buckling sensitive structures

31
Continuum Topology Optimization of Buckling-Sensitive Structures Salam Rahmatalla and Colby C. Swan Department of Civil and Environmental Engineering Center for Computer-Aided Design The University of Iowa Iowa City, Iowa 52242, USA ABSTRACT Two formulations for continuum topology optimization of structures taking buckling considerations into account are developed, implemented, and compared. In the first, the structure undergoing a specified loading is modeled as a hyperelastic continuum at finite deformations, and is optimized to maximize the minimum critical buckling load. In the second, the structure under a similar loading is modeled as linear elastic, and the critical buckling load is computed with linearized buckling analysis. Specific issues addressed include usage of suitable “mixing rules'', a node-based design variable formulation, techniques for eliminating regions devoid of structural material from the analysis problem, and consistent design sensitivity analysis. The performance of the formulations is demonstrated on the design of different structures. When problems are solved with moderate loads and generous material usage constraints, designs using compression and tension members are realized. Alternatively, when fairly large loads together with very stringent material usage constraints are imposed, structures utilizing primarily tension members result. Issues that arise when designing very light structures with stringent material usage constraints are discussed along with the importance of considering potential geometrical instabilities in the concept design of structural systems. Key words: topology optimization; structural optimization; geometrical nonlinearity; design sensitivity analysis (DSA); buckling; instability. Corresponding author. E-mail: [email protected]; Phone: 1(319)335-5831; Fax: 1(319)335-5660.

Upload: nagaraj-ramachandrappa

Post on 02-Aug-2015

33 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

Salam Rahmatalla and Colby C. Swan

Department of Civil and Environmental EngineeringCenter for Computer-Aided Design

The University of IowaIowa City, Iowa 52242, USA

ABSTRACT

Two formulations for continuum topology optimization of structures taking bucklingconsiderations into account are developed, implemented, and compared. In the first, the structureundergoing a specified loading is modeled as a hyperelastic continuum at finite deformations, and isoptimized to maximize the minimum critical buckling load. In the second, the structure under asimilar loading is modeled as linear elastic, and the critical buckling load is computed with linearizedbuckling analysis. Specific issues addressed include usage of suitable “mixing rules'', a node-baseddesign variable formulation, techniques for eliminating regions devoid of structural material from theanalysis problem, and consistent design sensitivity analysis. The performance of the formulations isdemonstrated on the design of different structures. When problems are solved with moderate loadsand generous material usage constraints, designs using compression and tension members arerealized. Alternatively, when fairly large loads together with very stringent material usageconstraints are imposed, structures utilizing primarily tension members result. Issues that arise whendesigning very light structures with stringent material usage constraints are discussed along with theimportance of considering potential geometrical instabilities in the concept design of structuralsystems.

Key words: topology optimization; structural optimization; geometrical nonlinearity; designsensitivity analysis (DSA); buckling; instability.

Corresponding author. E-mail: [email protected]; Phone: 1(319)335-5831; Fax: 1(319)335-5660.

Page 2: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

2

1. INTRODUCTION AND MOTIVATION ..............................................................................3

2. DISTRIBUTION OF MATERIALS ........................................................................................5

3. ANALYSIS FORMULATION.................................................................................................6

3.1 Hyperelastic Structural Analysis at Finite Deformations..........................................................6

3.2 Compliance Functional and Design Sensitivity Analysis ..........................................................9

3.3 Critical Load and Design Sensitivity Analysis .......................................................................11

4. LINEARIZED BUCKLING MODEL ......................................................................................12

4.1 Analysis ...............................................................................................................................12

4.2 The Optimization Problem ...................................................................................................13

5. ANALYSIS PROBLEM SIZE REDUCTION TECHNIQUE....................................................14

6. SENSITIVITY ANALYSIS RESULTS..................................................................................16

6.1 Nonlinear Compliance and Critical Load Functionals............................................................16

6.2 Linearized Buckling Eigenvalue Functional ..........................................................................16

7. DEMONSTRATIVE EXAMPLES ..........................................................................................17

7.1 Material Properties and Mixing Rules..................................................................................17

7.2 The X-Structure Problem.....................................................................................................17

7.3 The Circle Problem ..............................................................................................................18

7.4 The Fixed-End Beam Problem..............................................................................................20

8. DISCUSSION AND CONCLUSIONS ...................................................................................21

9. ACKNOWLEDGEMENTS ....................................................................................................22

10. REFERENCES .......................................................................................................................22

Page 3: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

3

1. INTRODUCTION AND MOTIVATION

Variable topology material layout optimization is a potentially useful tool in the design of

structures, mechanical parts, composite materials, and even micro-electro-mechanical systems

(MEMS). Its usage and a variety of formulations have been widely explored over the past fifteen

years for a considerable range of applications. Bendsoe and Kikuchi (1) and numerous subsequent

works by a wide range of investigators have introduced topology optimization methods dealing with

linear elastic material behaviors and geometrically linear structural behavior. To this point, only a

few works have addressed continuum structural topology optimization of nonlinear systems.

Among these are works using Voigt-Reuss continuum topology formulations and consistent

sensitivity analysis techniques for designing structures (2) and composite materials (3) featuring

general materially nonlinear behaviors. More recently, a few works have also been presented for

continuum topology optimization of elastic systems undergoing nonlinear, finite deformations.(4-7)

Consideration of geometrical and/or buckling instabilities is an important issue in the conceptual

design of sparse spatial structures. For example, in design of long-span bridges, tension structures

are typically optimal since they preclude potential buckling. If continuum topology optimization

were applied to obtain concept designs of sparse spatial structures and buckling were not

considered, the design results could rely excessively on compression, and thus constitute

unacceptable concept designs. In design optimization of discrete truss structures, buckling behaviors

can be avoided altogether by prescription of Euler buckling constraints on loads in individual

structural members (8,9). In continuum topology design formulations, however, this process is much

more challenging since it is very difficult to identify discrete structural members, their geometrical

properties, and their end support conditions from the vector of design variables. One promising

approach to addressing geometrical instabilities in continuum structural topology optimization (10) is

to model the structure as a linearly elastic system and to use the minimum critical buckling load

computed via eigenvalue analysis in either the objective function or as a design constraint.

On the other hand, in the design of MEMS with continuum structural topology optimization, it

is not uncommon for the system to be designed to undergo finite deformations even before

geometric instabilities might develop (11,12). In such cases, it is necessary to base the analysis and

Page 4: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

4

design on a more general framework that addresses finite deformations in the system and the

development of potential geometrical instabilities. In recent years, works dealing with continuum

topology optimization of structures to minimize the generalized compliance at finite deformations

have been proposed and demonstrated (4), although minimization of nonlinear compliance does not

necessarily address potential geometrical instabilities. To this point in time, only a limited number of

works(5-7) have dealt with this important issue of taking finite deformations of the system into

account, while also taking account of the associated instabilities that arise in the structure in a

consistent fashion.

In the body of this paper, a general hyperelastic continuum framework is developed for

structural analysis and continuum topology optimization. Within this framework, both the

generalized compliance of the system, and the minimum critical buckling load for the system can be

addressed. As an alternative, a linear elastic continuum topology formulation is also used along

with minimum critical buckling values computed by linearized buckling analysis. For both

formulations, expressions for design sensitivity of structural responses to material distribution

parameters are developed and verified. The two approaches have been implemented, tested and

compared on a variety of relevant design problems.

Although they are not the central focus of this work, two additional issues critical to achieving

the objectives of this work are addressed in this manuscript. The first is that of using a continuum

topology formulation based on nodal design variables as opposed to element-based design variables.

Since node-based design variables feature C0 continuity, this new framework is invulnerable to

checkerboarding instabilities, and thus does not require any spatial filtering techniques to preclude

such instabilities. The second issue addressed is that of removing regions devoid of structural

material from the structural analysis model. In the current context this is a vital consideration, since

“empty” elements can suffer excessively large deformations as the structural system undergoes large

deformations. The excessive deformation in the “empty” elements can lead to singular behaviors

that make it extremely difficult to solve the nonlinear system equations. In addition, since the

structural analysis problems considered herein are frequently nonlinear, the savings in computational

effort when void regions are eliminated from the structural model can be very considerable.

Page 5: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

5

2. DISTRIBUTION OF MATERIALS

In most continuum structural optimization formulations, there is some treatment of intermediate

cases where a specific region of a structure is not fully occupied by solid structural material and yet

not completely devoid of structural material either. In this work, description of these regions uses

amorphous mixtures (i.e. no specific micro arrangement of materials is assumed). The complete

undeformed spatial domain of the structure to be designed is denoted by SΩ ; its designable subset

by DΩ ; and its non-designable subset in which the spatial/topological arrangement of materials is

taken to be fixed by NΩ . The arrangement of a structural material in DΩ remains to be determined

and so this region is called designable. A set of single or multiple loading/boundary conditions to

which SΩ will be subjected are specified and a starting design )0(b which specifies the initial

material layout in DΩ is selected. For each set of loading/boundary conditions, the structure will be

analyzed as a boundary value problem.

The design of a structure is here considered to be the spatial distribution of the structural

material in Ωs. To describe very general structural material distributions in Ωs a volume-fraction

approach described earlier in [4] and alternatively described as the ``density'' approach by others, is

used. While preference is given here to final material layout distributions, which are nearly discrete,

such distributions are typically achieved using continuous formulations permitting mixtures to exist

throughout the design domain DΩ . By permitting mixtures, the structural material A and a

fictitious void material B are allowed to simultaneously and partially occupy an infinitesimal

neighborhood about each Lagrangian point X in DΩ . The volume fraction of structural material

phase A at a fixed Lagrangian point X in the design domain DΩ is denoted by )(XAφ and

represents the fraction of an infinitesimal volume element surrounding point X occupied by material

A . Natural constraints upon the spatial volume fractions for the two-material problem are:

1X0 ≤≤ )(Aφ 1X0 ≤≤ )(Bφ 1)()( =+ XX BA φφ (2.1)

Page 6: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

6

The last physical constraint of (2.1) states that the material volume fractions at X are not

independent and so one need only be concerned with the layout of structural material A .

Using the same mesh and basis functions that will be used to solve the structural analysis

problem described below, the spatial distribution of structural material in SΩ is expressed using the

following expansion:

SΩ∈∀Σ==

xxx )(Nφ)(φ ii

Numnp

1i(2.2)

where iφ are the nodal volume fractions, and )(Ni x are the nodal shape functions, and “Numnp”

denotes the number of nodes in the structural model. The design vector b describing the

arrangement of materials in the structural domain SΩ thus has the composition

( )Numnp321 ,......φφ,φ,φ=b . This approach yields a oC continuous design variable field that is not

susceptible to “checkerboarding” instabilities.

3. ANALYSIS FORMULATION

3.1 Hyperelastic Structural Analysis at Finite Deformations

The strong form of the nonlinear elliptic boundary value problems to be solved is:

Find 3]),0[(: ℜ×Ω TSu , such that the Kirchhoff stress field satisfies

0γρτ j0jij, =+ on SΩ ],,0[ Tt ∈∀ (3.1a)

subject to the boundary conditions:

(t)g(t)u jj = on gjΓ for ,1,2,3j = T][0,t ∈ (3.1b)

(t)hτn jiji = on hjΓ for ,1,2,3j = T][0,t ∈ (3.1c)

The Kirchhoff stress tensor τ is related to the Cauchy stress tensor σ via the relation στ J=

Page 7: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

7

where )det(F=J . As is customary, it is assumed that the Lagrangian surface Γ bounding the

Lagrangian structural domain SΩ admits the decomposition hjgj ΓΓΓ ∪= and ∅=Γ∩Γjj hg for

1, 2,3.j = For a given mesh discretization of Ωs whose complete set of nodes is denoted η, the

subsequent design formulation is facilitated by introducing a subset of nodes ηh at which non-

vanishing external forces are applied, and a subset of nodes ηg at which non-vanishing prescribed

displacements are applied.

Since the analysis problem is being solved in the context of topology optimization, it is assumed

that a local microscopic mixture of two generic materials A and B resides at each point X in the

structural domain SΩ . In (3.1a) τ represents the macroscopic Kirchhoff stress of the local

mixture, which is dependent upon the constitutive properties of the two material phases and the

mixing rule employed

],φ);();([)( BBAA FFX ττττ = (3.2)

where AF and BF are the respective deformation gradients for materials A and B at a given point

X , Aτ and B

τ are the stress tensors for the two materials at X , and φ is the volume fraction of

material A. For simplicity and efficiency, a power law mixing rule with an iso-deformation condition(6) is used here. Therefore a specific form of (3.2) is:

,)φ1(φ)( pp BAττXτ −+= (3.3)

where 1≥p is a fixed parameter of the mixing rule. In accordance with the iso-deformation

assumption,

XFFF

∂∂=== xBA (3.4)

where )(XuX +=x

It is assumed that both materials A and B obey isotropic hyperelastic constitutive laws. The

particular strain energy function E used here for both materials is that of Ciarlet (1988) wherein the

volumetric )(U and deviatoric )(W strain energy functions for the two materials A and B are

assumed to be decoupled and of the forms:

)()()( θWJUE +=F (3.5a)

Page 8: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

8

−−= )ln()1(2

1

2

1)( 2 JJKJU (3.5b)

]3)([2

1 −= θtrW µ (3.5c)

In the preceding expression, J is again the determinant of F ; K is a constant bulk modulus; µ is a

constant shear modulus; TFFθ = is the left Cauchy-Green deformation tensor; and θθ(2/3)J−= is its

deviatoric part. For this model, therefore, the Kirchhoff stress in a material τ is thus related to

deformation quantities in that material as follows:

θ1τ

∂∂+′= W

dev2(J)UJ (3.6)

The weak or variational form of the stress equilibrium problem can be obtained by restating the

original form (3.1a) as

0]dΩδuγρδu[τSΩ

Sjj0jiij, =+ (3.7)

from which integration by parts, usage of Green's Theorem and utilization of the natural boundary

conditions gives the virtual work equation

+=hS S Γ

hjj

Ω Ω

Sjj0Sijij dΓδuhdΩδuγρdΩδετ (3.8)

In (3.8) the expression on the left represents the internal virtual work )Wd( intδ , and that on the

right, the external virtual work )Wd( extδ . The differential of the internal virtual work can be written

as:

Ω Ω

+=S S

SjmimjiSijvijint dΩdετδε)dΩ(τdLδε)W(d δ (3.9)

indicating a decomposition into, respectively, a material stiffness term containing the Lie differential

of the Kirchhoff stress, and a geometric stiffness term.

Usage of a Galerkin formulation, in which the real and variational kinematic fields are expanded

in terms of the same nodal basis functions, and discretization of the time domain into a finite number

Page 9: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

9

of discrete time points, leads to the following force balance equations at each unrestrained node A in

the mesh as here at the thn )1( + time step:

0ffr =−= +++A

1nextA

1nintA

1n )()( (3.10)

where

Ω

+++ =S

S1nT

1nAA

1nint dΩ:)()( τBf (3.11a)

Ω Γ

+++ +=S h

h1nA

S1nA

0A

1next dΓNdΩNρ)( hγf (3.11b)

In (3.11), A1n+B represents the spatial infinitesimal nodal strain displacement matrix

))(N( AA1n 1

xB sxn+

∇=+ , and AN denotes the nodal basis function for the thA node. Under finite

deformations, (3.10) represents a set of nonlinear algebraic equations that must be solved in an

iterative fashion for the incremental displacement field nnn uuu −=∆ ++ 11)( for each time step of the

analysis problem.

When external forces applied to a structure are independent of its response, the derivative of the

thi residual force vector component at the thA node with respect to the thj displacement vector

component of the thB node is simply:

Ω Ω

Ω+Ω=S S

SilBkjk

AjS

Bkljk

Aji

ABil dNNdBcB δτ ,,K (3.12)

where jkc is the spatial elasticity tensor in condensed form. Assembly of this nodal stiffness operator

for all unrestrained nodes A and B gives the structural tangent stiffness matrix.

3.2 Compliance Functional and Design Sensitivity Analysis

The generalized compliance Π of a structural system undergoing a loading by a system of

varying body forces, applied tractions, and applied displacements can be expressed as follows:

Page 10: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

10

(3.13)dd)(dd)(ddΩ)(ρ(φ)(t) g

0

h

0

s

0

ττττττ Γ⋅−Γ⋅+⋅=Π ΓΓΩ

gtvhvγn

ttt

ghs

where )(τγ is a time-varying body force vector, )(τh a system of surface tractions applied to the

structure, and )(τg a field of velocities applied to the structure, where in all cases τ plays the role

of a parametric time variable. The generalized compliance of a structure can potentially be used to

design structures in a way that minimizes their susceptibility to buckling. For example, when a

structure goes through a point of instability, the incremental compliance at that point gets very large.

Consequently, it is possible that by minimizing the generalized compliance of a structural system,

one is also designing the structure to be less prone to buckling. For a wide variety of material

behaviors, the generalized compliance of a structural system can be computed using integration as

follows:

( ) ( )

( ) ( )

( ) ( ) ( ) ( )

[ ] (3.14c)

(3.14b)

(3.14a)-)(t

1

01

1

011

1

01

1

01N

21

21

21

21

=+

= ∈ ∈++++

=++

=++

∆Π=

∆⋅−∆⋅=

∆⋅∆⋅=Π

N

nn

N

n k j

jn

jn

kn

kn

j

N

n

jn

jn

k

N

n

kn

kn

h g

gh

η η

ηη

gfuf

gfuf

The first sum corresponds to work done by external forces applied to the structure, where specific

expressions for external forces were provided in (3.11b), and the second sum expresses the work

done on the structure by applied displacements. Displacement loading is applied to the structure

with prescription of non-vanishing displacements g at the nodal subset η g . Such nodes need not

necessarily lie on the external boundaries of SΩ , and the associated displacement field is simply

jj

j )(N)(g

gXxg η∈Σ= where jg are applied nodal displacements.

The design derivative of the generalized structural compliance at any given time instant Nt can

be expressed using adjoint sensitivity analysis (13) simply as:

(3.15)d

)(d)t( 1

0

1

=

+∆Π=

Π N

n

nN

d

d

bb

where

Page 11: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

11

( ) ( )

( ) ( )

(3.16b)d

d

(3.16a)d

d

d

d

d

d

11ka

11

11

11

21

21

21

21

21

jn

j

jn

k

nn

kn

kn

jn

j

jn

k

knk

nkn

knn

hg

hg

+∈

+

+++

+

+∈

+

+++

++

∆⋅+

∂∂

⋅+∆⋅∂

∂=

∆⋅+

∆⋅+∆⋅

∂=

∆Π

gb

f

br

uub

f

gb

f

bu

fub

f

b

ηη

ηη

In (3.16b), ( )kan 1+u is the solution of the following linear adjoint problem solved after the

equilibrium analysis problem is solved at the (n+1)th time/load step:

( ) (3.17).2111

k

n

kann +++ −=⋅ fuK

In solving (3.17), the current tangent stiffness operator specified in (3.12) is employed.

3.3 Critical Load and Design Sensitivity Analysis

An alternative approach to designing buckling-resistant structures is to apply displacement-

controlled loading to the structure and to maximize the critical internal force that can be generated

in response to this loading. The magnitude of the applied displacement loading is merely that which

induces the first instability in the structural model. The first instability point is the first point in the

load-deflection response of the structural model at which the tangent stiffness operator defined in

(3.12) becomes singular. A reliable algorithm for finding such singular points that has been used by

the authors in a variety of other applications (14,15) is presented in Figure 1. Essentially, this

algorithm involves gradual and iterative approaching of the structural model’s first critical point of

instability.

Once the critical point is found, the objective is to compute the magnitude of the resistance

force generated in the structure, and to design the structure so as to maximize the resistance. For

example, if the applied displacement loading KK

nK Ng

gXxg ).()( ∈Σ= is applied to the structure

and induces instability, then the magnitude of the critical resistance force generated in the structure

will be:

(3.17)f int

crit ∈ Ω

Ω⋅⋅

−=

gnK

TK

K

K dτBgg

Page 12: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

12

Accordingly, an optimization problem to maximize this critical force would be as follows:

( ) (3.18)0V-Vsuch thatand),(such thatfmax allowableMaterialintcrit ≤= 0ubr

b

The design derivative of the critical buckling force is computed using adjoint sensitivity analysis as

follows:

( )

(3.19)d

df intcrit

∈ Ω

∂∂⋅+Ω

∂∂⋅⋅

−=

gnK

aK

TK

K

K dbr

ubτ

Bgg

b

where ( )aKu is the adjoint displacement associated with the Kth node at which prescribed

displacements are applied. Specifically, it is the solution of the following adjoint problem:

( ) ( )(3.20)

f intcrit

uuK

∂∂

−=⋅ KaK

where K is tangent stiffness operator at the current state of the model.

4. LINEARIZED BUCKLING MODEL

4.1 Analysis

The full geometrically nonlinear method proposed above for maximizing the minimum

critical buckling load in continuum structural topology optimization is potentially very

computationally expensive. A potentially more efficient alternative might be to use linearized

buckling eigenvalue analysis to estimate the critical buckling load for the structure and to then

maximize this approximation. Neves, Rodrigues, and Guedes (10), have presented an approach to

include the critical load criterion in the continuum topology optimization model. Linearized buckling

eigenvalue analysis proceeds as follows: A prescribed force loading fext is applied to the structure

with its magnitude necessarily being less than that required to induce geometric instability in the

structure. Once the resulting linear, elastostatic displacement solution Niu Ru ∈= to the applied

loading fext is obtained ( )tL

exfuK =⋅ , where KL is the linearized stiffness matrix, then the following

eigenvalue problem is solved

Page 13: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

13

0ψbuGψbK L =+ ),()( λ (4.1)

In the preceding, Me R∈= bb is the vector of design variables; LK is the linear tangent stiffness

operator; ),( buG is the linearized geometric stiffness matrix;Gψψ

ψKψ LT

T

−=λ is an eigenvalue

denoting the magnitude by which fext must be scaled to create instability in the structure, and ψ is a

normalized eigenvector satisfying 1=ψKψ LT . For this model, it is assumed that linearized stiffness

operator KL is real, symmetric, and positive definite, while G is only assumed to be real and

symmetric. To avoid numerical difficulties in the solution of (4.1) associated with the indefinite

characteristics of G , it is common [Bathe (16)] to solve the modified eigenvalue problem that deals

with two positive definite matrices.

0ψKGK LL =++ ))(( γ (4.2)

where

)1

λγ −= .

4.2 The Optimization Problem

Once the linearized eigenvalue problem (4.2) is posed and solved, the design problem is

formulated to maximize the minimum buckling load ( λ ). The objective functional Ef to be

minimized for this problem would simply be the reciprocal of the lowest eigenvalue λ as follows.

)min(

1),(

λ=buEf (4.3)

The optimization problem is thus stated to minimize the reciprocal of the first (or minimum) critical

buckling load as follows

)max(min)1

(minmin0,,, ψKψ

Gψψ

T

T

Ef≠

−==ububub λ

(4.4)

subject to the normal bound constraints on the design variables (2.1), the linear, structural

equilibrium state equation ( )extL fuK0bu,r −⋅==)( , and a constraint on material resources.

Using adjoint design sensitivity analysis, the Lagrangian of the optimization problem (4.4) is:

)( bu,ru ⋅+= aEfL (4.5)

Page 14: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

14

where au is the adjoint displacement vector and the solution of the linear, adjoint problem

ψuG

ψuK L ∂∂= Ta . (4.6)

For a simple eigenvalue the design gradient of L is written simply as

br

bb ∂∂⋅+

∂∂= TaEf

d

Ld)(u (4.7)

where:

ψb

K

bG

ψb

L )1

(∂

∂+

∂∂−=

∂∂

λTEf

. (4.8)

In (4.7), the partial derivative of r with respect to b can be found from the equilibrium state equation

as follows:

bu

b

K

br L

∂∂−⋅

∂∂

=∂∂ extf

(4.9)

Consequently, the final expression for the design gradient of L can be written in the following form:

)()()1

(b

fu

b

Kuψ

b

K

bG

ψb

LL

∂∂−⋅

∂∂

+∂

∂+

∂∂−=

extTaT

d

dL

λ(4.10)

In the current formulation, a simple eigenvalue has been assumed. In cases where multiple

eigenvalues occur, computation of the sensitivities can be somewhat more involved since the

eigenvectors of the repeated eigenvalues are not unique, rendering the eigenvalues only directionally

differentiable (17,18). Frequently, however, multiple eigenvalues occur in symmetrical structures and

are actually attributable to the structural symmetry (19). In such cases, approaches that reduce the

design space in accordance with the symmetry (20) can render the repeated eigenvalues fully

differentiable in the reduced design space.

5. ANALYSIS PROBLEM SIZE REDUCTION TECHNIQUE

In continuum topology optimization of sparse structures with limited structural material

usage and realistically large design loads, geometrical instabilities become a definite possibility. If

modeled, geometrical instabilities in structural systems can result in finite deformations. The

Page 15: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

15

modeling of finite deformations in mixed solid-void grid-like meshes used in continuum structural

topology optimization can result in excessive distortion of void or low-density elements that can in

turn lead to numerical difficulties solving the structural analysis problem. Since the optimization

process in continuum topology optimization typically removes structural material from low stress, or

low-sensitivity areas, fairly substantial regions of low-density elements are common. As these

elements are highly compliant, they contribute very little to structural stability, while being subject to

excessive deformation that creates numerical difficulties. Therefore, it is sometimes advantageous to

identify these large regions of void and low-density elements and to remove them, at least

temporarily, from the structural analysis problem. An automated algorithm for identifying such

regions and removing them from the structural analysis problem is presented and discussed below. It

is worth noting that the procedure proposed and investigated here is reversible in that it permits low-

density regions of the structure to return as high-density structural regions even after they have

previously been removed from consideration during structural analysis.

The essence of the proposed analysis problem reduction technique can be captured in the

three steps listed below:

1. All finite elements in the structural analysis model that are devoid of solid material, ornearly so, are identified as “void” elements. (Typically, in the examples presented below,if an element’s volume fraction of solid material is less than or equal to .002, it isidentified as “void”.)

2. All nodes that are members only of “void” elements are identified as “prime” nodes. Thedegrees of freedom of such “prime” nodes are restrained, reducing the size of theanalysis problem.

3. If only “prime” nodes comprise an element, that element is then denoted as a “prime”element. Such “prime” elements are then neglected in the structural analysis problem sothat if they undergo excessive distortion it does not create any singularities in the systemof finite element equations. It is worth noting, that “prime” elements are those that aresurrounded by “void” elements.

A graphical description supporting the explanation of this technique for reducing the analysis

problem is presented in Figure 2.

The matter of reducing the analysis problem by neglecting significant regions of void

elements has been addressed in preceding works (5,6). The current reduction techniques have proven

to be both robust and efficient in all of the example problems presented in Section 7 below. The

techniques are especially powerful and effective when applied in design problems involving

Page 16: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

16

extremely sparse structures, since highly refined meshes are needed when very stringent material

resource constraints are imposed. When a fine mesh is employed with a very limited amount of

structural material, the proposed reduction techniques will allow for dramatic savings in computing

effort.

6. SENSITIVITY ANALYSIS RESULTS

6.1 Nonlinear Compliance and Critical Load Functionals

The sensitivity analysis formulation of Eqs. (3.15) and (3.16) for strain energy in the

structure is first applied to the problem of a hyperelastic cantilever beam undergoing finite

displacement-controlled loading of magnitude L/3 as shown below in Figure 3. Design sensitivities

associated with all nodal design variables were computed from (3.15) and (3.16) and also using

finite difference of the design variables. Nearly exact agreement was obtained between the results

computed with the semi-analytical expressions and the results computed with finite difference. The

design sensitivities at a sample of the nodes are presented in Table 1, showing agreement with

converged finite difference sensitivities to four digits of precision.

For the beam shown in Figure 3, a compressive load was also applied and full nonlinear

analysis was performed up until the structural model became unstable. The design gradients of the

magnitude of the critical buckling load for the structure were computed both by (3.19) and (3.20)

and also by finite difference in design space. The excellent agreement between the semi-analytical

results and the converged finite difference results are shown in Table 2, confirming the sensitivity

analysis expressions used.

6.2 Linearized Buckling Eigenvalue Functional

Next, the linearized buckling sensitivity analysis formulation of Eq. (4.13) was applied to the

problem of a linearly elastic cantilever beam to which a constant axial compression force P is

applied as shown in Figure 4. For the geometry, material properties, and applied loading, the

buckling mode for this beam is as shown in Figure 3b. Sensitivity of the critical buckling load factor

Page 17: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

17

λ expressions based on (4.13) and based on converged finite difference analysis are compared in

Table 2, and show again excellent agreement.

7. DEMONSTRATIVE EXAMPLES

7.1 Material Properties and Mixing Rules

In all example problems solved below, the initial starting designs always utilized a completely

solid structural domain. In addition, the solid structural material in all problems was isotropic with

Young’s modulus of 307 GPa and shear modulus of 118 GPa. Furthermore, the powerlaw mixing

rule with p = 4 was used in all computations. The nodal design variable formulation of Section 2

was employed without any spatial filtering of design variables and without any perimeter control.

7.2 The X-Structure Problem

In this problem, an X-shaped structure shown in Figure 5a is considered. The lateral by

vertical dimensions of the frame are (11x40). A displacement loading of 10≤d is applied to the

center node of this structure, and for all designs, the algorithm of Figure 2 is applied together with

geometrically nonlinear structural analysis to compute the critical internal buckling force that

develops in response to the applied displacement loading. The design problem solved is that posed

in (3.18), where only half of the original structural volume can be employed in the final, optimized

structure. The structure is meshed with 1200 bilinear continuum finite elements, and the

optimization problem is solved using sequential linear programming techniques. The design solution

to this particular version of the problem is shown in Figure 5b. Since it employs strictly tension as

opposed to compression, it is indeed the correct solution that one would expect.

While optimization of structures to maximize the minimum critical buckling load based on

fully nonlinear analysis was shown to be successful above, the computational expense can be

Page 18: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

18

considerable. As a potentially more efficient way to achieve the same objective, linearized buckling

analysis can also be used. Thus, for the same structural model shown in Figure 5a, a force of

magnitude 12ext 109.1 ⋅=f is applied to the center node the structural optimization problem was

solved once again as follows:

0.V-Vsuch thatand

(6.1)loadappliedanfor)(such that1

min

allowable

ext

=

f0ub,rb λ

With this linearized buckling criteria, the resulting design (Figure 5c) is virtually identical to that

achieved with nonlinear analysis and shown in Figure 5c. It is worth noting, however, that the

computational time required with the linearized buckling analysis was significantly smaller than that

required with the geometrically nonlinear analysis.

The preceding design problem was solved yet again to minimize the generalized compliance

of the structure (3.14) under the applied displacement loading up to the first point of instability. In

this particular instance, the resulting design obtained to minimize generalized compliance and shown

in Figure 5d, is comparable to the preceding designs obtained with nonlinear and linearized buckling

analysis.

7.3 The Circle Problem

Continuum topology optimization solutions are frequently used only as starting concepts that

are suggestive of potentially optimal structural forms. That is, the designs produced are then taken

to a second stage where more detailed shape and sizing analysis is performed. When used in design

of large-scale structures continuum topology solutions are frequently quite ``heavy'' in that the ratio

of volume occupied by structural material to the enclosed structural volume can be unrealistically

large. Specifically, the global volume fraction constraint used in many continuum topology

optimization methods is typically in the range of 0.10 - 0.50, whereas in many structures, the

structural material occupies only approximately %21− of the enclosed structural volume. Thus,

while buckling of compression members can be a real concern in design of sparse structures,

Page 19: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

19

continuum topology methods that produce unrealistically “heavy” designs are unable to address such

concerns in the concept design stage.

In order to address the development of geometric instabilities associated with buckling,

continuum topology optimum methods must be used with a finite deformation nonlinear formulation

or with a linearized buckling approach. In addition, sufficiently stringent material usage constraints

must be imposed which can in turn require highly refined structural meshes. To illustrate this point,

we consider the design of a structure to carry a radial point load applied at the center of the

structural domain to the fixed boundary of the circular domain of Figure 6a. The analysis problem

was solved by imposing a finite displacement 0.1Rδ ≤ at the center node of the structure and using

a nonlinear formulation to compute the resulting structural response. The design optimization

problem was first solved with a coarse mesh (Fig. 6a) to minimize the generalized compliance of the

structure up to the first point of instability. A structural material global volume fraction constraint

of %20 was imposed. The resulting solution shown in Figure 6b is quite “heavy” and supports the

applied load using both tension and compression. If during secondary design, however the members

are made much lighter, the possibility of buckling in the compression members would then become

apparent, and the design would need to be changed significantly.

In order to detect the onset of geometrical instability within continuum structural topology

optimization, it is sometimes necessary to solve problems of this type with much smaller volumes of

structural material or with loads and/or displacements large enough to generate instability. The

problem was solved once again to minimize the generalized compliance with the finer mesh shown in

Figure 6d and a structural material constraint of 2.5% of the total structural volume. The resulting

design is shown in Figure 6e (undeformed) and Figure 6f (deformed). Clearly, the structural model

detects instability in the light compression members, yet the problem formulation does not result in

complete elimination of the compression member.

As an alternative to solving this “Circle” problem using a generalized compliance objective

function, it was solved again using the minimum critical buckling loads as the objective function.

Using the moderately fine mesh shown in Figure 7a, the design problem was solved a number of

Page 20: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

20

times. It was first solved with material usage constraints of 25% of the structural volume, and with

applied an applied displacement 0.1Rδ ≤ , resulting in the design shown in Figure 7b (undeformed)

and 7c (deformed). Since the material usage constraint is generous, and since the loading is not

necessarily large enough to generate structural instability, a design using both compression and

tension is produced. When the design displacement loading is increased to 0.8Rδ ≤ and when the

material usage constraint is reduced to 5% of the total structural volume, the purely tensile designs

of Figure 7d (undeformed) and 7e (deformed) are obtained. Such designs are clearly preferred since

they will not be vulnerable to buckling instabilities. A similar solution (Figures 7f and 7g) can be

obtained by maximizing the minimum critical buckling load obtained using linearized buckling

analysis.

In the preceding design problems, relatively fine meshes were required in order to achieve

sparse structural designs. The computational cost associated with using fine mesh can be high.

However, when the problem reduction techniques of Section 5 are employed, the cost can be

reduced considerably. At the start of the design process, the structural analysis problems are indeed

quite large and expensive. However, as material is eliminated from the structural model as the

optimization process continues, the analysis problem gets progressively smaller, until at the optimal

design, the actual analysis problem can be quite modest in size. For example, in treating the circle

problem, the analysis computation time at the optimal designs was reduced by more than a factor of

ten using the proposed technique.

7.4 The Fixed-End Beam Problem

In this problem, a fixed end beam is loaded vertically with a large displacement applied to the

center of the upper edge of the beam as shown in fig (7a). The vertical and horizontal dimensions of

the beam are (12x40), and the structural material usage constraint for this problem is 10% of the

total structural volume. A similar problem was introduced and solved by Buhl and Sigmund (5) who

showed that their formulation yields completely different designs when using geometrically linear

analysis and when using geometrically nonlinear analysis. More recently Gea and Luo (7) have

studied a similar problem with three concentrated loads applied to the upper edge of the beam,

obtaining design results similar to those presented in [5]. In the present work both buckling

formulations introduced earlier are applied to this problem in concert with the size reduction

Page 21: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

21

technique mentioned in Section 5. In the nonlinear problem formulation, a large displacement of

9≤d is applied to the structure, while a force 12109.1 ⋅=f is applied in the linearized buckling

case. The continuum topology optimized material layout designs obtained are shown in Figures 7b

and 7c. The resulting topologies signify the ability of both methods to track the geometrical

instabilities occurring in the structure, by building two long tension members and two short

compression members. These results agree with those of [5] and [7].

In [5], the authors concluded that usage of a hyperelastic constitutive law such as that used

by Bruns and Tortorelli (4) appeared to result in unstable nonlinear analysis problems. In the present

work, the hyperelastic constitutive model of Section 3 has been used in the nonlinear formulation

together with the analysis problem size reduction technique, and has given very good results for the

tested problems.

8. DISCUSSION AND CONCLUSIONS

In this paper, the objective has been to develop continuum structural topology optimization

formulations that can be used to detect and avoid buckling instabilities in the conceptual design

stage of sparse structural systems. Toward this end, both a finite deformation hyperelastic treatment

of the structure, and linear elastic, linearized buckling treatment of the structure have been

developed and implemented. With the hyperelastic structural treatment, both generalized

compliance and the minimum critical buckling load were separately considered as objective

functions. Based on the example problems solved, selection of the minimum critical buckling load as

the objective function appears to be more effective at consistently achieving stable designs than the

generalized compliance of the structural system. The minimum critical buckling load as computed

from linearized buckling eigenvalue analysis was also considered as an objective function in this

work, and was found to give design results comparable to those produced from nonlinear stability

analysis. Since the formulation is based on linear structural analysis, it is much less computationally

expensive than the formulations based on nonlinear analysis.

A structural analysis problem size reduction technique has been successfully implemented

and tested on the numerous linear and nonlinear problems presented in this work. This technique

Page 22: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

22

reduces the problem size considerably, and at the same time removes the unstable elements

temporarily from the structural domain. In addition, the nodal volume fraction approach introduced

in this work has also been found very effective at eliminating the numerical instabilities that lead to

“checkerboarding” designs.

9. ACKNOWLEDGEMENTS

This research was funded in part by a grant from the University of Iowa CIFRE Program, and in

part by an NSF/DARPA Grant in the OPAAL Program. Dr. Richard K. Miller is also acknowledged

for helpful discussions and recommending the “circle problem” as a good test problem.

10. REFERENCES

1 Bendsoe, M.P., and Kikuchi, N., “Generating optimal topology in structural design usinga homogenization method,” Comput. Meth. Appl. Mech. and Engng., Vol. 71, 1988, pp. 197-224.

2 Swan, C.C. and Kosaka, I., “Voigt-Reuss topology optimization for structures with nonlinearmaterial behaviours,” Int. J. Numer. Meth. Engng. Vol 40, 1997, pp. 3785-3814.

3 Swan, C.C., and Arora, J.S., “Topology optimization of material layout in structured compositesof high stiffness and high strength,” Structural Optimization, Vol. 13 (1), 1997, pp. 45-59.

4 Bruns, D.A., and Tortorelli, D.A., “Topology optimization of geometrically nonlinear structuresand compliant mechanisms,” Proc. 7-th Symposium on Multidiciplinary Analysis and Optimization,AIAA/USAF/NASA/ISSMO, AIAA-98-4950, pp. 1874-1882, 1998.

5 Buhl, T. Pedersen, W., and Sigmund, O., “Stiffness design of geometrically nonlinear structuresusing topology optimization,” Struct. Multidisc. Optim., Vol. 19, 2000, pp. 93-104.

6 Pedersen, C. B. W., Buhl, T., and Sigmund, O., “Topology synthesis of large-displacementcompliant mechanisms,” Int. J. Num. Meth. Engng. Vol. 50, 2001, pp. 2683-2705.

7 Gea, H. C., and Luo, J., ”Topology optimization of structures with geometrical nonlinearities,”Computers & Structures, Vol. 79, 2001, pp. 1977-1985.

8 Rozvany, G.I.N., “Difficulties in truss topology optimization with stress, local bucklingand system stability constraints,” Structural Optimization, Vol. 11, 1996, pp. 213-217.

9 Achtziger, W., “Local stability of trusses in the context of topology optimization Part II:

Page 23: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

23

A numerical approach,” Structural Optimization, Vol. 17, 1999, pp. 247-258.

10 Neves, M.M., Rodrigues, H., and Guedes, J.M., “General topology design of structures witha buckling load criterion,” Structural Optimization, Vol. 10, 1995, pp.71-78.

11 Sigmund, O., “On the design of compliant mechanisms using topology optimization,” Mech. ofStruct. and Mach., Vol. 25 (4) , 1997, pp. 495-526.

12 Larsen, U., Sigmund, O., and Bouwstra, S., “Design and fabrication of compliant micro-mechanisms and structures with negative Poisson’s ratio,” J. of Microelectromechanical Sys. Vol.6, 1997, pp. 99-106.

13 Cardoso , and Arora, J.S., “Adjoint sensitivity analysis for nonlinear dynamic thermoelasticsystems,” AIAA J. Vol. 29 (2), 1991, pp. 253-263.

14 Swan C.C., and Seo Y.-K., “Limit state analysis of earthen slopes using dual continuum /FEMapproches.” Int. J. Num. Analat. Meth. Geomech. Vol. 23, 1999, pp. 1359-1371.

15 Swan C.C., and Seo Y.-K., “Stability analysis of embankments on saturated soils using elasto-plastic porous medium models,” J. Geotech. Geoenv. Eng., Vol. 127 (5), 2001, pp. 436-445

16 Bathe, K.J., “Finite Element Procedures,” New Jersey, Prentice Hall. 1996.

17 Seyranian, A.P., Lund, E., and Olhoff, N., “Multiple eigenvalues in structural optimizationproblems,” Structural Optimization, Vol. 8, 1994, pp. 207-227.

18 Ohsaki, M., and Uetani, K., “Sensitivity analysis of bifurcation load of finite-dimensionalsymmetric systems,” Int. J. Num. Meth. Engng. Vol. 39, 1996, pp. 1707-1720.

19 Pedersen, N.L., “Maximization of eigenvalues using topology optimization,” StructuralOptimization, Vol. 20, 2000, pp. 2-12.

20 Kosaka, I., and Swan, C.C., “A symmetry reduction method for continuum structural topologyoptimization,” Computers & Structures Vol. 70, 1999, pp. 47-61.

Page 24: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

24

.

Table 1. Compliance design gradients for displacement – loaded hyperelastic structure.Description DSA Algorithm Finite DifferencePower Law P = 3 P = 2 P = 3 P = 2

Nodes Finite Deformation of Hyperelastic Structure

#04 -2.1865D+07 - 2.8353D+07 - 2.1864D+07 - 2.8353D+07#14 - 3.8811D+05 - 5.0329D+05 - 3.8807D+05 - 5.0323D+05

#16 -6.5214D+07 - 8.4567D+07 - 6.5215D+07 - 8.4567D+07

Table 2. Design gradients for critical buckling loads and load factors.Description DSA Algorithm Finite DifferencePower Law P = 3 P = 2 P = 3 P = 2

Nodes Design Gradient for Nonlinear Buckling Load

#04 -4.8043D+06 -6.2300D+06 -4.8042D+06 -6.2299D+06

#14 -1.8278D+05 -2.3703D+05 -1.8277D+05 -2.3701D+05

#16 -1.5700D+07 -2.0359D+07 -1.5700D+07 -2.0359D+07

Nodes Design Gradient for Linearized Buckling Load Factor

#04 -1.4109D-01 -9.8048D-02 -1.4110D-01 -9.8051D-02

#14 -3.0288D-03 -2.1048D-03 -3.0289D-03 -2.1048D-03

#16 -2.5485D-01 -1.7710D-01 -2.5486D-01 -1.7711D-01

Page 25: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

25

P

δ

Pcr

δcr

a) b)

n=0; m=0;t0 = 0;∆t = ∆tbaseline;

tn+1 = tn +

Can r n+1 = 0 be solved, andis K n+1 positive definite?

n = n + 1

∆t = ∆t/4m = m +

Yes

No

m = mmax? Criticalstate

Yes

Figure 1. a) critical load and deflection associated with a structural response; and b) algorithm forfinding the first critical point of a structure’s response.

Page 26: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

26

Figure 2. Schematic to illustrate analysis problem reduction technique. Nodes with associateddesign variable values of zero are open circles; filled circles represent nodes with nonzero designvariable values; nodes with open squares denote “prime” nodes whose degrees-of-freedom arerestrained; elements with “S” are at least partially solid; elements with “V” are essentially devoid ofsolid material; elements with “P” are prime elements removed from consideration. a) shows a meshwith designated nodes and elements while b) shows the corresponding mesh used in analysis withsolid elements blackened, prime elements removed, and void elements retained. c) shows a meshwith an isolated solid node among a region of void nodes, and d) shows the resulting mesh in whichthere are no prime elements removed, thus avoiding formation of an island that is disconnected fromthe remainder of the mesh.

a) b)

c) d)

Page 27: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

27

Figure 4. Undeformed cantilever beam, and deformed shape in first buckling mode.

Figure 3. Undeformed and deformed cantilever beam under imposed displacement loading at node 14.Confirmed sensitivity of strain energy in the beam to that of selected nodal design variables is shownin Table 1.

Page 28: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

28

Figure 5. a) X-frame design problem where the structure of dimensions of 11 by 40 is loaded asshown and is modeled with a mesh of 1200 bilinear continuum elements. (b) Design solutionobtained by maximizing the minimum critical buckling load; (c) Design solution obtained bymaximizing the minimum critical load obtained using linearized buckling analysis; and (d) Designsolution obtained by minimizing the structural compliance up to the point of the first instability.

Page 29: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

29

Figure 6. a) circular structural domain modeled with coarse mesh, with rigidboundary conditions and point load or displacement applied at center; b) undeformeddesign solution obtained by minimizing structural compliance up to first critical pointwith structural material constraint at 20% of structural domain; c) deformedconfiguration of design solution; d) fine mesh and applied load; e) undeformed designsolution obtained by minimizing structural compliance up to first critical point withstructural material constraint at 2.5% of structural domain; and f) deformedconfiguration of design solution.

Page 30: Buckling sensitive structures

S. Rahmatalla and C.C. Swan

30

Figure 7. (a) Mesh of circle domain with rigid boundary restraints and central load. (b) Undeformedsolution obtained to maximize minimum critical nonlinear load with material constraint at 20% ofstructural volume and applied displacement loading 0.1Rδ ≤ ; (c) Deformed configuration ofassociated design; (d) Undeformed design obtained to maximize minimum critical nonlinear loadwith material constraint at 5% of structural volume and applied displacement loading 0.8Rδ ≤ ; (e)deformed configuration of associated design; (f) Design obtained to maximize minimum critical loadobtained with linearized buckling analysis; and (g) deformation associated with critical mode.

Page 31: Buckling sensitive structures

Continuum Topology Optimization of Buckling-Sensitive Structures

31

Figure 8. Fixed end beam problem. (a) Design problem with applied displacement loading for whichstructure is to be designed to maximize the minimum critical load computed via nonlinear analysis;(b) asssociated design solution; and (c) deformed shape of design solution. (d) Design problem withapplied force loading for which structure is to be optimized to maximize minimum critical loadcomputed with linearized buckling analysis; (e) associated design solution; and (f) deformed shapeassociated with buckling eigenmode.