test functions
distribution 𝜌(x) in the domain Ω . For a given material distri- (Hughes 2000).
bution 𝜌(x) , the optimization problem (31) finds the optimal
microstructure configuration maximizing the stress/deforma- Remark 3  We emphasize that from here on, displacement-
tion power for the known macroscale strains at each material type vector quantities with subscript n or (n + 1) such as
point x . Both statements are coupled through the macroscale un+1 denote the vector of unknown displacement-type coef-
strains and, therefore, through the displacement field solu- ficients at the corresponding load increment. Tensor quanti-
tion u(x,
̄ t) that satisfies the global equilibrium equations. ties with subscript n or (n + 1) such as macroscale strains
This interdependency makes the global equilibrium a con- En+1 or stresses 𝚺n+1 denote the corresponding tensor fields
stitutively nonlinear problem analogous to a typical initial approximated in terms of the corresponding displacement
boundary value problem with an inelastic constitutive law. finite element solution at the corresponding load increment.
Therefore, we propose to interpret the material optimization
problem as a reformulated elastoplastic constitutive law that The design vector [𝜌(x), m(x, t)]T for our example mul-
provides the locally optimal material response with respect tiscale configuration in Fig. 1 can now be defined in this
to the external loading history. The microstructure variable discrete setting as [𝝆, m]T  , where
m(x, t) can be thought of as an “internal state variable” anal-
ogous to any path-dependent history variable encountered 𝝆 = [𝜌1 , 𝜌2 , 𝜌3 , ..., 𝜌Ne ],
in elastoplasticity formulations. This interpretation will be m = [m0 , m1 , ..., mn+1 , ..., mnload −1 ],
used in Sect. 4 for devising the optimization algorithm for N ,1 x,j
the material optimization problem. mn+1 = [(m1,1
n+1
gp
, .., mn+1 ), ..., (.., mn+1 , ..), ...,
N ,N
(38)
1,N gp e
(mn+1e , .., mn+1 )],
3.2 Finite element discretization x,j x,j x,j x,j x,j
mn+1 = [𝜙A,n+1 , 𝜃A,n+1 , 𝜁A,n+1 , 𝛾C,n+1 ],
In the next step, we discretize our concurrent material and where x ∈ {1, .., Ngp }, j ∈ {1, .., Ne },
structure optimization formulation within the context of n ∈ {0, .., nload − 1}.
the finite element method. We use vector–matrix notation,

13
195   Page 10 of 31 T. Gangwar, D. Schillinger

The macroscale density 𝜌j is assumed to be constant in each


element and load increment, with j being the element index.
The microstructure design variable set m is defined at each
(macroscale) Gauss point and load increment with mn+1 as
the microstructure characterization set at load increment
(n + 1) . The microstructure configuration mn+1 at a Gauss
x,j

point x inside element j at load increment (n + 1) consists of


volume fraction 𝜙A,n+1 , orientation 𝜃A,n+1 elongation 𝜁A,n+1
x,j x,j x,j

for Material A, and volume fraction 𝛾C,n+1 of Material C. We


x,j

again emphasize that we use this definition of the multiscale


configuration in Fig. 1 for illustration purposes, this proce-
dure can easily be generalized to cover any other multiphase
hierarchical system.
The constitutive equation relating the macroscale stress
𝚺n+1 with the macroscale strains En+1 and En+1 at the Gauss
p
Fig. 2  Total mechanical work fw in the course of the deformation pro-
point x follows from (33) as cess
x,j p
𝚺n+1 = ℂ(mn+1 ) ∶ (En+1 − En+1 ), (39)
nload −1
1 ∑ ext
where the homogenized stiffness ℂ(mn+1 ) is evaluated for the
x,j max ∶ fw (𝝆) = (f + f ext T
n ) Δun+1
̄
𝝆 2 n=0 n+1
microstructure configuration mn+1 . To derive the incremental
x,j

form of the elastoplastic constitutive equations, the discrete s.t. ∶ r̄n+1 (𝝆, ū n+1 , m
̄ n+1 ) = 0 ∀n = 0, 1, ..., nload − 1
(42)
version of the maximum plastic dissipation principle at the Ne

Gauss point x from (35) is M(𝝆) = 𝜌j |Ωj | = Mreq = Mfrac × 𝜌C × |Ω|;
j=1
{ }
x,j p p
Dp [mn+1 , 𝚺n+1 ;En+1 ] = max 𝝉 ∶ (En+1 − Epn ) , (40) 𝜌j ∈ [𝜌min , 𝜌max ], ∀j = 1, 2, ..., Ne .
𝝉∈𝔼 Σn+1

In accordance with the notation introduced above in the con-


where the admissible stresses 𝝉 lie in a set 𝔼Σn+1 defined by
text of a finite element discretization, f ext is the external
the homogenized yield criterion 𝔉(𝝉, mn+1 ) evaluated at
x,j n+1
force vector, ū n+1 is the converged vector of the macroscale
x,j
mn+1 ∶ nodal displacements, and Δū n+1 ∶= ū n+1 − ū n is the incre-
ment of the displacement vector in load increment (n + 1) .
𝔼Σn+1 ∶= 𝝉 ∈ 𝕊 | mn+1 ∈ Ead (𝜌j ), 𝔉(𝝉, mn+1 ) ≤ 0 . (41)
{ }
x,j x,j
The force residual r̄n+1 is calculated utilizing the optimal
microstructure configuration m ̄ n+1 in load increment (n + 1) .
The macroscale plastic strain Epn is known from the equili- We note that the optimal microstructure configuration m ̄ n+1
brated solution state at load step n. We note that we write and ū n+1 are dependent on each other justifying the choice of
these constitutive relations in tensor notation given its direct ̄ notations introduced in Sect. 2.1. M(𝝆) is the total mass
(◻)
relation with continuum micromechanics principles stated of the occupying domain, and 𝜌j and |Ωj | are the density and
in Sect. 2.2. volume of element j.
The total mechanical work fw is the discrete version of the
3.2.2 Discrete form of the structure optimization problem second statement in (30), which is equivalent to the objec-
tive function proposed in Fritzen et al. (2016). This essen-
With the introduced definitions, we can write the discrete form tially is the area under the characteristic force-displacement
of the material and structure optimization formulation (30) curve approximated with the trapezoidal rule, as illustrated
and (31). In this work, we employ the trapezoidal rule for the in Fig. 2. The first condition in (42) ensures that the global
numerical evaluation of the integrals over the (quasi-) time equilibrium is satisfied in all load steps. The second and
domain, which is second-order accurate with respect to the third conditions of (42) are the discrete definitions of the
(quasi-)time step size. The discrete version of the structure macroscale admissible design variable set Aad  . The total
optimization problem (30) then becomes

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 11 of 31  195

available mass Mreq can be expressed in terms of the frac- We emphasize that all quantities at load increment n are
tion Mfrac with respect to the mass when the densest material known, and, therefore, Ψ(En − Epn ) does not play any role
occupies the complete domain. in this maximization problem. The optimized configura-
The force residual r̄n+1 at load increment (n + 1) is defined tion m ̄ n+1 is sought in the microscale design variable space
x,j

as mn+1 = [𝜙A,n+1 , 𝜃A,n+1 , 𝜁A,n+1 , 𝛾C,n+1 ] with constraints defini-


x,j x,j x,j x,j x,j

tions that follow from the admissible set Ead (27). From (44),
̄ n+1 ) ∶= f ext
r̄n+1 (𝝆, ū n+1 , m − f int
n+1 n+1 we can rewrite the material optimization statement as
Ne [ Ngp ]
∑ ∑ T (43) {
= f ext
n+1
− B 𝚺n+1 x ,
w m
x,j p
̄ n+1 = arg max 𝚺n+1 ∶ (En+1 − Epn ) + Ψ(En+1 −
j=1 x=1 x,j
mn+1 (𝜌j )
}
where wx contains the Gauss point weight and the deter- p
En+1 ) − Ψ(En − Epn )
minant of the Jacobian matrix for element j. We observe
x,j p
that the microstructure design variable set m is implicitly s.t. ∶ 𝚺n+1 = ℂ(mn+1 ) ∶ (En+1 − En+1 )
accounted for by the residual definitions in each load incre- 𝔉(𝚺n+1 , mn+1 ) ≤ 0
x,j

ment. The macroscale stress 𝚺n+1 at each Gauss point is


1p p
evaluated by solving the nonlinear elastoplastic constitutive Ψ(En+1 − En+1 ) = (E − En+1 ) ∶ (45)
2 n+1
relations (39) and (41) with known microstructure configu- x,j p
ℂ(mn+1 ) ∶ (En+1 − En+1 )
ration m̄ n+1 that solves the material optimization problem
x,j
x,j x,j
detailed in the following subsection. Therefore, the global 𝜌j = 𝜌A 𝜙A,n+1 + 𝜌M (1 − 𝜙A,n+1 );
equilibrium equation (43) is nonlinear and requires iterative x,j
𝜌M = 𝜌B (1 − 𝛾C,n+1 ) + 𝜌C 𝛾C,n+1
x,j
solution approaches such as the Newton–Raphson incremen-
x,j x,j
tal procedure (Simo and Hughes 2006; de Souza Neto et al. 𝜙A,n+1 ∈ [𝜙min
A
, 𝜙max
A
]; 𝜃A,n+1 ∈ [−𝜋∕2, 𝜋∕2];
2011). x,j x,j
𝜁A,n+1 ∈ [1, 𝜁 max ] 𝛾C,n+1 ∈ [𝛾Cmin , 𝛾Cmax ],
Structure optimization involving inelastic material mod-
els is potentially ill-posed in a force-controlled setting (Swan including all constraints defined through the stress admis-
and Kosaka 1997; Huang and Xie 2008; Schwarz et al. 2001; sible set 𝔼Σn+1 and microscale design admissible set Ead . The
Maute et al. 1998; Cho and Jung 2003). Therefore, we only first two conditions are essentially the elastoplastic constitu-
consider displacement-controlled loading in this article, incre- tive equations relating the macroscale stress with the mac-
mentally applied through prescribed displacements ū E (t) , see roscale strains via (39) and the constraint on the macroscale
Sect. 2.2.1. In a displacement-controlled setting, f extn+1
repre- stress defined through the homogenized yield criterion (41).
sents the discretized form of the loading potential resulting The third condition is the definition of the Helmholtz free
from the non-zero displacement boundary conditions. This energy in terms of microscale design variable mn+1 . The rest
x,j
assumption also simplifies the sensitivity calculations of the of the conditions follow in a straightforward manner from
objective function with respect to the design variables for the the constraints definitions in Ead  . The solution of (45) at
optimization algorithms, which we will discuss in Sect. 5.1. each Gauss point in each load increment yields the opti-
mized microstructure configuration set m ̄.
3.2.3 Discrete form of the material optimization problem We emphasize that in contrast to the equivalent strain
energy maximization for concurrent optimization problems
For a given material distribution 𝝆 and the macroscale dis- involving overall linear elastic multiphase hierarchical systems
placement solution vector ū n+1 , the material optimization prob- (Xia and Breitkopf 2014; Gangwar and Schillinger 2021), find-
lem for the Gauss point x inside element j for load increment ing the solution to the material optimization problem (45) is
(n + 1) follows from (31) as not straightforward. Both the macroscale plastic strain En+1
p

{
and the optimized microstructure m ̄ n+1 are unknown. Intui-
x,j
x,j p
m̄ n+1 = arg max max 𝝉 ∶ (En+1 − Epn )
x,j
mn+1 ∈Ead (𝜌j )
𝝉∈𝔼Σn+1 tively, the material optimization problem maximizes the area
} (44) under the homogenized elastoplastic stress–strain curve for
p
+ Ψ(En+1 − En+1 ) − Ψ(En − Epn ) . each material point. Multiple stress–strain curves are avail-
able at each load increment, defined by the the microscale
The first part of this equation directly comes from the incre- design variable mn+1 . This inter-dependency couples the his-
x,j

mental form of the principle of maximum plastic dissipation tory variable En+1 with mn+1 , necessitating a challenging novel
p x,j

outlined in (41). Similarly, the second part is the incremen- algorithmic treatment to tackle this maximization problem.
tal form of the Helmholtz free energy rate defined in (31).

13
195   Page 12 of 31 T. Gangwar, D. Schillinger

4 Algorithmic treatment of the material We define a Lagrangian functional that converts the


optimization problem constraint optimization problem (46a) into an uncon-
strained problem following (13):
For the algorithmic treatment, we interpret the material x,j p p
optimization problem as a reformulated constitutive law Ln+1 (𝝉, mn+1 , 𝛿; En+1 ) ∶= − 𝝉 ∶ (En+1 − Epn )+
(47)
at each material point that provides a locally optimal x,j
𝛿 𝔉(𝝉, mn+1 ),
mechanical response to the loading history. This interpre-
tation allows us to treat the microscale design variable where 𝛿 is in the cone of Lagrange multipliers defined
mn+1 as an additional internal state variable within the con-
x,j through (12). The solution to (46a) is given by a point
̂ n+1 , Δ𝛾n+1 ) that satisfies the Karush-Kuhn-Tucker
x,j
text of the classical formulation of plasticity. With this (𝚺n+1 , m
interpretation, we first exploit the principle of maximum optimality conditions for (47). The conditions entail
plastic dissipation to motivate a solution strategy for the
𝜕Ln+1
material optimization problem. We then cast this strategy p
= −(En+1 − Epn )+
into an algorithmic procedure that assumes the format of 𝜕𝝉
𝜕𝔉(𝝉, mn+1 ) ||
x,j
a typical return map algorithm for the integration of elas- |
Δ𝛾n+1 | = 0,
toplastic constitutive equations. Finally, we leverage con- 𝜕𝝉 |
tinuum micromechanics and the associated homogenized |𝚺n+1 ,m̂ n+1
x,j

elastoplastic constitutive relations, which enable further 𝜕Ln+1


x,j |
𝜕𝔉(𝝉, mn+1 ) | (48)
= Δ𝛾 | = 0,
simplifications that make our framework computationally x,j n+1 x,j |
𝜕 mn+1 𝜕mn+1 ||𝚺 ,m̂ x,j
feasible.
̂ n+1 ) ≤ 0, and
Δ𝛾n+1 ≥ 0, 𝔉(𝚺n+1 , m
n+1 n+1
x,j

x,j
4.1 The principle of maximum plastic dissipation Δ𝛾n+1 𝔉(𝚺n+1 , m
̂ n+1 ) = 0.
revisited The general structure of (48) is similar to the typical local
constitutive equations for plasticity (flow rule, loading/
As explained above, the first part of the material optimi-
unloading conditions) as described in (14). Equation (48)2
zation problem (44) is the incremental statement of the
represents the evolution of microstructure state in a particu-
maximum plastic dissipation principle. This part defines
lar load increment (n + 1) . All these equations together form
the interaction between the next stress state 𝚺n+1 and the
a coupled nonlinear system that requires a special compu-
optimized microstructure state m ̄ n+1 through the homog-
x,j
tational treatment. The solution m ̂ n+1 from (48) provides
x,j

enized yield criterion 𝔉(𝝉, mn+1 ) . Focusing on this part


x,j
important insights into the interaction of the plastic update
only, we can combine both statements in a single one as and the microstructure update. In the following, we will can
x,j
{
p
} utilize these insights to design an algorithmic framework
{𝚺n+1 , m
̂ n+1 } = arg max 𝝉 ∶ (En+1 − Epn ) , (46a) for solving the original material optimization problem (44).
x,j
(𝝉,mn+1 ) ∈ 𝔼Σn+1

where
4.2 Algorithmic procedure in the form of return
≤0 .
{ }
𝔼Σn+1 ∶= 𝝉 ∈ 𝕊,
x,j
mn+1 ∈ Ead (𝜌j ) |
x,j
𝔉(𝝉, mn+1 )
map algorithms
(46b) Analogous to the elastic–plastic operator split formulas
This maximization problem seeks the macroscale stress for inelastic constitutive equations, we define a trial elastic
state 𝚺n+1 and a solution m̂ n+1 within the modified admis-
x,j
state by freezing the plastic flow and microstructure evo-
sible space definition 𝔼Σn+1 . The solution m
̂ n+1 restricts the
x,j
lution state during the current load increment. It implies
search space for the solution m ̄ n+1 of the original material
x,j that the macroscale plastic strain and optimal microstruc-
optimization problem (45), which we will further detail in ture configuration state in the current load increment are
the subsequent discussion. We note that the interpretation of known and equal to that of the previous load increment, that
is En+1 = Epn , m ̄ n  . With these assumptions, the trial
p x,j x,j
the microscale design variable mn+1 as internal state variable ̂ n+1 = m
x,j

naturally arises from these statements. elastic state is

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 13 of 31  195

p
En+1 = Epn ⟹ Ee,tr ∶= En+1 − Epn ̄ n+1 } to the original material optimization problem
x,j
n+1 {Σn+1 , m
𝚺trn+1 ∶= ℂ(m
̄ x,j p
(49) (44) automatically satisfies (48). To see this, one can put
n ) ∶ (En+1 − En )
̄ n+1 , implying 𝔉(𝚺n+1 , m
̂ n+1 ) = 𝔉n+1 , then the dis-
x,j x,j x,j
m̂ n+1 = m
𝔉trn+1 ∶= 𝔉(𝚺trn+1 , m
̄ x,j ).
crete KKT condition Δ𝛾n+1 𝔉(𝚺n+1 , m
n x,j
̂ n+1 ) = Δ𝛾n+1 𝔉n+1 = 0
Here, En+1
e,tr
 , 𝚺trn+1 and 𝔉trn+1 denote the macroscale trial elastic in (48) results in Δ𝛾n+1 = 0 . This means that (48)1 implies
En+1 = Epn , and (48)2 is automatically satisfied with no
p
strain, trial elastic stress, and trial yield criterion, respec-
tively. Further, the convexity of the homogenized yield cri- restrictions on the solution space of the microstructure con-
figuration mn+1 . Therefore, with zero dissipation, the solu-
x,j
terion 𝔉 leads to the following important property Simo and
Hughes (2006): tion m ̄ n+1 of the material optimization problem reduces to the
x,j

𝔉trn+1 ≥ 𝔉n+1 , and 𝔉n+1 ∶= 𝔉(𝚺n+1 , m


x,j
strain energy maximization that follows from (45) as
̄ n+1 ), (50)
x,j 1 p
m
̄ n+1 = arg max (E − En+1 ) ∶
where 𝔉n+1 is the homogenized yield criterion calculated at 2 n+1
(51)
x,j
mn+1 ∈Ead (𝜌j )
the macroscale stress 𝚺n+1 and the optimal microstructure x,j p
ℂ(mn+1 ) ∶ (En+1 − En+1 ).
configuration m̄ n+1 after load increment (n + 1).
x,j

In conclusion, the solution m ̂ n+1 does not pose any restric-


x,j
4.2.1 Elastic trial state and plastic vs. microstructure
tions to the search space Ead for the optimized microstruc-
updates
ture configuration m ̄ n+1 in this case.
x,j

To support our discussion of the remaining two cases,


The trial state (49) and property (50) in combination with
Fig. 3 presents a geometric interpretation in a typical return-
(48) lead to three important cases defining the restrictions
mapping context. We note that in Fig. 3, the gray region rep-
posed by the solution m ̂ n+1 on the search space for the origi-
x,j
resents the family of available homogenized yield criterion
nal material optimization problem (45). In the following, we
envelops 𝔉(𝝉, mn+1 ) = 0 at each load increment, defined by
x,j
will discuss all three cases in detail.
the set of microscale design variables mn+1 .
x,j
Case 1 If 𝔉trn+1 < 0 , then property (50) entails 𝔉n+1 < 0 ,
Case 2 If 𝔉trn+1 > 0 , and if it is possible to find at least
implying a purely elastic step. In this case, the solution
one microstructure configuration m ̂ n+1 ∈ Ead (𝜌j ) such that
x,j

Fig. 3  Geometric illustration of solution strategy for the material optimization problem, based on the return-map algorithm interpretation

13
195   Page 14 of 31 T. Gangwar, D. Schillinger

̂ n+1 ) ≤ 0 , property (50) indicates that


x,j
𝔉tr(2) ∶= 𝔉(𝚺trn+1 , m 1
𝔉n+1 ≥ 𝔉n+1 ⟹ 𝔉n+1 < 0 . The solution is possible via
n+1 x,j p
m
̄ n+1 = arg max (E − En+1 ) ∶
tr(2)
2 n+1
(52)
x,j
𝔉(Een+1 ,mn+1 )≤𝔉tr(2)
n+1
the microstructure update provided that new microstructure
̄ n+1 ) ≤ 𝔉tr(2) ≤ 0.
x,j p
ℂ(mn+1 ) ∶ (En+1 − En+1 ).
state m ̄ n+1 follows the constraint 𝔉(𝚺trn+1 , m
x,j x,j
n+1
We call this case adaption to elastic state through micro- The constraint in this problem ensures that the state remains
structure evolution. We observe in Fig. 3a that due to the elastic and can be interpreted as a restriction on the search
current microstructure state m ̄ n (dashed line), the material
x,j
space for m̄ n+1 posed by the feasible solutions of (48). We
x,j
goes into plastic state; however, the material adapts itself to
write 𝔉 in the strain space to emphasize that the strain state
fall back to the elastic state by updating the microstructure
is known and problem (52) is a function of mn+1 only.
x,j
state to m ̄ n+1 (dotted line), while maximizing the total strain
x,j
C a s e 3 I f 𝔉trn+1 > 0   , and the problem
energy.
̂ n+1 ) = 0 does not have any solution.
x,j
Again, equation (48) leads to En+1 = Epn and Δ𝛾n+1 = 0
p 𝔉tr(2)
n+1
∶= 𝔉(𝚺trn+1 , m
following the discussion in Case 1. The solution m ̄ n+1 fol-
x,j It implies that no microstructure state can solve (48) with
the chosen trial elastic strain En+1
e,tr
 . Therefore, only a plas-
tic update is feasible, and En+1 ≠ Epn . This condition leads
lows as p

Fig. 4  Graphical solution of the material optimization problem in one dimension for the three possible cases

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 15 of 31  195

to Δ𝛾n+1 > 0 from (48)1 , and the microstructure evolution We cast these cases into an algorithmic frame analo-
condition from (48)2 entails 𝜕𝔉∕𝜕mn+1 = 0 . It means that
x,j
gous to a standard elastoplastic return map algorithm. We
the microstructure configuration remains unchanged, that is summarize the result in the following box.
̄ n  . It implies that the solution m ̂ n+1 of (48) restricts
x,j x,j x,j
m̄ n+1 = m
the search space for m ̄ n+1 to a single point, that is m̄ n  . With
x,j x,j
1. Given: E n+1 , E n , E p x,j
n , m̄n , ρj
̄ n  , the rest of the relations in (48) reduces to the
x,j x,j
m̄ n+1 = m tr
2. Compute elastic trial stress Σ n+1
typical elastoplastic constitutive equations with known stiff-
ness and yield criterion. Ep
n+1 = E n
p

This is illustrated in Fig.  3b, where we see that no n+1 := C(m̄n ) : (E n+1 − E n )
Σ tr x,j p

elastic update is possible for the current trial state; there-


n+1 := F(Σ n+1 , m̄n )
3. Check yield criterion Ftr tr x,j
fore, the microstructure state remains unchanged, and the
IF: Fn+1 ≤ 0
tr
next stress state 𝚺n+1 is solved with a standard return map CASE 1: Elastic update ∆γn+1 = 0; E pn+1 = E n
p
algorithm such as the closest point projection algorithm x,j
Microscale design m̄n+1 update through straightforward
(Simo and Taylor 1985). strain energy maximization

4.2.2 An analogue to the elastoplastic return map 1


arg max (E n+1 − E p x,j
n+1 ) : C(mn+1 ) :
algorithm mx,j
n+1 ∈Ead (ρj )
2
(E n+1 − E p
n+1 )
For an intuitive understanding, Fig. 4 presents a graphi-
cal solution of the material optimization problem for the n+1 > 0
ELSE IF: Ftr
tr(2)
example of a one-dimensional linear elastic-perfectly
x,j
CHECK IF: Fn+1 := n+1 , m̂n+1 )
F(Σ tr ≤
plastic model. The gray region in these graphs represents x,j
0; for m̂n+1 ∈ Ead (ρj )
the family of stress–strain curves for different microstruc- CASE 2: Evolve microscale design ensuring elastic
ture design configurations mn+1 . The material state (stress,
x,j material behavior
∆γn+1 = 0; E p x,j
n+1 = E n ; m̄n+1 is solution of
p
strain, microscale configuration) at load increment n is
known, and the macroscale strain En+1 at load increment 1
(E n+1 − E p x,j
arg max n+1 ) : C(mn+1 ) :
(n + 1) is given. Typical strain increments are infinitesi- F(E en+1 ,mx,j
n+1 )≤F
tr(2) 2
n+1
mal, and increments are large for illustration only. The
(E n+1 − E p
n+1 )
next material state from the material optimization prob-
lem warrants that the area increment (red shaded region ELSE:
CASE 3: Only plastic update possible =⇒ ∆γn+1 >
in the graphs) is maximized.
0; m̄x,j
n+1 = m̄n
x,j
The trial elastic stress state 𝚺trn+1 assumes the micro- p
Update E n+1 through closest point projection algorithm.
structure state and the macroscale plastic strain in this
4. Output: m̄x,j p
n+1 , E n+1 , and Σ n+1 = C(m̄n+1 ) :
x,j
load increment are equal to the previous load increment. p
(E n+1 − E n+1 ).
Case  1 results in a purely elastic update as shown in
Fig. 4a. Except for the first load increment, this leads to
a trivial solution for linear-elastic perfectly plastic mod-
els with the same microstructure configuration, that is
m
x,j
̄ n+1 = m ̄ n  . Case 2 in Fig. 4b is of particular interest.
x,j 4.3 A special choice: continuum micromechanics
The trial stress 𝚺trn+1 predicts a plastic update. However, it based homogenization
is possible to find material configurations m ̂ n+1 such that
x,j

̂ n+1 ) ≤ 0 . The material adapts itself by falling


x,j In this article, we focus on homogenized yield crite-
𝔉(𝚺trn+1 , m
rion based on the quadratic stress average that can be
back onto the elastic state via an appropriate update of
obtained within a continuum micromechanics framework
the microscale configuration m ̄ n+1 (denoted with the red
x,j
as reviewed in Sect. 2.2.2. To illustrate the simplifica-
dashed line), maximizing the total strain energy following
tions that can be achieved with this choice, we consider
(52). In Case 3, no material configuration allows an elas-
again the one-dimensional example from Fig. 4. Using the
tic state for the trial stress 𝚺trn+1 . Therefore, the material
homogenized yield criterion based on continuum micro-
configuration remains unchanged, and the stress–strain
mechanics, we arrive at Fig. 5 that graphically illustrates
state is updated through a return-mapping/closest point
the associated simplifications in the algorithmic procedure.
projection algorithm as shown in Fig. 4c.
In particular, the microscale configuration corresponding
to the maximum stiffness (maximum strain energy den-
sity) also results in the maximum strength properties for

13
195   Page 16 of 31 T. Gangwar, D. Schillinger

the homogenized response, which implies that adaption


to elastic state through microstructure evolution (Case 2)
cannot occur here. In the following, we will provide a
more detailed account of these simplifications.

4.3.1 Special properties induced through this choice

For illustration purposes, we fall back to our initial


example of a representative multiphase hierarchical sys-
tem defined in Fig. 1. We recall that Material C at the
microscale is a perfectly elastoplastic material that fol-
lows the von Mises failure criterion with yield strength
𝜎CY and bulk modulus 𝜇C  . For this example, we can then Fig. 5  Simplifications in the algorithmic procedure for material opti-
write the homogenized yield criterion 𝔉 given in (22) as mization induced by continuum micromechanics estimates. The opti-
a function of the stress 𝝉 and microscale design variable mized material configuration m ̄ n+1 remains unchanged with loading
x,j

mn+1 = [𝜙A,n+1 , 𝜃A,n+1 , 𝜁A,n+1 , 𝛾C,n+1 ]:


x,j x,j x,j x,j x,j history

≥ Ee,tr
x,j
𝔉(𝝉, mn+1 ) x,j x,j
√ Ee,tr ̄ n+1 ) ∶ Ee,tr
∶ ℂ(m ∶ ℂ(mn+1 ) ∶ Ee,tr
̄ n+1 ) ≤ 𝔉(𝚺trn+1 , mn+1 ).
n+1 n+1 n+1 n+1
x,j (54)
𝜕 ℂ(mn+1 ) ⟹ 𝔉(𝚺trn+1 , m
x,j x,j
𝝉:[ℂ(mn+1 )]−1 : :[ℂ(mn+1 )]−1 :𝝉
x,j x,j
=
𝜕 𝜇C (53)
√ It is straightfor ward to see from (54) that if
𝜙̄ C 𝜎CY
≤ 0, ̄ n+1 ) > 0 , then 𝔉(𝚺trn+1 , mn+1 ) ≤ 0 is not possible
x,j x,j
− 𝔉(𝚺trn+1 , m
3 𝜇C for any microstructure configuration mn+1 . Therefore, fol-
x,j

lowing our discussion in Sect. 4.2, we can conclude that an


where 𝜙̄ C is the equivalent volume fraction of Material C
adaption to the elastic state through microstructure evolution
computed as 𝜙̄ C = (1 − 𝜙A,n+1 ) 𝛾C,n+1 . For a detailed deri-
x,j x,j
(Case 2) is inconceivable for the continuum micromechanics
vation of the homogenized stiffness ℂ(mn+1 ) , we refer inter-
x,j
schemes outlined in this paper.
ested readers to Appendix 1 in Gangwar and Schillinger
(2021). These estimates hold the following three properties Property 2  An important conclusion from the previous
that form the basis for further simplifications in the algorith- section is that the microscale design update is possible in
mic procedure for the material optimization: an elastic step only. The elastic part of macroscale strain
Een+1 ∶= En+1 − En+1 at each Gauss point therefore entails
p

Property 1  The microscale configuration m ̄ n+1 correspond-


x,j
the optimal material orientation 𝜃̄A,n+1 for load increment
x,j
ing to the maximum stiffness (maximum strain energy den- (n + 1) . In the elastic step, the material optimization problem
sity) from (51) also maximizes the homogenized strength is essentially a strain energy maximization. The maximum
response. Figure  5 graphically represents this property strain energy is obtained for a general orthotropic material
for the one-dimensional case with a family of possible by aligning the material axis with the principal strain axes
stress–strain curves. Here, the microscale configuration for the elastic strains (Jog et al. 1994; Pedersen 1989).
corresponding to the higher linear-elastic slope leads to a
higher limit strength for the homogenized response. Thus, Property 3  If the external loading is monotonically increas-
the stress–strain curve for the configuration m ̄ n+1 acts as an
x,j
ing, the optimal material orientation 𝜃̄A,n+1 is the only micro-
x,j
envelope for all the stress–strain curves defined by the possi-
scale variable that may change in each load increment. We
ble microscale configurations mn+1 . Utilizing the definitions
x,j
denote the set of remaining microscale design variables as
∶= En+1 − Epn and 𝚺trn+1 ∶= ℂ(m
̄ n ) ∶ (En+1 − Epn ) , this
e,tr x,j
mn+1 = [𝜙A,n+1 , 𝜁A,n+1 , 𝛾C,n+1 ] . The optimal configuration
En+1 l(x,j) x,j x,j x,j

property can be summarized as


̄ n+1 for mn+1 remains unchanged throughout the loading
l(x,j) l(x,j)
m
history, that is m ∀n = 1, 2, ..., nload − 1 . We pro-
l(x,j) l(x,j)
̄ n+1 = m̄n
vide a proof of this property in A.

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 17 of 31  195

4.3.2 Simplification of the material optimization problem method for updating the design variables in each structure
optimization iteration, utilizing the computed sensitivities
The three special properties discussed above entail two (Sigmund 2001). Finally, we consolidate all developments
important simplifications. First, adaption to the elastic state into a single algorithmic framework.
through microstructure evolution (Case 2) cannot occur. Sec-
ond, except for the material orientation 𝜃A,n+1 , the optimized
x,j

material configuration remains unchanged throughout the


loading history. This implies that the material optimization 5.1 Sensitivity analysis
problem is solved for the first load increment only via the
strain energy maximization (51) for the optimized config- The format of the structure optimization problem (42) is
uration m̄ n+1 . Later, the optimized material orientation is
x,j
equivalent to the topology optimization formulation for
updated for each load increment by aligning the material elastoplastic structures presented by Fritzen et al. (2016),
axis with the principal strain axes of the elastic part of the Xia et al. (2017). They derive the sensitivities using the
macroscale strain tensor Een+1. path-dependent adjoint method (Buhl et al. 2000; Cho
We would like to emphasize the crucial role of the con- and Jung 2003). In the following, we provide a sketch of
tinuum micromechanics based estimates for the concurrent the derivation and highlight the important results. For a
material and structure optimization of multiphase hier- detailed derivation, we refer interested readers to Fritzen
archical systems. These estimates render both objective et al. (2016), Xia et al. (2017).
functions and constraint definitions of the material opti- The adjoint method begins with the construction of a
mization statement (45) and, therefore, the strain energy Lagrangian function fw∗ that satisfies the zero residual con-
maximization (51) as “discretization-free” analytical or straints r̄n+1 and r̄n at (quasi-)time tn+1 and tn for each term
semi-analytical expressions. This reduced problem is a of the trapezoidal rule stated in (42). With the Lagrange
straightforward constraint optimization problem that can multipliers 𝝀n+1 and 𝝁n+1 that are of the same dimensions
be solved with standard gradient-based methods. The solu- as the vector of unknowns ū n+1 , the Lagrangian function
tion to this material optimization problem is equivalent to fw∗ follows as
solving a set of (n + p) nonlinear equations with (n + p) nload −1{
variables, where n and p are the total number of micro- 1 ∑
fw∗ = (f ext ext T
n+1 + f n ) Δun+1
̄
scale design variables and the total number of equality 2 n=0
(55)
constraints, respectively. Thus, the continuum microme- }
T T
+ (𝝀n+1 ) r̄n+1 + (𝝁n+1 ) r̄n .
chanics enables us to handle computational challenge as
well as the complex constraint definitions in the optimi-
zation of multiphase hierarchical systems. For a detailed Since r̄n+1 and r̄n vanish at the equilibrium solution, the sen-
discussion on these aspects, please refer to our previous sitivity of fw∗ is same as that of fw , implying that
work (Gangwar and Schillinger 2021). 𝜕fw 𝜕f ∗
In conclusion, the total cost of solving all the material = w. (56)
𝜕𝜌j 𝜕𝜌j
optimization problems is equivalent to the case of an end-
compliance type optimization problem with a linear elas- The derivative of fw∗ with respect to the design variable 𝜌j
tic response at the material scales. These simplifications follows from (55) as
result in an enormous reduction in computational effort,
{
making our framework computationally tractable for the ( )
nload −1
𝜕fw∗ 1 ∑ 𝜕
elastoplastic case. = (f ext
n+1 + f ext T
n ) Δ u
̄ n+1 +
𝜕𝜌j 2 n=0 𝜕𝜌j
} (57)
T
𝜕̄rn+1 T 𝜕̄rn
(𝝀n+1 ) + (𝝁n+1 ) .
𝜕𝜌j 𝜕𝜌j
5 Comments on computer implementation
The derivative of r̄n+1 with respect to 𝜌j is evaluated follow-
ing the residual definition in (43). Substituting the definition
In this section, we provide an overview of our optimiza- of 𝚺n+1 given in (39) into (43), the derivative expression
tion framework with a focus on essential computer imple- becomes
mentation details. First, we derive the essential sensitivity
calculations of the objective function fw with respect to the
design variables 𝝆 for the structure optimization problem
(42). We then briefly touch upon the optimality criteria

13
195   Page 18 of 31 T. Gangwar, D. Schillinger

Result: Optimized solution vector [ρ̄, m̄]T second term of this equation are only computed for element
1 Initialize ρ0 , m0 ; j. For all remaining elements, 𝚺n+1 does not depend on 𝜌j .
2 Set design iteration counter i = 0; The second term is zero for all corresponding entries, main-
3 while ||ρi+1 − ρi ||/||ρi || > δtol do
taining dimensional consistency with the vector f ext .
4 Set up load increments ∆ūE n+1 for each loading step
n+1
index n = 0, .., nload − 1; We observe that the sensitivities of fw as expressed
5 Initialize load increment couter n = 0; in (57) and (58) require computationally extensive cal-
6 Initialize ū0 = 0 =⇒ E 0 = 0, and E p 0 = 0; culations of unknown derivatives. Therefore, our aim is
7 for n ≤ nload − 1 do to obtain the values of the Lagrange multipliers 𝝀n+1 and
Increment load ūEn+1 = ūn + ∆ūn+1 ;
E E
8
𝝁n+1 in such a way that these unknown derivatives can be
9 Set Newton iteration counter k = 0;
(0) eliminated from the sensitivity expression. To this end,
10 ūn+1 = ūn ;
(k)
we classify the degrees of freedom (DOF) into essential
11 while ||r̄ n+1 || < tol do (index E; associated with the Dirichlet boundary condi-
12 forall macroscale Gauss points do
13 Compute the macroscale strain tions) and free (index F; remaining). According to this
(k) (k) classification, we can partition vectors and matrices as
E n+1 = ∇s (ūn+1 );
(k) p,(k) shown for the following generic objects v and M :
14 Update the state variables Σ n+1 , E n+1 ,
[ E] [ EE EF ]
and m̄x,j
n+1 by solving the material v M M
optimization problem; v ∼ F and M ∼ . (59)
v MFE MFF
15 end
16 Evaluate the residual force vector
(k) (k) Since the displacements ū E on the Dirichlet boundary ΓD are
r̄ n+1 := f ext int
n+1 − f (Σ n+1 );
prescribed, they are independent of the current value of the
17 Set up the linear system:
tan,(k) (k) optimization variable 𝝆 . This observation leads to
K n+1 δ ū(k+1) = r̄ n+1 , and solve for
[ ] [ ]
δ ū(k+1) ; 𝜕Δū q E 0
𝜕 Δū q
18 Apply Newton correction to the displacements: = = 𝜕Δū q ,
F
(60)
(k+1) (k) 𝜕𝜌j 𝜕𝜌j Δū Fq
ūn+1 = ūn+1 + δ ū(k+1) ; 𝜕𝜌 j

19 k++;
20 end at an arbitrary load step index q = 0, ..., nload − 1 . With dis-
21 Calculate and store Lagrange multipliers λn+1 and placement-controlled loading, the only possible non-zero
µn+1 for senstivity calculations using converged entries in the global force vector f ext
q
are the reaction forces
state variables;
22 n++; f ext,E
p
 , that is
23 end [ ext,E ]
Compute the objective function fw (ρ) and sensitivities f
24
(61)
ext
fq = q .
∂fw /∂ρ; 0
25 Update density ρi+1 using the optimality criteria
algorithm; i++; The relations (60) and (61) lead to an educated choices for
26 end the vectors 𝝀n+1 and 𝝁n+1 such that the unknown derivatives
with respect to the design variables in (57) and (58) can
Algorithm  1 Concurrent structure and material optimization frame- be eliminated (see Fritzen et al. (2016); Xia et al. (2017)
work for elastoplastic structures with multiphase hierarchical materi- for details). The final expression for the sensitivity of the
als.
objective function fw with respect to the design variable 𝜌j is
nload −1 { Ngp [ x,j
𝜕fw∗ 1 ∑ ∑ 𝜕ℂ(m
̄ n+1 )
=− (𝝀n+1 )T BT (En+1
[Ngp x,j ] 𝜕𝜌j 2 n=0 x=1
𝜕𝜌j
𝜕̄rn+1 f ext
n+1
∑ T 𝜕ℂ(m̄ n+1 ) p ]
= − B (En+1 − En+1 )wx p
𝜕𝜌j 𝜕𝜌j x=1
𝜕𝜌j − En+1 )wx (62)
𝜕Δū n+1 [ ]}
− K tan
n+1 , (58) Ngp
∑ 𝜕ℂ(m
x,j
̄n )
𝜕𝜌j + (𝝁n+1 )T BT (En − Epn )wx .
𝜕𝜌j
𝜕̄rn+1 x=1
and K tan
n+1 = − ̄ .
𝜕 un+1 With the prescribed displacement increments Δū En+1 , the
choice of Lagrange multipliers 𝝀n+1 and 𝝁n+1 that lead to the
K tan is the global finite element stiffness matrix of the
n+1 above expression is
mechanical system at the equilibrium of load step (n + 1) .
We note that the quantities inside the square brackets in the

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 19 of 31  195

𝝀En+1 = −Δū En+1 and ( 𝛼ji )𝜂


tan,FF −1 tan,FE Bij = , (68)
𝝀Fn+1 = [K n+1 ] K n+1 Δū En+1 , Λi |Ωj |
(63)
𝝁En+1 = −Δū En+1 and
where Λi is the Lagrange multiplier corresponding to the
𝝁Fn+1 = [K tan,FF
n ]−1 K tan,FE
n Δū En+1 . total material mass constraint in design update i, and 𝜂 is a
damping parameter. The macroscale density is updated by
We note that the history of kinematic state variables E and means of the well-known optimality criteria method Sig-
Ep in (62) are known from the solution of the global equilib- mund (2001):
rium equations at each load increment. For the representa-
tive multiscale configuration in Fig. 1, the derivative of the ⎧ max(𝜌min , 𝜌i − 𝜇) if 𝜌i Bi ≤ max(𝜌min , 𝜌i − 𝜇)
= ⎨ min(𝜌ij + 𝜇, 𝜌max ) if min(𝜌ij + 𝜇, 𝜌max ) ≥ 𝜌ij Bij
j j j j
homogenized stiffness ℂ with respect to the element density 𝜌i+1

𝜌j in (62) can be evaluated by means of the chain rule. Fol- j
⎪ 𝜌i Bi otherwise
lowing Gangwar and Schillinger (2021), the expression is ⎩ j j
(69)
x,j x,j x,j
𝜕ℂ(m
̄ n+1 ) 𝜕ℂ 𝜕𝜙A,n+1 𝜕ℂ 𝜕𝛾C,n+1 To prevent a singular global stiffness matrix, the lower limit
= + , (64)
𝜕𝜌j x,j
𝜕𝜙A,n+1 𝜕𝜌j x,j
𝜕𝛾C,n+1 𝜕𝜌j 𝜌min on 𝜌j is limited by a small value, set in our case to 0.001.
The maximum possible element density 𝜌max depends on
where 𝜙A,n+1 and 𝛾C,n+1 relate to 𝜌j via (45). The partial deriv-
x,j x,j the density of the constituents at the microscales and the
atives of ℂ with respect to 𝜙A,n+1 and 𝛾C,n+1 are evaluated at
x,j x,j prescribed bounds in (45). 𝜇 is a small move parameter that
improves the stability, for instance by preventing multiple
the optimal microstructure configuration m
̄ n+1 by using finite
x,j
holes appearing and disappearing during optimization. The
difference approximations.
Lagrange multiplier Λi is updated using the bisection method
to satisfy the mass constraint. The design iterations stop
5.2 Structure optimization scheme when the density convergence criteria are met.

We utilize the algorithmic procedure for the structure opti-


5.3 General algorithm
mization that we outlined in our previous work Gangwar
and Schillinger (2021) and that we briefly summarize in the
Algorithm  1 consolidates all the developments into an
following. First, we define sensitivity numbers to rank the
algorithmic framework. It mainly consists of three blocks.
element sensitivities that are used to update the macroscale
The outer block represents macroscale structure optimiza-
design variables in each design iteration:
tion iterations with iteration index i, using the optimality
𝜕fc criteria method detailed in (69). It stops when the relative
𝛼j = −
𝜕𝜌j
. (65) change in macroscale density 𝝆 falls below the tolerance
𝛿tol , and the converged solution is the optimum macroscale
To avoid mesh dependency and checkerboard patterns, density 𝝆̄ . For a given macroscale density distribution, the
the sensitivity numbers are first smoothed with a filtering middle block solves the initial boundary value problem with
scheme defined as known load increments Δū En+1 at each load increment n. The
global equilibrium for each load increment is solved with the
∑ Nj
j� =1
gjj� 𝛼j �
Newton-Raphson method that uses the linearization of (43)
𝛼j = ∑N and gjj� = max {0, rmin − Δ(j, j )}, (66) (Simo and Hughes 2006). Here, (∙)(k) denotes the value of a
j n+1
j� =1
gjj� particular variable (∙) at the kth iteration at load step (n + 1) .
The Newton-Raphson scheme stops when the norm of the
where Nj is the set of neighboring elements for which center- residual force vector drops below a tolerance threshold 𝜖tol ,
to-center distance Δ(j, j ) to element j is smaller than the
� ′
and we adopt 𝜖tol = 10−5 in this article.
filter radius rmin . To improve convergence, the sensitivity The inner block solves the material optimization problem
numbers are further averaged with the sensitivity numbers described in Sect. 4 at each Gauss point with prescribed
of the previous design iteration as state variables at this iteration stage for each load incre-
ment. For the schemes based on continuum micromechan-
𝛼ji+1 → (𝛼ji+1 + 𝛼ji )∕2. (67) ics outlined in this paper, the material optimization problem
is solved for the optimized configuration m ̄ n+1 maximizing
x,j
The ratio of sensitivity numbers and the mass constraint are
combined to the strain energy via (51) at the first load increment only.
Thereafter, the material orientation is updated for each load

13
195   Page 20 of 31 T. Gangwar, D. Schillinger

increment by aligning the material axis with the principal points, resulting in 80 × 40 × 4 = 12, 800 material optimiza-
strain axes of the elastic part of the macroscale strain tensor tion problems in each load step.
according to our discussion in Sect. 4.3. The equations due As illustrated in Fig. 6, we consider a hierarchical sys-
to strain energy maximization can be solved with standard tem that consists of Material A, B, and C at two differ-
gradient-based constraint optimization methods such as the ent length scales. Their densities (in Kg/m3 ) are 𝜌A = 0 ,
quasi-Newton method of Broyden, Fletcher, Goldfarb, and 𝜌B = 0.5 , and 𝜌C = 1.0 , their Young’s moduli (in GPa) are
Shanno (BFGS) or sequential least squares programming EA = 0.0 , EB = 0.5 , and EC = 1.0 , and Poisson’s ratio of
(SLSQP) methods. all constituents is 0.3. Material C is elastoplastic with yield
strength 1 MPa. We assume that Material A forms cylin-
drical inclusions in the homogenized matrix of Material B
6 Numerical examples and C. At each Gauss point, the material microstructure is
parametrized by the volume fraction 𝜙A,n+1 , the orientation
x,j

In this section, we define two test examples with elastoplas- 𝜃A,n+1 , and the volume fraction 𝛾C,n+1 for load step (n + 1) ,
x,j x,j
tic multiphase hierarchical material models that are suitable which results in 38, 400 microscale design variables in each
to illustrate the computational efficiency and validity of our load step.
path dependent concurrent material and structure optimi- The minimum volume fraction of Material A is set to
zation framework. First, we consider a standard cantilever 𝜙min = 0.2 . The existence of the homogenized yield criterion
type benchmark problem and modify its material definitions A
𝔉 in (53) requires 𝜙̄ C = (1 − 𝜙A,n+1 ) 𝛾C,n+1 > 0 . It implies
x,j x,j
analogous to the multiscale configuration shown in Fig. 1.
that the bounds 𝜙max < 1 − h and 𝛾Cmin > h , where h is a small
Later, we demonstrate the potential of our framework for A
positive number. We restrict 𝜌min to 0.001 and 𝜌max to 0.799
biotailoring applications by solving a prototype problem
to satisfy these requirements. The total amount of material
that integrates a hierarchical material model for cereal stems
mass available is restricted to 40% of the maximum possible
Gangwar et al. (2021).
mass. As an initial condition at the macroscale, we assume
the maximum possible density 𝜌max in each element. At the
6.1 Cantilever benchmark problem
material level, we assume an initial microstructure configu-
ration 𝜙A = 0.0 , 𝜃A = 0.0 , and 𝛾C = 1.0 at each Gauss point.
6.1.1 Problem description and hierarchical design
In each design update, we reduce the target mass fraction by
0.025 until we reach the specified mass fraction Mfrac = 0.4 .
Figure 6 modifies the definition of the standard cantilever
The move parameter 𝜇 and the damping parameter 𝜂 are set
design problem to demonstrate the developed concepts in
to 0.05 and 0.5. The filter radius rmin is reduced linearly from
this article. The length and height of the macrostructure
rmin = 20 le to rmin = 4 le with design iterations for improv-
are 2.0 m and 1.0 m, respectively. The left edge is fixed,
ing the convergence of the structure optimization algorithm
and a displacement loading of u∗ = 7.5 mm is prescribed at
following Xia et al. (2017).
the central 10% of the right edge that we divide in six load
The structure optimization algorithm stops when the rela-
steps with a constant load increment of Δū E = 1.25 mm. We
tive change in the macroscale density field falls below the
discretize the macroscale structure with an 80 × 40 mesh
tolerance 𝜖tol = 10−3 . Figure 7 illustrate the convergence of
of 4-node plane strain quadrilateral elements, resulting in a
the macroscale design update. We notice that the algorithm
characteristic element size of le = 25 mm and 3, 200 mac-
takes 34 density updates to converge to the final design
roscale design variables. Each element contains four Gauss

Fig. 6  Cantilever benchmark
based on elastoplastic mul-
tiphase hierarchical materials

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 21 of 31  195

with converged objective function value 0.39022 N-mm.


Figure 8a and b plot the optimized macroscale density and
equivalent plastic strains overlaid on the density plot, respec-
tively. The plastic strains are concentrated at the clamped
end’s boundaries, consequently pushing the material towards
these regions. As highlighted in Fig. 8b, the sharp features
near the clamped end in the optimal density distribution
mimic the plastic front emphasizing its importance for the
final design.
Figure  9 illustrates the optimized morphology at the
mesoscale and the equivalent volume fraction of Material
B and C from the lowermost scale. The yellow color in
Fig. 9a represents the matrix material that results from the
homogenization of the lowermost scale, and the blue color
displays the volume fraction and orientation of Material A
inclusions. The inclusions follow the direction of the largest Fig. 7  Convergence of objective function fw (in N-mm) and mass
fraction with respect to number of macroscale design iterations
principal stress. The equivalent volume fractions of Mate-
rial A, B, and C at the macroscale are defined as: 𝜙̄ A = 𝜙A ,
𝜙̄ B = (1 − 𝜙A )(1 − 𝛾C )  , and 𝜙̄ C = (1 − 𝜙A )𝛾C   . Figure  9b plastic front, whereas these attributes are missing in the lin-
displays the equivalent volume fraction of Material B and ear elastic design.
C at the macroscale for the final design, where we use 60% Figure 12 quantitatively compares the structural perfor-
opacity for both. We can observe the regions dominated by mance of the elastoplastic design over the corresponding
Material B, C, and a mixing zone. The stiffer Material C linear elastic design. We subject the optimal configurations
is deposited in the regions anticipated to yield first, while in both cases to the same displacement loading of u∗ = 7.5
Material B dominates the transition zone. mm, whereas Material C is elastoplastic for both configu-
Figure 10 illustrates the evolution of the optimization rations (see Fig. 6). Figure 12 demonstrates the load–dis-
process by plotting equivalent plastic strains overlaid on placement curves with the equivalent plastic strains plots for
the density distribution and the equivalent volume fraction different load levels for both cases. Although the response
of Material B and C, all at selected design iterations. The at low load levels is practically identical, the load–displace-
evolution of the macroscale density and equivalent plastic ment curves start to deviate from each other at the higher
strains shows that the design process attempts to attenuate load levels, when elastoplastic behavior originating from
the plastic front. In this process, the algorithm pushes more material scales governs the overall response. The features
material towards the region close to the clamped boundary, in elastoplastic design highlighted in Fig. 8b plays a crucial
delaying yielding in this region. Material C is the stiffest role in attenuating the propagation of plastic front by delay-
material among the constituents and exhibits elastoplastic ing the plasticization of the hierarchical material system.
behavior. Its evolution is heavily influenced by the plastic Figure 12 demonstrates the differences in yielded regions
front, which leads for instance to sharp features in the mac- and the magnitude of equivalent plastic strains in both cases,
roscale density configuration. which clearly shows the superior behavior of the elastoplas-
tic design. Moreover, the relative percentage gain in the
6.1.2 Performance comparison with equivalent linear structural performance for the elastoplastic design, defined
elastic design as ((fw,nl − fw,lin )∕fw,lin ) × 100 , is 4% , where fw,nl and fw,lin are
objective function values for the elastoplastic and equivalent
We finally illustrate the impact of elastoplastic design on the linear elastic designs, respectively. This gain is expected to
structural performance by comparing with a corresponding grow with the applied load level u∗ . Thus, we conclude that
design that only assumes linear elastic material response at the linear elastic design and the plastic design are function-
the microscales and ignores any plasticity effects. To this ing differently, and it is important to consider plastic effects
end, Fig. 11 plots the optimized density distribution and the at different scales in multiphase hierarchical systems that
equivalent volume fraction of Material B and C for the lin- are expected to develop dissipation-based energy absorption
ear elastic design that assumes purely elastic properties of mechanisms against external impacts.
Material C at the lowermost scale. Comparing these plots
with Fig. 8 and 9b, we can find apparent differences in the
optimized layouts. The plastic design places more material
towards the clamped region with clear features imitating the

13
195   Page 22 of 31 T. Gangwar, D. Schillinger

Fig. 8  Macroscale density distribution and equivalent plastic strain distribution of the cantilever benchmark problem for a total prescribed dis-
placement of u∗ = 7.5 mm. Highlighting circles demonstrates that the plastic front influences the optimal density distribution

6.2 Towards predicting self‑adapting mechanisms


in plants

Biomaterials exhibit multiscale inelastic behavior and


develop dissipation-based energy absorption mechanisms
optimizing its hierarchical composition across material
microscales against the external biophysical stimuli through
natural evolution (Wegst et al. 2015; Fratzl and Weinkamer
2007; Bhushan 2009). A rational understanding of micro-
structure interdependencies with self-adapting mechanisms
will pave the way towards many biotailoring applications
with improved properties, for instance, in the context of the
targeted breeding of agricultural crops (Brulé et al. 2016;
Berry et al. 2004). A few studies have attempted the mul-
tiscale optimization of biological systems such as bone-
remodeling and bioinspired materials (Rodrigues et al. 1999;
Coelho et al. 2008; Radman et al. 2013). Several roadblocks,
however, such as high computational cost and non-trivial
problem decomposition in the case of elastoplastic behav-
ior have limited these approaches to simple linear elastic
problems with no more than two scales. With the follow-
ing prototype model, we demonstrate the potential of our
optimization framework in overcoming these roadblocks for
the computationally efficient modeling of self-adaption of
biomaterials.
Crop stem materials organize themselves hierarchically
across multiple length scales. The hierarchical scales in
crops range from base constituents such as cellulose, hemi-
cellulose, and lignin, to cell wall, functional tissues, cross-
section, and structure scale node morphology levels. In our
Fig. 9  Optimal material configuration for total prescribed displace- previous work, we experimentally profiled this hierarchi-
ment of u∗ = 7.5 mm cal organization through microimaging (micro-CT, light
microscopy, transmission electron microscopy) and chemi-
cal analysis, focusing on cereal stems (Gangwar et  al.

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 23 of 31  195

Fig. 10  Evolution of macro-
scale density configuration and
equivalent plastic strains (rain-
bow colormap, ×10−3 units )
and equivalent volume fractions
of Material B and C

2021), which we briefly summarize in Fig. 13. Exploiting behavior across different scales. We provide all implemen-
this data, we developed and validated a continuum micro- tation information relevant in the scope of this article in 1.
mechanics model of cereal stem materials that accurately Figure 14 summarizes the prototype model for the hier-
relates material composition with elastoplastic mechanical archical optimization of a cereal node region, given that

13
195   Page 24 of 31 T. Gangwar, D. Schillinger

Fig. 11  Final design of the cantilever benchmark in the equivalent linear elastic case (Material C purely elastic)

Fig. 12  Load vs displacement 1.5 2.7

curves of the final designs in


the equivalent linear elastic
and elastoplastic cases with
the equivalent plastic strains
overlaid on the optimal density 0 0

layout (in ×10−3 units, rainbow ×10-3


colormap)

5.3
Force (N)

Plastic design 12.6

Linear elastic design

Displacement (mm) 0

stem failure has been generally observed in this region. We non-mechanical, such as protecting against insects and
take the length and height of the macrostructure domain as regulating gas exchange. Thus, we only consider the solid-
10 mm and 5 mm, which is typical of a node dimension. pith region for hierarchical optimization. The microstruc-
The left edge is completely fixed, and the displacement in ture design variables mn+1 consist of the cell wall fraction
x,j

X direction on the right edge is fixed. The prescribed dis- 𝜙wall,n+1 in the parenchyma, the fiber fraction 𝜙fib,n+1 in the
par(x,j) x,j

placement in the Y direction on the right edge is u∗ = 0.4


vascular bundles, the vascular bundle fraction 𝜙vb,n+1 , and
x,j
mm. The displacement u∗ = 0.4 mm is divided in four load
the orientation 𝜃n+1 of the anisotropy axis of the solid pith
x,j
steps with increment Δū E = 0.1 mm. We discretize the
macroscale structure with a 60 × 30 mesh of 4-node plane- with respect to the global X direction (see Fig. 14). Lignin
strain quadrilateral elements. This macrostructure model in the parenchyma cell wall material exhibits elastoplastic
definition is equivalent to a cereal node cross-section that behavior. The parenchyma tissues and xylem-phloem ves-
undergoes combined shear and bending loads. sels in the vascular bundles are also responsible for food
Following our multiscale material model, the stem storage and nutrient-water transport. We incorporate these
cross-section consists of an outer-shell layer and a solid- biological constraints by adopting the bounds on the vol-
pith region. The primary functions of the outer-shell are ume fractions that we measured through microimaging

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 25 of 31  195

Fig. 13  Hierarchical structure of a cereal plant profiled through microimaging Gangwar et al. (2021)

Gangwar et al. (2021) in the material optimization prob- 2015). The plastic strains are concentrated at the end and
lem. At the structure scale, the total amount of material is middle regions due to the anticipated high shear deforma-
restricted by the reported average density, which can be tions. The optimal density layout puts material in areas to
interpreted as the limitation posed by the available biologi- attenuate plastic fronts, which can also be observed in the
cal energy required in the synthesis of biomass per unit of design evolution history in Fig. 15. These observations
stem material. again emphasize the role of the plastic front in the optimal
Figures 15 and 16 illustrate the design evolution his- structural layout.
tory and the final macroscale density with the equivalent Figure 17 plots the optimal microstructure configuration
plastic strains. The macroscale design algorithm takes 38 at different scales at the final load level. At the mesoscale,
density updates to converge to the final design. In the opti- the material orientation follows the stress flow direction in
mal layout, the branches from the left internode converge the main branches, while the morphology is more complex
to the central node region, and the branches of the right in the central node region. We also plot the optimal con-
internode emerges, a morphology that was also observed figurations at lower material scales in the main branches.
in real plants through micro-CT images (Ghaffar and Fan These results indicate the choice of a stronger solid-pith

Macrostructure Microscales
Vascular bundles

Parenchyma

Y Liquid
symplast Vessels
H = 5 mm

Fibers
60 × 30 θ
X Cell wall
material

Hemicellulose

Lignin
L = 10 mm Cellulose (elasto-plastic)

Fig. 14  Prototype model for the hierarchical optimization of a cereal node region

13
195   Page 26 of 31 T. Gangwar, D. Schillinger

material for the optimal mechanical response. The predicted 7 Summary, conclusions, and outlook
morphology is in qualitative agreement with what our col-
laborators in plant science have found via field experiments In this article, we established rigorous theoretical foun-
for lodging-resistant cereals (Gangwar et al. 2023). Based dations for an efficient concurrent material and structure
on these promising results, we believe that our optimiza- optimization framework for multiphase hierarchical sys-
tion framework can help pave the way towards efficient and tems with elastoplastic constituents at the material scales.
sustained biotailoring applications, supported by modeling In particular, we developed an efficient solution strategy
and simulation. for the material optimization problem based on the maxi-
mum plastic dissipation principle in the format of a typi-
cal return map algorithm for an elastoplastic constitutive
law. Finally, we integrated analytical expressions of the

Fig. 15  Convergence of the
objective function fw (in 10−3
N-mm) with respect to the num-
ber of design iterations, plotted
along with snapshots of the
macroscale density configura-
tion and the equivalent plastic
strains (in ×10−3 units, rainbow
colormap)

Fig. 16  Final design of the cereal node region with equivalent plastic strain distribution for total prescribed displacement u∗ = 0.4 mm

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 27 of 31  195

homogenized stiffness and the yield criterion that are field analysis (TFA) (Dvorak and Benveniste 1992; Brassart
derived via continuum micromechanics, enabling a com- et al. 2011). Moreover, the formulation can potentially be
putationally tractable implementation for elastoplastic extended for other path-dependent problems, where nonlin-
multiphase hierarchical systems. ear effects such as viscoplasticity, fracture, and damage orig-
We verified the validity and efficiency of our framework inate from the material microscales. Thus, our framework
with newly defined benchmark problems that, for the first helps push forward path-dependent concurrent material and
time, is computationally feasible via our framework. It con- structure optimization to consider nonlinearities exhibited
sists of a macroscale configuration in the form of a standard at the microscales, with a number of potential applications,
cantilever, but involves hierarchical material definitions with including multiscale additive manufacturing and architect-
elastoplastic constituents at the microscale. The optimized ing metamaterials (Meza et al. 2015; Sanders et al. 2021).
macroscale and microstructure configurations computed via
our framework demonstrated the importance of plasticity
effects that originate from the material microscales in devel- Appendix A Comments on continuum
oping dissipation-based energy absorbing mechanisms. In micromechanics‑based simplifications
addition, we applied our framework for investigating self-
adapting mechanisms in cereal plant structures, outlining We present a proof of Property 3 introduced in Sect. 4.3.
its potential for biotailoring applications. Our framework First, we write the strain energy maximization expres-
is a first attempt at a decomposed material and structure sions in index notation. We denote the orthonormal basis
formulation that optimizes the path-dependent macroscale that corresponds to the global coordinate system as {ep } .
mechanical response of elastoplastic multiphase hierarchi- We drop superscript (x, j) from variables for conciseness.
cal systems. We would also like to point out the limitations Problem (51) at load increment n with Een ∶= En − Epn can
of the presented framework. In this article, we restricted be rewritten in index notations as
ourselves to the class of inelastic hierarchical materials,
1
for which we can assume the existence of an associative max Een ∶ ℂ(mn ) ∶ Een
mn 2
flow rule, a rate-independent ideal elastoplastic response,
1 e
and an isothermal process at the macroscale. We also con- = max Epq(n) ep ⊗ eq ∶ Cpqrs (mn )ep ⊗ eq ⊗ er ⊗ es ∶
fined the quadratic stress average based micromechanical
mn 2
(70)
e
scheme for estimating the homogenized yield criterion by Ers(n) e r ⊗ es
an additional assumption that only one of the constituents 1 e e
= max Epq(n) Cpqrs (mn ) Ers(n) .
in the hierarchical material exhibits inelastic behavior with mn 2
a deviatoric stress-based yield criterion. These assumptions
restrict the plausible elastoplastic failure mechanisms that The principle coordinate system, that is co-linear with the
originate from the material microscales. We anticipate that principal strain directions, uses the Roman indexed basis
these mechanisms can be incorporated into our optimization {̂ei } . The transformation matrix Qpi between both systems
framework, for instance by combining the variational pro- depends on the optimal configuration 𝜃̄A,n from Property 2.
cedure for incremental homogenization and transformation

Fig. 17  Optimal microstructure Optimal microstrcutre at mesoscale Optimal microscales


configuration for total pre-
scribed displacement u∗ = 0.4
in main branch
mm

Parenchyma
Vascular bundles

Vessels
Fibers

Cell wall
Liquid material
symplast
Y

13
195   Page 28 of 31 T. Gangwar, D. Schillinger

Utilizing Qpi and its orthogonal property, we define the ten- Ê ij(n+1)
e
Ĉ ijkl (m
̄ ln ) Ê kl(n+1)
e

sor component transformations as
Ê ij(n+1)
e
Ĉ ijkl (mln+1 ) Ê kl(n+1)
e
(76)
ê i = Qpi ep ; Ê ij(n)
e
= e
Qpi Qqj Epq(n) ; ⟹ ̄ ln+1
m = ̄ ln .
m
(71)
Cpqrs = Qpi Qqj Qrk Qsl Ĉ ijkl ,

where Ê ij(n)
e
are the components of the elastic part of the
macroscale strain tensor in the principle coordinate system. Appendix B Details on the cereal prototype
Also, Ê ij(n)
e
= 0 if i ≠ j , and diagonal components ( i = j ) are model
the principle strain values. Similarly, Ĉ ijkl are the compo-
nents of the stiffness tensor in the principal coordinate sys- We briefly summarize the key components for implementing
tem. Using (71), we reformulate the maximization problem the hierarchical optimization of the cereal prototype model in
(70) in terms of mln = [𝜙A,n , 𝜁A,n , 𝛾C,n ] as Sect. 6.2. For further details on multiscale modeling of stiff-
ness and strength of crop stem material within continuum
1 ̂e ̂ micromechanics, we refer the interested reader to Section 3 in
max E Cijkl (mln ) Ê kl(n)
e
. (72)
mln 2 ij(n) Gangwar et al. (2021). In the current model, the microscale
design variables defined at each Gauss point at load step
With m̄ ln as the solution to this maximization problem, we
(n + 1) are the cell wall fraction 𝜙wall,n+1 in the parenchyma,
par(x,j)
can rewrite (72) as
the fiber fraction 𝜙fib,n+1 in the vascular bundles, the vascular
x,j

Ê ij(n)
e
Ĉ ijkl (m
̄ ln ) Ê kl(n)
e
≥ Ê ij(n)
e
Ĉ ijkl (mln ) Ê kl(n)
e
. (73) bundle fraction 𝜙vb,n+1 , and the orientation 𝜃n+1 of the anisot-
x,j x,j

ropy axis of the solid-pith material. The macroscale homog-


We assume that the macroscale loading increases monotoni-
enized stiffness tensor ℂ in the global coordinate system can
cally. Therefore, the elastic part of the macroscale tensor
be written as a function of the microscale design variables
Ê ij(n+1)
e
at load increment (n + 1) in the principal coordinate
mn+1 = [𝜙wall,n+1 , 𝜙fib,n+1 , 𝜙vb,n+1 , 𝜃n+1 ] following Equation
x,j par(x,j) x,j x,j x,j
system can be written in terms of the components Ê ij(n)e
with
(27) in Gangwar et al. (2021) as
an appropriate scaling. Exploiting the definition of the Kro-
necker delta 𝛿 , we write Ê ij(n+1)
e
in terms of scaling compo- x,j x,j
ℂ(mn+1 ) = [T(𝜃n+1 )]−1 ℂpith [T(𝜃n+1 )] with
x,j

nents ai𝛼 as par(x,j) x,j x,j


ℂpith (ℂpar (𝜙wall,n+1 ), ℂvb (𝜙fib,n+1 ), 𝜙vb,n+1 )
Ê ij(n+1)
e
= ai𝛼 𝛿j𝛼 Ê ij(n)
e
and ai𝛼 ≥ 0. (74)
{
x,j x,j
= (1 − 𝜙vb,n+1 ) ℂpar + 𝜙vb,n+1 ℂvb ∶ [𝕀+
}
As the scaling components are non-negative, we augment par
ℙcyl ∶ (ℂvb − ℂpar )]−1 ∶
the expression (73) and arrive at
{ }−1
x,j x,j par
ai𝛼 𝛿j𝛼 Ê ij(n)
e
Ĉ ijkl (m
̄ ln ) ak𝛽 𝛿l𝛽 Ê kl(n)
e (1 − 𝜙vb,n+1 ) 𝕀 + 𝜙vb,n+1 [𝕀 + ℙcyl ∶ (ℂvb − ℂpar )]−1 .

≥ ai𝛼 𝛿j𝛼 Ê ij(n)


(77)
e
Ĉ ijkl (mln ) ak𝛽 𝛿l𝛽 Ê kl(n)
e

(75) Here, ℂpar and ℂvb are the homogenized stiffness tensors of
⟹ Ê ij(n+1)
e
Ĉ ijkl (m
̄ ln ) Ê kl(n+1)
e
the parenchyma tissue and the vascular bundle tissue in the
≥ Ê ij(n+1)
e
Ĉ ijkl (mln ) Ê kl(n+1)
e
. solid-pith region, respectively. For the analytical expression
of these estimates, interested readers are referred to Sec-
This is the expression for the strain energy maximization tion 3.3 in Gangwar et al. (2021). The composition of all
problem at load increment (n + 1) analogous to (72) or (73). other RVEs in the multiscale material model of cereal stems
We emphasize that our definition of the admissible set is considered constant, and model parameters correspond-
Ead depends only on the macroscale density distribution and, ing to the Gopher oat variety are used (see Appendix B in
therefore, remains unchanged throughout the loading his- Gangwar et al. (2021)). T is a standard rotation matrix for
tory. In addition, Case 2 is not conceivable as laid out in our tensor transformations.
discussion of Property 1 in Sect. 4.3. Therefore, it implies Lignin exhibits elastoplastic material behavior at the con-
that the possible admissible solutions of the right hand side stituent level, and the macroscale limit state point corresponds
expression at increment (n + 1) in 75 are the same as those to the yielding of lignin. In this prototype model, we assume
at increment n. Replacing mln with mln+1 , we arrive at that only lignin in parenchyma cell wall material is elastoplas-
tic (see Fig. 14). In this case, the macroscale homogenized
yield criterion reads as

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 29 of 31  195

x,j
𝔉(𝝉, mn+1 ) = The setup of the structure optimization problem for this
√ prototype model resembles (42). The sensitivity analysis


√ x,j
x,j
𝜕 ℂ(mn+1 ) x,j
for the macroscale design updates follows from Sect. 5.1.
𝝉 ∶ [ℂ(mn+1 )]−1 ∶ ∶ [ℂ(mn+1 )]−1 ∶ 𝝉 The derivative of the homogenized stiffness ℂ with respect
𝜕 𝜇lig,par (78)
√ to the element density 𝜌j follows via the chain rule as
Y
𝜙lig,par 𝜎lig x,j par(x,j)
− , 𝜕ℂ(mn+1 ) 𝜕ℂ 𝜕𝜙wall,n+1
3 𝜇lig = par(x,j)
𝜕𝜌j 𝜕𝜙wall,n+1 𝜕𝜌j
where lignin follows the von Mises failure criterion with x,j x,j
𝜕ℂ 𝜕𝜙fib,n+1 𝜕ℂ 𝜕𝜙vb,n+1
yield strength 𝜎lig and the bulk modulus 𝜇lig . The equivalent .
Y
+ x,j
+ x,j
𝜕𝜙fib,n+1 𝜕𝜌j 𝜕𝜙vb,n+1 𝜕𝜌j
volume fraction 𝜙lig,par of lignin in parenchyma is computed
as 𝜙lig,par = 𝜙l
wall,par
× 𝜙wall,n+1 × (1 − 𝜙vb,n+1 ) . The lignin
par(x,j) x,j (80)
volume fraction 𝜙l
wall,par
in the parenchyma cell wall material The derivatives of ℂ with respect to the microscale design
is fixed as given in Gangwar et al. (2021). variables at the material level are evaluated by finite differ-
With these definitions in hand, we write the material opti- ence approximations. The move parameter 𝜇 and the damp-
mization problem for a Gauss point x inside element j for load ing parameter 𝜂 are set to 0.02 and 0.5. The filter radius rmin
increment (n + 1) following (45) as is reduced linearly from rmin = 15 le to rmin = 4 le with design
{ iterations.
x,j p
m̄ n+1 = arg max 𝚺n+1 ∶ (En+1 − Epn )+ Acknowledgements  This work has received funding from the Euro-
x,j
mn+1 (𝜌j )
pean Research Council (ERC) under the European Union’s Horizon
}
p 2020 research and innovation programme (Grant agreement No.
Ψ(En+1 − En+1 ) − Ψ(En − Epn ) 759001). This support is gratefully acknowledged. The authors are
x,j p also thankful to Pramesh Kumar (University of Minnesota) for helpful
s.t. ∶ 𝚺n+1 = ℂ(mn+1 ) ∶ (En+1 − En+1 ) discussions and comments.
𝔉(𝚺n+1 , mn+1 ) ≤ 0
x,j
Funding  Open Access funding enabled and organized by Projekt
p 1 p DEAL.
Ψ(En+1 − En+1 ) = (E − En+1 ) ∶
2 n+1
x,j p
ℂ(mn+1 ) ∶ (En+1 − En+1 ) Declarations 
par(x,j) x,j
𝜌j = 𝜌wall 𝜙wall,n+1 (1 − 𝜙vb,n+1 )+ Conflict of interest  The authors have no conflict of interest to disclose.
x,j x,j
𝜌fib 𝜙fib,n+1 𝜙vb,n+1 Replication of results  All the necessary information for possible repli-

≤ 𝜙wall,n+1 ≤
par,min par(x,j) par,max cation of all the results are presented in the manuscript. The manuscript
𝜙wall 𝜙wall ; also has a dedicated section on the detailed algorithmic treatment to

fib ≤ 𝜙fib,n+1 ≤ 𝜙max


x,j facilitate the reproduction of the results. The developed code (written
𝜙min fib ; in python) can be made available from the first author on request.

vb ≤ 𝜙vb,n+1 ≤ 𝜙max − 𝜋∕2 ≤ 𝜃n+1 ≤ 𝜋∕2,


x,j x,j
𝜙min vb ;
Open Access  This article is licensed under a Creative Commons Attri-
(79)
bution 4.0 International License, which permits use, sharing, adapta-
where 𝜌j is the given macroscale dry density for the finite tion, distribution and reproduction in any medium or format, as long
element with index j. The first three lines in the constraints as you give appropriate credit to the original author(s) and the source,
provide a link to the Creative Commons licence, and indicate if changes
definition represent the microstructure dependent constitu- were made. The images or other third party material in this article are
tive equations. The fourth statement connects 𝜌j with the included in the article's Creative Commons licence, unless indicated
microscale design variables via the rule of mixture. We otherwise in a credit line to the material. If material is not included in
adopt bounds for 𝜙wall,n+1  , 𝜙fib,n+1 and 𝜙vb,n+1 that are
par(x,j) x,j x,j the article's Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will
[0.01, 0.38] , [0.75, 0.90] , and [0.01, 0.16] , respectively. need to obtain permission directly from the copyright holder. To view a
The upper bound reflects the measured microscale parame- copy of this licence, visit http://​creat​iveco​mmons.​org/​licen​ses/​by/4.​0/.
ters reported in Gangwar et al. (2021). The strain energy
maximization part in the algorithmic treatment of the mate-
rial optimization problem is a constraint optimization prob- References
lem with nonlinear equality constraint. We utilize the
sequential least squares programming (SLSQP) method Allaire G, Aubry S (1999) On optimal microstructures for a plane
implemented in the SciPy library to solve this problem. shape optimization problem. Struct Optim 17(2–3):86–94

13
195   Page 30 of 31 T. Gangwar, D. Schillinger

Bendsøe MP, Sigmund O (1999) Material interpolation schemes in modeling of crop stem materials. Biomech Model Mechano-
topology optimization. Arch Appl Mech 69(9–10):635–654 biol 20(1):69–91
Berry P, Sterling M, Spink J, Baker C, Sylvester-Bradley R, Mooney Gangwar T, Susko AQ, Baranova S, Guala M, Smith KP, Heuschele DJ
S, Tams A, Ennos A (2004) Understanding and reducing lodging (2023) Multi-scale modelling predicts plant stem bending behav-
in cereals. Adv Agron 84(04):215–269 iour in response to wind to inform lodging resistance. R Soc Open
Bessa M, Bostanabad R, Liu Z, Hu A, Apley DW, Brinson C, Chen W, Sci 10(1):221410
Liu WK (2017) A framework for data-driven analysis of materials Gao H, Ji B, Jäger IL, Arzt E, Fratzl P (2003) Materials become insen-
under uncertainty: countering the curse of dimensionality. Comput sitive to flaws at nanoscale: lessons from nature. Proc Natl Acad
Methods Appl Mech Eng 320:633–667 Sci 100(10):5597–5600
Bhushan B (2009) Biomimetics: lessons from nature—an overview. Germain P (1973) The method of virtual power in continuum mechan-
Philos Trans R Soc A 367(1893):1445–1486 ics part 2: microstructure. SIAM J Appl Math 25(3):556–575
Blanchard R, Morin C, Malandrino A, Vella A, Sant Z, Hellmich C Ghaffar SH, Fan M (2015) Revealing the morphology and chemical dis-
(2016) Patient-specific fracture risk assessment of vertebrae: a tribution of nodes in wheat straw. Biomass Bioenerg 77:123–134
multiscale approach coupling x-ray physics and continuum micro- Gibson LJ (2012) The hierarchical structure and mechanics of plant
mechanics. Int J Num Methods Biomed Eng 32(9):e02760 materials. J R Soc Interface 9(76):2749–2766
Brassart L, Stainier L, Doghri I, Delannay L (2011) A variational for- Hellmich C, Ulm F-J, Dormieux L (2004) Can the diverse elastic prop-
mulation for the incremental homogenization of elasto-plastic erties of trabecular and cortical bone be attributed to only a few
composites. J Mech Phys Solids 59(12):2455–2475 tissue-independent phase properties and their interactions? Bio-
Brulé V, Rafsanjani A, Pasini D, Western TL (2016) Hierarchies of mech Model Mechanobiol 2(4):219–238
plant stiffness. Plant Sci 250:79–96 Hilchenbach CF, Ramm E (2015) Optimization of multiphase structures
Buhl T, Pedersen CB, Sigmund O (2000) Stiffness design of geomet- considering damage. Struct Multidisc Optim 51(5):1083–1096
rically nonlinear structures using topology optimization. Struct Hofstetter K, Hellmich C, Eberhardsteiner J (2005) Development and
Multidisc Optim 19(2):93–104 experimental validation of a continuum micromechanics model
Cho S, Jung H-S (2003) Design sensitivity analysis and topology opti- for the elasticity of wood. Eur J Mech-A Solids 24(6):1030–1053
mization of displacement-loaded non-linear structures. Comput Hofstetter K, Hellmich C, Eberhardsteiner J, Mang HA (2008) Micro-
Methods Appl Mech Eng 192(22–24):2539–2553 mechanical estimates for elastic limit states in wood materials,
Coelho PG, Fernandes PR, Guedes JM, Rodrigues HC (2008) A hier- revealing nanostructural failure mechanisms. Mech Adv Mater
archical model for concurrent material and topology optimisa- Struct 15(6–7):474–484
tion of three-dimensional structures. Struct Multidisc Optim Höller R, Aminbaghai M, Eberhardsteiner L, Eberhardsteiner J, Blab
35(2):107–115 R, Pichler B, Hellmich C (2019) Rigorous amendment of Vlasov’s
Da D (2019) Topology optimization design of heterogeneous materials theory for thin elastic plates on elastic Winkler foundations, based
and structures. Wiley, Hoboken on the principle of virtual power. Eur J Mech A 73:449–482
Da D, Cui X, Long K, Li G (2017) Concurrent topological design of Holstov A, Bridgens B, Farmer G (2015) Hygromorphic materi-
composite structures and the underlying multi-phase materials. als for sustainable responsive architecture. Constr Build Mater
Comput Struct 179:1–14 98:570–582
de Souza Neto EA, Peric D, Owen DR (2011) Computational methods Huang X, Xie Y (2008) Optimal design of periodic structures using
for plasticity: theory and applications. Wiley, Hoboken evolutionary topology optimization. Struct Multidisc Optim
Dvorak GJ, Benveniste Y (1992) On transformation strains and uni- 36(6):597–606
form fields in multiphase elastic media. Proc R Soc Lond A Hughes TJ (2000) The finite element method: linear static and dynamic
437(1900):291–310 finite element analysis. Dover Publications, Mineola
Egan P, Sinko R, LeDuc PR, Keten S (2015) The role of mechanics Jog CS, Haber RB, Bendsøe MP (1994) Topology design with
in biological and bio-inspired systems. Nat Commun 6(1):1–12 optimized, self-adaptive materials. Int J Numer Meth Eng
Feyel F, Chaboche J-L (2000) FE2 multiscale approach for modelling 37(8):1323–1350
the elastoviscoplastic behaviour of long fibre SiC/Ti composite Kalliauer J, Kahl G, Scheiner S, Hellmich C (2020) A new approach
materials. Comput Methods Appl Mech Eng 183(3–4):309–330 to the mechanics of dna: atoms-to-beam homogenization. J Mech
Fish J (2013) Practical multiscaling. Wiley, Hoboken Phys Solids 143:104040
Fratzl P, Weinkamer R (2007) Nature’s hierarchical materials. Prog Kato J (2010) Material optimization of fiber reinforced composites
Mater Sci 52(8):1263–1334 applying a damage formulation. PhD Thesis, University of Stutt-
Fritsch A, Hellmich C, Dormieux L (2009) Ductile sliding between gart, Germany
mineral crystals followed by rupture of collagen crosslinks: Kato J, Ramm E (2013) Multiphase layout optimization for fiber
experimentally supported micromechanical explanation of bone reinforced composites considering a damage model. Eng Struct
strength. J Theor Biol 260(2):230–252 49:202–220
Fritzen F, Xia L, Leuschner M, Breitkopf P (2016) Topology optimi- Laws N (1977) The determination of stress and strain concentrations
zation of multiscale elastoviscoplastic structures. Int J Numer at an ellipsoidal inclusion in an anisotropic material. J Elast
Meth Eng 106(6):430–453 7(1):91–97
Gangwar T, Schillinger D (2019) Microimaging-informed continuum Laws N (1985) A note on penny-shaped cracks in transversely isotropic
micromechanics accurately predicts macroscopic stiffness and materials. Mech Mater 4(2):209–212
strength properties of hierarchical plant culm materials. Mech Le B, Yvonnet J, He Q-C (2015) Computational homogenization of
Mater 130:39–57 nonlinear elastic materials using neural networks. Int J Numer
Gangwar T, Schillinger D (2021) Concurrent material and struc- Meth Eng 104(12):1061–1084
ture optimization of multiphase hierarchical systems within a Li P, Wu Y, Yvonnet J (2021) A SIMP-phase field topology optimiza-
continuum micromechanics framework. Struct Multidisc Optim tion framework to maximize quasi-brittle fracture resistance of
64:1175–1197 2D and 3D composites. Theoret Appl Fract Mech 114:102919
Gangwar T, Heuschele DJ, Annor G, Fok A, Smith KP, Schillinger Lipton R (1994) A saddle-point theorem with application to structural
D (2021) Multiscale characterization and micromechanical optimization. J Optim Theory Appl 81(3):549–568

13
Thermodynamically consistent concurrent material and structure optimization of elastoplastic… Page 31 of 31  195

Liu Z, Bessa M, Liu WK (2016) Self-consistent clustering analysis: an Simo JC, Taylor RL (1985) Consistent tangent operators for rate-
efficient multi-scale scheme for inelastic heterogeneous materials. independent elastoplasticity. Comput Methods Appl Mech Eng
Comput Methods Appl Mech Eng 306:319–341 48(1):101–118
Masson R (2008) New explicit expressions of the Hill polarization Suquet P (1997) Effective properties of nonlinear composites. In: Con-
tensor for general anisotropic elastic solids. Int J Solids Struct tinuum micromechanics. Springer, pp 197–264
45(3–4):757–769 Suquet P (2014) Continuum micromechanics, vol 377. Springer, Berlin
Maute K, Schwarz S, Ramm E (1998) Adaptive topology optimization Swan CC, Kosaka I (1997) Voigt-Reuss topology optimization for
of elastoplastic structures. Struct Optim 15(2):81–91 structures with nonlinear material behaviors. Int J Numer Meth
McCann MC, Buckeridge MS, Carpita NC (2014) Plants and bioen- Eng 40(20):3785–3814
ergy. Springer, Berlin Tadmor EB, Miller RE, Elliott RS (2012) Continuum mechanics and
Meza LR, Zelhofer AJ, Clarke N, Mateos AJ, Kochmann DM, Greer JR thermodynamics: from fundamental concepts to governing equa-
(2015) Resilient 3D hierarchical architected metamaterials. Proc tions. Cambridge University Press, Cambridge
Natl Acad Sci 112(37):11502–11507 Truesdell C, Noll W (2004) The non-linear field theories of mechanics.
Morin C, Vass V, Hellmich C (2017) Micromechanics of elastoplastic In: The non-linear field theories of mechanics. Springer, pp 1–579
porous polycrystals: theory, algorithm, and application to osteonal Wegst UG, Bai H, Saiz E, Tomsia AP, Ritchie RO (2015) Bioinspired
bone. Int J Plast 91:238–267 structural materials. Nat Mater 14(1):23
Nakshatrala PB, Tortorelli DA, Nakshatrala K (2013) Nonlinear struc- Wolff J (1986) The law of bone remodelling (Das Gesetz der Transfor-
tural design using multiscale topology optimization. Part I: static mation der Knocken). Springer, Berlin
formulation. Comput Methods Appl Mech Eng 261:167–176 Wu J, Sigmund O, Groen JP (2021) Topology optimization of multi-
Nguyen LH, Schillinger D (2019) The multiscale finite element method scale structures: a review. Struct Multidisc Optim 63:1455–1480
for nonlinear continuum localization problems at full fine-scale Xia L, Breitkopf P (2014) Concurrent topology optimization design of
fidelity, illustrated through phase-field fracture and plasticity. J material and structure within FE2 nonlinear multiscale analysis
Comput Phys 396:129–160 framework. Comput Methods Appl Mech Eng 278:524–542
Pedersen P (1989) On optimal orientation of orthotropic materials. Xia L, Breitkopf P (2015) Multiscale structural topology optimiza-
Struct Optim 1(2):101–106 tion with an approximate constitutive model for local material
Pichler B, Hellmich C (2011) Upscaling quasi-brittle strength of microstructure. Comput Methods Appl Mech Eng 286:147–167
cement paste and mortar: a multi-scale engineering mechanics Xia L, Fritzen F, Breitkopf P (2017) Evolutionary topology opti-
model. Cem Concr Res 41(5):467–476 mization of elastoplastic structures. Struct Multidisc Optim
Radman A, Huang X, Xie Y (2013) Topology optimization of func- 55(2):569–581
tionally graded cellular materials. J Mater Sci 48(4):1503–1510 Xia L, Da D, Yvonnet J (2018) Topology optimization for maximiz-
Ritchie RO, Buehler MJ, Hansma P (2009) Plasticity and toughness in ing the fracture resistance of quasi-brittle composites. Comput
bone. Phys Today 62(6):41 Methods Appl Mech Eng 332:234–254
Rodrigues HC, Jacobs C, Guedes JM, Bendsøe MP (1999) Global and Yuan Z, Fish J (2009) Multiple scale eigendeformation-based reduced
local material optimization models applied to anisotropic bone order homogenization. Comput Methods Appl Mech Eng
adaptation. In: Pedersen P, Bendsøe MP (eds) IUTAM sympo- 198(21–26):2016–2038
sium on synthesis in bio solid mechanics. Springer, Netherlands, Zaoui A (2002) Continuum micromechanics: survey. J Eng Mech
Dordrecht, pp 209–220 128(8):808–816
Rodrigues HC, Guedes JM, Bendsoe MP (2002) Hierarchical optimiza- Zhang Y, Xiao M, Li H, Gao L, Chu S (2018) Multiscale concur-
tion of material and structure. Struct Multidisc Optim 24(1):1–10 rent topology optimization for cellular structures with multiple
Sanders E, Pereira A, Paulino G (2021) Optimal and continuous mul- microstructures based on ordered SIMP interpolation. Comput
tilattice embedding. Sci Adv 7(16):eabf4838 Mater Sci 155:74–91
Schwarz S, Maute K, Ramm E (2001) Topology and shape optimiza- Zheng X, Lee H, Weisgraber TH, Shusteff M, DeOtte J, Duoss EB,
tion for elastoplastic structural response. Comput Methods Appl Kuntz JD, Biener MM, Ge Q, Jackson JA (2014) Ultralight, ultras-
Mech Eng 190(15–17):2135–2155 tiff mechanical metamaterials. Science 344(6190):1373–1377
Sigmund O (2001) A 99 line topology optimization code written in
Matlab. Struct Multidisc Optim 21(2):120–127 Publisher's Note Springer Nature remains neutral with regard to
Simo JC, Hughes TJ (2006) Computational inelasticity, vol 7. Springer, jurisdictional claims in published maps and institutional affiliations.
Berlin

13

You might also like