Está en la página 1de 8

Struct Multidisc Optim (2010) 42:2532

DOI 10.1007/s00158-009-0473-2

RESEARCH PAPER

Topology optimization of truss-like continua with different


material properties in tension and compression
Osvaldo Maximo Querin Mariano Victoria
Pascual Mart

Received: 3 August 2009 / Revised: 10 December 2009 / Accepted: 10 December 2009 / Published online: 7 January 2010
c Springer-Verlag 2009


Abstract The literature on the topology optimization of


truss-like continua has relatively few publications that look
at the problem of structures with different material properties/behaviour in tension and compression. After reviewing the literature on optimality criteria, ground structure
approach, growth method for truss structures, SIMP,
homogenization and heuristic methods of topology optimization, a method is proposed in this paper which consists
of identifying all negative (compressive) parts of a structure; modifying their modulus of elasticity; changing their
material properties to orthotropic, modifying the stresses or
strains and satisfying the optimality criterion of a constant
stress/strain ratio between the tensile and compressive parts
of the structure. The validity of this technique is demonstrated on several examples, including comparisons with
results by other methods
Keywords Dual material Stress constraints
Topology optimization Trusses

O. M. Querin (B)
School of Mechanical Engineering,
The University of Leeds, Leeds LS2 9JT, UK
e-mail: O.M.Querin@Leeds.ac.uk
M. Victoria P. Mart
Structures and Construction Department,
Technical University of Cartagena, Campus Muralla del Mar,
30202 Cartagena, Murcia, Spain
M. Victoria
e-mail: mariano.victoria@upct.es
P. Mart
e-mail: pascual.marti@upct.es

1 Introduction
Prager (Prager and Rozvany 1977) introduced the terms
truss-like continua for structures in which in some
regions members of infinitesimal cross-sectional area have
an infinitesimal spacing. Such structures are for example
Michell frames (Michell 1904). As Prager has declared,
while this type of structure is not practical, it is useful in
the evaluation of the efficiency of more practical designs
In the literature on topology optimization of truss-like
continua, relatively little attention has been paid to problems where the material behaves differently in tension and
compression. By behaviour we mean different stress limits or elastic modulus. There have been several approaches
using optimality criteria, ground structure, growth method,
SIMP, homogenization and heuristics to solve this problem.
The work of Michell (1904) has set the foundation for
research into the optimal layout of trusses, with materials
of the same or different strength in tension and compression. Michell proved that an optimum truss must follow
the orthogonal network of lines of maximum and minimum
strain in a constant magnitude strain field. This research was
further advanced by Hemp (1958, 1973) and Prager (1958).
Rozvany et al. (1993) showed that the optimal topologies
for plastic stress design and elastic compliance design are
the same, and the volume or weight of the latter is given
by the square of the volume of the former (multiplied by
a given constant), for a review see Rozvany et al. (1995),
p. 57.
Rozvany (1996, 1997) demonstrated that there was an
error in Michells (1904) derivation of optimality criteria,
and that the latter are only valid for a restricted class of
problems. Using three different methods, he derived the correct optimality criteria (with adjoint strains t = 1/t and
c = 1/c ), which confirmed those of Hemp (1973) and

26

demonstrated on a simple example that the new optimality criteria resulted in a much lower structural volume than
that of Michell. He also sketched the correct Michell cantilever with different permissible stresses in tension and
compression and gave a simple expression for the optimal
angle of members next to a line support (for a stress ratio of
1:3 these are 30 and 60 degrees).
Srithongchai and Dewhurst (2003), however expressed
appreciation for the early contribution of Prager (1958), in
which he established the optimality criteria that the strain
energy per unit volume, or per unit weight needed to be
constant in a structure with materials of different strength
in tension and compression. Their paper, like Rozvanys
compares Michell and what they called Prager structures.
Selyugin (2004) proposed a slight variation on the
adjoint strain optimality criteria of Hemp (1973) and
Rozvany (1996) for the optimization of bi-material structures with maximum stiffness. In it he proposed that the
layout should follow lines of maximum strains of magnitude k(t /t ) and k(c /c ), where k is a small positive
number but that the structural members should be designed
to all have equal strain energy per unit mass.
Dewhurst (2005) demonstrated that in order to achieve
an optimal layout of maximum stiffness and minimum volume/weight, the structural members should have a constant
strain with a fixed ratio between the tensile and compressive
members which is a function of their elastic moduli and densities. By achieving this, the structure would have a constant
strain energy per unit weight in all the members
Although these papers all dealt with structural layouts,
they formed the basis for the work of Nair (2005) and
Taggart et al. (2007) who extended this to continuum structures. They used a node based approach to carry out the
optimization, which is analogous to the Hard-Kill (Hinton
and Sienz 1995) or ESO (Xie and Steven 1993) methods,
and used the strain energy density as the optimality criteria.
The method used to incorporate the different stress limits in
tension and compression was based on modifying the modulus of elasticity of the material experiencing compressive
or biaxial stresses, whilst assuming an isotropic material
property, allowing an isotropic analysis of the structure.
Achtziger (1996) used the ground structure approach
coupled with a standard Linear programming (LP) solver to
derive the optimum truss topology with different stress limits in the tensile and compressive members. More recently
this has been extended to problems with up to 109 truss
elements in the ground structure by Gilbert et al. (2005),
which has resulted in a software application, Darwich et al.
(2008). Alternatively, Martinez et al. (2007) proposed a
growth method for the optimal design in a sequential manner of size, geometry and topology of plane trusses without
the need of a ground structure. This method is applicable
to single load case problems with stress and size constraints

O.M. Querin et al.

and to the problem of different stress limits in tension and


compression members.
Duysinx (1999) was the first to look at this problem for
continuous structures using SIMP (Zhou and Rozvany 1991;
Rozvany et al. 1992), for a review see Rozvany (2001 or
2009). He utilised the Raghava (Raghava et al. 1973) and
Ishai (Geli et al. 1981) quadratic criteria which predicts a
different behaviour in tension and compression. The application of these criteria was further studied by Duysinx and
Bruyneel (2000), Pereira et al. (2004), Rong et al. (2005),
and more recently Duysinx et al. (2008).
Desmorat and Duvaut (2003) extended classic compliance optimization to problems with nonlinear elastic materials with different behaviours in tension and compression.
They applied this process to long curvilinear fibre composite single ply membranes, where the orientation and the fibre
density are the local design parameters.
Guan et al. (1999, 2003) looked at the optimization of
what they called material-oriented structures, which use different materials to resist the tensile and compressive forces.
To do this, they used the principal stresses in each element coupled with the ESO method (Xie and Steven 1993).
This was further modified by Alfieri et al. (2007) who used
the concept of tension and compression dominated material,
where each individual element was considered to be dominated in tension or compression depending on the maximum
magnitude of the principal stress of that element.
Remodelling theories, which uses a macro model to represent fundamental properties of the bone have also been
applied to this problem. In Plfi (2004) the material orientation was aligned to be parallel with the principal stress directions in each finite element of the structure. Instead of this,
Nowak (2006) used the strain energy density (SED) distribution as the remodelling criterion. Material was moved
onto the surface of the structure to keep a constant volume.
The aim was to obtain an SED level on the surface and to
fulfil the minimum compliance assumption for the stiffest
design. With Cai and Shi (2008) being able to successfully
incorporate the effect of different mechanical properties in
tension and compression.
Graczykowski and Lewinski (2006a, b, 2007a, b) presented exact analytical solutions for Michell cantilevers
with different permissible stress in tension and compression in four long papers. Lewinski and Rozvany (2007,
2008a, b) published exact analytical solutions for threesided, L-shaped and square supports.
The aim of this paper is to propose an approach which,
by combining the works of Plfi (2004), Dewhurst (2005),
and Nair (2005), allows for the topology optimization of
continuous structures with different material properties in
tension and compression. This is achieved by converting
the material properties (where required) to orthotropic ones,
modifying the modulus of elasticity and stresses/strains of

Topology optimization of truss-like continua with different material properties in tension and compression

27

the compressive elements within the structure and satisfying the optimality criterion of a constant stress/strain ratio
in the structure. Three examples will be presented to show
the validity of the proposed approach.

2. The use of the same material for the tensile and compressive elements of the structure.

2 The optimality criterion

Since the material properties ( and E) are known for the


two materials, these are then used in the analysis of the
structure. Once the strains or stresses have been calculated
in each element, all of the negative terms are then multiplied by the factor k or k1 , which have been calculated using
the known material properties. The resulting strain or stress
field is then used for the optimization.

From Dewhurst (2005), an optimal layout of maximum


stiffness and minimum volume/weight is obtained if all
the structural members have a constant fixed strain ratio
given by
k=

T
=
C

T E C
C E T

1
2

(1)

where:

Case 1: Different Material for tensile and compressive


elements

Case 2: The Same Material for tensile and compressive


elements

constant fix ratio between strains in the optimum


structure
tensile and compressive maximum strains defining
optimal structure layouts
elastic moduli of tensile and compressive members
mass, weight, or cost per unit volume of tension
and compressive members

Since the same material is used throughout the structure,


the ratio of densities in (1) or (2) can be set equal to unity.
The factor k or k1 is specified and a fictitious modulus of
elasticity is calculated for the compressive elements of the
structure using

Depending on the method of analysis, it may be more convenient to work with stresses rather than strain. Converting
(1) by dividing by the respective modulus of elasticity gives
the equivalent equation, which relates the tensile and compressive stresses in the optimal structure by a new factor k1 :

By means of these moduli, the structure is analyzed and all


negative strains or stresses are multiplied by the factor k or
k1 . The resulting strain or stress field is then used for the
optimization.

k
T , C
ET , EC
T , C

k1 =

T
C


=

T E T
C E C

1
2

(2)

where:
k1
T , C

constant fix ratio between stresses in the optimum


structure
tensile and compressive maximum stresses in the
optimal structure

In order to obtain an optimum using the optimality criteria


of Dewhurst (2005), a fully stressed design solution of the
modified strain or stress field in (1) or (2) is required.

3 The modified analysis process


Two separate types of problems require consideration:
1. The use of different materials for the tensile and compressive elements of the structure (e.g. steel-reinforced
concrete), or

EC =

ET
= ET k2
k12

(3)

3.1 Orthotropic analysis


Since the structure has different real or artificially generated mechanical properties for its tensile and compressive elements, the analysis has to capture this. For a two
dimensional (2D) or three dimensional (3D) continua, an
orthotropic analysis needs to be carried out. However, when
the process is started it is impossible to determine which
components of the structure experiences tensile or compressive stresses or both. The following nine step analysis
process is required:
1. The entire design domain is given only the property of
the tensile material;
2. Carry out a structural analysis;
3. The design domain will then be under a biaxial (2D) or
triaxial (3D) stress state, for which the principal stresses
(1 , 2 and 3 for 3D) and their orientations can be calculated (Timoshenko and Goodier 1982). For the 3D
continuum case, three stress states can exist: (1) All
stresses are positive (1 > 2 > 3 > 0), (2) All
stresses are negative (0 > 1 > 2 > 3 ) and (3) One

28

4.
5.
6.
7.

8.

9.

O.M. Querin et al.

or more of the stresses are negative (1 > 2 > 0 > 3 )


or (1 > 0 > 2 > 3 ). For all elements satisfying
case: (1) the material properties are set as isotropic and
have the values of the tensile material, (2) the material
properties are set as isotropic and have the values of the
compressive material, and (3) the material properties set
as orthotropic with the directions of positive stress given
the values of the tensile material and the directions of
negative given the values of the compressive material.
Carry out a structural analysis with the new quasiorthotropic material properties;
Calculate the principal stresses and their orientation.
All of the principal stresses which are negative are
multiplied by the factor k1 ;
Carry out the optimization step on the structure using
the modified negative stresses together with the unmodified positive stresses;
The principal stresses and their orientations calculated
in step 5 are used to determine the new orthotropic (or
isotropic) properties for each element of the structure;
go to step 4 until the optimization process ends.

At the initial stages of this research, two finite element analyses were carried out after every optimization step. The
first to determine the sign and orientation of the principal
stresses and calculate the orthotropic properties for each element, the second to determine the principal stresses to use
for the optimization. However; apart from the first iteration;
it was found that changes in the structural domain due to
the optimization step have a negligible effect in the changes
of the direction and sign of the principal stresses. Hence
only one finite element analysis between optimization steps
is required.

where:
M
F
L
M Optimal

dimensionless weight or mass of a structure


applied force on the structure
horizontal length of the structure from the supports to the applied force
optimal mass or weight of the structure

If the same material is used for the tension and compression


members then this equation reduces to (5).
M =

VOptimal


F L 1T + 1C

where:
VOptimal

optimal volume of the structure

And if the average von Mises stress level is used, (5)


simplifies to
M =

VOptimal Av
2F L

Av the average stress level in the structure,


Three examples are presented to demonstrate the validity
of the proposed method: Short Prager Cantilever, Prager
Cantilever, and the Cantilever with circular support. For all
examples the modulus of elasticity used was E = 210,000
MPa. Rozvany (2009) suggests the use of a Poissons ratio

The method used to carry out the optimization in this work


was the Isolines Topology Design (ITD) method of Victoria
et al. (2009), where the von Mises stress is the design criterion. The reason for using it was that, although it uses a
finite element mesh, the emerging topology is very smooth
and the method has the capability to reintroduce or significantly redistribute material if the removal process worsened
the design.
For comparison with results in the literature, the final
weight of the structure was non-dimensionalised by dividing
it by the term F L (T /T + C /C ) used by Srithongchai
and Dewhurst (2003) to give
M Optimal


F L TT + CC

(6)

where:

4 Examples

M =

(5)

(4)
Fig. 1 Design domain for the Short Prager Cantilever

Topology optimization of truss-like continua with different material properties in tension and compression

Fig. 2 Optimal topology for the Short Prager Cantilever: a using the
method of Martinez et al. (2007), b this work

value of zero for getting solutions close to Michell structures. However, when this was tried there were problems
with the orthotropic elements so a Poissons ratio of v = 0.3
was used.
4.1 Short Prager Cantilever
This example given by Srithongchai and Dewhurst (2003)
represents a short cantilever of height W and width L with
a vertical force F acting along a distance L from the support
and a vertical distance D below the top of the support in
Fig. 1. Where the height to width ratio W/L = 1:0.5995.
In order to capture accurately the angles of the emerging
truss elements in the topology such that these are properly
aligned between the supports and load, a border around the
true design domain was placed here denoted by lengths B1
and B2 .
The dimensions used for this example are W = 1,000 mm
and L = 599.5 mm, represented by the dashed rectangle.
Fig. 3 Design domain for the
Prager Cantilever

29

The lengths of the border were B1 = 28.57143 mm and


B2 = 17.12857 mm. The entire design domain (solid outer
rectangle) was subdivided with a mesh of 370 222 elements. The distance to the vertical force F = 10 kN is D =
625 mm with the thickness of the domain t = 1 mm. The
constant fix ratio between stresses was k1 = 3.
Srithongchai and Dewhurst (2003) obtained the nondimensional weight of M = 0.9358 for the discrete 4
elements truss solution to this problem. The same problem was also optimised using the method of Martinez
et al. (2007), producing the layout of Fig. 2a with a nondimensional weight of M = 0.9351. The optimal topology
obtained using the method proposed in this paper is given in
Fig. 2b, with a resulting volume VOptimal = 69,341 mm3 ,
an average stress level Av = 173.5 MPa giving a nondimensional weight of M = 1.0034, which is 7.22%
greater than that obtained by Srithongchai and Dewhurst
(2003) and 7.30% greater than produced by the method of
Martinez et al. (2007). The resulting topology is very similar to that of the truss-produced optimum, and as expected,
the non-dimensional weight is slightly higher.
4.2 Prager Cantilever
This example, solved by Srithongchai and Dewhurst (2003)
for different tensile and compressive stress limits, was originally presented by Prager (1958) as a challenge to the reader
to solve its optimum characteristics. The original boundary
layout proposed by Prager is shown as the dashed rectangle
in Fig. 3. As in the previous example, a border denoted by
lengths B1,2,3,4 was placed around the design domain.
The dimensions used for this example are W = 1,000 mm
and L = 3,000 mm, represented by the dashed rectangle.
The lengths of the border were B1 = 310.0285 mm, B2 =
355.3001 mm, B3 = 77.4571 mm and B4 = 38.7285. The
entire design domain (solid outer rectangle) was subdivided
with a mesh of 215 402 elements. The distance to the vertical force F = 10 kN is D = 500 mm with the thickness of

B1

D
W
F

B2

B3

B4

30

O.M. Querin et al.

of M = 2.0928, which is 3.29% less than that produced


by Srithongchai and Dewhurst (2003) and 4.14% less than
produced by the method of Martinez et al. (2007). The
resulting topology is very similar to that of the trussproduced optimum with the same number of elements. The
non-dimensional weight is slightly lower than the theoretical, which is impossible. We believe that a possible reason
for this is due to the stress state of the continuous structure.
In the theoretical solution, the trusses experience a uniaxial
stress state and the joints between trusses dont contribute
at all to the weight. In the continuous structure, the trusslike elements may experience a near uniaxial stress state,
but the joint regions have a biaxial stress system. Since we
were unable to use a Poissons ratio of zero, the stresses
in the joints may have been lower than they should have
been, causing the non-dimensional weight to be less than
the theoretical optimum.
4.3 Cantilever with circular support

Fig. 4 Optimal topology for the Prager Cantilever: a using the method
of Martinez et al. (2007), b this work

the domain t = 1 mm. The prescribed constant ratio between


stresses was k1 = 2.
Srithongchai and Dewhurst (2003) gave the theoretical solution for the non-dimensional weight to be M =
2.1513. Interpolation of Fig. 8 from their paper gave the
non-dimensional weight for a discrete truss solution to be
M = 2.164. The same problem was also optimised using
the method of Martinez et al. (2007), producing the layout of Fig. 4a with a non-dimensional weight of M =
2.1831. The optimal topology obtained using the method
proposed in this paper is given in Fig. 4b, with a resulting volume VOptimal = 748,781 mm3 and an average stress
level Av = 167.7 MPa giving a non-dimensional weight
Fig. 5 Design domain for the
Cantilever with Circular
Support

This example is the classic cantilever beam with a fixed circular boundary and an applied load F. The design domainis
a rectangle of length L + 1.5R and of height W = L e/42
(Hemp 1973), Fig. 5. As for the previous examples, a border denoted by lengths B1,2 was placed around the design
domain.
The dimensions used for this example are L = 1740 mm,
R = L/10 = 174 mm, 1.5R = 261 mm, W =
1121.94136 mm and, represented by the dashed rectangle. The lengths of the border were B1 = 36.42932 mm,
B2 = 34.8 mm. The entire design domain (solid outer
rectangle) was subdivided with a mesh of 206 351 elements. The distance to the vertical force F = 10 kN is D =
560.97068 mm with the thickness of the domain t = 1 mm.
The prescribed constant ratio between stresses was k1 = 2.
No theoretical solution was found in the literature for
this problem when the elements of the structure experiencing tension and compression have different permissible

B1

R
W
F

B1

1.5R

B2

Topology optimization of truss-like continua with different material properties in tension and compression

31

parts of the structure. The success of the method has been


demonstrated by the relatively small errors in the nondimensional weights of the structures calculated by this
method compared to those obtained with classical layout
theory.
Acknowledgement The authors would like to thank Professor
George I. N. Rozvany for his invaluable advice in ways to improve
the paper.

References

Fig. 6 Optimal topology for the Cantilever with Circular Support:


a using the method of Martinez et al. (2007), b this work

stresses. So the method of Martinez et al. (2007) was used


to produce the optimal discrete truss layout of Fig. 6a which
has a non-dimensional weight of M = 2.3474. The optimal topology obtained using the method from this paper
is given in Fig. 6b, with a resulting volume VOptimal =
482,583 mm3 and an average stress level Av = 160.30 MPa
giving a non-dimensional weight of M = 2.2229, which is
5.30% less than produced by the method of Martinez et al.
(2007). The resulting topology is very similar to that of the
truss-produced optimum with the same number of elements.
As for the previous example, the non-dimensional weight is
slightly lower than the theoretical.

5 Conclusions
The method presented in this paper combines aspects of
the work of Plfi (2004), Dewhurst (2005) and Nair (2005)
in order to allow for the topology optimization of continuous structures with different material properties in tension
and compression. The method consists of a few simple
steps: identifying all negative (compressive) parts of a structure; modifying their modulus of elasticity; changing their
material properties to orthotropic; modifying the stresses
or strains; and satisfying the optimality criterion of a constant stress/strain ratio between the tensile and compressive

Achtziger W (1996) Truss topology optimization including bar properties different for tension and compression. Struct Optim 12(1):
6374
Alfieri L, Bassi D, Biondini F, Malerba PG (2007) Morphologic
evolutionary structural optimization. In: 7th world congress on
structural and multidisciplinary optimization, Seoul, Korea, paper
A0422
Cai K, Shi J (2008) A heuristic approach to solve stiffness design
of continuum structures with tension/compression-only materials.
In: Fourth international conference on natural computation, vol 1,
pp 131135
Darwich W, Gilbert M, Tyas A (2008) FORM: a practical layout optimization tool for civil and structural engineers. In: 8th world
congress on computational mechanics (WCCM8). 5th European
congress on computational methods in applied sciences and
engineering (ECCOMAS 2008), Venice, Italy
Desmorat B, Duvaut G (2003) Compliance optimization with nonlinear
elastic materials application to constitutive laws dissymmetric in
tension-compression. Eur J Mech A, Solids 22:179192
Dewhurst P (2005) A general optimality criterion for combined
strength and stiffness of dual-material-property structures. Int J
Mech Sci 47:293302
Duysinx P (1999) Topology optimization with different stress limit in
tension and compression. In: Proceedings (CD-Rom) of the third
world congress of structural and multidisciplinary optimization
(WCSMO3), Buffalo, NY, USA
Duysinx P, Bruyneel M (2000) Recent progress in preliminary design
of mechanical components with topology optimisation. Integrated
Design and Manufacturing in Mechanical Engineering: Proceedings of the Third IDMME Conference Held in Montreal. Kluwer,
Canada, pp 457464, (ISBN: 1-4020-0979-8)
Duysinx P, Van Miegroet L, Lemaire E, Brls O, Bruyneel M (2008)
Topology and generalized shape optimization: why stress constraints are so important? Int J Simul Multidisc Des Optim
2:253258
Geli S, Dolex G, Ishai O (1981) An effective stress/strain concept in the
mechanical characterization of the structural adhesive bonding.
Int J Adhes Adhes 1:135140
Gilbert M, Darwich W, Tyas A, Shepherd P (2005) Application
of large-scale layout optimization techniques in structural engineering practice. In: Proceedings of the 6th world congress of
structural and multidisciplinary optimization (WCSMO6), Rio de
Janeiro, Brazil, Paper 401
Graczykowski C, Lewinski T (2006a) Michell cantilevers constructed
within trapezoidal domains. Part I: geometry of Hencky nets.
Struct Multidisc Optim 32:347368
Graczykowski C, Lewinski T (2006b) Michell cantilevers constructed
within trapezoidal domains. Part II: virtual displacement fields.
Struct Multidisc Optim 32:463471

32
Graczykowski C, Lewinski T (2007a) Michell cantilevers constructed
within trapezoidal domains. Part III: force fields. Struct Multidisc
Optim 33:2746
Graczykowski C, Lewinski T (2007b) Michell cantilevers constructed
within trapezoidal domains. Part IV: complete exact solutions of
selected optimal designs and their approximations by trusses of
finite number of joints. Struct Multidisc Optim 33:113129
Guan H, Steven GP, Xie YM (1999) Evolutionary structural optimisation incorporating tension and compression materials. Adv Struct
Eng 2(4):273288
Guan H, Chen YJ, Loo YCH, Xie YM, Steven GP (2003) Bridge
topology optimization with stress, displacement and frecuency
constraints. Comput Struct 81:131145
Hemp WS (1958) Theory of the structural design. Report no 115,
College of Aeronautics, Cranfield
Hemp WS (1973) Optimum structures. Clarendon, Oxford
Hinton E, Sienz J (1995) Fully stressed topological design of structures
using an evolutionary approach. Eng Comput 12:229244
Lewinski T, Rozvany GIN (2007) Exact analytical solutions for some
popular benchmark problems in topology optimization II: threesided polygonal supports. Struct Multidisc Optim 33:337350
Lewinski T, Rozvany GIN (2008a) Exact analytical solutions for
some popular benchmark problems in topology optimization III:
L-shaped domains. Struct Multidisc Optim 35:165174
Lewinski T, Rozvany GIN (2008b) Analytical benchmarks for topology optimization IV: square-shaped line support. Struct Multidisc
Optim 36:143158
Martinez P, Marti P, Querin OM (2007) Growth method for size,
topology, and geometry optimization of truss structures. Struct
Multidisc Optim 33:1326
Michell AGM (1904) The limits of economy of material in framestructures. Philos Mag 8(47):589597
Nair AU (2005) Evolutionary numerical methods applied to minimum weight structural design and cardiac mechanics. PhD thesis,
Mechanical engineering and applied mechanics, University of
Rhode Island
Nowak M (2006) Structural optimization system based on trabecular
bone surface adaptation. Struct Multidisc Optim 32:241249
Plfi P (2004) Locally orthotropic femur model. J Comput Appl Mech
5(1):103115
Pereira JT, Fancello EA, Barcellos CS (2004) Topology optimization
of continuum structures with material failure constraints. Struct
Multidisc Optim 26:5066
Prager W (1958) A problem of optimal design. In: Proceedings of the
union of theoretical and applied mechanics, Warsaw

O.M. Querin et al.


Prager W, Rozvany GIN (1977) Optimization of the structural geometry. In: Bednarek AR, Cesari L (eds) Dynamical Systems
(Proc Int Conf Gainsville, Florida). Academic, New York,
pp 265293
Raghava R, Caddell RM, Yeh GSY (1973) The macroscopic yield
behavior of polymers. J Mater Sci 8:225232
Rong J, Tang G, Liang QQ, Yang Z (2005) A topology optimization method for three-dimensional continuum structures. In: 6th
world congresses of structural and multidisciplinary optimization
(WCSMO6), Rio de Janerio, Brazil, Paper 6531
Rozvany GIN (1996) Some shortcomings in Michells truss theory.
Struct Optim 12:244260
Rozvany GIN (1997) Some shortcomings in Michells truss theory.
Authors reply and corrigendum. Struct Optim 13:203204
Rozvany GIN (2001) Aims, scope, methods, history and unified terminology of computer-aided topology optimization in structural
mechanics. Struct Multidisc Optim 21:90108
Rozvany GIN (2009) A critical review of established methods of structural topology optimization. Struct Multidisc Optim 37:217237
Rozvany GIN, Zhou M, Birker T (1992) Generalized shape optimization without homogenization. Struct Optim 4:250254
Rozvany GIN, Zhou M, Gollub W (1993) Layout optimization by COC
methods: analytical solutions. In: Rozvany GIN (ed) Optimization of large structural systems (Proc NATO ASI, Berchtesgaden
1991). Kluwer, Dordrecht, pp 77102
Rozvany GIN, Bendsoe MP, Kirsch U (1995) Layout optimization of
structures. Appl Mech Rev ASME 48:41119
Selyugin SV (2004) Some general results for optimal structures. Struct
Multidisc Optim 26:357366
Srithongchai S, Dewhurst P (2003) Comparisons of optimality criteria for minimum-weight dual material structures. Int J Mech Sci
45:17811797
Taggart DG, Dewhurst P, Nair AU (2007) System and method for finite
element based topology optimization, WO/2007/076357 (patent
application)
Timoshenko SP, Goodier JN (1982) Theory of elasticity, 3rd edn,
section 77. McGraw-Hill, New York, pp. 223224
Victoria M, Marti P, Querin OM (2009) Topology design of twodimensional continuum structures using isolines. Comput Struct
87(12):101109
Xie YM, Steven GP (1993) A simple evolutionary procedure for
structural optimization. Comput Struct 49(5):885896
Zhou M, Rozvany GIN (1991) The COC algorithm, Part II: topological, geometrical and generalized shape optimization. Comput
Methods Appl Mech Eng 89:309336

También podría gustarte