You are on page 1of 12

Engineering Structures 30 (2008) 20572068

www.elsevier.com/locate/engstruct

Topology optimization of nonlinear structures under displacement loading


X. Huang , Y.M. Xie 1
School of Civil, Chemical and Environmental Engineering, RMIT University, GPO Box 2476V, Melbourne 3001, Australia
Received 12 July 2006; received in revised form 11 April 2007; accepted 17 January 2008
Available online 25 February 2008

Abstract
This paper presents topology optimization of geometrically and materially nonlinear structures under displacement loading. A revised bidirectional evolutionary optimization (BESO) method is used. The objective of the optimization problem is to maximize the structural stiffness
within the limit of prescribed design displacement. The corresponding sensitivity number is derived using the adjoint method. The original BESO
technique has been extended and modified to improve the robustness of the method. The revised BESO method includes a filter scheme, an
improved sensitivity analysis using the sensitivity history and a new procedure for removing and adding material. The results show that the
developed BESO method provides convergent and mesh-independent solutions for linear optimization problems. When the BESO method is
applied to nonlinear structures, much improved designs can be efficiently obtained although the solution may oscillate between designs of two
different deformation modes. Detailed comparison shows that the nonlinear designs are always better than the linear ones in terms of total energy.
The optimization method proposed in this paper can be directly applied to the design of energy absorption devices and structures.
c 2008 Elsevier Ltd. All rights reserved.

Keywords: Bi-directional evolutionary structural optimization (BESO); Geometrical nonlinearity; Material nonlinearity; Nonlinear finite element analysis; Energy
absorption

1. Introduction
Topology optimization methods enable designers to find the
best structural layout for a required structural performance.
Over the last three decades, many mathematical and heuristic
optimization methods have been developed [14]. So far,
most papers dealing with the optimization method have been
concerned with the optimization of structures with linear
material and geometric deformation behaviour. However, the
linear assumptions are not always valid, such as in the cases
of compliant design and energy absorption design. Using the
sensitivity/gradient based optimization methods, a number of
papers have considered topology optimization of geometrically
nonlinear structures [58]. Topology optimization of materially
nonlinear structures has also been conducted by several
researchers [912]. Jung and Gea [13] studied topology
optimization of both geometrically and materially nonlinear
Corresponding author. Tel.: +61 3 99253320; fax: +61 3 96390138.

E-mail addresses: huang.xiaodong@rmit.edu.au (X. Huang),


mike.xie@rmit.edu.au (Y.M. Xie).
1 Tel.: +61 3 99253655; fax: +61 3 96390138.
c 2008 Elsevier Ltd. All rights reserved.
0141-0296/$ - see front matter
doi:10.1016/j.engstruct.2008.01.009

structure using the generalized convex approximation (GCA).


The power-law material model and plasticity deformation
theory have been used to conduct the sensitivity analysis.
Difficulty in topology optimization of nonlinear structures
arises primarily from a large portion of the computational
overhead of the topology optimization. The computation time
in nonlinear finite element analysis depends strongly on the
size of the structures. Thus, computation efficiency is critical
especially for 3D problems. Moreover, in the mathematical
optimization method such as SIMP the material density is
penalized so that intermediate material densities invaluably
appear in the topologies. As the topology develops, large
displacements may cause the tangent stiffness matrix in
low-density elements to become indefinite or even negative
definite [5,19]. So some additional scheme must be devised
to circumvent the problem such as removing low-density
elements [19] or relaxing the convergence criterion [5]. It seems
that the bi-directional evolutionary structural optimization
(BESO) method [14,15] with discrete formulation has the
potential to overcome these problems.
BESO is an extension of evolutionary structural optimization
(ESO) [3,4] and allows for inefficient materials to be removed

2058

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

Fig. 1. (a) Typical load-displacement curve in nonlinear analysis and objective function of the optimization problem; (b) Sensitivity analysis of nonlinear structures
under displacement loading.

from the structure at the same time as the efficient ones are
added. However, the original BESO method may sometimes
lead to non-convergent and mesh-dependent solutions. Thus, it
is less robust than other mathematical optimization methods [2].
The BESO method with perimeter control [16] is shown
to be capable of obtaining somewhat convergent and meshindependent solutions because of one extra constraint (the
perimeter length) on the topology optimization problem.
However, predicting the value of the perimeter constraint
for a new design problem is a difficult task. It follows that
the original BESO method should be modified to deal with
aforementioned problems.
Consider a typical nonlinear load-displacement curve as
depicted in Fig. 1(a). If the structure is physically loaded
under force control, the displacement at the horizontal limit
point snaps through along the dashed line. Alternatively, the
trajectory due to displacement control can follow the whole
equilibrium path. Moreover, the displacement control is more
practical in applications such as the crushing distance in energy
absorption structures. So the objective of this paper deals with
optimization problems of this type of displacementloaded
nonlinear structures using a developed BESO method. To
maximize the structural stiffness within the prescribed design
displacement as shown in Fig. 1, the external work is selected
as the objective function. The sensitivity number is formulated
with energy-based functions using the adjoint variable method.
Several examples are presented to verify the proposed method.

Maximize
f (x) = W
"
= lim

#
n 

1X
T
UiT Ui1
(Fi + Fi1 )
2 i=1

Subject to g = V

M
X

Vj x j = 0

(1a)

(1b)

j=1

x j {0, 1}

(1c)

where F is the external force vector. n is the total number of


intervals of the design displacement. V j is the volume of an
individual element and V the prescribed structural volume.
The binary design variable x j denotes the absence (0) or
presence (1) of an element. M is the total element number in
the design domain.
For nonlinear analysis, the nonlinear equilibrium
R = 0 must be found using an iterative procedure whereas the
equilibrium of a linear problem is found from the solution of
a linear system of equations. The residual force, R, is defined
as the error between the internal force vector and external force
vector as
R = F Fint = 0.

(2)

The internal force vector can be expressed by


int

M Z
X

CeT B dv =

M
X

CeT Fe

2. Problem statement and nonlinear analysis

Consider a nonlinear structure subjected to a displacementload


up to a maximum, U , which is termed the design displacement,
the corresponding nonlinear forcedisplacement curve is depicted in Fig. 1(a). To maximize the structural stiffness (carried
load) at any displacement from 0 to the design displacement,
one may maximize the total external work, W , which is equal
to the total strain energy under quasi-static condition. Thus, the
optimization problem can be formulated with the volume constraint using the element as the design variable,

where Ce is a matrix which transforms the nodal force vector of


element to the globally nodal force vector and Fe is the nodal
force vector of element.
Normally, the equilibrium (2) is solved incrementally and
iteratively using the NewtonRaphson method. The details
on nonlinear analysis may be consulted in a nonlinear finite
element analysis book, e.g. [17]. The objective of this paper
is to present optimization formulation and sensitivity analysis
corresponding to nonlinear analysis.

(3)

2059

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

3. Sensitivity number
Consider a displacementload nonlinear structure, the
displacement is increased step by step up to the prescribed
displacement Un as shown in Fig. 1(b). The sensitivity of the
objective function with respect to a change in design variable x
is
"

n 
  dF
d f (x)
1X
dFi1
i
T
UiT Ui1
= lim
+
n 2
dx
dx
dx
i=1
!
#
T
n
dUi1
dUiT
1X
+

(Fi + Fi1 ) .
(4)
2 i=1 dx
dx
It is noted that the second term must be zero because the
dUT
dUT
variation of displacement dxi = 0 and dxi1 = 0 at controlled
freedoms and the external forces Fi = 0 and Fi1 = 0 at
other freedoms. An adjoint equation is introduced by adding a
series of vectors of Lagrangian multipliers i into the objective
function as
n h

1X
T
f (x) = lim
(Fi + Fi1 )
UiT Ui1
n 2
i=1
i
iT (Ri + Ri1 ) .
(5)
Because Ri and Ri1 are equal to zero, the modified objective
function (5) is same as the original objective function (1a).
Thus, the sensitivity of the modified objective function is

n 
  dF
1X
dFi1
d f (x)
i
T
T
= lim
+
Ui Ui1
n 2
dx
dx
dx
i=1


dRi1
dRi
+
.
(6)
iT
dx
dx
Within a small loading step, a linear forcedisplacement
relationship can be assumed:
Fi Fi1 = Kit (Ui Ui1 ).

dFi
dFi1
Fi
Fi1
+
=
+
.
dx
dx
x
x
Similarly, for the residual force we have

In order to eliminate the unknowns


i is chosen as

(8b)

dRi
dRi1
Ri
Ri1
+
=
+
.
(9b)
dx
dx
x
x
Substituting the above equations into Eq. (6) and utilizing Eq.
(2), Eq. (6) can be further simplified as
"

n
  F

d f (x)
1X
Fi1
i
T
T
= lim
Ui Ui1
+
n 2
dx
x
x
i=1

(10)

+ Fi1
x , adjoint variable
(11)

To verify the above sensitivity analysis equation, a simple onedimensional system is examined in the Appendix.
In the evolutionary structural optimization method, a
structure can be optimized by removing and adding elements.
That is to say that, the element itself is treated as the design
variable. Thus, when one element is totally removed from the
system, the variation of the objective function is
#
"
n 


1X
int
T
T
int
Ui Ui1 1Fi + 1Fi1 .
1 f (x) = lim
n 2
i=1
(13)
From Eqs. (2) and (3), the variation of internal force due to
removing one element can be written as
1Fiint = CeT Fie .

(14)

Substituting it into Eq. (12), the variation of the objective


function can be rewritten as
n 


1X
T
e
1 f (x) = lim
UiT Ui1
CeT Fie + CeT Fi1
n 2
i=1
n

n
X


e
E ie E i1
= E ne

(15)

i=1

where E ne is the total strain energy of the removed element. The


Eq. (15) means that the decrease of the total external work due
to removing one element is equal to the total strain energy of
the element in its final deformed state and irrelevant to the size
of displacement intervals. Thus, the sensitivity number of ith
element is defined as
i = E ni .

(9a)

Substituting i into Eq. (10), the sensitivity of the objective


function is expressed by
!
int
n 
 Fint
Fi1
d f (x)
1X
i
T
T
Ui Ui1
= lim
+
. (12)
n 2
dx
x
x
i=1

(7)

(8a)

Fi
x

!#

i = Ui Ui1 .

= lim

Thus, the variation of the external force is approximately


dFi
Fi
dUi
dUi1
=
+ Kit
Kit
dx
x
dx
dx
Fi1
dU
dUi1
dFi1
i
=
Kit
+ Kit
.
dx
x
dx
dx
Therefore,

Fint
Fint
Fi
Fi1
i +
i1
x
x
x
x

iT

(16)

Note that the sensitivity number is a direct measure of the


variation of the objective function due to element removal or
addition. To maximize the total external work, the evolutionary
process will be conducted by removing elements with the
smallest sensitivity numbers and adding the elements with the
highest ones. Mathematically such a procedure is known as the
hill-climb method or the steepest descent algorithm.
The above sensitivity number has a uniform expression for
any type of linear or nonlinear material model because the
element itself is used as the design variable. There is no surprise
that the above sensitivity number has the same expression

2060

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

as that for linear structures [4] although the formulation


process is totally different. However, this expression has
different meanings for linear and nonlinear structures. For an
elastoplastic structure, it denotes the total elastic and plastic
strain energies; while for a linear structure, it is only for the
elastic strain energy. It should also be noted that the above
sensitivity number is only the first-order approximation about
the variation of the objective function due to remove one
element.
The sensitivity number for adding material cannot be
obtained directly from the FEA results because the candidate
elements for addition are not included in the finite element
analysis. There are two existing algorithms for adding material:
linear extrapolation of the displacement field [14,16] and
adding the elements to the over-stressed region directly [15].
However, those algorithms without perimeter control would
cause mesh-dependent solutions. Here, a mesh-independent
Gaussian-weighted filter [18,19] is used to evaluate sensitivity
numbers for candidate elements for addition, although other
filters may be implemented, e.g. a linearly weighted kernel [2]
or a checkerboard suppression filter [20]. Thus, the sensitivity
number for element i is given by
N
P

i =

N
P

(17)
(ri j )

J =1

where N is the total number of elements in the mesh and (ri j )


is the weight factor given as
 
2 
exp ri2j /2 r3
,
(ri j ) =
2(r/3)


i N | ri j r , j = 1, 2, . . . , N
(18)
where ri j is defined as the distance between the centers of
the elements iand j and r is the filter radius specified by the
user. The above sensitivity filter serves two purposes: (1) to
extrapolate sensitivity number within the full design domain
and (2) to smooth the sensitivity number in the neighbourhood
of each element and prevent the problem of checkerboard
patterns and mesh-dependencies in topology optimization.
Despite the use of the mesh-independency filter, any sudden
jump in topology and objective function may cause the solution
to be non-convergent after the objective volume is satisfied. To
try to circumvent the problem we propose to further improve the
accuracy of the current sensitivity numbers by considering the
deformation history of each element. A simple way to achieve
this is to average the current sensitivity number with that of the
previous iteration as
in + in1
(19)
2
where n is the current iteration number. Then let in = i which
will be used for next iteration. Thus, the updated sensitivity
number includes all sensitivity information in the previous
iterations.
i =

Before we remove and add elements, the target volume


for the next iteration (Vi+1 ) needs to be given. Because the
objective volume (V ) can be more or less than the volume of
the initial guess design, the target volume in each iteration may
decrease or increase step by step until the objective volume is
reached. Therefore
Vi+1 = Vi (1 E R)

(i = 1, 2, 3 )

(20a)

where E R is called the evolutionary volume ratio. Once the


objective volume is reached, the volume will be kept constant
for the remaining iterations as
Vi+1 = V .

(20b)

Then sensitivity numbers of all elements, both solid (1) and


void (0) are calculated as described in Section 3. The elements
are sorted according to their values of the sensitivity number
(from the highest to the lowest). For solid elements (1), it will
be removed (switched to 0) if
th
i del
.

(21a)

For void elements (0), it will be added (switched to 1) if

(ri j )i

j=1

4. Element removal/addition criterion

th
i > add

(21b)

th and th are the threshold sensitivity numbers for


where del
add
th th . th and th
removing and adding elements and del
add
del
add
are determined by the following three simple steps:
th = th = , thus can be easily determined
1. Let add
th
th
del
by Vi+1 . For example, if there are 1000 elements in design
domain and 1 > 2 > 1000 and Vi+1 corresponds to a
design with 725 elements then th = 726 .
2. Calculate the admission volume ratio (A R), which is defined
the number of added elements divided by the total number
of elements in the current design. If A R A Rmax where
AR max is a prescribed maximum volume addition ratio, skip
th and th as in step 3.
step 3. Otherwise recalculate del
add
th by first sorting the sensitivity number of void
3. Calculate add
elements (0). The number of elements to be switched from 0
to 1 will be equal to A Rmax multiplied by the total number of
th is the sensitivity number
elements in the current design. add
th
of the element ranked just below the last added element. del
is then determined so that the removed volume is equal to
(Vi+1 Vi + the volume of the added elements).

It is noted that A Rmax is introduced to ensure that not too many


elements are added in a single iteration. Normally A Rmax is
greater than 1% so that it does not suppress the capability and
advantages of adding elements.
The new element removal and addition scheme ranks all
elements (void and solid) together, while in the original BESO
methods [14,15] elements for removal and those for addition
are treated differently and ranked separately, which is a bit
cumbersome and not very logical.
The cycle of finite element analysis and element removal
and addition continues until the objective volume (V ) is

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

2061

reached and the following convergence criterion (marginal


improvement in objective function) is satisfied:


P

N

C
C
(Wi j+1 WiN j+1 )
j=1

error =

(22)
N
P C
Wi j+1
j=1

Fig. 2. Design domain, initial guess design and support conditions for a beam.

where WiC is the complementary work for the structure in the


ith iteration, N is an integral number normally selected from
2 to 5, and is a allowable convergence error. As mentioned
above, the optimization problem for nonlinear structures may
not lead to a convergent solution. So the maximum number of
iterations, i max , should also be defined to stop the optimization
procedure.
5. Numerical implementation
The evolutionary iteration procedure of the present BESO
method is given as follows:
1. Discretize the design domain using an FE mesh for the given
boundary and loading conditions. Assign the initial property
values (0 or 1) of elements to construct initial design.
2. Perform FE analysis using commercial FEA software,
ABAQUS, on the design to obtain sensitivity information as
in Eq. (16).
3. Filter the sensitivity number to the design domain
and smooth the sensitivity information using the meshindependent Gaussian-weighted filter as in Eq. (17).
4. Average the sensitivity number using its history information
as in Eq. (19) then save the resulting sensitivity number for
next iteration.
5. Determine the target volume for the next iteration using
Eqs. (20a) or (20b).
6. Reset the property values of elements. For solid elements
(1), the property value is switched from 1 to 0 if Eq. (21a)
is satisfied. For void elements (0), the property value is
switched from 0 to 1 if Eq. (21b) is satisfied. Then construct
a new design using elements with property value 1 for the
next FE analysis.
7. Check the boundary and loading conditions for the new
design.
8. Repeat steps 27 until the objective volume (V ) is reached
and the convergence criterion, Eq. (22) or maximum
iteration is satisfied.
In step 7, any isolation of a boundary or load which changes
the nature of the problem must be avoided during the
optimization process, otherwise BESO may lead to an incorrect
solution [21]. It can be easily detected by FEA software or
BESO procedure. Once such a problem occurs, using a finer
mesh or replacing the removed elements with soft elements
to preserve the boundary or load can effectively circumvent the
problem.

Fig. 3. Evolutionary histories of external work and topology for the


optimization problem using linear finite element analysis.

6. Examples
6.1. Examples for a small design displacement
A beam with 400 mm long and 100 mm high is simply
supported at bottom corners as shown in Fig. 2. A small
displacement d = 5 mm is applied at the center of bottom
edge. It is assumed that the available material can only cover
30% volume of the design domain, and material has Youngs
modulus E = 200 GPa, Poissons ratio v = 0.3, yield stress
y = 300 MPa and plastic hardening modulus E p = 0.3E.
To save computation time, BESO starts from the initial guess
design with 30% material of the design domain which is also
shown in Fig. 2. BESO parameters are E R = 0, A Rmax = 2%,
filter radius r = 10 mm, = 0.01% and i max = 150. Thus,
only 30% of elements in the design domain are involved in finite
element analysis in each iteration.
To verify the proposed BESO method, the topology
optimization using linear finite element analysis was first
carried out to find the linear design. Fig. 3 shows the
evolutionary histories of the external work and topology. It
can be seen that the external work (non-dimensionized by the
external work of the initial guess design) gradually increases
until a convergent value is obtained in about 120 iterations.
Also, the topology also gradually develops to a stable and
mesh-independent design as shown in the last iteration. The
resultant topology is well-known and can be verified by other
mathematical optimization methods [2,22].

2062

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

Fig. 4. Evolutionary histories of external work and topology for the


optimization problem using nonlinear finite element analysis.

Fig. 5. Comparison of the applied load and external work between the linear
and nonlinear design for d = 5 mm.

Fig. 6. Evolutionary histories of external work for nonlinear optimization


problems under large design displacement (where the dotted lines indicate the
variation bounds of external work at final stage).

Fig. 7. Strainstress history for an element demonstrates that a defined


sensitivity number corresponds to two deformation stages (B and D) if the area
of OABEO is equal to OACDO.

6.2. Examples for a large design displacement


When we apply the proposed BESO method to the
above nonlinear (materially and geometrically) structure using
nonlinear finite element analysis with large strain theory, the
evolutionary histories of the external work and topology are
shown in Fig. 4. It is seen that the convergent external work and
topology are also obtained. The resultant topology is similar to
that using the linear finite element analysis due to a small design
displacement although the evolutionary topology histories are
different.
In order to sort out which design is better, the applied
force and external work versus displacement diagrams for
both designs are calculated using the nonlinear finite element
analysis and compared in Fig. 5. It shows that the nonlinear
design has larger external work at the design displacement, as
expected. Moreover, the force and external work of nonlinear
design is larger than these of linear design for most of the
deformation history. However, the nonlinear design is only
marginally better than the linear design.

The design problem definitions, with the exception of the


magnitude of the design displacement, are identical to those
presented in Section 6.1. Three different magnitudes of the
design displacement, 20, 50 and 100 mm are applied at the
center of bottom edge. The resultant evolutionary histories of
the external work are shown in Fig. 6. In general, the external
work in all cases tends to increase until it fluctuates between
two bound lines rather than converge to a constant value at the
final stage. This convergence difficulty arises due to the nature
of the problem defined in Eq. (1). First, there is no one-toone correspondence between the defined sensitivity number of
the element, Eq. (16) and its deformation state due to partial
unloading. In order to illustrate of this problem, Fig. 7 shows a
stressstrain relationship of an element. The sensitivity number
at point B is equal to that at point D if the area of OABEO
is equal to OACDO. That is to say that different deformation
states may have a same effect on the objective function, the total

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

2063

Fig. 8. Nonlinear optimized topologies for various design displacements: (a)


d = 20 mm; (b) d = 50 mm and (c) d = 100 mm.

strain energy of the structure. Second, the collapse mode of the


optimal topology especially for local buckling of members is
hard to determine beforehand although buckling is indirectly
taken into account by nonlinear finite element analysis. By
inspection, there are two collapse modes in the final stage which
are automatically found by the BESO method. Due to above
reasons, the solution is hardly convergent to a uniform solution
as in the linear or small displacement cases. In such cases,
we may select the design with the highest external work as
an optimal design. It should be noted that this type of optimal
design is sensitive to the variation of the geometry.
Fig. 8 shows the optimal designs when the design
displacement d = 20 mm, 50 mm and 100 mm respectively.
They are significantly different from linear design or small
displacementload design in Figs. 3 and 4. The members of
designs for large design displacement have variable crosssection to overcome local buckling which is considered by
nonlinear finite element analysis. However, the members of the
linear or small displacementload design have almost constant
cross-section because there is no local buckling involved. Also,
topologies for d = 20 mm, 50 mm and 100 mm have a different
layout of members which means that the nonlinear topology
depends strongly on the design displacement. Fig. 9 shows
the deformed configurations of nonlinear topologies under its
design displacement. It reveals significant differences of the
deformation mode among these designs.
Fig. 10 shows the external work-displacement diagram for
various designs. Most importantly, the nonlinear design at
its design displacement is better than other designs, e.g., the
nonlinear design for d = 50 mm has a higher external work
than the other designs when the displacement reaches 50 mm.
The improvement of the nonlinear design over the linear design
increases as a larger design displacement is used. However,
the nonlinear design at other displacement may be worse than
other designs, for example, the design for d = 50 mm has
smaller external work than the design for d = 20 mm or
the linear design when the displacement reaches 20 mm. It

Fig. 9. Deformed configurations of the nonlinear designs at its design


displacement: (a) d = 20 mm; (b) d = 50 mm and (c) d = 100 mm.

Fig. 10. Comparison of external work against displacement for various designs.

demonstrates that the objective function (external work) also


highly depends on the design displacement, and the design
displacement should be selected with caution. Fig. 11 shows the
different forcedisplacement diagrams for various designs. It
seems that the nonlinear design for a larger design displacement
has a larger buckling displacement. It is interesting to find that
the applied force of the design for d = 100 mm is much higher
than that of the design for d = 50 mm when the displacement
reaches to 100 mm although there is no significant difference
between their external works.
6.3. Examples for 3D structures
It is important to underline that most time-consuming part of
the optimization is spent on solving the equilibrium equations

2064

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

Fig. 11. Comparison of applied load against displacement for various designs.

both for linear and nonlinear structures. Thus it is critical for


large problems, especially in nonlinear and 3D structure, to
improve on the efficiency of the optimization method. Because
the analysis time depends strongly on the size of model, only
a portion of elements in the design domain is involved in
the finite element analysis in each iteration as in the above
examples is an efficient way to save the computation time. To
further demonstrate the efficiency of the proposed method, the
following examples on 3D optimization problem is given. The
following examples are conducted on a normal computer with
3.0 GHz CPU, and 1GB RAM.
The 3D design domain is shown in Fig. 12 and divided
into 100 10 25 mesh using 8-node brick elements. The

prescribed displacement d = 50 mm is applied on three points


as shown in Fig. 12. The nonlinear material is approximated
with the well-known RambergOsgood plasticity model with
Youngs modulus E = 20 GPa, Poissons ratio v = 0.4, the
yield stress y = 40 MPa, the yield offset = 0.002 and
the hardening exponent n = 3. The volume objective is only
cover 20% of the design domain and the initial guess design
with 5000 elements (which is satisfied the volume constraint) is
also shown in Fig. 12. The BESO parameters are E R = 0,
A Rmax = 2%, r = 10 mm, = 0.1% and i max = 100.
Thus, the topology develops without the change of total element
number.
The optimal designs from linear and nonlinear finite element
analyses are presented in Fig. 13(a) and (b) and show quite
different shapes. The total iterations and computation time are
40 and 74.3 min for linear optimization problem, and 29 and
107.3 min for nonlinear optimization problems respectively. In
this case, obtaining the nonlinear optimal design only takes 1.5
to 2 times the computation time to obtain the linear design.
We tested a single finite element analysis on the full design
and found that the analysis time was 2.5 min for the linear
analysis, however, more than 4 hours for the nonlinear analysis
with 40 displacement steps and 162 local iterations. It shows
the computational efficiency of the proposed method is obvious
especially for nonlinear optimization problems.
To further investigate these designs, we applied nonlinear
analysis to these two designs and the applied force (sum
of reaction forces at three points) and external work against
displacement are shown in Fig. 14. It can see that the linear
design is a little better than nonlinear design at the small
displacement range, and the nonlinear design is much better

Fig. 12. Design domain, initial guess design and support conditions for a 3D structure.

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

2065

The applied force of the nonlinear design is 44.2 kN which


is much higher than that of the linear design, 37.5 kN. The
improvement on the external work is only about 7.5%. So
for a small design displacement, a nonlinear design can be
replaced by a linear design unless it can not satisfy the
requirement of the application. Comparing with the initial guess
design with applied force, 25.4 kN and external work, 0.68 kJ,
the performance of the linear and nonlinear design has been
improved greatly.
7. Discussions and conclusions

Fig. 13. Comparison of topologies between linear and nonlinear design: (a)
linear design; (b) nonlinear design.

Fig. 14. Comparison of applied load and external work between the linear and
nonlinear design.

than the linear design at the large displacement. When d =


50 mm, the total external work of the nonlinear design, 1.14 kJ,
is higher than the linear design with the external work, 1.06 kJ.

In this paper, a modified BESO method for topology


optimization of nonlinear structures under a prescribed design
displacement has been proposed. In the original BESO method,
elements for removal and those for addition are treated
differently and ranked separately, which is a bit cumbersome
and not very logical. Due to short of length-scale constraint,
its solution may be mesh-dependent and non-convergent [2].
To improve the robustness of the optimization method, the new
BESO algorithm considers the sensitivity history and includes
a filter scheme. The new element removal and addition scheme
ranks all elements (void and solid) together. Thus, a meshindependent and convergent solution for linear optimization
problems has been obtained which verified the effectiveness of
the proposed BESO method.
The proposed BESO method has been applied to several
examples for topology optimization with both geometrical and
material nonlinearities. Due to partial unloading of material,
the defined optimization problem for nonlinear material is illposed and the solution may not always be convergent because
of oscillation between designs of two different deformation
modes. However, significant improvements over the initial
designs can be obtained using the proposed method. For the
complex problem of optimizing nonlinear structures, a much
improved design will be of practical importance rather than an
absolute optimum which could be only slightly better than the
improved design but may well be practically impossible to find.
The examples illustrate substantial differences in the topologies of the linear and nonlinear optimal designs. Unlike the linear design which is independent of the design displacement, the
nonlinear optimal design is highly dependent on the specified
design displacement. The results show that the nonlinear optimal design at its design displacement is always better than the
linear design in terms of the total energy, which is the objective
function. The improvement becomes more and more significant
as the magnitude of the design displacement increases.
The main advantage of the proposed BESO method is
its high computational efficiency because only a portion of
elements in design domain needs to be calculated in the finite
element analysis. Furthermore, the intermediate designs in the
BESO method have no low density elements the grey
area. This significantly improves the stability of nonlinear
finite element analysis.
The presented method may be directly applied to the design
of energy absorption devices and structures. The goal of the
energy absorption design is to maximize the absorbed energy

2066

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

Fig. 15. One-dimensional single degree-of-freedom two bar system.

within a prescribed crushing distance, which retains a minimum


volume, for survival of, for example, the occupants of a motor
vehicle. This would require the structure to have maximum
stiffness within the crushing distance as stated in Eqs. (1a)
(1c). Certainly, some other constraints such as the maximum
crushing force should be added in the energy absorption design.
The authors intend to explore this further in the future.
Acknowledgment
The authors wish to express their gratitude to the Australian
Research Council for the financial support for carrying out this
work.
Appendix
A one-dimensional system as shown in Fig. 15 is studied to
verify the sensitivity analysis, Eq. (12) for small deformation
elastoplasticity. The system is discretized with two bar
elements having length L and cross section A. The prescribed
displacement, u 2 , is applied at the free end. For simplicity,
element 1 consists of elastic material having the Young modulus
E and element 2 consists of elastoplastic material having the
Young modulus E, the yield stress Y and the linear hardening
modulus E p . Here, is chosen as the design variable. Three
loading steps 0 u 2A u 2B u 2C are assumed and the
corresponding stressstrain responses of the two bars are also
shown in Fig. 15.
A.1. Loading from 0 to u 2A
In this stage, both elements undergo elastic deformation and
the forces in element 1 and 2 are given by

EA
u 1
L
EA
(u 2 u 1 ).
F2 =
L
F1 =

Setting F1 = F2 , leads to:


u1 =

1
u2.
1+

Thus, the external work is equal to the incremental strain energy


of the system as
f 1 () =

1
1
F1A u 1A + F2A (u 2A u 1A ).
2
2

Then the derivative of f 1 () with respect to a change in design


variable is
d f 1 ()
1
=
d
2


F1A
F1A du 1A
1
du 1A
+
u 1A + F1A

u 1A d
2
d


1 F2A
F2A du 1A
+
+
2

u 1A d
du 1A
1
(u 2A u 1A ) F2A
2
d
EA
E A
2
=
u
u2
2L(1 + )3 2A 2L(1 + )3 2A
E A
E A
+
u 22A +
u2
3
2L(1 + )
2L(1 + )3 2A
EA
=
u2 .
2L(1 + )2 2A


2067

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

Ep
Ep
(u 2A + u 2B ) + Y 1
2L
E

Ep
EEp A
=
(u 2A + u 2B )
(E p + E)2 2L


Ep
+ Y 1
(u 2B u 2A ).
E

On the other hand,




1 F1A
F2A
EA 2
u 1A +
(u 2A u 1A ) =
u +0
2

2L 1A
EA
u2 .
=
2L(1 + )2 2A

Thus,


d f 1 ()
1 F1A
F2A
=
u 1A +
(u 2A u 1A ) .
d
2

(A.1)

A.2. Loading from u 2A to u 2B


In this stage, the plastic deformation occurs in element 2.
The forces in both elements are expressed by
F1 =

EA
u 1
L

F2 = A Y + E p

u2 u1
Y

L
E



Thus,
d f 2 ()
1
=
d
2

The external work in this step is equal to the incremental strain


energy of the system as

A.3. Unloading from u 2B to u 2C

1
(F1A + F1B )(u 1B u 1A )
2
1
+ (F2A + F2B )(u 2B u 1B u 2A + u 1A ).
2

Then, the derivative of f 2 () yields




1 F1A
F1B
F1A du 1A
F1B du 1B
d f 2 ()
=
+
+
+
d
2

u 1A d
u 1B d


1
du 1B
du 1A
(u 1B u 1A ) + (F1A + F1B )

2
d
d


1 F2A du 1A
F2B du 1B
+
+
2 u 1A d
u 1B d
(u 2B u 1B u 2A + u 1A )


du 1A
du 1B
1

+ (F2A + F2B )
2
d
d

2
E Ep A
Ep
=
(u 2A + u 2B )
(E p + E)3 2L


E E 2p A
Ep
+ Y 1
(u 2B u 2A )
E
(E p + E)3



Ep
Ep
(u 2A + u 2B ) + Y 1
(u 2B u 2A )

2L
E

E E 2p A
Ep
+
(u 2A + u 2B )
3
(E p + E) 2L


Ep
E2 E p A
+ Y 1
(u 2B u 2A ) +
E
(E p + E)3

(u 2B u 2A )

On the other hand,




F1B
1 F1A
(u 1B u 1A )
+
2



F2B
1 F2A
(u 2B u 2A u 1B + u 1A )
+
+
2

EA
=
(u 1A + u 1B )(u 1B u 1A ) + 0
2L

Ep
EEp A
=
(u 2A + u 2B )
2
(E p + E) 2L


Ep
+ Y 1
(u 2B u 2A ).
E

Setting F1 = F2 , gives



Ep
Ep
L
u 2 + Y 1
.
u1 =
E p + E L
E

f 2 () =



F1A
F1B
+

(u 1B u 1A ).

(A.2)

In this loading step, elastic unloading occurs in both


elements and the forces in both elements are
F1 =

EA
u 1
L

F2 = A 2B E

u2 u1
L



Here, it should be noted that both u 1 and 2B are the function


of the design variable, . Setting F1 = F2 gives
u1 =

1
L
(u 2 2B ).
1
E

The external work in this step is


f 3 () =

1
(F1B + F1C )(u 1C u 1B )
2
1
+ (F2B + F2C )(u 2C u 1C u 2B + u 1B ).
2

The derivative of f 3 () yields




d f 3 ()
1 F1B
F1C
F1B du 1B
F1C du 1C
=
+
+
+
d
2

u 1B d
u 1C d


1
du 1C
du 1B
(u 1C u 1B ) + (F1B + F1C )

2
d
d


1 F2B du 1B
F2C du 1C
+
+
(u 2C u 1C
2 u 1B d
u 1C d


1
du 1B
du 1C
u 2B + u 1B ) + (F2B + F2C )

2
d
d

2068

X. Huang, Y.M. Xie / Engineering Structures 30 (2008) 20572068

EA
u 2B + u 2C
L
2B (u 2C u 2B )
2
E
L(1 )3
A d B

(u 2C u 2B )
(1 )2 d


E A
L
u 2B + u 2C
+

2B
2
E
L(1 )3


u 2B + u 2C
E A
L
(u 2C u 2B )

2B
2
E
L(1 )3
A d B
(u 2C u 2B )
(u 2C u 2B ) +
(1 )2 d


u 2B + u 2C
E A
L

2B (u 2C u 2B )
2
E
L(1 )3


EA
u 2B + u 2C
L
2B (u 2C u 2B ).
=
2
E
L(1 )2
=

On the other hand,




F1C
1 F1B
+
(u 1C u 1B )
2



1 F2B
F2C
+
+
(u 2C u 2B u 1C + u 1B )
2

EA
(u 1B + u 1C )(u 1C u 1B ) + 0
=
2L


EA
L
u 2B + u 2C
=
2B (u 2C u 2B ).
2
E
L(1 )2
Thus,
d f 3 ()
1
=
d
2

F1B
F1C
+

(u 1C u 1B ).

(A.3)

In conclusion, the derivative of the total external work f () =


f 1 () + f 2 () + f 3 () can be expressed by
!
int
3
Fi1
Fiint
1X
d f ()
T
T
=
(U Ui1 )
+
.
(A.4)
d
2 i=1 i

References
[1] Bendse MP, Kikuchi N. Generating optimal topologies in structural
design using a homogenization method. Computer Methods in Applied
Mechanics and Engineering 1988;71:197224.
[2] Bendse MP, Sigmund O. Topology optimization: Theory, methods and
applications. Berlin, Heidelberg: Springer-Verlag; 2003.
[3] Xie YM, Steven GP. A simple evolutionary procedure for structural
optimization. Computers and Structures 1993;49:8856.

[4] Xie YM, Steven GP. Evolutionary structural optimization. London:


Springer; 1997.
[5] Buhl T, Pedersen CBW, Sigmund O. Stiffness design of geometrically
nonlinear structures using topology optimization. Structural and
Multidisciplinary Optimization 2000;19:93104.
[6] Gea HC, Luo J. Topology optimization of structures with geometrical
nonlinearities. Computers and Structures 2001;79:197785.
[7] Pedersen CBW, Buhl TE, Sigmund O. Topology synthesis of largedisplacement compliant mechanisms. International Journal for Numerical
Methods in Engineering 2001;50:2683705.
[8] Bruns TE, Sigmund O, Tortorelli DA. Numerical methods for
the topology optimization of structures that exhibit snap-through.
International Journal for Numerical Methods in Engineering 2002;55:
121537.
[9] Yuge K, Kikuchi N. Optimization of a frame structure subjected to a
plastic deformation. Structural and Multidisciplinary Optimization 1995;
10:197208.
[10] Bendse MP, Guedes JM, Plaxton S, Taylor JE. Optimization of structure
and material properties for solids composed of softening material.
International Journal of Solids and Structures 1996;33(12):1799813.
[11] Pedersen P. Some general optimal design results using anisotropic, power
law nonlinear elasticity. Structural Optimization 1998;15:7380.
[12] Mayer RR, Kikuchi N, Scott RA. Application of topological optimization
techniques to structural crashworthiness. International Journal for
Numerical Methods in Engineering 1996;39:1383403.
[13] Jung D, Gea HC. Topology optimization of nonlinear structures. Finite
Elements in Analysis and Design 2004;40:141727.
[14] Yang XY, Xie YM, Steven GP, Querin OM. Bidirectional evolutionary
method for stiffness optimization. AIAA Journal 1999;37(11):14838.
[15] Querin OM, Young V, Steven GP, Xie YM. Computational efficiency
and validation of bi-directional evolutionary structural optimisation.
Computer Methods in Applied Mechanics and Engineering 2000;189:
55973.
[16] Yang XY, Xie YM, Liu JS, Parks GT, Clarkson PJ. Perimeter control
in the bidirectional evolutionary optimization method. Structural and
Multidisciplinary Optimization 2003;24:43040.
[17] Crisfield MA. Non-linear finite element analysis of solids and structures.
New York: Wiley; 1991.
[18] Murio DA. The mollification method and the numerical solution of Illposed problems. New York: Wiley; 1993.
[19] Burns TE, Tortorelli DA. An element removal and reintroduction strategy
for the topology optimization of structures and compliant mechanisms.
International Journal for Numerical Methods in Engineering 2003;57:
141330.
[20] Li Q, Steven GP, Xie YM. A simple checkerboard suppression algorithm
for evolutionary structural optimization. Structural and Multidisciplinary
Optimization 2001;22:2309.
[21] Zhou M, Rozvany GIN. On the validity of ESO type methods in topology
optimization. Structural and Multidisciplinary Optimization 2001;21:
803.
[22] Sigmund O. A 99 line topology optimization code written in Matlab.
Structural and Multidisciplinary Optimization 2001;21:1207.

You might also like