Sei sulla pagina 1di 60

UNIVERSITY OF CALIFORNIA, SAN DIEGO

Weight Reduction Techniques Applied to Formula SAE Vehicle Design:


An Investigation in Topology Optimization

A thesis submitted in partial satisfaction of the

requirements for the degree Master of Science

in

Engineering Sciences (Mechanical Engineering)

by

Lucas V. Fornace

Committee in Charge:

Professor Frank E. Talke, Chair


Professor David J. Benson
Professor Hidenori Murakami

2006
ii

Copyright

Lucas V. Fornace, 2006

All rights reserved

ii
iii

The thesis of Lucas V. Fornace is approved:

_______________________________________

_______________________________________

_______________________________________
Chair

University of California, San Diego

2006

iii
iv

This work is dedicated to my younger sister and recent UCSD admit,

Gabrielle Angeline Fornace

May she continue along and improve upon the path set forth by her big brother, with
confidence and determination aplenty.

In addition, I would like to dedicate the completion of this thesis, and thus the
conclusion of my academic career to

Carrie Rose Martin

A dear friend whose companionship surely propelled my own drive and confidence,
and to whom I credit most of my successes. For this, I am forever grateful.

iv
v

“Any car which holds together for a whole race is too heavy.”

Colin Chapman, Founder, Lotus Engineering Co.

v
vi

TABLE OF CONTENTS

Signature Page……………………………………………………………………..….iii
Dedication…………………………………………………………………………......iv
Epigraph………………………………………………………………………………..v
Table of Contents…………………………………………………………………..….vi
List of Figures and Tables………..………………………………………………….viii
Acknowledgements………………………………………………………………….....x
Abstract………………………………………………………………………………..xi

I. Introduction…………………………………………………………………………..1
A. Formula SAE Competition……………………………...………………………..1
B. Recognition of Need for Reduced Vehicle Mass…...…...……………………….3
C. Concepts of Topology Optimization……………………...……………………...5
D. Problem Definition………………………………………...……………………..7

II. Pre-processing……………………………………………………………………..12
A. Load Prediction……………………………………………...………………….12
1. Multi-body Dynamics……………………………………..………………….12
2. Load Cases…………………………………………………..………………..14
B. Design Space…………………………………………………...……………….16

III. Topology Optimization…………………………………………………………...17


A. Design Space Mesh……………………………………………………………..17
1. Geometry Clean-up…………………………………………………………...17
2. Tetra-Meshing………………………………………………………………...17
B. Boundary Conditions……………………………………………………………19
1. Treatment of Bolt Holes………………………………………………………19
2. Constraints………………..…………………………………………………...19
3. Loads…………………………..……………………………………………...21
C. Optimization Statement…………………………………………………………23
1. Optimization Responses………………………………………………………23
2. Material Definition……………………………………………………………25
3. Manufacturing Constraints……………………………………………………25

IV. Post Processing…………………………………………………………………...27


A. Topology Optimization Results………………………………………………...27
B. Geometry Interpretation………………………………………………………...30

V. Design Validation………………………………………………………………….32
A. Finite Element Analysis………………………………………………………...32
B. Bell Crank Prototype Fabrication……..…………………………...…….……...33
C. Physical Testing………………………………………………………………...35

vi
vii

1. Hydraulic Test Fixture Design………………………………………………..35


2. Experimental Results…………………………………………………………40

VI. Conclusion………………………………………………………………………..42
A. Discussion of Experimental Results………………………………….………...42
B. Overall Vehicle Performance and Awards…………..…………………………44

Bibliography…………………………………………………………………………..47

vii
viii

LIST OF FIGURES AND TABLES

Figure 1: Diagram of a typical pushrod and bell crank suspension……………………9

Figure 2: Bell crank as installed on 2005 vehicle…………..………………………...11

Figure 3: Close-up view of bell crank as installed on 2005 vehicle………………….11

Figure 4: MBD model of rear suspension………………………………………….…13

Figure 5: Bell crank load case in left vehicle roll condition………………………….15

Figure 6: Bell crank load case in right vehicle roll condition………………………...15

Figure 7: Design space and spring/damper system.…………………………………..16

Figure 8: Geometry cleaning in preparation for meshing...…………………………..18

Figure 9: Tetra-meshed design space…………..…..…………………………………18

Table 1: Summary of bell crank constraints……………………………..…………..21

Table 2: Summary of bell crank load cases…………………………..….…………..22

Figure 10: Applied boundary conditions………………..…………………………...22

Figure 11: 2005 bell crank finite element analysis displacement results….…………24

Figure 12: Topology optimization results, side view………………………………...27

Figure 13: Topology optimization results, isometric view…………………………..28

Figure 14: Topology optimization results with draw constraint, isometric view……29

Figure 15: Optimized geometry interpretation, isometric view……………………...31

Figure 16: von-Mises stress distribution and exaggerated displacement.……………33

Figure 17: MasterCAM X tool paths………………………………………………...34

Figure 18: Bell crank prototype fabrication………………………………………….34

Figure 19: Bell crank prototype finished product……………………………………35

viii
ix

Figure 20: Hydraulic test fixture CAD assembly…………………………………….37

Figure 21: Hydraulic test fixture prepared for experiment…………………………..39

Table 3: Summary of experimental test results………………………………………41

Figure 22: Physical testing of 2006 optimized bell crank…………………………...41

Table 4: Comparison of 2005 and 2006 UCSD FSAE vehicle performance………...45

Figure 23: 2006 UCSD FSAE vehicle at competition………….……….…………...45

Figure 24: Close-up of bell crank on 2006 vehicle…………………………………..46

Figure 25: 2006 vehicle in motion…………………………………………………...46

ix
x

ACKNOWLEDGEMENTS

I would like to thank Professor Frank E. Talke for advising this work and for

allowing me the freedom to pursue this optimization project. His advice and

leadership was plentiful and much appreciated.

I would also like to acknowledge Terence Smith of Altair Engineering for his

relentless support of the relationship with UCSD and UCSD Formula SAE.

Furthermore, a thank you is in order to my colleagues in the CMRR for their

support; namely, Maik, Bart, Aravind, D.E., Ralf, John, Paul, Max, Mathias and

Thorsten. Thank you all for your suggestions and friendship.

Robert Shanahan deserves a huge thank you for his unparalleled dedication to

the Formula SAE team at UC San Diego as community advisor, as does Professor

Keiko K. Nomura for her contributions as Faculty Advisor. The team would not be

able to operate at the current level without their tremendous efforts.

Furthermore, I would like to acknowledge Billy Wight for his skilled work in

fabricating the prototype bell cranks for this project, not to mention just about every

other component on the 2004, 2005 and 2006 vehicles. In addition, I want to thank

Tom Chalfant and Dave Lischer in the MAE machine shop for their continued support

of the Formula SAE team at UCSD.

Lastly I want to thank my family and friends for putting up with my hectic

schedule, trusting my decisions, and supporting me along the way—a sincere thank

you to everybody who contributed.

x
xi

ABSTRACT OF THE THESIS

Weight Reduction Techniques Applied to Formula SAE Vehicle Design:


An Investigation in Topology Optimization

by

Lucas V. Fornace

Master of Science in Engineering Sciences (Mechanical Engineering)

University of California, San Diego, 2006

Professor Frank E. Talke, Chair

In the quest for reduced vehicle mass without sacrificed integrity, Computer

Aided Engineering (CAE) topology optimization software was investigated and

utilized in the design of the 2006 UC San Diego Formula SAE vehicle as a means to

determine the optimum material distribution within a component for a given set of

loading and boundary conditions. This paper looks at the design of a rear suspension

bell crank component using modern topology optimization techniques and compares

the end product to that of the 2005 model bell crank component, which was designed

using more traditional techniques. A hydraulic load cell system was created to

simulate the vehicle suspension forces and was used to physically test the original and

xi
xii

optimized parts to failure. Through the use of Altair OptiStruct® topology

optimization software, a weight savings of 24.3% coupled with an increase in yield

strength of 29.7% was realized in the optimized design of the 2006 bell crank.

xii
1

I. Introduction

A. Formula SAE Competition

Formula SAE is an international design competition held annually by the

Society of Automotive Engineers, and is described by the organization as follows:

The Formula SAE® competition is for SAE student members to conceive,


design, fabricate, and compete with small formula-style racing cars. The
restrictions on the car frame and engine are limited so that the knowledge,
creativity, and imagination of the students are challenged. The cars are built
with a team effort over a period of about one year and are taken to the annual
competition for judging and comparison with approximately 120 other
vehicles from colleges and universities throughout the world. The end result is
a great experience for young engineers in a meaningful engineering project as
well as the opportunity of working in a dedicated team effort [1].

This competition first began in Texas in the year 1981 with only four vehicle

entries. The event has since grown and moved to Detroit, Michigan, with a

consistently filled capacity of 140 vehicles that arrive annually to compete in the four

day affair. Over the years, the competition has seen a tremendous growth in

popularity. To satisfy the demand of the students, similar competitions have since

been created in Great Britain, Australia, Brazil, Italy and California. Even with the

addition of the new events, registration of the 140 available slots for the 2006 Detroit

competition filled in just 34 minutes. Needless to say, the event has grown quite

competitive.

The scoring of the competition is broken down into eight distinct events: the

first three of which are static, leaving the final five categories for dynamic

performance evaluation. Before any scoring begins, however, the student built

1
2

vehicles must first pass a thorough technical inspection which includes the vehicle

being laterally tilted to 61 degrees. This angle simulates a lateral acceleration

equivalent to 1.5 g’s to ensure that the car does not show any tendency to roll over or

leak fuel during hard cornering. Next the cars must demonstrate adequate braking

performance by being able to lock all four tires while driving at a moderate velocity.

Afterwards, the vehicle must pass a noise inspection to ensure compliance with the

sound level regulations. Fulfillment of this exercise marks the end of the technical

inspection.

Static scoring events include a marketing presentation, a cost feasibility

analysis, and most importantly, a design presentation in which the overall vehicle is

scrutinized by experienced automotive engineers. It is here that the students have a

chance to explain and defend their design rationale and field questions from the

various judges.

The dynamic scoring events include an acceleration event, a skid-pad event, an

autocross event, and an endurance event that is coupled with a fuel economy

measurement.

The acceleration event is a standing 75 meter sprint that is meant to

characterize the vehicle’s straight-line acceleration capabilities, which is a measure of

the vehicle’s inherent power-to-weight ratio and traction capacity.

To measure the car’s lateral acceleration abilities, a “skid-pad” circular path is

used in which the cars are in a constant radius turn at essentially steady state

2
3

conditions. This is a measure of the team’s ability to design and tune the vehicle

suspension for maximum lateral balance and traction.

The Autocross event is a timed driving event that takes place on a temporary

track typically laid out with pylons on a large paved surface in the formation of very

tight and technically challenging course. These autocross-style courses effectively

measure the overall performance package of the vehicle, including braking,

acceleration and cornering.

The largest and most heavily weighted of the contests is the endurance and fuel

economy event. Here the vehicles compete on a larger autocross-style course

approximately one kilometer in length. Twenty laps mark the duration of the event,

with a timed driver change on the 11th lap. In addition, the fuel level is measured, both

before and after the event, to obtain the effective fuel economy of the vehicle—further

adding to the challenge.

B. Recognition of Need for Reduced Vehicle Mass

Typical in high performance automotive and aerospace applications is the

demand for reduced vehicle mass while maintaining adequate performance and safety.

Formula SAE competition is no exception. The nature of the autocross style course

favors vehicles with good acceleration capabilities, and both the fuel economy and

acceleration events add to the desire for a lightweight vehicle design.

To increase the acceleration potential of a vehicle, or racecar in this case,

physics tells us that the traction force needs to be increased and/or the mass decreased

3
4

[2]. In Formula SAE competition, safety mandates require all of the engine’s air

supply to pass through a 20 mm diameter restrictor in an effort to limit the intake

volumetric flow rate. This puts a ceiling on the engine’s power capability, thus

reduction in vehicle mass becomes increasingly important, if not the only option in the

drive for improved acceleration performance. Furthermore, FSAE competition has no

minimum weight requirement, so the weight reduction benefits essentially have no

bound. Even in classes of racing where minimum weight mandates do exist, having

an underweight vehicle entry allows the engineer to put ballast in strategic locations,

such as down low on the floor pan or to a targeted location in an effort to aide or

correct vehicle handling and balance.

The aerospace world has an even greater desire for weight reduction, as any

given reduction in mass can be directly correlated to increased fuel savings and

increased vehicle range, as well as improved performance and agility [3]. Overall, as

long as the reduced mass does not pose any threats to the integrity of the design—be it

aircraft or racecar—there are rarely any drawbacks associated with decreased vehicle

mass.

For the 2006 UC San Diego Formula SAE team, the drive for weight reduction

stemmed from the 2005 competition, where the majority of the top-placing vehicles

were below 227 kg (500 lbs) in weight. The 2005 UC San Diego entry was slightly

heavy at 246 kg (543 lbs), thus it was made the primary goal to achieve a 2006 design

that was less than 227 kg (500 lbs). This would represent a weight loss of roughly

10% of the entire vehicle.

4
5

C. Concepts of Topology Optimization

Structural design problems require the fulfillment of specific objectives while

satisfying a set of performance constraints [4]. In the automotive and aerospace

worlds, this traditionally requires numerous iterations using finite element analysis

results that drive design changes, and the final design is often arrived at via a trial-and-

error approach [5]. This iteration process can prove to be a very time-consuming and

thus costly endeavor. While this technique has obviously been effective in meeting

design requirements, the final solution does not necessarily represent the best

solution—only one that has successfully met the objective. Furthermore, the quality

of the final design relies heavily on the quality and potential of the initial design

attempts. This is due to the fact that, as the design progresses, the freedom to make

significant changes diminishes over time. Therefore, in the interest of time and

quality, it is very important to have a good initial design solution early in the process.

Another drawback of the traditional design method is the fact that engineers

tend to think intuitively. Sometimes the optimum solution can be quite counter-

intuitive, and thus a great solution can go justifiably overlooked because it does not

seem plausible or reasonable. On the contrary, a design can include an overly

complicated network of reinforcing ribs, for example, that were deemed necessary by

the design engineer, when the ideal optimized solution is actually much simpler [6].

Topology optimization is a relatively new numerical method used to determine

the optimum shape and distribution of material within a given design space for a given

set of design constraints based on responses obtained from a finite element analysis

5
6

[7]. For structures under static loading, the basic finite element equation that is solved

can be expressed as

Kd=f (1.1)

where K is the effective stiffness matrix of the structure, d is the displacement vector,

and f is the load vector as applied to the structure. This equation is often referred to as

the matrix equivalent form of the Bubnov-Galerkin equation, and it represents

equilibrium of external and internal forces. Once the unknown nodal displacements

are solved for, Hooke’s law can be used to calculate the material stresses for

deformations in the elastic range. Hooke’s Law can be stated as

σ=Cε (1.2)

where σ is the stress vector, C is the material elasticity matrix, and ε is the strain

vector. Interested readers should consult [8] for further discussion of the linear static

finite element method.

The OptiStruct® topology optimization algorithm solves a structural

optimization problem in which the goal is to minimize an objective function subject to

constraint functions comprised of finite element responses [7]. Formally, the problem

can be written as follows:

Objective: Minimize W(x) (1.3)

Constraints: g(x)-gupper ≤ 0 (1.4)

Design Variables: xlower ≤ x ≤ xupper (1.5)

where the objective function W and the constraint functions g are structural responses.

Typical structural responses used to define the objective and constraint functions

6
7

include mass, volume fraction, compliance, frequency and displacement. Commonly,

the objective for a topology optimization study is to minimize the mass subject to

either nodal displacement or eigenfrequency constraints, with element density as the

design variable. By allowing the normalized element density to vary between 0 and 1,

the end result of the topology optimization should be a mesh in which the elements

take on a density value of either 0 or 1. A low density value represents a void, and a

high density value indicates solid material. By masking elements of low density, the

shape of the optimized structure is revealed. Overall, the optimization algorithm

involves discretizing the design space into a finite element mesh, calculating the

elemental material properties, and iteratively altering the material distribution

(element density) and re-calculating until convergence is reached at a solution that

best meets the objective function.

This technique can be used very early in the design stage to ensure that the

final design of the structure not only meets the requirements, but represents the

mathematically best solution based on the design constraints. Today the method is

widely accepted for bracket-type structures and has already proven to be a huge

benefit to many of the large aerospace and automotive corporations [5-6].

D. Problem Definition

With the need for weight reduction in mind as well as the desire to explore

topology optimization techniques, the general problem statement for the 2006 UCSD

Formula SAE team was to redesign and thus remove mass from an existing Formula

7
8

SAE vehicle component through the use of modern optimization software. The 2005

rear-suspension bell crank component was chosen as the test specimen for this project.

A bell crank is defined as a type of lever whose two arms form a right angle, or

nearly a right angle, having its fulcrum at the apex of the angle [9]. It received the

name from its first use which was to change the vertical pull of a rope into a horizontal

pull on the striker of a bell [10]. Bell cranks can also be found in the suspension

system of formula-style racecars. Because formula-style vehicles have their

suspension control arms exposed by protruding through the body panels, there are

aerodynamic reasons for relocating the spring and damper assembly to a location on

the vehicle that is contained within the bodywork. This set-up is distinctly different

from that of a standard production car (where the springs and dampers are traditionally

attached directly to the suspension control arms), and requires the use of a push or pull

rod and bell crank to transmit the suspension forces from the control arm to the spring

and damper. In addition to the aerodynamic benefits associated with relocating the

spring/damper assembly, the use of a bell crank gives the suspension design engineer

added control over the vehicle’s effective spring rate via manipulation of the bell

crank’s geometric parameters. A diagram showing a typical pushrod and bell crank

suspension is given in Figure 1.

8
9

Figure 1: Diagram of a typical pushrod and bell crank suspension [11].

The 2005 UCSD Formula SAE rear-suspension bell crank component was

chosen as the test specimen for the initial investigation of the topology optimization

for many reasons:

1) Because it transfers all of the suspension spring and damper forces for one

corner of the car as part of a four-bar linkage, the bell crank component is exposed to

high stresses, making the optimization results both critical and exciting.

2) Due to the 2005 rear suspension packaging (which remained the same for

2006), the bell crank geometry required a cut-out to clear the spring and damper

assembly. This feature makes the design problem more challenging, as the cut-out lies

in a potential load path between the spring connection and the connection for the rear

9
10

anti-roll bar pushrod. It was reasoned that this could possibly highlight a counter-

intuitive optimization result.

3) As part of a four-bar linkage, the bell crank is connected to truss-like

members of the suspension system, thus all of the loads can be thought of as simple

force vectors which can be predicted and represented with relative ease. This puts the

focus on the power of the software as opposed to the team’s ability to gather boundary

conditions (which can be quite complicated, as in the case of a wheel, for example).

4) Being that the rear suspension design is of the pushrod type, the rear bell

cranks are located in a position that is highly visible on the racecar, making it easy for

a person to identify the research component of interest (as opposed to the front

suspension bell cranks, which lie hidden under the drivers legs due to the pull-rod

style front suspension). Figures 2 and 3 show the rear bell cranks as designed and

installed on the 2005 competition-year vehicle.

5) Finally, the 2005 bell cranks were designed using a traditional trial-and-

error finite element approach, therefore making for a good comparison to the methods

utilizing topology optimization early in the design process.

After deciding upon a test article, the problem statement was to reduce the

mass of the rear bell crank component by at least 10% while maintaining the yield

strength of the previous design. Furthermore, due to the criticality of the component,

physical testing was deemed necessary to validate and objectively asses the integrity

of the design under static loading conditions similar to and beyond those seen in actual

use.

10
11

Figure 2: Bell crank as installed on 2005 vehicle (circled).

Figure 3: Close-up view of bell crank as installed on 2005 vehicle.

11
12

II. Pre-processing

A. Load Prediction

1. Multi-body Dynamics

To get a better understanding of the loading conditions at the component of

interest, a Multi-body Dynamic model (MBD) was created to accurately capture the

geometry of the vehicle’s rear suspension. Altair MotionView® software was used as

a mechanical system simulation tool with which the entire rear suspension of the

racecar was created in a 3-D environment. This enabled the extraction of the spring

and anti-roll bar force vectors on the bell crank component due to different vehicle

suspension conditions, such as full spring compression or full body roll.

The model was created using 3-D suspension and chassis coordinates from the

Formula SAE suspension design team inserted manually into MotionView® as points

in space [12]. The MBD software then easily allows for the creation of cylindrical

bodies between the points, and so the suspension linkages and control arms were made

in this way, and the 2005 model bell crank solid model file was imported as a graphic

file. The next step was to add joints between the bodies of interest, thus revolute

joints were used at the location of the suspension bearings to give the system the

necessary travel. Furthermore, both linear and torsional springs were added to the

model to act as the suspension and anti-roll bar springs, respectively. The UCSD

Formula SAE suspension design team specified the compression spring to have a

linear stiffness value of 701 N/cm (400 lb/in) and the anti-roll torsion bar to have a

12
13

stiffness of 4.91 N*m/deg (3.62 ft*lbs/deg) with an arm length of 6.35 cm (2.50 in), so

this data was programmed into the model accordingly. Figure 4 shows the MBD

model complete with the springs and linkages.

Figure 4: MBD model of rear suspension.

While the original intent of the MBD model was to solve for the dynamic loads

at the bell crank due to a time-dependent disturbance at the vehicle’s wheel, solver

capabilities and Formula SAE timeline constraints limited the analysis to only include

static forces. Therefore, damping and inertial forces were not accounted for in this

model, and only spring forces were captured. Obviously the components on the

vehicle are subjected to dynamic loads in actual use, so it was reasoned that a larger

13
14

factory of safety be designed into the parts to compensate for the unknown dynamic

forces.

2. Load Cases

Figure 5 shows the spring loads on the left bell crank due to a worst-case

scenario of the vehicle body rolling completely to the left such that the suspension

travel is at a maximum compression on the left side and maximum droop on the right.

In this case the linear spring is completely compressed and the anti-roll bar is at

maximum twist. Figure 6 depicts the load case in the event that the suspension is

loaded in the configuration that the vehicle is rolled over to the right. This would put

the linear spring in a relaxed position, with the anti-roll bar again at maximum twist,

but in the opposite direction. With the spring relaxed, there is no spring force at the

bell crank node C; however, the mass of the wheel system would be pulling on the

pushrod, and hence pulling on node C with roughly the weight of the vehicle’s

unsprung mass at one corner. As an engineering approximation, this force was

estimated to be 445 N (100 lbf) at a maximum. Thus the load case was updated to

include this force.

14
15

Figure 5: Bell crank load case in left vehicle roll condition.

Figure 6: Bell crank load case in right vehicle roll condition.

15
16

B. Design Space

Before the topology optimization study can begin, it is first necessary to

determine the maximum amount of volume that the geometry can safely occupy. This

is known as the design space, and it represents the volume that will be meshed into

finite elements and iterated upon while the optimization algorithm is working. If the

finalized component needs clearance or a pass-through for wires, for example, the

design space must reflect this so that the software does not try to use that space for

load-bearing elements. In the case of the UCSD Formula SAE bell crank, the design

space necessitated a notch for clearance to the spring, as well as cut-outs for the

attachment of the spring assembly and pushrods. This notch can be clearly seen in the

bell crank design space and spring/damper system of Figure 7. In addition, holes for

the bearing and bolt clearance are present.

Figure 7: Design space and spring/damper system.

16
17

III. Topology Optimization

A. Design Space Mesh

1. Geometry Clean-up

Part of the Altair Hyperworks® software package, Altair HyperMesh® was

used as the finite element meshing utility in preparation for the optimization study.

The design space previously created in a CAD program was imported as an IGES

(Initial Graphics Exchange Specification) file, and the geometry was “cleaned” to

prepare for meshing. This means that some of the lines in the imported model were

toggled from edge lines to suppressed (or manifold) lines so that they would not

represent an artificial edge that would force the finite elements to unnecessarily align

themselves to [13]. The misreading of lines happens at the locations of fillets and

radii features created in CAD models, as the features get falsely interpreted as distinct

surfaces in the IGES transformation. An example of the line types can be seen in

Figure 8, which shows the geometry cleaning phase for the design space meshing

process. Note: if viewed in color, the suppressed lines are depicted in blue, and the

edge lines are green.

2. Tetra-Meshing

Once the geometry was cleaned, the design space volume was filled with

tetrahedral elements using the auto-mesh features of HyperMesh®. This was done

with a volume-tetra element with a nominal minimum size of 2.54 mm (0.100 in), and

17
18

curvature and proximity adaptation enabled to refine the mesh in the regions of more

complex geometry. The resulting mesh that was used as the design space for the

topology optimization study can be seen in Figure 9.

Figure 8: Geometry cleaning in preparation for meshing.

Figure 9: Tetra-meshed design space.

18
19

B. Boundary Conditions

1. Treatment of Bolt Holes

As in any finite element analysis, proper boundary conditions (BC’s) are

crucial if the results are to be of any significance, and this is especially true for

topology optimization studies since these BC’s will be the basis for the resulting

distribution of material. An example of this lies in the treatment of the boundary

condition around the region of a bolted or pinned joint. Bolts and pins can only

transfer compressive load (unless, of course, they are bonded in place), thus they can

only push on another surface that is in contact. In a finite element model, however,

this phenomenon is somewhat difficult to capture as it requires the use of non-linear

gap elements which have a very low stiffness in tension to simulate the lack of

connectivity [7]. For the sake of time and simplicity, the bolt holes in this model were

filled with rigid “spiders” (RBE2 elements) at the acknowledged cost of reduced

accuracy of topological results in the vicinity of the bolt holes. These rigid spiders

connect all the perimeter nodes of the bolt hole to a single node in the center at which

the loads and constraints are applied.

2. Constraints

As installed on the vehicle, the rear bell crank pivots about a needle roller

bearing (node A of Figure 5) that is captured by a bolt through welded tabs to the

vehicle chassis. Furthermore, needle thrust bearings on both faces keep the bell crank

located with respect to the axis of the bearing bolt, and also serve to support any

19
20

moments out of the plane of the intended motion. Thus the only remaining degree of

freedom at the bearing is a rotation about the bolt, which in the local coordinate

system represents a rotation about the z-axis (refer to Figure 5 for coordinate system).

Because we have an idea of what the forces are at the spring nodes (nodes C

and D), it makes sense to apply known loads at these nodes, and constrain the pushrod

node (node B) in accordance with the actual set-up on the vehicle. The pushrod

transmits the suspension force from the lower control arm of the vehicle suspension to

the bell crank via spherical bearings (sometimes called ball-joints or rod-ends), thus

the node at which it connects to the bell crank is almost completely free to translate

and rotate. The pushrod is essentially a purely tension/compression member and so it

can only support translation along its axis. Because of this, the pushrod node at the

bell crank (node B) was allowed all of its degrees of freedom except translation in the

pushrod direction. Although this direction changes slightly as the suspension moves

through the range of motion, it is nominally at a position that corresponds to the y-

direction in the bell crank local coordinate system, and thus this degree of freedom

was constrained for the topology study. Defining DOFs 1, 2, and 3 as translational

degrees of freedom in x, y, and z; and DOFs 4, 5, and 6 as rotational degrees of

freedom about the x, y, and z axes; the constraints can be summarized as seen in Table

1. This set of constraints was denoted SBC and was used for all load cases. Refer to

Figure 5 for the bell crank node naming convention and coordinate system.

20
21

Table 1: Summary of bell crank constraints.

Bell crank
A B C D
Node
DOFs
12345 2 N/A N/A
constrained

3. Loads

Using the two previously determined load cases from the MBD model, we

have updated the finite element model to include these load vectors applied to the rigid

spiders of nodes C and D. Furthermore, two additional “out of plane” load cases were

created in which 445 N (100 lbf) transverse forces were applied to nodes C and D

perpendicular to the intended direction of travel. This was done to ensure that the

optimized geometry could accommodate realistic but unforeseen conditions and 445 N

(100 lbf) was chosen as an estimate for the worst-case condition. Table 2 shows a

summary of the four load cases used in the optimization of the 2006 UCSD Formula

SAE rear bell crank design. These values represent the forces seen at the left rear bell

crank during four distinct conditions. Note that there are no applied moments to the

system, and all forces are shown in units of Newtons with pound-force in parenthesis.

Load case 1 represents the vehicle in a worst-case right roll, load case 2 represents a

worst-case left roll, and load cases 3 and 4 are the assumed worst-case out of plane

loads. Figure 10 shows the aforementioned boundary conditions as applied to the

design space in preparation for the topology optimization study.

21
22

Table 2: Summary of bell crank load cases.

Load Const.
{A} {B} Cx Cy Cz Dx Dy Dz
Case Set
400N -187 311 -1748
1 0 0 0 0 SBC
(90 lbf) (-42) (70) (-393)
-3438 -921 -1259 1259
2 0 0 0 0 SBC
(-773) (-207) (-283) (283)
445 445
3 0 0 0 0 0 0 SBC
(100) (100)
-445 -445
4 0 0 0 0 0 0 SBC
(-100) (-100)

Figure 10: Applied boundary conditions.

22
23

C. Optimization Statement

1. Optimization Responses

The objective function of any optimization problem is to minimize or

maximize a certain response while meeting a prescribed set of constraints. For this it

is necessary to program the software to solve for the desired responses, and then

choose limits to these responses [14].

In the case of the bell crank optimization project, the objective function was to

minimize the mass while maintaining the integrity of the original design;

consequently, the response to minimize was chosen as the volume fraction. This value

represents the normalized volume of the design space after elements have been

iteratively eliminated or turned “off” in an effort to meet the design constraints, and

with a constant material density, the volume fraction directly correlates to the mass.

Obviously this mass minimization cannot continue without bound, thus it is necessary

to define responses that have upper and/or lower limits as constraints. For the design

of a highly stressed component it would be ideal to define the elemental stress level as

a response and limit the stress to a set factor of safety below the material yield stress.

Unfortunately, this response is not yet offered in the current version of HyperWorks®

7.0 (which was used for this research); however, it will be available in 8.0. In lieu of

this feature, the constraints applied in the optimization of the bell crank were

compliance driven as opposed to stress driven. Namely, three nodal displacements

were defined as responses, and the magnitude of these displacements was limited at a

level that compared with previous displacement results from a finite element analysis

23
24

of the 2005 model bell crank, the results of which can be seen in Figure 11. Because

the 2005 bell cranks went through an entire season of racing without any problems, it

was reasoned that the design was adequate enough to justify using the FEA

deformation results as a constraint for the new design. With this constraint set, the

optimized design would be at least as stiff as its predecessor.

Figure 11: 2005 bell crank finite element analysis displacement results.

24
25

2. Material Definition

The last input parameter required before implementation of the topology

optimization study is the material identification. In the case of the bell crank, the

material was specified as 7075-T6 aluminum; as a result, the material was defined as

linear isotropic (MAT1) and the values for Young’s modulus and Poisson’s ratio were

entered. These values used for the study were 71.7 GPa (1.04e7 psi) and 0.33,

respectively.

3. Manufacturing Constraints

Although the optimization problem is now fully defined, the resulting topology

design proposal will likely be very difficult to manufacture because the algorithm will

tend to make hollow structures with a lot of holes [15]. To ensure that the optimized

geometry can be realistically manufactured, the software includes algorithms for

implementing manufacturing constraints such as minimum size control and prescribed

draw directions. With minimum size control enabled, the optimization software will

not create geometry that is smaller than the desired size. This feature reduces the

number of small ribs and “blobs” that can complicate the interpretation and creation of

the optimized geometry. For the bell crank project, the minimum member size was set

to 2.54 mm (0.100 in), as this was the thinnest rib that the machinist felt comfortable

fabricating.

Because the bell crank was to be machined from solid aluminum, it was also

deemed beneficial to enable a draw direction, as if the part was going to be cast in a

25
26

mold. This eliminates the formation of a hollow structure, giving the part more of a

two-dimensional quality, thereby lending to its manufacturability. The bell crank

design space was given a single draw direction outward from the x-y plane in the z-

direction. For comparison purposes, however, optimization results were obtained with

and without the draw direction enabled, and the differences will be discussed in the

following chapter.

26
27

IV. Post Processing

A. Topology Optimization Results

The optimization analysis was conducted using Altair OptiStruct® 7.0 and the

jobs were run on a 3.40 GHz Pentium 4 processor with 512 Mb RAM. The average

run-time was about one hour. The resulting geometry proposal can be seen in Figure

12 with the previously stated boundary conditions, minimum member size control, and

symmetry constraint. Notice that there are clearly defined groupings of elements that

form ribs in an “x” pattern within the outer structure of the proposed bell crank design,

resulting in a rib geometry that is very different from that of the 2005 design. It is

especially interesting to see that the right hand portion of said rib group seems to curve

and meet the upper truss that forms the support for the anti-roll bar node. This was a

surprise, as intuitive thinking would say that typical truss members should be straight.

Figure 12: Topology optimization results, side view.

27
28

The isometric view of the proposal (Figure 13) reveals a hollow geometry that

would be very hard to machine from billet material. In particular, the design looks

like it would lend itself to fabrication from two separate plates spaced out from each

other. While this was an option for the design of the 2006 bell crank, the original goal

was to have the parts machined from billet aluminum; therefore, the study was run

again with the draw direction constraint enabled, and the results were much different.

Figure 13: Topology optimization results, isometric view.

28
29

Figure 14 shows the optimized material distribution when the same analysis

was run but also included a prescribed draw direction constraint. The algorithm

involved with this manufacturing constraint is complex and beyond the scope of this

Master’s Thesis, but suffice it to say that the constraint suppresses the formation of

cavities and undercuts [15]. The first obvious difference in this result is the lack of

two separate “plates” forming on each side, and instead the proposal is much more

solid looking—more closely resembling that of a cast and/or machined component.

The second important feature to notice is the reoccurring formation of an “x” shaped

brace in the center; however, this time the ribs are not separated into two distinct

planes. Instead of being spaced apart, the ribs have converged into one thin center

brace section.

Figure 14: Topology optimization results with draw constraint, isometric view.

29
30

An interesting result of both studies is the removal of material near the curved

notch region of the bell crank design space (see note of Figure 13). Because this cut-

out is in the path between the spring and anti-roll bar nodes, it was anticipated that all

of this material would remain in an effort to maximize the radius of the notch to

reduce the stress concentration factor in that region. Instead, almost all of the material

is removed, save for a small triangular brace for the anti-roll bar connection node.

It is important to note the lack of material around the bolt and bearing holes

due to the previously mentioned difficulties of accurately modeling the bolt

connection. Obviously, without any material surrounding the bolt-hole, the part could

not support a load placed at the bolt, thus one must pay attention to this fact in the

geometry interpretation phase of the topology study and design in the appropriate

amount of material around the connection holes.

B. Geometry Interpretation

Although the topology results appear reasonable, the design is definitely not

ready to hand over to the machine shop for fabrication. The results of the topology

studies are merely rough geometric proposals, and some interpretation is required to

create the final design. OptiStruct has the ability to export the topology results as an

IGES file using an export feature so that the geometry can be opened and perhaps

traced over in a CAD environment. For the interpretation of the 2006 UCSD bell

crank design, the optimized shape was created using SolidWorks® 2006 CAD

software. To capture the organic shape of the proposed design, splines were utilized

30
31

in the creation of the CAD model, and the interpretation can be seen in Figure 15. The

design retains the features proposed by OptiStruct: such as the curved “x” brace and

the truss-like anti-roll bar node support, and the final design is somewhat of a hybrid

inspired by both of the analyses. Appropriately sized fillets were utilized to reduce

stress concentrations and to aid manufacturability using standard tooling.

Figure 15: Optimized geometry interpretation, isometric view.

31
32

V. Design Validation

A. Finite Element Analysis

To validate the structural integrity of the newly designed bell crank, a finite

element analysis was performed using HyperMesh® 7.0 as a qualitative test to ensure

that the design did not have any inherent stress concentrations or fatal flaws. The

boundary conditions and mesh parameters were the same as used in the optimization

analysis, and the load case used was load case two of Table 2. This load case was

chosen as it represents the worst-case scenario of the linear suspension spring being at

maximum compression, and the anti-roll bar at maximum twist. Furthermore, this is

in the direction of loading that the bell crank will see due to impact loading, such as

the racecar driving over a pothole. Figure 16 depicts the von-Mises stress distribution

and 10-fold exaggerated displacement plot for the optimized 2006 design. Note that

the stresses are fairly homogenous within the structure. This is precisely the goal of

the topology optimization, and is the result of the algorithm which mimics the growth

behavior of biological load carriers like trees and bones, where the structures always

tend to grow into shapes that have homogeneous surface stress [15].

From the finite element results of Figure 16, there does not appear to be any

severe stress concentrations that would indicate a faulty design, and the highest stress

levels appear to be on the order of 140-210 MPa (20 to 30 ksi). With a material yield

stress of 500 MPa (73 ksi), these FEA results indicate a factor of safety of roughly 2.4

to 3.6.

32
33

Figure 16: von-Mises stress distribution and exaggerated displacement

B. Bell Crank Prototype Fabrication

With the initial finite element analysis validation complete and successful, it

was then necessary to build prototype bell crank components in preparation for the

next stage of the validation: physical testing. For this, MasterCAM X computer aided

manufacturing software was used to program the tool paths of a Hass VF-0 three axis

mill, and the machining required five mill operations and eight different tools. Using

7075-T6 alloy aluminum, four prototype components were fabricated. The tool paths

of the MasterCAM environment can be seen in Figure 17, as well as the fabrication

and finished product in Figures 18 and 19.

33
34

Figure 17: MasterCAM X tool paths

Figure 18: Bell crank prototype fabrication.

34
35

Figure 19: Bell crank prototype finished product.

C. Physical Testing

1. Hydraulic Test Fixture Design

As a means to further validate the design in a more practical sense, a hydraulic

load cell test frame capable of simulating the static suspension forces was designed

and assembled with the goal of testing the original and optimized components first to

the nominal load condition specified in load case two, and then to failure. Failure in

this case was defined as the yield point at which the bell crank would exhibit a

diminishing load carrying capacity. Before any testing could begin, however, the

35
36

experimental procedure first had to be established, and for this it was necessary to look

at the physical arrangement of the racecar suspension.

First of all, the anti-roll bar is a torsional type spring, and the force it exerts on

the anti-roll bar pushrod and thus bell crank is a function of its angle of twist, which is

a function of the vehicle suspension travel. Because the suspension has a finite

amount of travel (limited by the damper travel), the anti-roll bar force that can be

exerted to the bell crank is also finite, and this maximum level of anti-roll bar force is

precisely what is captured in load case two. Furthermore, aside from some small

structural value, the torsion bar spring does not have any associated damping

characteristics, thus dynamic impact will not drastically increase this force. The linear

suspension spring/damper assembly, however, is more complex. First, it does have a

viscous damper, thus dynamic travel will most definitely create a significant damping

force in addition to the static spring force. The spring/damper assembly also has a

finite amount of travel. In fact, it is what limits the travel of the vehicle suspension.

Under vehicle bottom-out conditions, for example, the damper shaft runs out of

compressive travel, and the bottom-out bump-rubber comes into contact with the

damper body, thus increasing the effective spring rate. If the bottom-out bump-rubber

should become completely compressed, the damper assembly would simply run out of

travel, thereby making the entire unit infinitely stiff (compared to the spring rate).

Because of this, the load that can be seen at the bell crank due to the linear spring

essentially has no bound, and will be responsible for the failure of the component in

actual use. Therefore it was decided for the experiment to first load the anti-roll bar

36
37

node to the worst-case load as depicted in load case two (1780 N), and then load the

linear spring node to the point of yield failure.

Hydraulic load cells with an acting bore diameter of 34.9 mm (1.375 in) were

used to supply the load to the connection pins associated with the linear spring as well

as the anti-roll bar attachment, as can be seen in the CAD model of Figure 20 below

[16]. In this manner hydraulic pressure could be directly correlated to force, so

pressure was recorded as a function of time for the physical tests. This was

accomplished by means of both analog pressure gauges as well as digital pressure

transducers, and the information was captured via videotape and LabVIEW data

acquisition, respectively. Hydraulic pressure was supplied with a hand pressure pump

capable of 10300 kPa (1500 psi).

Figure 20: Hydraulic test fixture CAD assembly.

37
38

In order to ensure that the experiments matched the conditions of both the

finite element model and the actual vehicle installation, care was taken in the

treatment of the test frame’s boundary conditions. For example, to mimic the

constraint at the pushrod attachment node (node B), a steel beam was used to represent

the pushrod, and it was loosely attached to the test frame base plate so that it could

rotate in a similar fashion as the actual vehicle component. In addition, the bell crank

bearing node (node A) was allowed to rotate only in the intended direction of motion,

and this was accomplished by capturing the bell crank with a bolt that was welded to

the test frame base plate. Furthermore, the hydraulic load cells were attached to the

bell crank using the same type of hardware that would be used to attach the

spring/damper assembly and anti-roll bar pushrod.

For the experiment to be a success, the bell crank components needed to be

tested to failure, thus the hydraulic cylinders had to be capable of supplying enough

force. With a nominal bore diameter of 34.9 mm (1.375 in), the cylinders have a

piston area of 958 mm2 (1.485 in2). With a maximum pressure rating of 10300 kPa

(1500 psi), the cylinders could each exert a maximum force of 9.92 kN (2230 lbf).

While this was obviously sufficient to supply the 1780 N (400 lbf) needed at the anti-

roll bar node, it was unknown how much force was needed at the spring node to fail

the part; therefore, two hydraulic cylinders were installed in parallel for this purpose,

doubling the maximum force available. A ball-valve was installed on the lines so that

the cylinder associated with the anti-roll bar node could first be pumped up to the test

pressure, and then closed while the other two cylinders were increasingly loaded.

38
39

Granted, it was realized that the pressure in the first cylinder would fall as the bell

crank deformed, but his was deemed negligible based on the very small deformation

seen in the finite element model. Furthermore, it was hypothesized that the small

force from this cylinder would be relatively insignificant as compared to the final

force at failure required from the tandem cylinders.

In preparation for the experiment, the test fixture was covered with a

polycarbonate safety shield, and the reservoir was filled with hydraulic fluid. Figure

21 shows the device loaded with a 2006 model bell crank ready to test.

Figure 21: Hydraulic test fixture prepared for experiment.

39
40

2. Experimental Results

To accurately compare the 2006 optimized bell crank design to the 2005

model, one of each component type was tested to static failure using the hydraulic test

fixture described in the previous section. Furthermore, both the 2005 and 2006 test

specimens were machined by the same person using the same equipment and alloy of

aluminum.

Weighing in at 185 grams, the 2005 model bell crank required a hydraulic

pressure of 6380 kPa (925 psi) before deforming plastically. Given the effective bore

area of the tandem cylinders, this pressure equates to a load of 12.2 kN (2750 lbf).

The final mass of the 2006 optimized rear bell crank design as tested was measured at

140 grams—a weight savings of 24.3%; moreover, the optimized design supported a

hydraulic pressure of 8270 kPa (1200 psi). This critical pressure represents a load of

15.9 kN (3560 lbf) and thus a 29.7 % increase in strength as compared to the 2005

model. Table 3 summarizes these results, and Figure 22 shows the 2006 design bell

crank after testing. Note the similarities in the deformation compared to the finite

element prediction of Figure 16.

40
41

Table 3: Summary of experimental test results.

Mass Yield Load

2005 model 185 grams 12.2 kN (2750 lbf)

2006 model 140 grams 15.9 kN (3560 lbf)

Percent change -24.3% +29.7%

Figure 21: Physical testing of 2006 optimized bell crank.

41
42

VI. Conclusion

A. Discussion of Experimental Results

Due to the very lightweight look and feel of the topology optimized bell crank,

many observers were initially dubious as to the structural integrity of the prototype

component, and justifiably so, as the parts did look fairly slim—especially when

compared to the already slender 2005 component. This meant that a lot was riding on

the physical testing of the bell crank—both literally and figuratively. Especially

considering the fact that during the time of physical testing, the 2006 Formula SAE

West competition was about a week away, and thus the results of the physical testing

would determine whether the final components found their way onto the 2006 UCSD

competition vehicle entry. Furthermore, the team’s overall confidence in topology

optimization also hinged upon the result of the physical tests. Needless to say, seeing

a nearly 30% increase in strength coupled with an equally impressive decrease in mass

surely had many people excited about the computational powers of the topology

optimization algorithm. And the parts made it on the car in time for competition—

anodized to match the wheels.

It is understood that these numbers might be on the high side of the weight loss

spectrum for similar studies, and this can most likely be attributed to the fact that the

2005 model to which the study is compared was not a very developed part. This is to

be expected for a student-run vehicle design team, where development time is at a

premium. Even so, some large companies are claiming similar weight savings

42
43

numbers. Volkswagon Design Development claimed a 23.0% weight reduction in the

optimized design of an engine bracket in a presentation given at the Optimization

Technology Conference (OTC) in Troy, Michigan in 2005 [6], and it seems that values

in the range of 10% -30% are fairly common amongst other users that have shared

their results.

Aside from the newfound confidence in topology optimization, another result

of the testing was the discovery of a relatively high measured factor of safety in the

tested bell crank component. While the lack of a complete and thorough

understanding of the dynamic loading conditions at the part necessitated an overly

designed part, the loads required at failure were higher than anticipated. Knowing this

information, some strength could have been sacrificed for further decreased mass;

namely, the bell cranks could have been crafted from magnesium as opposed to high

strength aluminum, bringing an additional 30% weight savings.

While the results of the study as presented in this thesis are indeed optimistic,

they merely represent the first attempt in applying topology optimization schemes in

vehicle design at UCSD. Hopefully future students will further exploit the powers of

the topology optimization algorithm through the use of increasingly developed and

demanding applications.

43
44

B. Overall Vehicle Performance and Awards

The 2006 UCSD Formula SAE vehicle represented the third successive vehicle

designed and developed by engineering students at the University of California, San

Diego, and it proved to be an entry of high caliber. Unfortunately, a fuel pump failure

during the final laps of the endurance/fuel economy event prevented the team from

placing well overall; however, many of the goals set forth by the team were met. For

example, the final wet weight of the 2006 entry was 223 kg (492 lbs), meeting the goal

of being under 227 kg (500 lbs), which was no small feat considering that the vehicle

was equipped with a supercharger (a feature which added roughly 7 kg (15 lbs) to the

vehicle package). Furthermore, a goal for the 75 meter acceleration event was set at

4.350 seconds, which would represent a drastic improvement from the team’s best of

4.634 seconds with the 2005 car [17]. Beyond anyone’s expectations, the 2006

vehicle shattered the goal by putting down an astonishing time of 4.137 seconds—a

time that not only exceeded our goals, but was good enough for 3rd place in the entire

field [18]. In fact, the time is one of the top ten fastest in Formula SAE competition

over the past five years—representing a field of over 700 vehicles! Table 4 shows a

performance comparison between the 2005 and 2006 UC San Diego Formula SAE

entries, and Figure 23 shows the 2006 vehicle in final form at the Formula SAE West

competition.

In addition to the satisfaction of meeting our own goals, the 2006 Formula

SAE team at UCSD was twice invited up to the stage to accept awards involving the

vehicle performance and design. For the acceleration event, the team was awarded the

44
45

3rd place prize for the Honda Acceleration Award. The team also won 1st prize in the

Altair Engineering’s William R. Adam Engineering Award, for the “top team that

exhibits innovative engineering concepts in vehicle design” [19], and this award was

given for the team’s effective use of topology optimization software in the design of

the rear suspension bell cranks. Figure 24 shows the anodized bell cranks as installed

on the 2006 vehicle at competition, and the racecar in motion can be seen in Figure 25.

Table 4: Comparison of 2005 and 2006 UCSD FSAE vehicle performance

Mass 75m Acceleration Peak Torque

2005 Vehicle 246 kg (543 lbs) 4.634 s 58 N*m (43 ft*lbs)

2006 Vehicle 223 kg (492 lbs) 4.137 s 71 N*m (52 ft*lbs)

Percent Change -9.4% -10.7% +21%

Figure 22: 2006 UCSD FSAE vehicle at competition.

45
46

Figure 23: Close-up of bell crank on 2006 vehicle.

Figure 24: 2006 vehicle in motion.

46
47

Bibliography
[1] SAE International. “Collegiate Design Series: Formula SAE Series.” 2006.
Online posting. <http://students.sae.org/competitions/formulaseries/>

[2] Fornace, L.V., Davis, A.E., Costabile, J.T., Hart, J.D., Arnold, D., “Traction
Control Systems: FSAE Vehicle Acceleration Optimization.” MAE 171B Final
Report. UC San Diego, 2005.

[3] Taylor, Rob. “F-35 Joint Strike Fighter Structural Component Optimization:
Lockheed Martin Aeronautics Company.” Optimization Technology
Conference. 27-28 September 2005. Troy, Michigan, USA.

[4] Talke, F.E., “Optimization: Computer Aided Analysis and Design.” Class
Notes. MAE 292. UC San Diego, Spring 2006.

[5] Schneider, Detlef, and Erney, Thomas. “Combination of Topology and


Topography Optimization for Sheet Metal Structures.” OptiCON 2000
Conference Proceedings.

[6] Meyer-Pruessner, Rainer. “Significant Weight Reduction by Using Topology


Optimization in Volkswagen Design Development.” Optimization Technology
Conference. 27-28 September 2005. Troy, Michigan, USA.

[7] Altair Engineering. “OptiStruct 7.0 User’s Guide.” HyperWorks 2004.

[8] Hughes, Thomas J.R., “The Finite Element Method, Linear Static and Dynamic
Finite Element Analysis.” New York: Dover Publications, 1987.

[9] “Webster’s New World Dictionary.” New York: The World Publishing
Company, 1967.

[10] Wikipedia. “The Free Encyclopedia: Bell Crank Definition.” 2006.


Online posting. <http://en.wikipedia.org/wiki/Bell_crank>

[11] Wright, P., “Formula 1 Technology.” Warrendale: SAE International, 2001.

[12] Altair Engineering. “Altair MotionView: Pre-and Post Processing for Multi-
Body Dynamics, Volume I.” HyperWorks Training Manual, 2004.

[13] Altair Engineering. “Altair HyperMesh: Introduction to FEA: Pre-Processing


Volume I.” HyperWorks Training Manual, 2004.

47
48

[14] Altair Engineering. “Altair OptiStruct: Concept Design Using Topology and
Topography Optimization.” HyperWorks Training Manual, 2004.

[15] Harzheim, Lothar, and Graf, Gerhard. “TopShape: An Attempt to Create


Design Proposals Including Manufacturing Constraints.” OptiCON 2000
Conference Proceedings.

[16] Ariely, A., Doring, P., Erkebaev, T., and Tawatao, M.L., “Suspension
Component Failure Simulation.” MAE 171B Final Report. UC San Diego,
2006.

[17] SAE International. “Collegiate Design Series: Formula SAE Results.” 2005.
Online post. <http://students.sae.org/competitions/formulaseries/fsae/>

[18] SAE International. “Collegiate Design Series: Formula SAE West Results.”
2005. Online post <http://students.sae.org/competitions/formulaseries/west/>

[19] SAE International. “2006 Award Winners: Formula SAE West Results.”
2006. Online post. <http://students.sae.org/competitions/formulaseries/west/>

[20] Milliken, William F., and Milliken, Douglas L. “Race Car Vehicle Dynamics.”
Warrendale: SAE International, 1995.

48

Potrebbero piacerti anche