ArticlePDF Available

Optimal design of functionally graded lattice structures using Hencky bar-grid model and topology optimization

Authors:

Abstract and Figures

Presented herein is a novel design framework for obtaining the optimal design of functionally graded lattice (FGL) structures that involve using a physical discrete structural model called the Hencky bar-grid model (HBM) and topology optimization (TO). The continuous FGL structure is discretized by HBM comprising rigid bars, frictionless hinges, frictionless pulleys, elastic primary and secondary axial springs, and torsional springs. A penalty function is introduced to each of the HBM spring’s stiffnesses to model non-uniform material properties. The gradient-based TO method is applied to find the stiffest structure via minimizing the compliance or elastic strain energy by adjusting the HBM spring stiffnesses subjected to prescribed design constraints. The optimal design of FGL structures is constructed based on the optimal spring stiffnesses of the HBM. The proposed design framework is simple to implement and for obtaining optimal FGL structures as it involves a relatively small number of design variables such as the spring stiffnesses of each grid cell. As illustration of the HBM-TO method, some optimization problems of FGL structures are considered and their optimal solutions obtained. The solutions are shown to converge after a small number of iterations. A Python code is given in the Appendix for interested readers who wish to reproduce the results.
This content is subject to copyright. Terms and conditions apply.
Vol.:(0123456789)
1 3
Structural and Multidisciplinary Optimization (2022) 65:276
https://doi.org/10.1007/s00158-022-03368-w
RESEARCH PAPER
Optimal design offunctionally graded lattice structures using Hencky
bar‑grid model andtopology optimization
Y.P.Zhang1 · C.M.Wang1· N.Challamel2· Y.M.Xie3· J.Yang3
Received: 19 February 2022 / Revised: 10 August 2022 / Accepted: 15 August 2022 / Published online: 16 September 2022
© The Author(s) 2022
Abstract
Presented herein is a novel design framework for obtaining the optimal design of functionally graded lattice (FGL) structures
that involve using a physical discrete structural model called the Hencky bar-grid model (HBM) and topology optimization
(TO). The continuous FGL structure is discretized by HBM comprising rigid bars, frictionless hinges, frictionless pulleys,
elastic primary and secondary axial springs, and torsional springs. A penalty function is introduced to each of the HBM
spring’s stiffnesses to model non-uniform material properties. The gradient-based TO method is applied to find the stiffest
structure via minimizing the compliance or elastic strain energy by adjusting the HBM spring stiffnesses subjected to pre-
scribed design constraints. The optimal design of FGL structures is constructed based on the optimal spring stiffnesses of
the HBM. The proposed design framework is simple to implement and for obtaining optimal FGL structures as it involves
a relatively small number of design variables such as the spring stiffnesses of each grid cell. As illustration of the HBM-TO
method, some optimization problems of FGL structures are considered and their optimal solutions obtained. The solutions
are shown to converge after a small number of iterations. A Python code is given in the Appendix for interested readers who
wish to reproduce the results.
Keywords Lattice structures· Functional grading· Hencky bar-grid model· Topology optimization
Abbreviations
FGL Functionally graded lattice
HBM Hencky bar-grid model
TO Topology optimization
GDM Greyscale density mapping
SIMP Solid isotropic material with penalization
GSO Ground structure optimization
OC Optimality criteria
FD Finite difference
1 Introduction
Additive manufacturing opens up new opportunities to fab-
ricate as-designed functionally graded lattice (FGL) struc-
tures by building up material layer-by-layer (Maconachie
etal. 2019). Lattice structures (also referred to as cellular
structures) have a clear network feature comprising basic
elements like solid struts or plates (Gibson 1989). Properly
designed FGL structures can achieve high performance
with great control in stiffness and strength, weight, energy
absorption, heat exchange and acoustic insulation (Zhang
etal. 2015; Ferro etal. 2022). Generally, elastic structures
with minimum compliance or minimum elastic strain energy
are the stiffest (Hassani and Hinton 1999; Kaveh etal. 2008;
Zhang etal. 2014). Topology optimization (TO) method is
commonly adopted to minimize the compliance or elastic
strain energy of FGL structures by iteratively updating the
material distribution subjected to loading, boundary condi-
tions and prescribed deflection or stress constraints (Panesar
etal. 2018; Liu etal. 2018; Li etal. 2018; Zhu etal. 2021).
Many design strategies have been proposed for obtain-
ing optimal designs of FGL structures with prescribed
mechanical properties. Most of the existing strategies may
Responsible Editor: Seonho Cho
* Y. P. Zhang
y.zhang16@uq.edu.au
1 School ofCivil Engineering, The University ofQueensland,
StLucia, QLD4072, Australia
2 Centre de Recherche, Université Bretagne Sud,
IRDL (CNRS UMR 6027), Rue de Saint Maudé,
BP92116,56321LorientCedex, France
3 School ofEngineering, RMIT University, Melbourne3001,
Australia
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 2 of 23
be grouped into two categories. One category is the grey-
scale density mapping (GDM) strategy. This design strategy
first uses standard TO algorithms such as the solid isotropic
material with penalization (SIMP) method (Bendsøe 1989;
Zhou and Rozvany 1991; Mlejnek 1992; Sigmund 2001;
Andreassen etal. 2011; Wang etal. 2022), the evolutionary
structural optimization method (Querin etal. 1998, 2000;
Huang and Xie 2007), or the level set method (Wang etal.
2003; Challis and Guest 2009; Challis 2010; van Dijk etal.
2013; Azari Nejat etal. 2022; Lin etal. 2022) with finite
element discretization or Isogeometric Analysis (Guerder
etal. 2022) to determine a greyscale density solution of
material distribution. Next, FGL structures with a selected
type-based unit cell are established according to the obtained
material distribution. The other category is the ground struc-
ture optimization (GSO) strategy. In this strategy, physical
structures such as trusses, beams or frames are used to rep-
resent the FGL structures. Gradient-based and derivative
free methods have been used to optimize the cross-section
dimensions, node positions and the structural topology (or
element connectivity) of truss, beam and frame subjected
to prescribed design requirements such as volume ratio,
deflection and stress (Miguel etal. 2013; Nguyen etal. 2013;
Zhang etal. 2015; Han and Lu 2018). By comparing these
two main strategies, GDM is more efficient because it needs
to optimize a smaller number of design variables associ-
ated with each structural element. In contrast, GSO requires
the optimization of a much larger number of design vari-
ables including material properties, node positions and the
structure topology (or element connectivity) (Miguel etal.
2013). However, the GSO strategy provides a clearer physi-
cal structural configuration without the need of designing the
internal structure and the connectivity for each lattice unit
cell (Panesar etal. 2018).
In this paper, we propose a new strategy for the optimal
design of FGL structures. This new strategy involves using a
recently proposed lattice structure model called the Hencky
bar-grid model (HBM) (Zhang etal. 2021a, b, 2022). A
Hencky bar model is a type of bar-spring structural model
that was pioneered by Hencky (1921). Owing to its simple
and clear physical representation and great flexibility, it had
been recently revisited and expanded to study various static,
buckling and dynamic problems of different structural types
(Wang etal. 2017, 2020; Zhang etal. 2018c, b, a, d, 2019a,
b, 2022). HBM comprises bars and springs whose stiffnesses
may be adjusted to allow for different material property
distributions. The optimization of the spring stiffnesses is
performed by using a gradient-based TO method to mini-
mize the compliance while satisfying required constraints.
The proposed design framework combines the advantages
of both GDM and GSO strategies. Firstly, HBM is easy to
couple with standard TO algorithms such as SIMP to effi-
ciently obtain optimal designs involving a smaller number
of design variables such as the spring stiffnesses of each grid
cell. Secondly, HBM is a simple bar-spring model which
does not need detail design to the internal structure and con-
nectivity for every lattice unit cells.
The layout of this paper is as follows: Sect.2 presents
the optimization problem definition. The new version of
the HBM is presented in Sect.3. Section4 explains how to
select the value of the HBM spring stiffnesses for solving a
linear elasticity problem. The implementation of the HBM
for TO is described in Sect.5. Section6 illustrates exam-
ples of an FGL beam and an FGL plate that are optimally
designed through the proposed HBM-TO method. Section7
shows a summary of the proposed HBM-TO method and
Sect.8 gives the concluding remarks. Appendix A contains
a simple python code that was used to obtain the optimal
solutions presented in Sect.6.
2 Problem denition
Consider an elastic FGL structure with length
𝛼L
, width
L
, and a uniform thickness
h
, which is illustrated in Fig.1.
A non-uniform Young's modulus
E(x,y)
and a constant
Poisson’s ratio
𝜈
are assumed. The FGL structure is simply
supported along its longitudinal edges and is subjected to a
central line load
P
. Body forces are not considered.
The problem at hand is to determine the optimal design
of the cross-section of FGL structures composed by either
large or small lattice units for maximum stiffness (Kaveh
etal. 2008), i.e. having minimum compliance (Hassani and
Hinton 1999) and then design a structure made by function-
ally graded material based on the material distribution given
by the optimal FGL structures. Figure1b and c illustrates
optimal design examples that differ in terms of lattice cell
size. Figure1d illustrates an optimal design example of a
functionally graded solid structure.
3 Hencky bar‑grid model
Hencky (1921) proposed a discrete structural model com-
prising rigid bars connected by frictionless hinges and elas-
tic rotational springs for solving the elastic buckling prob-
lem of columns under a compressive axial load. Since then,
there has been much work done on the so-called Hencky
bar-chain/net/grid model (HBM) for analysis of all kinds of
structural forms from beams to frames to arches to plates
(Wang etal., 2020). Recently, Zhang etal. (2021b) extended
the HBM for solving plane elasticity problems which has the
ability to model 2D structures for the full range of Poisson’s
ratio. In contrast, some previous lattice models such as the
models formulated by Born and Karman (1912) and Hren-
nikoff (1941) can only model 2D structures for Poisson’s
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 3 of 23 276
ratio less than 1/3 (Zhang etal. 2021a; Challamel etal.
2022).
In this study, we develop a slightly different version of
HBM. A comparison of the new version of HBM and its pre-
decessor reported in Zhang etal. (2021b) is given in Fig.2.
The main difference between the newly proposed HBM and
its predecessor is the construction of the internal (i.e. not at
the edge or corner) rigid bars. The new HBM has twice the
number of rigid bars and springs. This change is essential,
as it allows the determination of local stiffness matrix for
each HBM grid cell (see Eq.(27)) which will be required
by the topology optimization algorithm developed for the
adoption of HBM.
Figure3 shows the structure representation, a unit rigid
bar-grid representation and a lattice structure representa-
tion of the new HBM. As shown in Fig.3a, the new version
of HBM discretize a continuum plane structure into a rigid
bar-grid system with a cell size
that are connected by fric-
tionless hinges. Within a grid cell, four primary axial springs
connect two rigid bars placed at each side of the grid cell
with stiffnesses
and
kyy
in the x- and y- directions, respec-
tively, as shown in Fig.3b. Four secondary axial springs are
employed to model the Poisson effect with stiffnesses
kxy
and
kyx
in the y- and x- directions, respectively, as shown in
Fig.3b. Four torsional springs are installed at the four cor-
ners with stiffnesses
kS
as shown in Fig.3b. While primary
and secondary springs model the axial stiffness and the Pois-
son effect, respectively, torsional springs are employed to
resist in-plane shear forces. A lattice structure representation
related to HBM is shown in Fig.3c. It is worth noting that
the secondary axial springs of HBM are represented by some
curved bars. It is possible to use other structures to represent
the secondary axial springs.
For the sake of clarity, we shall demonstrate the topology
optimization of HBM by using the lattice structures com-
prising basic lattice cell as shown in Fig.3c. For a lattice cell
whose centre is located at
x
𝓁
=i+
1
2
and
y
𝓁
=j+
1
2
, its hori-
zontal and vertical axial bars have stiffnesses of
kxx
i+1
2
,j+
1
2
and
kyy
i+1
2
,j+
1
2
, respectively. Likewise, its curved bars located at the
corners and middle of the cell have stiffnesses of
kS
i+1
2
,j+
1
2
and
k
xy
i+1
2,j+1
2
(
or kyx
i+1
2,j+1
2
)
, respectively. We assume that the bars
in all lattice cells have a constant elastic modulus
E
and a
constant density
𝜌
but a varying cross-section area
A
i+1
2
,j+
1
2
.
Therefore, an HBM grid with stiffer springs results in a lat-
tice cell with a greater mass as shown in Fig.3d.
When a bar-grid cell as shown in Fig.3b is stretched or
shortened, the resulting axial forces
fx
,fy
are resisted by
both primary axial springs and secondary axial springs, i.e.
(1)
fx
i+1
2,j
=k
xx
i+1
2,j+1
2(ui+1,jui,j)+
1
2kxy
i+1
2,j+1
2
(
v
i,j
+
1
+v
i
+
1,j
+
1
v
i,j
v
i
+
1,j)
(2)
fy
i,j+1
2
=kyy
i+1
2,j+1
2(vi,j+1vi,j)+
1
2kyx
i+1
2,j+
1
2
(u
i
+1,
j
+u
i
+1,
j
+1u
i
,
j
u
i
,
j
+1)
p
x
y
z
P
L
αL
x
y
P
L
αL
x
y
(a)
(b)
(c)
(d)
P
L
αL
x
y
Fig. 1 Design FGL structure subjected to a central line load a 3D
Geometry; b illustration of an optimal design with large lattice cells;
c illustration of an optimal design with small lattice cells; d illustra-
tion of a structure made by functionally graded material based on the
optimal FGL design
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 4 of 23
where u, v are the in-plane displacements in the x- and y-
directions at the joints, respectively. The subscripts i and j
indicate the location of corresponding joints in the model.
When the new version HBM cell undergoes in-plane
shearing, the elastic torsional springs are deformed due to
the change of angles at each corner. In this way, the lumped
torque TS for one bar-grid cell is given by
where θ is the change of angle at each corner of the bar-grid
cell. The superscripts a, b, c, and d denote the corner points
at the bar-grid cell (see Fig.3b).
(3)
fx
i+1
2,j+1
=k
xx
i+1
2,j+1
2(ui+1,j+1ui,j+1)+
1
2kxy
i+1
2,j+
1
2
(
v
i
,
j
+1+v
i
+1,
j
+1v
i
,
j
v
i
+1,
j)
(4)
fy
i+1,j+1
2
=k
yy
i+1
2,j+1
2(vi+1,j+1vi+1,j)+
1
2kyx
i+1
2,j+
1
2
(
u
i
+
1,j
+u
i
+
1,j
+
1
u
i,j
u
i,j
+
1)
(5)
TSa
i+1
2
,j+1
2
=k
S
i+1
2
,j+1
2
𝜃
a
i+1
2
,j+
1
2
(6)
TSb
i+1
2
,j+1
2
=k
S
i+1
2
,j+1
2
𝜃
b
i+1
2
,j+
1
2
(7)
TSc
i+1
2
,j+1
2
=k
S
i+1
2
,j+1
2
𝜃
c
i+1
2
,j+
1
2
(8)
TSd
i+1
2
,j+1
2
=k
S
i+1
2
,j+1
2
𝜃
d
i+1
2
,j+
1
2
Based on small angle approximation, θ may be expressed
in terms of the in-plane displacements u, v as
The elastic strain energy
U
i+1
2
,j+
1
2
including the contribu-
tion of the torsional energy stored in the deformed springs
for a bar-grid cell is
In view of Eqs. (1)–(13) the elastic strain energy
U
i+1
2
,j+
1
2
can be reformulated in terms of in-plane displacements u, v,
(9)
𝜃
a
i+1
2
,j+1
2
tan 𝜃a
i+1
2
,j+1
2
=1
2
(u
i,j+1
u
i,j
𝓁+
v
i+1,j
v
i,j
𝓁
)
(10)
𝜃
b
i+1
2
,j+1
2
tan 𝜃b
i+1
2
,j+1
2
=1
2
(u
i+1,j+1
u
i+1,j
𝓁+
v
i+1,j
v
i,j
𝓁
)
(11)
𝜃
c
i+1
2
,j+1
2
tan 𝜃c
i+1
2
,j+1
2
=1
2
(u
i+1,j+1
u
i+1,j
𝓁+
v
i+1,j+1
v
i,j+1
𝓁
)
(12)
𝜃
d
i+1
2
,j+1
2
tan 𝜃d
i+1
2
,j+1
2
=1
2(u
i,j+1
u
i,j
𝓁+
v
i+1,j+1
v
i,j+1
𝓁
)
(13)
U
i+1
2,j+1
2
=
1
2(ui+1,jui,j)fx
i+1
2,j
+
1
2(ui+1,j+1ui,j+1)fx
i+1
2,j+1
+1
2(vi,j+1vi,j)fy
i,j+1
2
+1
2(vi+1,j+1vi+1,j)fy
i+1,j+
1
2
+1
2
𝜃a
i+1
2,j+1
2
TSa
i+1
2,j+1
2
+1
2
𝜃b
i+1
2,j+1
2
TSb
i+1
2,j+1
2
+1
2
𝜃c
i+1
2
,j+1
2
TSc
i+1
2
,j+1
2
+1
2
𝜃d
i+1
2
,j+1
2
TSd
i+1
2
,j+1
2
Fig. 2 Comparison of structure
representation between a prede-
cessor model and b new HBM
ll
l
l
i
j
0 1 2
0
1
2
ll
l
l
i
j
0 1 2
0
1
2
(a)
(b)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 5 of 23 276
A
l l l l l l l l l l l l l l l l
B
C
D
l
l
l
l
l
l
l
l
y
x
Fyy
Fyy Fyy Fyy+Fxy
F
yy
Fyy
Fyy
Fxy
Fxy
Fxy Fxy Fxy Fxy
Fyy+Fxy Fyy+Fxy
Fxy
Fxy Fxy
Fyy+Fxy
Fxy
Fxx
Fxx
Fxx+Fxy
Fxx
Fxx
Fxx+Fxy
bx
by
ui,j, vi,j ui+1,j, vi+1,j
ui+1,j+1, vi+1,j+1
ui,j+1, vi,j+1
torsional
spring
primary
axial
spring
secondary
axial spring
kS
frictionless
hinge
frictionless
pulley
kyy
kxx
kyx
kxy
l
l
ab
c
d
Fi+ ,j+
1
21
2
yyd Fi+ ,j+
1
21
2
yyc
Fi+ ,j+
1
21
2
yya Fi+ ,j+
1
21
2
yyb
Fi+ ,j+
1
21
2
xxb
Fi+ ,j+
1
21
2
xxc
Fi+ ,j+
1
21
2
xxd
Fi+ ,j+
1
21
2
xxa
Fi+ ,j+
1
21
2
xyd
Fi+ ,j+
1
21
2
xya
Fi+ ,j+
1
21
2
xya Fi+ ,j+
1
21
2
xyb
Fi+ ,j+
1
21
2
xyd Fi+ ,j+
1
21
2
xyc
Fi+ ,j+
1
21
2
xyc
Fi+ ,j+
1
21
2
xyb
primary
axial
spring
kyy
kxx
kyx
kxy
kS
torsional
spring
secondary
axial spring
l
l
HBM spring stiffness
(a)
(b) (c)
(d)
Fig. 3 New version of Hencky bar-grid model (HBM): a geometry and boundary conditions; b unit bar-grid cell representation; c lattice struc-
ture representation; d relation between HBM and lattice structure
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 6 of 23
Next, the work done by in-plane axial forces
Fxx
,
Fyy
, in-
plane shear force
Fxy
and the in-plane body forces
bx
and
by
lumped on the joints of each HBM cell as shown in Fig.3b
are given by
The total potential energy function is given by
where
Ω
is a set contains all the indexes of the bar-grid cells
for the HBM.
Now, the local stiffness matrix of an HBM cell is obtained
through a variational formulation, by taking
(14)
U
i+1
2,j+1
2
=
1
2k
xx
i+1
2,j+1
2[(ui+1,jui,j)2+(ui+1,j+1ui,j+1)2
]
+1
2kyy
i+1
2,j+1
2[(vi,j+1vi,j)2+(vi+1,j+1vi+1,j)2]
+1
2kxy
i+1
2,j+1
2(ui+1,jui,j)(vi,j+1+vi+1,j+1vi,jvi+1,j)
+1
2kxy
i+1
2,j+1
2(ui+1,j+1ui,j+1)(vi,j+1+vi+1,j+1vi,jvi+1,j
)
+1
2kyx
i+1
2,j+1
2(vi,j+1vi,j)(ui+1,j+ui+1,j+1ui,jui,j+1)
+1
2kyx
i+1
2,j+1
2(vi+1,j+1vi+1,j)(ui+1,j+ui+1,j+1ui,jui,j+1
)
+1
82kS
i+1
2,j+1
2[(ui,j+1ui,j+vi+1,jvi,j)2
+(ui+1,j+1ui+1,j+vi+1,jvi,j)2
+(ui+1,j+1ui+1,j+vi+1,j+1vi,j+1)2
+
(
ui,j+1ui,j+vi+1,j+1vi,j+1
)
2
]
(15)
W
i+1
2,j+1
2
=ui,j
(
Fxxa
i+1
2,j+1
2
+Fxya
i+1
2,j+1
2
+bxa
i+1
2,j+1
2
)
+vi,j(Fyya
i+1
2,j+1
2
+Fxya
i+1
2,j+1
2
+bya
i+1
2,j+1
2)
+ui+1,j(Fxxb
i+1
2,j+1
2
+Fxyb
i+1
2,j+1
2
+bxb
i+1
2,j+1
2)
+vi+1,j(Fyyb
i+1
2,j+1
2
+Fxyb
i+1
2,j+1
2
+byb
i+1
2,j+1
2)
+ui+1,j+1(Fxxc
i+1
2,j+1
2
+Fxyc
i+1
2,j+1
2
+bxc
i+1
2,j+1
2
)
+vi+1,j+1(Fyyc
i+1
2,j+1
2
+Fxyc
i+1
2,j+1
2
+byc
i+1
2,j+1
2
)
+ui,j+1(Fxxd
i+1
2,j+1
2
+Fxyd
i+1
2,j+1
2
+bxd
i+1
2,j+1
2)
+vi,j+1
(
Fyyd
i+1
2
,j+1
2
+Fxyd
i+1
2
,j+1
2
+byd
i+1
2
,j+1
2)
(16)
Π=
Ω
i=0
Ω
j=0(
Ui+1
2,j+1
2
Wi+1
2,j+1
2
)
Based on Eq.(14), the discrete equations of the derivative
of the elastic strain energy
U
i+1
2
,j+
1
2
for an HBM cell
(
i
+
1
2
,j+
1
2
) are given by
(17)
𝛿Π=0
(18)
𝜕
Ui+1
2,j+1
2
𝜕ui,j
kxx
i+1
2,j+1
2ui+1,jui,j1
2
kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2
vi,j+1+vi+1,j+1vi,jvi+1,j
1
42kS
i+1
2,j+1
22ui,j+12ui,j+vi+1,
j
v
i,j
+v
i
+
1,j
+
1
v
i,j
+
1
(19)
𝜕
Ui+1
2,j+1
2
𝜕ui+1,j
kxx
i+1
2,j+1
2ui+1,jui,j
+1
2kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2
vi,j+1+vi+1,j+1vi,jvi+1,j
1
42kS
i+1
2,j+1
22ui+1,j+12ui+1,j+vi+1,
j
v
i
,
j
+v
i
+1,
j
+1v
i
,
j
+1
(20)
𝜕
Ui+1
2,j+1
2
𝜕ui+1,j+1
kxx
i+1
2,j+1
2ui+1,j+1ui,j+1+1
2
kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2
vi,j+1+vi+1,j+1vi,jvi+1,j
+1
42kS
i+1
2,j+1
22ui+1,j+12ui+1,j+vi+1,
j
v
i
,
j
+v
i
+1,
j
+1v
i
,
j
+1
(21)
𝜕
Ui+1
2,j+1
2
𝜕ui,j+1
kxx
i+1
2,j+1
2ui+1,j+1ui,j+1
1
2kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2
vi,j+1+vi+1,j+1vi,jvi+1,j
+1
42kS
i+1
2,j+1
22ui,j+12ui,j+vi+1,
j
v
i
,
j
+v
i
+1,
j
+1v
i
,
j
+1
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 7 of 23 276
and
Likewise, the discrete equations of the derivative of the exter-
nal work
W
i+1
2
,j+
1
2
for an HBM cell (
i
+
1
2
,j+
1
2
) are given by
(22)
𝜕
Ui+1
2,j+1
2
𝜕vi,j
kyy
i+1
2,j+1
2vi,j+1vi,j1
2kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2
ui+1,j+ui+1,j+1ui,jui,j+11
42kS
i+1
2,j+1
2
2v
i
+
1,j
2v
i,j
+u
i,j
+
1
u
i,j
+u
i
+
1,j
+
1
u
i
+
1,j
(23)
𝜕
Ui+1
2,j+1
2
𝜕vi+1,j
kyy
i+1
2,j+1
2vi+1,j+1vi+1,j1
2kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2
ui+1,j+ui+1,j+1ui,jui,j+1+1
42kS
i+1
2,j+1
2
2v
i
+
1,j
2v
i,j
+u
i,j
+
1
u
i,j
+u
i
+
1,j
+
1
u
i
+
1,j
(24)
𝜕
Ui+1
2,j+1
2
𝜕vi+1,j+1
kyy
i+1
2,j+1
2vi+1,j+1vi+1,j+1
2kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2
ui+1,j+ui+1,j+1ui,jui,j+1+1
42kS
i+1
2,j+1
2
2v
i
+1,
j
+12v
i
,
j
+1+u
i
,
j
+1u
i
,
j
+u
i
+1,
j
+1u
i
+1,
j
(25)
𝜕
Ui+1
2,j+1
2
𝜕vi,j+1
kyy
i+1
2,j+1
2vi,j+1vi,j+1
2kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2
ui+1,j+ui+1,j+1ui,jui,j+11
42kS
i+1
2,j+1
2
2v
i
+1,
j
+12v
i
,
j
+1+u
i
,
j
+1u
i
,
j
+u
i
+1,
j
+1u
i
+1,
j
(26)
𝜕
Wi+1
2,j+1
2
𝜕ui,j
Fxxa
i+1
2,j+1
2
+Fxya
i+1
2,j+1
2
+bxa
i+1
2,j+1
2
;
𝜕Wi+1
2,j+1
2
𝜕vi,j
Fyya
i+1
2
,j+1
2
+Fxya
i+1
2
,j+1
2
+bya
i+1
2
,j+1
2
(27)
𝜕
Wi+1
2,j+1
2
𝜕ui+1,j
Fxxb
i+1
2,j+1
2
+Fxyb
i+1
2,j+1
2
+bxb
i+1
2,j+1
2
;
𝜕Wi+1
2,j+1
2
𝜕vi+1,j
Fyyb
i+1
2
,j+1
2
+Fxyb
i+1
2
,j+1
2
+byb
i+1
2
,j+1
2
(28)
𝜕
Wi+1
2,j+1
2
𝜕ui+1,j+1
Fxxc
i+1
2,j+1
2
+Fxyc
i+1
2,j+1
2
+bxc
i+1
2,j+1
2
;
𝜕Wi+1
2,j+1
2
𝜕vi+1,j+1
Fyyc
i+1
2
,j+1
2
+Fxyc
i+1
2
,j+1
2
+byc
i+1
2
,j+1
2
(29)
𝜕
Wi+1
2,j+1
2
𝜕ui,j+1
Fxxd
i+1
2,j+1
2
+Fxyd
i+1
2,j+1
2
+bxd
i+1
2,j+1
2
;
𝜕Wi+1
2,j+1
2
𝜕v
i,j
+
1
Fyyd
i+1
2,j+1
2
+Fxyd
i+1
2,j+1
2
+byd
i+1
2,j+
1
2
From Eqs. (18)–(29), the local stiffness matrix for a bar-
grid cell of the HBM is given by
where
with
and
k
xy
=1
2
(
kxy
i+1
2
,j+1
2
+kyx
i+1
2
,j+1
2)
;
(30)
[k]
i+1
2
,j+1
2
{u}
i+1
2
+j+1
2
={f}
i+1
2
,j+
1
2
(31)
[
k]i+1
2,j+1
2
=
axbc
xdebf d
aydf be d c
y
ax
bf deb
aydc
y
be
axbc
xd
aydf
sym.axb
ay
(32)
a
x=kxx
+k
S
2𝓁
2;ay=kyy
+k
S
2𝓁
2
(33)
b
=k
xy
2
+kS
4𝓁2
(34)
swcx
=−k
xx
;c
y
=−k
yy
(35)
d
=k
xy
2
kS
4𝓁
2
(36)
e=0
(37)
f=− k
S
2𝓁
2
(38)
{
u}i+1
2+j+1
2
=
ui,j
vi,j
ui+1,j
vi+1,j
ui+1,j+1
vi+1,j+1
ui,j+1
vi,j+1
;{f}i+1
2+j+1
2
=
Fxxa
i+1
2,j+1
2
+Fxya
i+1
2,j+1
2
+bxa
i+1
2,j+1
2
Fyya
i+1
2,j+1
2
+Fxya
i+1
2,j+1
2
+bya
i+1
2,j+1
2
Fxxb
i+1
2,j+1
2
+Fxyb
i+1
2,j+1
2
+bxb
i+1
2,j+1
2
Fyyb
i+1
2,j+1
2
+Fxyb
i+1
2,j+1
2
+byb
i+1
2,j+1
2
Fxxc
i+1
2,j+1
2
+Fxyc
i+1
2,j+1
2
+bxc
i+1
2,j+1
2
Fyyc
i+1
2,j+1
2
+Fxyc
i+1
2,j+1
2
+byc
i+1
2,j+1
2
Fxxd
i+1
2,j+1
2
+Fxyd
i+1
2,j+1
2
+bxd
i+1
2,j+1
2
Fyyd
i+1
2
,j+1
2
+Fxyd
i+1
2
,j+1
2
+byd
i+1
2
,j+1
2
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 8 of 23
In view of Eqs. (18)–(38) and by assembling the expres-
sions for all nodes in the system, the following linear system
of equation can be derived
where
and
The in-plane displacements of each joint
ui,j
and
vi,j
can
be calculated by solving Eq.(39). Next the axial forces
fx
i+1
2
,
j
and f
y
i,j+
1
2
and the torque
TS
1+1
2
,1+
1
2
are computed via the
obtained in-plane displacements
ui,j
and
vi,j
using Eqs.
(1)–(8).
4 Determination ofHBM spring stiness
In this section, we shall determine the spring stiffnesses
of HBM. For simplicity, it is assumed that each individual
HBM grid cell is homogeneous and isotropic. The inhomo-
geneous material properties of the structure are captured by
varying the material properties of different HBM cells. A
general version of the continuum equations of elastodynam-
ics (or Navier’s equations of elastodynamics) considering
inhomogeneous material properties and plane-strain can be
written as (Gurtin 1973)
and
(39)
[K]HBM{u}={F}HBM
(40)
[
K]HBM =
Ω
i=0
Ω
j=0
[k]i+1
2,j+
1
2
(41)
{
F}HBM =
Ω
i=0
Ω
j=0
{f}i+1
2,j+
1
2
(42)
{
u}=
u0,0
u0,1
unx1,ny1
v0,0
v0,1
vnx1,ny1
(43)
𝜕
𝜕x
[
(1v)E
(1+v)(12v)
𝜕u
𝜕x+vE
(1+v)(12v)
𝜕v
𝜕y
]
+𝜕
𝜕y
[
E
2
(
1+v
)
𝜕u
𝜕y+E
2
(
1+v
)
𝜕v
𝜕x
]
dΩ=𝜌ud
Ω
By considering the local homogeneous properties within
the area of each HBM grid cell, i.e.
x
[(
i1
2)
𝓁,
(
i+1
2)
𝓁
]
,y
[(
j1
2)
𝓁,
(
j+1
2)
𝓁
]
the general continuum equations of elastodynamics can be
simplified to
and
By expanding Eqs. (18) and (19) using Taylor’s series
following the procedures used in (Zhang etal. 2019c), we
obtain
and
By assuming the case of quasi-static, neglecting higher
order terms and comparing Eqs. (48) and (49) to the
(44)
𝜕
𝜕y
[
(1𝜈)E
(1+𝜈)(12𝜈)
𝜕v
𝜕y+𝜈E
(1+𝜈)(12𝜈)
𝜕u
𝜕x
]
+𝜕
𝜕x
[
E
2
(
1+𝜈
)
𝜕u
𝜕y+E
2
(
1+𝜈
)
𝜕v
𝜕x
]
dΩ=𝜌vd
Ω
(45)
E(x,y)=E
i+1
2,j+1
2
=E
e
;
𝜈
(x,y)=
𝜈i+1
2,j+
1
2
=𝜈e;
(i+1
2)𝓁
(
i1
2
)
𝓁
(j+1
2)𝓁
(
j1
2
)
𝓁
h
0
dΩ=Ve
(46)
(
1𝜈e
)
EeVe
(
1+𝜈e)(12𝜈e)
𝜕2u
𝜕x2+EeVe
2(1+𝜈e)
𝜕2
u
𝜕y2
+EeVe
2
(
1+𝜈
e)(
12𝜈
e)
𝜕2v
𝜕x𝜕y
=𝜌Veu
(47)
(
1𝜈e
)
EeVe
(
1+𝜈e)(12𝜈e)
𝜕2v
𝜕y2+EeVe
2(1+𝜈e)
𝜕2
v
𝜕x
2
+EeVe
2
(
1+𝜈
e)(
12𝜈
e)
𝜕2u
𝜕x𝜕y
=𝜌Ve𝜈
(48)
2
kxx
i+1
2,j+1
2
𝓁2𝜕2u
𝜕x2+kS
i+1
2,j+1
2
𝜕2u
𝜕y2
+kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2𝓁2+kS
i+1
2,j+1
2
𝜕2v
𝜕
x
𝜕
y
+O
𝓁4
=0
(49)
2
kyy
i+1
2,j+1
2
𝓁2𝜕
2
v
𝜕y2+kS
i+1
2,j+1
2
𝜕
2
v
𝜕x2
+kxy
i+1
2,j+1
2
+kyx
i+1
2,j+1
2𝓁2+kS
i+1
2,j+1
2
𝜕2u
𝜕
x
𝜕
y
+O
𝓁4
=0
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 9 of 23 276
continuum equation of motion Eqs. (46) and (47), it follows
that
A first-order central finite difference (FD) discretization
of Eqs. (43) and (44) results in
(50)
k
xx
i+1
2,j+1
2
=
(
1𝜈e
)
EeVe
2
(
1+𝜈
e)(
12𝜈
e)
(51)
k
yy
i+1
2,j+1
2
=
(
1𝜈e
)
EeVe
2
(
1+𝜈
e)(
12𝜈
e)
(52)
k
xy
i+1
2,j+1
2
=
𝜈
e
E
e
V
e
2
(
1+𝜈
e)(
12𝜈
e)
(53)
k
S
i+1
2,j+1
2
=EeVe𝓁
2
2
(
1+𝜈
e)
(54)
1
2
1𝜈
EV
21+𝜈12𝜈i1
2,j+1
2
+
1𝜈
EV
21+𝜈12𝜈i1
2,j1
2
u
i,jui1,j
21
2𝜈EV
21+𝜈12𝜈i1
2,j+1
2
vi,j+1vi,j+vi1,j+1vi1,
j
22
+𝜈EV
21+𝜈12𝜈i1
2,j1
2
vi,jvi,j1+vi1,jvi1,j1
22
+1
21𝜈EV
21+𝜈12𝜈i+1
2,j+1
2
+1𝜈EV
21+𝜈12𝜈i+1
2,j1
2
u
i+1,jui,j
2+1
2𝜈EV
21+𝜈12𝜈i+1
2,j+1
2
vi+1,j+1vi+1,j+vi,j+1vi,
j
22
+𝜈EV
21+𝜈12𝜈i+1
2,j1
2
vi+1,jvi+1,j1+vi,jvi,j1
22
+EV
21+𝜈i+1
2,j+1
2
ui,j+1ui,j+ui+1,j+1ui+1,j
22
+EV
21+𝜈i1
2,j+1
2
ui,j+1ui,j+ui1,j+1ui1,j
22
+EV
21+𝜈i+1
2,j+1
2
vi+1,j+1vi,j+1+vi+1,jvi,j
22
+EV
21+𝜈i1
2,j+1
2
vi,j+1vi1,j+1+vi,jvi1,j
22
EV
21+𝜈i+1
2,j1
2
ui,jui,j1+ui+1,jui+1,j1
22
+EV
21+𝜈i1
2,j1
2
ui,j+1ui,j+ui1,j+1ui1,j
22
+EV
21+𝜈i+1
2,j1
2
vi+1,jvi,j+vi+1,j1vi,j1
22
+EV
21+𝜈i1
2,j1
2
vi,jvi1,j+vi,j1vi1,j1
22
=1
4𝜌
Vi+1
2,j+1
2
+Vi1
2,j+1
2
+Vi+1
2,j1
2
+Vi1
2,j1
2
ui,j
The difference equation in y- direction is obtained from
Eq.(54) by swapping variables u, v and indices i, j. Now,
if we assemble four adjacent HBM grid cells like Fig.3b,
the difference equation of the middle node can be obtained
based on Eqs. (18)–(25) as follows:
The difference equation in y- direction is Eq.(55) by
swapping variables u, v and indices i, j. In view of Eqs.
(50)–(55), it can be seen that solving the linear algebraic
system given by the HBM proposed in this work is equiva-
lent to solving an inhomogeneous two-dimensional elasticity
problem using FD method.
Noting that the new HBM coincides with the old one
(Zhang etal. 2021b) when modelling homogeneous struc-
tures. So, it is possible to use the old HBM to model inho-
mogeneous structures, but the resulting energy functions
could not be readily applied to derive the local stiffness
matrix for each grid cell.
5 Topology optimization
The SIMP method is employed to optimize the HBM for-
mulated herein. The spring stiffnesses of each grid cell is
formulated as
(55)
1
2
k
xx
i1
2,j+1
2
+k
xx
i1
2,j1
2
ui,jui1,j
1
4kxy
i1
2,j+1
2vi,j+1vi,j+vi1,j+1vi1,j
+
kxy
i1
2,j1
2vi,jvi,j1+vi1,jvi1,j1
+
1
2kxx
i+1
2,j+1
2
+kxx
i+1
2,j1
2ui+1,jui,j
+
1
4kxy
i+1
2,j+1
2vi+1,j+1vi+1,j+vi,j+1vi,j
+
kxy
i+1
2,j1
2vi+1,jvi+1,j1+vi,jvi,j1
+
1
22kS
i+1
2,j+1
2ui,j+1ui,j+ui+1,j+1ui+1,j
+
kS
i1
2,j+1
2ui,j+1ui,j+ui1,j+1ui1,j
+
kS
i+1
2,j+1
2vi+1,j+1vi,j+1+vi+1,jvi,j
+
kS
i1
2,j+1
2vi,j+1vi1,j+1+vi,jvi1,j1
2
2
kS
i+1
2,j1
2ui,jui,j1+ui+1,jui+1,j1
+
kS
i1
2,j1
2ui,j+1ui,j+ui1,j+1ui1,j
+
kS
i+1
2,j1
2vi+1,jvi,j+vi+1,j1vi,j1
+
kS
i1
2,j1
2
vi,jvi1,j+vi,j1vi1,j1
=0
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 10 of 23
where
kxx
0
,k
yy
0
,kxy
0
,k
yx
0
and
kS
0
are the starting values of spring
stiffnesses for TO,
𝜌
is a non-dimensional variable intro-
duced for HBM-TO and
m
is the exponent for penalization.
In order to avoid singularity issues when solving the linear
system given in Eq.(39), all HBM spring stiffnesses are
constrained to have a minimum value of
𝜙min
k
0
at the void
areas. In view of Eqs. (56)–(61) and Eq.(14), the elastic
strain energy of a grid cell may be expressed as
where the starting value of elastic strain energy
U0
i+1
2
,j+
1
2
is
given by
(56)
kxx
i+1
2
,j+1
2
=𝜒 i+1
2
,j+1
2
k
xx
0
(57)
kyy
i+1
2
,j+1
2
=𝜒 i+1
2
,j+1
2
k
yy
0
(58)
kxy
i+1
2
,j+1
2
=𝜒 i+1
2,j+1
2
k
xy
0
(59)
kyx
i+1
2
,j+1
2
=𝜒 i+1
2,j+1
2
k
yx
0
(60)
kS
i+1
2
,j+1
2
=𝜒 i+1
2,j+1
2
k
S
0
(61)
𝜒
i+1
2,j+1
2
=𝜙min +𝜌
m
i+1
2
,j+1
2
(
1𝜙min
)
(62)
U
i+1
2
,j+1
2
=𝜒 i+1
2
,j+1
2
U
0
i+1
2
,j+
1
2
(63)
U0
i+1
2,j+1
2
=
1
2k
xx
0
[
(ui+1,jui,j)2+(ui+1,j+1ui,j+1)2
]
+1
2kyy
0[(vi,j+1vi,j)2+(vi+1,j+1vi+1,j)2]
+1
2kxy
0(ui+1,jui,j)(vi,j+1+vi+1,j+1vi,jvi+1,j)
+1
2kxy
0(ui+1,j+1ui,j+1)(vi,j+1+vi+1,j+1vi,jvi+1,j
)
+1
2kyx
0(vi,j+1vi,j)(ui+1,j+ui+1,j+1ui,jui,j+1)
+1
2kyx
0(vi+1,j+1vi+1,j)(ui+1,j+ui+1,j+1ui,jui,j+1
)
+1
82kS
0[(ui,j+1ui,j+vi+1,jvi,j)2
+(ui+1,j+1ui+1,j+vi+1,jvi,j)2
+(ui+1,j+1ui+1,j+vi+1,j+1vi,j+1)2
+
(
ui,j+1ui,j+vi+1,j+1vi,j+1
)
2
]
In view of Eqs. (63) and (16), the total strain energy is
The TO problem for HBM can be posed as
where
Vf
is the prescribed volume fraction and
V(𝜌 )
and
V0
are the material volume and design volume, respectively.
There are various methods for solving the aforementioned
optimization problem such as Optimality Criteria (OC) (Yin
and Yang 2001), Sequential Linear Programming (Dunning
and Kim 2015), and Method of Moving Asymptotes (Svan-
berg 1987). The OC method is adopted herein due to its
efficiency in solving optimization problems with a single
objective function (Sigmund 1997).
Following the works of Bendsøe (1989), Sigmund (1997)
and Sigmund and Petersson (1998) and assuming a constant
change of design volume, the design variable
𝜌
i+1
2
,j+
1
2
is
updated as follows:
If
max (
0, 𝜌 i+1
2,j+1
2
𝛿
)
𝜌 i+1
2,j+1
2
D𝛾
i+1
2
,j+
1
2
if
max
(0, 𝜌 i+1
2,j+1
2
𝛿
)
< 𝜌 i+1
2,j+1
2
D𝛾
i+1
2
,j+1
2
min (1, 𝜌 i+1
2,j+1
2
+𝛿
)
if
𝜌
i+1
2,j+1
2
D𝛾
i+1
2
,j+1
2
>min
(
1, 𝜌 i+1
2,j+1
2
+𝛿
)
where δ is a positive value to limit the change of design vari-
able
𝜌
i+1
2
,j+
1
2
between two successive iterations and γ is a
numerical damping exponent. These parameters are intro-
duced to avoid drastic change of density between adjacent
cells that are impractical (i.e. porous structures).
D
i+1
2
,j+
1
2
is
an auxiliary variable defined as
where λ is a Lagrangian multiplier which can be obtained
through the bisection method. In view of Eq.(64), the HBM
grid cell sensitivity matrices
c
can be written as follows
(64)
U
(𝜌 )=
Ω
i=0
Ω
j=0
𝜒 i+1
2,j+1
2
(
𝜌 i+1
2,j+1
2
)
U0
i+1
2,j+1
2
.
(65)
min
𝜌
(0,1]
(
U(𝜌 )
)
, subject to: Eq.(39)and
V(𝜌 )
V
0
=Vf
(66)
𝜌
i+1
2
,j+1
2
max
(
0, 𝜌 i+1
2
,j+1
2
𝛿
)
(67)
𝜌
i+1
2,j+1
2
𝜌 i+1
2,j+1
2
D
𝛾
i+1
2
,j+
1
2
(68)
𝜌
i+1
2
,j+1
2
min
(
1, 𝜌 i+1
2
,j+1
2
+𝛿
)
(69)
D
i+1
2,j+1
2
=−
1
𝜆
𝜕Π
𝜕 𝜌 i+1
2
,j+1
2
=−𝜆1ci+1
2+j+1
2
,
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 11 of 23 276
In order to avoid potential “chess board patterns”, a sim-
ple filtering technique formulated by Andreassen etal.
(2011) is adopted by replacing the HBM grid cell sensitivity
matrices
c
i+1
2
+j+
1
2
with
where
𝜑 (i,j,k,l)
is a weight factor defined as
The k and l indicate the location of corresponding joints
in the model. Equations(71) and (72) indicate that the filter
is a local averaging operator whereby the HBM grid cell
sensitivity matrices
c
i+1
2
+j+
1
2
take on an average value of all
c
weighted by
𝜑
within a radius
r
. The filtered HBM grid cell
sensitivity matrices obtained through Eqs. (71) and (72) are
used when updating
D
i+1
2
,j+
1
2
in Eq.(69).
6 Results anddiscussions
In this section, we first validate the proposed HBM through
modeling continuous functionally graded structures. We
then apply the HBM-TO to determine the optimal design
of a beam and a plate FGL structures subjected to central
line loading. The considered boundary conditions include
loads and prescribed displacements. In the following exam-
ple problems, we assume the material has Poisson’s ratio of
0.3 and set the starting value for the elastic modulus
E0
as 1.
6.1 Validation ofHBM forfunctionally graded plane
bodies
In this section, plane-stress and plane-strain elasticity prob-
lems are considered that are confined in a square design
domain under the loading and displacements boundary con-
ditions illustrated in Fig.1. The functionally graded Young’s
modulus is assumed to be
The accuracy of the proposed HBM is assessed by com-
paring the predicted in-plane displacements u, v, with results
(70)
c
i+1
2+j+1
2
=𝜕U
𝜕 𝜌 i+1
2
,j+1
2
=−m𝜌 m1
i+1
2,j+1
2
(
1𝜙min
)
U0
i+1
2,j+
1
2
(71)
c
i+1
2+j+1
2
1
Ω
k
Ω
l
𝜑 (i,j,k,l)
Ω
k
Ω
l
𝜑 (i,j,k,l)ck+1
2,l+
1
2
(72)
𝜑
(i,j,k,l)=max
(
r𝓁
(ik)2+(jl)2,0
)
(73)
E
(x,y)=2
[(
x
L1
2
)2
+
(
y
L1
2
)2]
obtained through the direct stiffness matrix method (Rao
2005). A general form of the local stiffness matrix for a unit
square FEM plane element and a square HBM cell with unit
bar length, can be expressed using Eq.(31). We assume iso-
tropic material properties within individual FEM elements
and HBM cells. The coefficients of the local stiffness matrix
are for plane-stress:
for plane-strain:
(74)
a
FEM =
1
1𝜈2
1
2
𝜈
6
;aHBM =
1
1𝜈2
3
4
𝜈
4
(75)
b
FEM =1
1𝜈2
1
8+
𝜈
8
;bHBM =1
8
1𝜈2
1
8+
𝜈
8
(76)
c
FEM =
1
1𝜈2
1
4
𝜈
12
;cHBM =−
1
2
1𝜈2
(77)
d
FEM =1
1𝜈2
1
8+3𝜈
8
;dHBM =1
1𝜈2
1
8+3𝜈
8
(78)
e
FEM =1
1𝜈2
1
4+
𝜈
12
;eHBM =
0
(79)
fFEM =
𝜈
6
1𝜈2
;fHBM =1
1𝜈2
1
4+
𝜈
4
(80)
a
FEM =1
1+𝜈12𝜈
1
22𝜈
3
;a
HBM
=1
1+𝜈

12𝜈
3
4𝜈
(81)
b
FEM =
1
8
(
1+𝜈
)(
12𝜈
)
;bHBM =
1
8
(
1+𝜈
)(
12𝜈
)
(82)
c
FEM =1
1+𝜈12𝜈
1
4
𝜈
2
;
c
HBM =1
1+𝜈

12𝜈
1
2+
𝜈
2
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 12 of 23
(83)
d
FEM =1
1+𝜈12𝜈
1
8+
𝜈
2
;
d
HBM =1
1+𝜈

12𝜈
1
8+
𝜈
2
(84)
e
FEM =1
1+𝜈

12𝜈
1
4+
𝜈
3
;eHBM =
0
(85)
fFEM =
𝜈
6
1+𝜈

12𝜈
;fHBM =1
1+𝜈

12𝜈
1
4+
𝜈
2
The abovementioned coefficients of the local FEM stiff-
ness matrix can be derived as detailed in Rao (2005). For
the plane-stress case, the coefficients of local HBM stiff-
ness matrix are obtained using the HBM spring stiffnesses
expression given by Zhang etal. (2021b). For the plane-
strain case, the coefficients of local HBM stiffness matrix
are computed by using Eqs. (50)–(53). It is interesting to
note that FEM and HBM have identical values for the coef-
ficients b and d which indicate that both FEM and HBM
have a similar physical interpretation of the Poisson effect.
In investigating the performance of HBM, several simula-
tions were run by adopting grid/mesh size
𝓁=
L
100
. For
Plane-Stress Problem
(a) FEM- (b) HBM-
(c) FEM- (d) HBM-
Fig. 4 A plane-stress body under a central point load. a horizontal displacement field u predicted by FEM; b horizontal displacement field u pre-
dicted by HBM; c vertical displacement field v predicted by FEM; d vertical displacement field v predicted by HBM
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 13 of 23 276
each test, we only change the local stiffness matrix based on
the choice of FEM/HBM or plane-stress/plane-strain. Fig-
ure4a and b presents the horizontal displacement u fields
predicted by FEM and HBM, respectively, for the plane-
stress problem under a central point load. Figure4c and d
present the corresponding vertical displacement v fields.
Figure5a–d presents the in-plane displacements for the
plane-strain problem under a central line load. Based on the
results shown in Figs.4 and 5, it can be seen that the in-
plane displacements given by HBM agree well with those
obtained through FEM for both the plane-stress and plane-
strain problems. These results show that the proposed HBM
is a simple and robust physical structural model that handle
well elasticity problems for functionally graded structures.
Plane-Strain Problem
(a) FEM- (b) HBM-
(c) FEM- (d) HBM-
Fig. 5 A plane-strain body under a central line load. a horizontal displacement field u predicted by FEM; b horizontal displacement field u pre-
dicted by HBM; c vertical displacement field v predicted by FEM; d vertical displacement field v predicted by HBM
P
L
y
zx
L
L
2
Design domain
Fig. 6 Design domain, loading and boundary conditions for FGL
beams
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 14 of 23
6.2 Optimal design ofFGL beams subjected
toacentral line load
The HBM-TO is first applied to the optimal design of an
FGL beam with a square cross-section (
𝛼=1
) under a cen-
tral line load. The bottom edges are fixed in y- direction.
The geometry and boundary conditions are shown in Fig.6.
The HBM-TO is carried out as follows.
1. Set the prescribed volume fraction
Vf=0.4
by initial-
izing all HBM design variable
𝜌
as 0.4;
2. Set parameters
𝛿=0.1
,
𝜙min =0.01
,
r=2
and
𝛾=0.5
;
3. Set the size of HBM grid
𝓁
as
L10
,
L20
or
L30
;
4. Select the exponent for penalization
m
as 0.5, 1 or 2;
5. Solve optimization problem defined by Eq.(65) and
using Eqs. (66)–(72);
6. The optimization process will stop when the change of
elastic strain energy
ΔU
is smaller than 0.01;
7. Finally, construct the FGL beam cross-section based on
the optimal
𝜒
.
Noting that the variation of the elastic modulus changes
the stiffness of the springs, thus the cell mass changes
accordingly. In actual design of FGL structures, the cross-
section area
A
, the elastic modulus
E
and the density
𝜌
of
lattice bars are dependent on the applied material.
Figure7 shows the optimal cross-section designs of FGL
structures obtained by HBM-TO. It can be seen that the solu-
tions are significantly affected by the size of the HBM grid
cell
𝓁
and the penalty exponent
m
. The cross-section mass
is more concentrated when
𝓁
is smaller and
m
is larger. In
contrast, the mass of cross-section becomes more distributed
when
𝓁
is larger and
m
is smaller. It can be seen from Fig.8
that the elastic strain energy converges after 25 iterations
and the solutions converge faster when
m
is smaller. The
converged elastic strain energy tends to be lower when
m
is smaller.
The optimally designed FGL beams can be classified into
two groups: “distributed” and “defined”; see also Fig.9. The
first group suggests the beams could be designed with func-
tionally graded properties such as a graded elastic modulus.
Fig. 7 Optimal designs of FGL
beams with a square cross-sec-
tion obtained by HBM-TO
=0.5 =1 =2
=10
=20
=30
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 15 of 23 276
The second group suggests that the beams could be designed
with a defined geometry. The observed results from Fig.9
show that the “distributed” FGL beams are stiffer than the
“defined” FGL beams. The former ones have a smoother
transition of mass which could help in preventing failure
initiations (Wang 1983; Mahamood etal. 2012). However,
the latter ones could be manufactured more readily.
Based on Eqs. (56)-(60), the variable
𝜒
describes the
distribution of the elastic modulus of the FGL beams. By
least-squares fitting
𝜒
from HBM-TO FGL beams using
𝓁
=
L
20
,m=
0.5
; and
𝓁
=
L
30
,m=
0.5
, one finds the fol-
lowing elastic modulus distribution function, i.e.
(86)
fb(x,y)=
0 if
fb(x,y)<0
1 if
fb(x,y)>1
fb(x,y)otherwise
;x[0, 1],y[0, 1
]
(a) ℓ=
10 (b) ℓ=
20 (c) ℓ=
30
Fig. 8 Convergence of elastic strain energy when optimizing FGL beams
Fig. 9 Two groups of HBM-TO
designed FGL beams
(a) Distributed (b) Defined
Fig. 10 Elastic modulus distribution for HBM-TO FGL beams with
square cross-section under a central line load
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 16 of 23
where
A contour plot showing the distribution given in Eq.(86)
is presented in Fig.10.
(87)
fb(x,y)=0.9 +0.9x114xy
+6.7y1.3x2+336.6x2y
2248.6x2y2+618xy233.1y2
+0.8x3445.3x3y+3261.3x3y2
5684.9x3y3+3781.4x2y3
938.9xy3+43.3y30.4x4
+222.7x4y1630.6x4y2+2842.4x4y3
1391.5x4y4+2783.0x3y41816.9x2y
4
+
425.4
xy
4
17.4
y
4
The functionally graded beams subjected to a central line
load with elastic modulus varying in x- and y-directions (as
shown in Figs.1d and 6) could be designed following the
distribution function given in Eq.(86) as follows
6.3 Optimal design ofFGL plate subjected
toacentral line load
For the second example, we apply HBM-TO to design the
cross-section of an FGL plate with a rectangular cross-sec-
tion under a central line load. The bottom edges are fixed in
the y- direction. The geometry and boundary conditions are
presented in Fig.11. In this example, the same parameters
of the previous example are used except
𝓁
=
L
10
or
L
20
in this
example.
Figure12 presents the optimal cross-section designs of
the FGL plates obtained from HBM-TO. Similar to what
observed in the previous example, results show that the
mass is more concentrated; thus, resulting in a more defined
geometry when
𝓁
is smaller and
m
is larger. It can be seen
from Fig.13 that the elastic strain energy converges after 80
iterations and the solutions with smaller
m
converge faster.
Similar to the HBM-TO designed FGL beams, the converged
elastic strain energy is lower when
m
is smaller. The results
of optimally designed FGL plates can be classified into
two groups. The first group suggests the FGL plates could
be designed with a more distributed elastic modulus and a
smoother transition of mass which could help in preventing
failure initiations (Wang 1983; Mahamood etal. 2012) as
shown in Fig.14a, while the other group (Fig.14b) suggests
(88)
E(x,y)=fb(x,y)E0
P
y
z
x
2LDesign domain
L
4L
Fig. 11 Design domain, loading and boundary conditions for FGL
plates
ℓ=10 ℓ=20
=0.5
=1
=2
Fig. 12 Optimal designs of FGL plate cross-section obtained by HBM-TO
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 17 of 23 276
the plates could be designed with a specific shape which
could be manufactured more readily.
By least-squares fitting
𝜒
from HBM-TO FGL plates
using
𝓁
=
L
10
,m=
0.5
; and
𝓁
=
L
20
,m=
0.5
, one finds the
following elastic modulus distribution function, i.e.
where
A contour plot showing the distribution given in Eq.(89)
is presented in Fig.15.
The functionally graded plates subjected to a central line
load with elastic modulus varying along x- and y-directions
(as shown in Figs.1d and 11) could be designed following
the distribution function given in Eq.(89) as follows
(89)
fp(x,y)=
0 if
fp(x,y)<0
1 if
fp(x,y)>1
fp(x,y)otherwise
;x[0, 4],y[0, 1
]
(90)
fp(x,y)=0.896 0.178x+2.007y
26.706xy +0.359x2+33.812x2y
155.349x2y2+115.226xy211.176y2
0.157x313.567x3y+63.271x3y2
90.855x3y3+220.799x2y3156.353xy3
+14.341y3+0.020x4+1.696x4y7.909x4y
2
+11.357x4x35.195x4y4+41.558x3y4
100.542
x
2
y
4+69.703
xy
46.264
y
4
(91)
E(x,y)=fp(x,y)E0
7 Summary
Developed herein is a novel strategy to obtain optimal
designs of FGL structures. It involves the use of the Hencky
bar-grid model and topology optimization. The HBM for-
mulated herein extends a previous model with twice the
number of internal rigid bars and springs as shown in Fig.2.
The new HBM enables the determination of the local stiff-
ness matrix for each HBM grid cell by using the expressions
of the stiffnesses of elastic primary axial springs, elastic
secondary axial springs and torsional springs that are
employed to connect the rigid bars. A TO with filtering pro-
cesses is presented based on the obtained local stiffness
matrices for optimizing HBM. The objective is to minimize
the strain energy of an HBM subjected to boundary condi-
tions and equilibrium and compatibility constraints by vary-
ing spring stiffnesses within each grid cell. The obtained
HBM with optimal material property distribution is then
used to build FGL structures (see Fig.3c and d, for exam-
ple). We assume that the horizontal and vertical axial bars
have the stiffnesses of
kxx
i+1
2
,j+
1
2
and
kyy
i+1
2
,j+
1
2
, respectively.
Likewise, its curved bars located at the corners and middle
of the cell have the stiffnesses
kS
i+1
2
,j+
1
2
and
k
xy
i+1
2
,j+1
2(
orkyx
i+1
2
,j+1
2)
, respectively. The bars in all lattice cells
have a constant elastic modulus
E
and a constant density
𝜌
but a varying cross-section area
A
i+1
2
,j+
1
2
. Therefore, HBM
grid cells with larger spring stiffnesses corresponds to lattice
structures with thicker bars and thus more mass.
Fig. 13 Convergence of the
elastic strain energy when opti-
mizing the FGL plates
(a) ℓ=
10 (b)ℓ=
20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 18 of 23
8 Concluding remarks
Based on the HBM-TO optimal designs of an FGL beam
and an FGL plate under a central line load, the following
observations may be drawn:
1. HBM-TO can efficiently find optimal FGL structures by
optimizing a small number of design variables such as
the spring stiffness of each grid cell;
2. HBM-TO does not need to perform extra optimization
for the structure and connectivity of each lattice unit
cell;
3. All optimal FGL structures studied herein can be effec-
tively obtained by HBM-TO with fewer than 80 itera-
tions;
4. The designed cross-section has more concentrated mass
and a more “defined” shape when using a smaller grid
cell size or a larger penalty exponent;
5. The “distributed” solutions are generally stiffer than the
“defined” ones and with a smoother transition of mass;
6. Polynomial functions that approximate the elastic mod-
ulus distribution for the optimal designed FGL beams
(Eqs. (88)) and FGL plates (Eqs. (91)) are presented for
the first time. These functions could be useful for fab-
rication, buckling and vibration analysis of these FGL
structures as shown in Fig.1d.
In this study, we only consider one type of lattice repre-
sentation as shown in Fig.3c. A more rigorous investigation
of the HBM lattice representation such as the effect of the
shape of the curved bars and the relation between the HBM
spring stiffness and the geometry of the unit lattice structure
could be performed in a future study (Cheng etal. 2019).
Owing to its physical representation, the HBM-TO can
account for local damage and local stiffening constrains
(Wang etal. 2009; Cui etal. 2011) as well as advanced
material constitutive behaviour (O’Brien 2008) by simply
adjusting the spring stiffnesses. Moreover, as a physical
model-based optimization framework, HBM-TO can be
applied to optimize micro- and nano-structures by setting
the grid length to the characteristic length of the intended
scale. Due to HBM having similar governing equations as
the ones given by standard FD method, the efficient alternat-
ing direction implicit method could be applied for dynamic
problems (Zhang etal. 2016, 2017). The proposed design
framework can be used for other lattice models (Wieghardt
1906; Born and Karman 1912; Hrennikoff 1941; McHenry
1943; Gazis etal. 1960; Suiker etal. 2001).
Appendix A
A Python 3 code is presented for solving the problems as
articulated in Sects.6.2 and 6.3. This code is highly inspired
by an open source code from Aage and Johansen (2013). The
open source libraries “numpy” (Harris etal. 2020), “scipy”
(Virtanen etal. 2020) and “matplotlib” (Hunter 2007) are
adopted in the following code.
(a) Distributed (b) Defined
Fig. 14 Two groups of HBM-TO designed FGL plates
Fig. 15 Elastic modulus distribution for HBM-TO FGL plates with
rectangular cross-section
𝛼
=
4
under a central line load
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 19 of 23 276
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 20 of 23
Disclaimer: The authors do not guarantee that the code
is free from errors, and they shall not be liable in any event
caused by the use of the code.
Acknowledgements The authors wish to thank the review editor and
the two anonymous reviewers for their insightful suggestions and com-
ments that help to improve the paper.
Funding Open Access funding enabled and organized by CAUL and
its Member Institutions.
Declarations
Conflict of interest On behalf of all authors, the corresponding author
states that there is no conflict of interest.
Replication of results All important details have been presented in the
paper. The results obtained in this paper can be reproduced by using
the python code given in Appendix A.
Open Access This article is licensed under a Creative Commons Attri-
bution 4.0 International License, which permits use, sharing, adapta-
tion, distribution and reproduction in any medium or format, as long
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
were made. The images or other third party material in this article are
included in the article's Creative Commons licence, unless indicated
otherwise in a credit line to the material. If material is not included in
the article's Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a
copy of this licence, visit http:// creat iveco mmons. org/ licen ses/ by/4. 0/.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 21 of 23 276
References
Aage N, Johansen VE (2013) Topology optimization codes written
in Python
Andreassen E, Clausen A, Schevenels M, Lazarov BS, Sigmund O
(2011) Efficient topology optimization in MATLAB using 88 lines
of code. Struct Multidisc Optim 43:1–16. https:// doi. org/ 10. 1007/
s00158- 010- 0594-7
Azari Nejat A, Held A, Trekel N, Seifried R (2022) A modified level
set method for topology optimization of sparsely-filled and slender
structures. Struct Multidisc Optim 65:85. https:// doi. org/ 10. 1007/
s00158- 022- 03184-2
Bendsøe MP (1989) Optimal shape design as a material distribution
problem. Struct Optim 1:193–202. https:// doi. org/ 10. 1007/ BF016
50949
Born M, Karman TV (1912) Über schwingungen in raumgittern. Phys
Zeit 8:297–309
Challamel N, Zhang YP, Wang CM, Ruta G, dell’Isola F (2022) Dis-
crete and continuous models of linear elasticity: history and con-
nections. Contin Mech Thermodyn, Under Review
Challis VJ (2010) A discrete level-set topology optimization code writ-
ten in Matlab. Struct Multidisc Optim 41:453–464. https:// doi. org/
10. 1007/ s00158- 009- 0430-0
Challis VJ, Guest JK (2009) Level set topology optimization of fluids
in Stokes flow. Int J Numer Methods Eng 79:1284–1308. https://
doi. org/ 10. 1002/ nme. 2616
Cheng L, Bai J, To AC (2019) Functionally graded lattice structure
topology optimization for the design of additive manufactured
components with stress constraints. Comput Methods Appl Mech
Eng 344:334–359. https:// doi. org/ 10. 1016/j. cma. 2018. 10. 010
Cui X, Xue Z, Pei Y, Fang D (2011) Preliminary study on ductile
fracture of imperfect lattice materials. Int J Solids Struct 48:3453–
3461. https:// doi. org/ 10. 1016/j. ijsol str. 2011. 08. 013
Dunning PD, Kim HA (2015) Introducing the sequential linear
programming level-set method for topology optimization.
Struct Multidisc Optim 51:631–643. https:// doi. org/ 10. 1007/
s00158- 014- 1174-z
Ferro N, Perotto S, Bianchi D, Ferrante R, Mannisi M (2022) Design of
cellular materials for multiscale topology optimization: applica-
tion to patient-specific orthopedic devices. Struct Multidisc Optim
65:79. https:// doi. org/ 10. 1007/ s00158- 021- 03163-z
Gazis DC, Herman R, Wallis RF (1960) Surface elastic waves in cubic
crystals. Phys Rev 119:533–544. https:// doi. org/ 10. 1103/ PhysR
ev. 119. 533
Gibson LJ (1989) Modelling the mechanical behavior of cellular mate-
rials. Mater Sci and Engg: A, 110:1-36. https:// doi. org/ 10. 1016/
0921- 5093(89) 90154-8
Guerder M, Duval A, Elguedj T, Feliot P, Touzeau J (2022) Isoge-
ometric shape optimisation of volumetric blades for aircraft
engines. Struct Multidisc Optim 65:86. https:// doi. org/ 10. 1007/
s00158- 021- 03090-z
Gurtin ME (1973) The linear theory of elasticity. Linear theories of
elasticity and thermoelasticity. Springer, Berlin, Heidelberg, pp
1–295
Han Y, Lu WF (2018) A novel design method for nonuniform lattice
structures based on topology optimization. J Mech Des. https://
doi. org/ 10. 1115/1. 40405 46
Harris CR, Millman KJ, van der Walt SJ, Gommers R, Virtanen P,
Cournapeau D, Wieser E, Taylor J, Berg S, Smith NJ, Kern R,
Picus M, Hoyer S, van Kerkwijk MH, Brett M, Haldane A, del Río
JF, Wiebe M, Peterson P, Gérard-Marchant P, Sheppard K, Reddy
T, Weckesser W, Abbasi H, Gohlke C, Oliphant TE (2020) Array
programming with NumPy. Nature 585:357–362. https:// doi. org/
10. 1038/ s41586- 020- 2649-2
Hassani B, Hinton E (1999) Homogenization and structural topology
optimization. Springer, London
Hencky H (1921) Über die angenäherte Lösung von Stabilitätsproble-
men im Raum mittels der elastischen Gelenkkette. Der Eisenbau
11:437–452
Hrennikoff A (1941) Solution of problems of elasticity by framework
method. ASME J Appl Mech 8:A169–A175
Huang X, Xie YM (2007) Convergent and mesh-independent solutions
for the bi-directional evolutionary structural optimization method.
Finite Elem Anal Des 43:1039–1049. https:// doi. org/ 10. 1016/j.
finel. 2007. 06. 006
Hunter JD (2007) Matplotlib: A 2D graphics environment. Comput Sci
Eng 9:90–95. https:// doi. org/ 10. 1109/ MCSE. 2007. 55
Kaveh A, Hassani B, Shojaee S, Tavakkoli SM (2008) Structural
topology optimization using ant colony methodology. Eng Struct
30:2559–2565. https:// doi. org/ 10. 1016/j. engst ruct. 2008. 02. 012
Li D, Liao W, Dai N, Dong G, Tang Y, Minxie Y (2018) Optimal
design and modeling of gyroid-based functionally graded cellular
structures for additive manufacturing. Comput Des 104:87–99.
https:// doi. org/ 10. 1016/j. cad. 2018. 06. 003
Lin Y, Zhu W, Li J, Ke Y (2022) A distance regularization scheme
for topology optimization with parametric level sets using cut
elements. Struct Multidisc Optim 65:88. https:// doi. org/ 10. 1007/
s00158- 021- 03098-5
Liu J, Gaynor AT, Chen S, Kang Z, Suresh K, Takezawa A, Li L, Kato
J, Tang J, Wang CCL, Cheng L, Liang X, To AC (2018) Current
and future trends in topology optimization for additive manufac-
turing. Struct Multidisc Optim 57:2457–2483. https:// doi. org/ 10.
1007/ s00158- 018- 1994-3
Maconachie T, Leary M, Lozanovski B, Zhang X, Qian M, Faruque O,
Brand M (2019) SLM lattice structures: Properties, performance,
applications and challenges. Mater Des 183:108137. https:// doi.
org/ 10. 1016/j. matdes. 2019. 108137
Mahamood RM, Akinlabi ET, Shukla M, Pityana S (2012) Function-
ally graded material: an overview. In: Proceedings of the World
Congress on Engineering 2012, Vol III (WCE). London
McHenry D (1943) A lattice analogy for the solution of stress prob-
lems. J Inst Civ Eng 2:59–82
Miguel LFF, Lopez RH, Miguel LFF (2013) Multimodal size, shape,
and topology optimisation of truss structures using the Firefly
algorithm. Adv Eng Softw 56:23–37. https:// doi. org/ 10. 1016/j.
adven gsoft. 2012. 11. 006
Mlejnek HP (1992) Some aspects of the genesis of structures. Struct
Optim 5:64–69. https:// doi. org/ 10. 1007/ BF017 44697
Nguyen J, Park S, Rosen D (2013) Heuristic optimization method
for cellular structure design of light weight components. Int J
Precis Eng Manuf 14:1071–1078. https:// doi. org/ 10. 1007/
s12541- 013- 0144-5
O’Brien GS (2008) Discrete visco-elastic lattice methods for seismic
wave propagation. Geophys Res Lett 35:L02302. https:// doi. org/
10. 1029/ 2007G L0322 14
Panesar A, Abdi M, Hickman D, Ashcroft I (2018) Strategies for func-
tionally graded lattice structures derived using topology optimisa-
tion for Additive Manufacturing. Addit Manuf 19:81–94. https://
doi. org/ 10. 1016/j. addma. 2017. 11. 008
Querin OM, Steven GP, Xie YM (1998) Evolutionary structural opti-
misation (ESO) using a bidirectional algorithm. Eng Comput
15:1031–1048. https:// doi. org/ 10. 1108/ 02644 40981 02441 29
Querin OM, Young V, Steven GP, Xie YM (2000) Computational
efficiency and validation of bi-directional evolutionary structural
optimisation. Comput Methods Appl Mech Eng 189:559–573.
https:// doi. org/ 10. 1016/ S0045- 7825(99) 00309-6
Rao SS (2005) The finite element method in engineering. Elsevier
Sigmund O (1997) On the design of compliant mechanisms using
topology optimization*. Mech Struct Mach 25:493–524. https://
doi. org/ 10. 1080/ 08905 45970 89454 15
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Y.P.Zhang et al.
1 3
276 Page 22 of 23
Sigmund O (2001) A 99 line topology optimization code written in
Matlab. Struct Multidiscip Optim 21:120–127. https:// doi. org/ 10.
1007/ s0015 80050 176
Sigmund O, Petersson J (1998) Numerical instabilities in topology
optimization: A survey on procedures dealing with checkerboards,
mesh-dependencies and local minima. Struct Optim 16:68–75.
https:// doi. org/ 10. 1007/ BF012 14002
Suiker ASJ, Metrikine AV, De Borst R (2001) Dynamic behaviour of
a layer of discrete particles, part 1: analysis of body waves and
eigenmodes. J Sound Vib 240:1–18. https:// doi. org/ 10. 1006/ jsvi.
2000. 3202
Svanberg K (1987) The method of moving asymptotes—a new method
for structural optimization. Int J Numer Methods Eng 24:359–373.
https:// doi. org/ 10. 1002/ nme. 16202 40207
van Dijk NP, Maute K, Langelaar M, van Keulen F (2013) Level-
set methods for structural topology optimization: a review.
Struct Multidisc Optim 48:437–472. https:// doi. org/ 10. 1007/
s00158- 013- 0912-y
Virtanen P, Gommers R, Oliphant TE, Haberland M, Reddy T, Cour-
napeau D, Burovski E, Peterson P, Weckesser W, Bright J, van
der Walt SJ, Brett M, Wilson J, Millman KJ, Mayorov N, Nel-
son ARJ, Jones E, Kern R, Larson E, Carey CJ, Polat İ, Feng Y,
Moore EW, VanderPlas J, Laxalde D, Perktold J, Cimrman R,
Henriksen I, Quintero EA, Harris CR, Archibald AM, Ribeiro
AH, Pedregosa F, van Mulbregt P, Vijaykumar A, Bardelli AP,
Rothberg A, Hilboll A, Kloeckner A, Scopatz A, Lee A, Rokem
A, Woods CN, Fulton C, Masson C, Häggström C, Fitzgerald C,
Nicholson DA, Hagen DR, Pasechnik DV, Olivetti E, Martin E,
Wieser E, Silva F, Lenders F, Wilhelm F, Young G, Price GA,
Ingold G-L, Allen GE, Lee GR, Audren H, Probst I, Dietrich JP,
Silterra J, Webber JT, Slavič J, Nothman J, Buchner J, Kulick J,
Schönberger JL, de MirandaCardoso JV, Reimer J, Harrington J,
Rodríguez JLC, Nunez-Iglesias J, Kuczynski J, Tritz K, Thoma
M, Newville M, Kümmerer M, Bolingbroke M, Tartre M, Pak
M, Smith NJ, Nowaczyk N, Shebanov N, Pavlyk O, Brodtkorb
PA, Lee P, McGibbon RT, Feldbauer R, Lewis S, Tygier S,
Sievert S, Vigna S, Peterson S, More S, Pudlik T, Oshima T,
Pingel TJ, Robitaille TP, Spura T, Jones TR, Cera T, Leslie T,
Zito T, Krauss T, Upadhyay U, Halchenko YO, Vázquez-Baeza Y
(2020) SciPy 1.0: fundamental algorithms for scientific comput-
ing in Python. Nat Methods 17:261–272. https:// doi. org/ 10. 1038/
s41592- 019- 0686-2
Wang SS (1983) Fracture mechanics for delamination problems in
composite materials. J Compos Mater 17:210–223. https:// doi.
org/ 10. 1177/ 00219 98383 01700 302
Wang MY, Wang X, Guo D (2003) A level set method for struc-
tural topology optimization. Comput Methods Appl Mech Eng
192:227–246. https:// doi. org/ 10. 1016/ S0045- 7825(02) 00559-5
Wang G, Al-Ostaz A, Cheng AH-D, Mantena PR (2009) Hybrid lat-
tice particle modeling: theoretical considerations for a 2D elastic
spring network for dynamic fracture simulations. Comput Mater
Sci 44:1126–1134. https:// doi. org/ 10. 1016/j. comma tsci. 2008. 07.
032
Wang CM, Zhang YP, Pedroso DM (2017) Hencky bar-net model for
plate buckling. Eng Struct 150:947–954. https:// doi. org/ 10. 1016/j.
engst ruct. 2017. 07. 080
Wang CM, Zhang H, Challamel N, Pan WH (2020) Hencky bar-chain/
net for structural analysis. World Scientific, Europe
Wang J, Wu J, Westermann R (2022) Stress topology analysis for
porous infill optimization. Struct Multidiscip Optim 65:92. https://
doi. org/ 10. 1007/ s00158- 022- 03186-0
Wieghardt K (1906) Über einen Grenzübergang der Elastizität-
slehre und seine Anwendung auf die Statik hochgradig statisch
unbestimmter Fachwerke. Verhandtlungen Des Vereinz z
Beförderung Des Gewerbefleisses Abhandlungen 85:139–176
Yin L, Yang W (2001) Optimality criteria method for topology optimi-
zation under multiple constraints. Comput Struct 79:1839–1850.
https:// doi. org/ 10. 1016/ S0045- 7949(01) 00126-2
Zhang W, Yang J, Xu Y, Gao T (2014) Topology optimization of ther-
moelastic structures: mean compliance minimization or elastic
strain energy minimization. Struct Multidisc Optim 49:417–429.
https:// doi. org/ 10. 1007/ s00158- 013- 0991-9
Zhang P, Toman J, Yu Y, Biyikli E, Kirca M, Chmielus M, To AC
(2015) Efficient design-optimization of variable-density hex-
agonal cellular structure by additive manufacturing: theory and
validation. J Manuf Sci Eng. https:// doi. org/ 10. 1115/1. 40287 24
Zhang YP, Pedroso DM, Li L (2016) FDM and FEM solutions to linear
dynamics of porous media: stabilised, monolithic and fractional
schemes. Int J Numer Methods Eng 108:614–645. https:// doi. org/
10. 1002/ nme. 5231
Zhang YP, Pedroso DM, Li L, Ehlers W (2017) FDM solutions to linear
dynamics of porous media: efficiency, stability, and parallel solu-
tion strategy. Int J Numer Methods Eng 112:1539–1563. https://
doi. org/ 10. 1002/ nme. 5568
Zhang H, Wang CM, Challamel N, Zhang YP (2018a) Uncovering the
finite difference model equivalent to Hencky bar-net model for
axisymmetric bending of circular and annular plates. Appl Math
Model 61:300–315. https:// doi. org/ 10. 1016/j. apm. 2018. 04. 019
Zhang H, Zhang YP, Wang CM (2018b) Hencky bar-net model for
vibration of rectangular plates with mixed boundary conditions
and point supports. Int J Struct Stab Dyn 18:1850046. https:// doi.
org/ 10. 1142/ S0219 45541 85004 63
Zhang YP, Wang CM, Pedroso DM (2018c) Hencky bar-net model for
buckling analysis of plates under non-uniform stress distribution.
Thin-Walled Struct 122:344–358. https:// doi. org/ 10. 1016/j. tws.
2017. 10. 039
Zhang YP, Wang CM, Pedroso DM, Zhang H (2018d) Extension of
Hencky bar-net model for vibration analysis of rectangular plates
with rectangular cutouts. J Sound Vib 432:65–87. https:// doi. org/
10. 1016/j. jsv. 2018. 06. 029
Zhang H, Challamel N, Wang CM, Zhang YP (2019a) Buckling of
multiply connected bar-chain and its associated continualized
nonlocal model. Int J Mech Sci 150:168–175. https:// doi. org/ 10.
1016/j. ijmec sci. 2018. 10. 015
Zhang H, Challamel N, Wang CM, Zhang YP (2019b) Exact and nonlo-
cal solutions for vibration of multiply connected bar-chain system
with direct and indirect neighbouring interactions. J Sound Vib
443:63–73. https:// doi. org/ 10. 1016/j. jsv. 2018. 11. 037
Zhang YP, Challamel N, Wang CM, Zhang H (2019c) Comparison of
nano-plate bending behaviour by Eringen nonlocal plate, Hencky
bar-net and continualised nonlocal plate models. Acta Mech
230:885–907. https:// doi. org/ 10. 1007/ s00707- 018- 2326-9
Zhang YP, Challamel N, Wang CM (2021a) Elasticity solutions for
nano-plane structures under body forces using lattice elasticity,
continualised nonlocal model and Eringen nonlocal model. Con-
tin Mech Thermodyn 33:2453–2480. https:// doi. org/ 10. 1007/
s00161- 021- 01031-1
Zhang YP, Wang CM, Pedroso DM, Zhang H (2021b) Hencky bar-
grid model for plane stress elasticity problems. J Eng Mech
147:04021021. https:// doi. org/ 10. 1061/ (asce) em. 1943- 7889.
00019 31
Zhang YP, Wang CM, Pedroso DM, Zhang H (2022) Hencky bar-
grid model and Hencky bar-net model for buckling analysis of
rectangular plates. In: Analysis and design of plated structures.
Elsevier, pp 75–107
Zhou M, Rozvany GIN (1991) The COC algorithm, part II: topologi-
cal, geometrical and generalized shape optimization. Comput
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Optimal design offunctionally graded lattice structures using Hencky bar‑grid model and…
1 3
Page 23 of 23 276
Methods Appl Mech Eng 89:309–336. https:// doi. org/ 10. 1016/
0045- 7825(91) 90046-9
Zhu J, Zhou H, Wang C, Zhou L, Yuan S, Zhang W (2021) A review
of topology optimization for additive manufacturing: Status and
challenges. Chinese J Aeronaut 34:91–110. https:// doi. org/ 10.
1016/j. cja. 2020. 09. 020
Publisher's Note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Strain energy (SE) is considered a direct indication of the structure stiffness, as the higher the strain energy of a certain volume indicates higher stored energy under loading (lower stiffness in other words). Many lattice topology optimisation studies use SE-based analysis to guide the optimisation process (Zhang et al., 2022), (Feng et al., 2022), (Zhang et al., 2007). So, the other optimisation trial used SE (calculated by FE model) to inversely interpolate VF values for lattice unit cell corners. ...
Article
The stiffness and toughness of topology optimised structures containing lattices under three-point bending is studied. A porosity constraint is introduced to control the proportion of lattice generated while optimising the beams for minimum compliance. A novel tetrahedron element-based lattice with tapered trusses and a near-isotropic elastic response is developed to accurately map the relative densities from topology optimisation to a 3D printable structure. Topology optimised solid structures (0 % porosity), used for benchmarking, are found to have high stiffness but can be susceptible to buckling. At the other extreme, structures comprising entirely of lattice (100 % porosity) are shown to have low initial stiffness and low residual toughness. An experimental parametric study reveals that porosity can be tailored between these two bounds to achieve both high stiffness and high residual toughness.
Article
Full-text available
This paper tracks the development of lattice models that aim to describe linear elasticity of solids and the field equations of which converge asymptotically toward those of isotropic continua, thus showing the connection between discrete and continuum. In 1759, Lagrange used lattice strings/rod dynamics to show the link between the mixed differential-difference equation of a one-dimensional (1D) lattice and the partial differential equation of the associated continuum. A consistent three-dimensional (3D) generalization of this model was given much later: Poincaré and Voigt reconciled the molecular and the continuum approaches at the end of the nineteenth century, but only in 1912 Born and von Kármán presented the mixed differential-difference equations of discrete isotropic elasticity. Their model is a 3D generalization of Lagrange’s 1D lattice and considers longitudinal, diagonal and shear elastic springs among particles, so the associated continuum is characterized by three elastic constants. Born and von Kármán proved that the lattice equations converge to Navier’s partial differential ones asymptotically, thus being a formulation of continuous elasticity in terms of spatial finite differences, as for Lagrange’s 1D lattice. Neglecting shear springs in Born–Kármán’s lattice equals to Navier’s assumption of pure central forces among molecules: in the limit, the lattice behaves as a one-parameter isotropic solid (“rari-constant” theory: equal Lamé parameters, or, equivalently, Poisson’s ratio υ=1/4\upsilon =1/4). Hrennikoff and McHenry revisited the lattice approach with pure central interactions using a plane truss; the equivalent Born–Kármán’s lattice in plane stress in the limit tends to a continuum with Poisson’s ratio υ=1/3\upsilon = 1/3. Contrary to McHenry–Hrennikoff’s truss, Born–Kármán’s lattice leads to a “free” Poisson’s ratio bounded by its “limit’ bound (υ=1/4\upsilon =1/4 for plane strain or 3D elasticity; υ=1/3\upsilon =1/3 for plane stress elasticity). Unfortunately, Born–Kármán’s lattice model does not comply with rotational invariance principle, for non-central forces. The consistent generalization of Lagrange’s lattice in 3D was achieved only by Gazis et al. considering an elastic energy that depends on changes in both lengths and angles of the lattice. An alternative consistent three-parameter elastic lattice is the Hrennikoff’s, with additional structure in the cell. We also discuss the capability of nonlocal continuous models to bridge the gap between continuum isotropic elasticity at low frequencies and lattice anisotropic elasticity at high frequencies.
Article
Full-text available
The optimization of porous infill structures via local volume constraints has become a popular approach in topology optimization. In some design settings, however, the iterative optimization process converges only slowly, or not at all even after several hundreds or thousands of iterations. This leads to regions in which a distinct binary design is difficult to achieve. Interpreting intermediate density values by applying a threshold results in large solid or void regions, leading to sub-optimal structures. We find that this convergence issue relates to the topology of the stress tensor field that is simulated when applying the same external forces on the solid design domain. In particular, low convergence is observed in regions around so-called trisector degenerate points. Based on this observation, we propose an automatic initialization process that prescribes the topological skeleton of the stress field into the density field as solid simulation elements. These elements guide the material deposition around the degenerate points, but can also be remodelled or removed during the optimization. We demonstrate significantly improved convergence rates in a number of use cases with complex stress topologies. The improved convergence is demonstrated for infill optimization under homogeneous as well as spatially varying local volume constraints.
Article
Full-text available
Isogeometric shape optimisation has been studied for over a decade. The present contribution focuses on applying this approach to aircraft engine blades design by integrating geometric description, structural analysis and shape optimisation in a single industrial framework. Starting with stacked cross-sections of the blade geometry as an input, we construct a B-spline analysis-suitable volumetric model of the blade, ensuring its geometric accuracy and parametrisation regularity. Using the multi-level approach inherent to Isogeometric Analysis, we perform structural shape optimisation with the same geometric representation of the blades in the analysis and optimisation; the final optimised model being directly CAD compatible. Several self-adjoint problems are addressed using gradient-based optimisation with full analytical sensitivities that are obtained in a compact form thanks to the isogeometric framework. Examples of interest are solved using suitable industrial loading cases and objective functions. The results demonstrate the efficiency of the method and its relevance for industrial aircraft engine blades design and optimisation.
Article
Full-text available
The regularization of the level set function is important to ensure numerical stability in the level set topology optimization. For the parametric level set method, introducing the distance potential functional is a popular way to regularize the distance of level set field and bypass the common re-initialization procedure in the conventional level set method. However, the conventional distance potential functional does not have sufficient control on the gradient of level set function and therefore it may lead to the lack of smoothness and regularity for the optimization results. This paper presents a modified potential functional which incorporates the information of the interfaces into the domain of functions so that the signed distance property can be enforced directly along the interfaces. In addition, the diffusion term is employed to further smooth the structural boundaries. The Compactly Supported Radial Basis Functions (CSRBF) are used to parametrize the level set function and the cut element method is adopted to discretize the governing equations. The linear elasticity problems including compliance minimization problems and compliant mechanism problem are tested to validate the proposed method.
Article
Full-text available
In structural optimization, the level set method is known as a well-established approach for shape and topology optimization. However, special care must be taken, if the design domains are sparsely-filled and slender. Using steepest descent-type level set methods, slender structure topology optimizations tend to instabilities and loss of structural cohesion. A sole step size control or a selection of more complex initial designs only help occasionally to overcome these issues and do not describe a universal solution. In this paper, instead of updating the level set function by solving a Hamilton–Jacobi partial differential equation, an adapted algorithm for the update of the level set function is utilized, which allows an efficient and stable topology optimization of slender structures. Including different adaptations, this algorithm replaces unacceptable designs by modifying both the pseudo-time step size and the Lagrange multiplier. Besides, adjustments are incorporated in the normal velocity formulation to avoid instabilities and achieve a smoother optimization convergence. Furthermore, adding filtering-like adaptation terms to the update scheme, even in case of very slender structures, the algorithm is able to perform topology optimization with an appropriate convergence speed. This procedure is applied for compliance minimization problems of slender structures. The stability of the optimization process is shown by 2D numerical examples. The solid isotropic material with penalization (SIMP) method is used as an alternative approach to validate the result quality of the presented method. Finally, the simple extension to 3D optimization problems is addressed, and a 3D optimization example is briefly discussed.
Article
Full-text available
A flexible problem-specific multiscale topology optimization is introduced to associate different areas of the design domain with diverse microstructures extracted from a dictionary of optimized unit cells. The generation of the dictionary is carried out by exploiting micro-SIMP with AnisoTropic mesh adaptivitY (microSIMPATY) algorithm, which promotes the design of free-form layouts. The proposed methodology is particularized in a proof-of-concept setting for the design of orthotic devices for the treatment of foot diseases. Different patient-specific settings drive the prototyping of customized insoles, which are numerically verified and successively validated in terms of mechanical performances and manufacturability.
Chapter
Full-text available
This chapter presents a novel numerical framework for elastic buckling analysis of rectangular plates with rectangular cutouts by using two physical structural models, the Hencky bar-grid model (eHBM) and the Hencky bar-net model (HBM). The eHBM comprises rigid bar-grids joined by elastic primary axial, secondary axial, and torsional springs, while HBM consists of rigid bar-nets connected by elastic rotational and torsional springs. The in-plane displacements at each bar-grid joint of eHBM or the deflections at each bar-net joint of HBM can be computed by solving a set of algebraic equations obtained from minimizing the sum of strain energy of all elastic springs and the total potential energy of the external loads. The in-plane stresses can be calculated from the obtained in-plane displacements. The buckling analysis is carried out in two steps. In the first step, eHBM is used to determine the prebuckling in-plane stress distributions in a rectangular plate under any applied in-plane loads. In the second step the computed in-plane stress distribution is applied as loads at the joints of the HBM for the buckling analysis. Several buckling problems involving rectangular plates and cutouts under various in-plane load conditions and boundary conditions are solved to illustrate the convergence, accuracy, and validity of the eHBM–HBM approach. The approach features a monotonic convergence of the buckling solutions to the continuum plate solutions with respect to decreasing grid size from below. Finally, we show that the eHBM–HBM approach can be easily used to optimize the locations of cutouts in rectangular plates for maximum buckling load.
Article
Full-text available
This paper presents exact elasticity solutions for nano-plane structures subjected to any distribution of inplane body forces. In deriving the plane stress solutions, three different models are used. They are a lattice elasticity model called the Hencky bar-grid model (eHBM), the continualised nonlocal plane model (CNM) and Eringen’s nonlocal plane model (ENM). eHBM is a physical structural model comprising a system of rigid bar grids with bars connected by axial and torsional springs. CNM is a nonlocal model derived by continualising the governing discrete equations of the eHBM. ENM is a stress gradient nonlocal model. The use of three models allows independent confirmation of the solutions as well as providing a better understanding of the phenomenological similarities between them. Based on the exact solutions for a nano-plane structure under a partial uniformly inplane body force, it is found that by setting the bar grid length of eHBM to be equal to the characterisitc length \ell of CNM and ENM, the maximum inplane displacements predicted by eHBM and CNM are in exact agreement when CNM small length scale coefficient c0=1/12c_{{0}}=1/\sqrt{12} . However, the ENM maximum inplane displacements are in agreement with eHBM solutions only when ENM’s small length scale coefficient e0e_{{0}} lies between 1/501/\sqrt{50} and 1/10.1/\sqrt{10} . These results confirm some phenomenological similarities among eHBM, CNM and ENM; with CNM being closely related to the eHBM physical structure model.
Article
Full-text available
A novel Hencky bar-grid model (eHBM for brevity) is developed to address plane stress elasticity problems. This model comprises rigid bars arranged in a grid and the bars are joined by frictionless hinges, frictionless pulleys, elastic primary and secondary axial springs, and torsional springs. Based on the energy approach, the inplane displacements at the joints are determined and the inplane stress resultants are obtained from the stress resultant-displacement relations. This paper calibrates the elastic spring stiffnesses for eHBM for the first time by matching with the finite difference governing equation for elasticity problems. Some rectangular plane elasticity problems were solved by using the newly developed eHBM. The solutions obtained from eHBM converge to the exact solutions for the continuum plane body with respect to decreasing eHBM segment size. It is shown herein that eHBM can readily handle any boundary conditions and furnishes accurate solutions for plane elasticity problems with any complex geometry such as a rectangular plane body with cutouts.
Article
Full-text available
Topology optimization was developed as an advanced structural design methodology to generate innovative lightweight and high-performance configurations that are difficult to obtain with conventional ideas. Additive manufacturing is an advanced manufacturing technique building as-designed structures via layer-by-layer joining material, providing an alternative pattern for complex components. The integration of topology optimization and additive manufacturing can make the most of their advantages and potentials, and has wide application prospects in modern manufacturing. This article reviews the main content and applications of the research on the integration of topology optimization and additive manufacturing in recent years, including multi-scale or hierarchical structural optimization design and topology optimization considering additive manufacturing constraints. Meanwhile, some challenges of structural design approaches for additive manufacturing are discussed, such as the performance characterization and scale effects of additively manufactured lattice structures, the anisotropy and fatigue performance of additively manufactured material, and additively manufactured functionally graded material issues, etc. It is shown that in the research of topology optimization for additive manufacturing, the integration of material, structure, process and performance is important to pursue high-performance, multi-functional and lightweight production. This article provides a reference for further related research and aerospace applications.