You are on page 1of 16

Finite Elements in Analysis and Des:gn 6 (1989) 97-112 97

Elsevier

EFFECTIVE USE OF NUMERICAL O P T I M I Z A T I O N IN S T R U C T U R A L DESIGN

G.N. VANDERPLAATS
VMA Engineering, Goleta, CA 93/17, U.S.A.

Abstract. Numerical optimization, applied to structural design, has past its twenty-fifth year of development and
is now recognized as a powerful and practical design tool. The only impediments to its expanded development
and application appear to be a lack of general knowledge of optimization by the engineering community and the
incorporation of this capability into commercial available structure analysis codes.
Here, the concepts of numerical optimization will first be outlined to provide a common basis for discussion.
Next, the features of finite element structural analysis, that make it uniquely suited for integration with
optimization will be described. An understanding of both optimization and finite element analysis leads to the
observation that there are several effective ways to couple the two technologies. The various approaches to
coupling analysis and optimization will be discussed and it will be shown that, from the engineer's point of view,
each is effective for its own reasons, depending on the resources available and the desired results.
Examples are offered to demonstrate the power and generality of these methods. It is concluded that
numerical optimization is a technology whose time has come.

Introduction

Structural optimization is now approaching its thirtieth birthday, having been introduced by
Schmit in 1960 [9]. During that time, three distinct technologies have advanced simultaneously
to the point were fully automated structural design is now a reality [14].
The first of these is the finite element method of analysis, which was in its infancy in 1960.
Today, numerous large scale finite element analysis programs are commercially available and
are almost universally accepted as the analysis method of choice for practical structures.
Analysis capabilities are well beyond simple linear static analysis and include dynamic
response, stability analysis and nonlinear elasto-plastic analysis as examples. Additionally, both
steady and unsteady aerodynamics and thermodynamics are often included to provide a
multidiscipline analysis capability.
The second technology that is important here is the dramatic increase in speed and capacity
of computers, without which most modern structural analysis, and especially design optimiza-
tion, would not be possible. For example, in 1960, Schmit solved a simple three-bar truss design
example, considering two design variables and stress limits on the members under two separate
loading conditions. This problem required approximately one half hour of CPU time on an
IBM 653 computer [10]. Today, this problem is solved in about 6 seconds on a microcomputer
and under one-hundredth of a CPU second on a supercomputer. Although some of this
efficiency is due to advances in optimization technology, most of it is due to the speed of
modern computers.
The third technology that is key to the present status of design optimization is the advances
in the optimization algorithms themselves. Optimization methods account for from one to three
orders of magnitude efficiency improvement for general applications. Furthermore, in struct-

0168-874X/89/$3.50 © 1989, Elsevier Science Publishers B.V.


98 G.N. Vanderplaats / Effectie use of numerical optimization in structural desgin

ural design, special techniques can be employed to gain from one to two orders of magnitude
further improvement for many problems. In addition to the improvements in the algorithms on
a theoretical level is the fact that these methods can now be used by non-experts in
optimization. Because some of these algorithms are mathematically complex, and because (as
opposed to the finite element method) this technology is not normally taught in engineering
schools, the ability to provide packaged optimization programs for the practitioner is an
important key to its successful application.
Here, the purpose is to discuss optimization methods relative to structural design. Because
these methods are not well known to practicing engineers, we will first provide a brief
description of what numerical optimization is and how it addresses the engineering design
problem. Following this, we will discuss the finite element method to show how these two
technologies interact. Here we will see that, in many cases, the special mathematical structure
of the problem can be used to create a very efficient design tool. However, it will also be shown
that the state of the art does not allow us to solve all of the problems we would like to solve by
using these special techniques and that, even when we can, considerable modifications to the
finite element program are needed. This suggests that, for many design tasks, we must rely on
the efficiency of the optimization methods themselves and our ability to formulate the problem
for efficient and reliable solution. Finally, we will consider several design examples to
demonstrate the power of optimization in modern structural design.

Basic concepts

Because numerical optimization methods are unfamiliar to many, a brief description of the
optimization problem and solution methods is offered here. It will be seen that these methods
offer an extremely general and versatile tool for engineering design. Even though most modem
optimization methods have their origins in the operations research community, significant
contributions have been made by engineers. These include basic algorithm development,
software and problem formulation techniques.
Engineering design requires finding the best combination of variables describing a system,
subject to a variety of conditions or constraints that must be satisfied. Here, by best, we mean
the design that is minimum cost, minimum weight, maximum performance, or perhaps
maximum reliability, depending on the particular application.
The general optimization problem is stated as, find the set of design variables, X, that will
Minimize F ( X ) , (1)
subject to:
g;(X)<~O, j=l,m (2)
x , L ~ x , ~ x i U, i=1, n (3)
The function, F(X), is referred to as the objective or merit function and is dependent on the
values of the design variables, X, which themselves include member dimensions or shape
variables of a structure as examples. The limits on the design variables, given in (3), are referred
to as side constraints and are used simply to limit the region of search for the optimum. For
example, it would not make sense to allow the thickness of a structural element to take on a
negative value. Thus, the lower bounds would be set to a reasonable minimum gage size. If we
wish to maximize F ( X ) , for example, maximize the fundamental frequency of a structure, we
simply minimize the negative of F ( X ) . Typically, we can consider up to about twenty variables
contained in X if no special features of the problem are taken into consideration and on the
order of 100 or more variables with specially written programs.
G.N. Vanderplaats / Effectie use of numerical optimtzation in structural desgin 99

The g j ( X ) are referred to as constraints and they provide bounds on various response
quantities. The most common constraint is the limits imposed on stresses at various points
within the structure. Then if 8 is the upper bound allowed on stress, the constraint function
would be written as

°'~k - 1 ~< 0, (4)

where subscripts i, j and k represent element, component and load condition, respectively.
Other typical constraints include displacement, frequency, local buckling, system buckling
load factor, dynamic response, aeroelastic divergence and aeroelastic flutter, as examples.
Noting that most structures are required to support numerous sets of loads, and constraints
such as stress limits must be imposed at hundreds or even thousands of locations within the
structure, it is clear that the number, m, of such constraints can become very large.
Note that, in describing the objective and constraints, we have made no assumptions about
their mathematical nature. In practice, some conditions are necessary. First, it is assumed that
these functions are continuous with continuous first derivatives with respect to X. Second, the
design variables themselves are assumed to be continuous so that we must usually optimize and
afterwards choose nearby member sizes from tables of available sections if we are limited to
these. These limitations are usually considered to be more theoretical and practical. Minor
discontinuities in function values are usually no problem and choosing members of nearly the
same dimensions from available groups after the optimization is complete usually produces a
very near optimum design.
Perhaps what is more notable is the assumptions we have not had to make. First, we have
not assumed anything about the optimum. That is, we have no preconceived idea of what
constraints will control the design. Thus, it is common for optimization to produce a design
that is not consistent with our intuition and experience. This is particularly true when dealing
with composite materials and when designing subject to dynamic response or frequency
constraints. Secondly, no assumptions have been made about the mathematical nature of the
problem except for continuity (there is an underlying assumption of mathematical convexity,
but this is true for design in general). The functions may be quite nonlinear and may be implicit
functions of the design variables. The fact that we do not have to provide simple, explicit
relationships between the design variables and the objective and constraint functions is key to
our ability to deal with really complex problems. Indeed, it is even possible to evaluate the
functions experimentally so that, except to perform the optimization sub-task, a computer is
not always needed.
In general, to use optimization, it is assumed that we have a computer program available to
analyze any proposed design, X. We must begin with some design, X °, which will be modified
by the optimization process in search of the best possible design which satisfies all constraints.
Usually, the initial design is our best estimate based on experience. The optimization program
will call the analysis program (usually as a subroutine) to evaluate the objective and constraint
functions. In practice, it may call an existing program such as ANSVS which calculates the
needed information as responses. The optimization program will then convert this information
to the standard form required by (1)-(3). The optimization will repeatedly perturb the design,
calling the analysis each time, in search of the one with minimum objective function which
satisfies the constraints.
There are a great many optimization algorithms available. However, in one form or another,
most of them are based on a simple iterative formula:
x q = x q-1 + a * S q, (5)
where q is the iteration number, S q is a vector search direction and a* is called the move
100 G.N. Vanderplaats / Effectie use of numerical optimization in structural desgin

parameter. This has the physical interpretation that S is a vector direction in n-dimensional
space, defining how we wish to simultaneously change the design variables, X, and a* is the
number of steps (fractional steps are allowed) in this direction. Together, the term a*S q
represent the perturbation, 8X q, to be made in the design at this iteration. The distinction
between most optimization algorithms is in how the search direction, S q, is chosen and how the
"one-dimensional" search to determine the step size, a*, is performed.
The direct application of (5) is often too expensive for practical use. To begin with, it is
desirable to be able to efficiently calculate the sensitivity of the objective and constraint
functions to changes in the design variables. Optimization methods that use this information
are referred to as gradient based methods and are usually considered to be the most efficient.
Such information can sometimes be calculated within the finite element analysis program itself,
but at considerable effort and usually only with respect to a reduced set of design variable
types. Alternatively, it is possible to use both the information directly available from the
analysis, and our engineering judgement, to provide the optimization program with the data
needed to direct the design. Therefore, various approximation methods are used to improve
overall optimization efficiency, whether or not sensitivity information is directly available. To
better understand the various possible approaches within the structural optimization frame-
work, we will first review the aspects of the finite element method that are relevant to
optimization. Following that, we will describe the two dominate approaches to using optimiza-
tion with finite element analysis and suggest that a long term solution may be best found in
some combination of these methods.

Finite element analysis as a basis for optimization

The most common form of the finite element analysis method is the displacement method,
where we wish to find the set of displacements, u, that solves the system of equations;

[K]u+ [C]ti + [ M ] i i = F ( t ) , (6)


where u is assumed to be a function of time. Depending on the problem at hand, this general
form may be added to or part may be deleted. For example, for linear static, elastic analysis,
equation (6) reduces to the familiar form:
[K]u=P, (7)

where K is the master stiffness matrix, being the sum of element stiffness matrices, and P
contains the applied loads. Vector P may contain multiple columns associated with multiple
loading conditions. On the other hand, when considering static aeroelastic analysis, an
aeroelastic influence coefficient matrix is added to the K matrix in (6).
As compared to optimization, the finite element method of analysis is unique in that the
displacement method is almost universally accepted today. The force method has also been
used, but is now quite uncommon in practical use. This is not to suggest that the displacement
method is always best for design. The force method is particularly attractive for design of
structures such as trusses which are usually not highly redundant. Noting that many candidate
structures must be analyzed, the force method only requires solution of the compatibility
conditions for reanalysis when the design variables have changed, a task that is usually quite
efficient. However, the force method requires solution of a full non-symmetric set of simulta-
neous equations, which are not efficient on most computers. The displacement method, on the
other hand, requires solution of a set of equations using a sparse, banded, symmetric matrix
and this form is well suited to computer solution.
G.N. Vanderplaats / Effectie use of numerical optimization in structural desgin 101

On the other hand, optimization attempts to model the more heuristic and creative aspects
of design and ideally will model the human mind's ability to extrapolate, using minimal
information, to an improved design. Therefore, it is not surprising that there remain a
multitude of optimization algorithms for solution of the design problem. To a degree, these
algorithms are as varied in their approach and complexity as the minds of good design
engineers. Nonetheless, just as with designers, some algorithms have surfaced as the methods of
choice for a wide range of optimization problems [15]. Perhaps the most significant develop-
ments of recent years then have been, not in the development of new algorithms for solving the
standard problem of equations (1)-(3), but instead in the area of problem formulation. It can
be argued that, by creating the proper mathematical problem, relative to structural design, the
actual optimization algorithm becomes secondary. Thus, we have seen considerable effort
devoted to creating approximations to the design task that are of very high quality with respect
to the original problem, but which are well suited for optimization. The basic idea here is to
somehow first approximate the real problem with explicit functions. These explicit functions
are then used for optimization, returning to the full finite element analysis only to update the
approximation. There are two fundamental approaches used here. The first is to use sensitivity
information calculated within the finite element program to direct the optimization, and the
second is to use the results of the analysis itself, together with some form of curve fitting, to
provide the approximate model to be used in optimization.

Sensitivity analysis

Because the structural responses are implicit, nonlinear, functions of the design variables, it
might be assumed that the gradients (sensitivity) of the structural responses with respect to the
design variables cannot be readily found. Indeed, research in structural optimization continued
for approximately five years after its inception in 1960 before sensitivity of displacements and
stress with respect to sizing variables was analytically calculated. Today, we can calculate the
gradient of displacement, stress, frequency, buckling load factor, and even aeroelastic quantities
such as divergence and flutter speed with respect to structural design variables. However, this
information is available at considerable cost in program development. That is, it is necessary to
either make fundamental changes within the finite element code itself or supply some relatively
complex external programming to get the needed information [16,19]. To understand what is
required, consider the relatively simple task of calculating the sensitivity of displacements with
respect to design variables for static elastic analysis. We can implicitly differentiate (7) with
respect to design variable X, to get upon re-arranging.

- [K]- - ,, , (S)

where the rate of change of the master stiffness matrix, K, is the sum of the rates of change of
the elemental stiffness matrices. Note that, once a static analysis has been performed, K - t is
available in decomposed form so that (8) requires only some additional forward and backward
substitutions for its solution, just as in the case of including new load vectors. This is equivalent
to adding a number of new right-hand sides to the basic analysis equation equal to the number
of design variables times the number of load conditions for which gradients are required. Once
the rates of change of displacements are calculated, the rates of change of stresses are recovered
from the basic stress-displacement equations. Thus, at least mathematically, sensitivity of
displacements and stresses are straightforward in static analysis. Only minor extensions are
needed for sensitivity of other responses such as frequencies and buckling load factors.
However, experience has shown that the addition of this capability to a general purpose finite
102 G.N. Vanderplaats / Effectie use of numerical optimization tn structural des'gin

element program can be a major task. Also, this capability is only applicable to routine sizing
problems. When we consider geometric changes, fiber orientation in composites and nonlinear
response, it becomes clear that conversion of an analysis program to one which provides
sensitivity for use in optimization is a major, long term investment in a technology that is only
now becoming accepted by the general design community.
While it is beyond the scope of this discussion, it must be noted that sensitivity information
can actually be provided externally to the finite element program. A method known as the
adjoint method can be used to produce what are called pseudo-load vectors which are treated
as additions to the original load vectors in analysis. The needed information is the rates of
change of the elemental stiffness matrices with respect to the design variables. A third
approach, known as the material derivative approach is useful when changing the external
shape of a continuum structure during design and this method can also be used externally to
the basic finite element analysis program. In each case, considerable coding is required by
experts in the field and the available design variables are still limited relative to all of those
possible. For a comparison of the various methods of sensitivity analysis, refer to [19].
In summary, if we are designing relative to the most common and routine variables, being
sizing and in some cases shape variables, it is possible to include sensitivity analysis within the
finite element code. However. if we wish to consider the full array of design variables and
response quantities in the optimization process, we must consider alternative methods to couple
optimization with finite element analysis.

Practical optimization

In designing structures of realistic size and complexity, two fundamentally different ap-
proaches are presently of interest. These will be briefly described here. Noting that neither
method is the best for all cases, the possibility of somehow combining these basic methods will
be discussed and further research in this area will be encouraged.

Using sensitivity information


One approach is to modify the finite element program to provide sensitivity of the response
quantities with respect to the design variables. Once this is done, a variety of approximation
techniques are available by which this information can be manipulated in order to efficiently
direct the design process [3,8,11]. This approach has the advantage that relatively large numbers
of design variables can be considered, but the disadvantage is that the designer is limited to the
specific set of design variables that arc included in the sensitivity analysis capability.
In this method, the basic program structure is:

(1) Analyze the initial proposed design as a full finite element analysis.
(2) Evaluate all constraint functions and rank them according to criticality. Retain only the
critical and potentially critical constraints for further consideration during this design cycle.
(3) Call the analysis again to calculate gradients of the retained set of constraints.
(4) Using these linear approximations, create an optimization problem to be solved by a
general purpose optimization code, and solve it. Here, the linearization could be used directly
or it may be modified in various ways as described in [3,8,11]. During this approximate
optimization, move limits are imposed on the design variables to insure the reliability of the
approximation.
(5) Update the analysis data and call the analysis program again to evaluate the quality of
the proposed design, if the solution has converged to an acceptable optimum, terminate.
Otherwise repeat from Step 2.
G.N. Vanderplaats / Effectie use of numerical optimization tn structural desgm 103

From these five steps, it is clear that the analysis and optimization tasks are quite closely
coupled. This provides the greatest possible efficiency. However, to a large degree, it limits the
choice of design variables and precludes the use of judgement and experience in directing the
design process.
Despite its relative lack of generality, this approach is attractive for its efficiency and ability
to deal with large numbers of design variables. Also, the state of the art is developed to the
point that a large percentage of common design problems can be formally included in a finite
element program. While most such capabilities under development presently include only sizing
variables, it is clear that they will be extended to include shape variables in the foreseeable
future. Therefore, these codes can be expected to provide powerful, efficient design tools for a
wide range of problems.

Without using sensitit, ity analysis

A fundamentally different approach to design optimization is to couple an existing finite


element analysis code directly with an optimization code, without the use of analytical
sensitivity information. This approach is attractive for several reasons. First, it can be done
quickly and with little or no modification to either code. Second, the state of the art in
optimization has improved dramatically in recent years, leading to acceptable efficiencies for
problems where the payoff from optimization is great. Finally, and perhaps most important,
this approach provides a great deal of flexibility in choosing design variables and objective and
constraint functions.
Using this approach, there are two methods currently available. The first is simply to let the
optimization program calculate all needed sensitivity information by finite difference. This can
be expensive since each set of gradient calculations requires a number of analyses equal to the
number of design variables in addition to the analysis of the nominal design. Also, since
optimization is iterative, this must be repeated several times. However, used judiciously, this
approach can be competitive, particularly if only two or three iterations of the optimization
process are needed. For example, using only three optimization iterations, it would be necessary
to perform approximately 3 x n + 15 finite element analyses. For a common 30 design variable
problem, this would mean 105 detailed analyses. This would perhaps be prohibitive if the
analysis model used thousands of degrees of freedom or if the analysis was nonlinear. On the
other hand, if a relatively small "'Design Model" is used, this approach can indeed be cost
effective. Also, while three optimization iterations may appear to be too few (usually we expect
to use 10-20 iterations), it should be remembered that the majority of the design improvements
are made in the first few iterations.
The second method which does not use sensitivities provided by the finite element code is to
analyze several candidate designs and use statistical or curve fitting techniques to create an
approximate optimization problem. This approach has the advantages of allowing the engineer
to use his experience and judgement in proposing designs, as well as reducing the number of
analyses necessary to reach an optimum and has been successfully applied to aerodynamic
design, which is relatively nonlinear [13]. This method can best be understood by considering a
simple one-variable problem. If we choose three candidate values for that variable and evaluate
the corresponding objective and constraint functions we can fit a quadratic curve to each
function. Then we can call the optimization program to minimize this approximated objective,
subject to bounds on the approximated constraints. Since we are now solving an explicit
approximate problem, the optimization cost is minimal. Furthermore, if the true problem is
approximately quadratic, this will provide a very high quality design with only three initial
analyses. Indeed, we could use only two analyses and start with a linear approximation. The
104 G.N. Vanderplaats / Effectie tt~e of numerical optimization in structural desgin

proposed optimum could then be analyzed to provide the third point for a quadratic
approximation.
In the general case, n + 1 analyses are needed to create a first-order approximation and
1 + n + ~n(n + 1) analyses are needed to create a second-order approximation. Experience has
shown that this approach is limited to perhaps ten design variables, and beyond that number, it
is usually more cost effective, and more reliable to use finite difference gradients in the
optimization.
It should be noted that this method does not rely on the use of standard linear or quadratic
approximations. In practice, any form of curve fit can be used, and so it should be chosen as
one that best models the actual physics of the problem. The overall idea here is to use our
experience and engineering judgment to create a high-quality approximation without the use of
analytically derived sensitivities.

Future needs

From the above discussion, it is apparent that neither of the basic approaches to optimiza-
tion is the best for all cases. Using sensitivity information generated by the finite element code
provides efficiency, but with a limited set of design variables, objective and constraint
functions. On the other hand, without such sensitivities, we have the desired high degree of
generality, but with an associated loss of efficiency. This all suggests that some combination of
the above two approaches can provide the best of all worlds.
To now, little or no research has been directed at such a combined approach. Therefore, it is
problematic whether such an approach is possible and, if so, whether it can be of acceptable
efficiency. However, it is clear that some effort in this area is warranted.
A second need that is becoming increasingly apparent is the development of methods and
software for convenient generation of the "design model". A wide variety of software is
available for creating analysis models, both automatically and interactively. However, the
mathematical model defining the design problem is different from this so, for fully automated
design to reach its full potential, considerable efforts are needed to make it easy to use by the
practicing engineer.

Examples

Four examples are offered here to demonstrate the state of the art in design optimization.
Each has been used elsewhere to demonstrate the various methods and programs. The first
example is for a case where sensitivities are provided by the finite element program. In the
second example, sensitivity information is calculated externally to the finite element program,
and in the last two examples, approximation techniques are used without direct calculation of
sensitivities.

E x a m p l e 1." Marine gear housing

Figure 1 shows a large marine gear housing. The housing has two planes of symmetry
although the loading conditions are non-symmetric. The finite element model of one fourth of
the structure is shown in Fig. 2 and consists of 1623 elements, 7239 displacement degrees of
freedom, and six load cases. The design program coupled the MSC/NASTRAN program, Version
63 [7} with the CONMIN optimization program [12]. The finite element program contains the
G.N. Vanderplaats / Effectie use of numerical optimization in structural desgin 105

.0 cm

67
Fig. 1. Marine gear housing.

calculation of specified sensitivities, so this represents the internal sensitivity capability


described above.
The design problem included thirty sizing variables, being plate thicknesses and stiffener
cross-sectional areas. There are 5620 nonlinear inequality constraints, including stress limits

Fig. 2. Finite element model.


106 G.N. Vanderplaats / Effectie ~tveof pumerical optimization in structural desgin

5500 I'
I (5360)

500(

"~ 4500
v

4000
(3850)

:-- 3500 (3350)

(3200)
(3130) (3110)

(2870)

2500'0 ~ ~ 3. . . 4. 5 v
ITERATION NUtIBER
Fig. 3. Iteration history.

under each loading condition. Also quite stringent deformation constraints were imposed,
including journal bearing out-of-roundness, transverse rotation, and center-to-center displace-
ments. Some of the deformation constraints were initially violated by nearly 100%, and so the
initial design was scaled up to give a near feasible starting point for optimization. The design
process consisted of the five steps described above for use with sensitivity information. This
problem is described in more detail in [18].
The iteration history is shown in Fig. 3, where one iteration consists of Steps 2 through 5
(Step 1 is for the initial design only). During the optimization, only 150 of the most critical
constraints were retained for sensitivity calculations and use in the approximate optimization
phase. The optimization required five such design cycles, and each cycle required 600 CPU
seconds on a Cray ls supercomputer. The total design thus required just under one hour of
computer time. While this design problem is considered to be of significant size, problems in
this general category of three or more times this size are considered well within the state of the
art in structural optimization.

Example 2." Engine connecting rod

Figure 4 shows an engine connecting rod that was designed for minimum weight [1]. The
structure has two planes of symmetry, and the finite element model is shown in Fig. 5. The
model consists of 105 solid elements, 928 nodal points and 2126 displacement degrees of
freedom for the one fourth model. The design model consisted of eight variables, b~-b8, shown
in Fig. 4. This problem was solved with MSC/NASTRAN coupled with the ADS optimization
program [17]. However, in this case the sensitivity information is not directly available from the
finite element program. Instead, this was calculated as a post-processing operation using the
G.N. Vanderplaats / Effectie tL~eof numerical optimization in structural desgin 107

- - Inilial design
.... Final design

Section A-A
~ X l x-A~ x

U I ' ~.--:tq;[ t . . . . f---~ I '~


AJ x6 Xo Fig. 4. Engine connecting rod (note:
Xl - Xs are design variables).

Fig. 5. Finite element model.

16000

14001?

12000~
E
E

~ 10000-

8000-

6ooo 1
o ~ ~
Itaration Number
Fig. 6. Iteration history.

material derivative method [4,19]. This method is quite useful for shape optimization problems
such as this. In practice, this method for calculating sensitivity information could be included
internally in the finite element program.
The iteration history is shown in Fig. 6, where an iteration consists of the same steps as in
the previous example, except the calculation of sensitivity information is external to the finite
element program. This design required twenty iterations, although it is clear that a practical
design was obtained with as few as six. This efficiency is quite good, considering the complexity
108 G.N. Vanderplaats / Effectie use of numerical optimization in structural desgin

fluid pressure being sensed diaphragm not shown


A
1 t t I l t l l t <

diaphragm
• . ~ header showing view
neaaer ~ for following pages
weld Jj ~ s14=~
- - electrical leads Fig. 7. Pressure transducer header.

of the three-dimensional shape optimization program and the fact that tree-dimensional shape
optimization is a relatively new area of application.

Example 3: Dual material pressure transducer header

Figure 7 shows a pressure transducer header that is designed for minimum cost, subject to
manufacturing and product compatibility side constraints [6]. Additional constraints were

h I = 1.32(''T'I'/4T'~)
(a) , w2

q 3 " A3h3 ~_.T~1/4


% - 1.421 h )

q 2 " A2h2 Stainless Steel


RasUc

Qweld
ql " A l h l *T

(b) pressure - 200 plfl


I
q.

_. \ /
t.

Fig. 8. Design model and analysis boundary conditions. (a) Thermal transient boundary conditions. (b) Stress analysis
boundary conditions.
G.N. Vanderplaats / Ef/ectie use of numerical optimization in structural desgin 109

I ~ Stainless Steel
~asUc

INITIAL DESIGN O P T I M U M DESIGN


Fig. 9. Initial and final designs.

imposed on stresses, temperature limits during welding, and debonding of the material
interface. There are 3 independent design variables, shown as h, Wl, and WE in Fig. 8. Figure 8
also shows the analysis model boundary conditions. This problem was solved using the ANSYS
program [2] using the approximation methods without sensitivity information. The initial and
optimum designs are shown in Fig. 9 and the thermal transient history is shown in Fig. 10 for
the initial and optimum designs.
The unique feature of this example is that the finite element analysis program and the
optimization program were coupled to provide the ability to solve a program that cannot
directly be solved by other structural optimization programs. The optimization required 23
detailed finite element analyses and reduced the objective function by 79 percent. The initial
design was provided by the customer. This is clearly cost effective, both from the standpoint
that the cost was greatly reduced and that the initial design violated several thermal constraints.

TEMP
('FI ('F) , Two nodes at ~ of
100o1-
900 1-
800 I- ( ~ ~ Two nodel at b o r d ~ of S.S.

700500I'-I-
600 l- ~ , ~ ~ / f l~astlc melting t e m p . 3OO
/ ,
_ _ - ~ - -
4OO I-- 2OO~ ~tdll'1~ i
300 I-
I/ 100
100 E/
t 1 t I i I
, , I , ' I (HOUr) .0o08 .0o16 .oo24 ,0032 .oo40 (HOURS)
•0002 .0004 .0006 .0008 .0010

Fig. 10. Thermal transient history. Left: initial design; right: near optimum design.
110 G.N. Vanderplaats / Effectie tt*e of numerical optimization in structural desgin

t2~ t4

560elements
2898 d.

DesigVari
n ables F i g . l 1. E n g i n e p i s t o n .

Example 4: Shape optimization of a piston

As a final example, the piston shown in Fig. 11 was designed for minimum weight [5]. The
figure shows the four design variables and the finite element model of the one-quarter model of
the structure. Three separate loading conditions are considered and stress constraints under
each loading condition, as well as geometric constraints, are imposed on the design. The initial
and final design are shown in Fig. 12, representing a 35% reduction in weight. Figure 13 shows
the iteration history. While a total of thirty finite element analyses are used, it is notable that
the optimization had converged to an acceptable design much sooner than this. This example
was created to demonstrate the optimization capability and is not intended to encompass all of
the details that could be included in a practical design. This problem was also solved by the

Initial Design Final Design

F i g . 12. L e f t : i n i t i a l d e s i g n ( q = t 2 = t 3 -
t 4 = 0 . 2 5 " , w e i g h t = 1.01 lb). R i g h t : f i n a l
design (q = 0.135", t, = 0.159", t3 =
0.260", t4 = 0 . 1 3 5 " , w e i g h t = 0 . 6 4 lb).
G.N. Vanderplaats / Effectie ~s'e of numerical optimization in structural desgin 111

1.0~
Light
0.4 1 I I I T 1 I
Fig. 13. Iteration history (arrow indicates
5.00 10.4 15.8 21.2 26.6 32 Loop convergence with default tolerances).

ANSYS program and again demonstrates the flexibility of the use of approximation techniques
using function values only.

ConcLusions

The purpose here has been to present optimization as a powerful and versatile tool for
structural design and to identify the present state of the art. Examples have been offered to
demonstrate that structural optimization is a reality and can be valuable in providing guidance
in the design process. This has come about from the simultaneous maturing of the finite
element method itself, computer technology, and optimization algorithms and software. It is no
longer necessary for the user of optimization to be a theoretician as well; he can use
optimization together with good design skills, to dramatically enhance both design quality and
productivity.
Two fundamental approaches to optimization have been offered. One is to modify the finite
element program to provide information directly useful to optimization. This provides the most
efficient design tool, but at the price of limiting the generality to which it applies. The second is
to couple the finite analysis program to an optimization program such that any input parameter
can be treated as a design variable and any response can be treated as an objective function or
constraint. This approach has the advantage of extreme generality at the expense of limiting the
number of design variables that can realistically be considered.
In the long term, it is apparent that some compromise between these two approaches is
desirable. That is, the finite element program may provide sensitivity where it is convenient but
otherwise this information may be obtained by other means Obviously, if all sensitivity
information is to be calculated by finite difference methods, then all objective and constraint
functions are allowed. However, this is expensive and so approximation methods whereby
several candidate designs are used to perform some type of curve fit to the functions becomes
the desired approach.
The basic issue then is whether it is possible to have the best of both worlds; use direct
sensitivity when possible, but augment this with other information when necessary. In any
discussion of structural optimization it is important to identify the needs for the future, and
this appears to be the logical need in the present context. The answer is not obvious since the
basic approaches seem to be conflicting. However, the question is a reasonable one that
deserves attention by the research community.
112 G.N. Vanderplaats / Effectie ~se of numerical optimization in structural desgin

In s u m m a r y , we r e p e a t the basic s t a t e m e n t t h a t o p t i m i z a t i o n is a t e c h n o l o g y w h o s e t i m e has


c o m e . C o n s i d e r i n g the i m m e n s e c o m p u t a t i o n a l p o w e r a v a i l a b l e t o d a y , as well as the relatively
m a t u r e state o f the art, the m o t i v e s to use o p t i m i z a t i o n are c o m p e l l i n g . O p t i m i z a t i o n d o e s not
r e p l a c e b a s i c e n g i n e e r i n g c r e a t i v i t y or d e s i g n j u d g m e n t . O n the c o n t r a r y , used by k n o w l e d g e a -
ble design engineers, o p t i m i z a t i o n c a n be the tool that p r o v i d e s the c o m p e t i t i v e edge!

References

[1] BOTKIN, M.E., R.J. YANG and J.A. BENNET, "Shape optimization of three-dimensional stamped and solid
automotive components", in The Optimum Shape: Automated Structural Design, Plenum Press, New York, pp.
235-262, 1986.
[2] DESAI.VO,G.J. and J.A. SWANSON, ANSYS Engineering Analysis System Users Manual, Swanson Analysis System,
Inc., June 1985.
[3] FLEURY, C., "Shape optimal design by the convex linearization method", in: The Optimum Shape: Automated
Structural Design, Plenum Press, New York, pp. 297-326, 1986.
[4] HAUG, E.J., K.K. CHOI and V. KOMKOV, Design Sensitivity Analysis of Structural Systems, Academic Press, New
York, 1984.
[5] IMGRUND,M.C., "Using a parametric design language to couple design optimization with analysis", Proc. Second
Optimization Users Conf, Santa Barbara, CA, August 1986.
[6] IMGRUND, M.C. and M.J. WHEELER, "Reducing design costs by integrating finite element and optimization
techniques", Proc. Spring Natl. Design Conf., Chicago, IL, March 24-27, 1986.
[7] MSC/NASTRANUser's Manual Version 63, The MacNeaI-Schwendler Corporation, Los Angeles, CA.
[8] SALAJAGHEH,E. and G.N. VANDERPLAATS,"An efficient approximation method for structural synthesis with
reference to space structures", Int. J. Space Struct. 2(3) pp. 165-175, 1986/1987.
[9] SCHMIT,L.A., "Structural design by systematic synthesis", Proc. 2nd Conf. on Electronic" Computation, ASCE, New
York, pp. 105-122, 1960.
[10] SCHMIT, L.A., "Structural synthesis--its genesis and development", AIAA J. 19(10) pp. 1249-1263. 1981.
[lll SCrtMIT, L.A. and H. MIURA, Approximation Concepts for Efficient Structural Synthesis, NASA CR-2552, 1976.
[12] VANDERPLAATS,G.N., "COMIN--a FORTRANprogram for constrained function minimization: User's Manual".
NASA TM X-62,282, 1973.
[13] VANDERPLAATS,G.N.. "Efficient algorithm for numerical airfoil optimization". AIAA J. Aircraft 16 (12), pp.
842-847, December 1979.
[14] VANDERPLAATS,G.N., "Structural optimization--past, present and future", AIAA J. 20(7), pp. 992-1000, July
1982.
[15] VANDERPI.AATS,G.N., Numerical Optimization Techniques for Engineering Design: with Applications, McGraw-Hill,
New" York, 1984.
[16] VANDERPLAA'rs,G.N. and H. MXURA,"Trends in structural optimization: some considerations in using standard
finite element software", Proc. 6th Vehicle Structural Mechanics Conf., Detroit, MI, April 1986.
[17] VANDF.RPLAATS,G.N. and H. SUGXMOTO,"A general-purpose optimization program for engineering design", Int.
J. Comput. Struct. 24(1), pp. 13-21, 1986.
[18] VANI)ERPI.AATS.G.N., H. MIURAand M. CHARGIN."Large scale structural synthesis", Finite Elements in Ana(vsis
and Design, 1(2), pp. 117-130, 1985.
[19] YANG. R.J. and M.E. BOTKIN. "The relationship between the variational approach and implicit differentiation
approach to shape design sensitivities", Proc 26th A IAA / A S M E / A S C E / A t l S Structures. Structural Dynamics and
Materials Conf., Orlando, FL, 1985.

You might also like