Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Archive of Applied Mechanics

https://doi.org/10.1007/s00419-024-02549-x

O R I G I NA L

Xiaozhe Ju · Kang Gao · Junxiang Huang · Hongshi Ruan ·


Haihui Chen · Yangjian Xu · Lihua Liang

A three-dimensional computational multiscale micromorphic


analysis of porous materials in linear elasticity

Received: 25 June 2023 / Accepted: 9 January 2024


© The Author(s), under exclusive licence to Springer-Verlag GmbH Germany, part of Springer Nature 2024

Abstract We present an extension of a multiscale micromorphic theory to three-dimensional problems for


porous materials, where a clear scale separation is not given. Following the multiscale micromorphic frame-
work of Biswas and Poh (J Mech Phys Solids 102:187–208, 2017), macroscopic governing equations of a
micromorphic continuum are derived from a classical continuum on the microscale by means of a kinematic
field decomposition. The macro–microenergy equivalence is guaranteed via the Hill–Mandel condition. For
linear elasticity problems, generalized elasticity tensors are determined via several RVE computations once
and for all, avoiding concurrent RVE computations for an online structural analysis. A three-dimensional
implementation is revealed in detail. A comparative study with direct numerical simulations and first-order
multiscale computations shows that the computational multiscale micromorphic method has sufficient accuracy
and computational efficiency, thus providing a powerful tool for design and practical engineering applications
of porous materials.

Keywords Porous materials · Micromorphic continua · Multiscale methods · Representative volume


element · Three-dimensional implementation

1 Introduction

Metal foams are a kind of functional porous materials, with ultra-light relative density, excellent impact resis-
tance and energy absorption characteristics. They have a wide range of applications in aerospace, automotive,
construction and flexible electronic devices [2]. Their effective mechanical properties mainly depend on the
mechanical properties, porosity and microscopic characteristics of the matrix [3]. Direct numerical simula-
tions (DNS) of those materials, including all the details of the microstructure, are unbearably computationally
intensive. Therefore, in order to facilitate the study of metal foams, a more efficient method that takes into
account both computational efficiency and computational accuracy is urgently needed.
Multiscale computational methods have been a research field in computational mechanics for a few decades.
They were developed based on the concept of representative volume element (RVE) [4], to find an equivalent
homogeneous material that can characterize the macroscopic properties of heterogeneous materials. For a

X. Ju · K. Gao · J. Huang · H. Ruan · Y. Xu (B) · L. Liang (B)


College of Mechanical Engineering, Zhejiang University of Technology, Hangzhou 310023, China
E-mail: xuyangjian571@163.com

L. Liang
E-mail: lianglihua@zjut.edu.cn

H. Chen
Zhejiang Huayun Information Technology Co., Ltd., Hangzhou 310000, China
X. Ju et al.

microscopic analysis, a direct way to predict the mechanical behavior of nonlinear heterogeneous materials is
the numerical method, e.g., finite element method (FE2 ) [5] or fast Fourier transform (FFT)-based methods
[6]. The relevant microscopic boundary value problem is solved at each macroscopic material point, and the
homogenization method [7–9] is used to predict the macroscopic equivalent mechanical behavior of hetero-
geneous materials. As a merit, such methods are able to consider 1. large deformations and rotations at all
scales, 2. arbitrary material behavior (physical non-linearity, time dependence, etc.) and 3. complex mesoscale
features (inclusions or grains). However, the huge amount of computational costs and storage requirements
make those methods infeasible for structural analysis. In order to improve the computational efficiency, differ-
ent model-order reduction methods have been developed, e.g., transformation field analysis [10], nonuniform
transformation field analysis [11,12], self-consistent clustering analysis [13], etc.
Metal foams can be regarded as a composite material composed of holes and metal matrix, where above
multiscale methods apply. For example, Gibson et al. [14] derived the elastic modulus expression of the metal
foam model by establishing an RVE. Based on the asymptotic homogenization method, Zhuang et al. [15] used
the finite element method to carry out elastic numerical analysis of porous materials and gave a comparative
verification with Gibson’s theoretical results. Liu et al. [16] established a simplified octahedral RVE model
and derived the formula for calculating the tensile strength of metal materials with high porosity.
However, the above-mentioned multiscale methods are based on the assumption of scale separation and are
often referred to as first-order computational homogenization [5,7,17], where merely first-order gradients of
displacement fields are involved. If there is a significant size effect in the considered material, the scale separa-
tion assumption of those multiscale methods is no longer valid, as also shown in experiments [18]. For problems
without a clear scale separation, Kouznetsova [19] proposed a second-order computational homogenization
method by introducing second-order gradients of macroscopic displacement fields. Microscopic equations
are homogenized to obtain a macroscale second-order gradient continuum. As a drawback, the second-order
gradient method generally requires the basic fields to satisfy C1 continuity. One has to resort to C1-continuous
elements or employing discontinuous Galerkin method [20,21] or Lagrange method with additional kinematic
constraints [22,23] to relax the C1-continuity requirement.
As an alternative, higher-order theory refines the kinematic description of materials by introducing addi-
tional degrees of freedom to account for the deformation of the microscopic continuum. Both micromorphic
theory and micropolar theory are high-order theories. The micromorphic theory was originally proposed by
Eringen [24]and Mindlin [25], with additional microscopic degrees of freedom (DoFs) attached to each mate-
rial point. Those microscopic DoFs describe full deformation of the underlying microstructure. The work [26]
focused on the Cosserat (or micropolar) theory, which merely contains additional rotational degrees of freedom
to describe the rigid rotation of the microstructure. Hence, the micropolar theory is regarded as a special case
of the micromorphic theory. Homogenization approaches toward a Cosserat continuum date back to [27] and
were further developed, e.g., by [28–32]. Our previous works further developed goal-oriented adaptive finite
element methods for linear elastic micromorphic continua in [33] and its inverse problem in [34], hyperelastic
micromorphic continua in [35] and micromorphic plasticity in [36]. In a multiscale context, a microscale
Cauchy continuum was homogenized toward a micromorphic continuum in the macroscale by Hütter [37],
where the microscopic equilibrium condition has to be appended with some non-standard body forces. Rokoš
et al. [38] developed a micromorphic computational homogenization framework for elastomeric metamaterials
exhibiting long-range correlated fluctuation fields. Biswas and Poh [1] proposed a computational micromor-
phic homogenization framework for composites of inclusion-matrix type, where a micromorphic continuum is
derived on the macroscale without any a priori assumptions. Computational micromorphic homogenization is
further extended to nonlinear problems in [39], auxetic tetra-chiral structures in [40] and architected materials
in [41]. Recently, neural network models are trained to replace the expensive RVE computations for an efficient
micromorphic homogenization in [42].
Currently, research on multiscale modeling of metal foams mainly focuses on first-order methods, whereas
research on higher order theories is relatively lacking. Meanwhile, in spite of some recent attempt [43], the
predictive capabilities of computational micromorphic homogenization have not been sufficiently explored
yet for three-dimensional problems. The computational micromorphic framework by [1] only requires the
basic field quantities to be C0-continuous and describes the microscopic average deformation by introducing
an additional macroscopic kinematic field. It characterizes nonlocal interactions between micro-mechanisms
within the RVE and thus enables a precise prediction of size effects, and also avoids the deficiencies of existing
second-gradient schemes, i.e., a dependency on the choice of RVE or an over-constrained fluctuation field, as
reported in the literature [1,39]. For those reasons, this paper aims to extend this framework by [1] toward
A three-dimensional computational multiscale micromorphic

Fig. 1 A two-scale problem in a Cartesian coordinate system

three-dimensional multiscale micromorphic analysis of metal foams, providing a reference for engineering
applications.
This paper is organized as follows. Firstly, we briefly introduce the theoretical basis of the micromorphic
multiscale framework, including kinematic relations, equilibrium equations and constitutive relations of linear
elasticity. Then, we describe the details of a three-dimensional numerical implementation in MATLAB. Finally,
several numerical examples are given to illustrate the effectiveness of the multiscale micromorphic framework
in the analysis of aluminum foams.

2 Theoretical basis

Porous materials have extremely complex microstructures, leading to extremely high modeling and computa-
tional costs with the DNS method. For this kind of problems, multiscale method is a very effective solution. The
first-order homogenization theory is limited to the problem that the macroscopic characteristic length is much
larger than the microscopic characteristic length. To overcome this limitation, in the following, we introduce
a micromorphic multiscale theory proposed by [1], describing the deformation of underlying microstructures
by using an additional kinematic field.

2.1 Multiscale problem and kinematic field decomposition

As shown in Fig. 1, following the work [1], the kinematic variables are divided into two parts: the macroscopic
slow field and the microscopic fast field, using a two-scale coordinate system. Here, x represents coordinates
on the macroscale, while y represents coordinates on the microscale. The microscopic displacement field
û i (x, y) is transformed into:
1
û i (x, y) = H̃i j (x)y j + û i∗ (x, y) + {∇k H̃i j (x) + ∇k H̃ ji (x)}y j yk . (1)
4
Here, H̃i j (x) represents the macroscopic slow field, while û i∗ (x, y) represents the rapidly fluctuating field.
Taking partial derivative of Eq. (1) with respect to y renders a microscopic strain field as
1
Ĥi j (x, y) = û i, j (x, y) = H̃i j (x) + û i,∗ j (x, y) + {∇k H̃i j (x) + ∇k H̃ ji (x) + ∇ j H̃ik (x) + ∇ j H̃ki (x)}yk .
4
(2)
Two macroscopic kinematic variables are obtained by imposing constraints on the microscopic kinematic field:

1
Hi j (x) = Ĥi j (x, y)d V̂ , (3a)
V̂ V̂
X. Ju et al.


1
H̃i j (x) = Ĥi j (x, y)d V̂ . (3b)
V̂I V̂I

Here, V̂ and V̂I are the volume of the RVE and the volume of the holes, respectively; the integration in Eq.
(3a) is thus performed over the microscale coordinates y. Further, Hi j (x) and H̃i j (x) represent the average
deformation of the RVE and the average deformation of the holes, respectively. We refer to H̃i j (x) as the micro-
scopic displacement gradient. In general, Hi j (x) and H̃i j (x) are asymmetric. Here, Hi j (x) is the macroscopic
displacement gradient, which is defined as

Hi j (x) = ∇ j u i (x). (4)

Substituting Eq. (2) into Eq. (3a), the following kinematic constraint equation is obtained:

1
Hi j (x) − H̃i j (x) = û i,∗ j (x, y)d V̂ , (5)
V̂ V̂
for the case
 that the origin of the chosen coordinates system coincides with the centroid of the RVE, where the
relation V̂ yd V̂ = 0 holds. Substituting Eq. (2) into Eq. (3b) similarly yields:

1 1
û i,∗ j (x, y)d V̂ = − {∇k H̃i j (x) + ∇k H̃ ji (x) + ∇ j H̃ik (x) + ∇ j H̃ki (x)} ȳk | I , (6)
V̂I V̂I 4

where ȳk | I denotes the coordinates of the centroid of the holes.

2.2 Homogenization of the microscopic continuum

The Hill–Mandel condition [44] indicates the equivalence of the macroscopic and the microscopic strain power,
i.e.,
 
P(x) = P̂(x, y) , (7)

where P, P̂ and   denote the macroscopic power density, the microscopic power density and the volume
averaging operator, respectively. Combining the kinematic decomposition (2), Eq. (7) further becomes
   1 
 
σ̂i j Ĥ˙ i j d V̂ = σ̂i j H̃˙ i j + σ̂i j û˙ i,∗ j + σ̂i j yk (∇k H̃˙ i j + ∇k H̃˙ ji + ∇ j H̃˙ ik + ∇ j H̃˙ ki )
1
P(x) =
V̂ V̂ 4
= σ Ḣ + ξ H̃ + ζ ∇ H̃˙ ,
ij
˙
ij ij ij i jk k ij (8)

where σ̂i j represents microscopic stress tensor, σ the (symmetric) macroscopic Cauchy stress tensor, ξ (gen-
erally non-symmetric) and ζ the generalized macroscopic stress tensors, respectively.
To identity those macroscopic stresses in form of surface integrals, based on the principle of minimum
potential energy, the total potential energy of the RVE is defined as
  
∗ 1 1 1
 = σ̂i j Ĥi j d V̂ − λi j ( û i n̂ j d Ŝ − Hi j ) − λ̃i j ( û i n̂ Ij d Ŝ − H̃i j ), (9)
2 V̂ V̂ ∂ V̂ V̂I ∂ Iˆ

where λi j and λ̃i j are the Lagrange multipliers associated with the constraints (3a) and (3b), respectively.
Furthermore, ∂ V̂ and n̂ j are the RVE boundary and its outward unit normal vector, respectively, while the hole
boundary and its outward unit normal vector are, respectively, denoted by ∂ Iˆ and n̂ Ij . To fulfill δ∗ = 0, the
Lagrange multipliers have to satisfy

λi j n̂ j = σ̂i j n̂ j , on ∂ V̂ , (10a)
λ̃i j n̂ Ij = (σ̂iIj − σ̂iMj )n̂ Ij , on ∂ Iˆ, (10b)
A three-dimensional computational multiscale micromorphic

where superscripts M and I denote the matrix and the hole domain, respectively. Based on the divergence
theorem and using Eqs. (10b), (8) is rewritten as
   
1 1
P̂(x, y) = σ̂i j n̂ j û˙ i d Ŝ + λ̃i j n̂ Ij û˙ i d Ŝ. (11)
V̂ ∂ V̂ V̂ ∂ Iˆ
Substituting Eq. (1) into Eq. (11) renders
  1   
λ̃ik n̂ kI y j d Ŝ H̃˙ i j +
1
P̂(x, y) = σ̂ik n̂ k y j d Ŝ + (σ̂iq n̂ q y j yk + σ̂ jq n̂ q yi yk )d Ŝ
V̂ ∂ V̂ ∂ Iˆ 4 V̂ ∂ V̂
   (12)
(λ̃iq n̂ qI y j yk + λ̃ jq n̂ qI yi yk )d Ŝ H̃˙ i j,k +
1 1
+ σ̂i j n̂ j û˙ i∗ d Ŝ + λ̃i j n̂ Ij û˙ i∗ d Ŝ.
∂ Iˆ V̂ ∂ V̂ V̂ ∂ Iˆ
Considering Eq. (10a) and the constraints (5) and (6), Eq. (12) further simplifies to
   
λ̃i j H̃˙ i j + (tˆi y j yk + tˆj yi yk )d Ŝ H̃˙ i j,k .
1 V̂I 1
P̂(x, y) = tˆi y j d Ŝ Ḣi j + (13)
V̂ ∂ V̂ V̂ 4 V̂ ∂ V̂
Comparing Eqs. (13) to (8), the macroscopic stress tensors are identified as

1
σi j = tˆi y j d Ŝ, (14a)
V̂ ∂ V̂
V̂I
ξi j = λ̃i j , (14b)
V̂ 
1
ζi jk = (tˆi y j yk + tˆj yi yk )d Ŝ, (14c)
4 V̂ ∂ V̂
respectively.
In the macroscopic domain V , the variation in the total potential energy W I reads
 
δW I = δW d V = (σi j δ Hi j + ξi j δ H̃i j + ζi jk δ∇k H̃i j )d V. (15)
V V

Combined with integral by parts and divergence theorem, Eq. (15) becomes
 
δW I = [−∇ j σi j δu i + (ξi j − ∇k ζi jk )δ H̃i j ]d V + (σi j n j δu i +ζi jk n k δ H̃i j )d S. (16)
V ∂V

In case of ignoring physical strength, the variation in external force work takes the following form

δW E = (ti δu i + m i j δ H̃i j )d S. (17)
∂V

Here, ti and m i j are the surface force and the high-order traction force acting on the Neumann boundary,
respectively. Furthermore, u and H̃ are the corresponding displacement and the deformation of the additional
kinematic field, respectively.
From the principle of virtual work δW I = δW E , we derive the macroscopic equilibrium differential
equation as:
∇ j σi j = 0i , (18a)
ξi j = ∇k ζi jk . (18b)
along with the boundary conditions:
ti = σi j n j , (19a)
m i j = ζi jk n k . (19b)
Clearly, the above equations are in perfect agreement with the conventional micromorphic theory.
X. Ju et al.

2.3 Scale transition for linear elastic micromorphic continua

In a numerical implementation, one could directly use Eq. (14) to compute the macroscopic stresses, as given in
Eq. (B.13) in “Appendix B.” However, for the linear problems under consideration, we seek for an alternative
way, where the RVE computation only needs to be done once and for all.
Due to linearity of the considered problems, the macroscopic constitutive relation of the linear elastic
micromorphic multiscale frame takes the form:

σi j = Ci1jkl Hkl + Ci2jkl H̃kl + E i1jklm ∇m H̃kl , (20a)


ξi j = Ci3jkl Hkl + Ci4jkl H̃kl + E i2jklm ∇m H̃kl , (20b)
ζi jk = E i3jklm Hlm + E i4jklm H̃lm + Di jklmn ∇n H̃lm , (20c)

similar to a classical micromorphic theory [33]. Here, C m , E n for m, n = 1, 2, 3, 4 and D are elasticity tensors
of order 4, order 5 and order 6, respectively. Note that the elasticity tensors E n vanish for a centrosymmetric
RVE. Their components can be extracted by several microscopic RVE computations. The following two elastic
tensors have symmetries as
Cimjkl = Ckli
m
j, (21a)
Di jklmn = Dlmni jk . (21b)
The elasticity tensors are determined numerically in a multiscale context as described in Appendix B. The
numerical approach starts from Eq. (14) to derive the expressions of the elasticity tensors in terms of a finite
element formulation.

3 A three-dimensional computational framework

The overall computational framework illustrated in Fig. 2 is implemented in MATLAB. The framework is
rather similar to the two-dimensional one in [1]. In the following, a three-dimensional extension is presented
in detail.

3.1 Periodic boundary condition of the RVE

In this section, we consider a three-dimensional RVE as illustrated in Fig. 3. Based on the divergence theorem,
Eq. (5) reduces to

1
û i∗ n̂ j d Ŝ = Hi j (x) − H̃i j (x), (22)
V̂ ∂ V̂
where ∂ V̂ and n̂ j denote the external boundary of the RVE and the outward normal vector, respectively. To
fulfill the constraint (22) with a generalized periodic boundary condition, the displacement field at analogous
points on the opposite surfaces of the RVE boundary has to fulfill the following relations:
1
û iF − û iBa = Hi j (y Fj − y Baj ) + (∇k H̃i j + ∇k H̃ ji )(y j − y j )(yk − yk ),
F Ba F Ba
(23a)
4
1
û iT − û iBo = Hi j (y Tj − y Boj ) + (∇k H̃i j + ∇k H̃ ji )(y j − y j )(yk − yk ),
T Bo T Bo
(23b)
4
1
û iR − û iL = Hi j (y Rj − y Lj ) + (∇k H̃i j + ∇k H̃ ji )(y Rj − y Lj )(ykR − ykL ). (23c)
4
Here, the superscripts F, Ba, T, Bo, L and R represent the front, the back, the top, the bottom, the left and
the right surface of the RVE boundary, respectively. Similarly, the displacement field at analogous points of
opposite edges fulfills the following relations:
1
û iL12 − û iL43 = Hi j (y L12
j − y L43
j ) + (∇k H̃i j + ∇k H̃ ji )(y j
L12
− y L43
j )(yk
L12
− ykL43 ), (24a)
4
A three-dimensional computational multiscale micromorphic

Fig. 2 A three-dimensional computational framework for multiscale micromorphic analysis of porous materials

1
û iL12 − û iL56 = Hi j (y L12
j − y L56
j ) + (∇k H̃i j + ∇k H̃ ji )(y L12
j − y L56
j )(yk
L12
− ykL56 ), (24b)
4
1
û i − û i = Hi j (y j − y j ) + (∇k H̃i j
L12 L87 L12 L87
+ ∇k H̃ ji )(y L12
j − y L87
j )(yk
L12
− ykL87 ), (24c)
4
1
û i − û i = Hi j (y j − y j ) + (∇k H̃i j
L23 L14 L23 L14
+ ∇k H̃ ji )(y L23
j − y L14
j )(yk
L23
− ykL14 ), (24d)
4
1
û iL23 − û iL58 = Hi j (y L23
j − y L58
j ) + (∇k H̃i j + ∇k H̃ ji )(y L23
j − y L58
j )(yk
L23
− ykL58 ), (24e)
4
1
û i − û i = Hi j (y j − y j ) + (∇k H̃i j
L23 L67 L23 L67
+ ∇k H̃ ji )(y L23
j − y L67
j )(yk
L23
− ykL67 ), (24f)
4
1
û i − û i = Hi j (y j − y j ) + (∇k H̃i j
L26 L15 L26 L15
+ ∇k H̃ ji )(y L26
j − y L15
j )(yk
L26
− ykL15 ), (24g)
4
1
û i − û i = Hi j (y j − y j ) + (∇k H̃i j
L26 L48 L26 L48
+ ∇k H̃ ji )(y L26
j − y L48
j )(yk
L26
− ykL48 ), (24h)
4
1
û iL26 − û iL37 = Hi j (y L26
j − y L37
j ) + (∇k H̃i j + ∇k H̃ ji )(y L26
j − y L37
j )(yk
L26
− ykL37 ). (24i)
4
X. Ju et al.

Fig. 3 Illustration of an RVE

Here, the twelve edges of the RVE in Fig. 3 are named after corner nodes. For example, node 1 and node 2
constitute the edge L12.
To enforce the above constraints, the displacements at the eight corners of the RVE are prescribed as
(c) (c) 1 (c) (c)
û i = Hi j y j + {∇k H̃i j + ∇k H̃ ji }y j yk , (25)
4
for c = 1, 2, · · · , 8. By means of interpolation of corresponding corner displacements, Eqs. (23) and (24) can
be further simplified as shown in Appendix A.

3.2 Imposing kinematic constraints on hole boundaries

For the hole boundary within the RVE, the following constraint is obtained by applying the divergence theorem
to Eq. (3b):

1
û i n̂ Ij d S = H̃i j , (26)
V̂I ∂ Iˆ
where ∂ Iˆ and n̂ Ij represent the boundary of the holes and its outward normal vector, respectively. In practice,
a discretization of the surface integral in Eq. (26) is required, see, e.g., [45] for a similar context. Therefore,
as illustrated in Fig. 12, a continuous area vector n I d S is converted into a discrete area vector Aq at node q.
For tetrahedral elements in three-dimensional, the nodal area vector is defined as

Ne
1 −→ −−→ −→ −−→
Aq = Oq − Oai × Oq − Obi . (27)
6
i=1

Here, Ne is the number of element surfaces sharing with the node q, while ai and bi represent the other two
nodes of the i-th element surface. Note that ai and bi have to be correctly ordered to ensure an outward normal
vector. Accordingly, Eq. (26) is discretized as
N
1
û q ⊗ Aq = H̃, (28)
V̂I k=1
A three-dimensional computational multiscale micromorphic

Fig. 4 Discretization of a surface integral on a hole boundary

where N represents the number of all boundary nodes related to the hole. By defining a (9 × 3)-dimensional
matrix as
⎡ ⎤
A1 0 0
⎢ 0 A2 0 ⎥
⎢0 0 A ⎥
⎢ 3⎥
⎢ A2 0 0 ⎥
⎢ ⎥
W =⎢ ⎢ 0 A1 0 ⎥ ⎥ , (29)
q
⎢ A3 0 0 ⎥
⎢ ⎥
⎢ 0 0 A1 ⎥
⎣ ⎦
0 A3 0
0 0 A2 q

where A1 , A2 and A3 denote the components of the nodal area vector Aq , Eq. (28) is rewritten in a matrix
form as
N
1
W û q = H̃ := [ H̃11 H̃22 H̃33 H̃12 H̃21 H̃13 H̃31 H̃23 H̃32 ]T . (30)
V̂I q
q=1

3.3 RVE computations

In a multiscale analysis, the main purposes of RVE computations are: 1. Compute the elasticity tensors C m , E n
( m, n = 1, 2, 3, 4 ) and D for the considered RVE (Their matrix counterparts are explicitly given in Appendix
B and computed similarly to the work [1]); 2. Compute macroscopic stress σ and macroscopic generalized
stresses ξ and ζ . Those elasticity matrices can be derived from the microscopic RVE stiffness matrix by the
condensation method [1].
X. Ju et al.

For linear elastic problems considered in this paper, those elasticity matrices are constant. Hence in practice,
they only need to be computed once and for all. Subsequently, the macroscopic stresses can be computed
following the constitutive relations (20). In this way, there is no need to perform separate RVE computations
for each material point of the macro model, thereby greatly improving the numerical efficiency of the multiscale
analysis.

3.4 Micromorphic finite elements on the macroscale

On the macroscopic level, the displacement field u and the microscopic displacement gradient field H̃ are
discretized in form of nodal interpolation:
u = N q, (31a)
H̃ = N q H̃ . (31b)

Here, N and N are the corresponding shape function matrices, respectively. Moreover, q and q H̃ are the

macroscopic and the microscopic nodal displacements, respectively. For three-dimensional problems, each
node includes three degrees of freedom related to displacements u and nine additional degrees of freedom
related to micro-deformations H̃ . Their corresponding gradients are discretized, respectively, as:
H = B q, (32a)
∇ H̃ = B H̃ q H̃ . (32b)

Here, B and B H̃ are the corresponding strain matrices, respectively.


Substituting Eqs. (31), (32) and the constitutive relations (20) into the above equations, the element stiffness
matrix is obtained as
    
K 11 K 12 q F1
= , (33)
K 21 K 22 q H̃ F2

where

K 11 = B T C 1 Bd V, (34a)
 
K 12 = BT C2 N dV + B T E 1 B H̃ d V , (34b)

 
K 21 = N T
C 3 Bd V + B H̃ T E 3 Bd V , (34c)

   
K 22 = N T C4 N dV + N T E 2 B H̃ d V + B H̃ T E 4 N d V + B H̃ T D B H̃ d V , (34d)
H̃ H̃ H̃ H̃
 
F1 = N T td S− BT σ dV , (34e)
  
F2 = T
N md S− N ξdV −
T
B H̃ T ζ d V . (34f)
H̃ H̃

The corresponding elasticity matrices are determined by Eq. (B.14) given in “Appendix B.”

4 Numerical examples

4.1 Example 1: Uniaxial tensile test of a perforated aluminum foam sheet

In this section, we first study the mechanical responses of a perforated closed-cell aluminum foam under uniaxial
tensile loading. The matrix material is aluminum, whose elasticity properties are given with a Poisson’s ratio
of 0.33 and a Young’s modulus of 70, 000MPa. As shown in Fig. 5a, the perforated rectangular metal foam
A three-dimensional computational multiscale micromorphic

Fig. 5 Example 1: Uniaxial tensile test of a perforated metal foam sheet

sheet has a periodic microstructure, where the unit cells (10mm in length) are periodically repeated with a
radius of 2.5mm for the spherical hole. This model represents a simplified model of aluminum foam. A uniform
displacement of u 1 = 0.04L is imposed to the right end of the sheet, while the left end is fixed. The DNS model
and the corresponding multiscale finite element model of the aluminum foam sheet (thickness L = 10 mm)
are shown in Fig. 5b, c, respectively. For multiscale finite element models, we further differ the micromorphic
procedure from the first-order one.
The contour plots of von Mises stresses (as usually defined in a conventional continuum theory in relation
to the macroscopic Cauchy tensor σ ) at point Q in Fig. 5a are depicted in Fig. 6 for different methods. Table 1
gives a comparison of average strains obtained from DNS, micromorphic and first-order multiscale analysis.
Taking the DNS result as a reference, the micromorphic prediction on the average deformation H̃11 of the
hole is much more accurate that the first-order one, while that on the average deformation H11 of the RVE is
slightly underestimated compared to the first-order prediction. This is quite reasonable, since in the multiscale
micromorphic theory the average deformation of the hole is treated as additional degrees of freedom, con-
tributing to the macroscopic equilibrium. As shown in Fig. 6, regarding the stress distributions, the prediction
quantity of the micromorphic procedure is much better than that of the first-order one, when compared to the
X. Ju et al.

Table 1 Example 1: Average deformation of the RVE at point Q

Average deformation DNS Micromorphic First-order


H11 3.044e−3 2.567e−3 2.603e−3
H̃11 5.156e−3 4.222e−3 4.150e−3

Fig. 6 Example 1: Contour plots of von Mises stresses computed by different methods

Fig. 7 Example 1: Von Mises stresses in case of scale separation between micromorphic and first-order multiscale models

DNS result in Fig. 6c. Compared with the first-order procedure, the micromorphic multiscale analysis provides
richer microscopic information (deformation of the holes) for the macroscopic model.
Now, we additionally study the case of scale separation. To this end, the size of the RVE for both multiscale
models (micromorphic and first-order) is 100-times downscaled, while the size of the macromodel remains
unchanged for a clear scale separation. Figure 7 shows the contour plots of von Mises stresses at point Q
in Fig. 5a predicted by the micromorphic and first-order multiscale method. Compared to the results in Fig. 6
without a scale separation, both stress distributions are rather close to each other in this case. This indicates that
the first-order multiscale theory is recovered by the micromorphic theory in case of a clear scale separation.
Moreover, we also conducted an investigation on a homogeneous RVE and found that the micromorphic
predictions become the same as the first-order ones. This indicates that the physical behavior of the microscopic
constituent is recovered for homogeneous cases in frame of the present micromorphic theory.
As a further investigation, we consider two different RVEs in Fig. 8 for the micromorphic procedure. As
shown in Fig. 9, the distributions of the macroscopic von Mises stresses are almost identical for both RVEs,
thus indicating that the micromorphic prediction is independent of the choice of the RVE. This is the same
conclusion as drawn in [1] for two-dimensional cases.
Furthermore, the computational efficiency of the above-mentioned three methods is comparatively studied.
On a computer with a CPU of Intel(R) Core(TM) i5-7300HQ, the computations took 301.1s, 17.6s and 8.3s
for the DNS simulation, multiscale micromorphic analysis and first-order multiscale analysis, respectively.
The CPU time of the micromorphic model is slightly longer than that of the first-order model, since the
micromorphic model has additional degrees of freedom. Compared with the DNS, both multiscale models
show a much higher computational efficiency.
A three-dimensional computational multiscale micromorphic

Fig. 8 Example 1: Different choices on the RVE

Fig. 9 Example 1: Contour plots of macroscopic von Mises stresses computed by the micromorphic approach with different
RVEs and the DNS

4.2 Example 2: Shear test of an aluminum foam sheet

Now, we further investigate a shear loading case of the aluminum foam sheet. As illustrated in Fig. 10, the
sheet is 160mm long, 20mm high and 10mm thick, while the radius of the circular hole is 4mm. Vertical
displacements are prescribed to the upper and the lower boundary of the sheet. More specifically, there hold

u 1 = 0, (35a)
 
2π x1
u 2 = A sin , (35b)
λ
u 3 = 0. (35c)

with a wavelength of λ = 80mm and an amplitude of A = −0.1mm.


The settings of the microscopic RVE models and material properties in this example are consistent with
those in Example 1. The computational results of the multiscale micromorphic model are compared with the
DNS and the first-order multiscale results. Figure 11 shows the variation in the macroscopic shear stress σ12
(in relation to the macroscopic Cauchy stress tensor σ ) for different x1 positions in an average meaning. Here,
we consider σ12 as the macroscopic shear stress averaged over the two unit cells for a same x1 position. In
case of weak scale separation, the prediction error of the first-order multiscale method is larger than that of the
micromorphic multiscale method. A relative error of 10.32% is reached for the first-order multiscale method,
when x1 = 80mm. However, the micromorphic multiscale method maintains a high prediction accuracy, with
an error of 0.54% at x1 = 80mm.

4.3 Example 3: Bending of an aluminum foam

This example considers a bending case of the aluminum foam sheet. The material parameters are the same
as Example 1. The geometry and boundary conditions for this example are illustrated in Fig. 12a, where the
radius of the spherical hole reads 2.5mm. The left end of the foam metal plate model is fixed in the x1 -direction
at the left end face and free in the remaining two directions, with a fully fixed corner node to avoid rigid
body motion. The prescribed displacements at the right end in x1 -direction linearly vary from 0 to 0.00625mm
in x2 -direction. The DNS model and the multiscale finite element model for the aluminum foam sheet are
illustrated in Fig. 12b, c, respectively.
X. Ju et al.

Fig. 10 Example 2: An aluminum foam sheet under shear loading

Fig. 11 Example 2: Comparison of macroscopic shear stresses σ12 computed by different methods

Figure 13 depicts the contour plots of von Mises stresses predicted by the DNS, the micromorphic and the
first-order multiscale models for the macroscopic positions A and B. The deformations in those contour plots
are 15-times upscaled. For both positions, the stress distributions predicted by the micromorphic approach are
much closer to the DNS results than those by the first-order approach. In view of the deformation, the first-order
approach fails to capture the rotation of the RVE as shown in the DNS results, especially for the location B,
where the rotation becomes obvious. In contrast, the deformation (especially the rotation) is correctly captured
by the micromorphic approach for both positions.

4.4 Example 4: Torsion of a rectangular bar

In this example, we consider a rectangular bar under a displacement-controlled torsion loading in Fig. 14. The
radius of the spherical hole reads 2.5mm. While the material parameters remain the same as Example 1, the
DNS and the multiscale modeling are also similar to above examples. The displacement boundary conditions
are applied at the nodes of the left and the right end, where each nodal displacement vector points along the
tangent direction with a magnitude linear to the distance R from the centroid. Formally, we have u = k · R
with k = 0.02.
Figure 15 depicts the contour plots of von Mises stresses at position M computed by three different methods,
where the deformations are 20-times upscaled for a better illustration. Compared to the first-order approach, the
stress distributions predicted by the micromorphic approach are much closer to the DNS results; especially, the
stress pattern around the hole is correctly captured. The deformation of the RVE computed by the micromorphic
A three-dimensional computational multiscale micromorphic

Fig. 12 Example 3: Bending of an aluminum foam

approach also seems to be more realistic than the first-order one, showing the advantages of the present
micromorphic approach against the first-order one.

5 Conclusions

In this work, we extended the multiscale micromorphic theoretical framework of [1] to three-dimensional
elasticity problems, developed a numerical simulation tool for porous materials and conducted a numerical
study on its effectiveness. The main conclusions are drawn as follows:
X. Ju et al.

Fig. 13 Example 3: Contour plots of von Mises stresses computed by different methods for positions A and B

Fig. 14 Example 4: Torsion of a rectangular bar

Fig. 15 Example 4: Contour plots of von Mises stresses computed by different methods at position M
A three-dimensional computational multiscale micromorphic

• A three-dimensional implementation of the micromorphic multiscale framework has been addressed in


linear elasticity, providing a basis for engineering design of porous materials and further related engineering
applications. A further extension to nonlinear problems is straightforward.
• In order to apply a generalized periodic boundary conditions to three-dimensional RVE models, periodic
constraint equations must be applied to angular nodes, edge nodes and surface nodes.
• The computational efficiency of the micromorphic multiscale method is much higher than that of the DNS,
while its efficiency is slightly lower than that of the first-order multiscale method.
• The RVE models used in this work are rather ideal and simple. An RVE should always be carefully chosen
for an accurate prediction in practice, especially for a real aluminum foam. For a systematic study on
adaptive choice of the RVE size in general, we refer to our previous work [46].
• Recent work of [43] offers novel possibilities to derive explicit expressions of elasticity constants for
multiscale micromorphic constitutive relations. A comparative study with computational approaches is of
large interest for future work.
• Various loading scenarios of aluminum foams were studied for illustration. The results confirmed that the
micromorphic multiscale method can accurately characterize both the macroscopic effective properties and
the microscopic field information of aluminum foams. In future works, this method can be further explored
to study the influence of different effects (like different porosities and micro-structural characteristics) on
the overall performance of the structure.

Acknowledgements The work described in this paper was supported by the Zhejiang Provincial Natural Science Foundation of
China (Grant No: LQ21A020002) and the National Natural Science Foundation of China (Grant Nos. 12002309, 52275164).

Author contribution XJ: Writing—original draft, conceptualization, methodology. KG: Writing—original draft, methodology,
software. JH: data curation, visualization. HR: software, validation. HC: validation. YX: conceptualization, supervision. LL:
conceptualization, supervision.

Declarations

Conflict of interest The authors declare no competing interests.

Appendix A: Technical details for three-dimensional periodic boundary conditions

In the following, several details are given for the three-dimensional periodic boundary condition in Sect. 3.1.
By means of interpolation of corresponding corner displacements in Eq. (25), Eq. (23) becomes

1 (4) (8) 1 (3) (7)


û iF − û iBa = (1 − h)(1 − g)(û i − û i ) + (1 + h)(1 − g)(û i − û i )
4 4
1 (1) (5) 1 (2) (6)
+ (1 − h)(1 + g)(û i − û i ) + (1 + h)(1 + g)(û i − û i ), (A.1a)
4 4
1 1
û iT − û iBo = (1 − h)(1 − k)(û i(5) − û i(8) ) + (1 + h)(1 − k)(û i(6) − û i(7) )
4 4
1 (3) (4) 1 (2) (3)
+ (1 − h)(1 + k)(û i − û i ) + (1 + h)(1 + k)(û i − û i ), (A.1b)
4 4
1 (7) (8) 1 (6) (5)
û iR − û iL = (1 − g)(1 − k)(û i − û i ) + (1 + g)(1 − k)(û i − û i )
4 4
1 (3) (4) 1 (2) (1)
+ (1 − g)(1 + k)(û i − û i ) + (1 + g)(1 + k)(û i − û i ). (A.1c)
4 4

2y 2y 2y
Here, h = l 1 , g = l 2 and k = l 3 represent the normalized coordinates of the analogous node pairs at the
front–back, top–bottom and left–right boundary pairs, respectively. Similarly, by means of linear interpolation
of corresponding corner displacements, Eq. (24) becomes
X. Ju et al.

1 (1) (4) 1 (2) (3)


(1 − h)(û i − û i ) + (1 + h)(û i − û i ),
û iL12 − û iL43 = (A.2a)
2 2
1 (1) (5) 1 (2) (6)
û iL12 − û iL56 = (1 − h)(û i − û i ) + (1 + h)(û i − û i ), (A.2b)
2 2
1 (1) (8) 1 (2) (7)
û iL12 − û iL87 = (1 − h)(û i − û i ) + (1 + h)(û i − û i ), (A.2c)
2 2
1 1
û iL23 − û iL14 = (1 − g)(û i(3) − û i(4) ) + (1 + g)(û i(2) − û i(1) ), (A.2d)
2 2
1 (3) (8) 1 (2) (5)
û iL23 − û iL58 = (1 − g)(û i − û i ) + (1 + g)(û i − û i ), (A.2e)
2 2
1 (3) (7) 1 (2) (6)
û iL23 − û iL67 = (1 − g)(û i − û i ) + (1 + g)(û i − û i ), (A.2f)
2 2
1 (6) (5) 1 (2) (1)
û iL26 − û iL15 = (1 − k)(û i − û i ) + (1 + k)(û i − û i ), (A.2g)
2 2
1 (6) (8) 1 (2) (4)
û iL26 − û iL48 = (1 − k)(û i − û i ) + (1 + k)(û i − û i ), (A.2h)
2 2
1 (6) (7) 1 (2) (3)
û i − û i = (1 − k)(û i − û i ) + (1 + k)(û i − û i ).
L26 L37
(A.2i)
2 2
which are similar to the formulas in a two-dimensional implementation in [1].

Appendix B: Numerical determination of the elasticity tensors

The elasticity tensors C m , E n and D are computed similarly to the work [1]. For our three-dimensional
implementation, the macroscopic elasticity operator is obtained by condensing the microscopic finite element
equations. In the microscale, the finite element equation system reads
K̂ û = r̂ , (B.1)

where K̂ is the global stiffness matrix of the RVE and r̂ represents the residual force. Then, the global stiffness
matrix is partitioned as follows
    
K̂ ii K̂ id û i r̂
= i , (B.2)
K̂ K̂ û d r̂ d
di dd
where the subscript d denotes the dependent degrees of freedom involved in the periodic boundary conditions
(23) and (24). There holds û d = Ĉ û i for the independent degrees of freedom û i , with the correlation matrix
Ĉ. By means of a static condensation [22], Eq. (B.1) reduces to

K̂ ∗ û i = r̂ ∗ , (B.3)
where
T T
K̂ ∗ = K̂ ii + K̂ id Ĉ + Ĉ K̂ di + Ĉ K̂ dd Ĉ, (B.4a)
T
r̂ ∗ i = r̂ i + Ĉ r̂ d . (B.4b)

In order to impose the constraint (30), dummy degrees of freedom with the subscript λ̃ are introduced as
W û I = H̃ = û λ̃ , (B.5)

where W and û I represent the coefficient matrix in Eq. (30) and degrees of freedom along the hole surfaces,
respectively. Next, we rewrite Eq. (B.5) as
     
−1 −1 û a û a
û Id = −Ŵ Ŵ Ŵ = Ĉ 1 Ĉ 2 , (B.6)
2 û1 2 û
λ̃ λ̃
A three-dimensional computational multiscale micromorphic

where û Id and û a denote the dependent degrees of freedom to be eliminated at the hole surface and the
remaining degrees of freedom in Eq. (B.3), respectively. As a result, Eq. (B.3) becomes
⎡ ∗ ⎤⎡ ⎤ ⎡ ⎤
K̂ aa 0 K̂ ∗ a I û a r̂ ∗ a
⎢ 0 I d

0 ⎦ ⎣ û λ̃ ⎦ = ⎣ r ∗ λ̃ ⎦ ,
⎣ (B.7)
K̂ ∗ 0 K̂ ∗ û Id r ∗ Id
Id a Id Id

where I is the identity matrix. On the basis of Eqs. (B.6), (B.7) is further condensed to
    
K̂ ∗∗ aa K̂ ∗∗ a λ̃ û a r̂ ∗∗ a
= ∗∗ , (B.8)
K̂ ∗∗ K̂ ∗∗ û λ̃ r λ̃
λ̃a λ̃λ̃

where
T T
K̂ ∗∗ aa = K̂ ∗ aa + K̂ ∗ a I Ĉ + Ĉ K̂ ∗ I + Ĉ K̂ ∗ I Ĉ , (B.9a)
d 1 1 da 1 d Id 1
T
K̂ ∗∗ a λ̃ = K̂ ∗ a I Ĉ + Ĉ K̂ ∗ I Ĉ , (B.9b)
d 2 1 d Id 2
T T
K̂ ∗∗ λ̃a = Ĉ K̂ ∗ I + Ĉ K̂ ∗ I Ĉ , (B.9c)
2 da 2 d Id 1
T
K̂ ∗∗ λ̃λ̃ = I + Ĉ K̂ ∗ I Ĉ , (B.9d)
2 d Id 2
T ∗
r̂ ∗∗ a = r̂ ∗ a + Ĉ r̂ Id , (B.9e)
1
T ∗
r̂ ∗∗ λ̃ = r̂ ∗ λ̃ + Ĉ r̂ Id . (B.9f)
2

The equation system (B.8) is next partitioned as


⎡ ⎤
K̂ ∗∗ pp K̂ ∗∗ pλ̃ K̂ ∗∗ p f ⎡ û ⎤ ⎡ r̂ ∗∗ ⎤ ⎡ f ⎤
⎢ ∗∗ ⎥ p p
⎢ K̂ K̂ ∗∗ K̂ ∗∗ ⎥ ⎣
û ⎦ = ⎣ r̂ ∗∗ ⎦ ≈ ⎣λ̃ ⎦ , (B.10)
⎣ λ̃ p λ̃λ̃ λ̃ f ⎦ λ̃ λ̃
K̂ ∗∗ K̂ ∗∗ K̂ ∗∗ û f r̂ ∗∗ f 0
fp f λ̃ ff

with the subscripts p and f denoting the corner and the internal free nodes, respectively. By a further conden-
sation [22], Eq. (B.10) finally reduces to
    
K̄ pp K̄ pλ̃ û
p = f . (B.11)
K̄ λ̃ p K̄ λ̃λ̃ û λ̃ λ̃

In the numerical implementation, the displacements of the corner nodes of the RVE are prescribed by Eq. (25)
as
(c) (c) 1 (c) (c)
û i = Hi j y j + {∇k H̃i j + ∇k H̃ ji }y j yk , c = 1, · · · , 8. (B.12)
4
Due to the relation û λ̃ = H̃ in Eq. (B.5), the RVE problem is solved after Eq. (B.11).
Under the periodic boundary conditions (A.1) and (A.2), only corner nodes contribute to Eq. (14). Accordingly,
the macroscopic stress tensors are computed by
8
1 (c) (c)
σi j = fi y j , (B.13a)
V̂ c=1
V̂I
ξi j = λ̃i j , (B.13b)

8
1 (c) (c) (c) (c) (c) (c)
ζi jk = ( f i y j yk + f j yi yk ). (B.13c)
4 V̂ c=1
X. Ju et al.

From Eqs. (B.11) and (B.13), the components for the elasticity tensors C m , E n and D in the macroscopic
constitutive relations (20) are computed by
8 8
1 (ab)
C1 = K̄ pp(i,k) y (a) y (b) ,
( j) (l)
(B.14a)
i jkl V̂ a=1 b=1
8
1 (a λ̃)
C2 = K̄ y (a) , (B.14b)
i jβ V̂ a=1 pλ̃(i,β) ( j)
8
V̂I (λ̃a)
C3 = K̄ y (a) , (B.14c)
α jk V̂ a=1 λ̃ p(α−, j) (k)
V̂I
C4 = K̄ λ̃λ̃(α,β) , (B.14d)
αβ V̂
1
8 8  
(ab) (ab)
E 1 i jklm = K̄ pp(i,k) y (a) y (b) (b)
y
( j) (l) (m)
+ K̄ y (a) (b) (b)
y y
pp(i,l) ( j) (k) (m)
, (B.14e)
4 V̂
a=1 b=1
8  
V̂I (λ̃a) (λ̃a)
E 2 α jkl = K̄ λ̃ p(α, j) y (a) y
(k) (l)
(a)
+ K̄ y (a) (a)
y
λ̃ p(α,k) ( j) (l)
, (B.14f)
4 V̂ a=1
8 8  
1 (ab) (ab)
E 3 i jklm = K̄ pp(i,l) y (a) y (a) (b)
y
( j) (k) (m)
+ K̄ y (a) (a) (b)
y y
pp( j,l) (i) (k) (m)
, (B.14g)
4 V̂ a=1 b=1
8  
1 (a λ̃) (a λ̃)
E 4 i jkβ = K̄ pλ̃(i,β) y (a) y (a)
( j) (k)
+ K̄ y (a) (a)
y
p λ̃( j,β) (i) (k)
, (B.14h)
4 V̂ a=1
8 8 
1 (ab) (ab)
D i jklmn = K̄ pp(i,l) y (a) y (a) y (b) y (b) + K̄ pp(i,m) y (a)
( j) (k) (m) (n)
y (a) y (b) y (b)
( j) (k) (l) (n)
16V̂ a=1 b=1

(ab) (ab)
+ K̄ pp( j,l) y (a) y (a) (b) (b)
y
(i) (k) (m) (n)
y + K̄ y (a) (a) (b) (b)
y
pp( j,m) (i) (k) (l) (n)
y y . (B.14i)

(ab)
with i, j, k, l, m, n = 1, 2, 3 and α, β = 1, 2, · · · , 9. Taking K̄ pp(i,l) as an example for notation, the super-
scripts a and b denote the partition in the matrix K̄ pp , whereas the subscripts i and l are indices for components.

References

1. Biswas, R., Poh, L.H.: A micromorphic computational homogenization framework for heterogeneous materials. J. Mech.
Phys. Solids 102, 187–208 (2017)
2. Zhang, S.: Research and application progress of foam metal. Powder Metall. Technol. 34(3), 6 (2016)
3. Wang, Z., Cao, X., Ma, H.: Effect of microstructure on meso-mechanical properties of porous materials. J. Taiyuan Univ.
Technol. 37(1), 1–4 (2006)
4. Hill, R.: Elastic properties of reinforced solids: some theoretical principles. J. Mech. Phys. Solids 11, 357–372 (1963)
5. Feyel, F., Chaboche, J.L.: Fe2 multiscale approach for modelling the elastoviscoplastic behaviour of long fibre SiC/Ti
composite materials. Comput. Methods Appl. Mech. Eng. 183(3–4), 309–330 (2000)
6. Moulinec, H., Suquet, P.: A numerical method for computing the overall response of nonlinear composites with complex
microstructure. Comput. Methods Appl. Mech. Eng. 157(1–2), 69–94 (1998)
7. Matouš K., Geers, M. G. D., Kouznetsova, V. G., Yvonnet, J.: Homogenization methods and multiscale modeling: nonlinear
problems. Encyclopedia of computational mechanics second edition, pp. 1–34 (2017)
8. Liang, L., Ju, X., Mahnken, R., Xu, Y., Zhou, W.: A nonuniform transformation field analysis for composites with strength
difference effects in elastoplasticity. Int. J. Solids Struct. 228, 111103 (2021)
9. Xu, Y., Ju, X., Mahnken, R., Liang, L.: NTFA-enabled goal-oriented adaptive space-time finite elements for micro-
heterogeneous elastoplasticity problems. Comput. Methods Appl. Mech. Eng. 398, 115199 (2022)
10. Dvorak, G.J., Benvensite, Y.: On transformation strains and uniform fields in multiphase elastic media. Proc. R. Soc. Lond.
437(A), 291–310 (1992)
11. Michel, J.C., Suquet, P.: Nonuniform transformation field analysis. Int. J. Solids Struct. 40, 6937–6955 (2003)
A three-dimensional computational multiscale micromorphic

12. Michel, J.C., Suquet, P.: A model-reduction approach in micromechanics of materials preserving the variational structure of
constitutive relations. J. Mech. Phys. Solids 90, 254–285 (2016)
13. Bessa, M.A., Liu, Z., Liu, W.K.: Self-consistent clustering analysis: an efficient multi-scale scheme for inelastic heterogeneous
materials. Comput. Methods Appl. Mech. Eng. 306, 319–341 (2016)
14. Gibson, L.J.: Cellular solids. MRS Bull. 28(4), 270–274 (2003)
15. Feng, M., Zhuang, S., Wu, C., Yuan, Z.: Numerical study of micromechanical properties of porous materials based on
homogenization method. J. Mater. Sci. Eng. 19(4), 5 (2001)
16. Fu, C., Liu, P., Li, T.: Tensile strength of metal materials with high porosity. Rare Metal Mater. Eng. 29(002), 94–100 (2000)
17. Zhao, S., Wang, X., Xu, Y., Wu, P., Liang, L.: Realization of elastic-plastic multiscale analysis and its application in particle
reinforced composites. Compos. Mater. Sci. Technol. 34(9), 10 (2017)
18. Onck, P., Andrews, E.W., Gioux, G., Gibson, L.J.: Size effects in ductile cellular solids. Part II: experimental results. Int. J.
Mech. Sci. 43(3), 701–713 (2001)
19. Geers, M.G. D., Kouznetsova, V.G., A. M, W.: Multi-scale second-order computational homogenization of multi-phase
materials: a nested finite element solution strategy. Comput. Methods Appl. Mech. Eng. 193(48–51), 5525–5550 (2004)
20. Becker, G., Nguyen, V.D., Noels, L.: Multiscale computational homogenization methods with a gradient enhanced scheme
based on the discontinuous galerkin formulation. Comput. Methods Appl. Mech. Eng. 260, 63–77 (2013)
21. Nguyen, V.D., Noels, L.: Computational homogenization of cellular materials. Int. J. Solids Struct. 51(11), 2183–2203 (2014)
22. Geers, M.G.D., Kouznetsova, V.G., Brekelmans, W.A.M.: Multi-scale second-order computational homogenization of multi-
phase materials: a nested finite element solution strategy. Comput. Methods Appl. Mech. Eng. 193(48), 5525–5550 (2004).
(Advances in Computational Plasticity)
23. Geers, M.G.D., Kouznetsova, V.V., Brekelmans, W.A.M.: Multi-scale constitutive modelling of heterogeneous materials with
a gradient-enhanced computational homogenization scheme. Int. J. Numer. Methods Eng. 54, 1235–1260 (2002)
24. Eringen, A. C.: . Mechanics of micromorphic materials. In: Applied Mechanics, pp. 131–138. Springer (1966)
25. Mindlin, R.D.: Micro-structure in linear elasticity. Arch. Ration. Mech. Anal. 16, 51–78 (1964)
26. Liu, Q., Li, X., Zhang, J.: A micro-macro homogenization approach for discrete particle assembly: Cosserat continuum
modeling of granular materials. Int. J. Solids Struct. 47(2), 291–303 (2010)
27. Adomeit, G.: Determination of elastic constants of a structured material. In: Kröner, E. (ed.) Mechanics of Generalized
Continua, pp. 80–82. Springer, Berlin (1968)
28. Forest, S.: Mechanics of generalized continua: construction by homogenizaton. Le Journal de Physique IV 8(PR4), Pr4-39
(1998)
29. Bigoni, D., Drugan, W.J.: Analytical derivation of Cosserat Moduli via homogenization of heterogeneous elastic materials.
J. Appl. Mech. 74(4), 741–753 (2006). (04)
30. Feyel, F.: A multilevel finite element method (fe2) to describe the response of highly non-linear structures using generalized
continua. Comput. Methods Appl. Mech. Eng. 192(28/30), 3233–3244 (2003)
31. Hütter, G.: On the micro-macro relation for the microdeformation in the homogenization towards micromorphic and microp-
olar continua. J. Mech. Phys. Solids 127, 62–79 (2019)
32. Jänicke, R., Steeb, H.: Wave propagation in periodic microstructures by homogenisation of extended continua. Comput.
Mater. Sci. 52(1), 209–211 (2012)
33. Ju, X., Mahnken, R.: Goal-oriented adaptivity for linear elastic micromorphic continua based on primal and adjoint consis-
tency analysis. Int. J. Numer. Meth. Eng. 112(8), 1017–1039 (2017)
34. Liang, L., Ju, X., Mahnken, R., Xu, Y.: Goal-oriented mesh adaptivity for inverse problems in linear micromorphic elasticity.
Comput. Struct. 257, 106671 (2021)
35. Xu, Y., Ju, X., Mahnken, R., Liang, L.: Goal-oriented error estimation and h-adaptive finite elements for hyperelastic
micromorphic continua. Comput. Mech. 69, 847–863 (2022)
36. Ju, X., Mahnken, R.: Goal-oriented h-type adaptive finite elements for micromorphic elastoplasticity. Comput. Methods
Appl. Mech. Eng. 351, 297–329 (2019)
37. Hütter, G.: Homogenization of a Cauchy continuum towards a micromorphic continuum. J. Mech. Phys. Solids 99, 394–408
(2017)
38. Rokoš, O., Ameen, M.M., Peerlings, R.H.J., Geers, M.G.D.: Micromorphic computational homogenization for mechanical
metamaterials with patterning fluctuation fields. J. Mech. Phys. Solids 123, 119–137 (2019)
39. Shedbale, A.S., Biswas, R., Poh, L.H.: Nonlinear analyses with a micromorphic computational homogenization framework
for composite materials. Comput. Methods Appl. Mech. Eng. 350, 362–395 (2019)
40. Poh, L.H., Biswas, R., Shedbale, A.S.: A micromorphic computational homogenization framework for auxetic tetra-chiral
structures. J. Mech. Phys. Solids 135, 103801 (2020)
41. Ganghoffer, J.-F., Wazne, A., Reda, H.: Frontiers in homogenization methods towards generalized continua for architected
materials. Mech. Res. Commun. 130, 104114 (2023)
42. Biswas, R., Yuan, Z., Poh, L.H.: Accelerated offline setup of homogenized microscopic model for multi-scale analyses using
neural network with knowledge transfer. Int. J. Numer. Meth. Eng. 124(13), 3063–3086 (2023)
43. Hütter, G.: Interpretation of micromorphic constitutive relations for porous materials at the microscale via harmonic decom-
position. J. Mech. Phys. Solids 171, 105135 (2023)
44. Hill, R.: Elastic properties of reinforced solids: some theoretical principles. J. Mech. Phys. Solids 11(5), 357–372 (1963)
45. Miehe, C.: Computational micro-to-macro transitions for discretized micro-structures of heterogeneous materials at finite
strains based on the minimization of averaged incremental energy. Comput. Methods Appl. Mech. Eng. 192(5–6), 559–591
(2003)
46. Mahnken, R., Ju, X.: Goal-oriented adaptivity based on a model hierarchy of mean-field and full-field homogenization
methods in linear elasticity. Int. J. Numer. Meth. Eng. 121, 277–307 (2020)
X. Ju et al.

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

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

You might also like