You are on page 1of 330

SAWE Paper No.

3602
Category Number 31.0
Rev D

MASS PROPERTIES AND ADVANCED AUTOMOTIVE DESIGN

By

Brian Paul Wiegand, P.E.

(Northrop Grumman Corporation, Retired)

For Presentation at the


th
74 SAWE International Conference
on Mass Properties Engineering
18-22 May 2015
Crown Plaza Hotel
Alexandria, Virginia, USA

Permission to publish this paper, in full or in part, with


credit to the Author and to the Society, may be obtained by request to:

Society of Allied Weight Engineers, Inc.


P.O. Box 60024, Terminal Annex
Los Angeles, CA 90060

The Society is not responsible for statements or opinions in


papers or discussions at its meetings. This paper meets all regulations for public
information disclosure under ITAR and EAR
SAWE Paper No. 3602
Category Number 31.0
Rev C

TABLE OF CONTENTS

Chapter: Page:

Table of Contents ........................................................................................................................ i


Abstract ...................................................................................................................................... ii
1 - Introduction: Mass Properties and Advanced Design ........................................................ 1
2 - Kinetic Energy and Effective Mass ..................................................................................... 7
3 - Energy Generation and Mass Distribution ...................................................................... 28
4 - An Advanced Mass Properties Design Concept .............................................................. 43
5 - Longitudinal Performance: Acceleration/Braking ............................................................. 77
6 - Lateral Performance: Steady State/Transient ................................................................... 100
7 - Vertical Performance: Shock/Road Contact/Vibration/Pitch/Bounce ............................ 125
8 - Gyroscopic Torque Reactions ......................................................................................... 154
9 - Fuel Economy .................................................................................................................. 161
10 - Safety ............................................................................................................................. 167
11 - Producibility & Marketability ........................................................................................ 171
12 - Conclusions .................................................................................................................... 178
References .............................................................................................................................. 181
Author’s Biographical Sketch ................................................................................................ 189
Appendices ............................................................................................................................. 190
A - Symbolism…………………………………………………………………191
B - Weight Accounting .. ................................................................................. 218
C - MAXDLONG.BAS Program Listing ........................................................ 224
D - Automotive Configurations: Wheel/Axle ................................................... 233
E - Reciprocating Piston Engine Inertial Energy Loss .................................... 245
F - Notes on Tire Behavior .. ............................................................................. 255
G - Transient State: Center of Oscillation .. .. …………………………………319
H – Road to Unsprung Mass Transmissibility .. .............................................. 322

i
SAWE Paper No. 3602
Category Number 31.0

ABSTRACT
The history of the automobile constitutes a long list of remarkable vehicles, each of which
is usually associated with some equally remarkable individual who provided the impetus behind
the vehicle’s creation. Henry Royce, Hans Ledwinka, Ferdinand Porsche, Enzo Ferrari, and Colin
Chapman constitute just a small sample of those who put the forceful stamp of their personal
design philosophies upon the automobile.
The design philosophies of such men reflected a focused viewpoint, such as an emphasis
on quality and refinement (Royce), efficient structure and aerodynamics (Ledwinka, Porsche),
exquisite engines of high specific power (Ferrari), or performance from extreme light weight
(Chapman). The emphasis on a particular aspect of vehicle design hardly means that other design
aspects were ignored, just that the emphasis was on some particular aspect(s) from which all else
evolved, whether by preference or necessity. Colin Chapman’s relentless concern with mass
properties, and in particular with light weight, was born of necessity, but would become a
personal obsession.
As a small start-up company competing against long established automotive concerns such
as Ferrari, Chapman’s Lotus Engineering Company 1 did not have the capability to gain advantage
through advanced engine design, or even the design of most of the other major mechanical
systems. Most Lotus components were commercially sourced, and so the only way a decisive
advantage could be obtained was through an uncompromising emphasis on gaining performance
“edges” from the remaining design aspects of structure, body, and suspension. Because the
automotive performance aspects of acceleration, braking, and handling are so dependent on
various vehicle mass properties the optimization of those mass properties became the “Holy
Grail” of Lotus design as directed by Colin Chapman.
In fact, with regard to light weight Chapman would come to demonstrate what could be
termed a mania. Although the automotive design philosophy of Colin Chapman was the closest
ever to being one that could be termed “designing from the viewpoint of the mass properties
engineer”, the key word is “closest”. There is much more to mass properties engineering than
simply minimizing weight, which Chapman often did with a disregard for reliability and even
safety. It was despite an unenviable record of DNF’s and driver casualties that Lotus still obtained
an outstanding record in automotive competition, and Formula One in particular.
This raises the fascinating question of what would result from applying an automotive
design philosophy based on a balanced professional mass properties viewpoint. The intent of this
paper is to show that a vehicle designed in true accordance with the balanced viewpoint of a
professional mass properties engineer may not only demonstrate superior acceleration, braking,
and handling, but superior ride, stability, fuel economy, and safety as well. If a design begins with
the first principles of how mass properties affect automotive performance in all its aspects 2, and
is optimized accordingly in an integrated manner, then the resulting advanced automotive design
may truly “go where none have gone before”.

1
“Lotus Engineering Company” was the official name at the company’s founding on 1 January 1952; since then
the company split into a number of divisions united in nomenclature only by the word “Lotus”.
2
Just how mass properties dominate all the major aspects of automotive performance has been the subject of a
series of previous papers by this author; see References [88] through [92].

ii
SAWE Paper No. 3602
Category Number 31.0

1 – INTRODUCTION: MASS PROPERTIES AND ADVANCED DESIGN

Historically there have been quite a few automotive designers/design managers who
placed an emphasis on mass properties, and in particular light weight, in their design
philosophies. Men like Ettore Bugatti, Ferdinand Porsche, Hans Ledwinka, Alex Issigonis,
and Dante Giacosa come to mind 3. However, Colin Chapman’s concern with achieving the
lightest weight possible for his race vehicles and road cars approached a mania, perhaps
equaled in intensity only by Pierre-Jules Boulanger, the design manager for the Citroen 2CV 4:

“Saving weight, of course, is part of every racing car designer’s


philosophy, but with Colin Chapman it amounted to a
monomania…”.

Anthony Curtis, 1970 5

However, Chapman’s concern went far beyond just achieving a minimum weight; he
relentlessly sought to optimize all relevant mass properties parameters driving the
automotive performance aspects of acceleration, braking, road holding, and maneuver. Other
than minimizing the total vehicle weight, he sought to minimize unsprung weight, and in
particular to minimize the weight and rotational inertia of the rolling unsprung
components. He also consistently sought to minimize the vehicle vertical center of gravity,
and he sought to modify other mass properties in accord with his design philosophy at the
time.
Colin Chapman’s early (c.1947-1962) automotive design philosophy called for a
forward bias in the vehicle longitudinal center of gravity and high vehicle yaw inertia; this
was in accord with his belief that directional stability required a grossly understeering vehicle.
His later (c.1963-1982) philosophy constituted a complete reversal; it called for a rearward
bias in the longitudinal center of gravity and low vehicle yaw inertia. This transformation
of design philosophy more or less paralleled the transition from front-engine to mid/rear-
engine race car design, so the changes in philosophy and vehicle configuration were
complementary and interrelated.

3
Ettore Bugatti (1881-1947) is famous for the fast sports and “Grand Prix” cars bearing his name. Ferdinand
Porsche (1875-1951) is best known for his design of the Volkswagen “Beetle” and for the sports cars bearing his
name. Hans Ledwinka (1878-1967) designed the fabulous T77, T87, and T97 models for Tatra. These vehicles
won him the admiration of Adolf Hitler and Ferdinand Porsche, among many others. Alec Issigonis (1906-1988)
designed the famous Austin/Morris Mini, as well as the popular Morris Minor and Austin 1100. Dante Giacosa
(1905-1996) was largely responsible for Fiat’s iconic “Topolino” as well as their FWD design standard.
4
Pierre-Jules Boulanger (1885-1950) was not only instrumental in the development of the “Très Petite Voiture”
that came to be Citroen’s classic 2CV, but in greatly hindering the Nazi war effort during the occupation of
France. He barely survived the Nazi occupation only to die in a car accident in 1950.
5
Reference [46], pg. 201. Anthony Curtis (1956 - ) is a well known writer on sporting subjects.

1
SAWE Paper No. 3602
Category Number 31.0

There was another mass properties parameter that Colin Chapman would consistently
seek to control, and that was the lateral center of gravity. Generally this meant keeping the
lateral center of gravity on the vehicle longitudinal center-line (except in the case of “circle”
track vehicles that turn only in one direction, such as Indy cars, where an offset CG could be
used to good advantage).
Despite the fact that Colin Chapman was perhaps the most sophisticated practitioner
of the science of mass properties analysis and control 6 in the automotive field, he did not
possess the balanced approach of a professional mass properties engineer. His ruthlessness in
pursuing superior automotive performance via mass properties optimization yielded results,
but at a cost that would be unacceptable today. He regularly overrode the best judgment of his
stress analysts and pared weight down to the point that the strength necessary to function
safely was compromised. Designing as light as possible, then waiting for things to fail in use
(there was virtually no budget for pre-production testing) so that the failed pieces would then
be appropriately strengthened, was the traditional Lotus (Chapman) way of obtaining an
optimum structure. 7
It might seem churlish to argue with success, and Lotus as directed by Colin Chapman
was very successful, but one is left to wonder just how much more Lotus could have
accomplished without the multitude of DNF’s and accidents that undoubtedly were the result
of such “empirical optimization”:

“I was becoming unhappy about the reliability of the car.


Vital bits kept falling off, like the steering wheel! He cut a lot
of corners, did Colin, and made things lighter than they
probably should have been.”
Cliff Allison,
Lotus Formula One driver, 1958 8.
“In one car at Monza… the mechanics found fourteen major
breaks in the chassis. I had breaks in the steering, wheels fell
off, wishbones breaking, anything that could happen to a
car.”

Innes Ireland,
Lotus Formula One driver, 1959 9.

6
Colin Chapman (1928-1982) utilized conceptual weight estimation, the setting of design target weights,
weight trade-off studies, weight reduction (often through material substitution), establishment of the value
of a pound, and weight control during development to a degree that was uncommon in the automotive world at
the time. He was also very astute regarding the management of fuel burn and gyroscopic reactions.
7
Reference [46], pp. 194, 383.
8
Ibid, pp. 195-196.

2
SAWE Paper No. 3602
Category Number 31.0

“Did I think the Lotus way of doing things was good? No. We
had several structural failures in those cars. But at the time I
felt it was the price you paid for getting something
significantly better.”
Dan Gurney 10
The most grievous cost of Chapman’s ruthlessness would have to be the injury and
death that resulted. The list of the killed and injured would include Stirling Moss (twice badly
injured, 1960 and 1962), Alan Stacey (d. 1960), Ricardo Rodríguez (d. 1962), Gary Hocking
(d. 1962), Mike Taylor (severely injured in 1960 when Lotus 18 steering column broke in
two, he sued Lotus for damages and won), Jim Clark (d. 1968), Mike Spence (d. 1968),
Bobby Marshman (d. 1964), Graham Hill (injured in 1969, but died piloting his private plane
in 1975), Jochen Rindt (d. 1970), and Ronnie Peterson (d. 1978). And this is only a partial list
of Lotus racing casualties, and includes none of that unknown number of victims of Lotus
road cars:

“Engines leaked oil prodigiously. Frames cracked apart in


everyday use. Due to the car’s dense packaging and tiny
footprint…service access was nightmarish. Early Elans
flooded when it rained and overheated in traffic, the Webers
leaked fuel directly onto the ignition system, and electrical
problems were legion.”

Sam Smith, 201211

“Lotuses are all rubbish. They always fall apart.”


Tony Vandervell, c. 1960 12

“We allowed the electrics to be developed on the line by the


electrical fitters that were making the harnesses, fitting the
lights. We never actually had a competent electrical
engineer…look at all the various loadings on the circuitry and
design it…We never had what I would call the correct
facilities for doing our own inspection.”
Mike Warner, Lotus Q/C, c. 1960’s 13

9
Reference [46], pg. 196.
10
Reference [25], pg. N/A (Internet).
11
Reference [78], pg. 71. This doesn’t even mention the additional problems of chassis rust and fiberglass body
cracking.
12
Reference [46], pg. 206. Guy Anthony "Tony" Vandervell (1898–1967) was an English industrialist and
founder of the Vanwall Formula One racing team.

3
SAWE Paper No. 3602
Category Number 31.0

Chapman’s pursuit of optimum mass properties led to much innovation in the


construction (space frame, monocoque), materials (magnesium wheels, honeycomb panels,
etc.), vehicle configuration (mid-engine), and so forth; but he was an innovator in other areas
as well, especially in the realms of suspension design and aerodynamics. It was perhaps the
value of his innovations that kept the negative aspects of his over-zealous weight reduction
from being too detrimental to the overall effort, allowing for his over-all success.

Colin Chapman once attributed his success as an innovator to a claimed ability to step
back from the present state of things and to return to first principles, in essence starting all
over again, and he gave credit for this ability to…

“…a professor at London University, Vaughan…who was


very fond of putting the theory forward. Far too many people
are prepared to take what has already been established and
improve on it or make small changes. Whereas I much
prefer to go right back to square one and try to assess what
the basic reason is for doing anything and find the most
basic, straightforward, and simple way of doing it.”
Colin Chapman 14

However, nothing could have been further from the truth, at least if the phrase
“returning to first principles” is taken to mean that he started from such basic relationships as
Newton’s Second Law of Motion, and from there pursued the theoretical development onward
to the much more complicated relationships which fully characterize automotive behavior.
Campbell was not a theoretician; he was not well prepared for such a role academically, nor
did he have the temperament or time for such activity. His whirlwind mode of operation was
to develop a general acquaintance with theory as advanced by the greatest authorities 15, then
utilize that limited understanding to recognize the potential in concepts inspired by numerous
sources, but to leave the detail development of those concepts to others who did possess the
necessary technical knowledge. That his actual starting point for developing a concept had
little to do with “first principles” has been fully testified to by those who worked with him:

13
Reference [46], pg. 19.
14
Ibid, pg. 23.
15
For instance, he would refer to the works of Sir Harry Ricardo on engines, Maurice Olley and William
Milliken on handling, G.E. Lind-Walker on directional stability, David Hodkin on steering, Robert Eberan von
Eberhorst on roll angles, and so forth.

4
SAWE Paper No. 3602
Category Number 31.0

“Colin just needed a starting point. He needed to see the efforts of


others before he knew just what he wanted. He would look over
the shoulder of a designer and make incredibly apposite
suggestions, or come up with elegant solutions to some knotty
problem aired in conference.”
Graham Arnold 16
“Colin was extremely sharp, clever and quick-witted, and
very charismatic, but not a really great engineer or designer.
He had a few principles which he pursued to great effect, such
as light weight…but he never had a really close grasp of
fundamental principles. Perhaps he might have had if he had
more time to learn. But in his hectic and unscrupulous life it
was quicker to rely on other people to whom he gave
minimum credit.”
Charles Bulmer 17

By truly undertaking a chain of reasoning beginning from first principles it is possible


to arrive at a result that Chapman would not have. And, if the direction of that chain of
reasoning were to be guided in accord with the sensibilities of a professional Mass Properties
Analysis and Control engineer, then the result might well turn out to be one that no one has
ever previously arrived at. Such is the case for the “advanced automotive design concept”
presented within this paper, which is developed from the most elementary mass properties
considerations to their logical conclusion.
However, although no one may have ever previously arrived at the “advanced
automotive design concept” exactly as it is presented in this paper, there have been a number
of “historical vectors” that have pointed towards it 18. Many of those “historical vectors” are
identified in this paper in the course of introducing the “advanced concept”. This is done in
order to lend the “advanced concept” such validation as can be obtained without an actual
empirical testing of the concept itself. To the greatest extent possible within the limited
resources of the author, the validity of the “advanced concept” is further buttressed by all the
theoretical analysis that can be brought to bear.
The “historical vectors” in question are the Reeves Octoauto (1911, road vehicle), the
Advanced Vehicle Systems Shadow (1970, Can Am racing), the Elf Tyrrell P34 (1975, F1
racing), the March 2-4-0 (1976, F1 racing), the Williams FW08D (1982, F1 racing), and the
Keio University Eliica (2004, road vehicle prototype). Each of these vehicles constitutes to

16
Reference [46], pg. 31. Graham Arnold was Lotus Cars Director of Sales from 1963 to 1971.
17
Ibid, pg. 385; Bulmer (1922-2012) was a great English automotive journalist, editor of The Motor.
18
The great German philosopher Georg Wilhelm Friedrich Hegel (1770-1831) would have appreciated this as
part of a “historical dialectic”, an uneven process through time toward some ideal.

5
SAWE Paper No. 3602
Category Number 31.0

varying degree an incomplete manifestation of the “advanced automotive design concept”.


They all showed to some degree the promise inherent in the “advanced concept”, but did not
enjoy great success or longevity due to the incompleteness of the development and/or
extraneous circumstances; none were developed with the assistance of a professional Mass
Properties Analysis and Control engineer. This paper will, through its presentation of the
case for the “advanced automotive design concept”, also serve to vindicate such historical
harbingers.

6
SAWE Paper No. 3602
Category Number 31.0

2 – KINETIC ENERGY AND EFFECTIVE MASS

“He who considers things in their growth and origin…will obtain


the clearest view of them”.
Aristotle (384-322 BC) 19

A true first principle starting point in automotive design could be the relationship
between a simple homogeneous point mass (“m”) traveling at some velocity (“V”) and the
kinetic energy (“KE”) required 20:

𝑲𝑬 = ½ 𝒎 𝑽𝟐 (EQ. 2.1)

However, an automobile is not a simple homogeneous body; it is a complex assembly


of variegated components, some of which are not just accelerated translationally, but
rotationally as well 21. Thus "m" is not just a matter of vehicle weight divided by the
gravitational constant (“m = W/g”) as would be the case with a simple homogeneous body.
To account for the overall effects of this variegated behavior the concept of "effective mass"
must be substituted for the fundamental mass "m"; to indicate this distinction requires the
employment of a new symbol “me” for the effective mass of an automobile:

𝑲𝑬 = ½ 𝒎𝒆 𝑽𝟐 (EQ. 2.2)

For most conventional automotive configurations the value of “me” may be


determined as per the following equation 22:

19
This has also been translated as “If you would understand anything, observe its beginning and its
development”.
20
The credit for this formulation is largely due to the feminine genius of Émilie du Châtelet (1706-1749), who
supplanted Newton’s “KE = m V” with “KE = m V2” based on her study of experimental data obtained by the
Dutch experimenter W.J. Gravesande (1688-1742). The expression becomes the familiar “KE = ½ m V2” when
subject to a modern derivation.
21
Reference [79], pg. 94-96.
22
Reference [89], pg. 7.

7
SAWE Paper No. 3602
Category Number 31.0

𝑾𝒕 𝑰𝟏+𝑰𝟐 (𝑻𝑹×𝑨𝑹)𝟐 +𝑰𝟑 (𝑨𝑹)𝟐 +𝑰𝟒


𝐦𝐞 = + (EQ. 2.3)
𝒈 𝑹𝑫𝟐 ×𝒈

The symbolism in this equation is as follows:

Wt = Weight of the vehicle (lb, kg).


g = Gravitational constant, which is required only when the
weight is expressed in force units and must be converted to
mass units. For example, when the units employed are in the
English feet, pounds, and seconds (FPS) system then “g” =
32.174 ft/s2.
I1 = Rotational inertia about front axle line (lb-ft2, kg-m2).
I2 = Rotational inertia about the crankshaft axis (lb-ft2, kg-
m2).
I3 = Rotational inertia about transmission 3rd motion axis,
plus 1st and 2nd transmission motion axis inertias translated
to the 3rd motion axis (lb-ft2, kg-m2).
I4 = Rotational inertia about rear axle line (lb-ft2, kg-m2).
TR = Transmission gear ratio (dimensionless).
AR = Axle gear ratio (dimensionless).
RD = Dynamic rolling radius at drive wheels (ft, m).

The equation for automotive effective mass may also include the gear set efficiencies,
transmission and differential, although the effect is small 23, and the exact manner of inclusion
depends upon the application 24. In the reference from which Equation 2.3 was drawn the
application was automotive longitudinal acceleration; the gear set efficiencies were accounted
for by inclusion in the equation for the accelerative force 25:

23
In some references the gear set efficiencies are neglected altogether.
24
See “Appendix B” of the published paper: B. Wiegand, SAWE 3634, “Mass Properties and Automotive Crash
Survival”, May 2015.
25
Reference [89], pg. 4.

8
SAWE Paper No. 3602
Category Number 31.0

𝑻×𝑻𝑹×𝑻𝑬×𝑨𝑹×𝑨𝑬
𝑭= − 𝑭𝑹 − 𝑫 (EQ. 2.4)
𝑹𝑫

The symbolism unique to this equation is as follows:

F = Accelerative force or thrust (lb, kg).


T = Engine torque (lb-ft, kg-m).
TE = Transmission efficiency (dimensionless).
AE = Axle efficiency (dimensionless).
FR = Tire rolling resistance force (lb, kg).
D = Aerodynamic resistance force (lb, kg).

When the concern is not with longitudinal acceleration, but with longitudinal
deceleration (braking), the effect of the gear set efficiencies may best be accounted for in the
effective mass equation like so 26:

𝑾𝒕 𝑰𝟏+𝑰𝟐 (𝑻𝑹×𝑨𝑹)𝟐 (𝑻𝑬×𝑨𝑬)+𝑰𝟑 (𝑨𝑹)𝟐 (𝑨𝑬)+𝑰𝟒


𝐦𝐞 = + (EQ. 2.5)
𝒈 𝑹𝑫𝟐 ×𝒈

Note that transferring the efficiencies from the force side of the “F = me a” equation to
the “me” side was not accomplished as a simple mathematical operation which would have
resulted in an increase in magnitude for “me”. Since the concern is to determine the amount of
kinetic energy requiring dissipation in the course of vehicle deceleration the value of “me” has
to decrease, not increase, when the effect of the gear set efficiencies is considered.
Also note that, for any calculation involving the kinetic energy to be dissipated during
hard braking, the “I2” term is generally dropped. This is because usually any energy stored
and/or generated “upstream” of the clutch or torque converter is cut off from the mainstream
of power flow under such conditions.

26
For acceleration the gear set efficiencies must serve to increase the amount of work energy required, and for
deceleration the gear set efficiencies must serve to reduce the amount of kinetic energy requiring dissipation.

9
SAWE Paper No. 3602
Category Number 31.0

A concrete appreciation of how the individual terms in the effective mass equation
contribute to the total effective mass can be obtained through study of an example set of
parameter values representative of an actual passenger car. For this purpose a 1958 Jaguar
XK150S was taken to be representative of conventional configuration 27 passenger cars (a
choice mandated by the fact that this is the only vehicle for which a full parameter value set is
readily available to this author 28):

Table 2.1 – REPRESENTATIVE PASSENGER CAR PARAMETER SET


Utilizing the data from Table 2.1, the effective mass (expressed as weight) in each
gear was calculated 29 with the absolute and relative magnitudes of each contribution noted.
As shown in the resulting Table 2.2, for the example test configuration considered the Jaguar
weighed 3640 lb (1651 kg), but in first gear it accelerated as if it had a total effective weight
(mass) of 5106 lb (2316 kg). The “I1” and “I4” terms, which essentially represent the wheels
/tires/brakes, contribute about 360.8 lb (163.7 kg) or 7.07% of that total. This is dwarfed by
the “I2” (essentially the engine’s rotating parts) contribution of about 1454.1 lb (659.6 kg) or
28.48%. The “I3” (essentially transmission/drivetrain parts) contribution is about 78.5 lb
(35.6 kg) or 1.54%, with the final remaining 62.92% of the effective total contributed by the
3212.6 lb (1457.2 kg) of purely translational (non-rotating) components.

27
“Conventional configuration” is herein intended to mean a traditional front engine, rear wheel drive
configuration as was prevalent throughout most of automotive history.
28
Reference [89], pp. 43-44. The parameter set on those reference pages was used as input for a longitudinal
acceleration simulation termed “Road Test No. 1”. All the output from the simulation program
“MAXGLONG.BAS” proved to be very accurate with respect to the published empirical results for three
different Jaguar test configurations.
29
The gear set efficiencies were neglected as they contribute little to this discussion.

10
SAWE Paper No. 3602
Category Number 31.0

Table 2.2 – BASELINE EFFECTIVE MASS BREAKDOWN (1984)


From this table it is immediately evident to any mass properties engineer that, for a
“weight reduction program” 30, it can be far more effective to reduce the weight of a rotating
mass than that of any purely translational component. If the weight reduction cost per
pound were about the same, then reducing the “I2” weight by 5 lb (2.3 kg) is more effective
and cost efficient than reducing the chassis weight by 5 lb. In first gear, a 5 lb general
reduction in “I2” weight may incur 31 a corresponding 38.0 lb (17.2 kg) reduction in effective
weight, resulting in a total effective weight reduction of 43.0 lb (19.5 kg)!
However, there are some reservations regarding such a weight reduction of the
rotating engine masses. For one thing, the effectiveness of such a reduction decreases rapidly
with higher (numerically lower) gear; in second gear an “I2” weight reduction of 5 lb only
results in a 16.5 lb (7.5 kg) total effective weight reduction, and by the time overdrive is
reached a 5 lb “I2” reduction represents a mere 7.0 lb (3.2 kg) total effective weight
reduction.
For another thing, for conventional internal combustion piston engines the magnitude
of the “I2” rotational inertia actually serves a purpose, so much so that the naturally
occurring inertia of the engine proper is generally augmented by the use of a flywheel. The
purpose of so much rotating mass is to even out the combustion and inertia pulses inherent
to reciprocating engine operation, and to provide enough stored rotational kinetic energy so as
to counteract the tendency to stall under load at low rpm 32. For racing purposes recourse to
light weight flywheels, and sometimes even to a “shaving” of the crankshaft counterweights,
may be made to improve acceleration, but such modifications tend to render an engine

30
It should be noted that the term “effective weight reduction” might be more appropriate.
31
Here and elsewhere in this discussion, where such reduction is a matter of a general consideration and not a
matter of specific design changes to a particular part, it will be assumed that a weight reduction will be
accompanied by a corresponding proportionate reduction in rotational inertia.
32
Reference [89], pp. 30-32. For detail information regarding design of rotational engine components there are
excellent sources such as Reference [23] or Reference [59]. Also see “Appendix F – Reciprocating Piston Engine
Inertial Energy Loss” of this paper.

11
SAWE Paper No. 3602
Category Number 31.0

unsuitable for road use. Clever devices such as the Suzuki variable inertia flywheel 33 allow
for some freedom from such compromise, but for the most part reduction of engine related
rotational inertia in order to improve performance involves considerations beyond the scope
of this paper.
This paper is based on the idea that it might be more effective, and very likely more
cost efficient, to seek the reduction of some rotating mass category other than “I2”. To
facilitate consideration of just where it may be best to seek reduction, the following Table 2.3
was prepared. This table shows the relative contribution of the various rotating mass
categories to the total vehicle effective mass in each gear:

Table 2.3 – BASELINE RELATIVE EFFECTIVE MASS CONTRIBUTORS


As noted, even if the “I2” weight/rotational inertia can be reduced significantly
without adversely affecting drivability and “out of the hole” 34performance, the benefit
decreases rapidly with gear (speed). An “I3” reduction would be much more consistent in
benefit, but would also be much less significant. However, the “I1+I4” rotating masses
promise both significant reduction and consistency across the speed range, so those masses
warrant a closer look as per the following Table 2.4:

Table 2.4 – BASELINE I1 & I4 MASS CONTRIBUTORS

33
Reference [89], pg. 59. As implemented on the 1983 Suzuki GR-650 motorcycle engine, this flywheel presents
a relatively high rotational inertia below 2,500 rpm, but above 3,000 rpm a centrifugal clutch disengages an
auxiliary mass thereby greatly reducing the inertia. This provides smooth running and high acceleration from a
vehicle standstill or low speed, but faster throttle response and greater acceleration in the higher vehicle driving
speed range.
34
To “jump out of the hole” constitutes drag racing parlance referring to the initial stage of hard acceleration
from a standstill when the “stored” rotational energy of various engine masses (“I2”) plays a large role.

12
SAWE Paper No. 3602
Category Number 31.0

Over 86% of the weight and 94% of the rotational inertia in this category is supplied
by the vehicle wheels/tires, with the remainder supplied by items closely associated with
and/or affected by the wheel/tire sizing 35. So, the idea of reducing effective mass, and thereby
improving performance, by reducing wheel/tire size quite naturally comes to mind.
Because such a mass properties based performance enhancement does come quite
readily to mind it is easy to find authoritative support for it. Pete Brock 36, noted automotive
designer and stylist, has been quoted as enumerating the benefits of wheel/tire reduction 37:

“First of all, there’s inertial weight. Every time you accelerate


a car, or brake it, you’re wasting a lot of energy because of the
rotational inertia of the tires and wheels. Second…is rolling
resistance…third is unsprung weight…Next is aerodynamic
drag…(then) the…(wheel/tire contribution to)vehicle weight
…last(ly)…little tires are cheaper…”.
Mr. Brock went on to claim that, given all the advantages of smaller wheels/tires, the
choice of a correspondingly “appropriate power plant” (i.e., smaller) could result in a
“cascade effect” of weight reduction throughout the entire vehicle design; the culmination
could be a “1100 lb (499 kg) four-passenger sedan” with “the performance of a 289 Cobra”.
Unfortunately, Mr. Brock’s approach to reducing tire/wheel size was to reduce the
width, which is unlikely to be the best way to go about it. In an industry news article
appearing some years subsequent to Mr. Brock’s comments, the giant French tire
manufacturer Michelin revealed the results of some physical exploration of the effects of
tire/wheel size reduction by decreasing the diameter 38. At the 2010 Challenge Bibendum in
Rio de Janeiro Michelin presented as its energy-efficient entry 39 a 175/70R10 tire which was
designed to replace a conventional sized 175/65R14 tire without any sacrifice in performance,
longevity, or load-carrying capacity 40. To demonstrate the benefits, Michelin provided a stock
Citroën C2 (14 inch wheels) and a modified Citroën C2 sporting the down-sized tires, and let
automotive journalists form their own opinions as to the efficacy of the innovation by driving
the vehicles around the Nelson Piquet International Autodrome; the overall journalist reaction
seemed to be favorable.

35
The wheel hubs would decrease proportionately, but what would happen to bearings, gears, and half-shafts is
more problematic.
36
Pete Brock has had a storied and varied career, initially as a stylist for General Motors, then as a manager and
designer for Shelby American, moving on to twenty years as an automotive photojournalist, and lastly as an
automotive consultant. Throughout much of his career he also managed his own automotive design firm and
competed in various levels of automotive racing.
37
Reference [77], pg. 33.
38
Reference [74], pg. 108.
39
Not only was the Michelin entry touted for saving energy in operation, but for economizing in the use of
petroleum-based artificial rubber in its construction.
40
This was achieved through careful manipulation of tire carcass construction, rubber compounds, and tread
pattern; reportedly there was even a 5% increase in load rating!

13
SAWE Paper No. 3602
Category Number 31.0

The transformation to the new tire/wheel paradigm came with various consequences.
The immediately measurable consequences was a vehicle weight reduction of about 88 lb (40
kg) and a vehicle overall height reduction of about 1 in (2.5 cm), but the smaller wheels
required the fitting of heavy duty brakes and fairings about the wheel well openings. Whether
this last was done to maximize the aerodynamic advantage afforded by the change, or to
improve the visual aspect of the modified vehicle, which tended to be that of a roller skate,
was not made clear, but the latter seems the more likely. Henny Hemmes, the Auto Channel
Senior European Editor, has said 41:
“Even on such a small car, the optical effect was a bit
silly, although that was compensated a bit by fender
extenders in a different color…may not expect 10 inch
tires any time soon…since cars are designed with large
wheel wells…(for) large wheel/tire…stylists will have more
freedom when designing new cars (for small wheel/tire set-
up)”.
This comment really made two good points: cobbling 10 inch wheel/tires onto an
existing automotive design did not allow for taking the most advantage of the change either
functionally (space utilization, etc.) or stylistically. In the public mind performance is firmly
associated with large diameter wheels and wide low aspect ratio tires, which is a perception
reinforced by current automotive styling and advertising. John O’Dell, the Senior Editor of
the Auto Observer, an Edmund’s publication, foresaw that overcoming this well established
prejudice on the part of the consumer may require an all new design paradigm 42:
“…Michelin’s 175/70R10…looks more like it was
designed for a small trailer or kiddie car than for a real
highway-going vehicle. It is short, fat, and sort of silly
looking – especially if fitted on today’s cars with outsize
wheel wells designed to house the 18-to-20 inch rim-and-
tire packages that have become de rigueur the world
over.”
“We can…predict that the 10 inch (wheel/tire) is going to
require an all new car design (in order) to make any
inroads (in the automotive market).”
Car and Driver magazine also has made note of how aesthetics may be trumping
function, and became determined to investigate into the truth of the matter 43:
“To examine the effects of installing larger wheels and
tires, also known as “plus-sizing”, we tested five different
wheel and tire combinations – ranging from 15 to 19

41
Reference [35], pg. N/A (Internet).
42
Reference [61], pg. N/A (Internet).
43
Reference [64], pg. 110.

14
SAWE Paper No. 3602
Category Number 31.0

inches – on a 2010 Volkswagen Golf and got a good sense


of what is gained and lost in the process.”
While the results of the Car and Driver test are indicative, it is important to remember
that the test was carried out within the constraints of the “plus-sizing” concept. “Plus-sizing”
refers to the modern practice of going up a wheel size while going down commensurately in
tire aspect ratio so that the overall diameter stays essentially the same, although the section
width tends to increase gradually. Therefore, given this constraint on the wheel/tire size
changes affected during this test, one can’t expect to observe dramatic consequences. What
consequences that were observable are summarized in the following Table 2.5:

Table 2.5 – CAR & DRIVER WHEEL/TIRE “PLUS-SIZE” TEST RESULTS


This table reflects the result of a progressive plus-sizing (…+1, +2, +3, +4) from the
stock wheel (15 × 6.0)/tire (195/65R15 91V) combination to the ultimate wheel (19 × 8.5)/
tire (235/35R19 91W) combination. Note that there was very little increase in overall
diameter throughout, as indicated by the very slight increase in speedometer error, so the
results are not fully indicative of all the effects that accompany large changes in rolling
radius. Yet still, as wheel/tire “size” went up, so did cost (+126%), 0-60 mph E/T (+4%), 0-
100 mph E/T (+5%), ¼ mile E/T (+1%), sound level (+1%), and unsprung weight (+35%!!).
What decreased was fuel economy, and that was by almost 10% (-9.4%)! The only
measurable improvement was an increase in lateral acceleration (+6%), and it is doubtful that
such a difference would ever be of notice in actual use of a street vehicle.
What presumably was of notice, and provided sufficient motivation for folks to shell
out good cold cash in exchange for something that essentially resulted in a degradation of
performance across the board, was “feel” and “looks”. By “feel” is not meant the increased
harshness and sound that often accompanies an increase in wheel size, but the more

15
SAWE Paper No. 3602
Category Number 31.0

responsive steering that resulted from decreased tire sidewall height 44, which has a solid basis
in reality. Although the really great motivator for the present use of oversize wheels and tires
is the subjective matter of appearance, historically there existed more objective factors. Larger
tires, when not limited by the constraints of the “plus up” system, may be generally associated
with greater load carrying capacity, longer wear, better rough or off road capability. Larger
diameter tires will also tend to have greater traction through an increased contact patch with
the road, although a better way to increase contact area may be via increased tire width.
However, early in automotive history there was not such a uniform availability of tire
widths and sizes across a broad spectrum of gradation as there is today 45. Just after the “turn
of the century” (1900) most passenger car tires had an outer diameter of 28 to 34 inches (71 to
86 cm) in 2 inch (5 cm) increments, a width of 3 to 4 inches (8 to 10 cm) in half inch (1 cm)
increments, and had an aspect ratio of 1.00 46. Even within these bounds there could exist large
gaps in availability. For instance, a “30×3½” tire would require a 23 inch diameter wheel, and
to go up in wheel size by 1 inch (“plus one”) would be impossible; the best that could be done
was perhaps to obtain a 25 inch diameter wheel, but that could result in the need to use a
“32×3½” tire, which was not a common size. Given this sort of problem, plus the fact that just
mounting a non-OEM wheel to a vehicle might be a challenge due to the lack of
standardization at the time, early experimentation in wheel/tire size was rather restricted.
But, although the present day freedom to modify wheel/tire sizes to suit personal
preference or need didn’t exist historically, opportunity to modify the wheel/tire set-up
presented itself in other ways, such as changing the accepted automotive paradigm. The four
wheel/two axle paradigm for passenger cars has existed for so long that it constitutes
“thinking out of the box” to recognize that such is not the only viable configuration possible.
Many of the earliest automotive attempts, starting with Cugnot’s steam tractor of
1769, were tricycle in layout; this layout greatly simplified the vehicle design, especially with
regard to steering, but inherent instability in roll led quickly to the dominance of the four
wheel/two axle paradigm. However, the limited durability and life of early tires led to some
interesting variations. One of the more interesting 47 was the “Octoauto” created by Milton
Othello Reeves (1864-1925) of Columbus, Indiana in 1911 48; the following figure is an

44
The decrease in sidewall height is largely the source of increased ride harshness, and also makes the expensive
flashy wheels more susceptible to contact damage with road irregularities and curbs.
45
Reference [26], pg. 190.
46
Reference [88], pp. 92-97. Early tire designations were generally universally given in English inch units as
“tire outer diameter” by “tire section width”; since the aspect ratio was almost always 1.0 it was easy to
determine the appropriate size wheel for a given tire (and vice versa), e.g.: a “28×3” tire would mount on a 22
inch wheel.
47
One of the most practical means for increasing the load capacity is the ubiquitous “dually axle” which
appeared very early, and is found today on everything from heavy duty pick-up trucks to tractor semi-trailers.
Such axles, which mount two wheel/tire units at each end, have also appeared from time to time in some more
unorthodox applications, such as race cars.
48
Reeves was not alone in his appreciation of the benefits of adding wheels/axles to the conventional
configuration of the day, there were many similar experiments. For example, a Liborio Rivas of Santa Rita, NM,
received United States Patent 1534810 in April 1925 for a vehicle very similar to the “Octoauto” (however,
unlike Reeves, Rivas never actually built his vehicle).

16
SAWE Paper No. 3602
Category Number 31.0

advertisement for the vehicle that appeared in a popular automotive magazine of the day, The
Horseless Age:

Figure 2.1 – 1911 REEVES “OCTOAUTO”ADVERTISEMENT


The chief advantage of the eight wheel/four axle layout was essentially that each
tire carried one-half the normal load. At a time when a tire could be expected to last no
more than 1,000 miles (1609 km), this was significant. As stated in the advertisement, tire
expense was halved and blowouts practically eliminated, which was the main intent.
However, other benefits were noted as well. The vehicle rode smoother, with less shock and
vibration, than other more conventional vehicles. A noted automotive journalist of the time,
Elbert Hubbard 49, wrote concerning the principles upon which the car was built 50:
“In the Reeves Octoauto the load is distributed over eight
wheels instead of four. In a four-wheeled auto, in the case
of hitting a rut, one wheel may for an instant carry half the
load, and it is this sudden jolt that causes tire trouble…your
tire reaches its limit and explodes…If you are running fast
49
Elbert Hubbard and his wife were drowned in the sinking of the Lusitania in 1915, about four years after his
review of the “Octoauto”.
50
Reference [56], pg. C-2.

17
SAWE Paper No. 3602
Category Number 31.0

you may lose control and the ditch gets you…so if you can
save your wheels from these severe jolts, you will prolong
the life of the tire, the car, and its occupants.”

Mr. Hubbard also wrote about his experiences on a test drive 51:

“We took the (busted up) pavement at the rate of 25 mph


(40.2 kph), oblivious of the ruts…were over 3 feet (0.9 m)
apart, but so evenly was the weight distributed that before
one wheel could really hit the bottom of a rut the wheel
behind was on firm footing and relieved the strain. This
taking ruts and bumps without a jar is something that no
man can possibly appreciate who has not experienced a ride
in an Octoauto…(the) Octoauto steers and controls exactly
the same as a four-wheeled car…(is) the easiest-riding car
in the world…and the easiest car in the world on tires.”

Despite Mr. Hubbard’s enthusiasm for the Octoauto he never bought one, and no one
else did either. The problem was that the Octoauto cost $3200 ($74,688 in 2010 value), which
was $2000 ($46,680 in 2010 value) more than the 1910 Overland 52 Model 40 on which it was
based, and that was already a fairly expensive car for the time ($28,008 in 2010 value).
Although the Octoauto (and its less expensive successor, the Reeves Sextoauto 53) failed
commercially, it gave some hints of what might be achieved through a radical reconsideration
of the conventional wheel/axle paradigm.
The Reeve’s Octoauto concept certainly was not beneficial from a mass properties
standpoint; the effective mass was significantly increased and the unsprung mass values
were not reduced. Such a multi-axle configuration also has inherent steering, drive train, and
stability problems the solution of which usually tends to incur weight penalties. The
Octoauto dealt with these problems in the simplest way possible, or just avoided them
altogether. Regarding the geometry of 8-wheel/4-axle layouts, such as the Octoauto, an
authoritative source, The Motor Vehicle, has this to say 54:
“On eight-wheel vehicles all four front wheels must be
steered. A similar requirement arises on…six-wheel
vehicles having a preponderance of weight at the front,
and…therefore…only a single axle at the rear. Both cases
are represented in Fig. 34.4.”

51
Ibid.
52
The Overland Automobile Company was founded in 1903. In 1908 the Overland Company was purchased by
John North Willys, and in 1912 became the famed Willys-Overland Company.
53
The failure of the Octoauto to sell prompted Reeves to construct a less expensive “Sextoauto” with only one
axle at front, but still with two rear axles. The absence of a second front axle, along with its steering mechanism,
promised a substantial reduction in cost. However, the new design sold no better than the first, and Reeves
eventually gave up the effort of trying to sell such radical vehicles.
54
Reference [59], pp. 874-875.

18
SAWE Paper No. 3602
Category Number 31.0

The “Fig. 34.4” spoken of in the above quote is hereby presented as Figure 2.2:

Figure 2.2 – EIGHT WHEEL/FOUR AXLE GEOMETRY

The Octoauto had just such a four-wheel steering mechanism at the front, with the first
axle steering synchronized with the second axle steering by means of a linkage such as the
link “V-U” shown in the figure. The Octoauto avoided rear drive synchronization problems
by simply supplying power to only the first (“F-F” in Figure 2.6) or stock “live” Overland
Model 40 beam axle; the second axle was a simple “dead” beam axle mechanically
unconnected to the first.
The front steering on the Octoauto was just a very simple matter as illustrated in
Figure 2.2 because the front axles were also simple dead beam axles; the associated weight
penalty was probably relatively small. When the situation is more complicated, say that of a
modern independent suspension, then the synchronization of the steering of two front “axles”
must also be more complicated. The 1976 Tyrrell P34 had exactly that kind of complicated
six-wheel steering geometry, and thereby incurred a significant weight penalty, yet still may
have experienced some overall mass properties benefits.

19
SAWE Paper No. 3602
Category Number 31.0

The 1975 Elf Tyrrell P34 had four undriven wheels forward on two separate axle lines
in tandem, plus two driven wheels aft on a common axle line. Designer Derek Gardner
created this unique Formula 1 55 race car supposedly in the hope of cleaning up the air flow (in
particular, decreasing front end “lift”), increasing front tire ground contact, and increasing
front swept brake area. He attempted to accomplish this through the use of four special size 10
in (25.4 cm) diameter front wheels 56 (on a narrow track which brought them inside the aero-
shielding effect of the rule-limited width of the front “shovel” nose) but with two more usual
sized 16 in (40.6 cm) wheels aft 57. All seemed good, except there was a weight gain of
approximately 56 lb (25.4 kg) 58, which may be attributable mainly to the extra steering
mechanism as shown in Figure 2.3 59 (note the greater sophistication than the “View V-U”
mechanism of Figure 2.2) :

Figure 2.3 – TYRRELL P34 STEERING


However, this weight gain may have been offset by a reduction in effective mass due
to the reduction in front wheel size, and there certainly was a reduction in unsprung mass.
Such reductions may have aided in this design vindicating itself very early on. Down in power
with respect to much of the competition (the Ferraris reportedly had about 40 hp/30 kW
more 60) the P34 seemed to have an advantage on the more twisty circuits. In 1976 a long
series of successes ( 2nd & 3rd at Monaco GP, 1st & 2nd at Swedish GP, 2nd at Watkins Glen,
2nd at Japanese GP) culminated in the P34 placing second and third overall for the season!
55
Reference [5], pg. 97: As early as 1969 Gardner had planned on the concept being a 4WD Indy car (power to
the 2nd and 3rd axles only), but then USAC banned 4WD. Apparently Gardner felt that for Formula racing 4WD
was less suitable, or maybe he was just clairvoyant, as by 1983 the FIA had also banned 4WD (and much else).
56
FIA standard practice is to refer to a “complete wheel diameter”, i.e., the outer inflated diameter of wheel plus
tire, which would make it possible to also refer to the P34 10 in (25.4 cm) wheel as a 15 in (38.1 cm) “wheel”.
57
Or, in FIA “complete wheel diameter” terms, this would be about a 26 in (66.0 cm) “wheel”.
58
Reference [69], pg. 93.
59
Reference [31], pg. 49.
60
Reference [5], pg. 97.

20
SAWE Paper No. 3602
Category Number 31.0

The steering and braking proved excellent, with only an initial “twitchy” characteristic which
was probably due to the relatively short wheelbase of 96.6 in (245.3 cm), and which was
corrected by remounting the batteries far forward so as to increase the yaw moment of
inertia 61.
For 1977 the P34 was redesigned with a wider front end, a longer wheelbase, a six (up
from five) speed transaxle, and “improved” aerodynamics. What definitely was not improved
were the special sized front tires; race car tire development in this period was very rapid, and
Goodyear could not justify the expense of keeping such a uniquely sized tire, utilized by only
one team, up to date. The additional weight (1400 lb/620 kg vs. 1310 lb/595 kg) and lack of
tire development doomed the P34 to a lackluster season, and by 1978 the 4-2-0 concept was
abandoned.

Figure 2.4 – THE TYRRELL P34


However, the P34 does not constitute the total saga of six wheels in Formula 1. In
1976 F1 designer Robin Herd, perhaps inspired by the Tyrrell Project 34 (P34) six-wheel F1
race car, constructed a six-wheel/three-axle variation on the standard March 761 he termed
the March “2-4-0” (Whyte’s Notation, see Appendix D). Herd reduced the rear “wheel” size
from the usual 26 in (66.0 cm) diameter to 20 in (50.8 cm) diameter (equal to the usual front
“wheel” diameter) 62 while doubling the number of rear wheels by having two rear drive axles
in tandem. Herd felt that this modification would improve the vehicle aerodynamics and
increase traction at the rear while avoiding the problems associated with unusual tire sizes 63.

61
Reference [31], pp. 49 & 104. The “twitchy” effect was possibly the result of the “Kz2 / (lf × lr)” ratio placing
the OC (oscillation center) too far aft of the LCG (see Chapter 6).
62
FIA practice to refer to “complete wheel diameter” can result in some confusion; the 20 in (50.8 cm)
“complete wheel diameter” may have actually had wheels (“rims”) of 13 in (40.6 cm) diameter.
63
Reference [97], pg. 95.

21
SAWE Paper No. 3602
Category Number 31.0

Unfortunately, March Engineering at the time was financially strapped, and the
development of a proper transaxle for the unusual dual drive axle configuration was not
possible. However, construction went ahead anyway as it was felt that the publicity generated
by such a novelty would result in some financial success even if there was no racing success.
The 2-4-0 started testing in late 1976 and immediately the transaxle casing revealed itself as
insufficiently stiff, allowing gears to slip out of mesh. The gearbox was crudely strengthened
(adding to the initial weight penalty), and the improved vehicle revealed itself as having
“incredible traction” in the wet, but further development was halted as problems continued.
The improved traction that was demonstrated by the 2-4-0 concept inspired British hill
climb competitor Roy Lane to think that the concept might still be valid for his very different
type of competition, where raw traction was much more significant. Lane bought a 2-4-0 rear
end and mated it to a March 771 forward structure. In the wet the car performed spectacularly,
but in the dry it did not do well. Reportedly the car was 270 lb (122.5 kg) “overweight”, but
perhaps even more significant was that the gears would often break because of the lack of a
differential between the two rear drive axles. The shaft between the two axles would “wind
up” as the car was driven in a turn, sometimes causing the car to “jump” sideways when the
shaft suddenly unwound. A proper transaxle incorporating a third between-axles differential
and making appropriate use of titanium was indicated, but was beyond Lane’s financial
capability, so the 2-4-0 concept died without the chance for vindication 64.

Figure 2.5 – THE MARCH 2-4-0


The problems experienced by the March 2-4-0 perhaps has as much to do with
steering as with propulsion (traction). To quote an authoritative source, The Motor Vehicle,
regarding such six-wheel configurations 65:

64
Reference [60], pg. 52.
65
Reference [59], pg. 874.

22
SAWE Paper No. 3602
Category Number 31.0

“With six-wheel vehicles perfect steering geometry is


unobtainable because the two rear axles always remain
parallel (not steered). An assumption is made…that the
vehicle turns about a center on an axis midway between
and parallel to…the two rear axles…There will…be some
tire scrub on both rear axles…accommodated by the lateral
flexibility…Even so, the two back axles are always placed
as close together as practicable…”
The situation can perhaps be better understood with reference to Figure 2.6. The
“View 1” of that figure shows the March 2-4-0 geometry; note that as per the rules for
establishing “Ackermann” steering geometry the provision of two rear axles results in the two
possible turn center locations of “1” and “3”, but making the “midway axis assumption” as
per the above results in the third turn center location “2”. The reason that this assumption has
any validity is that the less than “perfect” geometry of attempting to turn about centers “1”
and /or “3” results in intersecting tire travel paths. Since the rear tires can’t change in physical
location with respect to each other (i.e., actually intersect) due to the physical structure
holding them in a fixed space (spatial) relationship, the traction forces attempting to cause
motion along such intersecting paths instead only results in tire flexure and scrub. The forces
that would cause a turning motion about center “1” tend to be counteracted by the forces that
would cause a turning motion about center “3”, resulting in a compromise resultant motion
about center “2”.
Such compromise motion comes at the expense of some extra tire wear, but this wear
must often be within acceptable limits given the fact that such dual real axle configurations
are to be found on all manner of trailers, trucks, and busses, etc. However, in the extreme
situation of the March 2-4-0, a high power/ultra-fast/hard-turning race car, this conflict
between the two rear axles as to which would determine the turn path geometry (not ideal for
a race car in any case) was dramatically revealed via the shaft connecting differentials “A”
and “C”. The forces which otherwise would have been relegated to merely causing some tire
scrub now were able to cause a torsional distortion of the shaft, resulting in a build-up of
strain energy which found release resulting either in the curious sudden sideways jumping
motion of the vehicle, or in the shattering of differential gearing.

23
SAWE Paper No. 3602
Category Number 31.0

Figure 2.6 – SIX WHEEL/THREE AXLE GEOMETRY


If this conflict in turning between the two rear axles did indeed contribute to the
March’s problems, then the obvious solution diagramed in “View 3” of a third differential
“B” could be supplemented, or perhaps even supplanted, by the provision of a second steering
mechanism as illustrated in “View 2”. A concern with the addition of such an extra steering
mechanism was the weight penalty incurred by the complex modifications to the axle itself,
and the weight incurred by the need to synchronize the action of the new steering mechanism
with the steering of the front axle, which in this case was located so very far away.
The contest as to which turn center (“TC”) is to dominate is not the only such conflict
that can be engendered by the use of multi-axle layouts; as mentioned earlier there is a
possible stability problem that results from inadvertently creating multiple roll axis
possibilities. Traditional four wheel/two axle layouts produce two roll center (“RC”) locations

24
SAWE Paper No. 3602
Category Number 31.0

via the geometry of the front and rear suspension systems 66; these RCs define the “roll axis”
which usually is down at the front and rising to the rear as shown in the following Figure 2.7:

Figure 2.7 – DETERMINATION OF ROLL AXIS ORIENTATION


This “nose down” orientation of the roll axis insures that any roll moment (lateral
force through the CG times the distance to the roll axis) will be resisted more by the front
suspension than by the rear suspension. If the tire and roll resistance characteristics and tires
are the same front to rear, this means that the vehicle will tend to behave in an
“understeering” (as opposed to “oversteering”) fashion (i.e.: be directionally stable).
The key to determining the roll axis location was the fact that it only takes two points
to define a straight line, in this conventional case the front and rear RCs. But, if more axles
are added to the configuration then things can become somewhat indeterminate if care has not
been taken to keep all the roll centers on the same line. When Daimler Chrysler of Brazil
added a third axle to an existing bus design for increased load capacity the prototype IBC-
2036 exhibited “undesirable behavior” during lane change maneuver. To identify the cause
and find a solution to this problem Daimler Chrysler engaged the services of Debis Humaitá
IT Services Latin America Ltd.
IT Services studied the behavior of the IBC-2036 bus via a computer simulation using
a popular multi-body modeling program (ADAMS) in conjunction with the “Delft Tyre
Model” 67. They found that the roll center of the new third axle suspension (see Figure 2.8)
was 14.06 in (371 mm) below the line between the first and second axle roll centers. Ideally
66
Reference [21], pp. 157-159. The roll center determination depicted in Figure 3.3 is for the “geometric roll
center” (GRC); historically this has been taken as being equivalent to the “kinematic roll center” (KRC), the
“force roll center” (FRC), and the “moment roll center” (MRC), but there are real distinctions between these roll
center types.
67
The “Delft Tyre Model” is based on Prof. Pacejka's famous “Magic Formula” for describing tire
characteristics and is used throughout the automotive industry for the assessment and modeling of tire behavior.

25
SAWE Paper No. 3602
Category Number 31.0

the third axle should have been revised so its RC would fall on the line, but practical
considerations of design modification complexity and expense forced a compromise solution.
That compromise solution consisted of modifying the rear suspension to get the third RC to
fall 6.42 in (165 mm) above the line; the compromise RC is closer to the ideal position by
7.64 in (206 mm). This modification to the prototype bus brought the yaw “wiggle-wobble”
during a lane change maneuver to within acceptable limits (although still not ideal) for
Daimler Chrysler of Brazil 68.

Figure 2.8 – CONFLICTING ROLL AXES


The Reeves Octoauto probably inadvertently side stepped such problems as plagued
the IBC-2036 bus by virtue of the likelihood that all the axle suspensions were geometrically/
kinematically the same; the roll axis was probably a straight line in parallel with the ground
plane a few inches above the axle centers. Just how the Tyrrell P34 and the March 2-4-0 dealt
with the matter of roll center alignment for three axles is not known, but the probability is that
the designers were well aware of the situation and acted appropriately. In any case, the Tyrrell
P34 and the March 2-4-0 were not the last of the modern six-wheel/three-axle experiments in
Formula 1 racing 69.
In 1982 designer Patrick Head designed the Williams FW08B with a six-wheel/three-
axle configuration like the March 2-4-0, mainly with the intent of improving underside
aerodynamics (“ground effect”). Using lessons learned from Lane’s March hill climb special,
famed transmission firm Hewland Engineering Ltd assisted in the design of the transaxle.
Late that year a newer version termed the FW08D was tested at Donington Park and produced

68
Reference [19], pp. 4-5.
69
The experimental 1976 Ferrari 312T6 had six wheels, but only two axles, so it is not covered herein. The rear
of the Ferrari had four wheels mounted on a single rear axle in the manner of a “dually” pick-up truck. This was
in imitation of such arrangements as sported by various race cars of the 1920’s to 40’s; it made sense then given
the state of tire technology, but not in the 1970’s. The 312T6 never raced; it was well in excess of the FIA
maximum width for Formula1and apparently did not perform well.

26
SAWE Paper No. 3602
Category Number 31.0

lap times that were considered stunning, but all was for naught as the FIA soon banned any
configuration other than a four wheel configuration in Formula 1, along with 4WD and
“ground effects”. The FIA also set the F1 minimum weight limit to 540 kg (1190.5 lb) 70.
Consequently, the FW08D was retired without ever having turned a wheel in anger.
Thirteen years later, at the 1995 Goodwood Festival of Speed, in the hands of amateur race
driver Dr. Jonathan Palmer, the FW08D set an elapsed time that was only beaten four years
later by professional driver Nick Heidfeld in the famed 1998 McLaren MP4/13.

Figure 2.9 – THE WILLIAMS FW08D


The point is that all these vehicles: Octoauto, P34, 2-4-0, and FW08D constitute
design departures from the conventional paradigm which must have significantly affected the
translational/rotational inertias, resulting in changes to the vehicle effective mass and the
unsprung masses. In each case certain performance benefits were noted despite the design
changes being accomplished without a systemic mass properties approach, and in each case
design development was discontinued without the full potential benefit being realized. So
close, yet so far…

70
The 1982 Williams FW08D was in no danger of being under the minimum weight limit; the vehicle may
have had about 220 lb (100 kg) more avoirdupois than its adversaries. The 1976 March 2-4-0 may have been
anywhere between 90 lb (40.8 kg) and 270 lb (122.5 kg) overweight. The weight situation for the Tyrrell P34
seems clearer; the 1976 vehicle weighed 595 kg (1310 lb), which reflected a 56 lb (25.4 kg) steering mechanism
weight penalty, and the 1977 version weighed 620 kg (1400 lb). The FIA F1 minimum weight limit (which
includes driver and fuel) for 1976 was 575 kg (1268 lb), and was revised upward for 1982 to 580 kg (1279 lb).

27
SAWE Paper No. 3602
Category Number 31.0

3 – ENERGY GENERATION AND MASS DISTRIBUTION

Whether the wheel/axle layout of an automobile should be a tricycle configuration, or


a four wheel configuration, or even six or eight wheel, was a question that was “settled” fairly
quickly in favor of the “conventional” four wheel configuration; the question of how the
energy required to motivate the automobile was to be generated took a while longer. Around
1900 there were essentially four major contenders for the role of automotive propulsion
system: steam, electricity, internal combustion piston engine, and the hybrid 71. Each type of
propulsion system had its own particular impact on the character of the automobile, its
performance, and its mass properties.
Historical anomalies aside, like Leonardo DaVinci’s spring powered tricycle (c. 1478)
and Ferdinand Verbeist’s steam turbine driven five-wheel cart (c. 1672), the first vehicle that
could reasonably be called a “true” automobile was Nicolas Cugnot’s steam-powered piston-
engined tricycle of 1771. Cugnot’s “Fardier a Vapeur” failed to win the approval of its
intended military customer, and its performance was underwhelming anyway, but steam had
announced its candidacy for the propulsion system of the future. The genesis of the electric
automobile can probably be regarded as being the development of the battery by Alessandro
Volta in 1800 (although rechargeable batteries appropriate for practical electric car use
awaited development by Gaston Planté in 1856), but as there are so many contributors to, and
constructors of, early electrical “cars” it is probably best to only note that the electric car was
well established in Europe by around 1880 and in the United States by 1895. As to the first
piston engine, gasoline powered car, the credit for that conventionally goes to Karl Benz for
his Motorwagen of 1886 (German patent number 37435).
The situation with respect the first hybrid car is much clearer. In 1898 a young (23
years old) Ferdinand Porsche was given employment at the traditional carriage manufacturer
of Jakob Lohner & Company for the express purpose of developing a modern alternative to its
traditional product. Soon a remarkable departure from the usual horse-drawn conveyance was
unveiled in April of 1900 at the Paris World Exhibition: a carriage propelled by electricity.
The Elektromobile involved two front wheel electric hub motors (FWD!) powered by lead-
acid batteries. Shortly afterward the basic concept was greatly enlarged and modified to
include two more hub motors (“radnabenmotors”) at the rear wheels (4WD!) in response to a
special order from Englishman E.W. Hart of Luton. This vehicle was intended to be a record
breaking competition car, but was somewhat hampered in this endeavor by its 3,968 lb (1,800
kg) battery pack (!). It was exhibited by December 1900 in the National Cycle Exhibition at
the Crystal Palace in London under the name Toujours Contente (“Always Contented”) 72.

71
C. 1900 in the US about 40% of the autos were steam powered, 38% were powered by electricity, and 22%
were powered by gasoline engines. The European auto population probably had a similar composition, except
that the hybrid already had established a presence; it would take until 1911 for a hybrid to appear for sale in the
US (Woods Motor Vehicle Company “Dual Power” Model 44, US Patent 1244045). C. 1896 US engineer Harry
E. Dey constructed a flawed hybrid prototype of very advanced specification, but it never reached production.
72
The Belgian race car driver Camille Jenatzy had set the land speed record (LSR) of 65.75 mph (105.904 kph)
in 1899 with a Jeantaud-built electric powered streamliner named the Jamais Contente (“Never Contented”);
Hart probably intended to emulate this sort of achievement, although probably in more of a “trials” venue, but
his vehicle lacked the aerodynamic and light-weight refinement of Jenatzy’s LSR vehicle.

28
SAWE Paper No. 3602
Category Number 31.0

It was in this innovative spirit that Porsche took these advanced design trends to their
ultimate conclusion; he replaced the enormously heavy battery pack with a much lighter one
in conjunction with two small gasoline engine/generator units for a proof-of-concept
prototype called the Semper Vivus (the production vehicle would have a single
engine/generator); final drive was by electric hub motors at the front wheels. This
wonderfully prescient drive concept was called the Mixte (“Hybrid”); although in its details it
may be unlike most such concepts today, it is the first viable gasoline-electric hybrid
automobile 73. The Mixte concept ultimately went through many variations and considerable
development to result in a Lohner production model in 1903.

Although each of the four major contenders had its particular advantages, and
disadvantages, as a source of automotive propulsion, by the 1920’s the internal combustion
engine (ICE), in its reciprocating piston/gasoline powered form, was the overwhelming
winner of the automotive propulsion crown. However, interest in the other propulsion systems
did not just vanish, and new developments of one or the other (or even some relatively new
concept) has flared up from time to time, usually in accord with changing circumstances (fuel
shortages/price increases, government regulation, etc.).

The ascendency of the reciprocating piston 74 gasoline fueled engine in the propulsion
of the automobile had definite consequences with respect to the weight and mass
distribution of the automobile. With regard to weight, historically the ICE came to produce
much more power per unit weight (and volume) than the other contenders, which was a major
factor in its success. For one thing, the gasoline engine allowed a general overall lightening
of the entire automobile for the capability presented. However, there remained room for
improvement; automotive internal combustion engines generally need to be oversized in order
to have enough energy available for an acceptable rate of acceleration, but at cruising speed or
idle the extra engine capacity, and associate extra weight, generally represented a waste of
resources.
The nature of the piston ICE, and the transmission that necessarily accompanies it,
tends to result in a concentration of mass that greatly affects the mass distribution of the
entire vehicle. The exact effect(s) depends upon the location and orientation of the engine/
transmission; the following table illustrates the effect upon the weight, center of gravity,
moments of inertia, the XZ product of inertia, and the longitudinal principal axis angle
(“λ”) upon a “typical” (example) vehicle of various internal combustion engine/transmission
installations, everything else held constant (et ceteras parilis), which in reality is never quite
the case 75:

73
Reference [15], pg. 19. It should be noted that this source is replete with errors; the picture on page 18 is not
of a Mixte (gasoline/electric hybrid) but of the all electric Toujours Contente.
74
The repetition of the words “reciprocating piston” is done in recognition of the existence of other forms of
internal combustion engines (ICE), such as the turbine or the “rotary” (e.g., Wankel) engines. The word
“gasoline” is frequently used for the same reason; other fuels such as benzene, alcohol, hydrogen, and even
gunpowder (e.g.: Christiaan Huygens 1678, George Cayley 1807, etc.) have been used at times, with interesting
results.
75
Reference [92], pg. 24.

29
SAWE Paper No. 3602
Category Number 31.0

Table 3.1 – GASOLINE I.C.E. LOCATION/ORIENTATION EFFECT ON


CONVENTIONAL AUTOMOTIVE MASS PROPERTIES
The preceding table limits itself to front wheel drive (FWD) and rear wheel drive
(RWD) configurations. However, any of the configurations of Table 3.1 can be adapted to
operate as a four wheel drive (4WD)/all wheel drive (AWD) vehicle, which has a further
effect on the mass properties. Around 1969 it was estimated that the minimum weight
penalty of adding 4WD to Formula 1 race cars was about 90 lb (40.8 kg) 76, but obviously that
amount would vary given a different application (e.g.: passenger car, pick-up truck, etc.)/
level of technology/details of implementation.
The point is that the nature of the propulsion unit, its installation location/orientation,
and the exact manner in which the power is delivered to the drive wheels, has a notable
76
Reference [46], pg. 204. Actually, that estimate may have been very optimistic at the time; the Lotus Type 64
was at least 166 lb (75.3 kg) overweight, and most of that may have been due to the 4WD. Interestingly, F1
designer Robin Herd claimed a weight penalty of 90 lb (40.8 kg) in 1977 for his variation on 4WD as
implemented in his March 2-4-0, which was essentially a standard March 761 with four rear drive wheels instead
of two. However, in 1979 British hill climb contender Roy Lane found that his March 771 when modified with
the 2-4-0 rear end resulted in a 270 lb (122.5 kg) overweight.

30
SAWE Paper No. 3602
Category Number 31.0

impact on the vehicle mass properties. To state the matter somewhat differently, in what
might be considered an ancillary to that truism: the number of wheels to be driven may have a
significant effect on the mass properties of a vehicle not only in and of itself, but also
through consequent effect on the propulsion system, and vice versa. Table 3.1 illustrates that
point to a certain extent, but only for the traditional ICE propulsion system and the
conventional 4-wheel/2-axle layouts (FWD and RWD). The choice of wheel/axle layout
interacts with the propulsion system type to the extent that an unconventional wheel/axle
layout may practically “require” accompaniment by an unconventional propulsion system in
order to maximize the benefits.
One would expect for an unconventional propulsion system, such as installed in a
modern electric vehicle (EV) or hybrid electric vehicle (HEV), that the mass properties of
the vehicle would be considerably at variance with the possibilities depicted in Table 3.1.
Certainly with a pure (recharge via the commercial power grid) battery electric vehicle (BEV)
it would seem likely that conventional mass properties characteristics would be totally
irrelevant. This would be due to the greater freedom in location of an electric system’s
components, as afforded by the flexibility of interconnection by wire; such freedom could
allow for much greater leeway in obtaining optimum mass distribution. On the other hand,
one would expect modern hybrid car design to be much more restrained, with mass
properties not too unlike those characteristic of conventional internal combustion vehicles 77.
Such similarity in mass properties with past conventionality would increase with the degree
to which the hybrid design depends on its conventional internal combustion engine for
primary motivation, and with the degree to which the hybrid layout is forced to mimic the
conventional for reasons of space utility and/or because the chassis was originally designed
for conventional propulsion.
Consider the case of the 2012 Prius C, the most basic of the famed Toyota Prius
family of HEV’s which had its origin about 14 years previous. The Prius C (“C” stands for
“city”) is a subcompact of FWD “5-door” hatchback configuration which is very prevalent for
vehicles of this class. The design is totally biased toward practicality and fuel economy, with
104.5 ft3 (2.96 m3) of interior volume 78 and an EPA rated fuel economy of 53 mpg (4.4 L/100
km) city/46 mpg (5.11 L/100 km) highway. The 1.5 liter (91 cid) gasoline engine (4-cylinder
DOHC, Atkinson cycle, VVT, Auto Stop-Start) produces 73 hp (55 kW), and is augmented
by an electric motor of 60 hp (45 kW), which results in a rated total output of 99 hp (74
kW) 79. The engine, motor, generator, and CVT transaxle are located as a unit mass at the
front axle line; other components, such as the Inverter, Converter, Step-Up Transformer, and
Hybrid Control Computer are located nearby. The nickel-hydride battery pack, cooling

77
Ergo, much that has been learned about the mass properties of conventional vehicles is not necessary made
irrelevant by present and/or future advances in automotive design. The future is always still rooted in the past;
future technology just represents a modification and extension of past knowledge. In recognition of this
unbreakable link with the past Isaac Newton said: “If I have seen further than others, it is because I have stood
on the shoulders of giants”.
78
This apportions out to a passenger volume of 87.4 ft3 (2.48 m3) and cargo volume of 17.1 ft3 (0.48 m3) when in
full five passenger use.
79
Note that the rated combined power is not a simple addition of engine and motor power (rated electric motor
power is for a maximum output which is not long-term sustainable).

31
SAWE Paper No. 3602
Category Number 31.0

blower/duct, 12 volt conventional lead acid battery, and the 9.5 gallon (36 L) gas tank are
located aft under/near the rear seat.
Based on the total power output, the electric motor is only providing about 26% of the
C’s primary motivation, so the presence of an electrical propulsion system would not seem to
be of overwhelming dominance; this would seem to auger for a limited influence of the
electric on vehicle mass properties. Moreover, the massing of all major components at the
front axle, and the aft placement of the energy storage units (small gasoline tank and large
battery pack) in the same location where a large gasoline tank would conventionally lie,
further suggests that the mass properties of the Prius C might not be at great variance with
the past despite its modernity. This hypothesis may be tested by applying those mass
properties relationships (WER’s) as were developed for conventional FWD vehicles to the
Prius and noting how close those estimations come to the quoted mass properties values; if
the estimated and “actual” values are a reasonable match then the hypothesis regarding the
continuity of mass properties relationships would seem to be supported 80.
Unfortunately only two mass properties characteristics of the Prius C can be
validated at the time of this writing; the curb weight at 2,587 lb (1173 kg) and the weight
distribution at 61.3% front/38.7% rear 81, the latter meaning that the LCG is 38.9 in (98.7
cm) aft of the front axle line. Other vehicle characteristics pertinent to this investigation are 82:

Table 3.2 – 2012 TOYOTA PRIUS C DIMENSIONS

The following FWD weight estimation relationship (WER) produces 1077 kg (2375
lb) for the curb weight when the appropriate values (in meters) from Table 3.2 are “plugged
into” the equation 83:

80
Note that the possible results of such a “test” can’t be considered definitive; even if the estimated mass
property values are exact matches with the “actual” values that would only lend support to the hypothesis as it
falls far short of a “proof” due to the limited number of test cases.
81
These figures apply to a Prius C equipped with the optional 16×6 inch cast aluminum alloy wheels (shod with
Bridgestone Turanza EL400 D2 P195/50R16 83V low rolling resistance tires) and sunroof. The standard Prius C
wheels are 15 in with styled plastic wheel covers.
82
All specifications for the 2012 Toyota Prius C are per “www.edmunds.com”, the internet web site (founded
1997) for what was once the print automotive publishing firm of Edmunds Inc. (founded 1966).
83
Reference [88], pg. 23.

32
SAWE Paper No. 3602
Category Number 31.0

𝑾𝒕 = 𝟔𝟒𝟗. 𝟗𝟑𝟎𝟎𝟔𝟐𝟏 𝑳𝒐𝒂 − 𝟏𝟔𝟑𝟎. 𝟎𝟖𝟒𝟏𝟎𝟓 𝑾𝒐𝒂 + 𝟏𝟓𝟕𝟔. 𝟖𝟖𝟑𝟐𝟖𝟓 𝑯𝒐𝒂 +


𝟏𝟎𝟖. 𝟖𝟐𝟐𝟖𝟖𝟏𝟑 𝑾𝒃 + 𝟐𝟗𝟕. 𝟏𝟒𝟏𝟎𝟒𝟑𝟗 𝑻𝒇 + 𝟏𝟏𝟕𝟕. 𝟎𝟓𝟓𝟑𝟒𝟔 𝑻𝒓 −
𝟑𝟒𝟗𝟏. 𝟐𝟕𝟗𝟒𝟒 (EQ. 3.1)
The following FWD LCG estimation relationship (still called a “WER”) produces
1.052 m (41.4 in) for the LCG distance aft of the front axle line when the appropriate values
(in meters) from Table 3.2 are “plugged into” the equation 84:

𝑳𝒄𝒈 = 𝟎. 𝟎𝟔𝟖𝟑𝟐𝟗𝟑𝟕𝟔 𝑾𝒃 + 𝟐. 𝟒𝟎𝟕𝟏𝟕𝟖𝟎𝟖𝟗 𝑻𝒇 − 𝟏. 𝟒𝟏𝟓𝟒𝟗𝟒𝟗𝟒𝟐 𝑻𝒓 −


𝟎. 𝟑𝟒𝟎𝟏𝟕𝟏𝟑𝟓𝟖 𝑳𝒐𝒂 − 𝟎. 𝟕𝟓𝟎𝟗𝟐𝟎𝟎𝟕𝟓 𝑯𝒐𝒂 + 𝟏. 𝟓𝟎𝟎𝟏𝟑𝟑𝟒𝟓𝟏 𝑾𝒐𝒂 −
𝟎. 𝟔𝟗𝟒𝟔𝟓𝟒𝟗𝟎𝟒 (EQ. 3.2)
The WER determined values for curb weight and LCG are off by about -8.2% and
+6.5% respectively, which is a level of inaccuracy greater than expected of Equation 3.1
(+4%/-7%) and Equation 3.2 (+2.6%/-2.1%). The possibility suggests itself that the difference
between the estimated and quoted figures may be mainly due to the difference in the energy
storage type and the locations thereof, which is a hypothesis the validity of which may be
investigated via a simple weight and balance calculation:

Table 3.3 – WEIGHT & BALANCE “TEST” OF WEIGHT & LCG HYPOTHESIS

The hypothesis that the difference between the estimated and quoted figures may be
mainly due to the difference in energy storage type/locations would seem to be disproved. No
amount of reasonable “tweaking” of the weight and balance calculation resulted in a perfect
(or even good) match between the known Prius C mass properties and the WER estimates.
The only way the WER results could be made to match the known weight and balance
figures was to introduce an arbitrary 85 “delta weight” at a forward location which is outside

84
Reference [88], pg. 34.
85
This 99.1 lb (44.9 kg) weight is “arbitrary” by virtue of having no specific physical cause assignable to it.

33
SAWE Paper No. 3602
Category Number 31.0

of the body envelope 86. Essentially this just bears out the advisability of performing a
regression analysis for each distinct vehicle type in order to obtain an accurate WER
applicable to only that type, which is a point repeatedly emphasized in the source material
from which Equations 3.1 and 3.2 were drawn.

What is clear from this examination of the Prius C, other than the need for HEV WER
development (not to mention EV/BEV/NEV/PEV, PHEV, etc.), is that the mass
distribution of such an HEV is not totally unlike conventional FWD vehicles, but seems to
be naturally heavier and with an even more pronounced forward bias in the LCG 87. This
result may seem to be the opposite of what one would expect, at least with regard to weight,
because the FWD conventional vehicle WER’s as used herein were based on 1985-1995
technology, and present day vehicles are much more thoroughly optimized with respect to
weight. However, this weight optimization tends to get offset by “feature proliferation”
(e.g.: power windows, power door locks, heated 6-way adjustable power seats, 12-speaker
audio system, cruise control, tilt/telescopic adjustable steering, GPS navigation system,
TPMS, ABS, traction control, stability augmentation, climate control, air bags, remote
keyless entry, etc., etc.).

While present day weight optimization can be totally cancelled out by “feature
proliferation”, in the Prius C’s case it got overwhelmed to the tune of about 212 lb (96.2 kg)
by the additional weight penalty of HEV design. The HEV weight penalty has been, and
will continue to be, reduced with time (which is something that would have been much
appreciated by Ferdinand Porsche), but the penalty is still significant. Even more detrimental
than the weight penalty is the possible LCG penalty; the forward bias in LCG of
conventional FWD designs was detrimental to many aspects of automotive performance
(acceleration, braking, maneuver), and many present HEV FWD designs tend to be even
worse.

The mass distribution penalties inflicted by HEV design can be largely overcome by
pure electric designs, although at the present state of the technology weight and other
penalties (e.g.: range) become even more prominent. As of this writing there are a number of
modern EV (a.k.a BEV or PEV) designs that are operational, and some that have even
reached mass production (e.g.: Tesla). One of the most interesting and unusual BEV designs
falls within the former category: the experimental “Eliica” (Electric Lithium-Ion Car) as
developed by Prof. Hiroshi Shimizu and his team at Keio University of Tokyo, Japan.

86
The longitudinal location of 29.1 in (73.9 cm) forward of the front axle line is right at the leading edge of the
vehicle bumper.
87
The 2013 Volkswagen Jetta HEV represents a design similar to the 2012 Toyota Prius C, except that there is
more of an emphasis on “performance”, mainly acceleration (0-60 mph in 8.6 sec versus 11.9 sec). This
emphasis results in a more robust ICE under the hood, and consequently an even more extreme forward weight
distribution bias (rumored to be around 70% fwd/30% rr).

34
SAWE Paper No. 3602
Category Number 31.0

The Eliica represents a subsequent development of principles established by the prior


creation of an even larger and more massive experimental vehicle, the “KAZ” (Kieo
Advanced Zero-emissions) of 2002 88. The Eliica first made public appearance in 2004, and
soon after (2005) became a pair of very closely related vehicles, each optimized for a certain
aspect of automotive performance; Prototype Number 1 (“Speed”) model, and Prototype
Number 2 (“Acceleration”) model. The “Speed” model is essentially just a LSR (land speed
record) variation on the more practical (“everyday use”) vehicle termed the “Acceleration”
model, differing only in the electrical capabilities of the battery packs and in the final drive
gear ratio (3.757 vs. 6.923 respectively).

Figure 3.1 – 2004 KEIO UNIVERSITY ELIICA

Regardless of model type, the Eliica is still a relatively large and massive vehicle with
a length overall of 200.8 in (5100 mm) and a curb weight over 5,300 lb (2,400 kg); the
wheel/axle layout is reminiscent of the Reeves Octoauto, but that vehicle was longer at about
240 in (6096 mm) and heavier at about 6,500 lb (2,943 kg). With respect to size, the Eliica is
much more like a large United States sedan of 30 to 50 years ago, but significantly heavier 89.
The Eliica can carry up to five passengers (including driver), but is best suited for the
transport of four passengers. The wheels appear to be 16 in, for which the tire designation is
likely to be around P188/55R16 (KAZ spec). The wheels are attached via “independent”90
double-wishbone linkages to an aluminum platform chassis of about 6.5 in (165 mm) in
height.

88
The 2002 KAZ had a LOA of 263.78 in (6,700 mm) and a weight of about 6,570 lb (2,980 kg). In this respect
it is an almost perfect match for the 1910 Reeves Octoauto at 240 in (6,096 mm) long and a weight of about
6,500 lb (2,943 kg).
89
Compared to early 1980’s full size United States sedans the Eliica is about a foot (366 mm) shorter and about
1,500 lb (680 kg) heavier.
90
There is a hydraulic interconnection between each tandem wheel pair (pitch control), and transversely between
pairs (roll control), in order to “improve ride and cornering”.

35
SAWE Paper No. 3602
Category Number 31.0

Speaking of height, the Eliica HOA was originally given as 55.7 in (1414 mm), but
the extra support points of eight wheels and the flat underside allowed for a 2.0 in (49 mm)
further reduction in ground clearance, so the current official HOA is 53.7 in (1365 mm).
Since the suspension is highly variable with regard to ride height; the HOA, VCG, and
ground clearance are all a matter of the paramètre du jour.

The platform chassis is essentially a box structure, containing most of the propulsion
system components, similar to the long established “skateboard” concept 91. On top of the
aluminum platform structure a tubular steel superstructure is placed; this adds bending and
torsional stiffness, and provides support for the FRP (fiberglass) body which then encloses
all. The build sequence utilizing these three major structural components may be illustrated as
follows in Figure 3.2:

Figure 3.2 – ELIICA PLATFORM STRUCTURE + SUPERSTRUCTURE + BODY

The “skateboard” concept is notable for a claimed reduction in VCG (vertical center
of gravity), and much hoopla is made of this by the Eliica designers; curiously the Eliica
VCG was not quoted in the available literature, but is probably only 3-4 in (76-102 mm)
above the axle lines. This means it would be around 15 in (394 mm) above the ground plane,

91
Ford’s version of the “skateboard” concept is termed the “HySeries” which was featured in the 2007
“Airstream” concept car (PHEV, hydrogen fuel cell); GM’s “Hy-wire” version was introduced on the 2002
“Autonomy” concept car (HEV, hydrogen fuel cell).

36
SAWE Paper No. 3602
Category Number 31.0

or more than ½ ft (152 mm) below the point where a conventional ICE design of similar
proportions would place it 92:

Table 3.4 – CONVENTIONAL ICE LARGE SEDANS vs. ELIICA

“Skateboard” designs are also generally noted for a lack of LCG bias; the freedom of
mass placement inherent in an electrically propelled vehicle is utilized in the Eliica, as made
evident by the following plan view layout 93, to obtain a claimed near 50%/50% mass
distribution front/rear (which would place the LCG about 73.3 in, or 1862 mm, aft of the
first axle):

Figure 3.3 – ELIICA PROPULSION LAYOUT IN PLAN VIEW

However, with regard to the Eliica’s mass properties, not all aspects of the design are
beneficial; for reasons of space utilization the Eliica’s motors/gearboxes/inverters were
“essentially” (meaning not really within the wheel as much as hanging off the wheel hub,
jutting inboard) placed as hub motors, thereby adding at least 95 lb (43kg) to each unsprung

92
Reference [88], pp. 21 & 35.
93
Reference [39], pg. 48.

37
SAWE Paper No. 3602
Category Number 31.0

weight, which must adversely affect ride and road holding. To support this and other
statements made concerning the Eliica’s mass properties for which no direct quote is
obtainable, this author conducted the following weight and balance study of the Eliica:

38
SAWE Paper No. 3602
Category Number 31.0

Table 3.5 – ELIICA WEIGHT & BALANCE STUDY


39
SAWE Paper No. 3602
Category Number 31.0

In light of the Eliica’s AWD (all wheel drive), and common tire size all around, the
estimated low VCG of 15.4 in (391 mm) 94 and mid-point LCG of 74.2 in (1885 mm) should
greatly facilitate acceleration/braking/maneuver performance by rendering the tire-road
contact pressure distribution more uniform than otherwise would be the case. This would tend
to be substantiated by the acceleration performance quoted for the “Type 2” model Eliica 95:

0-100 kph (0-62.1 mph)________4.11 seconds

0-160 kph (0-99.4 mph)________7.04 seconds

Terminal velocity_____________190 kph (118.1 mph)

This acceleration performance when graphically displayed (alongside the performance


of a Porsche 911 Turbo) reveals itself to be of a very smooth and steady nature, free of the
“jerk” or inflection points typical of an internal combustion engine/sliding gear
transmission 96:

Figure 3.4 – ACCELERATION: ELIICA vs. PORSCHE 911 TURBO

94
To provide perspective, the 2013 Scion FR-S has a VCG of 18.1 in (459.7 mm). The Scion VCG is less than 1
in (25.4 mm) higher than the contemporary Lexus LFA, Porsche 911 GT3, and the Ferrari 360 (per Reference
[36]).
95
Reference [39], pg. 50.
96
Reference [95], pg. 7.

40
SAWE Paper No. 3602
Category Number 31.0

As noted, the Eliica wheel/axle layout is reminiscent of the Reeves Octoauto, so it


should not be surprising that the steering layout is also similar; both vehicles comply with the
“standard” steering geometry for 8-wheel/4-axle vehicles as depicted in Figure 2.2. There
must have been some concern during the design process that the Eliica might prove not be
able to turn within an acceptable radius because it is reported that some thought was given to
having the fourth axle “counter-steer”, like the rear end of an old “hook & ladder” fire
engine. However, the final steering layout is given as per the following figure (note the mid-
point location of the LCG), which shows the use of only Figure 2.2 type steering 97:

Figure 3.4 – ELIICA STEERING INTO A LEFT TURN MANEUVER

The point of this chapter was to emphasize that the choice of propulsion system, and
its orientation/location, interacts with the axle/wheel layout to greatly affect the mass
properties of a vehicle. When the propulsion system is ICE of a particular conventional
orientation/location (longitudinal front-engine, traverse mid-engine, etc.), and is constrained
to operate within a “standard” four wheel/two axle configuration, then there is a general
predictability of the vehicle mass properties, as per Table 3.1. The fact that such automotive
paradigms have a long history results in the existence of many examples of each type, which
in turn makes possible the formulation of reasonably accurate WERs, such as Equation 3.1
and 3.2, by regression analysis.

97
Reference [39], pg. 48.

41
SAWE Paper No. 3602
Category Number 31.0

However, as the course of automotive evolution leads further and further away from
the established propulsion and configuration norms, the applicability of the established WERs
based on the old paradigms diminishes. If an HEV is of a configuration reasonably akin to
those of the past, with the ICE merely “assisted” by the motor to enhance fuel economy
and/or performance, then established WERs might be pressed into service with some ad hoc
modification of the results to suit. However, as was demonstrated in the case of the Prius C,
any results thus obtained are not likely to be very accurate; it is always advisable to develop
new WERs whenever possible when dealing with “new” vehicle types.

If the case is one of modern purely electric vehicles, with no ICE involved, then the
break with past conventions is great enough to render established WERs irrelevant. The mass
properties of pure electric vehicles have to be estimated by a “bottom up” determination,
such as Table 3.5, until enough data exists regarding specific examples to allow for WER
formulation by regression analysis. Pure electric vehicles show great promise with regard to
an optimization of the CG location due to the flexibility of component arrangement, a
situation that can be referred to as “distributed mass”, and also with regard to drive
configurations that some have referred to as “distributed drive” 98. The characteristics of
“distributed mass” and “distributed drive”, working in synergy with each other, should
give EV’s extremely significant advantages over conventional vehicles.

However, electric vehicles have yet to sufficiently overcome the interrelated obstacles
of weight penalty, limited range, and recharging time. The promise, and the drawbacks, of
present pure electrical vehicle design is well illustrated by the Eliica; it has generally fine
performance due to the superior rpm/torque behavior characteristic of electric motors, coupled
with the advantages of distributed mass and distributed drive. However, the
propulsion/drive choice of eight full size wheels inflicted a considerable effective weight
penalty on the design, not to mention incurring an unsprung weight penalty which was
exacerbated by the hub motor location.

With appropriate emphasis on mass properties design, it should be possible to do


even better.

98
Reference [45], pp. 122-125.

42
SAWE Paper No. 3602
Category Number 31.0
Rev C

4 – AN ADVANCED MASS PROPERTIES DESIGN CONCEPT

In accord with the idea advanced in Chapter 2 of improving automotive performance


by reduction of effective mass, consider a conversion of the conventional four wheel/two
axle baseline vehicle configuration to an eight wheel/four axle configuration (a.k.a. “advanced
design concept”), wherein each of the eight wheels is one half the radius of the original four
wheels, with everything else kept as constant as possible. Since the load on each one of the
eight wheels is half the load on each one of the original four wheels, it may not matter that the
smaller tire will have a lower load capacity rating and, as the matter of load rating is closely
intertwined with that of speed rating, that the smaller wheels would have to rotate twice as
fast for the same vehicle speed. A thorough investigation of such implications of this
tire/wheel size reduction will be conducted immediately following; for now consider the
implications of Figure 4.1 99:

Figure 4.1 – BASELINE CONFIGURATION vs. “ADVANCED CONFIGURATION”

99
One of those implications is “static indeterminacy” for the new configuration. While it’s easy to determine the
reactions at the front and rear tire/road contact points for the baseline configuration (front/rear axle weight
distribution is 50.5%/49.5% for the 3640 lb/1651 kg condition) utilizing a static summation of the forces and
moments, doubling the number of reaction points requires the development of at least two more independent
equations for solution. While this can be done, for the most part an even distribution of static reaction forces will
be assumed throughout the initial part of this paper for convenience; this should not appreciably affect the
validity of any conclusions as the model will become more realistic as the concept is developed in this paper.

43
SAWE Paper No. 3602
Category Number 31.0

Note that, as depicted, the placement of the smaller wheel/tire sets is evenly spaced
longitudinally and pushed out to the maximum extent allowed by the existing vehicle
envelope, instead of being confined to within the original wheelbase. The small size of the
new wheel/tire sets allows for such “pushing out”, which improves the “stance” of the
vehicle 100. Note also that no spare “half-radius” wheel/tire is indicated in lieu of the original
“full-size” spare; with eight wheels and tires in use the need to carry such a spare wheel/tire is
questionable (although it does turn out the small tires are more highly loaded with respect to
their capacity).
The weight reduction associated with the possible elimination of the spare tire/wheel
is just the beginning of the mass property changes associated with the new configuration;
consider the mass properties of the “wheels” (tires and wheels) themselves (excluding, for
the moment, shafts, brakes, and all else that rotates with the wheel). To simplify the matter,
the “wheels” will be modeled as right circular cylinders, solid & homogeneous, for a weight
(“W”) and rolling inertia (“I”) comparison 101:

W = δ π t R2 (EQ. 4.1)

I = W R2 / 2 (EQ. 4.2)

Where:

δ = Tire/wheel density (lb/in3).

t = Tire/wheel width (in).

W = Tire/wheel weight (lb).

R = Baseline rolling radius (in).

r = Advanced configuration radius (in).

I = Tire/wheel rolling inertia (lb-in2).

Figure 4.2 – WHEEL MODEL

100
Reference [53], pg. 580, a Maurice Olley letter of 15 February 1959 notes “The greater the number of
independent rows of suspension the less the pitch stability.”. Earlier he alleges a six-wheel/three-row
suspension will have only two-thirds the pitch stability of a conventional four-wheel/two-row configuration. The
enlarged stance will not only provide general performance benefits, but run counter any such “instability”.
101
Reference [98], pg. 4.2.3. A “Right Circular Cylinder” model can’t possibly give accurate values for
tire/wheel weight and inertia, but is perfectly valid for a general, relativistic analysis.

44
SAWE Paper No. 3602
Category Number 31.0

So, for the four “full-size” wheels the weight (“W4”) and rolling inertia (“I4”) can be
expressed as:

W4 ∝ δ π R2 t

I4 ∝ W4 R2 / 2 = δ π R4 t / 2

And for the eight “half-radius” wheels the weight (“W8”) and rolling inertia (“I8”)
can be expressed as:

W8 ∝ δ π r2 t = δ π (R/2)2 t = δ π R2 t / 4π

I8 ∝ W8 r2 / 2 = δ π R2 t / 4 (R/2)2/2 = δ π R4 t / 32

Therefore, the weight and inertia of one “half-radius” wheel compared to one “full-
size” wheel is:

W8 / W4 = (δ π R2 t / 4) / δ π R2 t = 1/4

I8 / I4 = (δ π R4 t / 32) / (δ π R4 t / 2) = 1/16

Since the new configuration has eight wheels to the conventional four, the total
vehicle wheel weight and inertia comparisons are:

8W8 / 4W4 = (8δ π R2 t / 4) / 4δ π R2 t = 1/2

8I8 / 4I4 = (8δ π R4 t / 32) / (4δ π R4 t / 2) = 1/8

The indication is that having “wheels” one half the radius but twice the quantity would
reduce total vehicle “wheel” weight (not counting a possible spare tire/wheel deletion) by
about 50%, unsprung weight by about 25% 102, and “wheel” rolling inertia by about 87%;
this would suggest that change to the new “half-radius, twice the quantity” configuration
could lead to significant improvements in performance (acceleration/braking, lateral
acceleration, ride, and fuel economy).
Obviously this trend could be taken to a region of unprofitable extremes; if eight
“half-radius” wheels are better than the conventional four, then why not sixteen “quarter-
radius” wheels, et cetera. Given this progression, there soon would be a point of diminishing
returns, and ultimately at some point a decided reversal of benefit, but it’s difficult to say with
certainty where that point might be on this basis alone (although “eight” is possibly very near
that limit). So much depends on the details of the implementation of the new tire/wheel size,
and the practical consideration of tire availability trumps all else (reportedly it was the tire

102
Practical considerations would result in unsprung weight reduction of 14.2%, which is still significant (see
“Appendix B – Weight Accounting”).

45
SAWE Paper No. 3602
Category Number 31.0

manufacturer’s reluctance to develop the design of the uniquely sized front tires for the Elf
Tyrrell 6-wheeler that doomed that concept to oblivion, and not any basic conceptual flaw).
The magnitude of the automotive effective mass constitutes the very heart of this
paper’s subject matter, and the Dunlop RS4 6.00×16 tire/Jaguar 5x16 54-spoke wire wheel
rolling inertias constitute the predominate portions of the “I1” and “I4” values which
contribute so much to the baseline effective mass. As noted in Chapter 2, the exact “I1” and
“I4” contributions for the 1958 Jaguar XK150S baseline configuration 103 are taken as per
Table 2.4, which indicates the individual mounted (including wheel) baseline tire weight was
48.9 lb (22.2 kg) and rotational inertia was 40.6 lb-ft2 (1.71 kg-m2); these values were
obtained by a detailed “hand calc” estimation in 1984 104. Since the time of that estimation a
paper was published (1990) that presented some general rotational inertia estimation
formulae derived from regression analysis of empirical tire/wheel data; for just the deflated
tire the pertinent equation is given as 105:

𝑰𝒀𝒀 = 𝟏. 𝟓𝟑𝟕 × 𝟏𝟎−𝟐 𝑾 𝒅 − 𝟏. 𝟑𝟓𝟒 (EQ. 4.3)


And, for the inflated tire mounted on (i.e., including) a pressed steel wheel, the inertia
is calculated per 106:

𝑰𝒀𝒀 = 𝟑. 𝟎𝟑 × 𝟏𝟎−𝟒 𝑾 𝒅𝟐 + 𝟎. 𝟖𝟖𝟔 (EQ. 4.4)


Where, for both formulae, the symbolism is:

IYY = Tire and/or tire/wheel rotational inertia (lb-in-sec2)

W = Tire and/or tire/wheel weight (lb)

d = Tire deflated or inflated (no load) outer diameter (in)

Utilizing Equation 4.3 to estimate the rotational inertia of the baseline Dunlop RS4
6.00×16 tire (with inner tube), and Equation 4.4 for the tire/wheel (1958 Jaguar 16 in “center-
lock” 54-wire spoke) assembly, the following results are obtained:

𝑰𝒀𝒀 = 𝟏. 𝟓𝟑𝟕 × 𝟏𝟎−𝟐 𝟐𝟕. 𝟎 (𝟐𝟕. 𝟎𝟎) − 𝟏. 𝟑𝟓𝟒 = 𝟗. 𝟖𝟓 𝒍𝒃 − 𝒊𝒏 − 𝒔𝒆𝒄𝟐


= 𝟐𝟔. 𝟒𝟏 𝒍𝒃 − 𝒇𝒕𝟐 𝒐𝒓 𝟏. 𝟏𝟏 𝒌𝒈 − 𝒎𝟐

103
The Dunlop RS4, mounted on a 5½×16 pressed steel or a 5×16 “center-lock” wire wheel, was how the 1958
Jaguar XK150S was originally equipped. Since the technical data suite for this vehicle (and its tires) is fairly
complete it was chosen as the baseline vehicle to represent the “conventional” configuration. A modern baseline
might seem more relevant, but probably wouldn’t add much, if anything, to an investigation that is concerned
with relative change.
104
Reference [89], pg. 43.
105
Reference [50], pg. 4.
106
Ibid.

46
SAWE Paper No. 3602
Category Number 31.0

𝑰𝒀𝒀 = 𝟑. 𝟎𝟑 × 𝟏𝟎−𝟒 𝟒𝟖. 𝟗 (𝟐𝟕. 𝟎𝟒)𝟐 + 𝟎. 𝟖𝟖𝟔 = 𝟏𝟏. 𝟕𝟐 𝒍𝒃 − 𝒊𝒏 − 𝒔𝒆𝒄𝟐


= 𝟑𝟏. 𝟒𝟐 𝒍𝒃 − 𝒇𝒕𝟐 𝒐𝒓 𝟏. 𝟑𝟐 𝒌𝒈 − 𝒎𝟐

The tire result is only 73% of the 1984 tire estimate (36.2 lb-ft2, 1.52 kg-m2), while
the combined tire/wheel result is only 77% of the 1984 estimate (40.6 lb-ft2, 1.70 kg-m2);
however, the original estimates may be the more accurate. The regression analysis which
produced Equations 4.3 and 4.4 was performed on data representing modern tubeless P-metric
size radial tires, and not on obsolete Imperial size tube-type bias tires such as the Dunlop
6.00x16 RS4’s found on the 1958 Jaguar XK150S. The old Jaguar tire/wheel units differed
significantly from their modern counterparts, especially in having extra major components
such as an inner tube and “spinner” (the retainer for the Jaguar 16x5 54-spoke “center-lock”
wire wheel, which is a very structurally inefficient wheel). For this reason the original 1984
inertia estimates may be more accurate than the estimates obtained by use of Equations 4.3
and 4.4, and in any case this paper only deals with relative change. However, for all
estimation involving modern tire/wheel assemblies Equations 4.3 and 4.4 constitute the
preferred method, unless a detailed calculation or actual measurement is utilized.

Such a reduction in tire size as contemplated affects many tire characteristics other
than weight and rolling inertia, such as the vertical spring constant. Determining the effect
on the tire vertical spring constant mandates recourse to Rhyne’s Equation, which in its most
useful form (requiring as input only those parameter values readily found posted on a tire
sidewall) is 107:

𝑺𝑵 ×
𝑲𝒁 = 𝟎. 𝟎𝟎𝟎𝟐𝟖 𝑷𝒊 ��−𝟎. 𝟎𝟎𝟒 + 𝟏. 𝟎𝟑�𝑺𝑵 × � 𝟓𝟎
+ 𝑫𝑹 � + 𝟑. 𝟒𝟓 (EQ. 4.5)

Where:

KZ = Tire vertical stiffness (kg/mm).

Pi = Tire inflation pressure (kPa).

= Tire section aspect ratio (dimensionless).

SN = Tire nominal section width (mm).

DR = Wheel rim nominal diameter (mm).

107
Reference [26], pg. 194. While Eq. 4.5 is useful, it may not be the most accurate version of Rhyne’s
Equation, which would likely be its original form as shown on page 192 of the source, wherein only measured
data is used as input.

47
SAWE Paper No. 3602
Category Number 31.0
Rev C

The closest corresponding modern designation for the 6.00×16 baseline configuration
tires is 152/92R16. The best corresponding “half-radius” designation for the tires of the
“advanced” configuration is possibly 152/46R8. Per Rhyne’s Equation the vertical stiffness of
the two different size tires at various inflation pressures (within the “full-size” tire operating
range) is tabulated as follows 108:

Table 4.1 – TIRE VERTICAL STIFFNESS PER RHYNE’S FORMULA

That the “half-radius” tires may prove to be about 24.3% less stiff vertically than the
original tires at the same operating pressures suggests that the smaller tires will not present a
problem with regard to road induced high frequency vibration. The tire damping coefficient is
another significant factor in this regard, but this is traditionally neglected at the conceptual
stage of design 109. But, even without any exhaustive study, the short sidewall/low aspect ratio
of the half-size tires suggests the attainment of a more precise and responsive steering,
although achieved with a possible greater risk of tire and wheel damage upon sharp impacts,
and most certainly a greater risk of damage in curbside/sidewall encounters.

What the calculated stiffness values also provide is a means for making tire-to-road
contact area comparisons between the old and new configurations. This is very important, as

108
Rhyne’s Equation has the metric (mm) version of the “Michelin Formula” (the “-0.004 +1.03)SN” term)
embedded within it; that embedded formula was derived through a regression analysis of a large number of
conventional passenger car tires. As the “half-radius” tire is far from conventionally sized, a “brute force”
technique had to be employed wherein the term in question was replaced with the value “100.6” (mm) in order to
obtain reasonable “half-radius” tire vertical stiffness values.
109
Reference [93] deals with tire vertical stiffness and damping on pages 71-81 and gives some typical values;
note that rolling tires generally present lower values for stiffness (10-15%) and damping (50-70%) than non-
rolling. For automotive “simulations…the use of the rolling dynamic…(values)…is preferred”.

48
SAWE Paper No. 3602
Category Number 31.0
Rev C

the magnitude of the tire traction forces that can be generated by any automotive
configuration is determined at its most basic level by the amount of “rubber on the road”, and
acceleration levels in the longitudinal and lateral directions (acceleration/braking and
maneuver) depend on those forces.

Prof. Dixon presents a formula for estimating the tire/ground plane contact area as a
function of vertical stiffness, radius, and normal load in his book Suspension Geometry and
Computation. The following is essentially the same as that methodology; consider the
physical situation of tire under load:

Figure 4.3 – TIRE-ROAD CONTACT AREA vs. NORMAL LOAD

Under a “no load” condition the contact area between the tire and the ground plane is
essentially just a thin line of length equal to the width of tire tread times some indeterminate
thickness, but under some appreciable load the tire deflects vertically by some distance “d”
resulting in a contact area taken as roughly being proportional to the tread width “tw”
(assumed constant 110) times the length “Lc” which is taken as the contact portion of the
deformed arc length “S” (shown above in red) 111:

𝑨𝒄 = 𝑳𝒄 𝒕𝒘 = 𝟏. 𝟐𝟒 𝑹𝒊 𝑪𝒐𝒔−𝟏 [(𝑹𝒊 − 𝒅)/𝑹𝒊 ] 𝒕𝒘 (EQ. 4.6)

110
Reference [21], pg. 86. Taking the tread width of a modern automobile tire as constant under varying load is
a reasonable approximation, but for other types of tires (e.g.: motorcycle, aircraft) is totally incorrect.
111
Ibid, pp. 84-85. Using the full arc length as “Lc” would require a factor of “2”, but this author’s comparison
study of calculated areas with published empirical data resulted in the factor of “1.24” as a better value for use
within the working range of variation for passenger car tires. Pillai and Fielding-Russell, in their paper “Tire
Rolling Resistance From Whole-Tire Hysteresis Ratio” (Rubber Chemistry and Technology, May 1992), present
a similar formula for contact area: Ac = 1.85 d2/3 Ri1/3 tw. Upon investigation by this author the Pillai/Fielding-
Russell formula proved far less accurate than Eq. 4.6.

49
SAWE Paper No. 3602
Category Number 31.0
Rev C

Where:

Ac = Tire to ground plane gross contact area (in2)

Lc = Tire to ground contact area length (in).

tw = Tire tread width, assumed constant with load (in).

Ri= Tire no-load inflated radius (in).

d = Tire vertical deflection under load (in).

To use this equation requires calculation of the tire deflection “d” under normal load
“N” (lb) 112:

𝒅 = 𝑵�𝑲 + 𝒅𝟎 (EQ. 4.7)


𝒁

And also requires the width of the tire tread “tw” which, when lacking a measured
value, can be approximated per the formula 113:

𝒕𝒘 = 𝟎. 𝟎𝟑𝟗𝟑𝟕 (−𝟎. 𝟎𝟎𝟒 + 𝟏. 𝟎𝟑)𝑺𝑵 (EQ. 4.8)


Where for these last two equations:

KZ = Tire vertical stiffness (lb/in).

tw = Tire tread width, assumed constant with load (in).

= Tire section aspect ratio (dimensionless).

SN = Tire nominal section width (mm).

N = The normal load on the tire (lb).

d0 = Tire deflection function “y-intercept” value (in).

Although reasonable, Equation 4.6 has proven to not be very accurate in determination
of gross tire contact area values; tire deformation is more complex than the simple
visualization of Figure 4.3 suggests, but at this early level of design the accuracy is adequate.

112
Reference [72], pg. 6. This source attributes the equation to Nokian Tyres Ltd, but the equation far predates
that company’s founding in 1988.
113
Reference [26], pg. 193. This equation has been referred to as the “Michelin Formula”. It was derived by use
of a large sample of various passenger car tires, and thus is not reliable for tires with characteristics greatly
outside the range of that sample.

50
SAWE Paper No. 3602
Category Number 31.0
Rev C

Use of Equations 4.6, 4.7, and 4.8 in conjunction with the vertical stiffness data of
Table 4.1 produces the following table of tire to ground contact area 114. The “d0” values were
obtained by assuming a value of 0.2416 in. for the “full size” (0.1208 in. for the “half rad”) at
40 psi and then proportioning via the stiffness ratios. It should also be noted that the “Ri”
values in Table 4.2 where determined by taking “Ri”, at 40 psi inflation pressure, as being
equal to “DR + 2 SH”; with that as the reference all other “Ri” values were determined by use
of Equation F.32 (see Appendix F for explanation):

𝒅𝑹𝒊𝒏𝒇𝒍𝒂𝒕𝒆 = (𝑷𝟐 − 𝑷𝟏 ) 𝑹𝒊 𝟐 / 𝒕𝒕 𝑬

Table 4.2 – TIRE DEFLECTION & CONTACT AREA vs. INFLATION PRESSURE

Note that the “half-radius” tires will have about 44.3% less individual contact area
than the baseline “full-size” tires, but because there are twice as many tires on the concept
vehicle there will be a total of around 9.3% more tire/road contact area per vehicle 115. This
should translate into a proportional increase for maximum steady-state lateral acceleration,
and into a similar improvement for braking, but with possibly a decrease in initial acceleration
rate due to a loss of driving tire contact area (RWD being held constant). This acceleration
deficient could be countered by establishing the two rearmost axles of the “advanced concept”
as driven, but that would incur a weight penalty.

114
The calculated ground contact areas are not very accurate, and are “gross” areas as no attempt has been made
to account for the affect of tread patterns (“land/sea ratio”) which can reduce areas by up to about -40% to
achieve net figures. The total amount of contact area upon which an automobile depends for performance is
surprisingly small, as was demonstrated on page 72 of Reference [68]; a Michelin Pilot Sport Cup 245/35ZR19
inflated to 32 psi (221 kPa) under 606 lb (275 kg) load at a camber angle of 1.5 degrees had a net contact area of
only 13.3 in2 (85.8 cm2).
115
A simple line-by-line comparison of the gross vehicle tire-road contact areas of Table 4.2 indicates an
increase of +10.31% to +12.12%. However, the “9.3%” results from the “baseline” inflation pressure mix of 40
psi (275.79 kPa) front and 45 psi (310.26 kPa) rear, with the corresponding “advanced concept” inflation
pressure condition being 45 psi (310.26 kPa) all around.

51
SAWE Paper No. 3602
Category Number 31.0
Rev C

Figure 4.4 – “FULL-SIZE” vs. “HALF-RADIUS” TIRE/WHEEL


One immediate mass properties related consideration resulting from the contemplated
configuration change is rolling resistance; with more rubber in contact with the ground the
initial suspicion is that the new configuration will have more rolling resistance. The rolling
resistance force (“FR”) of a tire varies in accord with a resistance coefficient (“CR”) and the
normal load (“N”), where “CR” depends only on the road surface (historically resulting in
tables or charts of values ranging, say, from 0.04 paved to 0.40 deep sand):

FR = CR N (EQ. 4.9)
Although such a simple relationship has been used for crude calculations valid
(roughly) only for very low speed ranges; the rolling resistance (single coefficient) actually
always varies with speed as illustrated in Figure 4.5a 116:

Figure 4.5a – ROLLING RESISTANCE COEFFICIENT “CR” vs. VELOCITY

116
Reference [66], pg.122. The steel test drum coefficients appear to be “×100” and need to be multiplied by
“1.20” for asphalt and “1.30” for concrete.

52
SAWE Paper No. 3602
Category Number 31.0
Rev C

The point is that the rolling resistance coefficient is really more complex than just a
simple single value; the coefficient “CR” has both static (“CSR”) and dynamic (“CDR”)
components as per an equation that includes the vehicle velocity (“V”) in mph units and was
developed at the Institute of Technology in Stuttgart (c. 1938) 117:

CR = CSR + 3.24 CDR (V/100)2.5 (EQ. 4.10)

These two coefficient components are particular to specific tire type and also vary in a
known fashion with tire inflation pressure 118:

Figure 4.5b – STATIC & DYNAMIC COMPONENTS vs. INFLATION PRESSURE

This equation, when used for the calculation of the total vehicle tire rolling resistance
(hence “W” for vehicle weight replaces “N”) in a simplified situation wherein the normal
loads and the tire characteristics are all equal, becomes 119:

FR = CR W = CSR W + 3.24 CDR (V/100)2.5 W (EQ. 4.11)

117
Reference [81], pg. 34.
118
Reference [28], pg. 118. Figure 4.5b represents a re-plot of the “pick-off” data from the plot in this reference.
119
A slightly less simplified variation is “FR = CR (1 + V/100) W” which was a supposedly reasonable
approximation for use up to 80 mph (129 kph). Such equations were developed mainly for use with tables of
general roll resistance coefficients appropriate for passenger car bias ply tires inflated to 32-40 psi (221-276
kPa); various value sets/tables exist which depend on the type of vehicle/road surface considered, plus the
general level of tire technology (i.e., the point in time) in effect when the particular data was formulated, so care
must be exercised.

53
SAWE Paper No. 3602
Category Number 31.0
Rev C
Still, this relationship doesn’t give the necessary insight into the effect of vehicle
configuration changes such as that being considered; the “half-radius” tire “CSR” and “CDR”
values are unknown at this point. For the necessary insight into the effect on vehicle rolling
resistance of the configuration change considered, reference must be made to a relationship
developed at the University of Michigan by Prof. Samuel K. Clark (et al) of the Department
of Applied Mechanics and Engineering Science (c. 1974) 120:

𝑵 𝑺𝑯
𝑪𝑹 = 𝑪 �𝑺
𝑫𝒊 �
(EQ. 4.12)
𝑵

Where:
CR = Tire coefficient of rolling resistance (lb resistance/lb weight).
C = Tire material coefficient (in/lb).
Di = Tire “no-load” (inflated) outer diameter (in).
SH = Tire section height (in).
N = Normal load on tire (lb).
SN = Tire nominal section width (in).
SH/SN = Tire section aspect ratio ( ).

This means that, in general, the greater the tire diameter and/or the lower the aspect
ratio (decreasing “SH”) then the less the rolling resistance. Taborek, in his book Mechanics of
Vehicles (1957), presented a graph showing the general effect of increasing diameter on
pneumatic tire (bias-ply) rolling resistance coefficient (also indicates effect of road
surface) 121:

120
Reference [28], pg. 116. Note that this equation shows the roll resistance coefficient itself to vary directly
with the normal load (weight) borne by the tire. Pillai and Fielding-Russell present the equation: “CR = h d tw /
Ac” , where “h” is the “whole tire hysteresis ratio” which typically varies from 0.080 to 0.060. They also utilize
a “experimental…minitire” in their analysis designated as the HR78-8 (!). However, there is no theoretical
method given to determine “h” (although one could possibly be worked out), and no useful information given
regarding the HR78-8 tire (lack of a given nominal cross-section width is critical).
121
Reference [81], pg. 31. Reference [93], pg. 16. Wheels shod with pneumatic tires behave very differently
from rigid wheels; a tire/wheel can present much less rolling resistance than a rigid wheel of the same diameter,
and is much less sensitive to diameter variation, as Robert William Thomson (1822-1873) demonstrated long
ago (John Boyd Dunlop patented the pneumatic tire in 1888, but that patent was later invalidated in favor of
Thomson’s earlier patents such as his 1847 United States Patent Number 5104).

54
SAWE Paper No. 3602
Category Number 31.0
Rev C

Figure 4.6 – TIRE DIAMETER vs. ROLLING RESISTANCE COEFFICIENT (1957)

On the basis of this chart, using the inflated tire diameters from Figure 4.4, it is
projected (extrapolated!) that the rolling resistance per tire will increase by +103%! However,
this is assuming that, other than diameters, all else is equal, which is not the case… 122
For a “common sense” estimate, consider that the rolling resistance coefficient “CR”
should increase with the amount of rubber deformed at any one time, and with the number of
deformations occurring per unit time. At 40 psi (275.8 kPa) a “half-radius” tire of the new
configuration will have a gross road contact area of 13.82 in2 (89.17 cm2) versus the “full-
size” tire’s 24.71 in2 (159.42 cm2) while deflecting only about 62% (“dhr/dfs”) as much, but
will have to roll about twice as fast (“Rrfs/Rrhr”), so it would seem reasonable to expect a
change in rolling resistance coefficient for the new tire on the order of:
%Change CR = [((13.82/24.71)×(0.5772/0.925)×(12.59/6.18))-1]×100 = -28.9%

This is a far cry from the previous “estimate” of +103%!! Using the more
sophisticated methodology of Eq. 4.12, with values drawn from Table 4.2, the relationship of
the unconventional to the conventional for one wheel is:

⎡ ⎤
𝑪𝑹𝒉𝒓 ⎢ 𝑪 𝟒𝟓𝟓 √𝟒𝟔 ⎥
%𝑪𝒉𝒈 𝑪𝑹 =� �𝑪 − 𝟏� × 𝟏𝟎𝟎 = ⎢ 𝟏𝟑. 𝟓𝟐 � 𝟗𝟏𝟎 − 𝟏⎥ × 𝟏𝟎𝟎 = −𝟐𝟗. 𝟑%
𝑹𝒇𝒔 ⎢ 𝑪 𝟐𝟕. 𝟎𝟒 √𝟗𝟐 ⎥
⎢ ⎥
⎣ ⎦

122
Reference [93] utilizes this same figure (originally from Reference [81]), but says “…the effect of tire
diameter on…rolling resistance is…negligible on hard surfaces…” on pages 14-16. However, this comment
is probably made concerning only that portion of Figure 4.6 that could be considered “normal” tire diameter
variation; the “half-radius” tire is well to the left of “normal”.

55
SAWE Paper No. 3602
Category Number 31.0
Rev C

This indicates a decrease of -29.3% in rolling resistance coefficient per wheel. Given
the previous “common sense” estimate of -28.9%, this seems reasonable 123. If “Cr” for the
“full-size” tire is taken as 0.0101871, then “Cr” for the “half-radius” tire is 0.00722266.
However, what does this mean with respect to “CSR” and “CDR”? If the “full-size” “CSR” and
“CDR” are taken as 0.0065020 and 0.0064340, then maybe the “half-radius” tire “CSR” and
“CDR” can be reasonably taken as 0.006342 and 0.001503 124. However, there is at least one
other significant caveat: the basic coefficient of roll resistance “C” was taken as the same for
both sizes of tire; this coefficient is dependent on the material and structural characteristics of
the tire. Therefore, there is an implicit assumption that both size tires would be exactly the
same with respect to material and construction (which is not likely, but also may not be
important for a comparison on such a general level).
It would seem likely that this is the best rolling resistance estimate that can be made
on a constant and evenly distributed vehicle weight basis. That the rolling resistance of the
new configuration could decrease by such a significant amount may seem incredible; tire
manufacturers spend fortunes just to decrease rolling resistance by a couple of percent, and
then often to the detriment of some other critical tire characteristic(s), such as traction.
However, this is just the beginning of what may be a revelation in automotive design as
weight, rotational inertia, and rolling resistance are only some of the many ways that
tire/wheel size can affect the energy requirements necessary for automotive operation:
“This is one of the reasons (decreased unsprung weight) for the
current trend (in racing) towards smaller diameter wheels and
tyres…decreased aerodynamic disturbance is another… 125”
With regard to the aerodynamic aspect reference may be made to SAE Paper 2000-01-
0491: “…a simple addition of the wheels to a basic body proved to cause a high increase in
drag, two to three times that of the body alone…” 126. Conservatively (“two…times that of
the body alone”), this would mean that, with respect to the baseline vehicle which had a Cd of
0.42 (and a Cl of 0.20 127), the necessity of providing wheels and wheel wells to the basic body
form incurred about 0.280 of that total drag coefficient; as a smooth form without wheels the
XK150S body drag coefficient would have been about 0.140. This would seem to indicate
that the new configuration might very well have a new total Cd of about 0.280, which is a
reduction of -33%.

123
This result would find favor with famed Hudson/Packard suspension engineer William D. Allison, who long
contended that an eight wheeled car would have less rolling resistance than a four wheeled, everything else being
equal, which he based on a series of scale model studies.
124
These “CSR” and “CDR” values were determined using the methodology of SAE paper 2016-01-0445,
“Estimation of the Rolling Resistance of Tires”. However, that methodology does not apply with assurance to
the extreme case of the “half-radius” tire, so there was a good deal of “guesstimating” in that case.
125
Reference [17], pg. 63.
126
Reference [58], pg. 8. Alberto Morelli was referring to the relative effect on very low drag “advanced basic
shapes” for automotive bodies, not the average production road car shape, and certainly not the XK150S shape.
127
These were the aerodynamic coefficients used in Reference [89] by this author. An empirical study by
Debojyoti Mitra suggests that for the same ground clearance to wheelbase ratio a more modern car might have a
Cd of 0.35 and a CL of 0.14, which is about 17% and 30% less (respectively) than the XK150S values.

56
SAWE Paper No. 3602
Category Number 31.0

However, a more cautious paper presented by Pfadenhauer, Wickern, and Zwicker in


1996 claims that the wheels/tires/housings only account for about 35% of the total vehicle
Cd 128; to be conservative this figure will be used as the basis for a crude estimation of the new
configuration’s drag. Since the size of the wheels/tires/housings is to be halved in the new
configuration, it would seem reasonable to say that the new configuration Cd would be likely
to decrease by 0.0735 with respect to the original value, except that the new concept would
also have twice as many wheels. However, this may not be as severe a factor as might initially
be thought; the wheels are in longitudinal alignment within the airstream and in much closer
relation to each other than the wheels were in the original configuration (wheelbase of 102.0
in, or 259.1 cm); the overall aero disturbance may be less due to a wheel “drafting” 129 effect
(although the increased number of pressure cycles may militate against this somewhat).
The axle locations in the new configuration would be 49.7 in (126.2 cm) apart if
pushed out to the furthest extent of the body envelope (and evenly spaced); if the wheels are
kept within the original wheelbase then the distance would be a much smaller 34.0 in (86.4
cm), but the “stance” would be poorer. The rule for “poorly streamlined bodies” (and the
wheels would seem to fit that description) is that each body would have to traveling in a line
no more than two critical lengths (say, twice the tire “no-load” diameter) apart for the
phenomenon to take effect, which in this case means within 27 in (687 mm), so it’s likely
there will be little benefit due to a “drafting” effect. However, that there will be some Cd
improvement for the new configuration with respect to the original seems assured; the exact
amount just can’t be determined without resort to some empirical investigation or computer
simulation. As a very rough “guesstimate”, a reduction of about 367 drag counts (0.0367) will
be conservatively assumed as being reasonable in this regard (for a new Cd of about 0.38).
Cd reduction is only one of two ways the new configuration may affect aerodynamic
drag. The other way is through frontal area reduction; although the width of the “half-radius”
tires has been kept equal to that of the “full-size” tires, the improved stance of the new
configuration may allow a modest reduction in ground clearance. A minimal decrease of 1 in
(25.4 mm) would seem reasonable, and would result in a decrease in frontal area (“Af”) of
0.08 ft2 (0.0074 m2) , which is admittedly only about a -0.4% improvement considering that
the frontal area of the baseline vehicle is 18.53 ft2 (1.7215 m2) 130.

128
BMW aerodynamicist Karlheinz Ebbinghaus corroborated this estimate of the wheels/tires/housings aero
drag contribution: “Thirty to forty percent of a car’s aerodynamic resistance is created in the wheels and
housings”; see Reference [96], pg. 14.
129
In NASCAR racing a practice known as “drafting”, when one car follows closely in the wake of another, has
been known to reduce the aerodynamic drag of both vehicles. The exact effect depends very much on the
distance, measured in car lengths, between the vehicles, and on the degree of “streamlining” of those vehicles.
130
This is the author’s 1984 estimate; Reference [16], pg. 37 quotes a figure of 18.2 ft2 (1.691 m2).

57
SAWE Paper No. 3602
Category Number 31.0

In connection with this, there is a study by Prof. Debojyoti Mitra of the Sir Padampat
Singhania University which indicates that such a reduction in ground clearance would reduce
the Cd by about 183 drag counts (and would also increase the CL by about 319 counts!) per
the following charts: 131

Figure 4.7 –Cd & Cl vs. GROUND CLEARANCE/WHEELBASE RATIO


Taking all this into consideration, the total aero (and height) changes would result in a
new Cd of about 0.36, a new Cl of 0.23, a new Af of 18.45 ft2 (1.714 m2), and a new vertical
c.g. of about 24.5 in (62.2 cm) (ignoring other possible causes for VCG change at this point).
There is another area wherein the many tires of the new configuration may produce a
very significant impact, and that is within the realm of wet weather traction. A study by Metz
and Meyers found significant “path clearing” with a conventional four wheel vehicle (2012
Ford Taurus Police Interceptor) of 113 in (287 cm) wheelbase 132. The conclusions of that
study found the effect most significant during “high ay” maneuvers 133:

“In these runs the front wheels displayed significantly more


variations than the rear wheels, indicating that the clearing
effect hypothesis is realistic and that path clearing by the
front tires does assist in mitigate (sic) rear tire
hydroplaning”.
“The flat plate experiment demonstrated that essentially the
entire contact patch was in contact with the test conditions
described, again indicating significant path clearing by the
front tire.”

131
Reference [54], pg. 2679.
132
Reference [51], pg. 29.
133
Ibid, pg. 38.

58
SAWE Paper No. 3602
Category Number 31.0

This lends credence to some of the claims made by Italian manufacturer Ferruccio
Covini with regard to his six wheel sports car, the C6W. Covini claims that the four front
wheels presented less danger of loss of control due to a front tire blowout, provided additional
braking power, reduced unsprung weight, and “less risk of aquaplaning”.

Figure 4.8 – 2004 COVINI C6W


For the conventional baseline configuration of Figure 4.1 the “path clearing” or
“squeegee length” between tires (inflated to 40 psi front/45 psi rear) is equal to the wheelbase
minus one-half the front and rear tire/road contact patch lengths: 102.0 in – ½ (6.18 in + 5.88
in) = 95.97 in (2438 mm). This a considerable distance, yet Metz and Meyers found
significant path clearing for a vehicle of even greater wheelbase (113 in/2870 mm). For multi-
axle configurations such as the proposed “advanced design concept” (for which the axle lines
are located at equal intervals along the wheelbase) the squeegee length(s) would be equal to:
𝑳𝒘𝒃
𝑳𝒔 = − ½ 𝑳𝒄,𝒇𝒘𝒅 − ½ 𝑳𝒄,𝒓𝒓 (EQ. 4.13)
𝒏−𝟏

Where:
Ls = The “squeegee length” (in, mm).
Lwb = The wheelbase (defined as the distance between the most
forward and the most aft axles) (in, mm).
n = The number of axles.
Lc,fwd = The forward tire/road contact patch length (in, mm).
Lc,rr = The rearward tire/road contact patch length (in, mm).
For real “multi-axle” (more than two axles) vehicles there would be more than one
“squeegee length” magnitude, but not given the simplification of evenly distributed/constant
vehicle weight loads adhered to so far (see footnote on page 59). Obviously, the shorter the
squeegee length the better, not only for hydroplaning resistance but also aerodynamically as
“Ls” constitutes a reasonable means of determining if two wheels in tandem are within that
critical “drafting effect” distance (“2 Di”) of each other.

59
SAWE Paper No. 3602
Category Number 31.0
Rev C

There is a more definite way to explore the wet road performance of the new
configuration “half-radius” tires; the noted Czech researcher Koutný has criticized the
traditional determination of the critical velocity at which hydroplaning occurs (“Vh”) as being
“too general” (traditionally the critical velocity is taken as being proportional to the square
root of the average contact pressure 134). Koutný states the critical velocity is better determined
by135:

𝐋𝐜
𝐕𝐡 = �𝐭 (EQ. 4.14a)
𝐬

Where:
Vh = The critical velocity for hydroplaning (mm/sec).
Lc = The longitudinal dimension of the tire/road contact (mm).
ts = The “sink time” required for tire penetration to road surface (sec).
The simple appearance of this equation is deceptive, as the “sink time” must be
determined by numerical iteration of the following transcendental equation, which is not
directly definitive (as would be expected), but instead is inversely indicative (ts = h-1(t)):

𝒉𝟎
𝒉(𝒕) = (EQ. 4.14b)
𝒕 𝟑𝑷 𝒕 𝟑𝑷
𝒄𝒐𝒔𝒉�𝒂� 𝒄 �𝒄𝒐𝒔𝒉�𝒃� 𝒄 �
𝜹 𝜹

Where:
h0 = The height of the water surface above road roughness (mm).
hr = The average height of the road roughness (mm).
ht = The total “sink” distance before total contact: ht = h0 + hr (mm).
t = The time input variable which is incremented to solution (sec).
a = One half the tire/road contact length Lc (mm).
b = One half the tire/road contact width tw (mm).
Pc = The tire/road contact pressure: Pc = N/4ab (kN/mm2).
δ = The water density (kg/mm3).

134
Koutný may have had such equations in mind as Horne & Joyner: “𝑽𝒉 = 𝟏𝟎. 𝟑𝟓�𝑷𝒊 ” (see Reference [93],
pp. 66-67). However, note the use of tire inflation pressure “Pi”, not the contact pressure “Pc”; Horne & Joyner
probably did not subscribe to the “balloon theory” of tire behavior (see Appendix G), but they had no convenient
way of obtaining “Pc” and knew that “Pi” generally reflected “Pc”.
135
Reference [40], pp. 126-127. However, Koutný himself stresses that that the equation results are only
indicative due to model simplification (inadequate flow dynamics, ideal fluid, etc.).

60
SAWE Paper No. 3602
Category Number 31.0
Rev C
If the input “sink distance” to road contact is taken as “h0”, or if the input “sink
distance” to total contact is taken as “ht”, then the results would be appropriate for a “bald” or
“slick” tire. The presence of tread pattern can be accounted for by subtraction of the product
of void factor “fv” (dimensionless) times the tread depth “td” (mm) to get an adjusted input
“sink distance” (“ht = h0 + hr – fv td”). Note that the way Koutný presents the equation
suggests that the input “sink distance” is to be customarily taken as “h0” to calculate “Vh”.

Use of Equations 4.14a and 4.14b to compute the “Vh” for the “full-size” and “half-
radius” tires at 40 psi (275.8 kPa), with input data drawn from Table 4.2, results in 45.7 mph
(73.5 kph) and 30.9 mph (49.7 kph), respectively, for an input “sink distance” of 0.1 in (2.54
mm).

While these results are indicative of relative merit, they may not be correct in
magnitude. A traditional determination of the critical velocity (such as despised by Koutný)
may not highlight the difference between tires as well as Koutný’s equation(s) can, but
traditionally resulted in what were considered reasonable magnitudes (although water depth is
not taken into account!). The Horne and Joyner version of the “traditional” equation, but with
contact pressure “Pc” substituted for inflation pressure “Pi” (which makes the equation much
more realistic) is:

𝑽𝒉 = 𝟏𝟎. 𝟑𝟓 �𝑷𝒄 (EQ. 4.14c)

Where:

Vh = The critical velocity for hydroplaning (mph).

Pc = The tire/road contact pressure (psi).

This “traditional” determination of the critical velocity at which hydroplaning occurs


produces results of 62.8 mph (101.1 kph) and 59.4 mph (95.6 kph) respectively for the
previous two cases. As this is 17 and 28 mph greater than the Koutný results, possibly the best
values for “Vh” that can be obtained would be via an average of the Koutný and Horne-Joyner
results:

“Full-Size” Tire (N = 910 lb/ 413 kg, Pi = 40 psi/276 kPa):

Vh = (45.7 + 62.8)/2 = 54.3 mph, or 87.3 kph

“Half-Radius” Tire (N = 455 lb/206 kg, Pi = 40 psi/276 kPa):

Vh = (30.9 + 59.4)/2 = 45.2 mph, or 72.7 kph

61
SAWE Paper No. 3602
Category Number 31.0

Given the uncertainty concerning exactly what the magnitude of “Vh” might be for
any specific case, it is fortuitous that the previously discussed “path clearing” will be in effect,
and that modern road car tires generally have a tread pattern resulting in a “land/sea” ratio of
70% to 65% (void factor of 30% to 35%) or better, which should provide some margin for
error. Anyway, regardless of what the exact values for “Vh” are, the preceding should suffice
for the purposes of this comparison study.
It has been noted earlier that the matters of tire load capacity and speed rating are
somewhat interrelated. The load capacity of a tire depends upon the strength of its
construction, which means the carcass material and the number of plies of that material. The
sheer size of the tire, as measured in terms of volume, is also roughly proportionate to the load
that can be carried, as is the inflation pressure (the load capacity of a tire is also said to be
inversely proportional to the speed at which the tire runs, but that may just be an effect of size
and thus accounted for earlier). The “half-radius” tires of the “advanced” (from a mass
properties standpoint) vehicle concept are taken as being of the same construction with
regard to material, number of plies, etc., but the internal volume would only be ¼ of the “full-
size” tire, and load capacity has been traditionally linked to that volume.
Load capacity was one of the earliest concerns of tire manufacturers and consumers;
for over 100 years private organizations like the Tire and Rim Association (TRA) Inc.
(founded in 1903) set the standards for rating tires, and load capacity was foremost 136. Early
formulations for determining load capacity largely resulted from practical experience and
group consensus, but evolved with time as theoretical approaches and empirical observations
were utilized for refinement. In the United States, many TRA guidelines, such as those on tire
load capacity, would eventually be incorporated into NHTSA regulations; this has proved to
be a natural progression that would find its counterpart in many other nations. The traditional
(since 1936) TRA load capacity formula was developed by mathematician C.G. Hoover (who
later became staff director of the TRA) and has been continually refined since 137:

𝑳𝒄𝒂𝒑 = 𝟎. 𝟒𝟐𝟓 × 𝑲 × 𝑷𝒊 𝟎.𝟓𝟖𝟓 × 𝑺𝟏.𝟑𝟗 × (𝑫𝑹 + 𝑺) (EQ. 4.15)


Where:
Lcap = Tire load capacity at pressure “Pi” (lb).

K = Tire service factor (Ktruck&bus = 1.00, Kpassenger car = 1.10).

Pi = Tire inflation pressure (psi).

S = Tire section width "adjusted" for equivalent circular periphery (in).

DR = Nominal wheel rim diameter (in).

136
Reference [26], pg. 187.
137
Ibid, pg. 188.

62
SAWE Paper No. 3602
Category Number 31.0
Rev C

However, there is a new “proposed” (2005) replacement for this venerable old formula
which is supposedly more rational in its approach as it is based on a “relative deflection”
(“%d”) for a tire of known load capacity at some inflation pressure “Pi” 138:

𝒅
%𝒅 = (𝑺𝑯 −𝟏𝟕.𝟓)
× 𝟏𝟎𝟎 (EQ. 4.16)

The new formula requires utilizing that “%d” for a tire of unknown load capacity at
the same inflation pressure “Pi” 139:

⎡%𝒅×�𝑺𝑵× −𝟏𝟕.𝟓�⎤
𝟏𝟎𝟎
⎢ ⎥ 𝑺𝑵 ×
𝑳𝒄𝒂𝒑 =⎢
𝟏𝟎𝟎 ⎥ × �𝟎. 𝟎𝟎𝟎𝟐𝟖 𝑷𝒊 ��−𝟎. 𝟎𝟎𝟒 + 𝟏. 𝟎𝟑� 𝑺𝑵 × �
𝟓𝟎
+ 𝑫𝑹 � + 𝟑. 𝟒𝟓� (EQ. 4.17)
⎢ ⎥
⎣ ⎦

Where:

Lcap = Tire load capacity at inflation pressure “Pi” (kg).

%d = Relative deflection at inflation pressure “Pi”.

d = Tire deflection under load “N” with vertical stiffness “KZ” as


appropriate for pressure “Pi”, calculated as “d = N/KZ + d0” with “d0 =
0” (see Eq. 4.7) (mm).

SH = Tire section height (mm).

SN = Tire nominal section width (mm).

= Tire section aspect ratio, “truncate {[SH/SN] × 100}” (no units).

Pi = Tire inflation pressure (kPa).

DR = Wheel nominal diameter (mm).

The “17.5” mm (0.689 in) deduction from the tire section height “SH” in Equation
4.16 results in the maximum possible tire deflection for reasons best understood by a
consideration of tire geometry as presented in Figure 4.9:
138
Reference [26], pg. 196.
139
Ibid, pg. 200.

63
SAWE Paper No. 3602
Category Number 31.0
Rev C

Figure 4.9 – TIRE GEOMETRY AND DIMENSIONS

Utilization of Equation 4.15 with input drawn from Table 4.2 indicates that the “full-
size” tire would have a load capacity at 40 psi (276 kPa) of 1506 lb (683 kg) 140, while the
hypothetical “half-radius” tire would have a load capacity of around 599 lb (272 kg) at the
same pressure; all of which seems very reasonable. As there are eight of the new tires, this
capacity rating seemingly would allow for an “advanced design” vehicle weight up to 4790 lb
(2173 kg) at 40 psi (276 kPa). A reasonable operating weight for the baseline vehicle seems
to be around 3640 lb (1651 kg) 141, the weight would be greater if loaded to the baseline
GVWR, which is unknown but possibly 500 lb/227 kg more for a total of 4140 lb/1878 kg.
This would seem to indicate that tire pressures less than 32 psi (179.3 kPa) must be avoided if
there are to be no other restrictions for the new configuration on a load capacity basis:

140
The load capacity at maximum inflation pressure (45 psi/ 310 kPa) for this tire is 1613 lb (732 kg), which
would result in the “full-size” tire achieving a “Load Index” rating of “97” by today’s standards.
141
The 1958 Jaguar XK150S FHC (“Fixed Head Coupe”) curb weight is 3313 lb (1503 kg); see Appendix B.

64
SAWE Paper No. 3602
Category Number 31.0
Rev C

Table 4.3 – “FULL-SIZE” vs. “HALF-RADIUS” TIRE LOAD CAPACITIES

Now let’s consider the matter from the standpoint of the “proposed” methodology of
Equations 4.16/4.17. Assuming the 152/92R16 “full-size” tire @ 40 psi (276 kPa) load
capacity of 1506 lb (683 kg) is correct, such a load would give a “d0 = 0” absolute deflection
of 1.1315 in (28.74 mm), or a relative deflection of 23.5% (max deflection = 122.74 mm, or
4.8323 in). Applying that relative deflection to the 152/46R8 “half-radius” tire, one obtains an
absolute deflection of 0.485 in (12.32 mm) 142. Per the “proposed” methodology this indicates
a load capacity of 484 lb (219.5 kg). This is only 80.7% of the previous 599 lb (272 kg) result,
which indicates that at 40 psi (276 kPa) the eight “half-radius” tires would not be adequate.
Further investigation shows that even at 45 psi the 152/46R8 tires would not be adequate (by
about 25 lb or 11.3 kg). However, for the purposes of this paper the results of the traditional
Equation 4.15 will be utilized as there is considerable uncertainty regarding the application of
the “proposed” methodology.

Having touched on the matter of a possible load restriction for the hypothetical “half-
radius” tires, this is an appropriate time to address the concern that the small wheels of the
“advanced design” concept, which would have to run twice as fast rotationally as the wheels
of the baseline vehicle, may have some liability with regard to what is called the “standing
wave” 143 phenomenon. The standing wave phenomenon was addressed by researchers
McGivern and Shirk in the following manner 144:

142
The 152/46R8 maximum deflection is 2.064 in (52.42 mm), or 17.5 mm less than the section height.
143
This is a misnomer; the “standing wave” is actually traveling around the tire circumference at a velocity that
is equal to the tire peripheral velocity so the appearance is one of a stationary wave, but it really is a traveling
wave.
144
Reference [47], pg. 1.

65
SAWE Paper No. 3602
Category Number 31.0
Rev C

“…at high speeds automobile tires develop…“tread ripple” or


“standing wave”. This phenomena (sic)…is dangerous. It
produces an intense heating effect…with over-stressed
fibers…tire failure.”
“…tests indicate that this…is…a function of speed…tire
size…tire pressure and make…weight of tire…”
The McGivern/Shirk methodology is based on the observation that when the “standing
wave” phenomenon occurs it tends (hence the results are only approximate) to have a
wavelength “λ” equal to twice the tire section width “SN”. Based on a model of the tire tread
as a series of spring/mass systems, the critical velocity “Vc” when resonance occurs and the
wave phenomenon begins is calculated as 145:

(EQ. 4.18)
Where:
Vc = The tire “standing wave” critical velocity (mph).
SN = The tire section width (in).
KZ = The tire vertical stiffness (lb/in).
tw = The tire tread width (in).
tt = The tire tread thickness (in).
δ = The tire tread material density (lb/in3).
π = The value of “Pi” (3.14159265358979…).
For the original “full-size” tire, assuming a tread density “δ” of 0.05 lb/in3 (0.0014
kg/cm3) 146, a tread thickness “tt” of 0.75 in (19.05 mm) 147, and again using the values for an
inflation pressure of 40 psi (276 kPa) from Table 4.2 (“SN = 6.00 in, KZ = 1331 lb/in, tw =
3.96 in”), the “standing wave” velocity is calculated to be about 171 mph (275 kph). For the
“half-radius” tire, with the same input values except “KZ = 997 lb/in”, the “standing wave”

145
Reference [47], pg. 5. McGivern and Shirk actually made the determination of “Vc” a two step process: first
𝟏
the resonant frequency would be determined per “𝒇 = 𝟐𝝅 �𝒌⁄𝒎”, then the critical velocity would be determined
per “𝑽𝒄 = 𝒇 𝝀”. Equation 4.11 merely represents this author’s combination of the two equations of this multi-step
process into one simple calculation.
146
This is a conservative value for tread density; other researchers have used 0.0462 lb/in3 (0.0012788 kg/cm3)
or less.
147
This is a conservative value for tread thickness; other researchers have used 0.65 inches (16.51 mm) or less.

66
SAWE Paper No. 3602
Category Number 31.0
Rev C

critical velocity is now about 148 mph (238 kph), which seems reasonable relative the “full-
size” result.
Although a “Vc = 148 mph” doesn’t seem very limiting, the “half-radius” tires would
be operating at a point about 86% of their maximum load capacity at 40 psi (276 kPa), while
the “full-size” tires were only at 69% (both for a vehicle weight of 4140 lb/1878 kg). This
may indicate a need for caution; what could be done to improve the “Vc” for the new
configuration should the need arise, and what would be the impact on “Lcap” and other
characteristics? Per a Krylov and Gilbert graph showing variation in “Vc” per change in
inflation pressure, an increase in the inflation pressure of 5 psi (34.5 kPa) to a total pressure of
45 psi (310 kPa) might increase the “Vc” by about 9.5 mph (15.3 kph) 148, and certainly
wouldn’t hurt “Lcap”, but would be at the expense of some ride quality and those performance
characteristics dependent on tire/road contact area, which would decrease by about 5%.

Figure 4.10 – TIRE CRITICAL VELOCITY vs. INFLATION PRESSURE


Also per Krylov and Gilbert the critical velocity can be very sensitive to variation in
tread thickness 149. Using another of their graphs showing a variation in “Vc” per change in
tread thickness, it would seem that the old racer’s “trick” of “shaving the tread” by 1/8 in (3.2
mm) 150 might add 48 mph (77 kph) to the critical velocity (!) 151. This last would not only
148
Reference [41], pg. 4229. Of course this is not the recommended way to do this estimation, as it is only
indicative; the chart would have to reflect actual “half-radius” tire data for this to be a valid estimate.
149
Ibid, pg. 4230.
150
Most passenger car tires have a 10/32 in (7.9 mm) tread depth when new, of which 8/32 in (6.3 mm) is usable
as the legal (USA) minimum tread depth is 2/32 in (1.6 mm). 1983 SCCA Solo-II B-Stock National Champion
driver Roger Johnson stated he shaved his stock tires down to a 5/32 in (4 mm) remaining tread depth for
reduced heat generation and increased contact area through a flatter, wider tread width.
151
Again, this is only indicative; Figures 4.9 and 4.10 are for some unidentified but supposedly conventional
sized tire which differs considerably from the “half-radius” tires considered here.

67
SAWE Paper No. 3602
Category Number 31.0

improve high speed operation with respect to standing wave formation but also would
significantly reduce energy losses due to rolling resistance. Also on the “plus side”, there
would also be some reduction in tire weight and rotational inertia. Of course, tire life would
be adversely affected, and hydroplaning resistance might also suffer, but load capacity
shouldn’t be appreciably affected and those performance aspects dependent on contact area
might actually benefit somewhat…

Figure 4.11 – TIRE CRITICAL VELOCITY vs. TREAD THICKNESS

Shaving the tread thickness might have more of an effect on the critical velocity than
even Krylov and Gilbert may have anticipated. In a contemporary research paper by Jung-
Chul An and Jin-Rae Cho the following statement is made regarding their comparison of such
“simple” critical velocity models with detailed finite element models (FEM), and with
empirical test results 152:

“One interesting observation is that the critical speeds


predicted by the simple tire model are vastly overestimated,
which implies that the disregard of the detailed tread grooves
results in a…overestimation…”

152
Reference [2], pg. 127.

68
SAWE Paper No. 3602
Category Number 31.0

Supporting this statement was the fact that, for a P205/60R15 tire inflated to 137.9 kPa
(20.0 psi) and under a 475 kg (1047 lb) load, the “simple model” used by An and Cho
returned a “Vc” value of 127 mph (205 kph), while their “detailed FEA model” resulted in
113 mph (182 kph); empirical testing showed the tire to have an actual “Vc” of 118 mph (190
kph) 153. For this P205/60R15 tire the McGivern/Shirk equation (Eq. 4.18) returned a value of
149 mph (240 kph) when using “realistic” values for “δ” (0.043352 lb/in3) and “tt” (0.6299
in). Using those same “realistic” values, critical velocity equations developed by researchers
Krylov/Gilbert returned 135 mph (217 kph) for the P205/60R15. All this would seem to bear
out An and Cho’s statement that “simple models” tend to overestimate the standing wave
critical velocity.
However, when using “conservative” values for “δ” (0.05 lb/in3) and “tt” (0.75 in),
McGivern/Shirk produced 127 mph (204 kph) and Krylov/Gilbert produced 119 mph (192
kph). Note that this Krylov/Gilbert result is almost exactly equal to the empirical result of 118
mph (190 kph), and that the McGivern/Shirk result is now much closer. This would seem to
indicate that, when lacking the resources for FEA modeling or empirical testing, the use of
“simple models” such as that developed by McGivern/Shirk or Krylov/Gilbert is justified, as
those models will give reasonably accurate results when appropriately “conservative” values
for density and tread thickness are used.
The methodology developed by Krylov/Gilbert is more sophisticated than that of
McGivern/Shirk, so the natural assumption would be to expect a better indication of the “Vc”
value by Krylov/Gilbert, which is an assumption that seems to be borne out by the
P205/60R15 result. The Krylov/Gilbert equations are 154:

𝟏
𝐤 𝟒
𝛌= � � (EQ. 4.19)
𝟒𝐄𝐉
𝟏
𝐄𝐉 𝟐
𝑪𝒄𝒓 = 𝟐𝛌 � � (EQ. 4.20)
𝛍
𝟏
𝑺 𝟐
𝑪𝒎𝒆𝒎 = �𝝁� (EQ. 4.21)

𝑪𝒄𝒓𝒊𝒕 = �𝑪𝒄𝒓 𝟐 + 𝑪𝒎𝒆𝒎 𝟐 (EQ. 4.22)

Where:

153
Reference [2], pg. 128.
154
Reference [41], pg. 4227. This multi-step process could also be combined into a single step, but insight into
exactly what Krylov and Gilbert are doing would be lost, which is that they are using for “Ccrit” the combination
of two different modeling effects, “Ccr” (beam without tension) and “Cmem” (membrane under tension).

69
SAWE Paper No. 3602
Category Number 31.0
Rev C
λ = Standing wave phenomenon wavelength (m).
Ri = Tire inflated (no load) radius (m).
b = Tire belt width (m).
h = Tire belt thickness (m).
E = Young’s modulus (N/m2).
k = Tire foundation stiffness (N/m2).
ρ = Tire mass density (kg/m3).
Pi = Tire inflation pressure (Pa).
J = Tire belt second moment of area = h3b/12 (m4)
S = Tire tension = PiRib (N).
μ = Tire belt mass per unit length = ρhb (kg/m).
These equations give a value of 118 mph (190 kph) for the critical velocity of the
“half-radius” tire, which is a significant variation from the previous McGivern/Shirk result of
148 mph (238 kph) 155. Use of the Krylov/Gilbert formulae, in lieu of the previously used
graphs, indicates that increasing the inflation pressure by 5 psi (34.5 kPa) to a total pressure
of 45 psi (310 kPa) would result in an increase in “Vc” by about 5 mph (8 kph) to a total of
123 mph (198 kph), and that “shaving the tread” by 1/8 in (3.2 mm) would result in a further 5
mph (8 kph) increase, for a grand total of 128 mph (206 kph). So, if 118 mph (190 kph)
should prove to be an accurate “Vc” estimate for the “half-radius” tire 156, the matter could be
ameliorated somewhat by increased inflation pressure (if possible) and/or tread shaving.
A consideration of a possible need for tread “shaving” leads naturally to a concern
with tire wear, or “tire life”, as noted earlier. The Pneumatic Tire gives the following simple
equation for tread wear 157:

𝑹𝑾 = 𝑲 𝑺𝒏 (EQ. 4.23)
Where:

155
To eliminate any uncertainty, it is hereby noted that these values were determined using the “conservative”
values for “δ” (0.05 lb/in3) and “tt” (0.75 in), which will be the case from this point and on.
156
It should be noted that per Reference [26] such calculations as carried out herein regarding “standing wave”
critical velocity constitute only a possibility, not a certainty, as small variations in construction can have
disproportionate consequences in this regard. Tire manufacturers have learned how to significantly offset the
calculated onset of the “standing wave” phenomenon through employment of various design details.
157
Reference [26], pg. 257. The original metric formula has been modified to be in the English Units System,
and was perhaps initially too “simple” in that it utilized “force” instead of “stress”, but in the form given herein it
seems reasonable enough for the present usage. Curiously there is “rule of thumb” regarding the inverse
relationship between the UTQG tread wear index “TW” (60 to 600, meaning 18,000 to 180,000 miles rated life)
and the basic (maximum) coefficient of traction “b” (1.22 to 0.86) as “b = 2.25/TW0.15”.

70
SAWE Paper No. 3602
Category Number 31.0
Rev C
RW = Rate of tread wear, possibly within 0.002 to 0.008 (in/1000 mi).
S = Tire/road contact area stress (lb/in2).
K = Tread compound/road pavement value, for a typical passenger car tire
/road situation possibly around 1.927306x10-6 (in/lb-1000 mi).
n = Exponent inversely varying with tire radius size and usage (accel/
braking, and/or cornering), possibly 1.90 to 2.30 (dimensionless).

Accordingly, computations of tire “mileage” (a.k.a. “life”), symbolized as “M” which


is in 1000 mile units, may be made per the following:

𝑴 = 𝟎. 𝟐𝟓 𝒊𝒏/𝑹𝒘 (EQ. 4.24)

Most passenger car tires seem to have an exponent “n” value between 1.90 and 2.30;
for the “half-radius” tires considered herein an exponent value of 2.25 seems appropriate if
the “full-size” tire exponent value is taken as 2.00. A comparison of the “half-radius” tire
wear rate (“RWhr”) versus the “full-size” tire (“RWfs”):
𝑹𝑾𝒉𝒓 𝑲 𝑺 𝒏𝒉𝒓
�𝑹 = 𝒉𝒓 𝒉𝒓 �
𝑾𝒇𝒔 𝑲𝒇𝒔 𝑺𝒇𝒔 𝒏𝒇𝒔

Making the appropriate substitutions using values from Table 4.2 at 40 psi (276 kPa)
inflation pressure results in:
𝟐.𝟐𝟓
𝑹𝑾𝒉𝒓 𝟏. 𝟗𝟐𝟕𝟑𝟎𝟔𝒙𝟏𝟎−𝟔 (𝟑𝟐. 𝟗𝟐𝟑) �
�𝑹 = = 1.91
𝑾𝒇𝒔 𝟏. 𝟗𝟐𝟕𝟑𝟎𝟔𝒙𝟏𝟎−𝟔 (𝟑𝟔. 𝟖𝟐𝟕)𝟐.𝟎

So, per this application of Equation 4.23, the “half-radius” tires would experience near
twice the baseline wear rate 158 and thereby last about one-half as long as the “full-size”159,
which is bad news if a need to “shave” the tread of the “half-radius” tires should materialize.
However, since this conclusion was arrived at in a somewhat arbitrary fashion using a poorly
documented methodology, a “sanity check” is called for. The significant factors are the
tire/road contact patch stress levels and the fact that the “half-radius” tire will have to rotate

158
The corresponding 40 psi (276 kPa) wear rates are 0.00500418 in/1000 mile versus 0.002613894 in/1000 mile
(0.0790 mm/1000 km and 0.0413 mm/1000 km, respectively).
159
Per Equation 4.24 that would be a “half-radius” tire mileage life of 49,958 miles (80,400 km) versus a “full-
size” tire life of 95,643 miles (153,923 km).

71
SAWE Paper No. 3602
Category Number 31.0
Rev C
twice (2.04) as many times as the “full-size” tire for the same distance travelled. Since the
stress on the “half-radius” tire is only 89% as much as the “full-size” tire the wear rate would
seem to be 1.82 times as much (0.89×2.04), which is better than the above 1.91 times as
much; the reality may very well lie somewhere between those values (1.86?).
As noted earlier in Chapter 2, automotive authority Pete Brock proposed obtaining
improved automotive performance through a similar reduction of tire/wheel weight and
rotational inertia as that which has been investigated herein, but through reduced width 160,
not reduced radius. To be thorough, it should be noted that there is merit to Mr. Brock’s
suggestion, at least enough merit such that tire companies such as Michelin and Bridgestone
have seriously investigated this possibility. Unmentioned in the previous recounting of the
2010 Challenge Bibendum was the fact that Michelin did not just present one energy-efficient
entry, the small radius 175/70R10 tire, but also a narrow width/large radius 195/55R21 tire.
The literature is short on specifics on just how well the Michelin 195/55R21 tire did
with regard to the many affected aspects of automotive performance 161, and in any case it is
not sufficiently representative of the narrow/tall paradigm, so a brief theoretical analysis of
the possibilities concerning a “half-wide” 76/184R16 tire was conducted for this paper. Given
the theoretical load capacity limitation of such a tire, it would take a minimum of six wheels
(three axles) to handle the presumed baseline vehicle 4140 lb (18781 kg) GVWR. Given the
possibility of such a tire, the question also arises as to what would be the effects of a “half-
size” (“half-width” and “half-radius” of baseline: 76/92R8) tire, which would necessitate the
use of at least sixteen wheels (8 axles).
As a small step toward greater realism in the accounting of the full effects of such
configuration changes, the total vehicle weight and other mass properties were no longer
held constant. This was done in a simplistic manner but should serve as to clearly indicate all
the possible effects, even when the tire/wheel type is the same but the configuration varies
(i.e., six “half-wide” wheels vs. eight “half-wide” wheels, etc.).
The resulting four basic tire/wheel types are illustrated in Figure 4.12; note that the
normal loads are now reflective of the new vehicle weights indicative of the use of such
tires in the configurations specified. A short general summary of the performance
characteristics (as has been studied so far) of these four tire types as employed in six vehicle
configurations is presented in Figure 4.13, with details presented in the following Tables 4.4
to 4.6. These tabulations serve the interest of completeness by not only indicating the
possibilities inherent in varying the tire/wheel types, but also to a limited extent of varying the
configurations in which such tire types may be employed 162.

160
Narrower than usual tires, often also utilizing relatively high inflation pressures and low hysteresis tread
compounds, constitute a common way of achieving low rolling resistance energy losses. Per Reference [96], pg.
14, the 2014 BMW i8 hybrid will ride on 195/50R20 tires front/215/45R20 tires rear, in the attempt to reduce
“weight and rolling resistance”.
161
Michelin did claim a rolling resistance reduction of -17%, largely the result of a high-silica tread compound.
162
Obviously the total possibilities approach infinity, and the consequent/collateral effects can be profound, but
the tables may provide inspiration and hopefully lead to the investigation of possibilities yet unimagined.

72
SAWE Paper No. 3602
Category Number 31.0
Rev C

Figure 4.12 – “FULL-SIZE”, “HALF-WIDE”, “HALF-RADIUS”, “HALF-SIZE”


TIRE/WHEELS

73
SAWE Paper No. 3602
Category Number 31.0
Rev C

Rev C

Figure 4.13 – SIX VARIATIONS IN TIRE/WHEEL/AXLE CONFIGURATION

74
SAWE Paper No. 3602
Category Number 31.0
Rev C

The first of the tabulations has to do with the characteristics of the four tire types as
employed in the six configurations. For example, the estimated load capacity of the “half-
wide” (76/184R16) tire indicates that for a vehicle of the baseline GVWR a minimum of six
tires would be required, but a configuration employing eight such tires is also considered. The
consequent deflections “d”, rolling radii “Rr”, tire/road contact area lengths “Lc”, contact
areas “Ac”, and rolling resistance coefficients “Cr” are all appropriate for the baseline weight
of 3640 lb (1651 kg) as modified by the change in tire type/number; in each case this
resulting weight (“Wt”) is divided by the number (“n”) of wheels to determine the
average normal load (“N = Wt/n”):

Table 4.4 – TIRE (40 psi, 275.8 kPa) CHARACTERISTICS


The second of these tabulations has to do with the mass properties of the tire/wheel
types 163 and the consequent mass properties (weight, vertical c.g., and effective mass) of
the vehicle those types are employed on. In this table certain configurations achieve what
might be considered the best “scores” in certain categories, and those “winning scores” are
highlighted. For instance, the “advanced design” (“Config. No. 4”), which is shod with 8
“half-radius” (152/46R8) tires, has the lowest total tire/wheel weight and lowest total
tire/wheel rotational inertia, and more importantly has the lowest vehicle weight and the
lowest vehicle effective mass.

Table 4.5 – VEHICLE/TIRE (40 psi, 275.8 kPa) MASS PROPERTIES

163
Remember that the tire/wheel weights and rotational inertias include the tire, wheel, hub, brake, and
bearings.

75
SAWE Paper No. 3602
Category Number 31.0
Rev C

The last of this set of tabulations has to do with vehicle performance in certain
categories as a consequence of the results of the previous two tables. For instance, the “full-
size” tires of the Baseline (Configuration Number 1) will have the greatest resistance to
“standing wave” phenomena, while the vehicle itself will have the lowest aerodynamic lift (in
a tie with Configuration Number 2). Perhaps most interesting, at least for those involved in
the design of ultra fuel efficient vehicles, is that Configuration Number 6 has both the lowest
roll resistance, the lowest aerodynamic drag, and the greatest lift 164 (not to mention the least
unsprung weight).

Table 4.6 – VEHICLE PERFORMANCE (40 psi, 275.8 kPa)

As can be seen from the tables, the “half-wide” tire that would be so enamored of by
Mr. Brock does have much to recommend it, but for mass properties and related reasons the
“half-radius” tire seems to be the better choice. While the “advanced design” concept (a.k.a.
Configuration Number 4), which employs the “half-radius” tire, does not obtain the highest
score in any category other than mass properties, it may well represent the best overall
design compromise. The veracity of that statement will be revealed in the next three chapters
wherein vehicle performance in terms of motion in the longitudinal (acceleration, braking),
lateral (steady-state and transient maneuver, directional stability, roll-over), and vertical (road
contact/shock/vibration, ride pitch/roll) directions will be investigated. The values listed in
Tables 4.4, 4.5, and 4.6 are indicative, not definitive, in determining which configuration
might be best for a particular purpose; firm decisions have to be made on a more exhaustive
basis.

164
Minimum aerodynamic lift is generally considered a desirable trait for ground vehicles which derive all
primary motivating (acceleration, braking) and control (steering, stability) forces from road contact (hence the
employment of aerodynamic “down force” for race vehicles). However, for some applications (ultra-high fuel
economy vehicles) there exists the possibility that deliberately generated lift forces can be put to productive use
by reducing the energy lost to tire hysteresis.

76
SAWE Paper No. 3602
Category Number 31.0

5 – LONGITUDINAL PERFORMANCE: ACCELERATION/BRAKING

In 1984 this author presented the paper “Mass Properties and Automotive
Longitudinal Acceleration” at the 43rd Annual International Conference of the SAWE in
Atlanta, Georgia 165. That paper dealt with the effect, as determined by a rigorous computer
simulation, of the variation in various mass property parameters (weight, longitudinal
center of gravity, vertical center of gravity, and rotational inertias) on automotive
longitudinal acceleration; the program used was named “MAXGLONG.BAS”. A major
advantage of such a computer simulation is that the effect of interconnected parameters can be
examined independently in a way that is not possible in the physical world; that capability
will now be used to explore the feasibility of the proposed “advanced design”.

The validation of that program was made via simulated acceleration runs which
corresponded to documented empirical cases; all of the cases involved various configurations
of the 1958 Jaguar XK150S. Run #1 involved a “Fixed Head Coupe” (four-seat hardtop)
equipped with the Moss 4-speed manual “Type A” transmission and possessing a test weight
of 3640 lb (1651 kg); tire pressures were set to 40 psi (276 kPa) front and 45 psi (310 kPa)
rear. Run #2 involved another “Fixed Head Coupe” (FHC) with the Moss 4-speed manual
“Type B” transmission at a test weight of 3584 lb (1626 kg); tire pressures were set to 23 psi
(159 kPa) front and 26 psi (179 kPa) rear. Run #3 involved a “Roadster” (rudimentary soft-
top two-seater) also equipped with the Moss “Type B” transmission at a test weight of 3492
lb (1584 kg); tire pressures were again set to 40 psi (276 kPa) front and 45 psi (310 kPa) rear.
In most other respects these cars were essentially identical. The program validation was
successful; the simulated time-to-velocity plots closely corresponded to the empirical plots. A
summary tabulation of the results follows as Table 5.1 166:

Table 5.1 – SUMMARY “MAXGLONG.BAS” PROGRAM VALIDATION RUNS

165
Reference [89].
166
Ibid, pg. 20.

77
SAWE Paper No. 3602
Category Number 31.0

In Chapter 2 of this current paper the vehicle parameter set for “Run No. 1”, with
emphasis on the rotational inertia values, was presented as being a representative passenger
car data set, i.e., the “baseline” vehicle. Table 2.1 presented the test weight, rotational
inertias, basic engine/transmission/clutch/differential characteristics, and basic rear (driven)
tire/wheel data. Table 2.2 revealed how the vehicle weight, and the rotational inertias as
affected by the drivetrain gear ratios, resulted in various effective mass (weight) in gear
values. Now, initially with as little collateral change as possible, the information in those
tables will be modified to reflect the “advanced design” with its reduced rotational inertias.

Colin Campbell, in his book The Sports Car, stated that for a powerful sports/racing
car, such as one of Jim Hall’s Chaparrals 167, approximately 60% of the thrust (in first gear) is
used to rotationally accelerate engine components, transmission components, drive shafts,
wheels, tires, and brake components 168. Such a significant change in vehicle performance as
that example would suggest is not expected herein; a much smaller proportion of the baseline
vehicle’s overall mass is rotational. Also, not all the baseline’s rotating components are being
reduced, just some (highlighted below) that contribute to the “I1” and “I4” inertias (compare
with Table 2.4):

Table 5.2 – REVISED I1 & I4 MASS CONTRIBUTORS (RUNS #5 to #7)

167
Reference [12], pg. 263. Chaparral Cars Inc. was founded in 1962 by Hap Sharp and Jim Hall to build and
race cars, usually of rather innovative design. After much success in Can-Am, Indy, and other racing venues, the
company folded in the 1970’s.
168
Ibid, pp. 264-265.

78
SAWE Paper No. 3602
Category Number 31.0

The changes are essentially as predicted in Chapter 4, a 50% reduction in weight (-


104.1 lb, or -47.2 kg) and 87% reduction in rotational inertia (-150.4 lb-ft2, or -6.3 kg-m2).
At this point no change has yet been incorporated for power delivery to the drive wheels,
although some change will be necessary. Also at this point there is no steering mechanism
weight penalty, nor any weight reduction benefit for spare tire removal, etc. As noted
earlier, at this point all collateral changes associated with the “half the radius, twice the
number” tire/wheel concept are held to a minimum so that the effect of the “half-radius”
tire/wheels will be clear and stand alone; a following investigation will involve all the
associated changes: aerodynamic, suspension, propulsion, et cetera.

However, the new “advanced design” vehicle will require at least one significant
change early on. It is a change required when the drastically reduced rolling radius of the
“half-radius” drive wheels is incorporated into the acceleration model; such a change alters
the overall gear ratio of the whole drive system. Think of the driving wheels as just one more
gear in mesh with the infinitely larger “driven gear” of the earth itself; given the disparity in
relative inertia it is the vehicle that moves while the earth seems unaffected. Therefore such
a reduction in rolling radius is like going to a numerically much higher final drive ratio.

The inline six-cylinder engine of the “baseline” vehicle has a “redline” of 5500 rpm
which necessitates non-optimum shifts upon reaching that rpm; the speeds in gear with the
“half-radius” 152/46R8 tires @ 45 psi (310.3 kPa) would be around 15, 27, 40, and 51 mph
(24, 44, 64, and 82 kph), resulting in a top speed of only about 65 mph (105 kph) 169. That
doesn’t sound like a recipe for sparkling performance; a change to the final drive ratio has to
be made to compensate for the reduction in rolling radius. If the baseline rear axle ratio of
4.09 is changed to the near half-value (46.6%) of 1.9078, then the resultant speeds in gear
would be around 32, 58, 85, and 109 mph (51, 93, 136, and 175 kph), terminating in a top
speed of about 140 mph (225 kph) 170; now that seems more sporting! These “advanced
design” shift points are essentially the “baseline” vehicle’s shifts during a WOT (“wide open
throttle”) acceleration run. With the new rolling radius and new rear axle ratio the propulsive
force available (traction permitting) at each shift point will be about the same for both
configurations, only the effective mass (weight) will be less for the “advanced design”!

169
These speeds were calculated for a static rolling radius of 6.33 in (160.8 mm). The drive wheel static radius
value used for the baseline “Test Run #1” in 1984 was 13.57 in (344.7 mm) (which seemed reasonable at the
time as Reference [12] on page 294 gave 13.6 in as the radius). For “Test Run #1a” revised baseline vehicle CG
and tire parameters were used; the static rolling radius was about 12.77 in (324.4 mm). Note that the “advanced
design” drive wheel static rolling radius of “6.33” is near half-sized; it is about 46.6% of the “13.57”, and 49.6%
of the “12.77”.
170
Horsepower determined the terminal velocity for the 1958 Jaguar XK150S, not the rpm redline. At low tire
inflation pressure settings (23 psi front/26 psi rear) terminal velocity was around 110 mph (177 kph), while at
high settings (40 psi front/45 psi rear) terminal velocity was around 132 mph (212 kph); higher speed was not
possible without more power and/or reductions to the aero drag and rolling resistance.

79
SAWE Paper No. 3602
Category Number 31.0

Consequently, the “baseline” Tables 2.1 and 2.2 are recalculated as the following
“advanced design concept” Tables 5.3 and 5.4:

Table 5.3 “ADVANCED CONCEPT” PARAMETER SET

Table 5.4 “ADVANCED CONCEPT” EFFECTIVE MASS BREAKDOWN (RUN #7)

Note that while the actual (scale) weight reduction is only -104 lb (-47 kg), there has
been a reduction in the total effective weight through the gears of about -160 lb (-73 kg)
with respect to Table 2.2 resulting from this crude estimation 171.

171
If it were not necessary to double the number of wheels in order to handle the load, then the simplified effect
(no collateral changes) of halving the wheel radii would be an ideal scale weight reduction of 156 lb (71kg),
for a total effective mass (weight) reduction of about 251 lb (114 kg).

80
SAWE Paper No. 3602
Category Number 31.0
Rev C

The effect on longitudinal acceleration of some of these parametric changes (weight,


inertias) can now be determined by input to the automotive acceleration simulation computer
program “MAXGLONG.BAS”. However, that program, as most other such analysis routines,
was written for analysis of conventional 4-wheel/2-axle automotive configurations. Therefore,
in order to realistically (including all collateral effects) analyze the unconventional 8-wheel/4-
axle “advanced concept” configuration, that configuration must be converted at some point
into an “equivalent conventional configuration” for input into the “MAXGLONG.BAS”
program.

The conversion of the “advanced concept” into an equivalent conventional


configuration is accomplished by simple moment balance calculations as illustrated in Figure
5.1. Note that a conventional front axle location is hypothesized to exist at longitudinal station
“10.785”, which is exactly mid-point between the forward axle pair. A conventional rear axle
is correspondingly hypothesized to exist at longitudinal station “110.115”, which is between
the aftward axle pair. A longitudinal moment balance calculation about each of these
equivalent conventional axle locations establishes the equivalent tire (vertical spring rate) and
vehicle characteristics (wheelbase, LCG) necessary for the equivalent conventional model
to behave in a manner characteristic of the actual “advanced concept”:

Figure 5.1 – “ADV CONCEPT” (RUN #7) EQUIVALENT CONVENTIONAL CONFIG

81
SAWE Paper No. 3602
Category Number 31.0
Rev C
In the case depicted in Figure 5.1 the equivalent tire vertical spring rates are
determined to be 1994 lb/in (35.6 kg/mm) forward and 2194 lb/in (39.2 kg/mm) aftward
(which just happens to be double the corresponding original fore and aft tire stiffnesses in this
case). The conversion also reveals that the “advanced concept” equivalent (“effective”)
wheelbase is only 99.33 inches (252.3 cm), or 2.67 inches (6.8 cm) less than the original
baseline vehicle 172. Also, the equivalent LCG is 42.225 inches (107.3 cm) aft of the front
axle, whereas the baseline LCG was 53.01 inches (134.6 cm) aft of the front axle. As a
consequence of these stiffness and LCG changes the static rolling radii, deflections, and other
characteristics of the “half-radius” tires had to be recalculated for the Run #6 input.

The results of the initial acceleration runs are tabulated in the following Figure 5.5.
Run #1 is the original baseline vehicle run made in 1984, while Run #1a is a replay of that run
utilizing revised tire and vehicle CG data. This replay of the baseline run was necessary to
eliminate the effect of new tire parametric estimation methodology (and revised baseline
vehicle weight accounting), from obscuring the effect of the contemplated mass properties
and associated changes; Run #1a is the appropriate benchmark against which all subsequent
acceleration runs are to be compared.

Run #4 173 is the first of those subsequent runs: it represents the effect on acceleration
of the full ideal reduction in baseline vehicle tire/wheel weight of -156 lb (-71 kg), for a total
effective weight impact of -281 lb (-127.5 kg). The salutary effect on acceleration represents
the ideal of simply reducing baseline “tire/wheel” weight by 3/4th and rolling inertia by
15/16th without any consideration of consequent effects/design changes. Acceleration
improves throughout the speed range, which itself increases by 1 mph (1.6 kph) to 134 mph
(215.6 kph).

Run #5 represents an initial concession to realism; four “half-radius” wheels would be


inadequate in load capacity, so the number of wheels (still “full-size” dimensionally) is
increased to eight. This results in a weight reduction of only -104 lb (-47 kg), for a total
effective weight impact of -219 lb (-99.3 kg); acceleration is adversely affected but
significant improvement over the baseline remains.

Run #6 represents yet another step toward realism: incorporation of the “half-radius”
tire data and conversion of the “advanced concept” to an equivalent conventional
configuration. All this revealed a serious lack of traction at the drive wheels due to the LCG

172
This “equivalent” (or “effective”) wheelbase change is indicative of a reduction in pitch stability, which is
something that tends to occur as the number of support points within a wheelbase increases. If the “advanced
concept” wheels had not been “pushed out to the maximum extent allowed by the existing vehicle envelope” as
made possible by the reduction in wheel radius, then the reduction in equivalent wheelbase would have much
greater than a mere 2.67 inches (6.8 cm).
173
Runs #2 and #3 are the other 1984 baseline vehicle program validation runs.

82
SAWE Paper No. 3602
Category Number 31.0

being further forward than the baseline (thereby reducing the normal load on the drive
wheels) plus the reduced ground contact area of the “half-radius” drive wheels.

Therefore, in order to get Run #6 underway without wheelspin or stalling the engine, it
was necessary to stipulate that both rear axles be driven, and that the basic coefficient of
traction be of a magnitude more appropriate for modern tire construction than the 1958
baseline. Even with those changes in place the acceleration run started poorly, and it took
until after 20 mph (32 kph) before improvement over the baseline could reassert itself. The
new computed top speed is 135 mph (217.3 kph), representing another improvement of 1 mph
(1.6 kph) in top speed, which is mainly a consequence of a 1% theoretical improvement in the
differential efficiency, which accompanied the reduction in gear ratio mandated by the “half-
size” tire radius. Vehicle ground clearance, height overall, VCG, and aerodynamic center of
pressure (CP) all changed in accord with the incorporation of “half-radius” tire data and the
conversion to an equivalent 4-wheel/2-axle reference axis system. Weight reduction is -104
lb (-47 kg), with a total effective weight reduction of -162 lb (-73.5 kg), which is very close
to the previous crude “advance concept” estimate of -104/-160 lb.

Table 5.5 – “BASELINE” vs. “ADVANCED CONCEPT” ACCEL RUNS, PART 1

Continuing toward an increasingly more realistic “advanced concept”, a second batch


of acceleration runs are summarized in Table 5.7. Run #7 was modeled to include certain
secondary effects as noted earlier in Tables 4.4 to 4.6: namely the changes to the aerodynamic
lift/drag coefficients, small frontal area reduction, plus rolling resistance coefficient
reductions; all of which mean less resistance to motion. These changes boosted acceleration
performance back to where it was for Run #4, although mainly at the high end; top speed is
now more than 7 mph (11.3 kph) greater than “baseline” 174!

174
Top speed is now also determined by engine rpm redline, not horsepower as previously. Change of the final
drive ratio (differential ratio, rolling radius) constitutes the governing factor for top speed determination from
this point on.

83
SAWE Paper No. 3602
Category Number 31.0

Run #8 represents the most realistic appraisal of the “advanced concept” acceleration
performance; all consequent design effects are now included. These involve weight and CG
changes due to the suspension/drivetrain impacts of converting to an eight wheeled design
with both rear axles driven and the forward three axles steered. Also, as the “half-radius”
wheels provide insufficient space for effective outboard brakes 175, inboard braking became
part of the design concept; this rationalization of the braking system included a sizing increase
and provision for half-shafts at all the wheels whether driven or not. The
drive/steering/braking mass property penalties reduce the weight/effective weight benefits
from that of the ideal “half-radius” design. This reduction of weight/effective weight benefits
is reflected mainly via decreased performance in the low end of the acceleration run.

Run #9 presents a “tweak” to the “advanced concept”; it is not an unavoidable


consequence of that concept. Due to the CG having migrated far forward, it is now proposed
to mitigate that movement by mounting the CW rotating (as viewed from the front) engine in
a “sidewinder” position with the rotation vector pointing toward the vehicle “right-hand” side.
This will not only mitigate the forward CG bias, but will also mitigate vehicle roll and lateral
“weight transfer” in maneuver due to a counteracting gyroscopic reaction 176. There is much
else can be done to enhance the performance of the “advanced concept”, but those changes
would constitute such radical departure from the baseline vehicle as to render all connection
with it moot, and thus are outside the scope of this paper. However, there will be some
discussion of what those further changes could be in the “Conclusions” chapter of this paper.

The weight and balance analysis of Appendix B details the mass properties
impacts of converting the “baseline” vehicle of Run #1a to the “advanced concept” vehicle of
Runs #8/#9 (Run #7 is presented only for comparison; it has no MOI/POI estimation) 177. The
weight penalties of having two rear axles driven, three front axles steering, and inboard
braking all around, as well as the weight penalty for the engine relocation of Run #9, are all
incorporated therein. The obtaining of adequate braking and drive for the “advanced concept”
also resulted in some rotational inertia penalties to the ideal “rolling” masses of Table 5.2;
those changes to inertia groups “I1” and “I4” are presented in Table 5.6 and the final effective
weight changes are shown in Table 5.7. Because of all the concessions to reality, reflected in
the aforementioned weight penalties, the final reduction in effective mass (weight) is much
less than that theoretically anticipated in Chapter 4.

175
This lack of space for effective I/B braking is based on an assumption that conventional disc or drum brakes
would be employed; crown brakes reputedly offer about 25% greater capability for the same wheel confinement.
176
Reference [91], pp. 51-52.
177
The determination of the revised (from 1984 Run #1) “baseline” mass properties (to Run #1a) and all
subsequent Advance Concept permutations is detailed in Appendix B. All weight, center of gravity, MOI, and
POI estimates are made in accord with the formulae established in Reference [88].

84
SAWE Paper No. 3602
Category Number 31.0

Table 5.6 - I1 & I4 MASS CONTRIBUTORS, FINAL REVISION (RUN #9)

Table 5.7 – EFFECTIVE MASS: BASELINE (RUN #1a) TO ADV CONCEPT (RUN #9)

85
SAWE Paper No. 3602
Category Number 31.0

The numerical results for Run #7, #8, and #9 are presented in Table 5.8. Essentially all
the Run #7 changes interacted to reduce parasitic energy losses (even the increase in lift, as it
reduced rolling resistance by virtue of its effect on normal loads at the tires 178); acceleration
got off to a poor start, but the reduced energy losses asserted themselves resulting in the best
quarter mile ET. Run #8 acceleration decreased due to the realistic weight accounting which
incorporated the suspension/drive/braking/steering weight/rolling inertia penalties; but
much improvement over the baseline remained, particularly at the high end. Run #9’s initial
acceleration should have improved due to the engine relocation, but instead remained
essentially the same because the CG shift benefit was offset by the use of more conservative
tire traction values 179.

Table 5.8 - “BASELINE” vs. “ADVANCED CONCEPT” ACCEL RUNS, PART 2

Taking the configuration of Run #9 as the ultimate manifestation of the “advanced


concept”, it might seem disappointing that ultimately, after all the convolutions of concept
development, the weight savings proved to be only -11 lb (-5 kg), for a total effective weight
reduction of only about -80 lb (-36.3 kg). Such small mass property returns might seem
hardly worth all the effort, especially given the modest improvement in acceleration/top
speed. However, the “advanced concept” has other performance advantages yet to be
investigated and made manifest; investigation with regard to those other performance areas
(lateral motion, vertical motion, fuel economy, safety, producibility/marketability) will be the
objective of the next few chapters.

178
Per Tom Froling, General Motors “Lead Development Engineer – Aerodynamics”: “Most production cars
are designed with some lift at speed in order to slip through the air for greater fuel economy…” (Reference
[27], pp. 11-13).
179
This was the result of the commitment on this author’s part of being as realistic in the analysis of the
“advanced concept” as possible throughout.

86
SAWE Paper No. 3602
Category Number 31.0

The numerical acceleration performance information as presented in Tables 5.5 and


5.8 is hard to fully absorb. Presenting the runs as time-to-velocity plots in a single graph
visually illustrates the acceleration performance development of the “advanced design
concept” in a more comprehensible fashion, hence the following Figure 5.2:

Figure 5.2 - “BASELINE” vs. “ADVANCED CONCEPT” ACCELERATION RUNS

87
SAWE Paper No. 3602
Category Number 31.0

Having spent so much effort with regard to acceleration, the following braking
analysis might seem superficial. Braking is just the reverse of acceleration, an observation
which in itself tends to give some cause for optimism regarding the performance of the
“advanced concept” in this area; just as a reduction in effective mass led to improved
acceleration, it also should lead to improved deceleration. The simplest (2-dimensional)
depiction of a braking vehicle of weight “Wt” initially moving with a velocity “V”, but
sliding to a stop in a distance “d”, is:

Figure 5.3 – SIMPLEST DEPICTION OF BRAKING

The only force in line with the vehicle displacement is the dynamic friction force “f”,
so this is the only force doing any work. The total work done by “f” between Point 1 and Point
2 is “f × d”, and since “f” is directionally opposite the displacement this represents negative
work. If the vehicle comes to a complete stop (“V = 0”) at Point 2, then the friction work done
at that point has completely dissipated (equaled) the vehicle’s kinetic energy as it existed at
Point 1, where the full braking effort was initialized 180:

𝑾𝒐𝒓𝒌 = 𝑲𝒊𝒏𝒆𝒕𝒊𝒄 𝑬𝒏𝒆𝒓𝒈𝒚 (EQ. 5.1)

𝟏
𝒇 × 𝒅 = 𝒎𝑽𝟐 (EQ. 5.2)
𝟐

180
Driver reaction time of about 0.75 seconds would add about 66 ft (20.1 m) to the vehicle stopping distance as
calculated for an initial velocity of 60 mph.

88
SAWE Paper No. 3602
Category Number 31.0

Since “f” is just simple Coulomb friction, we may substitute “Wt μ” for “f”, where
“μ” is just the appropriate coefficient of friction for tire rubber sliding on asphalt (which
according to one source can have a value between 0.6 - 0.8 181) and “Wt/g” for “m”:

𝟏 𝑾𝒕
𝑾𝒕 𝝁 𝒅 = 𝑽𝟐 (EQ. 5.3)
𝟐 𝒈

Solving for “d”, this reduces to:

𝟐
𝒅 = 𝑽 �𝟐𝝁𝒈 (EQ. 5.4)

Taking “V” equal to 30 mph (48.3 kph), a.k.a. 44 ft/sec (13.4 m/s), “μ” = 0.7 and “g”
= 32.174 ft/sec2 (9.807 m/sec2), results in a distance “d” of:

d = (44)2/ (2 (0.7) 32.174) = 42.98 ft, or 13.10 m


The baseline 1958 Jaguar XK150S reportedly could stop from 30 mph (48.3 kph) in
33 ft (10 m) 182, which means that “μ” would have to have been about 0.91. And “therein lies
the rub”; the simplistic Equation 5.4 can produce reasonable stopping distance values for any
vehicle in any situation (dry, wet, etc.) by merely adjusting the value of “μ” to suit, but there
is little of engineering use to be learned thereby as the complexity of the actual situation is not
actually modeled. If the values of “μ” used are not to be merely convenient, giving realistic
looking stopping distances, but to also reflect actual tire properties, then use of Eq. 5.4
becomes less enlightening.

To begin with, braking to a stop requires dissipating all the kinetic energy of the
183
vehicle , which means not just the energy associated with the translationally moving mass,
but all the rotationally moving mass as well. Taking Equation 5.2 and substituting the “me” of
Equation 2.5 (dropping the “I2” term for a “clutch disengaged” condition 184) for “m” brings
the math model a little closer to reality:

𝟏 𝑾𝒕 𝑰𝟏+𝑰𝟑 (𝑨𝑹)𝟐 (𝑨𝑬)+𝑰𝟒


𝒇×𝒅= � + � 𝑽𝟐 (EQ. 5.5)
𝟐 𝒈 𝑹𝑫𝟐 ×𝒈

181
Reference [85], pp. 104-107. Other sources put the dry sliding (a.k.a.: kinetic, dynamic) rubber on concrete
coefficient as 0.6 to 0.85, with a value of 0.80 most common. All this may be dated, as modern (post-2000) low
profile ( 65 or less) passenger car tires may have a dry sliding coefficient of up to around 0.90 (see Figure
5.5).
182
Reference [101], pg. 118.
183
This holds except in the case of regenerative braking when the kinetic energy, or at least a portion of it, is
stored for re-use.
184
This avoids the matters of: 1) initial input of engine energy, 2) transition to “engine braking” as the vehicle
begins to drive the engine (which then acts as a pneumatic pump).

89
SAWE Paper No. 3602
Category Number 31.0

Another step toward reality involves “f”; there is no such single force acting on the
vehicle to slow it down. There are “friction” forces at the tire/ground contact points, which
result from true friction forces at the brake surfaces generating rotation resisting torques “Tf”
and “Tr”, and there are the ubiquitous rolling resistance forces (not shown), plus aerodynamic
drag (which is often neglected). This more complicated situation contrasts sharply with the
simplicity of Figure 5.3:

Figure 5.4 – MORE REALISTIC DEPICTION OF BRAKING

Even this figure is not completely realistic. However, it does show the aero forces “D”
(drag) and “L” (lift) acting at the “CP” (center of pressure). More importantly, it now shows
the individual “friction” forces at the front and rear (2-dimensional model) tire/road contact
areas, and indicates how those forces are influenced by longitudinal “weight transfer” (which
is modified somewhat due to the aero moments) resulting from the “ma × h” moment. The
previous Equation 5.5 may be revised in accord with this greater realism:
𝟏 𝑾 𝑰𝟏+𝑰𝟑 (𝑨𝑹)𝟐 (𝑨𝑬)+𝑰𝟒
�𝒇𝒇 + 𝒇𝒓 + 𝑭𝑹𝒇 + 𝑭𝑹𝒓 + 𝑫� × 𝒅 = 𝟐 � 𝒈𝒕 + � 𝑽𝟐 (EQ. 5.6)
𝑹𝑫𝟐 ×𝒈

This equation presents another indication as to how the simplicity of Equation 5.4
could still be made to produce seemingly reasonable results: the amount of the kinetic energy
to be dissipated was underestimated, but so was the number of dissipating forces.
The aerodynamic forces are a function of velocity 185:

𝑪𝒅 𝑨𝒇 𝑽𝟐�
𝑫 = 𝟖𝟒𝟏 (EQ. 5.7)

185
The “841” factor is appropriate when the “D, Af, V” are in units of pounds, square feet, and feet per second,
respectively. If “V” is to be in mph, then the factor should be “391”.

90
SAWE Paper No. 3602
Category Number 31.0

𝑪𝒍 𝑨𝒑 𝑽𝟐�
𝑳 = 𝟖𝟒𝟏 (EQ. 5.8)

The front and rear rolling resistance forces are functions of velocity and the front and
rear normal loads “Nf” and “Nr”, as discussed earlier in Chapter 4:

FRf = CSf Nf + 3.24 CDf (V/100)2.5 Nf (EQ. 5.9)

FRr = CSr Nr + 3.24 CDr (V/100)2.5 Nr (EQ. 5.10)

The front and rear “friction” forces are also determined by velocity (which will be
handled in an ad hoc manner as no established mathematical relation to velocity is known to
this author 186) and normal load:

𝒇𝒇 = 𝑵𝒇 𝝁𝒇 (EQ. 5.11)

𝒇𝒓 = 𝑵𝒓 𝝁𝒓 (EQ. 5.12)

However, further complication results from the fact the “friction” coefficients “μf” and
“μr” are themselves a function of normal load 187 or, more precisely, contact area pressure
“Pc” 188:

𝝁𝒇 = 𝒂 𝑷𝒄𝒇 𝒏 (EQ. 5.13)

𝝁𝒓 = 𝒂 𝑷𝒄𝒓 𝒏 (EQ. 5.14)

Some typical tire specific values for “a” and “n” might be “15.7369” 189 and “-
0.67791”, respectively. Since “Pc = N/Ac”, the obvious substitution is:
𝒏
𝑵
𝝁𝒇 = 𝒂 � 𝒇�𝑨 � (EQ. 5.15)
𝒄𝒇

𝒏
𝑵
𝝁𝒓 = 𝒂 � 𝒓�𝑨 � (EQ. 5.16)
𝒄𝒄𝒓

186
In Reference [93], on page 17, Prof. Wong states “In preliminary performance calculations, the effect of
speed (on rolling resistance) may be ignored…”. In the case of tire traction, in so much as it varies in
relationship to speed, a similar rationale seems to be generally exercised. For actual empirical data on how
traction may vary with velocity see Reference [20], pg. 12.
187
Reference [28], pg. 57. “Increasing vertical load is known to…reduce normalized traction levels
(Fx/Fz)…”. A more concise statement would say that increasing contact pressure reduces traction coefficient.
188
Reference [40], pg. 61. This author’s ad hoc accounting for the velocity effect on traction resulted in the
combined “Pc ” and “V” function: “μp = a Pcm1 Vm2”.
189
This is the value of “a” for pressure units in psi; for pressure units in kPa use “58.2587” for “a”.

91
SAWE Paper No. 3602
Category Number 31.0

Of course, the contact areas will also vary in accord with normal loads as per the
discussion in Chapter 4:

𝑨𝒄𝒇 = 𝟏. 𝟐𝟒 𝑹𝒊𝒇 𝑪𝒐𝒔−𝟏 ��𝑹𝒊𝒇 − 𝒅𝒇 �/𝑹𝒊𝒇 � 𝒕𝒘𝒇 (EQ. 5.17)

𝑨𝒄𝒓 = 𝟏. 𝟐𝟒 𝑹𝒊𝒓 𝑪𝒐𝒔−𝟏 [(𝑹𝒊𝒓 − 𝒅𝒓 )/𝑹𝒊𝒓 ] 𝒕𝒘𝒓 (EQ. 5.18)

This in turn necessitates the determination of the deflections “df” and “dr” with
normal load (again, refer to Chapter 4):

𝑵𝒇
𝒅𝒇 = �𝑲 + 𝒅𝟎𝒇 (EQ. 5.19)
𝒁𝒇

𝑵𝒓
𝒅𝒓 = �𝑲 + 𝒅𝟎𝒓 (EQ. 5.20)
𝒁𝒓

An initial concentration on the normal loads would seem to be key; the normal loads,
which result from the static longitudinal weight distribution as modified by dynamic
“weight transfer”, aerodynamic drag and lift 190:

𝑵𝒇 = {𝑾𝒕 (𝑾𝒃 − 𝑳𝑪𝑮) + 𝒎𝒆 𝒂 (𝑽𝑪𝑮) − 𝑫 (𝑽𝑪𝑷) − 𝑳(𝑾𝒃 − 𝑳𝑪𝑷)}⁄𝟐𝑾𝒃 (EQ. 5.21)

𝑵𝒓 = {𝑾𝒕 (𝑳𝑪𝑮) − 𝒎𝒆 𝒂 (𝑽𝑪𝑮) + 𝑫(𝑽𝑪𝑷) − 𝑳(𝑳𝑪𝑷)}⁄𝟐𝑾𝒃 (EQ. 5.22)

The situation is even more complex than is apparent from the simple equations
presented so far. At the heart of this complexity is the fact that the situation is not one of
simple Coulomb “friction” (which was first indicated when the “μ” coefficients were
revealed as being functions of normal load). Non-sliding tire/road interaction is better termed

190
The variables “Nf, Nr, Wt, me” are in lb units, while all the dimensions are in inches and the acceleration is in
gravities.

92
SAWE Paper No. 3602
Category Number 31.0

“traction”, to distinguish it from the “true” 191 sliding friction (skidding) case into which
traction may devolve. Exactly how “traction” further differs from “friction” is perhaps best
illustrated by the diagram of how “μ”, itself determined by the normal loads (and influenced
by velocity), in turn determines what amount of apparent slippage “%S” between tire and
road will be tolerated (usually around 10-15%) before true slip (i.e.: wheelspin, locked
wheels, skid, etc.) ensues:

Figure 5.5 – TIRE BRAKING TRACTION COEFFICIENT VS. %S, VELOCITY

“Slip” is perhaps a misnomer; it certainly is misleading; it would probably be better to


use the word “Stretch”. The tire segment in contact with the road and under traction is not in
general motion with respect to the road (although some portions may be); on the whole the
tire contact area may be regarded as “static” (“V = 0”) with respect to the road. What actually
happens is that, as each tire segment rolls into contact with the ground, there is a longitudinal
stretching of the tread, followed by a contraction as the segment rolls back up out of contact.

191
Even sliding tire/road contact is not “true” Coulomb friction as the coefficient is still dependent on normal
load, contact area, and velocity (as can be seen in Figure 5.5).

93
SAWE Paper No. 3602
Category Number 31.0

It is this cyclical stretching of the tread that gives the appearance of “slip”, which is defined
as 192:

%𝑺 = �𝑹𝝎�𝑽 − 𝟏 � × 𝟏𝟎𝟎% (EQ. 5.23)

The greater the traction force the tire is required to transmit, the greater the
“%S(tretch)”. And, just as with Coulomb friction, when a certain limit has been reached the
“static” (traction) condition quickly devolves into a “dynamic” sliding case. However, unlike
Coulomb friction, even the tire sliding coefficient varies with normal load, contact area,
velocity, and temperature (a condition of thermal equilibrium is assumed for the purposes of
this paper, so temperature is not dealt with).
In Figure 5.6, for the conditions of acceleration and braking, the vertical force “Fz”
(the resultant of the contact area pressure distribution and equal to “N”) times the offset arm
“d” constitutes the rolling resistance 193.

Figure 5.6 – CONTACT PRESSURE DISTRIBUTION: ACCELERATION, BRAKING

192
SAE J670 2008, pg. 39. In the braking case the wheel seems to be turning slower than the velocity which
would result in: “V > Rω” or “%S” being negative. For acceleration the opposite, “V < Rω”, is true, so the
“%S” is then positive. Note that there are variations on the “%S” formulation; see Reference [52], pages 39-41.
193
Reference [93], pp. 18 and 25. Note that the vertical offset “d” is depicted as being smaller in acceleration.

94
SAWE Paper No. 3602
Category Number 31.0

Because the longitudinal force “Fx” and its shear distribution influences the pressure
distribution, the matters of traction and rolling resistance are intertwined as shown in Figure
5.7, with rolling resistance dropping to a minimum under light acceleration 194.

Figure 5.7 – ROLLING RESISTANCE vs. ACCELERATION, BRAKING

Up to now the “braking” subject matter has concerned itself mainly with the behavior
of the tires, not the brakes themselves. This is because, to paraphrase Colin Campbell,
regardless of the effectiveness of the braking mechanism, “the upper limit on the rate of
retardation is always set by the grip of the tyres on the road” 195. Or, as Richard Radlinski
of the NHTSA put it: “If a brake is ‘big’ enough to lock a wheel, (then) the issue of
stopping …focuses on tire properties and not the brake…” 196. The science of brake design
is well established, and for the purpose of comparing the conventional four wheeled
configuration to the eight wheeled configuration comparatively little need be said with regard
to the actual braking mechanism(s):
1) All eight wheels of the “advanced design” are to be equipped
with brakes equal in design sophistication to the brakes of the
baseline configuration and sized appropriately for the given
application.

194
Reference [26], pg. 497. Figure 12.15 of source agrees with the above Figure 5.7, but similar figures on page
115 of Reference [4] and page 16 of Reference [93] show the minimum rolling resistance as occurring under
braking. However, Reference [26] is the most authoritative in this regard, and that source says “…minimum
rolling resistance does not occur under freely rolling conditions, but rather under a small driving torque.”.
Regardless, the effect is minor, and in general both increasing acceleration and increasing deceleration torques
cause increasing rolling resistance.
195
Reference [12], pg. 245.
196
Reference [65], pg. 1.

95
SAWE Paper No. 3602
Category Number 31.0

2) Just as the four wheel design required proper brake


proportioning between front and rear axles, there must now
be a proper proportioning between four axles to avoid “lock
up”.

In the past, proportioning was accomplished by mechanical means at a fixed


compromise setting of 60/40 to 65/35 197. Later, more modern designs involved hydro-
mechanical variation of the brake proportioning during deceleration. However, now all that is
obsolete due to the prevalence of automatic, generally electronic, anti-skid braking systems
(ABS) that simply won’t allow lockup. It is thus postulated for this analysis that both the
baseline and the “advanced concept” are adequately equipped in this regard, and that both are
braking at an optimal rate given the normal load and traction coefficient, hence the
“Maximum Braking” relationships shown initially within Figure 5.4 198:

𝑻𝒇
𝒇𝒇 = 𝑵𝒇 𝝁𝒇 = �𝑹 (EQ. 5.24)
𝒇

𝑻𝒓
𝒇𝒓 = 𝑵𝒓 𝝁𝒓 = �𝑹 (EQ. 5.25)
𝒓

The “advanced concept” brakes were initially to be resized in accord with the other
rolling masses inherent in “I1” and “I4”. Ultimately, however, as a practical concession the
brakes were held to be the same diameter as the baseline (so the total friction area is doubled,
not halved, obviating any wear/fade concerns). Also, the same hydraulic pressure results in
the same retarding torque about the axle. However, the torque divided by the “half-size”
rolling radius tends to result in a retarding ground contact force twice the original case, so the
brakes need not be applied as hard to achieve the same effect.
Now that the particular governing relations of the braking case have been established,
the natural presumption (of some, anyway) would be to take all the partial relations from Eq.
5.7 to Eq. 5.25 and, through substitution into the basic energy relation Eq. 5.6, develop one
immense and very complicated general equation. Then, by input of the appropriate variable
values, that general equation could be solved for the comparison of the braking performance
of the “baseline” vs. the “advanced concept”.
However, that is not possible; such a braking performance general equation would
prove to be transcendental in nature, not admitting to a direct solution. However, the braking

197
This was based on the belief that if any set of wheels were to “lock up” it would be best if it were the front
set. That way steering would be lost, but at least the vehicle would tend to slide straight ahead and the driver
could see what he was about to hit.
198
Reference [28], pg. 53. Gillespie initially presents the equation as “f = (T - Iwαw)/R” but relents as the wheel
rotational inertia “Iw” may be “…simply lumped in with the vehicle mass for convenience in calculation. In
that case the torque and brake force…” relation is reduced to “f = T/R”. It is this assumption of perfect brake
proportioning which makes the functioning of “MAXDLONG.BAS” possible.

96
SAWE Paper No. 3602
Category Number 31.0

performance problem, as with the acceleration performance problem before it, does admit to a
numerical or iterative solution. To that end Equations 5.6 through 5.25 were utilized to create
a computer simulation of automotive braking performance. The listing of that program,
entitled “MAXDLONG.BAS”, is provided in Appendix C, along with the input/output values
for the “baseline” and “advanced concept” runs. Again, the absolute accuracy of the program
output is relatively unimportant; what is significant is the relative veracity of the “baseline”
vs. the “advanced concept” results for comparative study.
Although, as stated, the “absolute accuracy of the program output is relatively
unimportant”, the “MAXDLONG.BAS” program can’t be totally capricious in its operation.
To ascertain that the program is at least somewhat reasonable in its output, a comparison of
calculated vs. empirical would be indicated. However, the only “hard data” obtainable
regarding the braking performance of the baseline 1958 Jaguar XK150S was the 30-0 mph
(48.3-0 kph) distance of 33 ft (10 m) reported earlier 199. To “flesh out” the data comparison
the maximum allowable stopping distances for passenger cars as embedded in the present UK
Highway Code (seemingly representing a 1950’s level of performance!) were also plotted
with the following result:

Figure 5.8 – “MAXDLONG.BAS” PROGRAM “VALIDATION”

199
The “MAXDLONG.BAS” simulation of the Jaguar 30-0 mph braking returned a stopping distance value of
39.7 ft (12.1 m), which would have been fine in the 1950’s. The quoted 33 ft (10 m) test distance, if correct,
would indicate that the Jaguar truly was special, more in accord with 1960’s expectations.

97
SAWE Paper No. 3602
Category Number 31.0

Even though the 1958 Jaguar XK150S was considered an astounding performer for its
day, it is not terribly surprising that the less than demanding requirements of the UK Highway
Code should almost exactly overlap the plot of the simulated 1958 braking performance
(which also suggests the simulation was somewhat conservative). To add perspective, there is
a lower curve which represents what can be expected of the improved technology of more
modern vehicles as per an April 2011 edition of Consumer Reports 200. Essentially there are
just two data sets (plots) which can be regression fitted with 2nd order polynomial curves (and
with an apparent high degree of accuracy 201); the resulting parabolic equations are displayed
near the corresponding data set plots.
In accord with the goal of obtaining a comparison of baseline versus “advanced
concept” braking performance the “MAXDLONG.BAS” simulation was employed to
generate “advanced concept” stopping distance data; the results were regression modeled and
plotted as follows 202:

Figure 5.9 – “BASELINE” vs. “ADVANCED CONCEPT” BRAKING


PERFORMANCE

200
No more detailed citation can be give as the information was gleaned from an internet website not designed to
provide that sort of reference data.
201
That parabolic models should so closely fit the brake distance data should not seem remarkable given the
essence of the situation (“stopping distance ∝ ⅟2 m V2”).
202
Note that, due to some instability in the “MAXDLONG.BAS” program (characterized by oscillating
deceleration values) the “advanced concept” stopping distances for velocities above 80 mph (129 kph) were
determined by extrapolation using the fitted parabolic curve.

98
SAWE Paper No. 3602
Category Number 31.0

As was the case with longitudinal acceleration performance, the “advanced concept”
proves itself to be better than the “baseline”, although not overwhelmingly so. What
improvement there is may be mainly attributed to the increased tire-to-road contact area and
reduced effective mass of the more radical vehicle. As noted earlier, such modest
improvements might not be considered worth the effort of designing and manufacturing
something as radical as the “advanced concept” if such improvements were the only
benefits…

99
SAWE Paper No. 3602
Category Number 31.0

6 – LATERAL PERFORMANCE: STEADY STATE/TRANSIENT

“…by designing a vehicle to accommodate tires of smaller


overall diameter, the c.g. height is lowered and SSF is increased;
this approach would seem to be impractical because of
performance, ground clearance, and marketing issues.”

Joseph D. Walter 203

The “Static Stability Factor”, or SSF, is only one of the many ways that tires can
affect lateral performance. Steady state lateral acceleration (aymax, roll gain, yaw angle,
over/under steer)/directional stability (understeer coefficient, characteristic/critical
speeds, steer angle, static margin) and transient lateral response (rise/decay time,
oscillation center) are some more of the ways lateral performance is influenced by the tires.
And for all of these ways mass properties plays a major role.

The SSF is a mathematically derived figure of merit indicating vehicle resistance to


those lateral forces that may cause rollover. Although rollover is a relatively unlikely mode
of automotive accident, it still merits serious concern because when it does occur then there is
a disproportionately high chance of fatality or serious injury. The NHTSA used to (NCAP
2001-2003) rate new vehicles for rollover resistance based solely on SSF 204, but more
recently the rollover resistance rating has been modified to include empirical test results 205.
However, the SSF is still the most basic and ready means of comparison between vehicles:

𝒕
𝑺𝑺𝑭 = (EQ. 6.1)
𝟐𝒉𝒄𝒈

Where:

SSF = A figure numerically equal to the lateral acceleration for overturn (in

“g’s”), simplistically calculated without regard for roll, etc.

203
Reference [26], pg. 609.
204
Reference [90], pg. 35.
205
However, the appropriate vehicle test, which is known as the “Fishhook”, has been in use only sparingly since
2004 due to its expense. Most vehicles receive their “Fishhook” ratings based on “similarity” to those few other
vehicles which actually have been tested. Other than the SSF and the “Fishhook”, the NHTSA may give some
credit in its rollover rating for the presence of electronic “active rollover protection” (ARP).

100
SAWE Paper No. 3602
Category Number 31.0
Rev C

t = The average vehicle track width, front plus rear divided by two.

hcg = The vehicle center of gravity height above the ground plane.

As the SSF constitutes the simplest mode of lateral performance comparison between
the “baseline” 1958 Jaguar XK150S and the “advanced concept” based upon that XK150S, it
will be the first addressed; the comparison works out as follows (all dimensions are in inches):

Table 6.1 – SSF: “BASELINE” vs. “ADVANCED CONCEPT” (RUN #9)

That the “advanced concept” should score about +14% better than the “baseline” is
encouraging, but that needs to be put into perspective. Per the NTSHA, passenger cars in the
United States averaged an SSF of 1.38 in 1978, but by 2003 averaged 1.41 206. Therefore, the
SSF scores for the “baseline” and the “advanced concept” are not particularly praiseworthy by
modern “standards”; however, for the purposes of this comparison study only the +14%
improvement is relevant.

The SSF is numerically equal to the lateral acceleration in “g’s” that a vehicle can be
subjected before rollover (as calculated without consideration of body roll and other dynamic
effects). Another measure of lateral performance is the steady state level of acceleration a
vehicle can be expected to attain before slide sets in; this also can be quasi-statically estimated
in a very simplified way (no roll consideration) by the following equation 207:

𝟐𝒎𝑾𝒉𝒄𝒈 𝟐 𝒎𝑾
−𝟏 +�𝟏−𝟒� ��� 𝟐 �−𝒃�
𝒕𝟐
𝒂𝒚𝒔𝒍𝒊𝒅𝒆 = (EQ. 6.2)
(𝟒𝒎𝑾𝒉𝒄𝒈 𝟐 ⁄𝒕𝟐 )

Where the symbolism is:

206
Reference [86], pg. 14. Of course, this apparent improvement over time is largely illusionary. Vehicles such
as SUV’s, pickup trucks, vans, and mini-vans have increasingly come to be used as “passenger cars” in the 1978-
2003 time period, and if the low SSF scores of such vehicles were to be included with those of traditionally
configured passenger cars then the result would be far from improved.
207
Reference [90], pg. 35. The equation is very optimistic; roll effects can considerably decrease lateral
acceleration capability due to longitudinal load transfer, camber effects, etc. However, because roll effects are
not taken into account, it is a good indicator of low speed (no roll) handling (initial oversteer or understeer).

101
SAWE Paper No. 3602
Category Number 31.0
Rev C

𝒂𝒚𝒔𝒍𝒊𝒅𝒆 = Lateral acceleration limit, slide (g’s).

b = The basic tire coefficient of traction potential (dimensionless).


m = The rate of tire traction potential decrease with load (lb-1).
W = The vehicle weight at the axle considered (lb).
hcg = The vehicle height of the center of gravity above the ground plane (in).
t = The track width of the axle considered (in).

Because there are two axles (the “advanced concept” in its equivalent conventional
configuration has just two axles) Equation 6.2 has to be calculated twice, once for each axle
(instead of just once using averaged data values as was the case for Equation 6.1). This is not
just because each axle will generate different “g” values (with the lowest value constituting
the limit), but because the relationship between the two axle values will provide further
fundamental lateral performance information. For the “baseline” and the “advanced concept”
(equivalent conventional configuration) the lateral acceleration limits per Equation 6.2 were
determined to be as in the following table:

Table 6.2 – LATERAL ACCELERATION: “BASELINE” vs. “ADVANCED


CONCEPT” (RUN #9)

According to this simplified analysis, the “baseline” vehicle will only be able to reach
about 0.70 g’s before sliding off the turn radius rear end first. However, the “advanced
concept” will pull 0.83 g’s before it also heads into the boondocks, but front first. This latter
behavior is considered characteristic of an “understeering” (US) vehicle, while the former is
an “oversteering” (OS) characteristic. The difference between these two modes of lateral
behavior is more than just a matter of giving the driver a clear view of what he is about to
crash into; the “understeering” characteristic is associated with directional stability, and
“oversteering” with a lack of stability. However, these results can only be regarded as
indicative at best...

102
SAWE Paper No. 3602
Category Number 31.0
Rev C
Directional stability is a more complex matter than can be dealt with by just one
simple formulation such as Equation 6.2 208. The significance of the results as displayed in
Table 6.2 is that the “advanced concept” displays almost a +19% improvement over the
“baseline” with regard to maximum lateral acceleration level obtainable, and is initially
understeering while the “baseline” is initially oversteering. These changes are attributable to
the increased contact area and lateral stiffness of the “advanced concept” tires, the removal of
a front-to-rear inflation pressure differential, plus the reduced VCG and the increased track.

One of the ways the analysis was simplified was that it did not include a consideration
of the effect of vehicle roll. The “baseline” vehicle has a “double wishbone” IFS with a “live”
(driven) beam suspension in the rear. The front and rear suspension spring rates (at the wheel)
are reputed to have been 106.4 lb/in (19.00 kg/cm) and 128.8 lb/in (23.00 kg/cm)
respectively 209, with corresponding track widths of 51.75 in (131.4 cm) front, 51.25 in (132.7
cm) rear, and a rear “spring base” of 38.0 in (101.9 cm) 210. The “baseline” vehicle also had a
front anti-roll bar (ARB) of approximately 84.7 lb-ft/deg (114.8 Nm/deg) stiffness; this anti-
roll bar roll stiffness is directly additive to the front axle roll stiffness as calculated per the
following formulae for an independent suspension 211:

𝒌𝒄𝒔 𝒕𝟐
𝒌𝒓𝒐𝒍𝒍 = 𝟏𝟑𝟕𝟓 (EQ. 6.3)

Where:

kroll = Independent suspension roll stiffness, lb-ft/degree.


kcs = Combined stiffness of suspension spring at the wheel + tire, lb/in.
t = The track width at the axle line, in.

208
It has been stated (source unknown) that vintage vehicles of the 1958 Jaguar type often tended to exhibit
slightly “oversteering” behavior initially, but as speed increased became definitely “understeering”. That the
Jaguar may have had some inherent initial oversteer is evidenced by the fact that the manufacturer specified a
higher inflation pressure for the rear tires. Reference [28], page 354 notes: “…it was common practice in the
past to control the low-g directional behavior of passenger cars through the specification of different inflation
pressures for front and rear tires…”. In any case, the fact is that the directional stability of any vehicle can be
affected by speed, loading, and turn radius. Stability can vary in degree, and even change completely in character
(“understeer” to “oversteer”, and vice versa) during the course of operation, if proper design diligence has not
been observed (race cars being an exception, as some “designed-in” variations in stability can be advantageous
given a skilled driver).
209
Reference [91], pg. 5.
210
A beam axle has two “spring bases”; one is the track and the other is the lateral distance along the axle
between the suspension spring mounting points; for an independent suspension the “spring base” is just the track.
211
Reference [52], page 603. Note that the combined tire and suspension spring rate “kcs” would be calculated
using the formula for springs in series: “kcs = (kt × ks)/(kt + ks)” (as opposed to the formula for springs in
parallel: “kcs = kt + ks”).

103
SAWE Paper No. 3602
Category Number 31.0
Rev C

For a non-independent suspension, the roll stiffness at that axle line is 212:

�𝒌𝒔 𝒕𝒔𝒃 𝟐 ��𝒌𝒕 𝒕𝟐 �


𝒌𝒓𝒐𝒍𝒍 = (EQ. 6.4)
𝟏𝟑𝟕𝟓�𝒌𝒔 𝒕𝒔𝒃 𝟐 +𝒌𝒕 𝒕𝟐 �

Where:

kroll = Non-independent suspension roll stiffness, lb-ft/degree.


ks = Effective suspension spring stiffness at the mount to axle, lb/in.
tsb = “Spring base” distance between spring mounts along the axle, in.
kt = The spring stiffness of a tire on the axle, lb/in.
t = The track width at the axle, in.
The “baseline” and “advanced concept” parameter values for the determination of the
front and rear roll stiffnesses are as per the following table:

Table 6.3 - TIRE + SPRING STIFFNESS (Kcs), ANTI-ROLL BAR (ARB) STIFFNESS,
TRACKS (t), SPRING BASE (tsb), SPRUNG (Ws) & UNSPRUNG (Wus) WEIGHTS,
ROLL CENTERS (RC)

This table contains enough information to determine that the “baseline” vehicle has
front and rear axle roll resistances of 276.6 lb-ft/deg (375.0 Nm/deg) and 167.8 lb-ft/deg
(227.5 Nm/deg), respectively. The corresponding values for the “advanced concept” calculate

212
Reference [52], pg 604. Note that here the effective suspension spring stiffness “at the mount” to the axle, not
the more common “at the wheel”, is required. Some “reverse” calculation will be required to get the effective
spring stiffness “at the mount” values if only “at the wheel” spring rates are known.

104
SAWE Paper No. 3602
Category Number 31.0
Rev C
to be 297.9 lb-ft/deg (403.9 Nm/deg) and 187.4 lb-ft/deg (254.1 Nm/deg). The “roll gain”
values can now be determined by iterative solution (varying “θs”) of the roll equation 213:

θs kroll = Ws ay hr Cos θs + Ws hr Sin θs + Wusf ay rrf + Wusr ay rrr (EQ. 6.5)


Where:

θs = The sprung mass roll angle, degrees.


kroll = The total vehicle roll resistance, lb-ft/deg.
Ws = The weight of the sprung mass, lb.
ay = The lateral acceleration, g’s.
hr = The sprung mass roll moment arm, ft.
Wusf = The front axle unsprung mass weight, lb.
rrf = The front axle unsprung mass vertical c.g.
(approx. the rolling radius 214), ft.
Wusr = The rear axle unsprung mass weight, lb.
rrr = The rear axle unsprung mass vertical c.g.
(approx. the rolling radius), ft.

The roll moment arm “hr” is the difference between the sprung mass c.g. height and
the roll axis height immediately under the c.g., times the cosine of the angle “ϕ” the roll
axis makes with the horizontal (this last is a correction which is sometimes neglected). The
resulting roll angle from a particular lateral acceleration value establishes a vehicle’s “roll
gain” (“θs/ay”) 215. The necessary “hr” value is obtainable through a study of the vehicle front
and rear suspension geometry, which for the “baseline” and “advanced concept” vehicles may
be as depicted per the following figure 216:

213
Reference [90], pg. 17.
214
The rolling radius is commonly taken as a reasonable approximation of the vertical c.g. of the unsprung
mass.
215
The “roll gain” itself varies slightly with roll, and therefore for comparisons the best values to use are those
calculated or measured at the same lateral acceleration value for both vehicles.
216
Reference [28], page 263 for the RC of the Hotchkiss rear suspension, and page 266 for the RC of the
“negative” double swing arm front suspension. In 1969 the Buick division of General Motors introduced such a
“negative” system, termed “Tru-Track”, on the LeSabre, Wildcat, and Electra models claiming improved
directional stability (Machine Design, 10 October 1968, pg. 37).

105
SAWE Paper No. 3602
Category Number 31.0
Rev C

Figure 6.1 – “BASELINE” AND “ADVANCED CONCEPT” SUSPENSION


GEOMETRY

The “baseline” IFS utilizes a torsion bar type spring which acts at the inboard pivot
point of the lower suspension arm, and the front ARB is not depicted in the illustration for
clarity. The “advanced concept” (in its “equivalent conventional configuration” form)
suspension geometry was derived from that of the “baseline” with only such change as
absolutely necessary. Again, this was done to minimize the effects of “non-mass properties”
related changes obscuring the result of what is intended to be a purely mass properties study.

The roll gain values may now be obtained, which for the “baseline” vehicle turns out
to be 15.6 deg/g, and for the “advanced concept” 14.3 deg/g. That that the “advanced
concept” should have a lower roll gain than the “baseline” is the consequence of its
significantly increased total roll resistance: 485.3 lb-ft/deg (658.0 Nm/deg) vs. 444.4 lb-ft/deg
(602.5 Nm/deg) 217.

217
The roll moment arm values, corrected to account for the angle of the roll axis with respect to the horizontal,
of both vehicles are nearly equal, with the “baseline” at 20.66 in (0.5248 m) and the “advanced concept” at 20.99
in (0.5331 m).

106
SAWE Paper No. 3602
Category Number 31.0
Rev C
To lend some perspective to these roll gain results, they are plotted on a graph that
already displayed the roll gain character of some modern (c. 1984) sporting vehicles, along
with a very prosaic type modern passenger car (c. 1980) 218:

Figure 6.2 – ROLL GAIN RESULT COMPARISON

Although the “baseline”, and “advanced concept” derived therefrom, would seem to
have to be considered “sports” vehicles as that was the nature of the 1958 Jaguar XK150S, it
can be seen from the above graph that, at least with respect to roll gain, they have little in
common with modern sports cars. Instead, the “baseline”/”advanced concept” vehicles have a
roll gain more like that of the prosaic modern passenger car, a 1980 Ford Fiesta S; this
constitutes a testimony as to how far vehicle design has progressed since 1958. However,
again it is only the comparison between the “baseline” and the “advanced concept” that
matters for the purposes of this paper.
218
The Ford Fiesta S roll gain value of 9.2 deg/g is this author’s calculation (Reference [90], pg. 21); Colin
Campbell presents a figure of 7.8 deg/g (Reference [13], pg. 192) for the same vehicle.

107
SAWE Paper No. 3602
Category Number 31.0
Rev C

The “advanced concept” lateral acceleration capacity was earlier shown to possibly be
almost 19% greater than that of its “baseline” parent (0.83 vs. 0.70 g’s), and the roll gain
comparison is supportive of that result. However, a more detailed study of the lateral
acceleration capacity that is inclusive of some of the complications resulting from vehicle roll
(which are not accounted for in Equation 6.2) may be made. This is accomplished by
calculating the normal load, and the consequent lateral traction potential, at each tire as lateral
acceleration is incrementally increased. First, the roll angle “θs” for a particular lateral
acceleration “ay” is easily estimated now that the roll gain (“RG”) values are known 219:

𝜽𝒔 = 𝒂𝒚 × 𝑹𝑮 (EQ. 6.6)

With the roll angle known the normal loads may be determined 220:

𝑾𝒇 𝒕𝒇
𝑵𝒊𝒇 = − 𝜽𝒔 𝒌𝒇𝒓𝒐𝒍𝒍 / � � (EQ. 6.7)
𝟐 𝟐𝟒

𝑾𝒇 𝒕𝒇
𝑵𝒐𝒇 = + 𝜽𝒔 𝒌𝒇𝒓𝒐𝒍𝒍 / � � (EQ. 6.8)
𝟐 𝟐𝟒

𝑾𝒓 𝒕
𝑵𝒊𝒓 = − 𝜽𝒔 𝒌𝒓𝒓𝒐𝒍𝒍 / � 𝒓 � (EQ. 6.9)
𝟐 𝟐𝟒

𝑾𝒓 𝒕
𝑵𝒐𝒓 = + 𝜽𝒔 𝒌𝒓𝒓𝒐𝒍𝒍 / � 𝒓 � (EQ. 6.10)
𝟐 𝟐𝟒

From which the lateral traction potential at each axle may be determined:

𝑭𝒇 = �𝒃 − 𝒎𝑵𝒊𝒇 �𝑵𝒊𝒇 + �𝒃 − 𝒎𝑵𝒐𝒇 �𝑵𝒐𝒇 (EQ. 6.11)

𝑭𝒓 = (𝒃 − 𝒎𝑵𝒊𝒓 )𝑵𝒊𝒓 + (𝒃 − 𝒎𝑵𝒐𝒓 )𝑵𝒐𝒓 (EQ. 6.12)


Each of the axle traction potentials is then to be compared with the centrifugal force
each axle is required to resist if the vehicle is to maintain its course around the skidpad 221:

𝑪𝒆𝒏𝒕𝒓𝒊𝒇𝒖𝒈𝒂𝒍 𝑭𝒐𝒓𝒄𝒆, 𝑭𝒓𝒐𝒏𝒕 𝑨𝒙𝒍𝒆 = 𝑾𝒇 𝒂𝒚 (EQ. 6.13)

𝑪𝒆𝒏𝒕𝒓𝒊𝒇𝒖𝒈𝒂𝒍 𝑭𝒐𝒓𝒄𝒆, 𝑹𝒆𝒂𝒓 𝑨𝒙𝒍𝒆 = 𝑾𝒓 𝒂𝒚 (EQ. 6.14)

219
Again, use of these roll gain values result only in approximations as the roll gain values are dependent on a
particular vehicle/suspension orientation (CG, IC, RC locations, etc.) which will change with roll.
220
Reference [90], pp. 17-20.
221
Reference [28], pg. 199-201.

108
SAWE Paper No. 3602
Category Number 31.0
Rev C
These equations make possible a more realistic analysis of vehicle behavior when
undergoing steady state lateral acceleration. That analysis shows that the “baseline” maximum
lateral acceleration is about 0.56 g’s and the “advanced concept” maximum lateral
acceleration is about 0.71 g’s, which is about 27% greater. The analysis also indicates that the
“baseline” initial oversteer becomes terminal understeer, and that the “advanced concept”
initial understeer becomes terminal oversteer. However, those indications have yet to involve
consideration of the “slip angle” functions…

The “roll angle” is only part of the “attitude” that a vehicle adopts when undergoing
lateral acceleration in a steady state turning maneuver, there is also the “yaw angle”. The yaw
angle (“β”) of a vehicle is a function of the relationship between the “slip angles” (“ψ”)
adopted by the tires at the front vs. those at the rear. To make a determination of those “slip
angles” requires specific tire information such as the following 222:

Figure 6.3 – 6.00×16 LATERAL LOAD, NORMAL LOAD, & “SLIP ANGLE”

The relationship depicted in Figure 6.3 is also affected by speed, temperature, and tire
inflation pressure, but speed has a relatively minor (caution is indicated for the “half-radius”)
effect and the tires will be assumed to have reached thermal equilibrium. As for the effect of
inflation pressure on the “cornering power” (“lby/deg”): it would typically vary about
plus/minus “2 lb/deg/psi” around the “normal” inflation pressure (per Colin Campbell) 223.

222
Reference [12], pg. 165. “Slip angle” is a misnomer; a better term might be “drift angle”.
223
Ibid. “Cornering Power” is yet another misnomer; a more correct term for the slopes of the near-linear
portions of “Fy, ψ” curves is the term “Cornering Stiffness” (“CS”), which is in fairly common usage.

109
SAWE Paper No. 3602
Category Number 31.0
Rev C

The “advanced concept” tires are very low profile; so a more suitable (high cornering
stiffness) chart for the “half radius” tire than the previous had to be found to approximate the
behavior of such tires 224:

Figure 6.4 – 215/60R15 LATERAL LOAD, NORMAL LOAD, & “SLIP ANGLE”

At the same lateral acceleration of 0.384 g’s the “baseline”/“advanced concept” loads
and “slip angles” for the individual wheels are estimated from Figures 6.3-6.4 to be 225:

Table 6.4 – NORMAL (Fz)/LATERAL (Fy) LOADS & “SLIP ANGLES” (ψ) @ .384 G’s

224
Reference [52], pg. 29. Admittedly, this chart is not as desirable as one would like for this purpose; charts for
152/46R8 tires are understandably rare.
225
Reference [53], pg. 257, states that “ψf > ψr” at just one point (“ay”) is not proof of general understeering
behavior (but is indicative at that point…BPW).

110
SAWE Paper No. 3602
Category Number 31.0
Rev C

The lateral acceleration value of 0.384 g’s was used for this comparison as this is a
good point for a fair comparison that favors neither configuration. The maximum lateral
accelerations of 0.56 and 0.70 g’s represent significant reductions (-20% and -16%) from the
0.70 and 0.83 g’s obtained with the simpler method of Equation 6.2 226.
A “bicycle model” analysis can quickly be made utilizing the average “slip angles”
front and rear to calculate the respective “yaw” angles 227 of the vehicles at 0.384 g’s lateral
acceleration while on a 100 ft (30.5 m) radius skidpad. The formulae for the steer (“δ”) and
yaw (“β”) angles are 228:
𝑳𝒘𝒃�
𝜹 = 𝟓𝟕. 𝟑 𝑹 + 𝛙𝒇 − 𝛙𝒓 (EQ. 6.15)
𝒍
𝜷 = 𝟓𝟕. 𝟑 𝒓�𝑹 − 𝛙𝒓 (EQ. 6.16)
Where:
δ = The steering angle (degrees).
β = The yaw angle at the vehicle CG (degrees).
Lwb = The vehicle wheelbase (in, cm, etc.).
lr = The longitudinal distance from CG to rear axle (in, cm, etc.).
R = The turning radius (in, cm, etc.).
Ψf = The average front axle “slip angle” (degrees).
Ψr = The average rear axle “slip angle” (degrees).
Use of these equations and information from Tables 6.3 and 6.4 for a 100 ft (30.48 m)
radius (“R”) skidpad for the “baseline” vehicle results in:

δ = 57.3 (102/1200) + 4.82 – 3.05 = 6.64 deg


β = 57.3 (48.99/1200) – 3.05 = -0.71 deg
For the “advanced concept” the results are:

δ = 57.3 (99.33/1200) + 0.80 – 0.69 = 4.85 deg


β = 57.3 (53.41/1200) – 0.69 = 1.86 deg

226
Reduction in the indicated maximum lateral acceleration level is to be expected when roll effects are
considered, and changes in oversteering/understeering behavior are possible with variation in “speed” (lateral
acceleration).
227
The term “vehicle side slip angle” seems to be preferred to “yaw angle” by the SAE, possibly because “β” is
constant during steady-state, but there is still an angular yaw velocity “ω” (differing reference systems used!).
228
Reference [28], pp. 202 and 208. These formulae are simply derived from the steering geometry.

111
SAWE Paper No. 3602
Category Number 31.0
Rev C

Alternatively, use of the “bicycle model” for each vehicle (where the front and rear
“slip angles” are averages of the front and rear angles from Table 6.4 229), plus the fact that the
velocity vectors at each key point are at 90 degrees to the turning radius at the point, allows
for determination of all the angles involved via the “Law of Sines” 230:

Figure 6.5 – STEADY STATE TURNING GEOMETRY CALCULATIONS

Now it is possible to illustrate (also not to scale) the comparative turning situations of
the two vehicles at 0.384 g’s on a 100 ft (30.5 m) skidpad:

229
The sign convention of SAE J670 2008 (“FIGURE 3”) on page 17 indicates the “β” angle is negative when
the turning vehicle velocity vector “V” (through the CG) lies to the “outside” of the vehicle centerline...
230
Reference [52], pg. 131. Obviously Figure 6.5 is not drawn to scale as then the “TC’s” would be well off the
page, so the angles depicted are only representative (leave your protractor in the desk!).

112
SAWE Paper No. 3602
Category Number 31.0
Rev D

Figure 6.6 – STEADY STATE YAW: BASELINE vs. ADVANCED CONCEPT

The directional stability of both configurations seems assured, but the “baseline”
vehicle had an inherent initial tendency to oversteer which its designers countered with higher
tire inflation pressures at the rear, greater roll resistance at the front, and a roll axis steeply
inclined nose down to produce more lateral load transfer at the front 231. The result was an
understeering character which held all the way to maximum lateral acceleration. The
“advanced concept” displayed an initial tendency to understeer which remained through 0.384
g’s and beyond but may change at some point to terminal oversteer…Luckily there are means
to investigate this further…

The “Understeer Gradient” (a.k.a. “Understeer Coefficient”, a misnomer),


symbolized as “Kus” and expressed in radians or degrees per “g”, presents a more general way
of classifying a vehicle’s stability condition than comparing specific drift (“slip”) angle values
“ψf” and “ψr”. If a vehicle is stable (understeering) then “Kus > 0”, if neutral (NS) then “Kus =
0”, or if unstable (oversteering): “Kus < 0”. This coefficient can take various forms 232, of
which one of the simpler is:

𝑾𝒇 𝑾𝒓
𝑲𝒖𝒔 = � − � (EQ. 6.17)
𝒈 𝑪𝒔𝒇 𝒈 𝑪𝒔𝒓

To determine “Kus” for the “baseline” and “advanced concept” vehicles use the axle
weight loads “Wf, Wr” and the axle (twice the individual tire values) cornering stiffness “Csf,
Csr” values as shown in Tables 6.2 and 6.4. The resulting “Kus” values for “baseline” and
“advanced concept” are 7.38 deg/g and 0.39 deg/g respectively, which indicates that both

231
Other factors (camber thrust, roll steer, lateral compliance steer, etc.) associated mainly with the IFS, but not
dealt with in this simplified study, would tend to further enhance the understeering characteristic.
232
Reference [28], pg. 203. A much more sophisticated treatment is presented in Reference [52], pp. 161-164.

113
SAWE Paper No. 3602
Category Number 31.0
Rev D

vehicles are understeering. However, the “baseline” vehicle is much more understeering than
the “advanced concept”, and one of the consequences of that situation may be seen through a
study of Equation 6.15:
𝑳𝒘𝒃�
𝜹 = 𝟓𝟕. 𝟑 𝑹 + 𝛙𝒇 – 𝛙𝒓
This expression may be rewritten:

𝑳𝒘𝒃� 𝑾𝒇 𝑾𝒓
𝜹 = 𝟓𝟕. 𝟑 𝑹 + � �𝑪𝒔𝒇 − �𝑪 � 𝒂𝒚
𝒔𝒓

Where “Kus” may be substituted for “[Wf/Csf-Wr/Csr]” and “V2/Rg” may be


substituted for “ay”:
𝑳𝒘𝒃� 𝑽𝟐�
𝜹 = 𝟓𝟕. 𝟑 𝑹 + 𝑲 𝒖𝒔 𝑹𝒈 (EQ. 6.18)

Equation 6.18 relates how the steering angle “δ” is affected by the wheelbase “Lwb”,
turn radius “R”, Understeer Gradient “Kus” (and all which comprises it), and velocity “V”.
Since the “baseline” has the relatively large “Kus” value of 7.38 deg/g it is apparent that any
small change in velocity will require a fairly large correction at the steering wheel in order to
keep the vehicle on course around the skidpad. However, the “advanced concept” vehicle,
which appears close to being neutral in behavior with its small “Kus” value of 0.39 deg/g,
requires only the mildest of steering corrections to stay on course.
The “Characteristic Speed”, provides a means of comparing designs with respect to
the degree of understeer present 233. Mathematically, it is the speed at which the steering
angle “δ” to make a turn of radius “R” 234 is equal to “2lwb/R” (that is, the steering angle is
twice that determined by the low speed “Ackermann” 235 steering geometry “lwb/R”):

𝟓𝟕.𝟑 𝒈 𝒍𝒘𝒃
𝑽𝒄𝒉𝒂𝒓 = � (EQ. 6.19)
𝑲𝒖𝒔

For the “baseline” vehicle “Vchar” works out to 31.5 mph (50.7 kph), while the
“advanced concept” works out to be 134.3 mph (216.2 kph). This indicates that the “baseline”
vehicle is extremely oversteering; it has been stated that traditional highly oversteering

233
Ergo, both vehicles must be essentially understeering.
234
The turn radius “R” is measured from the turn center “TC” to the vehicle CG.
235
The credit for development of this type of steering geometry would more properly be given to Erasmus
Darwin, Georg Lankensperger, or Charles Jeantaud; Rudolf Ackermann was merely an opportunistic swindler.

114
SAWE Paper No. 3602
Category Number 31.0
Rev C

“…North American cars have a characteristic speed as low as 40 mph…” 236. The “Vchar”
is the point at which steering response reaches a maximum for the understeering vehicle, and
goes to infinity (!) for an oversteering vehicle 237.
There is also a stability factor known as the “Critical Speed”, which is a speed at
which a vehicle can switch from a controllable initial condition (but still oversteering) to a
completely unstable (very oversteering) condition:

−𝟓𝟕.𝟑 𝒈 𝒍𝒘𝒃
𝑽𝒄𝒓𝒊𝒕 = � (EQ. 6.20)
𝑲𝒖𝒔

Note that the system of units conversion factor “57.3” is now negative; this is in
recognition of the fact that an oversteering vehicle will have a negative “Kus”. Since neither
vehicle in this study has a negative “Kus” this particular stability factor is not applicable
(“…there is no critical speed for an understeer car…a neutral steer car does not have a
real critical speed…” 238).
Comparison of the “Steady State Steering Angle”, as required at the front wheels to
make a steady state turn of radius “R”, is another measure of under/over steer:

𝒍𝒘𝒃
𝜹 = 𝟓𝟕. 𝟑 + 𝑲𝒖𝒔 𝒂𝒚 (EQ. 6.21)
𝑹

Note that this is yet an alternative way of writing Equation 6.15, just as Equation 5.18
was earlier. It has already been established (Figure 6.5) that for a steady state turn of 100 ft
(30.5 m) radius at 0.384 g’s the “baseline” to “advanced concept” steering angle comparison
is 6.64 deg to 4.85 deg, and that the “advanced concept” will require far less steering wheel
input for course correction. However, the above Equation 6.21 gives “Steady State Steering
Angle” results of 7.70 deg to 4.89 deg which again indicates the “advanced concept” is less
understeering (at 0.384 g’s constant lateral acceleration)…
Lastly there is “Static Margin”, which is a determinant like “Kus” for classifying a
vehicle’s stability condition as stable (understeering): “SM > 0”, neutral: “SM = 0”, or
unstable (oversteering): “SM < 0”. However, the SM is a dimensionless fraction with the

236
Reference [52], pg. 181.
237
Ibid, pp. 183-184.
238
Ibid, pg. 178.

115
SAWE Paper No. 3602
Category Number 31.0
Rev C

advantage of relating the LCG placement to the “Neutral Steer Point” (NSP) for that
condition (LCG – NSP = SM ˟ lwb) 239:

𝒍𝒓 𝑪𝒔𝒓 −𝒍𝒇 𝑪𝒔𝒇


𝑺𝑴 = (EQ. 6.22)
𝒍𝒘𝒃 �𝑪𝒔𝒇 +𝑪𝒔𝒓 �

The “baseline” SM is about 0.15 and the “advanced concept” SM is about 0.05.
Again both vehicles are identified as oversteering, but the “advanced concept” LCG is much
closer to the NSP than is the “baseline”.

To summarize the stability situation: the “baseline” vehicle by virtue of its slightly
aftward LCG placement (in the “test” configuration) is initially unstable, but this positioning
favored acceleration and braking performance. To ensure sufficient directional stability
throughout the operational envelope the “baseline” designers “pulled out all the stops”: a tire
inflation pressure differential with the higher pressure at the rear, the use of IFS with a beam
rear axle, a roll resistance differential with the greater resistance at the front, and the steeply
angled (12.28o) nose-down roll axis; all favored a high degree of understeer. This was
achieved, but at a loss in lateral acceleration and steering responsiveness that would be
unacceptable today. That the “advanced concept” is able to better the “baseline” acceleration
and braking performance, while impressively improving on that vehicle’s lateral acceleration
and responsiveness, is the result of its superior mass properties, but also of various collateral
changes (e.g., roll axis angle now 6.18o). The “collateral changes” occurred despite a serious
attempt to keep changes unrelated to mass properties at a minimum; it proved impossible to
achieve zero collateral changes given the drastic mass property design revisions.

So far the concern has been with steady-state turning performance, but every turning
action has three phases: initiation of the turn, turning, and termination of the turn. Only during
the middle phase can a steady state condition be achieved; the turn initiation and termination
are both transient states. On a winding road, when weaving in and out of traffic, or especially
on a slalom test course, a vehicle may spend more time in a transient state than in steady-state.
Therefore, the transient state is every bit as important as the steady-state, perhaps even more
so. However, a complete analysis of the transient state is more complex than that of the
steady-state, and is beyond the scope of this paper. However, it would be negligent to consider
a design paradigm change such as that proposed without at least a perfunctory look at the
transient performance consequences…

239
This is all very similar to aircraft stability theory wherein the relationship between the center of gravity and
the center of pressure is expressed in terms of percent of the “Mean Aerodynamic Chord” (MAC). Typical
passenger car “SM” values may range between +0.05 to +0.07 (Ref. 52, pg. 209).

116
SAWE Paper No. 3602
Category Number 31.0

Implicit throughout the turning discussion so far was the Dynamic Equilibrium force
balance for the vehicle wherein the tire traction forces at the axles resist the imaginary
centrifugal force vector through the vehicle CG:

𝑭𝒇 + 𝑭𝒓 = 𝒎 𝒂 (EQ. 6.23)
The question of exactly how much force did “Ff” or “Fr” represent was made moot by
reducing Equation 6.23 to an equivalent set of force balances, one at each axle, where the
vehicle mass “m” is replaceable by the front and rear axle weight loads “Wf” and “Wr” as
the lateral acceleration vector “ay” is now in gravity units.:

𝑭 𝒇 = 𝑾𝒇 𝒂 𝒚 (EQ. 6.24a)
𝑭𝒓 = 𝑾𝒓 𝒂𝒚 (EQ. 6.24b)
This is all so very simple because in the steady-state condition there is no additional
complication of a rotational acceleration “α” of the vehicle about the CG, which would
require the addition of a Dynamic Equilibrium moment balance:

𝑭𝒇 𝒍𝒇 − 𝑭𝒓 𝒍𝒓 = 𝑰 𝜶 (EQ. 6.25)
In the transient state both the translational force balance and the rotational moment
balance equations have to be satisfied simultaneously, so now it is a little more difficult to
establish what the necessary tire traction force values “Ff, Fr” will be in order to accomplish
the maneuver. Some of the difficulty rests with the need to first determine the exact value of
the rotational inertia to use in the determination of those force values. One might assume
that the vehicle would tend to rotate about its CG, in which case the appropriate rotational
inertia value would be “Iz”, the vehicle’s yaw inertia through the CG. If the vehicle were
floating in space, with no constraints upon its motion, this would be true, but for a ground
vehicle initiating or terminating a turn the initial center of rotation may be far away from the
vehicle CG. This initial point of rotation is known as the Oscillation Center “OC” 240; as the
vehicle progresses through the transient state the axis of rotation will migrate from the OC
until it coincides with the CG at the achievement of steady-state.
The OC is directly determinable from the vehicle dimensions and the vehicle
longitudinal center of gravity and yaw inertia “Iz”, or more specifically the yaw radius of
gyration “Kz” (“Iz = Kz2 m”). With the location of the vehicle LCG defined by the
dimensions “lf” and “lr”, the mass properties contribution to the transient response is neatly

240
Reference [4], pg.32. Reference [53], pg. 253, refers to this point as the “center of percussion”.

117
SAWE Paper No. 3602
Category Number 31.0
Rev C

encapsulated by the quantity “Kz2/(lf x lr)” 241, which is known as the “Dynamic Index in Yaw”
or “DIY”, which is similar to the sprung mass “Dynamic Index in Pitch” or “DIP” 242.

Just as the DIP, or “Kys2/(lf x lr)”, is an important factor in the pitch motion of the
vehicle sprung mass, the quantity “Kz2/(lf x lr)” is an important factor in the yaw motion of the
entire vehicle. Although the quantity “Kz2/(lf x lr)” is officially (SAE) called the “Dynamic
Index in Yaw”, instead of being represented by the initials “DIY” the quantity “Kz2/( lf x lr)”
is often represented by the more convenient symbol “χ” (Greek lower case letter “chi”):

𝛘 = 𝐊 𝐳 𝟐 ⁄(𝐥𝐟 × 𝐥𝐫 ) (EQ. 6.26)


“χ” is very important as it allows for a determination of the location of the
“oscillation center” (OC) about which the vehicle will initially pivot, which in turn allows for
a determination of the magnitude of the yaw inertia “Ioc” that the tire traction forces will have
to initially overcome. For the value of this yaw motion factor “χ” there are three possibilities
which can be depicted as follows: 243:

Figure 6.7 – TRANSIENT ROTATION CENTER & ROTATIONAL INERTIA

241
In SAE J670 2008 standard symbolism this quantity is symbolized as “kzz2/a×b”.
242
Reference [91], page 55.
243
Reference [4], pages 32-33. Reference [53], pg. 253 notes that for “classic” cars (c. 1930) the “χ” ratio tended
to be around 0.775 to 0.675 in value (which was fairly desirable), and on pg. 251 it is noted that many “modern”
cars (c. 1955-?) exceed 1.0. On pg. 301 it is stated that 1920-1930 larger closed cars (were) close to 0.57 to 0.62.

118
SAWE Paper No. 3602
Category Number 31.0

For possibility (1) the yaw inertia “Ioc” tends to be relatively small, equivalent to “Iz
+ Wt (lr χ)2” 244, and the OC is at the distance “lr χ” aft of the c.g., which is “lr (χ-χ 0.5)”
forward of the rear axle. As soon as the front wheels begin to steer the tendency to pivot about
the OC will generate lateral reaction forces at the rear tires in the same direction as the forces
will be when the steady state condition is reached; there will be no reversal of forces from
transient to steady state. This case is supposed to represent a relatively short and smooth
transient period. For perspective, consider that 1930 automobiles tended to have a “χ” factor
value of 0.675 to 0.775 245.

For possibility (2) the yaw inertia “Ioc” tends to be relatively larger and equivalent to
“Iz + Wt lr2” 246 with the OC right at the rear axle. The tendency to pivot about the rear axle
means that steering angle input at the front wheels will not immediately generate lateral
reaction forces at the rear tires; there will be some small lag time.

For possibility (3) the yaw inertia “Ioc” tends to be relatively much larger, equivalent
to “Iz + Wt (lr χ)2” 247, and the oscillation center is at distance “lr χ” aft of the c.g., which is “lr
(χ–χ 0.5)” aft of the rear axle. As the front wheels begin to steer the tendency to pivot about the
OC will generate lateral reaction forces at the rear tires in the opposite direction from how the
forces will be when the steady state condition is reached; there will be a reversal of forces
from transient to steady state. This case would represent a long and fluctuating transient
period; this has been described as creating an uneasy sensation of “floating” at the rear of the
vehicle.

So, the first phase of initiating a turn can be considered as typically starting with the
driver cranking some steering angle “δ” in at the front wheels, which gives rise to some “slip
angle” “ψ” at the front wheels and one of the consequent situations as depicted in Figure 6.7.
The lateral forces and turning moments will initially be in accord with the OC and the
rotational inertia “Ioc”, but in a short interval of perhaps around 0.2 seconds the turn axis will
migrate from the OC to the CG with consequent changes in rotational inertia, moments, and
forces. Arrival at steady-state yaw velocity does not signal end of the transient phase; there is
now some angular momentum which, in conjunction with the lateral compliances inherent in
the tires and suspension, means there will be some overshoot of the steady-state position.

244
Reference [4] says that in this case the “yaw inertia” is “Wf (lf χ 0.5)2 + Wr (lr χ 0.5)2”, which for the “baseline”
vehicle (χ = 0.74) amounts to 48,513 lb-ft2, or just about exactly equal (100.033%) to the “known” vehicle “Iz”
of 48,497 lb-ft2. Something is definitely wrong if the words “yaw inertia” are to be taken as meaning “Ioc”, so
this author has substituted his own expressions for the “Ioc”, which in this case for the “baseline” vehicle yields a
reasonable 81,664 lb-ft2.
245
Reference [53], pg 253. Generally this coincided with “Lwb/Leff” of 0.75 to 0.70 (inverse relationship to “χ”).
246
Reference [4] says the “yaw inertia” is “…as if the front and rear axle loads represented equivalent masses
actually concentrated at the respective axle lines…”; this author has substituted “Ioc” and the expression “Iz +
Wt lr2” in place of that quote.
247
Reference [4] gave “Wf (lf χ 0.5)2 + Wr (lr χ 0.5)2” for the “yaw inertia”.

119
SAWE Paper No. 3602
Category Number 31.0

It should not be surprising that there would be some oscillatory behavior observable
with all this fluxing of forces and moments inflicted upon a damped spring-mass system, and
a decaying oscillation is exactly what ensues after the overshoot. Figure 6.8, a plot of the
vehicle yaw velocity “𝜔” versus time for the transient phase from the commencement of a
turn up to attainment of the steady-state condition, illustrates how transient behavior has two
aspects, “rise” and “decay”, which sum to the total vehicle “response time”:

Figure 6.8 – TRANSIENT YAW RESPONSE TO TURN INITIATION

For some of the same specific vehicles as previously noted in Figure 6.2, the over plot
of transient responses is illustrated by Figure 6.9:

Figure 6.9 – COMPARATIVE TRANSIENT YAW RESPONSE FOR FOUR


VEHICLES

120
SAWE Paper No. 3602
Category Number 31.0

Of course, along with this yaw oscillation there would be some accompanying roll
oscillation 248, and maybe even a little pitch oscillation. All these oscillations result in
fluctuating demands upon the tire contact patches, shifts in suspension linkage, and confusing
sensory inputs to the driver: “…oscillation…is critical to the feel of the car. …a high decay
rate might...be described as ‘twitchiness’…a vehicle that would wiggle around as the driver
turned into a corner” 249. It would seem best that this transient phase be as short in duration
(“response time”), small in magnitude (“yaw velocity”), and smooth as possible.

It is because the lateral forces in the transient state can be very different in magnitude,
and even in direction, from the forces in the steady-state that the handling can also be
different; an normally stable vehicle can enter an unstable condition known as “transient
oversteer”. In a research paper authors Renfroe, Semones, and Roberts took a hard look at the
transient handling of the 1995 Ford E-350, a 15 passenger capacity van 250. Such a van tested
well on the skidpad per the SAE J266 protocol, benignly understeering (although at just over
the maximum lateral acceleration limit it could lose traction at the rear, exiting the skidpad
rear end first in classic terminal oversteer style 251). The performance nature of such a high
load capacity (what aeronautical engineers would call a high “payload fraction”) vehicle can
vary greatly dependent on the payload magnitude and distribution. Given an adverse payload
condition that moved the van c.g. upward and rearward (while increasing the weight), it
would not be surprising that a harsh transient condition brought on by a sharp steering input
could result in an oversteer loss of control.

Apparently such loss of control in the transient state had been known to occur, which
prompted Renfroe et al to seek to find a simple metric that could predict such transient
oversteer. They began with what has been termed the “Ackermann Yaw Rate” (AYR) which
utilizes only the vehicle velocity “V” (ft/sec), steering angle “δ” (deg), and wheelbase “lwb”
(ft) for computation:
𝑽 𝐭𝐚𝐧 𝜹
𝑨𝒀𝑹 = 𝟓𝟕. 𝟑 (EQ. 6.27)
𝒍𝒘𝒃

248
Reference [28], page 319. An example roll oscillation plot (Fig. 9.6) is shown, which is very similar to the
yaw plots presented in this paper. Transient yaw motion is complicated by accompanying transient roll motion,
just as in the steady-state interaction of these two motions. And, just as terminal transient oversteer is a problem,
so is transient rollover. However, transient roll is neglected in this paper’s discussion; for a more complete
discussion see Reference [53], pp. 223-267, and Reference [28], pp. 317-324.
249
Reference [29], pp. 37-40. Figures 6.2, 6.8, and 6.9 are based upon this source.
250
Reference [67], pp.1-3
251
Understeering to the point of maximum lateral acceleration, and then experiencing terminal oversteer, may be
an indication that it would be wise to consider the test vehicle a special case which possibly may present
transient oversteer difficulty.

121
SAWE Paper No. 3602
Category Number 31.0
Rev C

Even though the AYR is a yaw rate calculated for a condition of zero “slip angle” at
the tires, it may be considered an “ideal” to which the “…actual (i.e., measured) yaw rate
can…be compared…to quantify the transitional oversteer or understeer of the
vehicle… 252”. The authors go on from that statement to show how the AYR compared to the
measured yaw rate from a series of standard test maneuvers (“J-Turn”, “Fishhook”) of the E-
350 van in various load conditions.

Given the preliminary review of the physics of the transient state, the comparative
transient behavior of the “baseline” vs. the “advanced concept” may now be considered.
Utilizing the weight “Wt” and yaw inertia “Iz” the vehicle yaw radii of gyration are
obtained:

Base: 𝑲𝒛 = �𝑰𝒛 ⁄𝑾𝒕 = �𝟒𝟖𝟒𝟗𝟕⁄𝟑𝟔𝟒𝟎 = 3.65 ft, or 1.11 m

Adv: 𝑲𝒛 = �𝑰𝒛 ⁄𝑾𝒕 = �𝟒𝟔𝟏𝟔𝟖⁄𝟑𝟔𝟐𝟗 = 3.57 ft, or 1.09 m

These radii (converted to inches) and the vehicle CG location dimensions “lf” and “lr”
allow for determination of the vehicle DIY, a.k.a. “χ”, factors (“χ = Kz2/( lf x lr)”):

Base: 𝝌 = (𝟑. 𝟔𝟓 × 𝟏𝟐)𝟐 ⁄(𝟓𝟑. 𝟎𝟏 × 𝟒𝟖. 𝟗𝟗) = 0.74

Adv: 𝝌 = (𝟑. 𝟓𝟕 × 𝟏𝟐)𝟐 ⁄(𝟒𝟓. 𝟗𝟐 × 𝟓𝟑. 𝟒𝟏) = 0.75

So both vehicles are well established as being of the first case: “χ < 1”, which is
thought to “…represent a relatively short and smooth transient period” (also these values
are under 0.80 which some authorities used to regard as the maximum advisable for a
passenger car 253). The location of the OC’s and determination of the “Ioc” rotational inertia
values via the relations of case (1):

Base: OC = CG + lr χ = 53.01 + 0.74 (48.99) = 89.26 in, or 226.7 cm

Ioc = Iz + Wt (lr χ)2 = 48497 + 3640 (0.74 × 48.99/12)2 = 81,718 lb-ft2, or


3444 kg-m2

Adv: OC = CG + lr χ = 45.92 + 0.75 (53.41) = 85.98 in, or 218.4 cm

Ioc = Iz + Wt (lr χ)2 = 46168 + 3629 (0.75 × 53.41/12)2 = 86,606 lb-ft2,


or 3650 kg-m2

252
Ibid, pg. 2.
253
Reference [53], pg. 251.

122
SAWE Paper No. 3602
Category Number 31.0
Rev C

At 0.384 g’s on the 100 ft (30.5 m) radius (“R”) skidpad the actual steady-state
yaw rate (a.k.a. “yaw velocity”, “ω”) and the “ideal” AYR for “baseline” and “advanced
concept” are:

Base:

𝛚 = 𝟑𝟔𝟎��𝟐𝛑𝐑� � = 𝟑𝟔𝟎��𝟐𝛑𝟏𝟎𝟎� �
�𝐚 𝐲 𝐠 𝐑 �𝟎. 𝟑𝟖𝟒 (𝟑𝟐. 𝟏𝟕𝟒) 𝟏𝟎𝟎

= 20.14 deg/sec
𝑽 𝐭𝐚𝐧 𝜹 𝟑𝟓.𝟏𝟒 𝐭𝐚𝐧 𝟔.𝟔𝟒
𝑨𝒀𝑹 = 𝟓𝟕. 𝟑 𝒍𝒘𝒃
= 𝟓𝟕. 𝟑 (𝟏𝟎𝟐⁄𝟏𝟐)
= 27.58 deg/sec

Adv:

𝛚 = 𝟑𝟔𝟎��𝟐𝛑𝐑� � = 𝟑𝟔𝟎��𝟐𝛑𝟏𝟎𝟎� �
�𝐚 𝐲 𝐠 𝐑 �𝟎. 𝟑𝟖𝟒 (𝟑𝟐. 𝟏𝟕𝟒) 𝟏𝟎𝟎

= 20.14 deg/sec
𝑽 𝐭𝐚𝐧 𝜹 𝟑𝟓.𝟏𝟓 𝐭𝐚𝐧 𝟒.𝟖𝟓
𝑨𝒀𝑹 = 𝟓𝟕. 𝟑 𝒍𝒘𝒃
= 𝟓𝟕. 𝟑 (𝟗𝟗.𝟑𝟑⁄𝟏𝟐)
= 20.64 deg/sec

The ratio “AYR/ω” is greater than unity for both vehicles (“baseline”: 1.369,
“advanced concept”: 1.025), although only by the slightest of margins for the “advanced
concept”, which is to say that it is very close to the “ideal” 254. If the “baseline” vehicle took
about 0.35 seconds to reach the yaw velocity “ω” of 20.14 deg/sec, then the average yaw
acceleration “α” would have had to have been about 57.5 deg/sec2. Simply based on a
comparison of the vehicle rotational inertias (“Icg+Ioc”) it would seem that the “advanced
concept” rise time would be at least 6% longer (and possibly much more) but, taking the
lateral stiffness of the tires into account, its decay time could be nil. Comparison of the plotted
transient yaw response of the “baseline” vs. the “advanced concept” could be very much like
the respective response of the Corvette Z51 vs. Datsun 280ZX in Figure 6.9 initially, and then
more like the reverse (Datsun 280ZX vs. Corvette Z51) during the decay phase.

254
That the value of the yaw rate ratio “AYR/ω” can serve as an indicator of US, NS, or OS conditions seems
reasonable, but (per Reference [67], pg. 2) certain reputable researchers, Michael Blundell and Damian Harty,
have reportedly identified “AYR/ω < 1” as an “oversteering” condition, but “AYR/ω > 1” is the case for the
vehicles in question. It would seem reasonable to conclude that “AYR/ω > 1” indicates an “understeering”
condition.…

123
SAWE Paper No. 3602
Category Number 31.0

All of this calculation might have been ultimately more useful if Renfroe et al had
come up with a generally applicable metric for the determination of terminal transient
oversteer. However, they came up with a yaw rate metric specific to the van in question,
suggesting that electronic monitoring of vehicle yaw rate with respect to this limiting metric
could serve as the basic algorithm for electronic stability control 255. However, the van’s
handling deficiencies were actually overcome by the far more prosaic (and reliable) method of
increasing the rear suspension’s lateral cornering stiffness; the van’s standard rear axle was
replaced with a “dually”.

To the extent possible due to the limited methodology of this paper, both the
“baseline” and “advanced concept” have displayed acceptable steady-state and transient
behavior, with the latter showing some improvements over the former. The calculations as
presented in this chapter represent a methodology useful only for conceptual design and the
exposition of principles. Truly accurate determination of automotive lateral behavior requires
recourse to complex tire models such as the “Pacejka Magic Formula” 256, which in the form
of appropriate computer program modules (such as those produced by Delft-Tyre) may be
incorporated into commercial automotive dynamic simulations such as Dynacar, CarSim, or
ADAMS/Car. However, the 2002 version of the “Pacejka Magic Formula” utilizes about 118
different parametric inputs, and nearly half of them have to be empirically derived for the
particular tires involved. Even a “simplified” version of the “Magic Formula” requires 16
parameters for the tires alone 257, and that count increases rapidly if different tire
types/inflation pressures/et cetera are specified for the same vehicle. All this was beyond the
scope of this paper, and the resources of this author, so the much simpler methodology as
presented was used, which is more conducive to human understanding anyway.

255
Gyroscopic yaw/lateral acceleration rate sensors, and possibly “weight-on-wheels” (WOW) sensors, have all
been proposed to serve as key elements in electronic stability systems to counter terminal oversteer/rollover.
However, this author is strongly opposed to any idea that such artificial electronic stability should ever come to
serve in lieu of inherent physical stability for passenger cars. See Reference [3], pp. 66-68, for more information.
256
Hans Bastiaan Pacejka is a Professor Emeritus at Delft University of Technology. Pacejka, et al, has
developed a tire behavior model(s) termed the “Magic Formula”. It is so termed because there is little theoretical
basis in physics for the governing equations; they are mainly based on regression analysis of empirical data. A
tire is characterized by 10-to-20 coefficients for each force (lateral, longitudinal, torque) produced at the tire-
road contact area. The Pacejka tire model is the de facto standard for use in professional vehicle dynamics
simulations.
257
Reference [80], Table 1.

124
SAWE Paper No. 3602
Category Number 31.0
Rev C

7 – VERTICAL PERFORMANCE: SHOCK/CONTACT/VIBRATION/PITCH &


BOUNCE

The primary function of a suspension system is the “isolation” of the sprung mass
from road shock. How well a design does in this regard has been traditionally determined at
the conceptual design stage by use of a simple “Quarter-Car Model” as illustrated on the left
hand side of the following figure:

Figure 7.1 – TRADITIONAL vs. ADVANCED CONCEPT ROAD SHOCK


“QUARTER-CAR MODEL”

However, if a “Quarter-Car Model” is appropriate for a conventional four wheel


vehicle such as the “baseline”, then it seems reasonable that an “Eighth-Car Model” is
appropriate for an unconventional vehicle such as the “advanced concept” 258. However, the
tires of the “advanced concept” are only about 82% front/75% rear as stiff as the “baseline”,
and the “advanced concept” suspension is only 50% as stiff as the “baseline”, with only the
slightest change (2%) in overall sprung mass between the two 259, so the reasonable
expectation is that the “advanced concept” road shock magnitude would be slightly less than
half that of the “baseline”. However, that is far from what happens when using an “Eighth-Car
Model”; the reduction in road shock transmitted to the sprung mass is only about -14% (front
axles). So, at the risk of indulging in a “self-fulfilling prophecy”, the “Quarter-Car Model”
will be used for the analysis of the “advanced concept” as well, although it is a special one
(the right hand side of Figure 7.1).

258
The “Equivalent Conventional Configuration” model used earlier as a “stand in” for the “advanced
configuration” was appropriate for the longitudinal acceleration/braking study of Chapter 5 and the lateral
handling/stability study of Chapter 6, but for the “Quarter-Car Model” now used for vertical suspension study
remember that the “advanced configuration” tire and suspension spring rates shown in Table 6.3 are double the
actual “advanced configuration” rates now to be used in this chapter.
259
The total “baseline” sprung mass weight is 3118 lb (1414.3 kg) versus the “advanced concept” sprung mass
weight of 3181 lb (1442.9 kg); see Table 6.3.

125
SAWE Paper No. 3602
Category Number 31.0

For a “Quarter-Car Model” the appropriate equations for evaluating the situation are
determined as follows 260. First, the period for one complete oscillation “τ” of the sprung
mass “ms” mounted on a suspension with an effective vertical spring rate at the wheel “kcs”261
is:

τ = 2π √(ms/kcs) (EQ. 7.1)


This period “τ” is a characteristic of the free oscillation of the suspension-spring/sprung-
mass system, of which the system’s natural frequency “fs” is the reciprocal, and therefore the simple
sprung mass frequency equation can written as:

𝟏
fs = 1/τ = √(kcs/ms) (EQ. 7.1a)
𝟐𝛑

Knowledge of the system’s frequency “fs” and amplitude “A” (the road bump height) allows
for a determination of the maximum vertical acceleration “amax” that will be inflicted upon the sprung
mass 262:

amax = (2π fs)2 A (EQ. 7.2)


A simple substitution for “fs” results in an even simpler formula:

amax = (kcs/ms) A (EQ. 7.3)


For the “baseline” vehicle the effective spring rate at a front wheel for a “single wheel
bump” (when the ARB joins in) is 141.994 lb/in (25.357 kg/cm), but for a “double wheel
bump” (both wheels on an axle impacted at the same time, so the ARB pivots freely) is
98.524 lb/in (17.594 kg/cm). The effective spring rate at a rear wheel is 118.4497 lb/in
(21.153 kg/cm). 263. The sprung mass (traditional “Quarter-Car Model”) is 1.9865 lb-sec2/in
(766.9669 lb, 347.89 kg) front, and 2.0514 lb-sec2/in (792.0331 lb, 359.26 kg) rear 264.
Therefore, the system period “τ”, and maximum acceleration “amax” of the “baseline”
suspension spring-sprung mass system, are as follows:

260
Reference [91], pp. 17-18.
261
The “kcs” stands for the combined suspension spring and tire vertical spring in series, which is slightly
“softer” than suspension spring rate alone.
262
See SAE 670 2008, Vehicle Dynamics Terminology, and SAE J6a, Ride and Vibration Data Manual.
263
The “baseline” combined (spring and tire in series) front spring rate is 98.524 lb/in, but for a single wheel
bump the ARB contributes an additional 43.47 lb/in for a total 141.994 lb/in; the combined (spring and tire in
series) rear spring rate is simply 118.4497 lb/in (see Table 6.3) as there is no ARB at the rear.
264
See “Appendix B - Weight Accounting”.

126
SAWE Paper No. 3602
Category Number 31.0
Rev C

Front Single Wheel Input (Single At Wheel Stiffness With ARB):

τ = 2π √(ms/ks) = 2π √(1.9865/141.994) = 0.7432 sec

amax = (ks/ms) A = (141.994/1.9865) 2 = 142.96 in/sec2, or 0.37 g’s265


Front Double Wheel Input (Single At Wheel Stiffness Without ARB):

τ = 2π √(ms/ks) = 2π √(1.9865/98.524) = 0.8922 sec

amax = (ks/ms) A = (98.524/1.9865) 2 = 99.19 in/sec2, or 0.26 g’s


Rear Single or Double Wheel Input (Single Wheel Stiffness):

τ = 2π √(ms/ks) = 2π √(2.0514/118.4497) = 0.8269 sec

amax = (ks/ms) A = (118.4497/2.0514) 2 = 115.48 in/sec2, or 0.30 g’s


For the “advanced concept” the effective spring rate at a front wheel (first two axles)
for a “single wheel bump” (the ARB joins in) is 72.4744 lb/in (12.9425 kg/cm), but for a
“double wheel bump” (both wheels on an axle impacted at the same time) is 50.7394 lb/in
(9.0610 kg/cm). The effective spring rate at a rear wheel (last two axles) is 60.8290 lb/in
(10.8628 kg/cm) 266. The sprung mass (“Advanced Concept Quarter-Car Model”) is 2.2948
lb-sec2/in (886.00 lb, 401.88 kg) front, and 1.8247 lb-sec2/in (704.50 lb, 319.56 kg) rear 267.
Therefore, the system period and maximum accelerations inflicted on the sprung mass are as
follows:

Front Single Wheel Input (Axles #1 and #2 Single At Wheel Stiffness With ARB):

τ = 2π √(ms/ks) = 2π √(2.2948/72.4744) = 1.1181 sec


amax = (ks/ms) A = (72.4744/2.948) 2 = 63.16 in/sec2, or 0.16 g’s
Front Double Wheel Input (Axles #1 and #2 Single At Wheel Stiffness Without ARB):

τ = 2π √(ms/ks) = 2π √(2.2948/50.7394) = 1.3362 sec


amax = (ks/ms) A = (50.7394/2.2948) 2 = 44.22 in/sec2, or 0.11 g’s

265
The value of gravity is taken as 386.088 in/sec2 for calculations in the English IPS units system.
266
The “advanced concept” spring rates used here are the actual values and not the “equivalent conventional
configuration” values as per Table 6.3 (although it’s easy to determine the actual values from Table 6.3). This is
because for road shock analysis the individual wheels, or at least individual axles, have to be considered.
267
See “Appendix B - Weight Accounting” for all mass property values used in this chapter.

127
SAWE Paper No. 3602
Category Number 31.0
Rev C

Rear Single or Double Wheel Input (Axle #3):

τ = 2π √(ms/ks) = 2π √(1.8247/60.8290) = 1.0882 sec


amax = (ks/ms) A = (60.8290/1.8247) 2 = 66.67 in/sec2, or 0.17 g’s
Rear Single or Double Wheel Input (Axle #4):

τ = 2π √(ms/ks) = 2π √(1.8247/60.8290) = 1.0882 sec


amax = (ks/ms) A = (60.8290/1.8247) 2 = 66.67 in/sec2, or 0.17 g’s
The “advanced concept” sprung mass shock levels would seem to be about ½ the
“baseline” levels, which “makes sense”, but also the severest shock location has shifted from
the front for the “baseline” to the rear for the “advanced concept”. The “shock” significance
of the system period “τ” is that it is indicative of how fast the cycle time (about one tenth of
the period) of any automatic damping control system must be in order to be effective 268.
Although damping was omitted from the shock analysis for simplicity, the presence of a
“shock absorber” would tend to increase the shock levels. To minimize this adverse effect
many vehicles are now equipped with automatic damping control systems which adjust the
damping to an optimum level best suited to the driving conditions of the moment. For the
“baseline” maximum cycle times of 0.074-0.089 seconds front/0.083 seconds rear are
indicated, and for the “advanced concept” maximum cycle times of 0.111-0.133 seconds
front/0.109 seconds rear are indicated. However, this is not quite the advantage for the
“advanced concept” that the reduction in shock levels is 269.

Table 7.1a – ROAD SHOCK RESULTS

268
BWI (Beijing West Industries) Global Technology and Business Development Manager (Paris) Olivier
Raynaud has been quoted as saying “It is important to achieve times below 20 ms for the…damper
transitions…” (Reference [103], pg. 12).
269
The indicated cycle time “advantage” may likely be negligible given the ongoing improvements in automatic
control cycle times; “slow” cycle times used to bedevil early ABS and other control systems. In particular,
modern rheologically controlled damping systems have a marked advantage over older adjustable valve
(“orifice”) types.

128
SAWE Paper No. 3602
Category Number 31.0

Shock attenuation may be the primary function of any suspension system, but for
“high speed” automobiles the maintenance of tire contact with the road must rank as a very
close second. In this regard there is the oft-encountered condition of a dip in the road surface
as depicted in the following illustration:

Figure 7.2 – TRADITIONAL vs. ADVANCED CONCEPT ROAD CONTACT


“QUARTER-CAR MODEL”

It’s easy to see that if the suspension spring/unsprung-mass system is not quick
enough to react in time, or flexible enough to reach the bottom of the depression, then contact
between tire and road will be lost. The road surfaces of the real world are often a series of
such undulations, interspersed with bumps and other irregularities. The question concerning
the suspension spring/unsprung mass system is: what is the minimum length “l” and
maximum depth “d” that the system can traverse at speed “V” without losing contact? Using
the equations of SHM for the suspension spring/unsprung mass system the following
relationships are derived 270:

l = (V τ / 2) √mus/ms (EQ. 7.4)

d = ds (1 + mus/ms) (EQ. 7.5)

Note the appearance of the unsprung-to-sprung mass ratio “mus/ms” (“MR”); this is
a key parameter that will appear frequently throughout any study of suspension performance;
generally the smaller the “MR” value then the better the performance. Note also the
reappearance of our old friend “τ”, the system period of oscillation, which is a measure of

270
Reference [91], pp. 22-23. Note that the analysis model is now based on the suspension spring only, with no
consideration of tire stiffness, as it is the motion of the unsprung mass on the suspension spring which provides
the major “road-profile-following” capability.

129
SAWE Paper No. 3602
Category Number 31.0
Rev C

system “quickness”. And, of course, there is the deflection “ds” of the suspension spring
under the “Quarter-Car” sprung weight load.

Using some of the information from the previous analysis, the “baseline” has a
“Quarter-Car Model” sprung mass of 766.97 lbs (1.986508 lb-sec2/in, 347.89 kg) front and
792.03 lbs (2.051432 lb-sec2/in, 359.26 kg) rear. The corresponding unsprung mass is 107 lb
(48.53 kg) front and 154 lbs (69.85 kg) rear. The effective spring stiffness is 149.87 lb/in
(26.8 kg/cm, with ARB)/106.4 lb/in (19.0 kg/cm) at the front wheels, and 128.8 lb/in (23.0
kg/cm) at the rear wheels. The vehicle approaches a dip at 30 mph (44 ft/sec, 48.3 kph, 13.4
m/sec); in order for the tires to keep in contact with the road the length of the dip can be no
narrower/depth no greater than:

Front Single Wheel (with ARB) Input:

ds = 766.97/149.87 = 5.1175 in (12.9985 cm)


τ = 2π √(1.986508/149.87) = 0.7234 sec
l = (44 ft/sec 0.7234 sec/2) √(107/766.97) = 5.94 ft, or 71.33 in (181.1782 cm)
d = 5.1175 in (1 + 107/766.97) = 5.8315 in (14.812 cm)
Front Double Wheel (w/o ARB) Input:

ds = 766.97/106.4 = 7.2083 in (18.3091 cm)


τ = 2π √(1.986508/106.4) = 0.8585 sec
l = (44 ft/sec 0.8585 sec/2) √(107/766.97) = 7.05 ft, or 84.66 in (215.0364 cm)
d = 7.2083 in (1 + 107/766.97) = 8.214 in (20.8636 cm)
Rear Single or Double Wheel Input:

ds = 792.033/128.8 = 6.1493 in (15.6192 cm)


τ = 2π √(2.051432/128.8) = 0.7930 sec
l = (44 ft/sec 0.7930 sec/2) √(154/792.033) = 7.69 ft, or 92.31 in (234.4674 cm)
d = 6.1493 in (1 + 154/792.033) = 7.345 in (18.6563 cm)
Compare this with the “Advanced Concept Quarter-Car Model” for which the sprung
mass of 443.00 lb (1.147407 lb-sec2/in, 200.94 kg) and an axle #1 and #2 unsprung mass of
44.25 lb (20.07 kg), with an effective spring stiffness with ARB of 74.935 lb/in (13.382

130
SAWE Paper No. 3602
Category Number 31.0
Rev C
kg/cm) and without ARB of 53.20 lb/in (9.500 kg/cm) at the front wheels. The comparable
figures for the rear axles are sprung mass of 352.25 lb (0.91236 lb-sec2/in, 159.78 kg),
unsprung mass at axle #3 of 70.185 lb (31.83 kg) and at axle #4 of 65.685 lb (29.79 kg), and
an effective spring stiffness of 64.4 lb/in (11.500 kg/cm); all of which gives road contact
parameters of:

Front Single Wheel (w/ ARB) Input (Axles #1 and #2):

ds = 443.00/74.935 = 5.9118 in (15.0160 cm)


τ = 2π√(1.147407/74.935) = 0.7775 sec
l = (44 ft/sec 0.7775 sec/2) √44.25/443.00 = 5.41 ft, or 64.87 in (164.7698 cm)
d = 5.9118 in (1 + 44.25/445.00) = 6.5023 in (16.5158 cm)
Front Double Wheel (w/o ARB) Input (Axles #1 and #2):

ds = 443.00/53.2 = 8.3271 in (21.1508 cm)


τ = 2π√1.147407/53.20 = 0.9227 sec
l = (44 ft/sec 0.9227 sec/2) √44.25/443.25 = 6.42 ft, or 76.991 in (195.5571 cm)
d = 8.3271 in (1 + 44.25/443.25) = 9.1588 in (23.2634 cm)
Rear Single or Double Wheel Input (Axle #3):

ds = 352.25/64.4 = 5.4697 in (13.8930 cm)


τ = 2π√0.91236/64.4 = 0.7479 sec
l = (44 ft/sec 0.7479 sec/2) √70.185/352.25 = 7.34 ft, or 88.13 in (223.8502 cm)
d = 5.2795 in (1 + 70.185/352.25) = 6.5596 in (16.6614 cm)
Rear Single or Double Wheel Input (Axle #4):

ds = 352.25/64.4 = 5.4697 in (13.8930 cm)


τ = 2π√0.91236/64.4 = 0.7479 sec
l = (44 ft/sec 0.7479 sec/2) √65.685/352.25 = 7.10 ft, or 85.26 in (216.5604 cm)
d = 5.4697 in (1 + 65.685/352.25) = 6.4897 in (16.4838 cm)

131
SAWE Paper No. 3602
Category Number 31.0
Rev C

At first glance the results may seem a little disappointing. A very impressive
improvement in tire-to-road contact may have been expected for the “advanced concept” over
the “baseline”, but while the minimum length numbers show improvement the maximum
depth numbers show improvement only for the “advanced concept” front axles.

However, the “d” and “l” numbers by themselves only tell part of the story due to the
unusual nature of the “advanced concept”. When a tire on the “baseline” configuration loses
contact with the road then 25% of the total tire-to-road contact is lost, but when a tire on the
“advanced concept” loses contact with the ground then only about 12.5% of the total tire-to-
road contact is lost. However, if the longitudinal length “l” that causes the loss of tire-to-road
contact is greater than the “advanced concept” wheel-to-wheel distance of 49.66 inches, then
after the amount of time has elapsed that it takes for the “advanced concept” to transverse the
difference will the loss of contact will also become about 25%, but then for a much shorter
duration.

A tabulation and analysis of the results aids in understanding, hence Table 7.1b:

Table 7.1b – ROAD CONTACT RESULTS

Largely due to the approximately 15% general overall reduction in the unsprung
mass-to-sprung mass ratio, the “advanced concept” displays about a 36% overall
improvement in road contact capability “l” over the “baseline” (calculated in terms of

132
SAWE Paper No. 3602
Category Number 31.0
Rev C

“percent contact loss × duration of loss”), and an overall improvement of only about 0.14%
in the depth capability “d” 271.

The “bottom line” is that the “advanced concept” does indeed have eight wheels on
four axles to the “baseline” four wheels on two axles, and that fact all by itself is enough to
ensure better tire-to-ground contact. Add to this the previously indicated ability of the
“advanced concept” to maintain road contact in wet conditions, due to both improved
hydroplaning resistance and “squeegee” action, and any performance aspect that depends
upon the quality of tire-to-road contact promises to be significantly improved.

This is because tire-to-road contact is the sine qua non of automotive behavior. With
good contact between road and tire; acceleration, braking, maneuver, and directional stability
all become much more assured. The performance of the “advanced design” with regard to all
those aspects of vehicle behavior has so far been indicated to be superior, but perhaps not as
superior as might actually be the case. That is because the analysis models used assume
perfect ground contact conditions, which in the real world is often not the situation. The real
world empirical performance of the “advanced concept” may prove in many ways to be even
better than predicted by the analysis.

For instance, with twice as many ground contact points for the “advanced concept” as
for the “baseline”, localized adverse road surface conditions are likely to only have ½ the
effect. Consider the case of a localized condition which reduces front axle traction to near
zero during the negotiation of a high “g” turn; such a happenstance could suddenly reduce
directional stability so as to cause a high speed exit from the turn, nose first. Upon
encountering the exact same circumstance, the “advanced concept” would still have three
axles generating directionally stabilizing forces; it might very well proceed safely through the
turn, perhaps requiring only a slight directional correction at the steering wheel.

Road shock and contact are just two of the many performance aspects concerned with
vertical motion. The attenuation of road surface induced vibration, which is very closely
related analytically to attenuation of vibration internally induced by imbalance and engine
pulsations, also comes under the general heading of vertical motion. The matter is one of
input and the response to that input. The equations for the automotive suspension response
factors, a.k.a. the “transmissibility factors” or “gains”, are conventionally derived from the
equations of motion of the free-body representations of the sprung and unsprung mass
components of the “Quarter-Car Model” suspension model.

271
If variation in “d” is considered to be of the same significance as variation in “l”, then perhaps it could be said
that in total there is a 36.14% improvement. Another way of evaluating the quality of road contact maintenance
of a suspension system, but for small scale input as opposed to bumps and dips, would be to determine the
“gain” or “transmissibility” ratio of the unsprung mass system response to road vibration input.

133
SAWE Paper No. 3602
Category Number 31.0
Rev C

This brings this investigation back once again to the question of the appropriateness of
the analysis models utilized. In the previous “road shock” and “road contact” investigations
an adaptation of the conventional “Quarter-Car Model” was utilized for the analysis of the
“advanced concept”. This seemed justifiable based on the fact that, although the “Advanced
Concept Quarter-Car Model” had two wheels, only one wheel at a time encountered the road
“bump” or “dip” 272. However, in the vibration case road surface induced input affects both
wheels of the “Advanced Concept Quarter-Car Model” at the same time:

Figure 7.3 – “ADVANCED CONCEPT QUARTER-CAR MODEL” DIFFERENCES

However, with the “advanced concept” vibration model of Figure 7.3 there is a
significant difference between the model and reality; the road vibration input to both wheels is
not synchronized as the model would seem to indicate. The trailing wheel does experience the
exact same random input as the leading wheel, but at a time lag of “L/V”. A more realistic
representation of the “advanced concept” suspension system would be a much more
complicated, involving two identical but out of phase inputs. However, a way to avoid such
complexity would be to use an “Eighth-Car Model”, which is conventional looking but with
the unsprung mass set to ½ of what it would be for a “Quarter-Car Model”, and with the
remaining parameter values set as appropriate for one particular wheel at a time; this will not
be perfect but should give results accurate enough for the purpose of this study. The revised
model as appropriate for the vibration analysis of the “advanced concept” may therefore be
depicted as:

272
This is not completely true in the road contact case as was considered in the previous analysis; when the road
ripple length “l” is greater than the distance from the leading axle to the second axle (49.66 in or 126.24 cm),
then a second wheel will be affected by the dip before the first has fully cleared it (see Table 7.1b).

134
SAWE Paper No. 3602
Category Number 31.0
Rev C

Figure 7.4 – “ADVANCED CONCEPT EIGHTH-CAR MODEL”

With the “baseline” represented by a “Quarter-Car Model”, and the “advanced


concept” now represented by an “Eighth-Car Model”, the respective suspension parameters
for vibration analysis may be appropriately tabulated for use as follows:

Table 7.2 – VIBRATION PARAMETER VALUES

The relative difference in the parametric values between “baseline” and “advanced
concept” serves as an early indicator as to how the vibration analysis might go. There is a
significant general decrease in the unsprung-to-sprung mass ratio for the “advanced
concept” at the front (-28%) with a small variation in the rear (+2.5% to -4.0%). Relative to
the “baseline”, there are modest changes (-5.4% front to +7.8% rear) in the “advanced
concept” damping coefficients 273, but the “advanced concept” spring constant ratios are
greatly reduced by about 50% with respect to the “baseline”.

273
The “shock absorber” damping coefficients were determined per “𝑪𝒔 = 𝝃𝟐�𝒌𝒄𝒔 ⁄𝒎𝒔 “for “ξ = 0.4”.

135
SAWE Paper No. 3602
Category Number 31.0

What all this means with regard to the vibration gain values can be determined in both
a qualitative and a quantitative manner. Qualitatively, a plot of three of the most significant
vibration gain values over the 0-to-20 Hz range of input frequency typically behave as
follows 274:

Figure 7.5 – SPRUNG MASS GAIN VALUES vs. FREQUENCY FOR ROAD, SPRUNG
MASS, AND UNSPRUNG MASS INPUTS

The sharp peak(s) around 1.22 Hz (ω = 7.66 radians/sec) are indicative of the sprung
mass resonance, whereas the peak at about 13.4 Hz (ω = 71.25 radians/sec) is coincident with
the unsprung mass resonance. The effect on the “sprung mass response to road input” gain
of varying the unsprung mass may be illustrated in both (for clarity) linear and semi-log
plots 275:

274
Reference [91], pg. 37. There is a fourth gain value, the “unsprung mass response gain to road input”,
which is presented in Reference [48], pg. 5, and in Reference [93], pg. 481. This gain value would have bearing
on the quality of road contact, but with respect to the small scale irregularities of the road surface as opposed to
the large scale bumps and dips. For more information see Appendix H.
275
Ibid, pp. 39 & 40.

136
SAWE Paper No. 3602
Category Number 31.0

Figure 7.6 – “SPRUNG MASS RESPONSE TO ROAD INPUT” GAIN AS UNSPRUNG


MASS IS VARIED

Reduction of the unsprung mass reduces the peak transmissibility values, although
beyond some “high speed” point (12 to 13 Hz or so in the above illustrations) will actually
cause somewhat of a transmissibility increase. However, this is a point at which the vehicle is
entering a realm where all the transmissibility values are pretty low and getting lower.

It has been noted that the “advanced concept” spring constant ratios are greatly
reduced with respect to the “baseline”. What this means with regard to the “sprung mass
response to road input” gain values can be determined qualitatively from the following
“typical” plot 276:

Figure 7.7 – “SPRUNG MASS RESPONSE TO ROAD INPUT” GAIN AS STIFFNESS


RATIO IS VARIED

276
Ibid, pg. 42.

137
SAWE Paper No. 3602
Category Number 31.0

Lastly, although there are modest changes (-5% front to +8% rear) in the “advanced
concept” damping coefficients relative to the “baseline”, the damping ratio “ξ” is held
constant at 0.4 in both cases; therefore no change in “sprung mass response to road input”
gain is anticipated in this regard. How the “sprung mass response to road input” gain is
influenced by variation in the damping ratio is indicated by the following “typical” plots
(again linear and semi-log) 277:

Figure 7.8 – “SPRUNG MASS RESPONSE TO ROAD INPUT” GAIN AS DAMPING


RATIO IS VARIED

Based on these typical plots of “sprung mass response to road input” vibration
transmissibility the transition from “baseline” to “advanced concept” would qualitatively
appear to be somewhat beneficial, with a decrease in vibration transmissibility up to about 13
Hz, followed by a slight increase in values from there on (the “high speed” effect as shown in
Figures 7.6 and 7.7).

The transition effect on “sprung mass response to sprung mass (i.e., internal,
generally engine related) input” should be somewhat similar; “typical” plots representing how
that transmissibility is affected by the unsprung mass ratio, stiffness ratio, and the damping
ratio are as follows 278:

277
Ibid, pp. 43-44.
278
Ibid, pp. 46-47.

138
SAWE Paper No. 3602
Category Number 31.0

Figure 7.9 -“SPRUNG MASS RESPONSE TO SPRUNG MASS INPUT” AS MASS RATIO VARIES

Figure 7.10 -“SPRUNG MASS RESPONSE TO SPRUNG MASS INPUT” AS SPRING RATIO VARIES

Figure 7.11 – “SPRUNG MASS RESPONSE TO SPRUNG MASS INPUT” AS DAMPING RATIO
VARIES

139
SAWE Paper No. 3602
Category Number 31.0

The “sprung mass response to sprung mass input” transmissibility will be essentially
unaffected in the transition from “baseline” to “advanced concept” by the unsprung mass-to-
sprung mass ratio (understandably, as only the unsprung mass is actually varying in that
plot), and as the damping ratio is held constant at 0.4 there will no effect on that score either.
However, Figure 7.10 indicates that there will be a general decrease in “low speed” values for
this transmissibility factor as a result of the increased stiffness ratio.

Now the point has been arrived at for an attempt to quantify these expected
improvements. First let’s consider the input values. The road input depends upon vehicle
speed (frequency)/road surface roughness (magnitude), while the sprung mass internal input
frequency/magnitude depends mainly on engine rpm 279, so there will be no change with
respect to those two inputs.

However, the unsprung mass input to the sprung mass depends upon the rotational
speed of the unsprung rotating mass, and a simple study suggests that such input would
increase significantly for the “advanced concept” in both frequency and magnitude. A simple
“axle level” input “F(t)” due to imbalance might be represented as 280…

F(t) = m r 𝜔 2 (EQ. 7.6)

…where “m” represents the mass value of the imbalance, and the radial distance from the
mass “m” to the rotation axis is “r”, which will be taken as being proportional to the rolling
radius. Lastly, “ω” (rad/sec) is the angular velocity of the rotating unsprung mass, which is
inversely proportional to the rolling radius. The force “F(t)” translates into an acceleration
(a(t) = F(t)/mus) for a corresponding oscillating acceleration input into the sprung mass. The
corresponding sprung mass response is the input acceleration times the appropriate gain
value (𝒁̈𝒔 ⁄(𝑭𝒖𝒔 ⁄𝒎𝒔 )) over the range of frequencies considered. The rolling radius for the
“baseline” is about 12.7 in (32.26 cm), while the rolling radius for the “advanced concept” is
6.34 in (6.34 cm). Since “ω = V/2πr”, the “F(t)” ratio of “advanced concept” to “baseline”
is…

279
Due to the necessary final drive ratio change, the engine will be running at essentially the same rpm per
vehicle velocity for the “advanced concept” as for the “baseline”. However, due to the engine re-orientation
within the chassis, the engine rotary excitation of the sprung mass will now tend to generate a response about
the pitch axis as opposed to the former case about the roll axis. Considering the difference in magnitude of the
sprung mass “Iy” with respect to “Ix” and the pitch vs. roll stiffness, the response should be of about 3% less
magnitude and frequency (i.e., improved).
280
Reference [91], pg. 48.

140
SAWE Paper No. 3602
Category Number 31.0

F(t)Adv/F(t)Base = m rA ωA2/ m rB ωB2 = (1/rA) / (1/rB)

= rB /rA = 12.7 / 6.34 = 2.00

…which indicates that any vibration input resulting from imbalance of the rotating
unsprung mass would have twice the magnitude and be at twice the frequency as the
“baseline”. However, the rotating unsprung mass of the “advanced concept” is about
29.9%-37.9% that of the “baseline”, and the magnitude of any imbalance mass “m” is likely
to be proportional, so the final result is that sprung mass vibration input resulting from
imbalance of the rotating unsprung mass is likely to have about 60-to-76% of the
magnitude but at twice the frequency of a similar imbalance for the “baseline”. The power of
the unsprung mass input is therefore likely to be greater by 44% front to131% rear...

The change in vibration inputs in the transition from “baseline” to “advanced concept”
constitutes one half of the vibration problem; the second half is what happens to the vibration
“gains” or transmissibility factors in the transition. These “gain” values must be less than
unity to indicate attenuation instead of amplification; the smaller the value the better. The
following equations for the transmission to the sprung mass of road and internal vibration
input are as given by Prof. Dukkipati in his book Road Vehicle Dynamics 281:

𝒁𝒔̈ 𝟏
� �= 𝟐 �(𝑲𝒕 𝑲𝒔 𝑷 + 𝑲𝒕 𝑪𝒔 𝑸𝝎)𝟐 + (𝑲𝒕 𝑪𝒔 𝑷𝝎 − 𝑲𝒕 𝑲𝒔 𝑸)𝟐
𝒁𝒓̈ 𝑷 +𝑸 𝟐

(EQ. 7.7)

𝒁𝒔̈ 𝒎𝒖𝒔 𝝎𝟐
� �= 𝟐 ��(𝑲𝒔 + 𝑲𝒕 − 𝒎𝒔 𝝎𝟐 )𝑷 + 𝑸𝑪𝒔 𝝎�𝟐 + �𝑪𝒔 𝑷𝝎 − 𝑸(𝑲𝒔 + 𝑲𝒕 − 𝒎𝒔 𝝎𝟐 )�𝟐
𝑭𝒔 ⁄𝒎𝒔 𝑷 +𝑸 𝟐

(EQ. 7.8)

281
Reference [22], pp. 266-271. It was originally intended to use the three transmissibility equations presented in
Reference [28], and reflected by this author in Reference [91], but there seems to be a “typo” in those equations
which is beyond the ability of this author to find and correct. As an unfortunate result, calculation of the “sprung
mass response to unsprung mass input” gain is not represented herein, but the other two basic gain factors are.

141
SAWE Paper No. 3602
Category Number 31.0

Where “P” and “Q” are:

𝑷 = 𝒎𝒔 𝒎𝒖𝒔 𝝎𝟐 − (𝑲𝒔 𝒎𝒖𝒔 + 𝑲𝒕 𝒎𝒖𝒔 + 𝑲𝒔 𝒎𝒔 )𝝎𝟐 + 𝑲𝒔 𝑲𝒕

𝑸 = 𝑲𝒕 𝑪𝒔 𝝎 − 𝑪𝒔 (𝒎𝒔 + 𝒎𝒖𝒔 )𝝎𝟑


The various vehicle parameters that serve as input to these equations are:

ω = the angular frequency in radians per second (ω = 2π Hz).

ms = the sprung mass (lb-sec2/in).

mus = the unsprung mass (lb-sec2/in).

Cs = the damping coefficient for the suspension damper (lb-sec/in).

Ks = the spring rate for the suspension spring (lb/in).

Kt = the spring rate for the tire (lb/in).

A complete vibration analysis would require that the gain values be evaluated for all
possible loading conditions over the entire possible frequency range and be multiplied by the
appropriate (type & frequency) input values to obtain the response spectrum; the sprung
mass response to road input case is shown (the other cases are similar) 282:

Gzs(f) = [Hzs(f)]2 Gzr(f) (EQ. 7.9)


Where:

Gzs(f) = Sprung mass vertical acceleration PSD response at a frequency.

Hzs(f) = Sprung mass vertical gain or transmissibility factor.

Gzr(f) = Road vertical acceleration PSD input at a velocity (frequency).

The result is an enormous number of sprung mass response acceleration spectral


density plots of “g/Hz” per frequency “Hz” that would cover the entire operational envelope
of the vehicle considered. Areas of the operational envelope that generated unacceptable
response could be identified and corrective actions (e.g.: modification of damping
coefficients, spring rates, tire pressures, etc.) taken, giving rise to yet another round of

282
Reference [91], pg. 38.

142
SAWE Paper No. 3602
Category Number 31.0

analysis as an acceptable design is closed in on via the converging 283 spiral of the design
cycle.

Such a proper and lengthy analysis is beyond the scope of this paper. However, since
the performance of the “baseline” vehicle with regard to vibration transmission can be
assumed to have been acceptable 284, the vibration performance of the “advanced concept” can
be evaluated in a limited way by comparison with the “baseline” for just the single “two-up”
test condition(s) as obtained in Appendix B and reflected in the parameters listed in Table 7.2.
The results are plotted as follows:

Figure 7.12 – FRONT SUSPENSION SPRUNG MASS RESPONSE TO ROAD &


INTERNAL INPUT

The “advanced concept” front suspension shows sharply decreased peak


transmissibility for both road (3.16 to 1.93) and (2.94 to 1.66) internal input, but in the
frequency range of about 6 to 15 Hz the internal input gain is somewhat elevated.

283
Hopefully.
284
Although perhaps not acceptable by modern standards; the 1958 Jaguar XK150S would probably be
considered a bit on the harsh side of the NVH spectrum today.

143
SAWE Paper No. 3602
Category Number 31.0

Figure 7.12 – FRONT SUSPENSION W/ARB SPRUNG MASS RESPONSE TO ROAD


& INTERNAL INPUT

The front suspension with anti-roll bar (ARB) shows essentially the same result, only
as the suspension is stiffer all the peak transmissibility values are higher.

Figure 7.12 – REAR SUSPENSION SPRUNG MASS RESPONSE TO ROAD &


INTERNAL INPUT

144
SAWE Paper No. 3602
Category Number 31.0

The “advanced concept” rear suspension (axles #3 and #4 indistinguishably over-plot)


again shows sharply decreased peak transmissibility for both road (3.29 to 1.70) and (3.07 to
1.40) internal input, but in the frequency range of about 5 to 13 Hz the internal input gain is
again somewhat elevated.

The somewhat elevated gains in the “higher” frequency range was anticipated from
the qualitative analysis, as was a modest decrease in peak values; the large decrease (around
50%) comes as a pleasant surprise. However, the “sprung mass response to unsprung mass
input” is not addressed here, and that is a significant drawback in view of the expect increase
in unsprung mass vibratory input. There is some reason 285 to hope for a decrease in peak
transmission at the front (-13.5%), but at the rear an increase (+12.5%) seems likely, which
means possible +38%/+147% front/rear increases in the unsprung mass disturbance of the
sprung mass; this is another area (in addition to tire wear and “standing wave”) for concern
in the design of the “advanced concept”.

The last aspect of vertical motion concerns the ride motions of pitch/bounce. This is a
subject dealt with at length by this author in a previous paper, and the “baseline” (1958 Jaguar
XK150S) was specifically investigated 286. This investigation will now be repeated with some
updating of the parameter values, followed by a similar analysis of the “advanced concept”
for comparison.

For the 1958 Jaguar XK150S the sprung weight is about 3118 lb (1414.3 kg, sprung
mass is 3118/32.174 = 96.91 lb-sec2/ft), with an “lf” of 4.32 ft ( 1.32 m), an “lr” of 4.18 ft
(1.27 m), and a wheelbase “l” of 8.5 ft (2.59 m). The pitch mass moment of inertia “J” is
1130.5 lb-ft-sec2 (1532.7 kg-m2), so the pitch radius of gyration squared is 11.6654 ft2 (“K”
= 3.42 ft, or 1.042 m). The spring constants at the axles 287 front and rear are 2364.0 lb/ft
(3518.0 kg/m) and 2841.6 lb/ft (4228.8 kg/m), respectively. The equations for coupled motion
are employed as they will degenerate into the simpler uncoupled equations if the Coupling
Coefficient “CC” is zero (which is seldom the case):

285
The “sprung mass response to unsprung mass input” gain is inversely proportional to the sprung mass and
directly proportional to the spring stiffness.
286
Reference [91], pp. 64-66.
287
The spring constants at an axle are double the spring constants at a wheel (in pitch/bounce there is no anti-roll
bar involvement). This is a necessary consequence of using a “Half-Car Model” with a total car sprung mass
value as is the custom (really a “Full-Car Model”).

145
SAWE Paper No. 3602
Category Number 31.0

a = (kr+ kf)/Ms = (2841.6+2364.0)/96.91 = 53.72


b = CC = (krlr - kflf)/Ms = (11877.9-10212.5)/96.91 = 17.185 (EQ. 7.10)
c = (krlr2+kflf2)/J = (49649.6+44117.9)/1130.5 = 82.94
(c – a) = 29.23
DI = Dynamic Index = K2/(lf lr) = 3.422/(4.32 x 4.18) = 0.65 (EQ. 7.11)
The principal modes of vibration frequencies are:

f1 = √c+(b2/(K2(c-a))) = √82.94 + (17.1852/(11.6654(29.23)) (EQ. 7.12)


= 9.15 rad/sec, or 1.46 cps (ideally, this would be < 1.3 cps)
f2 = √a-(b2/(K2(c-a))) = √53.72-(17.1852/(11.6654(29.23)) (EQ. 7.13)
= 7.27 rad/sec, or 1.16 cps
Now to find the nodes input these principal mode frequencies (in “radians/sec”) into:

X = b/(𝜔2-a)
X1 = 17.185/((9.15)2-53.72) = 0.57 ft (0.1737 m) (EQ. 7.14)
X2 = 17.185/((7.27)2-53.72) = -19.84 ft (-6.0472 m) (EQ. 7.15)
Which may be plotted, remembering that the c.g. is the origin, according to the rule
that forward from the c.g. is positive while rearward is negative. It is noted that the
Dynamic Index in Pitch is 0.65, so the Conjugate Points of Percussion are not at the axle
lines and therefore not very significant in the ride dynamics of this vehicle. The Spring
Center (“SC”) location is determined:

α = krl / (kf + kr) = 2841.6 (8.5)/(2364.0+2841.6) = 4.64 ft (1.4143 m) (EQ. 7.16)


β = kfl / (kf + kr) = 2364.0 (8.5)/(2364.0+2841.6) = 3.86 ft (1.1765 m) (EQ. 7.17)
The Conjugate Point “H” (forward) and “J” (rear) locations from the c.g. are “r” and
“s” respectively:

c2 = α x β = 4.64 x 3.86 = 17.91 ft2 (1.6639 m2)


e = α-lf = 4.64-4.32 = 0.32 ft (0.0975 m)
r = (K2-e2-c2)/2e + √ ((K2-e2-c2)2+4K2e2) / 2e (EQ. 7.18)

146
SAWE Paper No. 3602
Category Number 31.0

= (11.6654-0.322-17.91)/0.64+√((11.6654-.322-17.91)2+
4(11.6654)0.322)/2(0.32) = 0.57 ft (0.1737 m)
s = K2/r = 11.6654/0.57 = 20.41 ft (6.2210 m) (EQ. 7.19)

The frequencies about points “H” and “J” are:

fh = 1/2π √(2kf(α-e-r)2 + 2kr(β+e+r)2)/(Ms(K2+r2)) (EQ. 7.20)


= ((2(2364.0)(4.64-0.32-0.57)2+ 2(2841.6)(3.86
+0.32+0.57)2)/(96.91(11.6654+0.572)))0.5/2π
= 2.06 cps
fj = 1/2π √(2kf(α-e+s)2 + 2kr(s-β-e)2)/(Ms(K2+s2)) (EQ. 7.21)
= ((2(2364.0)(4.64-0.32+20.41)2+2(2841.6)(20.41
-3.86-0.32)2)/(96.91(11.6654+20.412)))0.5/2π
= 1.64 cps
When plotted against a side elevation of the Jaguar, all these points would look as
follows:

Figure 7.13 – 1958 JAGUAR XK150S KEY VIBRATION POINTS


What we have learned from all this is that the Jaguar’s principal vibrations are a bit
stiff, as we would expect for a sports car of this era, at 1.46 cps pitch and 1.16 cps bounce.

147
SAWE Paper No. 3602
Category Number 31.0

The Coupling Coefficient was not zero, so the two motions are linked to a certain extent,
which is not generally desirable 288. Also, the Dynamic Index in Pitch is 0.65, which is
appropriate for 1958, but means that some heterodyning can occur 289, and that the Conjugate
Points of Percussion will not play any role in this design. The analysis as performed on the
“baseline” is intended for a vehicle of conventional configuration, which the “advanced
concept” configuration definitely is not. Without resorting to more sophisticated analysis
techniques, the solution to this conundrum is to analyze the “advanced concept” in its
“equivalent conventional configuration form” as introduced in Chapter 5:

Figure 7.14 – “ADVANCED CONCEPT” PITCH/BOUNCE MODEL


This model, when subjected to the same pitch/bounce analysis as was performed on
the baseline Jaguar XK150S, results in:

a = (kr+ kf)/Ms = (3162.1+2446.6)/98.85 = 56.74


b = CC = (krlr - kflf)/Ms = (3162.1×4.61 -2446.6×3.67)/98.85 =
(14577.3- 8979.0)/98.85 = 56.81
c = (krlr2+kflf2)/J = (67201.4+32952.9)/1039.5 = 96.35
(c – a) = 39.61

288
The motions are only slightly coupled, and some authorities claim that a little coupling is desirable in special
cases such as when the “DI” is equal to 1.
289
In the 1920’s researchers Rowell and Guest had independently identified an interaction between the front and
rear suspension frequencies (“heterodyning”) as the fundamental cause of the poor ride (sudden lurches and
jumps) experienced in certain vehicles of the period; the Dynamic Index in Pitch was shown to be an important
tool for controlling this phenomenon.

148
SAWE Paper No. 3602
Category Number 31.0

DI = Dynamic Index = K2/(lf lr) = 3.242/(3.67 × 4.61) = 0.62


The principal modes of vibration frequencies are:

f1 = √c+(b2/(K2(c-a))) = (96.35 + (56.812/(10.4976(39.61)))0.5


= 10.20 rad/sec, or 1.62 cps (ideally, this would be < 1.3 cps)
f2 = √a-(b2/(K2(c-a))) = (56.74-(56.812/(10.4976(39.61)))0.5
= 7.00 rad/sec, or 1.11 cps
Now to find the nodes input these principal mode frequencies (in “radians/sec”) into:

X = b/(𝜔2-a)
X1 = 56.81/((10.20)2-56.74) = 1.20 ft (0.3658 m)
X2 = 56.81/((7.00)2-56.74) = -7.33 ft (-2.2342 m)
Which may be plotted, remembering that the c.g. is the origin, according to the rule
that forward from the c.g. is positive while rearward is negative. It is noted that the
Dynamic Index in Pitch is 0.65, so the Conjugate Points of Percussion are not at the axle
lines and therefore not very significant in the ride dynamics of this vehicle. The Spring
Center (“SC”) location is determined:

α = krl /(kf + kr) = 3162.15 (8.28) / (2446.6+3162.15) = 4.67 ft (1.4234 m)


β = kfl /(kf + kr) = 2446.6 (8.28) / (2446.6+3162.15) = 3.61 ft (1.1003 m)
The Conjugate Point “H” (forward) and “J” (rear) locations from the c.g. are “r” and
“s” respectively:

c2 = α x β = 4.67 x 3.61 = 16.85 ft2 (1.5654 m2)


e = α-lf = 4.67-3.67 = 1.00 ft (0.0929 m)
r = (K2-e2-c2)/2e + √ ((K2-e2-c2)2+4K2e2) / 2e
= (10.4976-1.002-1.6639)/2.00+√((10.4976-1.002-16.85)2+
4(10.4976)1.002)/2(1.00) = 1.23 ft (0.3749 m)
s = K2/r = 10.4976/1.23 = 8.56 ft (2.6091 m)
The frequencies about points “H” and “J” are:

fh = 1/2π √(2kf(α-e-r)2 + 2kr(β+e+r)2)/(Ms(K2+r2))

149
SAWE Paper No. 3602
Category Number 31.0

= ((2(2446.6)(4.67-1.00-1.23)2+ 2(3162.15)(3.61
+1.00+1.23)2)/(98.85(10.4976+1.232)))0.5/2π = 2.28 cps
fj = 1/2π √(2kf(α-e+s)2 + 2kr(s-β-e)2)/(Ms(K2+s2))
= ((2(2446.6)(4.67-1.00+8.56)2+2(3162.15)(8.56
-3.61-1.00)2)/(98.85(10.4976+8.562)))0.5/2π = 1.59 cps
When plotted against a side elevation of the “advanced concept”, all these points
would look as follows:

Figure 7.15 – 2014 ADVANCED CONCEPT KEY VIBRATION POINTS


What we have learned from all this is that the “advanced concept” principal
vibrations are maybe even a little stiffer at 1.62 cps pitch and 1.11 cps bounce. The Coupling
Coefficient is not zero, so the two motions are linked to a certain extent, and is even further
from zero than the “baseline”. Also, the Dynamic Index in Pitch is 0.62, just slightly less
than the “baseline”, and that means that some heterodyning is even more likely occur, and that
the Conjugate Points of Percussion still will not play any significant role.
The “advanced concept” increased stiffness in pitch (1.62 cps vs. 1.46 cps), and its
increased pitch node-to-driver distance (2.46 ft, or 0.75 m) over the “baseline” (1.75 ft, or
0.53 m), has significant implications. The frequency would indicate a somewhat stiffer ride in
pitch (the baseline was a bit stiff to begin with), as values under 1.3 cps (and over 0.5 cps) are
generally sought after. However, the concern noted in Chapter 4 that the new configuration
might exhibit “pitch instability” (too soft) was successfully nullified by the increased stance,
although perhaps too much so. The detrimental significance of all this for the front seat
occupants, which results from the combined effect of greater range of motion plus a more

150
SAWE Paper No. 3602
Category Number 31.0

rapid rate of that motion, is that the ride in pitch might be nearly 60% worse than the already
objectionable “baseline”.
However, so much could vary with regard to spring rates, locations, and mass
properties in the course of going from a concept to a real finished design that all this is hardly
a disqualifier from proceeding, but it is yet another concern to keep in mind if an “advanced
concept” design were to actually be pursued.
A chapter on “Vertical Performance” would not be complete without some discussion
of “flat ride”. “Flat ride” was a term coined by famed automotive dynamicist Maurice Olley
to label a desirable relationship between the front and rear suspensions of a conventional
automotive configuration. “Flat ride” required that, at a certain vehicle speed “V”, the
distance between the front and real suspensions (the wheelbase “lwb”) be such that the time
lag “tθ” (= V/lwb) set the phase relationship (“θ”) of the pitch oscillations, resulting from the
stimulus of front and rear suspensions from some road irregularity, so that they tended to
cancel out. Of course, the front and rear bounce frequencies figured in this too, and it was
determined that for most passenger cars the rear should be about 15-to-20% “stiffer” than the
front.
For a 1958 Jaguar XK150S (the “baseline”), which was a sports car with a top speed
of about 132 mph (212 kph), the “flat ride” target speed seemed to be around 66.5 mph (107.0
kph), as has been demonstrated by this author elsewhere 290. To accomplish this, the rear
suspension was 13% “stiffer” than the front which, as opposed to the reverse, resulted in a
much smoother pitch motion:

Figure 7.16 – “FLAT RIDE”

290
Reference [91], pp. 70-72.

151
SAWE Paper No. 3602
Category Number 31.0

How “flat ride” would be determined, let alone accomplished, for a multi-axle
configuration such as the “advanced concept” is less clear, but is demonstrably possible.
Military vehicles, such as tanks, are essentially just multi-axle vehicles that also supply their
own road surface (the “tracks”). For a tank pitch stability is even more important than for a
passenger car, as it is not just for reasons of ride comfort, but because a tank is a gun platform
and target acquisition requires the platform to be as steady as possible 291.

Figure 7.17 - MULTI-AXLE (TRACKED) VEHICLE


The ride motions of such vehicles are carefully studied to ensure that such conform to
rigorous standards such as the US Army AMM-75 Ground Mobility Model. When adjustment
of spring rates, dampers, and wheel locations simply won’t achieve standard performance
levels, then the alternative is an interconnected suspension, which may be either “passive” or
“active” (improvements in electronic control systems have “recently” led to the latter being a
favorite option).
Even for conventional passenger cars interconnected suspension to control roll (anti-
roll bar) and pitch (Moulton Hydrolastic/Hydragas, etc.) have long been used. The lowly
Citroen 2CV used a mechanical interconnection between fore and aft suspensions to achieve a
passive pitch control 292. Today such simple early passive systems have given way to full
active pitch/roll control on a great many road vehicles, the only restraints being cost,

291
Reference [93], pg. 475. Historically, tanks were required to be stationary before firing. However, even as
early as WW II certain German tank crews specialized in firing on the run, and with great success. Modern
electronic gun stabilization and aiming systems have rendered such feats less miraculous.
292
Reference [13], pp. 165-173, 180-183.

152
SAWE Paper No. 3602
Category Number 31.0

complexity, and weight 293. Whatever “challenges” the “advanced concept” might provide in
this regard should be easy to meet, but will incur some of those penalties….

293
A noted authority (Reference [4], page 205) estimates that a coupled suspension system may incur a 7%
suspension system weight penalty with respect to an uncoupled system.

153
SAWE Paper No. 3602
Category Number 31.0

8 - GYROSCOPIC TORQUE REACTIONS

A performance aspect that comes under the heading of “vertical performance”, but in
the context of this paper deserves to be considered in a chapter of its own, has to do with
“gyroscopic torque reactions”. These reactions occur when the axis of rotation of a spinning
mass is forced to change angular orientation (“forced precession”). This topic has already
been touched on in Chapter 5 when the “advanced concept” engine location/orientation had to
be adjusted from that of the “baseline” in order to maintain a reasonable vehicle center of
gravity location 294.

However, that adjustment involved the gyroscopic reactions of the rotating masses of
the engine; now the concern will be the gyroscopic reactions of the rotating unsprung
masses. Early in the history of the automobile a common configuration was that of a tricycle;
a single front wheel simplified the steering mechanism while drive was provided through the
rear wheels. However, that soon gave way to the four-wheeled, front engine, rear wheel drive
(RWD) paradigm that was to dominate for over sixty years. Early versions of this paradigm
tended to have the engine just aft of a simple beam axle, supplying propulsion via a driveshaft
to a “live” rear axle. Up until about 1915-1920 this was a perfectly suitable arrangement; the
power of the engines, condition of the roads, and design of the suspension ensured reasonably
safe operation at the prevailing low speeds.

But major developments came rapidly back then. Better roads and more powerful
engines, accompanied by better braking and tires, pushed general speed levels up. Brakes had
initially been supplied only at the rear axle, and early tires were more like bicycle tires than
the more robust specimens of today. With increases in speed and the unsprung rotating
inertia, the latter due to “balloon” tires and front brakes, an unsettling and dangerous
phenomenon began to be experienced: “shimmy” and “tramp” 295. Although there had been
disturbing vibrations encountered earlier in the development of the automobile, those had
tended to be mild and occurring at around 12 to 15 mph (19 to 24 kph). This new
phenomenon of “shimmy” would commence suddenly at speeds of 60 to 70 mph (97 to 113
kph) on smooth roads, perhaps after a single wheel encounter with a small dip or bump. A
left/right “flapping” of the front wheels would commence, similar in motion to the earlier
problem of “wobble”, but now much more violent and accompanied by a large scale up/down
rotational oscillation of the whole front axle known as “tramp”. Palliatives that had been used

294
Incidentally, this action also results in a possible enhancement of maneuver performance, although one that
does not directly reveal itself in an analysis such as conducted in Chapter 6.
295
SAE J670 2008, pg. 63 defines “shimmy” as “…accompanied by appreciable tramp…”.

154
SAWE Paper No. 3602
Category Number 31.0

to treat earlier occurrences of unsettling vibrations, involving stiffness and damping changes,
now proved insufficient 296.

“Shimmy” became the automotive crises of the day. Unless a solution to the problem
could be found all automotive progress would be stalled. Men like W.A.Robotham of Rolls
Royce, and Maurice Olley of Cadillac, went to work in their laboratories using “Bump Rigs”
to do resonance testing of various suspension designs. What they ultimately learned included
not only the solution to “shimmy”, but the scientific understanding of automotive shock,
vibration, and ride motions; the dawn of automotive dynamics had arrived 297.

The mechanism of “shimmy” proved to be fairly complex, but to simplify it for this
discussion let it suffice to say that it involved the unsprung rotational inertias at the front
axle. The tire, wheel, brake drum, and anything rotating in unison, constituted a fairly large
gyroscopic mass spinning about an axis coincident with the axle line. Any forced change in
the angular orientation of the spin axis, a “forced precession”, would cause a “kick back” or
torque reaction in accord with the relation 298:

T = I ωr ωp (EQ. 8.1)

Where:
T = Gyroscopic torque reaction due to precession (lb-ft).

I = Rotational inertia of rolling mass (lb-ft-sec2).

ωr = Angular velocity of rolling mass (radians/second).

ωp = Angular precession velocity due to turn or camber change (rad/s).

The increasing unsprung rotational inertia (“I”) and greater speeds (mainly “ωr”,
but also “ωp” to some extent) had resulted in an ever increasing reaction (“T”) until it became
too much for the suspension and steering to control. Since the front axle unsprung rotational

296
Reference [53], pp.611-612.
297
Ibid, pp. 613-619. H.S Rowell had published his work on automotive ride motions by 1923, with J.J. Guest
following by 1928. Robotham began his “Bump Rig” study in 1925; Olley began his study with a “Bump Rig” in
1930. By 1932 Olley had built his famous “K2-Rig” (a modified Cadillac limousine) for related studies of ride
motions.
298
Reference [18], pp. 255-257.

155
SAWE Paper No. 3602
Category Number 31.0

masses were connected to each other by the beam axle and the steering mechanism, any
vertical “forced precession” of the spin axis on one side would be doubled by the other side
acting in unison; the resulting gyroscopic reaction causing a powerful “kick back” through the
steering mechanism. The result would be a violent oscillation at the steering wheel while the
vehicle seemed to hop, skip, and jump down the road. Regaining control of the vehicle was
difficult, and many did not succeed (seasoned drivers utilized the handbrake, which worked
on the rear wheels only, as their best option for regaining control). An authority on machine
dynamics, Erskine Crossley, expressed the situation this way299:

“…note the cause of front wheel shimmy. One wheel runs over a
stone or a hole in the road. To prevent the rise or drop (which involves
an angular change in the wheel axis) the wheel will precess right or left,
so that…vibration is set up from the initial twist and subsequent caster
action.”

Gyroscopic torque adversely affects vehicle performance in ways other than


“shimmy”, as Crossley goes on with the comment 300:

“The gyroscopic torque from any pair of rolling wheels rounding a


corner is…shown to augment the usual centrifugal effect. However…it
is generally reckoned to be so much less than centrifugal effects that it
is ignored…”

With regard to “shimmy”, men like Maurice Olley soon realized that by going to an
independent front suspension (IFS) the principal link between the left and right rotating
masses would be severed, thereby cutting the magnitude of any gyroscopic reaction in half.
As a bonus, it would then be possible to move the engine forward to a location right between
those masses, thus improving the ride (the engine relocation increases the vehicle pitch
radius of gyration, thus improving the Dynamic Index in Pitch). The fact that this single
action “killed two birds with one stone” made the additional expense of IFS acceptable to
management, and IFS became part of the automotive paradigm that exists to the present day.

However, gyroscopic reactions resulting from the forced precession of the unsprung
rotational inertias still continue to fight suspension and steering movement to this day, but
now generally only present a problem in racing where every advantage must be exploited.
Hence Colin Chapman’s concern as expressed in Chapter 1. Colin was not concerned with
some possible resurrection of “shimmy” and “tramp”, but with improving control and
cornering power, and he did so by diligently working to reduce the rolling unsprung mass to
the greatest extent possible.
299
Ibid, page 255.
300
Ibid, pages 256-567.

156
SAWE Paper No. 3602
Category Number 31.0

The transition from “baseline” to “advanced concept” represents a significant


reduction in the unsprung masses/rotational inertias. The determination of the exact
improvement in control and cornering is beyond the ability of this author, but the decrease in
the magnitude of the gyroscopic reactions can readily be calculated. A determination of “ωp”
can be made from a consideration of the unsprung mass-suspension spring systems, and a
determination of “ωr” will be made by assuming a 60 mph (96.6 kph) encounter with a 2 in
(5.1 cm) bump 301. Given the respective unsprung masses from Appendix B, the rotational
inertias may be estimated:

Table 8.1 – ROLLING MASS SUBJECT TO FORCED PRECESSION

Since the front axles are IFS, the front axle values given in Table 8.1 represent only
one wheel. However, the rear axles are of the beam type, and therefore the rear axle values
represent the entire rotating mass at the axle (both wheels, etc.). Figure 8.1 allows for the
determination of the angular velocities of the rolling masses using the rolling radii:

ωrf = [60 / (5280/3600)] / (12.63 2 π /12) = 13.31 RPS, or 83.63 rad/sec

ωrh = [60 / (5280/3600] / ( 6.31 2 π / 12) = 26.63 RPS, or 167.35 rad/sec


A kinematic study of the suspensions as depicted in Figure 8.1 reveals that rolling over
the 2 in (5.1 cm) bump will cause about a 5 degree camber change for the “baseline” IFS, and
a 3 degree change for the “advanced concept” IFS. The “half-radius” 152/46R8 tire will climb
and surmount the bump quicker than the “full-size” 152/92R16 tire; this is behavioral
difference that must also somehow be accounted for. To do so, it will be assumed that the
time it takes for a tire to climb and surmount the bump will be equal to the time it would take
the vehicle to traverse a distance equal to the tire section height “Sh”. The “full-size” tire “Sh”

301
Reference [91], pp. 49-50.

157
SAWE Paper No. 3602
Category Number 31.0

is 5.52 in (14.02 cm), and the “half-radius” tire “Sh” is 2.76 in (7.01 cm), therefore the two
elapsed times are:

tf = (5.52/12) / ((60) (5280/3600)) = 0.00523 seconds

th = (2.76/12) / ((60) (5280/3600)) = 0.00261 seconds


Therefore, the average precession angular velocities will be:

ωpf = 5/0.00523 = 956.0 deg/sec, or 152.12 rad/sec

ωph =3/0.00261 = 1148.4 deg/sec, or 182.94 rad/sec


Now the respective torque reactions may be obtained:

“Baseline”: Tb = I ωrf ωpf =0.45 83.63 152.12 = 5724.81 lb-ft

“Advanced”: Ta = I ωrh ωph =0.04 167.35 182.94 = 1224.60 lb-ft

Figure 8.1 – IFS CAMBER CHANGE FOR 2 INCH BUMP

Because the camber change is so dependent on details of suspension design, and


because the time intervals were so crudely determined, let’s see what happens assuming the
same camber change of 4 degrees (0.63662 radians) for both IFS systems over the same time
interval of 0.004 seconds, for which the torque reaction results are:

158
SAWE Paper No. 3602
Category Number 31.0

“Baseline”: T = 0.45 83.63 rad/sec 159.16 rad/sec = 5989.75 lb-ft

“Advanced”: T = 0.04 167.35 rad/sec 159.16 rad/sec = 1065.42 lb-ft


These results still look very favorable for the “advanced concept” at about 18% of the
“baseline” reaction, whereas the previous result was about 21% of the “baseline” reaction.

At the rear the camber changes involved in going over the 2 in (5.1 cm) bump are
more easily determined from the simple geometry of a beam axle to be 2.235 degrees
(0.35568 radians) for the “baseline” and 2.151 degrees (0.34233 radians) for the “advanced
concept”:

Figure 8.2 – REAR BEAM AXLE CAMBER CHANGE FOR 2 INCH BUMP

Here the resulting reactions are (note that the rolling mass angular velocities have
changed slightly due to the different rolling radii at the rear):

“Baseline”: T = 1.07 82.77 rad/sec 68.01 rad/sec = 6023.23 lb-ft

“Advanced”: T = 0.10 166.04 rad/sec 131.25 rad/sec = 2179.35 lb-ft


So at the rear the “advanced concept” has about 36% of the “baseline” torque reaction.

The curious thing about all this is that it suggests the possibility of a return to a beam
axle front suspension. If the “baseline” was to retain its IFS, but the “advanced concept” was
to revert to a beam axle front suspension, then the camber change would be 2.132 degrees
(0.33939 radians) and the torque results would be:

“Advanced”: T = (0.04 × 2) 167.35 red/sec 130.03 red/sec = 1740.88 lb-ft

159
SAWE Paper No. 3602
Category Number 31.0

The “advanced concept” still enjoys the advantage of a gyroscopic torque reaction that
is only about 30% that of the “baseline” IFS even while utilizing the stronger and simpler
(less expensive) beam axle. Of course, there would an increase in unsprung mass causing the
loss of some of its advantage in road contact in dip, but then adverse camber changes in roll
would be eliminated 302…

A radically different paradigm opens up many new possible directions….

302
Reference [4], pg. 163. Most of the listed limitations of front beam axles still apply, but to a greatly decreased
extent (e.g., “half-elliptic springs” need not be employed, and front wheel lock clearance is less of a concern
with “half-radius” tires).

160
SAWE Paper No. 3602
Category Number 31.0

9 - FUEL ECONOMY

“The case for reducing fuel consumption by reducing


…weight, in turn reducing translational and rotational
inertia drag on acceleration, as well as rolling resistance
and …, has been well argued.” 303

John Fenton, BSc, MSc

Automotive “fuel economy” is dependent on how the energy produced (or stored)
within a vehicle is used. Per a Transportation Research Board (TRB) “Special Report 286”,
produced with funding by the DOT, the energy usage for a late model midsize passenger car
in the and highway and urban driving cycles is allocated as follows 304:

Table 9.1 – TRB CONVENTIONAL AUTO ENERGY USAGE

This table raises some immediate questions. The only losses that the table identifies
which can be readily attributed to vehicle “weight” (mass) are the percentages of energy lost
to deceleration (braking) and rolling resistance; where is the energy consumed in accelerating
the vehicle to speed? Is the energy used to accelerate to be considered as being the same as
that energy consumed by deceleration (braking)? Regarding this conundrum the following
authoritative quote applies 305:

“A quantitative measure of the effect of weight saving on fuel


economy…carried out…by a UK manufacturer…showed that

303
Reference [23], pg. ix (Preface).
304
Reference [99], pg. 40. Prof. Marc Ross, of the University of Michigan, seems to be in general agreement
with the “city” side of this table with values of “62, 18, 3, 2, 5, 5, 5” (Reference [71], pg. 16).
305
Reference [23], pg. ix (Preface).

161
SAWE Paper No. 3602
Category Number 31.0
Rev C

the proportions of fuel used to accelerate the mass of a typical


European car , when driven to the USA EPA
‘Highway/Urban’, and to the ‘EEC 15’ cycles, were 18/38,
and 25 percent respectively. Thus it was deduced that a one
third reduction in weight may be expected to yield some 6-12
percent fuel economy.” 306

The first thought that may occur regarding this quote is “How does it reconcile with
the information of Table 9.1”? Assuming that the “UK manufacturer” that carried out the
study was indeed capable of isolating the fuel consumed due to acceleration from all other
“engine drive” consumption, the Transportation Research Board energy consumption
breakdown can be modified as follows:

Table 9.2 – MODIFIED TRB CONVENTIONAL AUTO ENERGY USAGE

It is reassuring to note that both sources agree that it is the urban driving cycle which
consumes the most energy via the acceleration/deceleration of the vehicle, which makes sense.
What also makes sense is that the energy used to accelerate the vehicle is greater than the
energy dissipated in deceleration through the brakes by reason of the difference in effective
mass values involved in each case, as explained in Chapter 2. The “acceleration” fuel
consumption figures resulting from the UK study were assumed “hidden” in the engine drive
percentages of Table 9.1. Of course, the quote which led to this line of action should perhaps
not be taken at full face value either, as the original source is not available for study.

306
Mr. Gregg Peterson, Senior Technology Specialist for Lotus Engineering, at the 6th Annual Lightweight
Materials Summit held at the Hilton Double Tree Suites in Detroit, MI, on 26-27 August 2013; stated that a 30%
mass reduction would provide an 18%/24% increase in fuel economy. When it comes to fuel economy, this is
what passes for a close consensus.

162
SAWE Paper No. 3602
Category Number 31.0
Rev C

However, it does seem certain that a lot of energy is consumed for acceleration, and
that it can get “hidden” in the “engine drive mode” because it is not as easy to estimate
separately as is the aerodynamic losses, rolling resistance, etc. Moreover, even with the
“acceleration” values deducted from the “engine drive” losses, the greatest single source of
energy loss is still the internal combustion engine (ICE) itself. The ICE in its present
conventional reciprocating-piston gasoline-fueled form only manages to transform around
40% to 50% of the chemical energy it guzzles into mechanical energy that feeds into the
drivetrain, and the drivetrain extracts its own penalty, so that less than half of the chemical
energy produced actually does some useful work in moving the vehicle 307.

Because so much is lost to the thermal and mechanical inefficiencies (these two
sources of inefficiency are often confused and lumped together) of the reciprocating-piston
ICE (see Appendix F), the greatest opportunity for significant fuel economy improvement has
traditionally been with the engine and engine operation. This is why so much effort has been
expended over the last forty or more years to improve ICE efficiency: fuel injection (FI), high
energy ignition (HEI), variable valve timing (VVT) and lift, computer control, lean burn,
stratified charge, variable displacement, automatic intermittent (stop and start) engine
operation 308 (which directly addresses the “standby” energy loss), kinetic energy
regeneration/storage (which not only improves efficiency by “recycling” energy, but
addresses the “oversize engine for acceleration performance” problem). The late, great
“Smokey” Yunick 309 spent the last years of his life pursuing greater fuel economy through the
development of an ultra-efficient “adiabatic” ICE 310, which indicates another avenue of
possible improvement. And finally, thermal and friction losses aside, reciprocating engines
lose a great deal of energy to cyclically accelerating pistons, rings, wrist pins, con-rods,
valves, etc., up to speed then decelerating to a stop, only to begin accelerating in the opposite
direction (see Appendix F); these are engine losses but also mass acceleration (inertia) losses.

More recently, the methodology for improving fuel efficiency has placed greater
emphasis on mass properties, and in particular weight reduction. The effort to increase the
efficiency of the ICE itself seems to have run its course, and may now be entering a realm of
diminishing returns. ICE fuel efficiency has been enhanced greatly, but at a considerable cost
in complexity, maintenance and expense. The reappearance of hybrids, which add to the

307
Reference [99], pg. 40.
308
Reference [7], pg. 70. The 1980 Zenco Corporation’s “Fuel Saver” supposedly improved mpg by 7.5% as
tested on the EPA urban drive cycle. Although the “Fuel Saver” was not a commercial success due to its retrofit
nature, intermittent engine operation is an inherent feature of some present day production vehicles.
309
Henry “Smokey” Yunick (1923-2001) was an innovative mechanic and inventor best known for his work in
developing vehicles for NASCAR racing.
310
Reference [44], pp. 71-73. The Yunick “adiabatic” engine was a very high temperature, lean burn piston
engine made possible by the use of ceramic coatings and “plasma” spark ignition; Smokey’s engine design
patent number 4862859 (United States) expired in 2002; Smokey himself expired in 2001.

163
SAWE Paper No. 3602
Category Number 31.0
Rev C

complexity of the ICE the complexity of an onboard electrical power generation station, is
only a “half-way house” toward getting rid of the ICE altogether. The same may well be the
course of weight reduction, as the reductions being considered seem to mainly come from
the use of exotic materials such as carbon fiber composites, honeycomb panels, high strength/
ultra-light metals, plastics, and other stratagems that may be equally unwise in a commercial
setting. A recent (July, 2014) SAE journal article reports how “composite and hollow steel
coil springs are featured on the Ford lightweight concept vehicle” 311; there has to be a good
deal of desperation to compel a move so expensive for so little benefit.

A simple paradigm change such as advanced in this paper incurs none of the liabilities
of such desperate measures which are intended to shore up the conventional configuration.
With respect to the conventional “baseline”, the “advanced concept” has less aero drag, less
rolling resistance (due to reduced tire resistance coefficients and reduced normal load), plus
less acceleration/deceleration losses (due to reduced effective mass). How this would be
reflected in improved fuel consumption figures can only be crudely estimated given the
methodology of this chapter, but the attempt will now be made.

The reduction in aerodynamic drag is the easiest to determine; the drag coefficient
went from 0.42 for the “baseline” to 0.36 for the “advanced concept”, with an accompanying
reduction in frontal area from 18.53 ft2 (1.72 m2) to 18.45 ft2 (1.71 m2). That results in a total
drag reduction of -14.66%, which per the Table 9.2 fuel consumption factors results in a
Highway/City driving cycle improvement of -1.61%/-0.44% reduction in fuel consumption.

With regard to the rolling resistance the matter is a little more complex; the static and
dynamic resistance coefficients have changed, but so have the normal loads due to weight,
weight distribution, and aerodynamic lift. The summary of all the relevant data may be
tabulated as per Table 9.3:

Table 9.3 – “ADVANCED” vs. “BASELINE” CITY/HIGHWAY ROLL RESISTANCE

311
Reference [104], pg. 39.

164
SAWE Paper No. 3602
Category Number 31.0
Rev C

With the rolling resistance forces (lb) calculated in accord with the average speeds
(“Vivo”) of the EPA Highway/City fuel economy driving cycles, the “advanced concept”
registers a clear Highway/City -28.95%/-9.01% reduction in rolling resistance with respect to
the “baseline”. Again referring to the fuel consumption factors of Table 9.2, this indicates a
further reduction in fuel Highway/City consumption of -2.03%/-0.36%.

Next, a consideration of braking energy consumption is required as the braking effect


on Highway/City fuel consumption is explicitly called out in Table 9.2. The energy requiring
dissipation through braking is calculated as per Equation 5.5; note that it is directly
proportional to the braking effective mass “me”. For the “baseline” and the “advanced
concept” the input values used in the braking analysis of Chapter 5 (see Appendix B for mass
property tabulations) result in the following braking “me” values:

𝟏 𝟑𝟔𝟒𝟎 𝟖𝟓.𝟗𝟔+𝟑.𝟓𝟑(𝟒.𝟎𝟗)𝟐 𝟎.𝟗𝟕+𝟖𝟔.𝟐𝟕


𝑩𝒂𝒔𝒆 𝒎𝒆 = � + 𝟏𝟑.𝟔𝟏−𝟎.𝟖𝟒 𝟐
� = 𝟓𝟗. 𝟑𝟕 𝒔𝒍𝒖𝒈𝒔,
𝟐 𝟑𝟐.𝟏𝟕𝟒 � � ×𝟑𝟐.𝟏𝟕𝟒
𝟏𝟐
𝒐𝒓 𝟖𝟔𝟔. 𝟓𝟏 𝒌𝒈

𝟏 𝟑𝟔𝟐𝟗 𝟏𝟏.𝟓𝟓+𝟑.𝟓𝟑(𝟐.𝟎𝟒𝟎𝟐)𝟐 𝟎.𝟗𝟗+𝟏𝟏.𝟕𝟓


𝑨𝒅𝒗 𝒎𝒆 = � + 𝟔.𝟖𝟏−𝟎.𝟒𝟒 𝟐
� = 𝟓𝟖. 𝟒𝟖 𝒔𝒍𝒖𝒈𝒔,
𝟐 𝟑𝟐.𝟏𝟕𝟒 � � ×𝟑𝟐.𝟏𝟕𝟒
𝟏𝟐
𝒐𝒓 𝟖𝟓𝟑. 𝟓𝟏 𝒌𝒈
On this basis the “advanced concept” should dissipate -1.5% less energy through
braking than the “baseline”, which translates per Table 9.1 to Highway/City fuel consumption
decreases of -0.03% /-0.09%, respectively.

Lastly, the varying effective mass due to gearing would result in different acceleration
“me” values as well. Most of the Highway acceleration would be done in the higher gears:

3rd Gear: 3974Adv/4054Base = 0.980, or -2%

4th Gear: 3891Adv/3972Base = 0.980, or -2%

It is also reasonable to assume that most of the City acceleration would be done in the
lower gears:

1st Gear: 5216Adv/5296Base = 0.985, or -1.5%

2nd Gear: 4204Adv/4284Base = 0.981, or -1.9%

165
SAWE Paper No. 3602
Category Number 31.0
Rev C

As decided earlier, the factors to be applied toward reduction in acceleration fuel consumption
are .18 Highway and .31 City, which when applied to the indicated effective mass reductions
result in an indicated Highway/City fuel consumption improvement of -0.36 %/-0.65% .

The complete results of this fuel economy investigation are summarized in Table 9.4:

Table 9.4 – FUEL ECONOMY SUMMARY

If the “baseline” vehicle had test EPA fuel economy mpg figures of 11.9 City/18.3
Highway with a 14.8 Combined (0.55 × 11.9 + 0.45 × 18.3) mpg, then by this comparison the
“advanced concept” vehicle would have estimated EPA mpg figures of 12.4 City/18.6
Highway, for a 15.2 Combined (0.55 × 12.4 + 0.45 × 18.6) mpg.

Although the total fuel economy savings are significant, the portion of those savings
due to just mass properties effects alone seems disappointingly small at -0.41%/-0.74%.
This paper is conservative in the estimation of fuel economy improvements in that only the
estimated improvements that can be clearly demonstrated are quoted in Table 9.4; there is a
definite possibility that the actual improvements could be greater, but for that determination a
more sophisticated methodology is required.

166
SAWE Paper No. 3602
Category Number 31.0

10 – SAFETY

Automotive safety has two major aspects, the “active” and the “passive”. The “active”
aspect of automotive safety is the subject of accident avoidance, which involves such things
as braking, acceleration, cornering, stability, visibility, steering, et cetera. The “passive”
aspect of automotive safety involves accident minimization, the reduction of the
consequences of accidents when they inevitably occur.

It has been the “active” or performance aspects of the “advanced concept” that have
been the primary focus of this paper. The first clear indication of superior active safety on
behalf of the “advanced concept” occurred in Chapter 4 with the discussion of the “squeegee
length” (“LS”) and the critical velocity at which hydroplaning occurs (“He”); with regard to
both these important wet weather safety parameters the “advanced concept” has been shown
to be overwhelmingly superior to the conventional configuration; this is an unexpected benefit
of what had started out to be merely a exploration of expected improvement in acceleration
and deceleration through a mass properties driven design. Offsetting this improvement
somewhat is the possible reduced tire load capacity (“Clap”)/“standing wave” critical velocity
(“VC”)/mileage life (“M”). These are factors that can be dealt with through careful tire design,
reduced GVWR, and an acceptance of possible greater operational expense due to more
frequent tire replacement.

With regard to longitudinal performance, acceleration and braking, sizable


improvements were anticipated at the onset of this study. These improvements were to come
about through the reduction in effective mass, but practical considerations (the doubling of
the number of wheels necessitated by tire load capacity, and the retention of conventional size
brakes but relocated inboard, as necessitated by brake capacity) resulted in greatly reduced,
but still significant improvements. Such improvements, although diminished, still represent an
improvement in “active” safety, especially the braking.

The investigation of the lateral performance proved to hold yet more surprising results.
Analysis involving the “Static Stability Factor” (“SSF”), the NTSHA measure of vehicle
resistance to rollover, indicated an increase of about +15% due to reduced VCG and increased
track. This is a very significant improvement in safety, given that accidents involving rollover
are especially dangerous, and that improvements in SSF supposedly were on the order of
about +2.2% over the period 1978-2003 312. An admittedly simplified lateral acceleration
capability analysis showed an improvement of almost +50% (!!) for the “advanced concept”
over the “baseline”, which would seem almost “too good to be true” if it were not for the fact
the calculation is so simple. Also, the “advanced concept” shows an improved roll gain of -
8.9%, which is very believable. The “advanced concept” stability analysis showed it to be
understeering, but much more near neutral, with sharper steering. The “advanced concept”
transient response would likely be slightly longer in “rise time” (+4%) than the “baseline”,
but with zero “decay time”! Furthermore, it should be noted that the lateral
312
Reference [86], “Exhibit 2”. However, that improvement was for “passenger cars”, which ignores the present
tendency to use light trucks and vans in lieu of genuine passenger cars. Also, this author suspects the rate of
improvement has decreased over time due to the “law of diminishing returns”.

167
SAWE Paper No. 3602
Category Number 31.0

acceleration/transient analysis methodology did not include the possible benefits of the
gyroscopic torque reaction reductions (engine rotating mass and unsprung rotating mass),
nor did it include the real but hard to quantify benefit of having double the number of tire/road
contact areas, so there is reason to expect possible further improvements over those
determined herein when applying a more sophisticated analysis or with empirical testing.

The vertical performance analysis demonstrated a -50% reduction in road shock,


which may not be considered a safety issue in the same way that braking is, but a reduction in
road shock can plausibly be said to be an aid to control. On the other hand, the indicated
+29% improvement in maintenance of tire/road contact over dips is definitely safety related.
Like road shock, vibration can to some extent be safety related, so the fact that the road-
induced and internally-induced vibration peak values were halved is significant. However,
there was some increase in the internal vibration gain values in the 5.5 to 14 Hz range. Also,
the impact of a suspected increase in unsprung mass vibration input could not be
quantitatively evaluated due to lack of an appropriate equation for gain determination, but a
crude estimation indicates an increase of effect on the sprung mass +38% front/+147% rear.
Ride motions fall into the same somewhat ambiguously safety-related category as road shock
and vibration, and the somewhat stiff motions (1.62 Hz in pitch and 1.11 Hz in
bounce)/“poor” pitch node location (2.5 ft forward from the front seat occupant CG) are
matters readily dealt with in detail design, as would be any “flat ride” tuning of the
suspension.

Qualitatively indicated, but quantitatively undetermined, the beneficial gyroscopic


torque reactions resulting from the ICE reorientation within the chassis of the “advanced
concept” was noted in Chapter 5 (Acceleration Run 9). Another torque reaction improvement,
this one due to reduction of the rolling unsprung inertia, was quantitatively evaluated as
averaging about a -75% improvement over the “baseline”. Both types of gyroscopic torque
reductions are certain to have a beneficial effect on cornering and steering, and therefore a
beneficial effect on safety.

That concludes al that can be said with regard to the “advanced concept” and “active”
safety in avoiding an accident. With regard to the “passive” safety aspect, which concerns
providing occupant protection when “active” safety fails, the two most basic design
considerations have been succinctly stated by a noted authority 313:

313
Reference [24], pg. 1. Historically this sort of observation goes back to Hugh DeHaven, the father of crash
safety study, in the early 1900’s. Automotive manufacturers have been able to reasonably comply with the first
requirement for passive safety, but adequately complying with the second requirement seems to elude them,
especially with regard to roof structure. For nearly fifty years automakers have fought a dogged rear guard action
against the implementation of reasonable roof crush safety standards by utilizing the most sophistic of
arguments; yet in automotive sports every governing body that sanctions stock car events requires roll bars/roll
cages to minimize roof crush in the interest of driver safety.

168
SAWE Paper No. 3602
Category Number 31.0

“…(1) reduction of decelerations on the occupant…by


intentional deformability of the structure; (2) limitation of
passenger's compartment deformation to avoid passenger
crushing...should have basic importance...necessary to ensure a
"survival space"…in the event of severe crashes…”

The “advanced concept” has had no demonstrated impact in this regard as the
performance concerns of this paper mainly impact the “active” safety aspect. However, there
is a general observation that can be made. The “crumple zones” at the front and rear of
conventional automotive designs are compromised in their impact attenuation function by the
fact that those zones must be stiff enough to withstand the suspension (and engine) loads fed
into the chassis structure in those areas. In the “advanced concept” the suspension load input
locations are doubled, but the loads are essentially halved. The difference this could make in a
crash can be graphically illustrated as follows:

Figure 10.1 – THE CRASH: “ADVANCED CONCEPT” vs. CONVENTIONAL

169
SAWE Paper No. 3602
Category Number 31.0

As drawn both vehicles hit an unyielding barrier at sufficient velocity to cause a


frontal deformation of about 50 in (127 cm), resulting in idealized force-deformation (“F-x”)
curves as shown. The grey areas under the curves represent the work done in the course of
deformation, which in each case is equal to the energy kinetic (“½meV2”) of the vehicle at
impact. Because of the difference in effective mass, the “advanced concept” would have less
KE, which would be an advantage just in itself, but that is not the main point being made
here. Each vehicle’s occupants would be subjected to crash deceleration levels which would
vary in accord with “F/me” 314, and due to inherent differences in structure the “advanced
concept” would tend to subject the occupants to higher deceleration levels than the “baseline”
for about the first 22 in (56 cm) of crush. This is tantamount to saying that the “baseline”
vehicle might be easier on the occupants in lower speed crashes. However, in higher speed
crashes where the deformation would tend to be the full 50 in (127 cm) shown (or beyond),
then deceleration advantage would definitely tip in favor of the “advanced concept”. Overall,
the inherent nature of the “advanced concept” would allow for more crash deceleration level
tailoring of the vehicle structure, at least in the front, and would naturally tend to be more
accommodating for higher speed crashes, which generally tend to be the more deadly.

In summary, the “advanced concept” represents a significant advance in accident


avoidance; with a potential for improved energy absorption should an accident be
unavoidable.

314
This would seem to indicate that the lesser “me” vehicle, the “advanced concept”, would experience greater
deceleration levels, but remember that less “me” also means less KE, so….

170
SAWE Paper No. 3602
Category Number 31.0

11 – PRODUCIBILITY & MARKETABILITY

The single most practical concern regarding the reducibility of the advanced design
concept is the ability to obtain suitable tires. Certain ground breaking automotive designs of
the past have also required unusual tire sizes; the Citroen 2CV and the BMC (Austin/Morris)
Mini come to mind.

The 1948 Citroen 2CV, as designed under the direction of Pierre Jules Boulanger 315,
required unusually narrow 125R400 low pressure (20 to 23 psi) tires (mounted on 3-lug
4Jx15.75 pressed steel wheels) 316. The acquisition of such tires presented no problem because
the whole project had been started prior WW II by Pierre Michelin, head of Michelin Tire
(Citroen’s holding company since 1934). For the 1959 British Motor Corporation’s Mini
designer, Sir Alec Issigonis, the acquisition of an appropriate tire was a bit more difficult; he
initially tried to convince the Dunlop Tire Company to build a 5.2×8 size tire 317, but
ultimately settled for a larger but still diminutive 5.2×10 tubeless bias-ply (which Issigonis
secretly had in mind as a fall-back position well before negotiations started). This tire came to
be known as the Dunlop model “C41”; it was originally mounted on a 4-stud 3.5×10 pressed
steel wheel.

A preliminary search of present day tire availability reveals a number of existing tires
sized for 8 in wheels, but such tires are generally constructed for trailers, lawn
tractor/mowers, golf carts, and the nose wheels of certain light aircraft. Closer in suitability
are the tires intended for use in the ultra-high speed class of “go-kart” racing known as
“super-kart”, while perhaps the absolute closest in suitability are those tires intended for use
in a special class of circle track racing known as “Bandoleer”. This last, a youth orientated
type of racing, utilizes 178/50R10 tires designed to operate at normal loads of around 300 lb
(136 kg) and speeds up to about 70 mph (113 kph). None of these tire types, with the
exception of the trailer tires, are certified by the United States DOT for use on the street.
However, the rate of increase in the number of tire sizes available to the public for use on
passenger cars gives some cause for hope; around 1950 there were only 21 different sizes of
passenger car tire available in the United States, by 1975 that had grown to 91 different sizes,
and by 2003 there were 254 sizes 318.

That ultra-small automotive tires can be made which will stand up to the most rigorous
usage has been demonstrated in the past. There is the curious case of the Advanced Vehicle

315
Boulanger was possibly the only automotive design manager nearly as fanatical as Colin Chapman in the
matter of automotive weight reduction; the original 2CV weighed only about 1091 lb (495 kg), and was notable
for its low unsprung weight as well.
316
This became 125R380 (125R15) after 1960.
317
Note that this seems very close to the 152/46R8 “half-radius” tires of the “advanced concept”.
318
Reference [26], pg. 190.

171
SAWE Paper No. 3602
Category Number 31.0

Systems (AVS) Shadow, which debuted in the now defunct Canadian American (Can-Am)
Challenge racing series in 1970. Can-Am racing was possibly the least restrictive form of
automotive racing in the modern era; it encouraged designers to find innovative ways to push
the performance envelope. Don Nichols, founder of AVS in 1968, sought an unorthodox
designer who could take full advantage of the situation, and found one in Englishman Trevor
Harris.

Trevor’s design goal was to reduce aerodynamic drag to an absolute minimum, which
he did not so much by sophisticated design absorbing thousands of man hours in wind tunnel
studies, but by simply reducing frontal area to an absolute minimum. The design of the
consequent AVS Shadow Mk-I was that of an oversize go-kart powered by a Chevrolet V8
and riding on tiny wheels: 10 inch (254 mm) diameter wheels at front with 11 inch (279 mm)
wide tires, and 12 inch (305 mm) diameter wheels at the rear shod with 16 inch (406 mm)
wide tires. The vehicle was so low and flat that the driver essentially rode lying on his back;
the Shadow was referred to by some as the “two-dimensional” car 319.

Figure 11.1 – 1970 AVS SHADOW MKI

The design was very fast in a straight line; the theoretical top speed was over 250 mph
(402 kph), and in practice for a race at Mossoro in June of 1970 driver George Folder hit 190
mph (306 kph) on the long straight, beating everyone else by a full 20 mph (32 kph).
However, the car was to quickly display the cooling and handling problems 320 that were to
handicap it throughout its short career.

The unusual 11x10 and 16x12 tires for this ferocious, if ill fated, beast were custom
made by Firestone to withstand the rigorous demands of Can-Am racing, even at the projected

319
Reference [49], pg. N/A (Internet).
320
Such problems stemmed mainly from suspension, brakes, and the unusual front/rear weight distribution of
25/75. The fact that the driver’s position was extremely uncomfortable and adversely affected operation couldn’t
have helped the situation.

172
SAWE Paper No. 3602
Category Number 31.0

speed of 250 mph (402 kph). Although the Shadow would be haunted by numerous problems
through the two seasons it was campaigned, the tires did not add to those difficulties.

At this point it should be relatively easy to accept that the “advanced concept” would
be producible if suitable tires are obtainable, and that such procurement is quite possible.
After all, other than the tires everything else is dead conventional, just arranged in a different
way. There are no exotic materials or complex electronic controls necessary, nor does the
propulsion system require miraculous advances in battery or engine efficiency for the
“advanced concept” to bring its promised benefits to fruition. Many successful projects have
gone on to detail design with far more numerous and complex unresolved issues looming
ahead (although that is not a recommended practice). The “advance concept” may even
provide the opportunity to perhaps employ a lower level of technology, meaning less cost and
complexity, in certain areas (such as use of beam suspension, which is feasible due to the
relative decrease in load/gyroscopic torque reaction) while still providing meaningful
improvements in performance, fuel economy, and safety.

The matter of “advanced design” marketability also hinges mainly on one key factor:
appearance. An “advanced concept” vehicle that looks as it has been depicted so far in the
course of this paper is unlikely to find a niche in anyone’s heart. The history of the
automobile is replete with designs that had considerable technical merit, but which failed to
find favor with the buying public simply because of a lack of emotional appeal.

The Chrysler Airflow and the Ford Edsel are two commonly cited examples of failure
in the marketplace, although with somewhat different causation. The Airflow had its genesis
in aerodynamic studies begun in 1927 321 stimulated by the pioneering work being done in
Europe by the likes of Paul Jaray, Edmund Rumpler, and Hans Ledwinka. The Europeans had
achieved very aerodynamic teardrop body shapes largely though the use of a totally new
drivetrain paradigm, the rear engine. The goal of Chrysler engineers Fred M. Seder, Owen R.
Skelton, and Carl Brier was the same 322, but would be achieved differently. Per Brier’s
memoir the rear engine possibility had been considered, but ultimately rejected, due to
concerns regarding the effects of a rear engine on stability/handling and because of the
development expense of all new drivetrain components.

The configuration that the Chrysler team settled on was driven by those expense and
mass properties concerns. It would be a variation on the conventional front engine/RWD
configuration then prevalent, utilizing existing hardware but in a somewhat revised

321
Chrysler Engineering’s Carl Breer hired Bill Earnshaw, a Dayton, Ohio-based consulting engineer and
friend of the aerodynamic pioneering Wright Brothers, to conduct preliminary aerodynamic research. In his 1960
memoir Breer gives a much more fanciful tale for the inspirational basis behind Chrysler aero research.
322
Some credit for the Airflow/Airstream development should also go to Oliver Clark, Chrysler’s Chief Body
Engineer.

173
SAWE Paper No. 3602
Category Number 31.0

positioning. The engine would move forward from just behind the front axle line to a position
above that axle line, improving the ride through the effect on the Dynamic Index in Pitch of
increasing the pitch radius of gyration, and also improving the directional stability/handling
characteristics through a more forward CG. The engine’s CG was still behind the beam axle,
but now only slightly so and somewhat elevated. However, the engine longitudinal axis was
tilted about five degrees down at the rear to reduce the height of the driveshaft tunnel
intrusion into the passenger compartment (incidentally also slightly countering the increased
VCG). The engine relocation allowed movement of the entire passenger compartment
forward about 20 inches (51 cm), with the consequence that the rear seat was no longer
positioned right over the rear axle 323. The wheelbase/LOA was maintained, but the body was
widened at the front seat and narrowed at the rear seat, in exact reversal of the common
practice of the time. The result of this was a vehicle body with a broad, smooth nose and a
gradual tapering down toward the rear. This general shape and many detail improvements
resulted in a Cd of around 0.50 to 0.55. While this is not impressive today, at the time most
contemporary US cars had a Cd of 0.65 to 0.70.

However, in 1934 most US highways at the time were traversed at speeds of 45 mph
(72 km/h) or less, and streamlining provided little benefit at those speeds; besides, gasoline
was cheap. More significantly, the vehicle appearance was considerably different from the
norm, looking blunt and bulbous. It was totally an engineering design effort, Chrysler’s Art
and Color Department (styling) had been allowed no input on the radical body shape. When
the car debuted at the New York Auto Show on 6 January 1934, a poll revealed that the
appearance had aroused strong reactions from public of a “love it or hate it” sort. This
suggests that the car’s appearance, while significant, was not the sole determinant of its
demise.

Unfortunately, there were many other factors. Walter P. Chrysler had rushed the car
into production to meet the New York show date, and despite taking deposits from a good
number of the show goers, it would be a long time before any got their new vehicle. Not
helping the compressed production schedule was the fact that the Airflow was radical not only
in appearance, but in construction as well. Airflows utilized a structure type termed “bridge
and truss” in lieu of the conventional “body on frame”. “Bridge and truss” tended to be a sort
of “space frame”/“monocoque” hybrid attempt to retain the ease of conventional construction
the while reaching new heights of structural efficiency. This and other new aspects of Airflow
design helped slow production ramp-up and resulted in many early flaws reaching soon-to-be-
disgruntled customers. Worse, despite the new supposed structural efficiency, the resulting

323
This shift has been referred to as having some relationship to Chrysler’s “Cab Forward” design slogan of
1991; that later forward shift in passenger compartment location was also the result of some revision in front
engine relocation, but involved the adoption of FWD and a reorientation of the engine to a lateral position.

174
SAWE Paper No. 3602
Category Number 31.0

vehicles missed their ambitiously light design weight by about 450 lb (205 kg) 324, and
rumors circulated as to the new construction being dangerously weak 325. Lastly, despite an
intense effort to hold production costs down, the Airflows would cost 20 to 25 percent more
than the models they replaced, and this was during the Depression.

The Airflow was ultimately to succumb to poor sales, although a lawsuit for patent
infringement brought by Jaray Streamline Corporation of America regarding the aerodynamic
design couldn’t have helped either. The last Airflow left the factory in October 1937, bringing
the production total to just around 55,655 cars. Although Chrysler Corporation had long
established itself as being a technologically advant garde company, it would be over a decade
before that pioneering streak would dare reassert itself.
While the Chrysler Airflow was an engineering effort, representing a legitimate
attempt to advance the state of the art; the Ford Edsel was anything but. The Edsel was the
misbegotten child born of an unholy union of marketing and advertising. The impetus for its
creation was the idea that Ford’s premier brand, Lincoln, should move up-market to compete
with Cadillac, with the Continental model breaking out as a ne plus ultra. The Edsel was not
intended to be just one car, but a whole new range of cars, a new in-house brand to fill the
supposed vacuum made by Lincoln’s move up. The resulting Ford marketing hierarchy would
be: Ford, Edsel, Mercury, Lincoln, and Continental. The fact that the Edsel models would be
competing mainly against Ford’s upper level model and most of Mercury’s lower level
models seems to have gone unnoticed.
The launch of the Edsel line was preceded by one of the most massive marketing and
advertising campaigns ever made. Every feature of the new car line was subject to intensive
market research and analysis except, curiously, the styling. There were no prototype “show
cars” to gauge public reaction, no “sneak peeks” for the automotive press, no “test marketing”
with real potential customers. Perhaps the attractiveness of the styling was considered beyond
question; after all, it was supposedly the work of the talented Roy Brown Jr 326. Of course,
Roy Brown’s original design had undergone all the subsequent in-house “group think” and
“armchair quarter-backing” for which big corporations are renown; the final result, although
still respectable, bore little relation to the fine-lined graceful design Brown had started with.
Although the styling of any new vehicle is of premier importance, the name is also
very significant in generating an emotional resonance with the public. Again, as with the
styling, the process for determining an important aspect of salability was curious. Initial

324
The design goal had been to achieve a weight savings of 200 lb (90 kg) with respect to conventional “body
on frame”, but production changes resulted in a 250 lb (115 kg) weight penalty with respect to the conventional.
325
In the summer of 1934, Chrysler hired Barney Oldfield’s “Hell Drivers” racing team to roll Airflows in a sand
pit as part of an hourly demonstration of structural strength at the Chicago World’s Fair. Chrysler also combated
the rumors in dramatic fashion by demonstrating that the Airflow remained drivable even after being pushed off
a 110-foot (33-meter) cliff. It’s interesting to think about this when considering some modern manufacturer’s
claims that meaningful roof crush standards are unachievable.
326
Roy Brown Jr. (1916-2013) was a 1937 graduate of the Detroit Art Academy. He worked initially at General
Motors, then at Ford in both the US and the UK. He is credited with the design of the dashboard of the 1939
Cadillac, 1951 English Ford Zephyr, the 1955 Ford Futura (a show car later converted by George Barris into
TV’s “Batmobile”), the 1957 Edsel, the 1962 English Ford Cortina, and the 1966 Ford Econoline.

175
SAWE Paper No. 3602
Category Number 31.0

corporate tendency was to name the new brand “Edsel”, in honor of Edsel B. Ford, son of
corporate founder Henry Ford. Objections from the Ford family led to an amusing name
game. The advertising firm of Foote, Cone, & Belding was engaged to come up with a name;
they came up with 6,000.
Perhaps in desperation, the Pulitzer Prize winning poet Marianne Moore 327 was
unofficially invited to make suggestions; the result provides a bright spark of light in a dark
tale of corporate blundering. Among the suggestions she offered: “Mongoose Civique”,
“Pastelogram”, “Turcotinga”, “Varsity Stroke”, “Andante con Moto”, "Resilient Bullitt",
"Intelligent Whale", and “Utopian Turtletop”. Ultimately the decision came down: “Edsel”.
Along with the decision to call the new brand “Edsel”, the Foote, Cone, & Belding list
of infinite names was pared down to just four for the christening of the Edsel model line:
Citation, Corsair, Pacer, and Ranger, with the station wagon variants on those models named
Bermuda, Villager, and Roundup. The following massive advertising blitz even included a
television special, The Edsel Show. The key theme was that the Edsel was to be an “entirely
new kind of car”. The months of flack that preceded the introduction of the new brand on 4
September 1957 whipped the public into a frenzied anticipation of something akin to the
Second Coming of Christ. The banal reality of the awaited debut was destined to disappoint…
Although the hype promised something astoundingly different, the reality was that the
Edsel was a dead conventional collage of existing automotive components. The Edsel line did
not have dedicated manufacturing plants, but instead were built right alongside the Ford or
Mercury models from which they were derived. Most of the “unique” aspects of the Edsels
were frivolous, such as the “rolling dome” speedometer and more “idiot lights”. The
“Teletouch” push button transmission shifter in the center of the steering wheel was fairly
innovative 328 but troublesome; many later customers would opt for the standard column shift.
The only really advanced features were pretty much limited to the self-adjusting brakes, but
those had been pioneered by others much earlier.
Initially the Edsel seemed a great success. Sales were less than expected but still very
brisk. Soon however, as the fervor wore off and reality set in, sales plummeted. By 1958 only
four models were offered, and by 1959 that was down to two; the brand was on life support.
When the plug was pulled on 19 November 1959 the final production tally was 118,287 units
(including 2,846 “1960” units), about halfway to the sales breakeven point. Ford lost about
$350 million, or $2,831,563,927 in 2014 dollars.
There were numerous reasons for the Edsel’s failure, as was the case for the Airflow,
but the public disappointment has to be paramount. The styling was perhaps the most
disappointing aspect, though it wasn’t as bad as many have suggested. The styling was
reasonably handsome example of a mode in line with the times, but the public had been
anticipating something very different. There was also the dead conventional engineering; the

327
Marianne Moore (1887-1972) was a graduate of Bryn Mawr College, majoring in history, law, and politics.
She authored numerous publications, winning the Helen Haire Levinson Prize, the Pulitzer Prize, the National
Book Award, and the Bollingen Prize. From 1925 to 1929 she served as editor of the literary and cultural journal
The Dial. She was known for a sly wit that pricked the inflated egos of many of the pompous and overbearing.
328
The 1937 Cord 810 had a push button “pre-selector” transmission, but the buttons were mounted on the dash
and were considerably less troublesome.

176
SAWE Paper No. 3602
Category Number 31.0

mediocre performance of yet another big front engine RWD with IFS front and a beam axle
rear was the same as millions of cars already in production. Lastly there was the name; an
“Edsel” simply evoked the same sort of emotional response as would have been the case had
the car been named the “Poindexter”, and for a new car brand there was no long established
brand name loyalty to offset the dullness. The Edsel may have failed for insubstantial reasons,
but reasons which are significant in the marketplace 329.
For the “advanced concept” to be marketable reasons of that type have to be
addressed. As noted in Chapter 2, simply mating conventional automotive architecture with
such small wheels (not to mention an unusual number of them to boot) results in an
appearance certain to found “silly” and reminiscent of a “kiddie car” or “small trailer”.
Present day styling trends emphasize large wheels and tires, even to the point of decreasing
performance, as demonstrated by the Car and Driver “plus-sizing” study. The styling has to
be of “an all new car design” paradigm in accord with the all new configuration paradigm,
and it has to appeal to the emotions; an emotional commitment generally has to precede (or
supersede) the rational for mass appeal to be attained.
That same respect for the emotional has to also be maintained when it comes to all the
other sensory aspects of the new vehicle, not only the sight of it, but the sound, feel, and even
smell. And high on that list of aspects of emotional appeal is the name. To call the new
“advanced concept” configuration the “Centipede” or the “Half Mass” might have some
appeal, but only to an engineer.
As an indication of the possible new styling directions the “advanced concept” might
take, the following illustration is presented:

Figure 11.2 – ADVANCED CONCEPT “ANDROMEDA”

329
Severenson, Aaron; “Family Cars, Model Histories”, Internet, 29 May 2011. Anon., “Edsel”, Wikipedia,
Internet, 2014.

177
SAWE Paper No. 3602
Category Number 31.0

12 – CONCLUSIONS

“…friction would be off the chart when the thing turns......or tries to!
And the weight of four more tires, wheels and related suspension and
steering bits would so negatively impact fuel economy that it would
more than offset the reduction in rolling resistance. "Rolling
resistance" is just one small variable in the overall efficiency
equation...”
Bob Lutz 330
14 June 2014

The Bob Lutz quote was made as a direct comment on the “advanced concept” as
presented in this paper (admittedly he didn’t have any details), and it is quite understandable.
A “half-baked” attempt to trump the automotive world with some totally “out-of-the-box”
concept, one that violates every aspect of the conventional paradigm, occurs with an almost
periodic persistence over time (and a good number of such attempts have been documented in
this paper). For those who actually have served professionally in the automotive sector, it
becomes almost a reflex action to deny the feasibility of any seemingly outrageous concept
out-of-hand. However, the “advanced concept” as put forth in these pages is quite different
from those “half-baked” attempts, and in this chapter the many reasons why the “advance
concept” is so different will be reviewed.

There have been many historical vectors pointing toward the eventual emergence of
the “advanced concept” for a long time. The 1911 Reeves Octoauto, the 1970 AVS Shadow
MkI, the 1975 Elf Tyrrell P34, the 1976 March 2-4-0, the 1982 Williams FW08B, the 2004
Keio University Eliica, the 2004 Covini C6W…all displayed some beneficial aspect of the
“advanced” configuration, but in an incomplete and unsystematic way. All of the signs of the
hidden potential were there: the improved ride (Octoauto), the reduced drag (Shadow,
Tyrrell), the improved traction (Williams), the superior handling (Tyrrell, Williams), the
superb wet weather and braking performance (Covini, etc.)…

Within these pages, though the lens of mass properties, the “ultimate” configuration,
to which all those historical vectors have pointed, has been focused on and developed. The
attempt has been made to explore every possible performance aspect, and to coordinate the
330
Robert Anthony Lutz (1932- ) has served as Executive Vice President for BMW, Executive Vice President
of Ford Motor Company, Head of Chrysler Global Product Development, and Vice Chairman of Global Product
Development at GM. He retired from GM on 2 May 2010, and now heads his own consulting firm. In an E-mail
exchange this author made reference to the advantages of the “advanced concept”, and thereby received the
comment as quoted.

178
SAWE Paper No. 3602
Category Number 31.0

result into a synergistic whole insofar as the limited methodology employed allows. That the
“advanced concept” can simultaneously exhibit improved acceleration and braking, superior
handling, smoother ride, and greater fuel economy should at least seem to be a reasonable
possibility at this point; after all the analysis presented only a completely closed mind would
still reject it out-of-hand.

Most of these indicated performance advantages were anticipated long before the
actual commencement of this investigation as they flow simply and directly from the mass
properties changes envisioned. However, the enormous improvement in wet weather
drivability, and the possible reduction in high speed crash severity, came as pleasant surprises.
Looking at all the projected areas of improvement together, it would seem that the one aspect
of automotive performance which has benefited the most would be one not initially thought
of: safety. In both its passive and active aspects, automotive safety would seem to be greatly
advanced.

However, as the methodology employed has been somewhat limited, and since there
have been some recognized areas for concern; there is still a need to proceed with caution. No
matter what the anticipated benefits of any new concept, the end result ultimately depends on
the execution of that concept. Proper steering geometry 331/suspension geometry 332/drive
distribution 333 must be carefully insured. Brake proportioning 334 and pitch/roll control 335 are
other key design elements that require special attention for the “advance concept” to come to
fruition. And the most crucial element of all for the success of the “advanced concept” is the
acquisition of appropriate tires 336.

Lastly, it should be noted that the “advanced concept” as presented herein does not
represent the pinnacle of achievement, and would seem even more compelling when subjected
to further development. The “distributed traction” characteristic of the “advanced concept”
lends itself naturally to take best advantage of the benefits of “distributed drive” as proposed
by Hao, Chen, and Wang in their paper, and as demonstrated by Keio University with their
Eliica. That the automobile of the future will have complete electronic control over braking,

331
See the Tyrrell P34 steering arrangement of Figure 2.3; for a four-axle vehicle proper steering geometry
requires the inclusion of the first three axles.
332
The conflicting roll centers of the initial IBC-2036 design serve as an example of one way in which
suspension design can go awry. The suspension at each axle of a four-axle design must be made to work in
unison.
333
The March 2-4-0 serves as an example of incorrect drive distribution, and also to some extent of poor
steering geometry, and the consequences.
334
With modern ABS adapted to all eight wheels this becomes less problematic than it would have been
traditionally.
335
The Eliica points the way by obtaining such control through a hydraulically interconnected suspension; a
fully active suspension is just a natural progression.
336
The AVS Shadow and the Tyrrell P34 demonstrate that appropriate tires can be obtained (and the Tyrrell P34
also demonstrates what can happen when that appropriateness can’t be maintained).

179
SAWE Paper No. 3602
Category Number 31.0

drive, roll, pitch, ride, and stability seems certain, and the mass properties driven
configuration of the “advanced concept” represents a complementary means to physically
employ that new technology to greatest advantage. Many have gone a part of the way to this
goal, but at present it still remains a place “where none have gone before.”

180
SAWE Paper No. 3602
Category Number 31.0

REFERENCES

[1] Allen, R.W.; T.J. Rosenthal, and H.T. Szostak, “Steady State and Transient Analysis of
Ground Vehicle Handling”, Automotive Crash Avoidance Research, Warrendale, PA;
SAE SP-699, 1987.

[2] An, Jung-Chul, and Jin-Rae Cho; “Tire Standing Wave Simulation by 3D Explicit Finite
Element Method”, International Conference on Computational & Experimental Science,
Vol. 2, No. 3, pp. 123-128, Tech Science Press; Duluth, GA; 2008.

[3] Ashley, Steven; “Spin Control for Cars”, Mechanical Engineering, pp. 66-68, June 1995

[4] Bastow, Donald; Car Suspension and Handling, Plymouth, UK; Pentech Press Ltd., 1980.

[5] Bedard, Patrick; “Everything You Always Wanted to Know About Six”, Car and Driver,
pp. 97-98, September 1976.

[6] Beer, Ferdinand P.; and E. Russell Johnston Jr., Vector Mechanics for Engineers: Statics,
New York, NY; McGraw-Hill Book Company, 1962.

[7] Bolon, Paul; “Stop-Go Fuel Saver“, Popular Science, pg. 70, July 1980.

[8] Brach, Raymond M.; “Adhesion, Hysteresis, and the Peak Longitudinal Tire Force”,
Brach Engineering LLC, 2006.

[9] Brach, Raymond; and Matthew Brach; “The Tire Force Ellipse (Friction Ellipse) and Tire
Characteristics”, SAE Paper 2011-01-0094; Warrendale, PA, 2011.

[10] Brooke, Lindsay; “Chrysler Sees the ICE Future”, Automotive Engineering, pp. 16-19,
SAE, October 2013.

[11] Campbell, Colin; Design of Racing Sports Cars, Cambridge, Mass.; Robert Bentley Inc.,
1976.

[12] Campbell, Colin; The Sports Car, Cambridge, Mass.; Robert Bentley Inc., 1969.

[13] Campbell, Colin; New Directions in Suspension Design, Cambridge, Mass.; Robert
Bentley Inc., 1981.

181
SAWE Paper No. 3602
Category Number 31.0

[14] Canudas-DeWit, Tsiotra, Velenis, Basset, and Gissinger; “Dynamic Friction Models for
Tire/Road Longitudinal Interaction”, Vehicle System Dynamics, pp. 189-226, 2002.

[15] Chapman, Giles; The Illustrated Encyclopedia of Extraordinary Automobiles, New York
City, NY; DK Publishing, 2009.

[16] Clark, R.M. (Ed.); Jaguar Sports Cars 1957-1960; Cobham, UK; Brooklands Books, 1980.

[17] Costin, Michael, and David Phipps; Racing and Sports Car Chassis Design, Cambridge,
Mass.; Robert Bentley Inc., 1967.

[18] Crossley, F.R. Erskine; Dynamics in Machines, New York, NY; The Ronald Press
Company, 1954.

[19] da Cunha, Rodivaldo Henrique; Álvaro Costa Neto, Marcelo Prado, Odilon T. Perseguim,
and Daniel M. Spinelli; “Handling Analysis of a Three-Axle Intercity Bus”, Daimler-
Chrysler do Brasil Model IBC-2036 Intercity Bus Report by Debis Humaitá IT Services
Latin America Ltd, c. 2000.

[20] Dreher, and Yager; “Frictional Characteristics of 20×4.4, Type VII, Aircraft Tires
Constructed with Different Tread Compounds”, NASA Technical Note D-8252, 1976.

[21] Dixon, John C.; Suspension Geometry and Computation, Chichester, UK; John Wiley &
Sons Ltd., 2009.

[22] Dukkipati, Rao V.; and Jian Pang, Mohamad S. Qatu, Gang Sheng, Zuo Shuguang; Road
Vehicle Dynamics, Warrendale, PA; SAE R-366, 2008.

[23] Fenton, John; Handbook of Vehicle Design Analysis, Warrendale, PA; SAE R-191, 1999.

[24] Franchini, E.; “Crash Survival Space is Needed in Vehicle Passenger Compartments”,
SAE 690005, February 1969.

[25] Friedman, Dave; Indianapolis Racing Memories 1961-1969, Motorbooks International;


Osceola, WI, 1997.

[26] Gent, Alan N., and Joeseph D. Walter; The Pneumatic Tire, US DOT/NHTSA, DOT HS
810 561; Washington, DC; 2006.

[27] Gehm, Ryan; “Designing for Downforce”, Automotive Engineering, SAE International,
February 2014.

182
SAWE Paper No. 3602
Category Number 31.0

[28] Gillespie, Thomas D.; Fundamentals of Vehicle Dynamics, Warrendale, PA; SAE R-114,
1992.

[29] Grable, Ron; “Formula One Chevy: A Technical Overview of the World’s Best Handling
Production Car”, Motor Trend, pages 37-40, March 1983.

[30] Harned, Johnston, and Scharpf; “Measurement of Tire Brake Force Characteristics as
Related to Wheel Slip (Antilock) Control System Design”, SAE Paper 690214, 1969.

[31] Hartford, Bill; “What’s Behind Those Four Little Wheels?”, Popular Mechanics, pp. 47-
49 and 104, January 1977.

[32] Harris, Cyril, and Charles E. Crede; Shock and Vibration Handbook, New York, NY;
McGraw-Hill, 1961.

[33] Hayes, Russell; Lotus, The Creative Edge, Sparkford, UK; Haynes Publishing, 2007.

[34] Hayes, D.F.; and A.L. Browne, The Physics of Tire Traction, New York, NY; Plenum
Press, 1974.

[35] Hemmes, Henny; “New Michelin Tires Designed for EV’s and Hybrids”, The Auto
Channel, theautochannel.com, 29 June 2010.

[36] Hong, Patrick; “2013 Scion FR-S”, Road & Track, 9 December 2011.

[37] Horner, Jim; High-Performance Wheel & Tire Handbook, Osceola, WI; Motorbooks
International, 1988.

[38] Kasprzak, Edward M.; and David Gentz, “The Formula SAE Tire Test Consortium – Tire
Testing and Data Handling”, Warrendale, PA; Society of Automotive Engineers
International, SAE 2006-01-3606, 2006.

[39] Kawakami, Kiyomoto; and Hidetoshi Tanabe, Hiroshi Shimizu, Yoshida Hiroichi;
“Advanced Redundancy Technology for a Drive System Using In-Wheel Motors”,
WEVA Journal, Vol. 1, 2007.

[40] Koutný, F.; Geometry and Mechanics of Pneumatic Tires, Zlin, The Czech Republic;
2007.

[41] Krylov, V.V.; and O. Gilbert, “Physically Explicit Theory of Standing Waves in Tyres at
High Vehicle Speeds”, Proceedings of ISMA Including USD, pg. 4219-4232, 2010.

183
SAWE Paper No. 3602
Category Number 31.0

[42] Lamar, Paul; “More About Cornering Power”, Road & Track, October 1969; pp. 127-130.

[43] Lamar, Paul; “Aerodynamics & the Group Seven Racing Car”, The Aerodynamics of
Sports & Competition Automobiles, Proceedings of AIAA Symposium, 20 April 1968,
Los Angeles, CA; pp. 25-38.

[44] Lindsley, E.F.; “Smokey’s Hot Engines”, Popular Science, pp. 71-73, July 1980.

[45] Liu, Hao; Xinbo Chen, and Xinjian Wang; “Overview and Prospects on Distributed Drive
Electric Vehicles and Its Energy Saving Strategy”, Przeglad Elektrotechniczny
(“Electrical Review”), ISSN 0033-2097, R. 88 NR 7a, pp. 122-125, 2012.

[46] Ludvigsen, Karl; Colin Chapman – Inside the Innovator, Sparkford, UK; Haynes
Publishing PLC, 2010.

[47] McGivern, James G., and Ivan A. Shirk; “Standing Waves in Tires”, SAE Paper 770873;
Warrendale, PA; 1977.

[48] Mehmood, Arshad; Ahmad Ali Kahn, and Ayaz Mehmood; “Optimization of Suspension
Damping Using Different Mathematical Car Models”, International Journal of Mechanical
Engineering, Vol. 3, Issue 19, October 2013.

[49] Melissen, Wouter; “Shadow AVS MK I Chevrolet”, Ultimate Car Page,


ultimatecarpage.com, 19 November 2007.

[50] Metz, L. Daniel; Cheryl K. Akouris, Craig S. Agney, and Mark C. Clark; “Moments of
Inertia of Mounted and Unmounted Passenger Car and Motorcycle Tires”, SAE Paper
900760; Warrendale, PA; 1990.

[51] Metz, L. Daniel, and Duane R. Meyers; “Estimating Path Clearing Effects During
Potential Hydroplaning Through Use of Vehicle CAN Bus Data”, SAE Paper 2013-01-
0408; Warrendale, PA; 2013.

[52] Milliken, William F., and Douglas L. Milliken; Race Car Vehicle Dynamics, Warrendale,
PA; SAE R-146, 1995.

[53] Milliken, William F., and Douglas L. Milliken; Chassis Design, Principles and Analysis,
Warrendale, PA; SAE R-206, 2002.

184
SAWE Paper No. 3602
Category Number 31.0

[54] Mitra, Debojyoti; “Design Optimization of Ground Clearance of Domestic Cars”,


International Journal of Engineering Science and Technology”, Vol. 2(7), pp. 2678-2680,
2010.

[55] Lord Montagu of Beaulieu, Jaguar, Sparkford, UK; Haynes Publishing, 1979.

[56] Moore, Sam; “Forward Thinking Reeves Develops the Octoauto”, Farm and Dairy
(News), pg. C-2, 3 January 2013.

[57] Moran, Tim; “The Radial Revolution”, Invention & Technology Magazine, Volume 16,
Issue 4, pp. 28-39, Spring 2001.

[58] Morelli, Alberto; “A New Approach to Advanced Automotive Shapes”, SAE Paper 2000-
01-0491, March 2000.

[59] Newton, K.; W. Steeds, and T.K. Garrett, The Motor Vehicle; Jordan Hill, UK;
Butterworth-Heinemann, SAE R-298, 1996.

[60] Nye, Doug; “Racing On the Up and Up”, Road & Track, pp. 51-52, July 1980.

[61] O’Dell, John; “Tire Efficiency – Michelin’s Examining the Short and Tall of It”, Auto
Observer, edmunds.com, 3 June 2010.

[62] Paine, Michael; Michael Griffiths, and Nimmi Magedara; “The Role of Tyre Pressure in
Vehicle Safety, Injury, and Environment”, Caringbah, Austrailia; Road Safety Solutions,
24 April 2007.

[63] Pillai, and Fielding-Russell, “Tire Rolling Resistance From Whole-Tire Hysteresis Ratio”,
Rubber Chemistry and Technology, May 1992.

[64] Quiroga, Tony; “Tech. Dept.: Effects of Upsized Wheels and Tires Tested”, Car and
Driver, pg. 110, May 2010.

[65] Radlinski, R.; “Braking Performance of Heavy US Vehicles”, SAE Technical Paper
870492, 1987.

[66] Reimpell, Jörnsen; Helmut Stoll, and Jürgen W. Betzler; The Automotive Chassis; Jordan
Hill, UK; Butterworth-Heinemann, SAE R-300, 2001.

185
SAWE Paper No. 3602
Category Number 31.0

[67] Renfroe, David A.; Paul T. Semones, and Alex Roberts; “Quantitative Measure of
Transient Oversteer of Road Vehicles”, Engineering Institute LLC, Paper Number 07-
0217, c. 2005.

[68] Reynolds, Kim; “Science Friction – How to Corner the Fastest 911 Ever Built”, Motor
Trend, pp. 66-72, December 2011.

[69] Rosinski, Jose; “A Journalist’s Ultimate Track Test”, Motor Trend, pp. 92-96, August
1977.

[70] Ross, Peter; Lotus, The Early Years, Luton, UK; Coterie Press Ltd, 2004.

[71] Ross, Marc; “Fuel Efficiency and the Physics of Automobiles”, Contemporary Physics,
Vol. 38, Issue 6, pp. 381-394, 1997.

[72] Saarilahi, M.; “Soil Interaction Model, Appendix Report No. 6: Modeling of the Wheel
and Tyre 2, Tyre Stiffness and Deflection”, Part of a Deliverable Set of Work Packages to
the Finnish Government by the University of Helsinki under contract to Develop a
Protocol for Eco-Efficient Wood Harvesting, 1999-2002.

[73] Sears, F.W.; and M.W. Zemansky, University Physics, Reading, Mass.; Addison-Wesley
Publishing Co., 1967.

[74] Simanaitis, Dennis; “Tech Tidbits: Bailey Meets Bibendum”, Road & Track, pg. 108,
October 2010.

[75] Singer, Ferdinand L.; Strength of Materials, New York, NY; Harper & Row Publishers,
1962.

[76] Stater, Brian; “Dark Clouds Taint Lotus Founder Colin Chapman”, The Telegraph, 14
December 2002.

[77] Smith, Kevin; “Hold Everything! Pete Brock says Skinny Tires Work Better!”, Motor
Trend, pg. 33, May 1997.

[78] Smith, Sam;”The Lotus Elan: 1962-1973”, Road & Track, December 2012, pp. 70-71.

[79] Steeds, W.; Mechanics of Road Vehicles, London, Iliffe & Sons, 1960.

186
SAWE Paper No. 3602
Category Number 31.0

[80] Szostak, Allen, and Rosenthal; “Analytical Modeling of Driver Response in Crash
Avoidance Maneuvering Volume II: An Interactive Model for Driver/Vehicle
Simulation”, United States NHTSA DOT Report HS-807-271, April 1988

[81] Taborek, Jaroslav J.; Mechanics of Vehicles, Cleveland, OH; Penton Publishing, 1958.

[82] Taylor, R.K.; L.L. Bashford, and M.D. Schrock; “Methods for Measuring Vertical Tire
Stiffness”, Transactions of the American Society of Agricultural Engineers, Vol. 43(6),
pp. 1415-1419, 2000.

[83] Thomson, William T.; Vibration Theory and Applications, Prentice-Hall Inc., Englewood
Cliffs, NJ; 1965.

[84] Vable, Madhukar; Mechanics of Materials, Houghton, MI; Michigan Technological


University, 2012.

[85] Walker, Jearl; “The Amateur Scientist”, Scientific American, pp. 104-107, February 1989.

[86] Walz, Marie C.; “Trends in the Static Stability of Passenger Cars, Light Trucks, and
Vans”, NHTSA Technical Report DOT HS 809 868, June 2005.

[87] Wiegand, B.P.; “The Basic Algorithms of Mass Properties Analysis and Control (Rev.
A)”, Los Angeles, CA; SAWE #2067, 2011.

[88] Wiegand, B.P.; “Automotive Mass Properties Estimation”, Los Angeles, CA; SAWE
#3490, 2010.

[89] Wiegand, B.P.; “Mass Properties and Automotive Longitudinal Acceleration”, Los
Angeles, CA; SAWE #1634, 1984.

[90] Wiegand, B.P.; “Mass Properties and Automotive Lateral Acceleration”, Los Angeles,
CA; SAWE #3528, 2011

[91] Wiegand, B.P.; “Mass Properties and Automotive Vertical Acceleration”, Los Angeles,
CA; SAWE #3521, 2011.

[92] Wiegand, B.P.; “The Mystery of Automotive POI Values”, Weight Engineering, pp. 20-
24, SAWE,Winter 2011.

[93] Wong, Jo Yung; Theory of Ground Vehicles, Hoboken, NJ; John Wiley & Sons Inc.,
2008.

187
SAWE Paper No. 3602
Category Number 31.0

[94] Woodrooffe, and Burns; “Effects of Tire Inflation Pressure and CTI on Road Life and
Vehicle Stability”, Proceedings of the International Forum for Road Transport
Technology, pp. 203-221.

[95] Yoshida, Hiroichi; and Hiroshi Shimizu, “Development of High-Performance Electric


Vehicle ‘Eliica’”, Koyo Engineering Journal, English Edition No. 168E, 2005.

[96] Zenlea, David; “First Drive BMW i8”, Automobile, pp. 12-14, November 2013.

[97] ___________, “Six Another Way”, Motor Trend, pg. 95, August 1977.

[98] ___________, Weight Engineer’s Handbook, Society of Allied Weight Engineers; Chula
Vista, CA; 1968-1986.

[99] ___________, “Tires and Passenger Vehicle Fuel Economy”, Special Report 286,
Transportation Research Board; Washington, DC; 2006.

[100] ___________, “General Dynamics Stryker MGS”, Weight Engineering, pp. Cover, 2,
44; SAWE, Winter 2012-13.

[101] ____________,”The Jaguar XK150S Fixed-Head Coupe”, The Motor, 19 August 1959, pg.
118.

[102] _________“Cliff Allison, Formula 1 Racing Driver for Lotus and Ferrari”, The
Independent, 9 April 2005.

[103] __________“Hardware Variation Remains a Challenge for Modular Platform


Solutions”, Automotive Engineering, 1 July 2014.

[104] __________ “Advancing Structural Composites”, Automotive Engineering, SAE, pp.


36-40, July 2014.

188
SAWE Paper No. 3602
Category Number 31.0

AUTHOR’S BIOGRAPHICAL SKETCH

Brian Paul Wiegand, now retired, was a Senior Weight Engineer and Mass Properties
Handling Specialist for the Mass Properties Analysis and Control Group of Northrop
Grumman Corporation, Bethpage, NY. He is a 1972 graduate of Pratt Institute, Brooklyn, NY,
and a licensed Professional Engineer registered in the State of New York (#58470). He
continues to be an active member of the Society of Allied Weight Engineers (SAWE) and of
the Society of Automotive Engineers (SAE), and in 2009 he reaffirmed his long-standing
interest in automotive dynamics by attending the SAE Seminar “Vehicle Dynamics for
Passenger Cars and Light Trucks” (Troy, MI; August 11-13). He has presented five SAWE
papers, two award winning: “Mass Properties and Automotive Longitudinal Acceleration”
(SAWE #1634, 1984), “The Basic Algorithms of Mass Properties Analysis & Control”
(SAWE #2067, 1992), “Automotive Mass Properties Estimation” (SAWE #3490, “Special
Merit Award”, 2010),”Mass Properties and Automotive Vertical Acceleration” (SAWE
#3521, “L.R. ‘Mike’ Hackney Award”, 2011), and “Mass Properties and Automotive Lateral
Acceleration” (SAWE #3528, 2011). He has also published a number of SAWE Journal
(Weight Engineering) articles, including two award winning: “The Weight and C.G.
Implications of Obtaining Maximum Automotive Lateral Acceleration Levels” (Winter 1982-
83), “The Mystery of Automotive POI Values” (Winter 2011),”Ancient Mass Properties
Engineering” (Winter 2011-12, “Best Journal Article 2011-12”), and “An Uncertainty
Regarding C.G. Uncertainty” (Winter 2012-13, “Best Journal Article 2012-2013”). He has
published a SAE paper “Estimation of the Rolling Resistance of Tires” (SAE #2016-01-0445,
presented at the SAE World Congress in Detroit on 14 April 2016). He also is the recipient of
the 2011 SAWE President’s Award for contributions to the society, and two Northrop
Grumman “TAP” (Timely Awards Plan) awards.

189
SAWE Paper No. 3602
Category Number 31.0

APPENDICES

190
SAWE Paper No. 3602
Category Number 31.0

APPENDIX A – SYMBOLISM

A The amplitude of a vibration, which is the maximum “plus” or “minus”


deflection about the “rest” position.
ABS “Anti-skid braking system” refers to automatic systems for brake modulation
intended to prevent wheel “lock-up” under hard and/or low traction braking
conditions.
ACC The “Advanced Concept Configuration” is an automotive configuration which
has twice as many axles/wheels as the conventional configuration, and with the
tire/wheels sized to be one-half the radius of the conventional; suspension
spring stiffness is reduced accordingly. The axles are placed at equally spaced
intervals along a “wheelbase” sized to the maximum extant allowable
longitudinally within the body envelope. The “tracks” are likewise sized to the
maximum extant allowable laterally within the body envelope. The resulting
“distributed traction” configuration may be conventionally powered, but would
be most likely achieve its full performance potential with a “distributed drive”
system for power (see Reference [30]).
ADAMS “Automated Dynamic Analysis of Mechanical Systems” is a popular multi-
body dynamics simulation software package. ADAMS was originally
developed by Mechanical Dynamics Inc., but has been acquired by the MSC
Software Corporation. ADAMS is modular, allowing linkage with other
software packages and other customization to meet specific industry
requirements. Some existing “semi-customized” modules are ADAMS/Car,
ADAMS/Engine, ADAMS /Aircraft, and ADAMS Machinery.
AE The axle (final drive) efficiency (dimensionless).

AIAA The “American Institute of Aeronautics and Astronautics” is the professional


society for the field of aerospace engineering. The AIAA was founded in 1963
from the merger of two earlier societies, and today has about 35,000 members.

AR The conventional automotive configuration rear axle (final drive) gear ratio
(dimensionless).

ARB Stands for “Anti-Roll Bar”, which is a suspension component that increases
the roll stiffness while causing relatively little (or no) increase in vertical
spring rate. The first ARB patent was awarded to S.L.C. Coleman of
Canada on April 22, 1919.

191
SAWE Paper No. 3602
Category Number 31.0

ARP “Active Rollover Protection” refers to a system that detects a condition of


impending rollover and selectively applies the vehicle brakes so as to create a
counteracting roll moment. Future ARP systems may create a counteracting
roll through the action of a fully active suspension.

Aspect ratio, in the context of this paper, is usually tire cross-section height
divided by the nominal width and multiplied by a hundred: “SH/SN × 100”
(dimensionless).

AVS “Advanced Vehicle Systems”, founded in 1968 by Don Nichols, was the
corporate builder of the innovative Canadian-American (Can-Am) Challenge
racing series car the AVS Shadow in 1970.

AYR The “Ackermann Yaw Rate” is computed as per Equation 6.27 for a turning
vehicle (radians/second) and, when “compared” to the actual yaw rate “ω” in
the dimensionless ratio “AYR/ω”, can be used as a stability indicator.

AWD “All wheel drive”, generally meaning a constant drive configuration of 4WD
(no RWD option).

Ac The tire-to-ground gross contact area (in2, cm2).

Acf The tire-to-ground gross contact area at the front axle (in2, cm2).

Acr The tire-to-ground gross contact area at the rear axle (in2, cm2).

Af The frontal area of a vehicle (ft2, m2).

Ap The plan area of a vehicle (ft2, m2).

a Usually stands for an acceleration (ft/sec2, in/sec2, m/sec2, g’s), but can also
symbolize the distance between the front axle and the vehicle c.g. as per SAE
J670 2008 (in, cm). Also can mean ½ of the tire/road contact area length (mm)
as in Equation 4.14b, or can be a units conversion factor of value “15.7369”
(English) or “58.2587” (Metric) for use in Equations 5.15 and 5.16 for the
calculation of tire traction coefficients “μf” and “μr”. This symbol can also
represent the maximum traction force possible in the longitudinal direction for
an ellipse model of tire traction.

amax The maximum acceleration of a body, particularly one in SHM (ft/sec2,


m/sec2)

ax Longitudinal acceleration, usually in gravity units (g’s).

192
SAWE Paper No. 3602
Category Number 31.0

ay Lateral acceleration, usually in gravity units (g’s).

ayoverturn The vehicle lateral acceleration value at which overturn commences (g’s),
which may be roughly estimated per Equation 3.6 of Reference [63]:

𝒂𝒚𝒐𝒗𝒆𝒓𝒕𝒖𝒓𝒏 = 𝒕⁄𝟐𝒉𝒄𝒈

ayslide The vehicle lateral acceleration value at which slide commences (g’s), which
may be roughly estimated per Equation 6.2.

az Vertical acceleration, usually in gravity units (g’s).

α Greek lower case “alpha”, usually symbolizing an angular acceleration


(deg/sec2, rad/sec2), or the distance from the front axle to the “Spring Center”
as per Equation 7.16 (ft).

BEV “Battery Electric Vehicle”, meaning a vehicle that depends solely on batteries
for motivation and requires recharging from an external source.

BLMC The “British Leyland Motor Corporation” was a British automotive


conglomerate formed in 1968. It encompassed most of the British motor
industry and was partially nationalized in 1975. Continued decline led to a
dismemberment and sell off to various foreign companies in 2005.

BMC “British Motor Corporation Ltd” was formed in 1952 with the merger of the
Morris and Austin car companies; it later became part of the ill-fated British
Leyland Motor Corporation (BLMC).

BMW “Bayerische Motoren Werke” (Barvarian Motor Works) AG is a German


automotive manufacturer founded in 1916.

BNP “Bakker-Nyborg-Pajecka” or “Magic Formula” is the predominant model of


tire behavior; variations of the “Magic Formula” are used professionally by
competitive racing organizations, tire manufacturers, and automotive
corporations the world over.

BRDC The “British Racing Driver’s Club” is an organization which represents the
interests of professional racing drivers from the UK. The club was founded
April 1928 by Dr. J. Dudley Benjafield, one of an informal group of British
racing drivers known as the "Bentley Boys". By 1929 the club was organizing
and sanctioning various motor sport events.

193
SAWE Paper No. 3602
Category Number 31.0

b Symbolizes ½ of the tire/road contact width (mm) as in Equation 4.14b, or the


tire belt width (m) as used in the Krylov/Gilbert equations, or the basic tire
coefficient of traction potential (dimensionless). This symbol can also
represent the maximum traction force possible in the lateral direction for an
ellipse model of tire traction.

β The Greek lower case letter “beta” meaning yaw angle at the vehicle CG
(degrees), or the distance from the rear axle to the “Spring Center” as per
Equation 7.17 (ft).

C The University of Michigan formula (Equation 4.12) basic tire material


coefficient (in/lb), or a damping coefficient (lb-sec/in).

CART “Championship Auto Racing Teams” was founded in 1979 by team owners
who disagreed with the USAC. CART went bankrupt at the end of 2003, but
some CART team owners acquired the assets and renamed it the Champ Car
World Series (CCWS). Financial difficulties caused CCWS to file for
bankruptcy before the 2008 season.

CC The “Coupling Coefficient”, expressed as “(krlr-kflf)/ms”, is the common


factor found in both the sprung mass bounce and pitch equations; when this
coefficient is zero the bounce and pitch motions are independent of each other,
otherwise they are linked to a degree proportionate to the magnitude of the
“CC”; see Equation 7.10.

Cc The “Cornering Coefficient” is the normalized “Cornering Stiffness”: “Cs/N”


(lb/deg/lb).

Ccr This represents the critical speed for “standing wave” (misnomer) phenomenon
for a tire when modeled as a beam without tension as per Equation 4.20 (m/s).

Ccrit The Krylov/Gilbert tire “standing wave” critical velocity determined as a


combination of “Ccr” and “Cmem” as per Equation 4.22 (m/s).

Cd The aerodynamic coefficient of drag (dimensionless).

CD The Stuttgart Institute of Technology formula dynamic coefficient of tire


rolling resistance (lb/lb-mph2.5); see Equation 4.10.

C/D Car and Driver magazine is a major US reviewer and tester of new production
automobiles; it was founded as “Sports Car Illustrated” in 1955.

194
SAWE Paper No. 3602
Category Number 31.0

CDf The Stuttgart formula dynamic coefficient of tire rolling resistance for the tires
on the front axle (lb/lb-mph2.5).

CDr The Stuttgart formula dynamic coefficient of tire rolling resistance for the tires
on the rear axle (lb/lb-mph2.5).

CG, c.g. The center of gravity of a vehicle or other object.

Cmem This represents the critical speed for “standing wave” (misnomer) phenomenon
for a tire when modeled as a membrane under tension as per Equation 4.21
(m/s).

CFf The centrifugal force at the front axle which must be resisted by tire traction
forces if a turn is to be completed without “plow off” (lb).

CFr The centrifugal force at the rear axle which must be resisted by tire traction
forces if a turn is to be completed without “spin out” (lb).

CFRP Carbon fiber reinforced plastic.

Cl The aerodynamic coefficient of lift (dimensionless).

Cs The Stuttgart Institute of Technology formula static coefficient of tire rolling


resistance (lb/lb), or the suspension damping coefficient (lb-sec/in), or the
“cornering stiffness” (a.k.a. “cornering power”, which is a misnomer) (lb/deg) .

Csf The “cornering stiffness” (a.k.a. “cornering power”, which is a misnomer) at


the front axle (lb/deg), or the Stuttgart formula static coefficient of rolling
resistance (lb/lb) for the tires at the front axle, or the suspension damping
coefficient at the front axle (lb-sec/in).

Csr The “cornering stiffness” (a.k.a. “cornering power”, which is a misnomer) at


the rear axle (lb/deg), or the Stuttgart formula static coefficient of rolling
resistance (lb/lb) for the tires at the rear axle, or the suspension damping
coefficient at the rear axle (lb-sec/in).

CR The total coefficient of tire rolling resistance (lb/lb); see Equations 4.10 and
4.12.

CP The aerodynamic “center of pressure”.

CCP The “Conjugate Centers of Percussion”, a.k.a. “Double Conjugate Points”, etc.,
are a set of sprung mass oscillation node points with special properties; when

195
SAWE Paper No. 3602
Category Number 31.0

the “Dynamic Index in Pitch” is equal to one these points are located exactly at
the front and rear axles and govern the sprung mass ride motions.

CTI “Central Tire Inflation” refers to an onboard automotive tire inflation system;
such a system has long been a feature of certain military vehicles.

Ct The tire damping coefficient, which is usually around a tenth of the suspension
damping coefficient in value (lb-sec/in).

Ctf The tire damping coefficient for the tires at the front axle (lb-sec/in).

Ctr The tire damping coefficient for the tires at the rear axle (lb-sec/in).

CVT “Continuously Variable Transmission” refers to a transmission that can shift


smoothly through an infinite number of gear ratios between “high” and “low”
gear. CVT’s usually have low efficiencies and low torque capacity, but still
can provide good fuel economy by enabling the engine to run at its most
efficient rpm over a wide range of vehicle speeds.
CW Clock-wise motion.

CCW Counter-clock-wise motion.

Cα The “Cornering Stiffness” alternate symbolism, see “Cs” (lb/deg).

D The aerodynamic drag force (lb, kg), or a tire rolling diameter (in, cm).

DFV “Double Four Valve” Cosworth V8 engine.

DI The “Dynamic Index in Pitch”, an important indicator of ride motion, is equal


to “Kys2/(lf×lr)”, which in SAE J670 2008 standard symbolism is given as
“kyy2/a×b”; see Equation 7.11 (dimensionless).

DIP The “Dynamic Index in Pitch”, see “DI” (dimensionless).

DIY The “Dynamic Index in Yaw”, equal to “Kz2/(lf×lr)”, see “χ” (dimensionless).

Di The tire “no-load” inflated outer diameter (in, cm).

DNF When a race vehicle crashes, breaks down, or otherwise fails to complete a
race, then it is listed in the finishing order as “DNF” (did not finish).

DOF “Degrees of Freedom”, within the context of this paper, refers to the number of
ways a physical body can move; for a simple homogeneous mass there are 6-
DOF: three directions and three rotations.

196
SAWE Paper No. 3602
Category Number 31.0

DOHC “Dual Over-Head Cams”.

DOT The “Department of Transportation” is a Cabinet level department of the


United States government established by act of Congress on 15 October 1966.
It has the responsibility for the regulation of all matters relating to
transportation, and consists of twelve agencies including the FAA (Federal
Aviation Administration), MARAD (Maritime Administration), FRA (Federal
Railroad Administration), NHTSA (National Highway Traffic Safety
Administration), etc.

DR Wheel rim nominal diameter (in).

d A distance or diameter, or a tire’s vertical deflection under normal load as per


“d = N/Kz + d0” for a particular inflation pressure, or the maximum depth of a
depression in a road surface (in, cm) for which tire/road contact is maintainable
as per Equation 7.5.

d0 The tire deflection function “y-intercept” value (in, cm); see Equation 4.7.

d0f The tire deflection function “y-intercept” value (in, cm) for the tires at the front
axle.

d0r The tire deflection function “y-intercept” value (in, cm) for the tires at the rear
axle.

df The tire deflection (in, cm) for the tires at the front axle.

dr The tire deflection (in, cm) for the tires at the rear axle.

ds The static deflection of a spring/mass system (in, cm).

δ Greek lower case letter “delta”, usually symbolizing the steering angle at a
wheel (degrees, radians); or a material density such as that of water
(kg/mm3)/tire tread (lb/in3)/etc., or the “Steering Angle” directional stability
indicator which is calculated as per Equation 6.21 (degrees) for a turn at some
radius “R” and lateral acceleration “ay”.

%d A relative (with respect to the maximum deflection possible, usually taken as


tire section height less 17.5 mm) tire vertical deflection at a particular inflation
pressure; see Equation 4.16.

E Young’s modulus (psi, N/m2), a.k.a. the “Modulus of Elasticity” (misnomer).

197
SAWE Paper No. 3602
Category Number 31.0

ECC The “Equivalent Conventional Configuration” is a form into which the


“Advanced Concept Configuration” can be translated for input into analysis
routines created for evaluation of conventional configurations.

ECM The “Equivalent Conventional Model” is essentially the same as the


“Equivalent Conventional Configuration” (ECC).

EGR “Exhaust Gas Recycling” is a technique used to improve ICE exhaust gas NOx
emissions through recycling some of the exhaust to modify the stoichiometric
mix for combustion.

Eliica “Electric Lithium-Ion Car” was an 8-wheeled electric vehicle developed by the
Kieo University of Tokyo, Japan, circa 2004; it had a LOA of 200.8 in (5100 mm)
and a weight of about 5300 lb (2400 kg). It succeeded the less sophisticated KAZ
electric vehicle.

EPA The United States “Environmental Protection Agency” is an agency of the


federal government created by an executive order signed by President Richard
Milhouse Nixon on 2 December 1970. The EPA is entrusted to protect the
public health as it is affected by the environment through regulations based on
laws passed by the US Congress.

EV “Electric Vehicle”.

ξ The Greek lower case letter “xi” usually symbolizes the “damping ratio”,
which is the ratio of the suspension damping coefficient to the “critical
damping coefficient” as per “ 𝝃 = 𝑪𝒔 ⁄𝑪𝒄𝒓𝒊𝒕 = 𝑪𝒔 ⁄𝟐�𝒌𝒄𝒔 ⁄𝒎𝒔 ” (dimensionless).

F A force, as in “F = ma” (lb, kg, N).

F1 Signifies “Formula 1”, which is a class of international automotive racing


under the aegis of the FIA.

FEA Finite element analysis.

FEM Finite element modeling.

FHC “Fixed Head Coupe” is English terminology meaning “hardtop” (American


terminology).

FI “Fuel Injection” is a more precise means of supplying fuel to the ICE


combustion chamber than traditional carburetion.

198
SAWE Paper No. 3602
Category Number 31.0

FIA The “Fédération Internationale de l’Automobile” is a governing body for


international automotive racing established on 20 June 1904 and headquartered
in Paris, France.

FOCA The “Formula One Constructor’s Association” founded by Bernie Eccleston.

FPF Part of a chain of Coventry Climax engine development: FW, FWA, FWB,
FWF, etc.

FPS The English System of Units in feet, pounds, and seconds.

FR The total vehicle tire rolling resistance force as per Equation 4.11 (lb).

FRP “Fiberglass Reinforced Plastic”.

FR A tire rolling resistance force (lb, kg).

FRf A tire rolling resistance force at the front axle (lb, kg).

FRr A tire rolling resistance force at the rear axle (lb, kg).

FVA “Four Valve (Type) A” was a Cosworth straight-four engine.

FW In 1950 the English firm Coventry Climax unveiled a high specific power
(hp/lb) engine produced in response to a government contract for a light
weight portable fire pump system; the engine was designated the “Feather
Weight”. The engine caught the attention of Colin Chapman and others in the
automotive industry; this resulted in a line of famous high specific power
automotive engines. The FW was a straight 4-cylinder SOHC of 1020 cc
displacement and 38 hp (28 kW) with a bare weight of 180 lb (81.6 kg).

FWA Around 1953, this was the first in a line of automotive engines derived from
the original Coventry Climax portable fire pump engine; it was designated the
“Feather Weight Automobile”. It was a straight 4-cylinder SOHC of 1098 cc
displacement and 71 hp (53 kW).

FWB This was the successor to the Coventry Climax “FWA” engine. It was a
straight 4-cylinder SOHC of 1460 cc displacement and 108 hp (81 kW).

FWC This was a limited production (3 units) variant of the Coventry Climax
“Feather Weight” series of straight 4-cylinder SOHC engines intended for
use by Dan Gurney in the US and Lotus at Le Mans in 1957-1958.

199
SAWE Paper No. 3602
Category Number 31.0

FWD “Front wheel drive”, which is a common drive configuration wherein only the
two front wheels supply propulsion.

FWE This descendent of the Coventry Climax FW straight 4-cylinder SOHC engine
was designed for use in the Lotus Elite sports car as a 1216 cc varient, but
eventually saw use by a number of sports racing firms in stages of tune from
75 hp (55 kW) to 105 hp (77 kW).

FWMV The V8 version of the Coventry Climax “Feather Weight” line of engine
development: “FW”, “FWA”, “FWB”,”FWC”,”FWE", “FWF”, “FWMV”.
The initial 1.5L FWMV in 1961 made 174 hp (130 kW), but successive
generations resulted in 213 hp (159 kW) by 1964.

FWMW The 3L FWMW flat-16 engine was to supersede the FWMV in 1965, but that
proved to be a failure in development, ultimately leading to Coventry Climax
withdrawal from Formula 1 engine manufacture.

4WD “Four wheel drive”, which is a drive configuration generally with an option to
shift to RWD.

fs The natural frequency of vibration for a suspension system as calculated per


Equation 7.1a (cps, Hz).

Fx A horizontal, longitudinally directed force; usually an acceleration or a braking


force (lb).

Fy A horizontal, laterally directed force; usually a turning or a directional


disturbance force (lb).

Fz A vertical, upwardly directed force; usually the counter-force to the normal


load “N” at a tire/road contact patch which, times the offset arm “d” to the tire
vertical centerline, constitutes the tire’s rolling resistance (lb).

Ff The lateral traction potential force, or some other force, at the front axle (lb).

Fr The lateral traction potential force, or some other force, at the rear axle (lb).

Fs A force directly acting on the vehicle sprung mass, usually internal (e.g.,
engine torque reaction), but may be external (e.g., aerodynamic buffeting, gust,
etc.).

F(t) A force varying with time (lb, kg).

200
SAWE Paper No. 3602
Category Number 31.0

f1 The frequency of oscillation about principal node point number “1” on the
sprung mass per Equation 7.12 (Hz, cps).

f2 The frequency of oscillation about principal node point number “2” on the
sprung mass per Equation 7.13 (Hz, cps).

fh The frequency of oscillation about conjugate node point number “H” on the
sprung mass per Equation 7.20 (Hz, cps).

fj The frequency of oscillation about conjugate node point number “J” on the
sprung mass per Equation 7.21 (Hz, cps).

ff The braking force at the wheels of the front axle per Equation 5.11 (lb).

fr The braking force at the wheels of the rear axle per Equation 5.12 (lb).

GB Stands for “Great Britain”, which constitutes England, Scotland, and Wales.
When Northern Ireland is included then the nation is referred to as the “United
Kingdom”.

GM General Motors Corporation, founded 16 December 1908, was once the largest
manufacturing concern in the world, but went bankrupt and was saved by a US
Government “bail out” in 2009. GM is still among the world’s largest
automotive manufacturers, employing about a quarter of a million people and
doing business in 157 nations.

GPS “Global Positioning System” is a space-based satellite navigation system that


provides location and time information in all weather conditions.

GRC The “geometric roll center” or instantaneous center of rotation of a sprung


mass as determined by a static kinematic study of the suspension, usually taken
as being equivalent to the “kinematic roll center” (KRC), the “force roll
center” (FRC), and the “moment roll center” (MRC), although there are
distinctions between these roll center types.

GRP Glass reinforced plastic was one of the earliest of modern composite materials.

GVWR “Gross Vehicle Weight Rating” is the maximum operating weight condition
of a vehicle as specified by the manufacturer.

Gzr(f) The road vertical acceleration PSD input at a velocity (i.e., frequency).

Gzs(f) The sprung mass vertical acceleration PSD response at a frequency.

201
SAWE Paper No. 3602
Category Number 31.0

g The gravitational constant, needed when a mass expressed in force units must
be converted to actual mass units: “m = W/g” (32.174 ft/s2, 386.088 in/sec2,
9.80665 m/sec2).

HEI “High Energy Ignition” supplies about 40,000 volts, compared to about 18,000
volts for the traditional condenser/points/coil/ballast resister system.

HEV “Hybrid Electric Vehicle”, meaning an electric vehicle that uses an onboard
ICE/generator for recharging in operation.

Hz Hertz, or cycles per second, is a measure of oscillation frequency named in


honor of Heinrich Rudolf Hertz (1857-1894), who was the first to conclusively
prove the existence of electromagnetic waves.

Hzs Vertical vibration frequency of the sprung mass in Hz.

Hzs(f) The sprung mass vertical gain or transmissibility factor at a frequency.

h A height, or a tire section height, or the vertical distance from the sprung mass
CG to the roll axis (in, cm). Also can stand for a tire belt thickness as per in
Equations 4.19-4.22 (m).

h0 The height of the water surface above road roughness peaks as in Equation
4.14b (mm).

hcg The height of the center of gravity above the ground plane (in, cm).

hr The length of a line that is orthogonal to the roll axis and runs from there
onward to the sprung mass CG (in, cm).

ht The total “sink” distance before tire makes firm contact with road as in
Equation 4.14b (mm).

h/w The tire section aspect ratio, or any height to width ratio (dimensionless).

I A mass moment of inertia about some axis (lb-in2, lb-ft2, kg-m2, etc.).

I/B Inboard, or close to the vehicle longitudinal centerline.

Icg A mass moment of inertia through the CG; usually the pitch or yaw inertia of
a vehicle (lb-in-sec2, lb-ft2, kg-m2).

Ioc A mass moment of inertia through the Oscillation Center in yaw (lb-ft2, kg-
m2).

202
SAWE Paper No. 3602
Category Number 31.0

ICE “Internal combustion engine”, of which there are many types (piston, turbine,
etc.), with subtypes operating on any of a number of combustible fuels
(gasoline, diesel, alcohol, etc.).

IPS The English System of Units in inches, pounds, and seconds.

I1 The rotational inertia of certain components about the front axle line as in
Equation 2.3 (lb-ft2, kg-m2)

I2 The rotational inertia of certain components about the crankshaft axis as in


Equation 2.3 (lb-ft2, kg-m2)

I3 The rotational inertia of certain components about the transmission 3rd motion
axis as in Equation 2.3 (lb-ft2, kg-m2)

I4 The rotational inertia of certain components about the rear axle line as in
Equation 2.3 (lb-ft2, kg-m2).

I4 The unit rotational inertia of “full-sized” tires/wheels (lb-ft2, kg-m2), see


Chapter 4.

I8 The unit rotational inertia of “half-radius” tires/wheels (lb-ft2, kg-m2), see


Chapter 4.

IFS Independent front suspension.

IRS Independent rear suspension.

Ix, Ixx The mass moment of inertia in roll (lb-in-sec2, lb-ft2, kg-m2).

Iy, Iyy The mass moment of inertia in pitch (lb-in-sec2, lb-ft2, kg-m2).

Iz, Izz The mass moment of inertia in yaw (lb-in-sec2, lb-ft2, kg-m2).

J A mass moment of inertia (lb-in-sec2), usually the same as “Iy”; or the tire
belt second moment of area “J = h3b/12” (m4); see Equations 4.19-4.22.

K The tire load capacity service factor: “Ktruck&bus = 1.00, Kpassenger car = 1.10” as
used in Equation 4.15, or a tread compound/pavement wear rate value as used
in Equation 4.23 (typical passenger car tire value around 1.947904×10-6 in/lb-
1000 mi).

KAZ “Kieo Advanced Zero-emissions” was an 8-wheeled electric vehicle developed


by the Kieo University of Tokyo, Japan, circa 2002; it had a LOA of 263.78 in (6700

203
SAWE Paper No. 3602
Category Number 31.0

mm) and a weight of about 6,570 lb (2980 kg). It preceded the vastly more
sophisticated Eliica vehicle.

KE The “kinetic energy” as possessed by a mass in motion per “KE = ½ m V2”


(ft-lb, N-m).

KUS The “Understeer Coefficient”, or “Understeer Gradient”, as per Equation 6.17


is an indicator of the level of vehicle directional stability (degrees, radians).

Kv Tire vertical stiffness as used in Equation G.21 (lb/in).

Kx The roll radius of inertia (ft, in, m, cm).

Ky The pitch radius of inertia (ft, in, m, cm).

Kz The yaw radius of inertia as used in Equation 6.26 (in), or a tire vertical
stiffness as used in Equation 4.18 (lb/in).

Kzf, ksf, kf The suspension spring rate at a wheel on the front axle of a vehicle (lb/in).

Kzr, ksr, kr The suspension spring rate at a wheel on the rear axle of a vehicle (lb/in).

k Tire foundation stiffness as used in Equation 4.19 (N/m2).

kcs The combined effective vertical stiffness at a wheel of the suspension spring
and the tire vertical spring rate (lb/in).

ks The effective vertical stiffness of a suspension spring, either at the wheel or at


the spring-mount on a beam axle (lb/in).

ksf The vertical stiffness of suspension spring at a front wheel (lb/in).

ksr The vertical stiffness of suspension spring at a rear wheel (lb/in).

ks1 A spring stiffness (lb/in).

ks2 Another spring stiffness (lb/in).

kt The vertical stiffness of a tire (lb/in).

ktf The vertical stiffness of a tire at the front axle (lb/in).

ktr The vertical stiffness of a tire at the rear axle (lb/in).

kfroll The roll stiffness at the front suspension, usually calculated (IFS) per Equation
6.3 (lb-ft/degree).

204
SAWE Paper No. 3602
Category Number 31.0

krroll The roll stiffness at the rear suspension, calculable for a beam axle as per
Equation 6.4, otherwise per Equation 6.3 (lb-ft/degree).

kroll The vehicle/suspension roll stiffness as calculated per Equation 6.3 and/or per
Equation 6.4 and combined (lb-ft/degree) .

L The aerodynamic lift force (lb, kg).

Lcap The tire load capacity at a particular inflation pressure “Pi” as per Equation
4.15 (lb).

LWB, lwb The “wheelbase”, or longitudinal distance between the front and the rear axles
(in, ft, cm, m).

Lc The longitudinal dimension of the tire-to-ground contact area (in, cm).

Leff Per Reference [38] this signifies the LOA less the protruding bumpers (classic
configuration). For modern vehicles with bumpers fully integrated into the
body proper the “Leff” is probably equivalent to the “LOA”.

Ls The “squeegee length” between tires as per Equation 4.13 (in, mm).

LCG The vehicle longitudinal center of gravity (in, cm).

LCP The vehicle longitudinal center of pressure (in, cm).

Lc,fwd The forward tire/road contact patch length (in, cm).

Lc,rr The rearward tire/road contact patch length (in, cm).

LOA Length over-all, which for modern cars may be considered equivalent to “Leff”.

L/H Left hand, as opposed to right hand orientation.

LSR Refers to the “Land Speed Record” and the type of vehicles used to set such
records.

l A length, especially the minimum length of a dip in a road surface which is


traversable by a suspension without tire/road loss of contact per Equation 7.4
(ft, in, cm).

lf The longitudinal distance from the front axle line to the vehicle CG (in, cm).

lr The longitudinal distance from the rear axle line to the vehicle CG (in, cm).

205
SAWE Paper No. 3602
Category Number 31.0

λ The Greek lower case letter “lamda”, generally symbolizing the angle between
the longitudinal principal axis and the longitudinal design axis as illustrated in
Table 3.1 (degrees). It can also stand for a wavelength, such as the wavelength
of “tire standing wave” phenomenon as per Equation 4.19 (m).

M Tire “mileage” or “life” expectancy in units of 1000 as per Equation 4.24


(1000 mile).

MAC The “Mean Aerodynamic Chord” is the average cross-section nose-to-tail


(chord) distance of an airfoil (ft, in, m, cm).

MAE A Cosworth 997 cc, 100-110 hp engine design for 1965 Formula 3
competition.

MG The “MG Car Company Ltd” was a British automobile manufacturer


founded in the 1920’s by Cecil Kimber acting for his employer William
Morris; MG stands for Morris Garages. In 1935 Morris “sold” MG into his
holding company, Morris Motors Ltd, before issuing shares in Morris Motors
to the public in 1936. The company underwent many subsequent changes in
ownership: becoming the Nuffield Organization, then merging into The British
Motor Corporation Ltd in 1952, then becoming the MG Division of BMC in
1967, and then becoming part of the 1968 merger into the British Leyland
Motor Corp. By 2000 MG was part of the MG Rover Group which entered
receivership in 2005; the assets were purchased by the Chinese firm Nanjing
Automobile Group with production since 2007 located in China, and some
limited production in the UK as MG Motor.
MGS “Mobile Gun System” is a reference to a type of self-propelled artillery, such
as the Stryker vehicle by General Dynamics.
MIRA The “Motor Industry Research Association” (Ltd) was founded 1949 in the UK
and is now a major automotive test facility with aerospace, rail, and other
branches.
MNC “Modified Nicolas-Comstock” is yet another unified model of longitudinal and
lateral tire behavior. The modifiers of the original “Nicolas-Comstock” model,
Raymond and Matthew Brach, claim the “MNC” model used in conjunction
with the de facto standard “BNP” model accurately represents tire behavior for
moderate %Slip and Slip Angle, but becomes less reliable at the limits due to
the usual problems associated with linearized models of non-linear behavior.

206
SAWE Paper No. 3602
Category Number 31.0

The “BNP-MNC” model tends to underestimate the traction forces at the limit
of tire performance.
MOI MOI (Moment of Inertia) is the summation of all the individual mass
elements of a body times the individual element radial distance (squared) to
some axis of revolution:

𝑰 = � 𝒓𝟐 𝒅𝒎

MR Generally used to refer to the unsprung-to-sprung mass ratio (“mus/ms”),


which is a key parameter in suspension design (dimensionless). Also can mean
“motion ratio” or “moment ratio”, which are key factors in effective spring
constant determination (dimensionless).

M/T Motor Trend magazine is a major automotive publication (US) founded in


1949.

m Mass, as in “F = ma” (slugs, kg), or the rate of lateral tire traction potential
decrease with load “μ = b-mN” (lb-1, when “N” is in lb units).

me “Effective mass” is the mass that a vehicle appears to have depending upon its
behavior in acceleration or deceleration; it is something other than its simple
“scale” mass (m=W/g) and may be calculated as per Equations 2.3
(acceleration) and 2.5 (deceleration); it is often expressed in weight units.

ms The “sprung mass” is that portion of the vehicle that may considered to be
protected from the road irregularities by the suspension.

mus The “unsprung mass” is that portion of the vehicle that may be considered
directly affected by the road irregularities.

mus/ms The “unsprung-to-sprung mass ratio” is a key factor in many automotive


performance determinations

mpg “Miles per gallon” is one of most commonly recognized measures of fuel
economy, but it is important to recognize that the UK and US gallons are
different (1 “Imperial” gallonUK = 1.200949 gallonsUS).

N The normal load, usually the load borne by a tire (lb).

NASA The United States “National Aeronautics and Space Administration” is the
agency entrusted by the US Government with the responsibility for the nation’s
civilian space program and general aeronautics/aerospace research. NASA was

207
SAWE Paper No. 3602
Category Number 31.0

established by the National Aeronautics and Space Act on 29 July 1958,


replacing its predecessor the National Advisory Committee for Aeronautics
(NACA).

NASCAR The “National Association for Stock Car Auto Racing” was founded by Bill
France Sr. in 1947. It sanctions and governs various auto racing events, and is
the largest sanctioning body for “stock car” racing in the United States.

NEV “Neighborhood Electric Vehicle” is an electric vehicle of low range and speed
suitable only for use on secondary routes close to home.

Nf The normal load at a front tire (lb, kg)

Nif The normal load at an “inside” front tire in a turn as per Equation 6.7 (lb, kg).

Nof The normal load at an “outside” front tire in a turn as per Equation 6.8 (lb, kg).

Nr The normal load at a rear tire (lb, kg).

Nir The normal load at an “inside” rear tire in a turn as per Equation 6.9 (lb, kg).

Nor The normal load at an “outside” rear tire in a turn as per Equation 6.10 (lb, kg).

NHTSA The National Highway Traffic Safety Agency is a major part of the US DOT.
It describes its mission as “Save lives, prevent injuries, reduce vehicle
related crashes”. The NHTSA is responsible for creating and enforcing
Federal Motor Vehicle Safety Standards (FMVSS) and fuel economy
regulations.

NOx Nitrogen oxides are one of the components of ICE exhaust considered to be
environmental pollutants, along with carbon monoxide and unburned
hydrocarbons.

NS Neutral steering.

NSP “Neutral Steer Point” is the automotive analog to the aeronautical “Neutral
Stability Point”.

NTE “Not to Exceed” is a contractual obligation, usually referring to a limiting


weight or cost.

NVH Noise, vibration, and harshness.

n A number, as in the number of axles on a vehicle, or the exponent in the tire


wear rate Equation 4.23 which varies directly with usage (acceleration/braking

208
SAWE Paper No. 3602
Category Number 31.0

and/or cornering) and inversely with tire radius (possibly 1.90 to 2.30,
dimensionless). May also represent the exponent in the tire traction Equations
5.13 and 5.14 (around -0.67791 in value, dimensionless).

O/B Outboard, or far from the vehicle longitudinal centerline.

OC A point of initial rotation in a turn known as the “Oscillation Center”, which


migrates toward the LCG as the transient condition progresses toward steady-
state.

OEM “Original Equipment Manufacturer”.

OHC “Overhead Cam”.

OS “Oversteering”.

PC The tire-to-road contact pressure (psi, kN/mm2).

PCf The tire-to-road contact pressure at the tires on the front axle (psi, kN/mm2).

PCr The tire-to-road contact pressure at the tires on the rear axle (psi, kN/mm2).

PEV “Plug-in Electric Vehicle”, meaning an electric vehicle that must be recharged
via the electrical power grid.

PHEV “Plug-in Hybrid Electric Vehicle”, meaning an electric vehicle that uses an
ICE/generator for recharging in operation, but which can also be recharged via
the electrical power grid.

Pi Tire inflation pressure (psi, kPa).

POI POI (Product of Inertia) is a measure of the degree of asymmetrical


distribution of mass in a particular plane. The mathematical POI definition for
each of the three planes (XY, XZ, and YZ) is:

𝑷𝒙𝒚 = ∫ 𝒙𝒚 𝒅𝒎

PSD “Power Spectrum Density” is how the strength (i.e., power) of a transmission
is distributed over the frequency domain.

P&W Pratt & Whitney is an American aerospace manufacturer with global


operations. Its first engine, the 425 hp (317 kW) R-1340 Wasp radial, was
completed on Christmas Eve 1925; on its third test run it easily passed the
Navy qualification test. Today Pratt & Whitney gas turbine and jet engines are
widely used in both civil and military aviation.

209
SAWE Paper No. 3602
Category Number 31.0

ϕ Greek lower case letter “phi” generally used to symbolize the roll axis angle
with the horizontal (degrees, radians).

ψ Greek lower case letter “psi” generally used to symbolize the “slip angle”
(misnomer) between a tire’s direction of movement with the tire’s longitudinal
axis (degrees, radians).

ψf The average “slip angle” (misnomer) at the front axle tires (degrees, radians).

ψr The average “slip angle” (misnomer) at the rear axle (degrees, radians).

Q.E.D. “Quod Erat Demonstrandum” (“Thus It Is Proved”) is the usual Latin phrase
found at the end of a proof of a mathematical hypothesis.

Q/C “Quality control” is a product oriented function that focuses on defect


identification. Today it is usually paired with “quality assurance” (“QA/QC”)
which is a function that focuses on defect prevention.

R A radius, such as a tire rolling radius (in, cm), or the radius of a turn (ft, m)
such as a skidpad radius (“R” is measured from the TC to the vehicle CG).

RAF The Royal Air Force. Colin Chapman spent some time in the RAF as a result
of having been in a UK version of the US ROTC while attending college.

RC “Roll center”, which is the point about which the sprung mass would tend to
rotate in roll due to the kinematics of the suspension.

RD The dynamic rolling radius of a tire, usually taken at the drive wheels (ft, in, m,
cm).

R&D “Research & Development”.

RG The “roll gain” is a measure of a vehicle’s resistance to roll under lateral


acceleration as determined by iterative solution of Equation 6.5 (deg/g).

R/H Right hand, as opposed to left hand, orientation.

R&T Road & Track magazine is an automobile enthusiast’s magazine founded at


Hempstead, NY, USA in 1947; it was perhaps the most technically orientated
of all major automotive publications.

ROTC The “Reserve Officers Training Corps” is a US military training program for
college students.

210
SAWE Paper No. 3602
Category Number 31.0

Ri The inflated, no load, tire radius (in, cm).

Rif The inflated, no load, tire radius at the front axle (in, cm).

Rir The inflated, no load, tire radius at the rear axle (in, cm).

Rr The rolling (inflated, under load) tire radius (in, cm).

Rw The rate of tire tread wear, possibly around 0.002 to 0.008 (in/1000 mi).

r A small radius, such as a tire rolling radius (in, cm). Also the distance from the
sprung mass CG to the front conjugate node point of percussion “H” as per
Equation 7.18 (ft).

rrf The tire rolling radius at the front axle, which is often taken as an
approximation of the front axle unsprung mass vertical c.g. height (ft).

rrr The tire rolling radius at the rear axle, which is often taken as an
approximation of the rear axle unsprung mass vertical c.g. height (ft).

rpm “Revolutions per minute”, which might more correctly be termed “rotations
per minute”, generally refers to the angular speed of an engine or motor.

ρ Greek lower case letter “rho” for tire mass density as used for Equations 4.19-
4.22 (kg/m3).

S Stands for the tire section width as “adjusted” for an equivalent circular
periphery (in) for use in Equation 4.15, or for tire belt tension force “S =
PiRib” (N) in Equation 4.21, or for the tire-to-road contact area stress (lb/in2)
in Equation 4.23.

SAE The “Society of Automotive Engineers” (now “SAE International”) was


initially founded in 1904 as the “Society of Automobile Engineers”, mainly as
the result of the advocacy efforts of two journalists: Peter Heldt of The
Horseless Age, and Horace Swetland of The Automobile. By 1916 the focus
was broadened beyond the automobile to include all forms of transportation.

SAWE The “Society of Allied Weight Engineers” is the premier professional society
for engineers involved in mass properties analysis and control. The Society
of Aeronautical Weight Engineers was organized in 1939 in Los Angeles,
California, and was incorporated as a nonprofit organization April 2, 1941. As
membership grew to include engineers associated with shipbuilding, land
transportation, and other allied industries and technologies, the Society name

211
SAWE Paper No. 3602
Category Number 31.0

was changed on January 1, 1973 to the Society of Allied Weight Engineers,


Inc.

SC The “Spring Center” is a special point in the automobile 2-DOF bounce/pitch


model where the two motions are uncoupled statically but still coupled
dynamically.

SCCA The Sports Car Club of America is a descendent of the Automobile Racing
Club of America founded in 1933 by brothers Miles and Sam Collier, which
was dissolved in 1941 at the entry of the US into WW II. In 1944 it was
reconstituted as the SCCA, and began sanctioning road racing events in 1948.

SHM “Simple Harmonic Motion” is a type of periodic motion where the restoring
force is directly proportional to the displacement. It can serve as
a mathematical model for a variety of motions, such as the oscillation of a
mass on a spring.

SM “Static Margin” is an indicator for classifying a vehicle’s directional stability


as stable (“understeering”), neutral, or unstable (“oversteering”); it is a
dimensionless fraction (see Equation 6.22) which can be used to locate the
NSP with respect to the LCG (“NSP = LCG – (SM×Lwb)”). Typical values
for passenger cars may range between +0.05 to +0.07 (understeering).

SH Tire nominal cross-section height (in, mm).

SN Tire nominal cross-section width (in, mm).

SUV “Traditional” American automobiles were long objects of criticism as being


too large, too heavy, too clumsy, and too inefficient for the purpose of serving
as passenger cars. Today many people in the US use “Sport Utility Vehicles”
for the passenger car function, which tend to be larger, heavier, clumsier, and
just as inefficient as their traditional predecessors.

SSF The “Static Stability Factor” is a rough indication of a vehicle’s overturn


threshold (numerically equal to “ayoverturn”) as it is calculated without reference
to roll or other dynamic effects; see Equation 6.1 (dimensionless).

STP The STP (“Scientifically Treated Petroleum”) Company was founded in 1953,
but was acquired by Studebaker-Packard Corporation in 1961. Soon afterward
the flamboyant Andy Granatelli was named CEO of the automotive products
company, with the intent he should popularize the brand through the

212
SAWE Paper No. 3602
Category Number 31.0

sponsorship of automotive competition. It was a job that Granatelli did so well


that the subsidiary has long survived the parent company.

%S The apparent “slip” (misnomer) of a tire corresponds to the level of


longitudinal traction; for braking it is defined as per Equation 5.23
(dimensionless).

s The distance from the sprung mass CG to the rear conjugate point of
percussion “J” as per Equation 7.19 (ft).

T Torque, generally that produced by an engine or motor, but also could


represent the gyroscopic reaction of a spinning mass to a forced precession as
per Equation 8.1 (lb-ft).

TC The turn center.

TE The transmission efficiency for a particular gear ratio (dimensionless).

TPMS Stands for “Tire Pressure Monitoring System”, which by DOT regulation is
now required on all new cars sold in the US.

TR The transmission (gear) ratio (dimensionless).

TRA Tire and Rim Association Inc. was founded in 1903 to set standards for the
rating of tires, of which the standard for load capacity was foremost. Many of
the industry standards have since been incorporated into federal regulation by
the US DOT.

TRB “Transportation Research Board” is one of six major divisions within the
United States National Research Council, which serves as an independent
advisor to the President, Congress, and federal agencies on matters of scientific
and technical importance. The National Research Council is administered by
the National Academy of the Sciences, the National Academy of Engineering,
and the Institute of Medicine.

TV Television.

t A thickness, such the thickness of a tire tread, or the width of a tire tread (in,
cm). Also, a vehicle track width or lateral distance from wheel centerline to
wheel centerline along the axle (in, cm).

ts The “sink time” required for tire penetration through standing water to the road
surface as used in Equation 4.14b (sec).

213
SAWE Paper No. 3602
Category Number 31.0

tsb The “spring base” distance for a beam axle is the distance along the axle
between the spring-mounts as used in Equation 6.4, but for an independent
suspension it is the track as used in Equation 6.3 (in).

tt The tire tread thickness as in Equation 4.18 (in).

tw The tire tread width, assumed to stay constant with load for passenger car tires,
as used in Equation 4.18 (in).

twf The tire tread width of the tires on the front axle (in, cm).

twr The tire tread width of the tires on the rear axle (in, cm).

tθ A lag time that sets the phase relationship “θ” between the conflicting sprung
mass pitch oscillations resulting from the front and rear suspensions impacting
the same road irregularity, “tθ = V/lwb” (seconds).

τ Greek lower case letter “tau”, usually representing the period of oscillation
(sec/cycle) of a spring/mass system, which is the inverse of the frequency, as
per Equation 7.1 (seconds).

θs The sprung mass roll angle due to lateral acceleration “ay” as determined by
iterative solution of Equation 6.5 (degrees).

UCL The “University College of London” is part of the venerable University of


London college system. It was at UCL that Colin Chapman received his
education as a structural engineer.

UK The “United Kingdom”, meaning England, Wales, Scotland, and Northern


Ireland, plus three British Crown Dependencies and fourteen British Overseas
Territories.

US The “understeering” condition of automotive handling, or the “United States”.

USA The “United States of America”.

USAC The “United States Automobile Club” was formed by famed Indianapolis
Motor Speedway owner Tony Hulman in 1956. It serves as the sanctioning
body for a number of “open wheel” racing series in the US, including Indy,
Sprint, and Midget cars.

μ The tire belt mass per unit length “μ = ρhb” as used in Equations 4.20-
4.21(kg/m).

214
SAWE Paper No. 3602
Category Number 31.0

μf The tire traction coefficient for the tires on the front axle of a vehicle (lb/lb).

μr The tire traction coefficient for the tires on the rear axle of a vehicle (lb/lb).

V Vehicle velocity, often called “speed” (mph, ft/sec, kph, m/sec).

VARI “Vacuum Assisted Resin Infusion” is a process used to make GRP structures.
In the automotive world VARI offers many advantages over traditional hand
lay-up, but the most significant gain is that up to six bodies a day, as opposed
to just one, can be produced from a set of molds.

Vc The tire critical velocity for “standing wave” onset as determined by Equation
4.18 (mph).

Vchar The “characteristic speed” is an index for comparing the degree of understeer
present in vehicles; see Equation 6.19 (in/sec, ft/sec). Traditional highly
oversteering US cars had characteristic speeds as low as 40 mph (64 kph).

Vcrit The “critical speed” is a key speed for cars of initially controllable oversteer; it
is the point of transition to uncontrollable oversteer as per Equation 6.20
(in/sec, ft/sec).

Vh The tire critical velocity for hydroplaning as determined by Equation 4.14a


(mm/sec) and/or Equation 4.14c (mph).

VCG The vehicle vertical center of gravity (in, cm).

VCP The vehicle vertical center of pressure (in, cm).

VVT “Variable Valve Timing” alters the timing of ICE valve open/close cycle
during operation to improve performance/fuel economy/emissions. This is in
contrast to the traditional mechanically fixed valve timing.

V8 Refers to a reciprocating piston type ICE of eight cylinders arranged in a “V”


formation.

V16 Refers to a reciprocating piston type ICE of sixteen cylinders arranged in a


“V” formation. Historically there have been far fewer examples of V16 engine
construction than V8. Some “V16” engines in chronological order: Bugatti
U16 (1915, actually two straight-8’s side-by-side), Marmon V16 (1931-1933),
Cadillac Series 452 V16 (1930-1937, actually two straight-8’s in a “V”), Auto
Union Type C GP V16 (1933-1938), Cadillac Series 90 V16 (1938-1940),
Alfa Romeo Tipo 162/316 V16 (1938), BRM V16 (1954-1955, actually two
V8’s in tandem), BRM H16 (1966, actually two stacked flat-8’s), Cizeta-

215
SAWE Paper No. 3602
Category Number 31.0

Moroder V16T (1991-1995, actually two V8’s in a transverse mounting), and


Bugatti Veyron W16 (2000- ).

W A weight, often the vehicle weight or a portion of that weight as borne at an


axle (lb, kg).

Wb The vehicle wheelbase, same as “lwb” (in, cm).

W4 The weight of the “full-sized” tires/wheels (lb, kg).

W8 The weight of the “half-radius” tires/wheels (lb, kg).

w A width, perhaps a tire section width (in, cm).

WER A “Weight Estimating Routine”, which is a mathematical formula that can be


used to estimate the mass properties, most often the weight, of an object.

WEVA The World Electric Vehicle Association (WEVA) is an international


organization launched in 1990 with the objective of promoting the research,
development and dissemination of electric vehicles.

Wf The weight load at the vehicle front axle (lb).

Wr The weight load at the vehicle rear axle (lb).

Ws The weight of the vehicle sprung mass (lb).

Wt A total vehicle weight, usually a “curb weight plus driver” or a “test weight”
(lb, kg).

Wusf The front axle unsprung mass weight (lb).

Wusr The rear axle unsprung mass weight (lb).

WOW “Weight On Wheels” is an aircraft landing gear load sensing system usually
based on “shock absorber” stroke movement using inductive sensors from
which the load is inferred. Some recent variations on this traditional system
measure load more directly and precisely using strain gauge technology.

WW II The Second World War.

ω The Greek lower case letter “omega” usually signifies an angular velocity or a
frequency (radians/second).

ωr The angular velocity of a spinning mass (radians/second).

216
SAWE Paper No. 3602
Category Number 31.0

ωp The angular velocity of precession of the rotational axis of a spinning mass


(radians/second).
X, x A coordinate, usually along the longitudinal axis, of some point per some
reference system.

X1 The distance from the sprung mass CG to the first principal node point “1”
per Equation 7.14 (ft).

X2 The distance from the sprung mass CG to the second principal node point “2”
per Equation 7.15 (ft).

χ Greek lower case letter “chi”, usually symbolizing the value “𝐊 𝐳 𝟐 ⁄(𝐥𝐟 × 𝐥𝐫 )”
(dimensionless). It may sometimes be used to refer to the unsprung-to-sprung
mass ratio “mus/ms” (also dimensionless).

Y, y A coordinate, usually along the lateral axis, of some point per some reference
system.

Z, z A coordinate, usually along the vertical axis, of some point per some reference
system.

ZF “ZF Friedrichshafen AG” is a company famous for the production of


transmissions, differentials, and other automotive components. It was founded
in 1915 in Friedrichshafen, Germany by Ferdinand von Zeppelin, to
produce gears for Zeppelins and other airships. “ZF” is a German abbreviation
for “Zahnradfabrik” (Gear Factory).

𝐙𝐬̈ The vertical acceleration of the vehicle sprung mass (ft/sec2, m/sec2).

𝐙𝐫̈ The vertical “acceleration” of the road surface at speed (ft/sec2, m/sec2).

217
SAWE Paper No. 3602
Category Number 31.0
Rev. C

APPENDIX B – WEIGHT ACCOUNTING

For this paper the very first necessary mass properties task was to establish the
“baseline” mass properties. The Autocar of 18 September 1959 gave the 1958 Jaguar
XK150S FHC "kerb weight" (English spelling) as 3248 lb, and a “test weight” of 3640 lb,
all of which seems reasonable. However, as stated elsewhere 337, one must exercise caution
when using curb weight figures that may not conform to the SAE standard. As it turns out,
the Autocar “kerb weight” was for a road ready vehicle, but with only about 5.0 “Imperial”
(UK) gallons of gasoline onboard. The XK150S fuel tank capacity was 14.0 UK gallons,
which is equivalent to 16.8 US gallons. So the addition of 9.0 UK gallons, or 10.8 US gallons,
of fuel would be required to achieve a SAE standard curb weight condition. Since gasoline
has a density of 6.0 lb/gal 338, this means about 65 lb must be “added” to the Autocar “kerb
weight” to obtain a SAE standard curb weight figure. Then, from the curb weight condition,
the test weight condition was derived as shown using standard weight accounting
methodology339:

Figure B.1 – “BASELINE” CURB & TEST WEIGHT MASS PROPERTIES

337
Reference [88], pg. 8.
338
Reference [98], pg. 5.16. The c.g. location and other mass properties associated with this fuel increment
were determined via rough calculations using a 1:18 scale model of the vehicle as a longitudinal location guide.
339
Reference [87], pg. 3.

218
SAWE Paper No. 3602
Category Number 31.0
Rev. C

The Autocar road test weight included two occupants, a driver and a “test recorder”.
The latter’s function was to use a clipboard equipped with an array of stopwatches for the
recording of the various elapsed time-to-speed readings, a function now relegated to
electronic devices. The mass properties of the occupants were determined in accord with
SAWE practice, but using the average of two different seating positions to approximate
automotive seating 340:

Table B.1 – CREW MASS PROPERTIES DETERMINATION

Lastly, all vehicle mass properties of Figure B.1 not given in The Autocar 341 road test
summary were determined in accord with the “weight” estimation routines (WER’s) of
Reference [61], with the exception of “Pxz” which was estimated in accord with Reference
[65]. In particular, the “kerb weight” “Z” was estimated per Eq. 5.13, “Ix” per Eq. 7.37, “Iy”
per Eq. 7.39, “Iz” per Eq. 7.8b, “Pxy” by virtue of symmetry, “Pxz” per Eq. 5, and “Pyz” also
by virtue of symmetry.

The next mass properties task was to determine the “baseline” suspension unsprung
mass properties using established methodology342:

Table B.2 – “BASELINE” UNSPRUNG MASS PROPERTIES DETERMINATION

340
Reference [98], pp. 20.5-20.11.
341
The longitudinal weight distribution was given; the lateral weight distribution was assumed.
342
Reference [88], pp. 29, 57.

219
SAWE Paper No. 3602
Category Number 31.0
Rev. C

Using the suspension unsprung mass properties the total suspension mass
properties were obtained but, since there is no established methodology for making such a
determination, the relationship of the total suspension mass properties to the suspension
unsprung mass properties of a 1994 Ford Taurus suspension (as detailed in Appendix C of
Reference [61]) was used as a rough guide:

Table B.3 – “BASELINE” TOTAL SUSPENSION MASS PROPERTIES

Knowledge of the “baseline” test weight and unsprung mass properties also allows
for determination of the “baseline” sprung mass properties:

Table B.4 – “BASELINE” SPRUNG WEIGHT MASS PROPERTIES

The knowledge of the “baseline” mass properties obtained so far (total vehicle as
used for acceleration Run #1a, total suspension, unsprung suspension, and sprung vehicle)
allows for the modification of the “baseline” into the “advanced concept”. This modification
was preformed in successive steps to arrive at the “advanced concept” mass properties as
appropriate for acceleration Runs #7, 8, and 9 (the mass properties for the “advanced
concept” acceleration Runs #4, 5, and 6 were determined by less “realistic” more theoretical
means) as per the following “detail” estimation:

220
SAWE Paper No. 3602
Category Number 31.0
Rev. C

Table B.5 – “ADVANCED CONCEPT” TEST WEIGHT MASS PROPERTIES

Note that the “detail” (more realistic, less theoretical than that of acceleration Runs #4,
5, 6) development of the “advance concept” mass properties begins with the “baseline”
verification acceleration Run #1a condition. The suspension, spare tire, differential, engine,
and radiator are “removed”; the resulting hulk is “lowered” 1 in (0.08 ft, 0.0254 m) and the

221
SAWE Paper No. 3602
Category Number 31.0
Rev. C

“advanced concept” suspension, differentials, and radiators are “added”. The engine is also
“reinstalled” in its new transverse location. The result is the mass properties as appropriate
for the final “advanced concept” configuration of Run #9; the mass properties as appropriate
for the configurations of acceleration number Run #7 and #8 were obtained similarly as noted
in the table.

Since it is important for certain analysis to not only have the suspension mass
properties as divided into the “sprung” and “unsprung” categories, but also into “rolling” and
“non-rolling” categories as well, the following “detail” suspension weight accounting was
carried out for the “baseline” vehicle:

Table B.6 – “BASELINE” SUSPENSION MASS PROPERTIES DETAIL

222
SAWE Paper No. 3602
Category Number 31.0
Rev. C

The corresponding suspension “detail” for the “advanced concept” is as follows:

Table B.7 – “ADVANCED CONCEPT” SUSPENSION MASS PROPERTIES DETAIL

223
SAWE Paper No. 3602
Category Number 31.0

APPENDIX C – MAXDLONG.BAS PROGRAM LISTING

The MAXDLONG.BAS program created for the braking performance analysis of


Chapter 5 was written in GW-Basic; the program listing follows, along with sample input and
output.

MAXDLONG.BAS PROGRAM LISTING

224
SAWE Paper No. 3602
Category Number 31.0

225
SAWE Paper No. 3602
Category Number 31.0

226
SAWE Paper No. 3602
Category Number 31.0

227
SAWE Paper No. 3602
Category Number 31.0

228
SAWE Paper No. 3602
Category Number 31.0

229
SAWE Paper No. 3602
Category Number 31.0

MAXDLONG.BAS INPUT (BASELINE)

230
SAWE Paper No. 3602
Category Number 31.0

MAXDLONG.BAS OUTPUT (BASELINE)

231
SAWE Paper No. 3602
Category Number 31.0

232
SAWE Paper No. 3602
Category Number 31.0

APPENDIX D – AUTOMOTIVE AXLE/WHEEL CONFIGURATIONS

The conventional automotive configuration of two axle lines and four wheels, with
each wheel located in a corner of the automotive plan view, is one of only a large number of
possible wheel/axle configurations, but has prevailed for so long that this fact is often
forgotten. The choice of which wheels steer and which wheels are driven compounds the
number of configuration variations possible. When other configuration variations are also
considered, such as varying the vehicle tire/wheel size/type fore to aft or even side to side (as
on some circle track racers), or whether the engine is to be front, mid, or rear located, then the
number of all possible configurations becomes infinite.

In an attempt to develop a convenient short-hand to designate unconventional


automotive configurations, recourse has been made to the Whyte system for the classification
of steam locomotive configurations. This system was devised by the engineer Frederick
Whyte around 1900, and was applied to automobiles by the race car designer Robin Herd
circa 1976. Whyte's system counts the number of leading wheels, then the number of driving
wheels, and finally the number of trailing wheels, the numbers being separated by dashes;
Herd termed his 6-wheel variation of the standard March 761 F1 racer the “2-4-0”:

Figure D.01 - THE MARCH “2-4-0” IN ROTHMAN’S INTERNATIONAL LIVERY 343

Although the Whyte system is not all-encompassing when it comes to automotive


configuration description, it can be useful when employed in conjunction with such other
descriptive terms (e.g.: FWD, 4WD, Mid-Engine, etc.) as needed. Utilizing the Whyte
system, the eleven most common basic automotive axle/wheel configurations might be
classified as follows:

343
Rothman’s International was a British tobacco company which did considerable sponsorship of Formula 1
racing in the 1970’s.

233
SAWE Paper No. 3602
Category Number 31.0

Figure D.02 - THE ELEVEN MOST BASIC AUTOMOTIVE WHEEL/AXLE


CONFIGURATIONS

The single wheel “0-1-0” or “unicycle” configuration might seem to be a very odd
choice for automotive design, but there have been a surprising number of vehicles based on
this configuration. Perhaps the earliest known example is the Rousseau “Monocycle” of 1869,
with more or less continuous attempts appearing from that point in time to the present day.
One of the more prominent examples of modern unicycle design would be the “Wheelsurf”
which is marketed as a recreational vehicle for use at beachfront resorts. However, the most
historically significant example of the “unicycle” configuration would have to be the Purves
“Dynosphere” of 1932. Apparently Dr. John Archibald Purves wanted to reduce the
automobile to just its essential characteristic: the wheel. Supposedly this would result in
unprecedented efficiency; what it did result in was a huge directionally unstable juggernaut
that dwarfed conventional vehicles. Despite much work and at least two prototypes, the
“Dynosphere” was not successful.

234
SAWE Paper No. 3602
Category Number 31.0

Figure D.03a – THE 2005 “WHEELSURF” RECREATIONAL VEHICLE

Figure D.03b – THE 1932 PURVES “DYNOSPHERE”

Only slightly less strange than the “unicycle” configuration is the two-wheeled
configuration called a “dicycle”. This oddity consists of two wheels side-by-side on a single
axle line of relatively narrow track. Otherwise the configuration tends to be functionally much
like the unicycle configuration, although with obvious advantages in load carrying capacity,
directional stability, and steering. Curiously, despite any such seeming advantages, relatively

235
SAWE Paper No. 3602
Category Number 31.0

few dicycle type vehicles 344 have been built with respect to the unicycle, or any of the other
possible configurations. However, for those curious as to what concrete form such a rare
configuration might take, consider the “Gyrauto” designed by Italian engineer Ernest
Fraquelli circa 1935:

Figure D.04 – THE 1935 “GYRAUTO” OF ERNEST FRAQUELLI

Around 1947 Fraquelli sold his invention, and all the rights thereto, to a Belgian living
in Brussels named Edouard Vereycken, who subsequently patented it (Belgium patent number
473,555 filed 29 May 1947). However, Vereycken had no better luck than Fraquelli in
convincing the world that two wheels are better than four.

The attraction of such contraptions as the “Dynosphere” and the “Gyrauto” might stem
from the fact that, in general, the larger the wheel utilized for road transport the smoother the
ride. The following chart 345 shows the vertical movement with time of two different rigid
wheels encountering a one-inch step at 10 mph; one wheel is of a 9.8-inch radius in size, and
the other wheel is 14.57 inches in radius. Although both wheels must raise a total of one inch
to clear the step, note that the smaller wheel rises at a greater rate than the larger one:

344
Horse drawn chariots, sulkies, and carts not withstanding; for such vehicles the horse(s) provides at least a
third “point” of ground contact.
345
Reference [4], pg. 4.

236
SAWE Paper No. 3602
Category Number 31.0

Figure D.05 – VARIATION IN RIGID WHEEL RADIUS vs. BUMP

Crude calculations based on this chart indicate that the smaller wheel might be subject
to something on the order of 35 g’s vertical acceleration, compared to the larger wheel’s 30
g’s. If one were to rely on wheel radius alone, then to reduce the vertical acceleration to
something acceptable for human comfort (say, less than 1 g), a wheel on the order of a 46-
inch radius might be required. Fortunately, Napoleonic coachbuilders did not have to rely on
wheel size alone to mitigate road shock; flex in the chassis, a crude suspension, and thick
upholstery to cradle sensitive aristocratic derrières were the order of the day (as well as very
large wheels).

At any rate, it was perhaps this ancient knowledge, that the larger the wheel the better
the ride, which provided some of the motivation to develop vehicles that were essentially just
one gigantic wheel or set of wheels. It may also have to do with some inherent human drive to
reduce matters to a pristine essential, and the wheel is the most essential element, the sine qua
non, of ground transportation. One might draw a parallel with the desire of some individuals
in aviation to reduce air transport to just one big set of wings, although there appears to have
been some rather more rational considerations involved in the aero case. At any rate, large
flying wings have found acceptance, at least in the instance of the B-2 bomber, while
unicycles and dicycles have yet to make a regular appearance on public roads and highways.

However, there is one popular dicycle incarnation that merits some mention. Although
technically not an automobile, the Segway PT (“Personal Transporter”) is definitely
automotive. Introduced to the public in 2001 by its inventor, Dean Kamen, the Segway is an
electrically powered, gyroscopically stabilized (in pitch) dicycle that is ridden by its operator
in a standing position. Completely open to the elements and limited to a maximum speed of

237
SAWE Paper No. 3602
Category Number 31.0

12.5 mph, Segway use is mainly limited to sidewalks, malls, and the grounds of tourist
attractions.

Figure D.05a – SEGWAY ON POLICE PATROL

Seemingly almost as odd as the unicycle and dicycle would be the numerous attempts
to base an automobile on the bicycle configuration. Examples of this type range from Count
Shilovski’s Gyrocar of 1914 to the Ford Motor Company’s beautiful Gyron show car of 1961.
Note that we are not discussing motorcycles here, but true automobiles with fully enclosed
bodywork. Resistance to lateral roll is achieved via gyroscopic stabilization, even when
momentarily at a standstill (on such occasions the operator of a true motorcycle will stabilize
the vehicle against roll over by using his feet) 346.

346
For long term standing in place, as when parked, gyroscopically stabilized cars often employed small
“outrigger” or “bicycle training wheel” type arrangements.

238
SAWE Paper No. 3602
Category Number 31.0

Figure D.06 – 1914 SHILOVSKI GYROCAR

Generally the attraction of utilizing this configuration seems to have been the
aerodynamic benefits of minimizing frontal area; although Count Shilovski seems to have
been concerned primarily with developing a military vehicle that could utilize narrow off-road
trails. What the Count did not seem to understand was the fact that off-road vehicles generally
require more contact area with the ground, not less. Count Shilovski aside, there is a
reasonable rationale for the development of gyroscopically stabilized bicycle configuration
designs.

Figure D.07 – 1961 FORD GYRON

239
SAWE Paper No. 3602
Category Number 31.0

This brings us to the familiar tricycle (a.k.a. “delta”) configuration, which is a single
front wheel and two rear wheels on a common axle line (a motorcycle with a sidecar
configuration may be similar, but the lead wheel is offset from the center line). This is
perhaps one of the most common configurations extant, starting with Cugnot’s artillery tractor
of 1769, to such modern examples as the Carver, Ventura, and the Mercedes Life Jet.

Figure D.08 – 2007 CARVER

Many of the earliest automotive designs were of the tricycle configuration as this
greatly simplifies the design of a steering system, and even today in many developing
countries tricycle designs may dominate the automotive landscape 347, but this design choice
tends to die out in favor of the four-wheel, or “conventional”, configuration. However, when
and where industrial and/or economic vitality is compromised, such as in Europe after WW
II, then resurgence in the popularity of such tricycle designs may be observed.

Next there is the configuration that may be termed the “reverse tricycle” (a.k.a.
“tadpole”). This is a configuration that historically has been almost as popular as the tricycle
itself, and for much the same reasons. Perhaps some of the most significant and well known
examples of this configuration are the Buckminster Fuller’s Dymaxion and Morgan Super
Sports vehicles of the 1930’s, with numerous later, though less well known, examples such as
the Flitzer (circa 1945) and the Scootercar (circa 1962) popping up in Europe post-WW II.
Even today there are modern attempts to harness the supposed advantages, generally
aerodynamic and economic, of this configuration such as Fuel Vapour Technology’s Alé or
the now defunct Aptera:

347
Reference [23], pp. 464-467.

240
SAWE Paper No. 3602
Category Number 31.0

Figure D.09 – 2007 APTERA

In the transition between the three wheeled and the conventional four wheeled there is
to be found that strange hybrid configuration, the “rhomboid”. Examples of this configuration
given concrete form would include such vehicles as Pinin Farina’s “X” of 1960, Dr. Kesling’s
Yar of circa 1978, and the Rhombus of Hunan University circa 2006. The rational for such
designs generally involves supposed aerodynamic benefits.

Figure D.10 – 1960 PININ FARINA “X”

241
SAWE Paper No. 3602
Category Number 31.0

Regarding the “conventional” configuration of four wheels on two axle lines, one line
forward and one line aft, not much need be said at this point; examples of this configuration
are everywhere and most technically inclined people are at least somewhat cognizant of the
characteristics of this design. For identification of configurations to be found beyond this
familiar point the Whyte terminology used to describe early locomotive configurations must
be adopted as there is no alternative nomenclature.

Next up would be the “4-2-0”, which would be four wheels forward on two separate
axle lines in tandem, plus two wheels aft on a common axle line. The most famous example of
this type would be designer Derek Gardner’s Elf Tyrrell Formula 1 race car of 1976, plus
various and seemingly less purposeful applications such as the Covini Engineering C6W
sports car (circa 2007) and even Lady Penelope’s perambulating contraptions (initially Rolls
Royce, then Ford) as featured in the popular British children’s TV program “The
Thunderbirds”.

Figure D.11 – LADY PENELOPE’S THUNDERBIRD

Next would be the “2-4-0” configuration, which is just a reversal of the “4-2-0”; the
single axle is now forward and the two axles are now aft. Examples of this type range from
the Kurtis Kraft - Offenhauser KK500G “Pat Clancy Special” Indianapolis Speedway car of
1948, to designer Robin Herd’s March Formula 1 race car of 1977.

242
SAWE Paper No. 3602
Category Number 31.0

Figure D.12 – 1948 KURTIS KRAFT – OFFENHAUSER KK500G “PAT CLANCY


SPECIAL”

The last two basic configurations are often difficult to categorize even by the Whyte
method, due to possible variations in steering and drive wheels there can be 23 sub-types (see
Figure D.02). A common 6-wheel configuration is merely two wheels on a common axle line
occurring at three more or less equally spaced longitudinal locations, while the 8-wheel
variation is much the same except the axles are at four equally spaced locations. Examples of
such arrangements would include such vehicles as the extremely successful German Leichte
und Schwerer Panzerspahwagen (“light and heavy armored-scout-car”) designs of WW II,
or the modern General Dynamics Stryker MGS 348. Other possible examples of such
configurations include the Ferrari 312T8 Formula 1 race car of 1976 (never developed), the
circa 2002-2007 University of Keio KAZ and Eliica experimental electric cars, which may
have be considered a sort of a 4-4-0 car that morphed into a 0-8-0, and even the prophetic but
commercially unsuccessful Reeves Octoauto of 1911, which could be designated a 4-2-2.

348
Reference [100], pg. 44.

243
SAWE Paper No. 3602
Category Number 31.0

Figure D.13 - 1942 SCHWERER PANZERSPAHWAGEN (Sd Kfz 231)

This completes a brief historical overview of what may be considered the most basic
vehicle configuration variations, but it is by no means exhaustive. For instance, for every
“basic” configuration identified there can be infinite variations in the relative size of the
wheels and/or variations in track, front to rear. Also, considerations of which wheels are to be
driven, which wheels are to be steered 349, and which wheels are to provide braking, also
provide for infinite variation. And, of course, there is the inter-acting matter of propulsion
type and location, which generally has significant effect on center of gravity location and all
other vehicle mass properties.

349
Reference [59], pp. 874-875.

244
SAWE Paper No. 3602
Category Number 31.0

APPENDIX E – RECIPROCATING PISTON ENGINE INERTIAL ENERGY LOSS

Perhaps the most curious aspect of the modern piston engine is that it utilizes a
reciprocal motion more reminiscent of the inefficient reciprocating motion of nature (people,
monkeys, birds, fish, etc.) than some of mankind’s more efficient creations (wheeled). The
reciprocating motion characteristic of the piston engine has been dismissively referred to as
“monkey motion”, and with good reason. This fact has long been recognized, and much effort
has expended to find a rotary substitute for the reciprocating, such as the Wankle engine or
the gas turbine, but as of this writing the reciprocating engine still reigns supreme for
automotive propulsion.

As noted in Chapter 2, the engine “rotating” (which includes the reciprocating pistons,
valves, etc.) masses contribute tremendously to the automotive effective mass, especially in
the lower gears, yet a simple minded reduction of those masses is not possible. The engine
rotating mass term “I2” includes the flywheel, for which a certain amount of mass is
essential to its function as an energy storage device, and the crankshaft, for which the
inclusion of several large heavy counterweights is also essential to smooth out the inertia
pulses of reciprocation. It is because of such complications that the reduction of the “I2”
contribution to the effective mass was not considered in this paper, despite being very
desirable.

However, that does not mean that attempts to reduce the “I2” term have not been
attempted in the past (which usually involved such things as “slipper” pistons, titanium con-
rods, trimmed counterweights, light-weight flywheel, etc.), or that such efforts are not
ongoing today. The October 2013 issue of the SAE journal Automotive Engineering reports
that Chrysler has been engaged in a 3-year, $30 million dollar R&D program under United
States Department of Energy contract to improve gasoline fueled, reciprocating piston engine
efficiency by 25% over the present norm 350. Among such things as duel-stage
turbochargers 351, multi-fuel operation, and cooled EGR with secondary air injection, the
project also investigated the elimination of heavy balance shafts 352, and the consequent
parasitic energy losses, through the use of a special crankshaft utilizing dynamic
counterweights (“pendulums”) at each crankshaft “throw”; whether this approach will be
successful enough to make it to production remains to be seen.

350
Reference [10], pg.16. The “norm” was represented by a 2009 Chrysler 4.0 L port-injected V6, which served
as the project baseline engine.
351
The duel-stage turbocharger set-up minimizes what is known as “turbo-lag” through some clever mass
properties engineering; the smaller first stage turbo has an especially low rotational-inertia compressor to
minimize “spool-up” time.
352
The 1928-1934 Duesenberg J was powered by a 419.7 cid DOHC straight-eight “hemi” engine. To the
crankshaft of this remarkable engine was bolted two containers, each partially filled with 16 oz (0.4536 kg) of
mercury. The sloshing of the mercury within the containers provided significant vibration damping, although
with a toxic risk that would not be allowed today.

245
SAWE Paper No. 3602
Category Number 31.0

Determination of the rotational inertia of a piston engine as commonly used for


automotive propulsion is a variable inertia problem whose solution is complex but
manageable. The variation in the rotational inertia with the angle of rotation is primarily the
effect of the reciprocating motion of the piston and connecting rod. According to the Shock
and Vibration Handbook (Harris and Crede) the rotational inertia “J” of a piston engine may
be approximated by the following equation 353:

𝑾𝒑 𝒉
𝑱 = 𝑰𝒄𝒓𝒂𝒏𝒌𝒔𝒉𝒂𝒇𝒕 + 𝑵 � + 𝑾𝒄 �𝟏 − �� 𝑹𝟐 (EQ. E.01)
𝟐 𝟐

Where:

J = piston engine rotational inertia (lb-in2).

Icrankshaft = piston engine crankshaft rotational inertia (lb-in2).

N = piston engine number of cylinders.

Wp = weight of piston and wristpin with some allowance for oil (lb).

Wc = weight of connecting rod (lb).

h = con-rod C.G. location as the fraction “h′/l” of rod length (see Figure F.01).

Figure E.01 – SCHEMATIC DIAGRAM OF CONNECTING ROD

353
Reference [32], Equation 38.2.

246
SAWE Paper No. 3602
Category Number 31.0

The calculation of “Icrankshaft” may be accomplished by the usual methods of “weight


accounting”. Such “usual methods” may constitute the traditional but tedious “hand calc”
technique of breaking down the crankshaft into standard volumes 354, multiplying by the
material density, and “summing”; or the utilization of the “mass properties analysis”
function of CATIA or whatever 3D CAD/CAM system the crankshaft may be modeled in.
The example engine for this exposition is the Jaguar XK150S 3.4L straight-six engine of
1958, so of necessity the traditional and tedious “hand calc” method was employed 355,
resulting in the following values:

Icrankshaft = 151.11 lb-in2

Wp = 2.52 lb

Wc = 1.52 lb

h = 0.272

R = 2.1 in

“Plugging” these values into Equation E.01 produces the following result for the
rotational inertia “J” of the crankshaft/con-rods/piston assembly:

𝟐. 𝟓𝟐 𝟎. 𝟐𝟕𝟐
𝑱 = 𝟏𝟓𝟏. 𝟏𝟏 + 𝟔 � + 𝟏. 𝟓𝟐 �𝟏 − �� 𝟐. 𝟏𝟐
𝟐 𝟐
= 𝟐𝟏𝟗. 𝟐𝟎 𝒍𝒃 − 𝒊𝒏𝟐
This approach obscures the variable inertia nature of the piston engine rotating
mass, which is of considerable significance with regard to the engine induced sprung mass
vibration problem, and with regard to the problem of energy loss (vibration, sound, light, and
heat are all forms of energy loss) which adversely affects engine efficiency. To account for
the inertia variation with rotation this author developed an appropriate equation based on a
consideration of the inertia forces resulting from the angular acceleration “α1” as per the
following single cylinder free-body diagrams:

354
Reference [32] presents some specialized formulae for the calculation of the inertia of the crankshaft webs
which may be used instead, and it is stated that the formulae are also applicable to marine propellers with blades
of “ogival” section.
355
Jaguar XK150S 3.4L engine “engineering” drawings may be found in Reference [55].

247
SAWE Paper No. 3602
Category Number 31.0

Figure F.02 – SINGLE CYLINDER INERTIA ABOUT CRANKSHAFT AXIS

The torque “T” about the crankshaft axis is equal to the sum of all the inertial
resistances:

𝒍 − 𝒉′ 𝟐 𝑰𝟐 𝜶𝟐
𝑻 = 𝑰𝟏 𝜶𝟏 + 𝒎𝟐 � � 𝑹 𝜶𝟏 + �𝒎𝟑 𝒂𝟑 + � � 𝐬𝐢𝐧 𝜽� 𝑹
𝒍 𝒍
Substitute “α1 R sin(θ)/l” for “α2” and substitute “α1 R” for “α3”:

𝒍 − 𝒉′ 𝟐 𝑹 𝟐
𝑻 = 𝑰𝟏 𝜶𝟏 + 𝒎𝟐 � � 𝑹 𝜶𝟏 + 𝒎𝟑 𝜶𝟏 𝑹 + 𝑰𝟐 𝜶𝟏 � � 𝒔𝒊𝒏𝟐 𝜽
𝟐
𝒍 𝒍
Divide through by “α1”:

′ 𝟐
𝑻� = 𝑰 + 𝒎 �𝒍 − 𝒉 � 𝑹𝟐 + 𝒎 𝑹𝟐 + 𝑰 �𝑹� 𝒔𝒊𝒏𝟐 𝜽
𝜶𝟏 𝟏 𝟐
𝒍 𝟑 𝟐
𝒍

This is the effective rotational inertia (“Ieff = T/α1”) in terms of “θ”, but calculation
is more convenient in terms of crankshaft angle “ψ” as it is the crankshaft rotation which
causes the inertial flux. The relationship of the crankshaft angle “ψ” to connecting rod angle
“θ” and how the effective “R’” varies with “ψ” may be determined from the following
diagram:

248
SAWE Paper No. 3602
Category Number 31.0

Figure F.03 – CRANKSHAFT ANGLE vs. CONNECTING ROD ANGLE

So now we may substitute “R cos(ψ)” for “R” and “sin-1(R/l) cos(ψ)” for “θ”, also we
may change “T/α1” to “Ieff” or “J” (for better unity in symbolism with Equation E.01):

𝒍 − 𝒉′ 𝟐
𝑱 = 𝑰𝟏 + 𝒎𝟐 � � 𝑹 + 𝒎𝟑 (𝑹 𝐜𝐨𝐬 𝝍)𝟐
𝒍
𝟐
(𝑹 𝐜𝐨𝐬 𝝍)𝟐 −𝟏
𝑹
+ 𝑰𝟐 �𝒔𝒊𝒏 �𝒔𝒊𝒏 � � 𝐜𝐨𝐬 𝝍��
𝒍𝟐 𝒍
This is valid for only a single cylinder engine, for an engine of “N” cylinders the
equation becomes:

𝒍−𝒉′ 𝟐𝝅𝑲 𝑹 𝟐
𝑱 = 𝑰𝟏 + 𝑵𝒎𝟐 �
𝒍
� 𝑹𝟐 + ∑𝑲=𝑵 𝟐 𝟐
𝑲=𝟏 �𝒎𝟑 𝑹 𝒄𝒐𝒔 �𝝍 − 𝑵
� + 𝑰𝟐 � 𝒍 � 𝒄𝒐𝒔𝟐 �𝝍 −
𝟐𝝅𝑲 𝑹 𝟐𝝅𝑲
𝑵
� 𝒔𝒊𝒏𝟐 �𝒔𝒊𝒏−𝟏 � � 𝐜𝐨𝐬 �𝝍 −
𝒍 𝑵
��� (EQ. E.02)

Where:

249
SAWE Paper No. 3602
Category Number 31.0

J = rotational inertia of crankshaft/con-rods/pistons assembly (lb-in2).

I1 = crankshaft rotational inertia (lb-in2).

N = number of cylinders.

m2 = weight of connecting rod (lb).

l = distance along connecting rod between centers of rotation (in).

h′ = distance along connecting rod between crankshaft center and CG of connecting


rod (in).

R = crankshaft “throw” distance (in).

K = counter variable.

m3 = weight of piston, piston rings, wristpin, plus “h′/l” fraction of “m2” (lb).

ψ = crankshaft angle (radians).

I2 = rotational inertia of con-rod about the wristpin center (lb-in2).

An equation such as Equation E.02 is best evaluated by means of a computer program,


hence the following listing of the Commodore BASIC language program (c. 1982)
“CYLINERT.BAS” (“N” is input at line 10, plotting info is required lines 100 to 150, all
other input values are embedded in lines 2000 to 2005)…

CYLINERT.BAS

250
SAWE Paper No. 3602
Category Number 31.0

251
SAWE Paper No. 3602
Category Number 31.0

252
SAWE Paper No. 3602
Category Number 31.0

The output from this program, using values appropriate for the Jaguar XK150S 3.4L
straight-six engine of 1958, is as follows… 356

Figure E.01 - CYLINERT.BAS OUTPUT

356
Note this analysis concerns itself only with what may be termed “primary” forces and couples; there are
“secondary” forces and couples of a lesser magnitude not addressed; see Reference [59] pp. 25-39.

253
SAWE Paper No. 3602
Category Number 31.0

The inertial flux with rotation can clearly be seen for “N = 1, 2, 4, 5”. It is the
combination of these inertial pulses plus the power pulses resulting from the ignition cycle
that compose most, but hardly all 357, of the piston ICE vibration output (exhaust pulsations
running through the exhaust system are also significant). Note that for “N = 6” the inertial
flux seems completely nonexistent, which explains the traditional popularity of big straight-
six engines such as those found in early XK and E-Type Jaguars.

Lastly, interpretation of the “N = 6” plot indicates that the internal rotational inertia
of the 1958 Jaguar 3.4L engine is about 220 lb-in2, which is in close agreement with the 219.2
lb-in2 obtained by use of Equation F.01. Unfortunately the determination of the rotational
inertia of the crankshaft/con-rod/piston assembly is only the first step in obtaining the total
engine rotational inertia about the crankshaft axis “I2”. To this first step value must be added
the inertial contribution of the camshafts and valve train 358, flywheel, and various pulleys,
sprockets, belts, chains, engine accessories. For the 1958 Jaguar 3.4L engine the inertia total
was determined to be:

Crankshaft 220.00 lb-in2


Camshafts/Valve Train 2.95 lb-in2
Flywheel 716.94 lb-in2
Crankshaft Pulley 6.91 lb-in2
Pulley Nut 0.09 lb-in2
Drive Sprocket 0.35 lb-in2
Idler Sprockets 0.70 lb-in2
Fan Belts 1.45 lb-in2
Camshaft Chain 2.96 lb-in2
Fan 15.03 lb-in2
Water Pump 0.25 lb-in2
Distributor 0.05 lb-in2
Generator 8.50 lb-in2
“I2” Total 976.18 lb-in2, or 6.78 lb-ft2
or 0.2857 kg-m2

357
The piston/con-rod/crankshaft vibration in the Y-Z plane as considered herein may be thought of as the
“primary” vibration, but there are secondary and maybe even tertiary/quaternary/etc. vibrations as well.
Formulae for the estimation of such vibrations may be found in the literature.
358
Reference [83] has an example on page 16 may be helpful in the evaluation of the valve train inertia.

254
SAWE Paper No. 3602
Category Number 31.0

APPENDIX F – NOTES ON TIRE BEHAVIOR

The primary forces which determine the dynamic behavior of aircraft are aerodynamic
forces generated by pressure differentials acting over the aerosurface areas. In contrast, the
primary forces which determine the dynamic behavior of automobiles are friction forces
generated by contact pressure acting over the tire-to-road contact areas.

It is the tires that transmit the forces that accelerate, decelerate, and maneuver the
automotive road vehicle. It is the tires that play a major role in isolating the vehicle, its cargo
and passengers, from the shock and vibration effects of road surface irregularities. Last, but
not least, the tires play an absolutely critical role in providing vehicle directional stability.
What tires do is necessary and very complex, so much so that in nearly 125 years of
development no adequate substitute has been found for the pneumatic-elastic rubber and cord
structure known as the tire. The tire has prevailed over all those years, undergoing
innumerable improvements and refinements, despite still not being fully understood in its
mechanisms and behavior.

Most elementary study of friction is limited to what is termed the dry or Coulomb 359
friction case. This model is limited to simplified tangential (shearing, sliding) contact between
two relatively rigid, smooth bodies; the model situation is typically depicted as follows 360:

Figure F.01 – COULOMB FRICTION SITUATION

The resulting Coulomb friction relationship is mathematically modeled as:

F=μN (EQ. F.01)


359
Charles-Augustin de Coulomb (1736-1806). Although the name of this famed French physicist and engineer
is the one most commonly associated with the “classic” dry friction model, Leonardo da Vinci (1452-1519),
Guillaume Amontons (1663-1705), Pieter van Musschenbroek (1692-1761), and Leonhard Euler (1707-1783) all
made prior contributions to the study of friction. Sometimes the “classic” friction model is referred to as
“Amontons-Coulomb” friction.
360
Reference [6], pp. 273-277.

255
SAWE Paper No. 3602
Category Number 31.0

The normal load “N” is often just the weight of the simple block of material typical of
such illustrations, and the friction force “F” available to resist any motion is related to the
normal force by the friction coefficient “μ”. This coefficient is considered to have two distinct
sets of values, static and dynamic 361.

Table F.01 – TYPICAL COULOMB FRICTION COEFFICIENTS

“F = μ N” is not a fundamental law of physics like “F = m a”, but instead is a highly


circumscribed description of physical behavior that is accurate enough to be useful only
within certain specified limitations:

1. Relatively rigid bodies, such as


per most common engineering
materials (e.g.: steel, aluminum,
copper, glass, etc.).
2. Reasonably smooth, dry contact
surfaces (no lubrication).
The reason “F = μ N” is not a universally applicable is because the actual nature of
friction is much more complex than indicated by the useful simplicity of the Coulomb friction
relationship, and not fully understood 362. The tangential forces between bodies in contact may
arise from a number of contributing phenomena:

1. Van der Waal’s forces.


2. Electrostatic forces.
3. Mechanical forces.
4. Hysteresis forces.

361
Reference [73], pg. 35. Table G.01 is a somewhat simplified version of a table which is presented in the
referenced material.
362
At least not for tires in 1974 per Reference [34], page 73.

256
SAWE Paper No. 3602
Category Number 31.0

This list may not be complete, but the point is that the seeming nature of the friction
relationship is dependent upon numerous phenomena of which certain type(s) may become
overwhelmingly dominant depending on the exact nature of the circumstances, thereby
allowing for a gross simplification such as “F = μ N” only when given the limiting
circumstances noted. The exact nature of contact involves the material composition of the
bodies in contact, the roughness of the surfaces in contact, the velocity of the surfaces with
respect to each other, the magnitude of the contact area, the condition of the surfaces in
contact (any lubrication or contamination 363), and the temperature of the contact surfaces.

Given the circumscribing conditions associated with Coulomb friction, the dominant
phenomena are such that the resulting friction force potential is seemingly independent of the
contact area, independent of velocity, independent of temperature, isotropic in the plane of
contact, and with the static and dynamic friction coefficients independent of the normal load.
However, for tires the situation is completely the opposite in all those respects. To emphasize
this seeming behavioral difference from the commonly understood Coulomb friction
case, tire-to-road friction is better referred to as “traction”.

TIRE TRACTION: MATERIAL

As noted, tire traction is notably very dependent upon contact area. This behavior is
resultant of the fact that tires principally consist of “rubber” (including “natural” rubber and
various artificial or synthetic “rubbers”); rubber has characteristics very different from the
more prosaic engineering materials such as those listed in Table F.01.

Rubber is not a “rigid” material like iron or steel; it is a very “elastic” material. That is
to say that when a tensile force is applied to a length of rubber it will stretch to many times its
original length, and when the force is relieved the rubber item will return to that original
length without any permanent deformation. Robert Hooke (1635-1703) was one of the
earliest to experiment with material samples subjected to tensile loads, and he noted an initial
proportional relationship between force and deflection for samples of most engineering
materials, “Ut tensio sic vis” 364:

F ∝ ΔL (EQ. F.02)
However, for studies of force and deflection to be really useful, and to reflect the
underlying properties of just the materials from which the test samples were constructed, the

363
For example, any dust or grit on the contact surface constitutes a contamination which will affect the friction.
364
Reference [75], pg. 28. The phrase “As strain, so force” (literal English translation of the original Latin)
became well known when Hooke published his structures research in 1678.

257
SAWE Paper No. 3602
Category Number 31.0

force “F” and the deformation “dL” parameters need to be “normalized” using the original
cross-sectional area “A” and original length “L” like so:

Normalized Force = Stress = F/A (EQ. F.03)

Normalized Deflection = Strain = ΔL/L (EQ. F.04)

Now that the effects of variation of cross-sectional area and of length of the test
sample have been “divided out”; all that is left is the properties of the materials themselves.
Possibly one of the first to “realize” this was Thomas Young (1773-1829), although he did so
for a specific unique case of the pressure at the base of a column (due to its own weight load)
divided by the unit compression; not as a result of a general study of normalized force-
deflection diagrams such as:

Figure F.02 – TENSION STRESS-STRAIN DIAGRAM (STEEL)

However, given Young’s observations in his course material, which he published after
he retired from his post as Professor of Natural Philosophy at the Royal Institute of England
in 1803, it was not long before the ratio of stress to strain in Hooke’s proportional region
(from point 0 to point 1) became known as the “Young’s Modulus” in his honor. Later this
nomenclature was superseded by the term “Modulus of Elasticity” (symbol “E”), although
this is a misnomer (a more accurate terminology would have been “Modulus of Stiffness”, but

258
SAWE Paper No. 3602
Category Number 31.0

that would lack the inherent implication of applicability within the elastic realm) 365. Using the
Modulus of Elasticity, “Hooke’s Law” (Equation F.02) regarding the linear relation between
force and deflection (now normalized) becomes:

σ=Eε (EQ. F.05)

Previous Figure F.02 presented a tensile stress-strain diagram (Quadrant I) more or


less typical in shape of many traditional ductile engineering materials such as copper, bronze,
steel, etc., in that there are well defined transition points such as Figure F.02 points 1 through
5 366. Figure F.02 is most representative of steel in this regard; other traditional engineering
materials may be missing one or more of these points depending upon the degree to which the
material is “ductile” or “brittle”, etc. However, most such materials behave very much the
same in compression as in tension; the “first quadrant” plot of tensile behavior would merely
tend to be an inverted reverse “third quadrant” reflection of the first quadrant for the
proportional region, with ultimate strength being much higher due to the lack of crack
propagation in compression.

Generally when a stress-strain diagram is presented it is only as a Quadrant I plot. A


figure showing all the quadrants of the stress-strain diagram is presented as Figure F.03. For
Figure F.03 the copper Modulus of Elasticity was taken as about 21.0x106 psi, with a yield
point of 53.6x103 psi. For cast iron, which is a “brittle” material, the Modulus of Elasticity
was taken as being about 15.9x106 psi, with an Ultimate Strength in tension of about 35.0x103
psi.

It is because “E” is generally much the same for most engineering materials in
compression as it is in tension, and because testing in compression presents more practical
problems (buckling, damage to the testing machinery), that the compression stress-strain
diagram (Quadrant III) is generally not done. And, of course, Quadrants II and IV have no
physical meaning here.

365
Reference [75], pg. 31. Maybe an even better term would be “Ricardi’s Modulus”, as Giordano Ricardi in
1782 had determined the Modulus of Elasticity for brass and steel from his vibration study of chimes made from
those materials. It is from such vibration studies that the most accurate modulus values are obtained, and such is
a common technique today for most serious modulus determinations. However, despite his advanced technique,
Ricardi’s findings failed to attract much attention at the time.
366
Point 1: Proportional Limit. Point 2: Elastic Limit. Point 3: Yield Point. Point 4: Maximum Stress. Point 5:
Rupture. Not all points will be present for every common engineering material; see how copper varies from steel,
and copper and steel from cast iron.

259
SAWE Paper No. 3602
Category Number 31.0

Figure F.03 – TENSION & COMPRESSION STRESS-STRAIN DIAGRAM

However, shear and pressure stress-strain testing is commonly done for the
determination of the “Shear Modulus” (symbol “G”) 367 (which is virtually the same as
“Torsion Modulus”) and for the “Bulk Modulus” (symbol “B”) 368, respectively. Elasticity,
Shear, and Bulk constitute the basic trio of material properties moduli (there is generally a
plethora of other more specialized “moduli” that are beyond the scope of this discussion), and
are interrelated to each other by a quantity known as Poisson’s Ratio (“ν”) while in the linear
elastic range 369, which for most engineering materials ranges from 0.25 to 0.35 (while rubber
behavior is not linear, it can be approximated as linear at points over a “wide” range, for
which “ν” is taken as 0.50, which is the theoretical maximum for the ratio):
367
The material sample of height “H” undergoes a shear loading “F” causing “layers” of area “A” to slide past
each other an infinitesimal distance “∆x”: G = (F/A)/(∆x/H).
368
The material sample of original volume “V” undergoes an external pressure change “∆P” over its entire
surface, and consequently experiences some decrease in volume “∆V”: B = ∆P/(∆V/V).
369
Reference [75], pg. 39. Reference [84], pg. 11. Siméon Denis Poisson (1781-1840) defined the ratio “ν” in
terms of the orthogonal unit deformations (strains) within the proportional limit (point 1 of Figure F.02). For a
derivation of Equation F.07 see pages 360-362 of Reference [75].

260
SAWE Paper No. 3602
Category Number 31.0

𝑬
𝑮= 𝟐(𝟏+𝛎)
(EQ. G.06)

𝟑𝐁−𝟐𝐆
𝛎= 𝟔𝐁+𝟐𝐆
(EQ. G.07)

In contrast to the “traditional” engineering materials, rubber is very different in its


behavior. Rubber belongs to a strange class of materials known as “elastomers” which, as
noted earlier, can stretch many times the original length and then return to that original length
without permanent deformation. It is because rubber is so different in behavior that it is
seldom, if ever, depicted in a stress-strain comparison such as follows:

Figure F.04 – RUBBER vs. STEEL STRESS-STRAIN BEHAVIOR

Even in Figure F.04, the rubber stress values had to be exaggerated and the strain
understated for the sake of the visual presentation. If plotted to a true scale, it would be
difficult if not impossible to visually discern the rubber plot from the strain axis for most of
the elongation. If the elongation under load, and the subsequent unloaded contraction, of a
rubber sample were plotted on a scale more appropriate to the material’s unique behavior,
then the result would be more like:

261
SAWE Paper No. 3602
Category Number 31.0

Figure F.05 – RUBBER STRESS-STRAIN DIAGRAM, HYSTERESIS

In this diagram 370 rubber (probably natural rubber or “cis-1,4-polyisoprene”) is


stretched to over seven times its original length (“loading”), yet when relieved of load
(“unloading”) returns to its original length with no permanent deformation. Note that there
also that over no portion of this stress-strain plot is there any true linear portion which
conforms to “Hooke’s Law”: “σ = E ε”. However, rubber materials can still be assigned a
Modulus of Elasticity (“E”) which generally represents a linearized approximation of the
behavior in the mid-range area; such an “E” is taken as being equal to three times the Shear
Modulus (“G”), courtesy of Poisson’s ratio (“ν = 0.50”) and Equation F.06:

E=3×G (EQ. F.08)

The diagram reveals yet another curious aspect of the nature of rubber: high
“hysteresis”. Most materials subjected to cyclical stress at sufficient levels will exhibit some
conversion of mechanical energy to thermal energy, but rubber does so abundantly at
relatively low stress levels; the area between the “loading” and the “unloading” curves
represents the magnitude of this energy loss.

370
This diagram is very similar in appearance to the load-deflection curve of an Austin/Morris Mini rubber
“spring” per Reference [4], page 205.

262
SAWE Paper No. 3602
Category Number 31.0

It is the curious properties of rubber which make it so valuable a material for the
manufacture of mounts for vibrating machinery, suspension spring-damper units, tires 371, and
many other things. And where the properties of one kind of rubber don’t quite fit the use
intended, there are now many compounds of “natural” rubber 372, along with innumerable
totally synthetic variants, which fill most needs.

Having noted how unusual rubber is in its elasticity, hysteresis, and other properties;
we now have some clues as to the reasons for that unusual frictional behavior known as tire
traction. At the small scale level of the contact area between tire and road the situation looks
like (greatly enlarged):

Figure F.06 – VIEWS OF FRICTION & TRACTION MICRO-BEHAVIOR

371
The original Alec Issigonis designed Austin “Mini” (project code ADO15) of 1959-1964 utilized an Alex
Moulton suspension with rubber cones constituting the spring elements. The cones also supplied some damping,
but conventional hydraulic dampers were required to provide some assistance in that regard.
372
Rubber in its absolutely natural state tends to be very difficult to use due to extreme variation in its physical
characteristics with temperature, ageing, etc. Natural rubber only became truly useful after the discovery of the
“vulcanization” process by Charles Goodyear in 1839. This process involved the chemical mixing of rubber with
sulfur using heat; the resulting compound has vastly stabilized physical properties. In 1906 S.C. Mote of the UK
introduced the addition of carbon black to the rubber compound used for automobile tires; carbon black
improved the strength and thermal conductivity of rubber, and thereby increased tire durability. It also changed
the appearance of tires from the classic off-white of natural rubber to the black now not only associated with tires
but with most things rubber. However, “white wall” tires would continue to be produced for quite a while
afterward in homage to the “classic” look of earlier tires.

263
SAWE Paper No. 3602
Category Number 31.0

Note how on the left in the Coulomb friction case the relative rigid and smooth nature
of the contact surfaces allow for the rigid block to only make contact with the supporting
surface only at the high spots. The block slides along the supporting surface riding on only
those high areas, and increasing normal load does not greatly change the nature of the contact.
Due to the smoothness such contact tends to be very close, such that the principal resistance to
tangential motion is due to attractive forces between molecules. There may be a small amount
of mechanical shearing off of some of the high points (resulting in wear), but mainly
mechanical shear and hysteresis forces play a very minor role.

In contrast, in the traction case depicted on the right, the elastic nature of rubber
allows for the rubber to conform to the roughness of the road surface. Although this
mechanical “interlock” of the tire tread rubber to the road surface results in a situation much
like a set of locked gears in mesh, the “gear teeth” do not necessarily “break off” (shear) in
order to allow relative tangential motion between surfaces. Some rubber particles may shear
off as a result of the tangential strain, which constitutes wear of the tire tread, but the extreme
elasticity of rubber allows for a sort of “flow” of the rubber “teeth” up over road surface
“hills” and down into the following “valleys”. Of course, such motion means a cyclical
elongation-contraction occurs, resulting in a tangential resistance to motion and dissipation of
energy as heat due to rubber hysteresis 373.

Increasing normal load forces an increasing amount of rubber down over the “hills”
and into the “valleys”, causing a greater and greater mechanical “interlock”, but to an ever
decreasing degree. The result is an ever increasing lateral force potential (traction) with
normal load, but to an ever decreasing extent. Thus the force ratio (coefficient of traction) is
not a constant as with the Coulomb friction model, but instead is a function of the normal load
“N”:

μ = f(N) (EQ. F.09)


Such traction would also vary significantly with contact area, as increasing contact
area would mean more mechanical interlock, like more teeth in mesh for a gear set. To
consider the traction behavior of rubber as a material the effect of contact area has to be
“divided out” (Pc = N/Ac). That is, the coefficient of traction has to be looked at as a function
of the resulting contact pressure “Pc”; the form of such a rubber traction function may be as
determined by Koutný:

373
Reference [8], pp. 3-5. The “Kummer Theory of Tire Friction” attributes tire traction to hysteresis,
mechanical interlock, and molecular attraction (van der Waals) mechanisms. Electrostatic forces are essentially
ruled out because humidity does not seem to be a significant factor affecting tire traction; this may be the result
of carbon black, a major component of “modern” tires, possibly causing static charge dissipation due to its
electrical conductivity, preventing any significant charge build-up (this should be easy to confirm by
experiment).

264
SAWE Paper No. 3602
Category Number 31.0

μ = a Pcn (EQ. F.10)

Where “a” may have a value such as “15.7369” (for “P” in units of psi, “58.2587” for
units of kPa) and “n” a value of “-0.67791”:

μ = 15.7369 Pc-0.67791 (EQ. F.11)


Graphically, such a rubber traction-pressure function may be depicted as follows 374:

Figure F.07 – RUBBER TRACTION COEFFICIENT vs. CONTACT PRESSURE

It should be noted that because rubber traction is so dependent upon the material being
forced down into the irregularities of the opposing surface that it follows that the smoother
that surface the more dependent the traction is on normal load 375. Also, the harder the rubber
compound the more the coefficient of traction becomes less dependent on contact pressure,
and more like that of steel. However, such hard compounds are not the sort that tires are made
of, and normally need not be of concern. However, it should also be noted that the behavior of

374
Reference [40], pg. 61. Reference [26], pg. 535. Unfortunately, even the testing of a simple rubber block of
shallow thickness (height) will still result in some distortion and consequent contact area variation. However,
increasing the contact area to thickness (height) ratio of the rubber test sample can minimize such variation to the
point where it becomes negligible. Reference [34], page 75 states “…there is evidence that, in the case of
contact between rubber and hard surfaces (such as road pavements), the friction coefficient is a function
of…the normal pressure”. This seems to indicate that a general awareness of the tire traction coefficient
dependence on normal load was just starting to permeate the industry circa 1971; Colin Chapman understood
this much earlier.
375
Reference [26], pg.422.

265
SAWE Paper No. 3602
Category Number 31.0

rubber can vary significantly with temperature, and that with very low temperatures tires can
still behave as if they were made of very hard rubber, so Antarctic expeditions beware!

TIRE TRACTION: STRUCTURE

Rubber as a material is isotropic in its frictional behavior, yet when rubber is


embodied in the “macroscopic” structure known as a tire then the resulting behavior is
anisotropic! Tire structural behavior is often thought of as being essentially that of a simple
“balloon” 376, even by people who should know better 377. The idealized balloon is a simple
membrane owing its shape solely due to the pressure of an entrained gas; the behavior is
simple and isotropic. Tires, however, possess a definite shape even without inflation; tires are
not a simple membrane but a complex structure of various rubber compounds, cords, and
belts. The very shape of a tire promises anisotropic behavior, behavior quite different from
that of a simple balloon whose spherical shape tends to be the same from any aspect.

Figure F.08 – TIRE STRUCTURE vs. BALLOON STRUCTURE

For instance, consider an automobile at curb weight condition riding on a set of tires
inflated to 32 psi, front and rear. The addition of passengers and cargo up to the vehicle’s

376
It is possible that the mistaken concept of how a tire works obtained credence in the public mind with the
advent of “balloon tires” in the 1920’s. The term “balloon tire” referred to the lower pressure (circa 40 psi) and
lower aspect ratio tires which supplanted the previous generation of tall, narrow, high pressure (circa 90 psi)
tires. There were also internal changes with the tire carcass which morphed from a woven cloth construction to
one of bias-ply cords, with the material tending to change from natural fibers to synthetics.
377
Reference [26], page 2 provides an excellent example: a tire inflated to 35 psi is said to require 10 square
inches of tire-road contact area to support 350 pounds. This is not a correct illustration of tire behavior.

266
SAWE Paper No. 3602
Category Number 31.0

GVWR, and a subsequent check of the tire inflation pressures, reveals that despite the
significantly greater tire contact pressures required to support the increased load, the tire
inflation pressures do not appear to have changed at all 378. This is because the air pressure
within the tire does not directly support the load; the inflation pressure merely maintains the
positioning of the complex tire carcass system of cords and belts so that those members in
tension will, like the cables of a suspension bridge, support the load 379. Admittedly this is
more difficult to envision and understand, hence the common “balloon” misconception.

The complex tire structure has a number of such major impacts on behavior, and one
of those impacts is in the vertical, or “Z” axis, direction. That is, how the structure affects the
normal load/deflection behavior. Overall the tire load/deflection curve resembles the rubber
load/deflection curve:

Figure F.09 – TIRE FORCE/DEFLECTION RELATIONSHIP

Note that within a range of tire normal loads (and deflections), termed the “working
range”, the load/deflection relationship may be linearized, and that linear relationship may be
represented by what some have called the Nokian formula 380:

𝒅 = 𝑵�𝑲 + 𝒅𝟎 (EQ. F.12)


𝑽

378
There may well be pressure changes on a scale beyond the common pressure gauge’s sensitivity to detect, but
such changes are very small and a consequence of volumetric changes resulting from tire deflection.
379
Reference [57], pg 29. Reference [37], pg. 51. In the latter reference the author seems reluctant to give up the
simple notion that the inflation pressure times the contact area will equal the load, but recognizes that it is
incorrect somehow. Also see Reference [93], pg. 4.
380
Reference [79], pg. 218. Figure F.09 is similar to Prof. Steed’s “Fig. 8.1.2”, but is different in that it is not for
a specific tire, and the “zero load” is truly zero (not the weight of the wheel and tire).

267
SAWE Paper No. 3602
Category Number 31.0

The significant thing here is that, while a block of rubber would have just one such
relationship, a tire has an infinite number of such relationships depending on inflation
pressure. Figure F.09 shows just the relationship for one particular inflation pressure, but
typically a whole family of such “curves” may be plotted indicating a whole range of possible
variation, as was done for a certain 5.60×13 bias ply tire:

Figure F.10 – TIRE FORCE/DEFLECT vs INFLATION PRESSURE

An authoritative source presents the situation for this same 5.60×13 bias ply tire
somewhat differently and with a few more inflation pressure values 381:

Figure F.11 – TIRE FORCE/DEFLECT vs INFLATION PRESSURE

381
Reference [13], pg. 37. Figure F.10 displays the same information as the referenced source’s “Fig. 2.9” for the
15, 25, and 35 psi inflation pressures, but is totally redrawn in a more conventional style.

268
SAWE Paper No. 3602
Category Number 31.0

If the 5.60 × 13 tire vertical spring rates for the various working ranges were plotted
for the corresponding inflation pressures the result may be considered typical:

Figure F.11a – TIRE VERT STIFFNESS vs INFLATION PRESSURE

Note the dependence of both the working ranges and vertical stiffness values upon the
inflation pressures. The working range is capped by the maximum load, or load capacity, at
the maximum inflation pressure as determined by the TRA Load Capacity Formula:

𝑳𝒄𝒂𝒑 = 𝟎. 𝟒𝟐𝟓 × 𝑲 × 𝑷𝒊 𝟎.𝟓𝟖𝟓 × 𝑺𝟏.𝟑𝟗 × (𝑫𝑹 + 𝑺) (EQ. F.13)

And the vertical stiffness (which is very closely related to the load capacity) at a
pressure is determined by the Rhynes Equation:

𝑺𝑵 ×
𝑲𝑽 = 𝟎. 𝟎𝟎𝟎𝟐𝟖 𝑷𝒊 ��−𝟎. 𝟎𝟎𝟒 + 𝟏. 𝟎𝟑�𝑺𝑵 × � 𝟓𝟎
+ 𝑫𝑹 � + 𝟑. 𝟒𝟓 (EQ. F.14)

The accuracy of this equation is indicated by the following investigation:

269
SAWE Paper No. 3602
Category Number 31.0

Figure F.11b – RHYNE’S EQUATION vs EMPIRICAL MEASUREMENT

Further structural effects result in anisotropic behavior on the horizontal “X-Y” plane,
while for a simple block of rubber, or even a rubber balloon, such behavior is isotropic.
Consider the cases of a tire providing longitudinal direction (X) traction and a tire providing
lateral direction (Y) traction 382…

Figure F.12 – TIRE LONGITUDINAL & LATERAL MACRO-BEHAVIOR

382
SAE J670 2008 “Tire and Wheel Axis Systems – Z-Up” on page 11. The “Tire and Wheel Axis Systems –
Z-Up” corresponds with the “Vehicle Axis System – Z-Up” on page 8: “+X” is forward, “+Y” is left, and “+Z”
is up. However, the “Tire and Wheel Axis Systems – Z-Up” origin is the intersection of the tire vertical
centerline with the “road plane”; the vehicle “Vehicle Axis System - Z-Up” origin is stated as being a “vehicle
reference point” which may be “the vehicle center of gravity”, among other possible points.

270
SAWE Paper No. 3602
Category Number 31.0

Even when rolling freely a tire has a longitudinal pressure distribution biased toward
the leading edge of the contact patch, as shown. Therefore, the resultant force “Fr” of this
pressure distribution is positioned at some distance “e” forward of the tire center line; the
moment “Fr × e” represents the rolling resistance. When a longitudinal traction force “Fx” is
generated due to an acceleration or braking torque (“Fx = T / r”) there is a distortion of the
contact area, but the magnitude of the area stays essentially the same 383. However, when a
tire is subjected to a lateral load such as to generate a reaction lateral traction force “Fy”
then there is a decrease in overall contact area due to distortion causing the tread “inner”
edge to “curl up” off the road surface. The extent to which this “curl up” occurs is dependent
on the loading magnitude, inflation pressure, tread shape, and the internal construction; bias
ply tires generally exhibit more “curl up” than radial tires as a result of the defining structural
details: the orientation of the cords, the use of tread belts, etc. 384.

TIRE TRACTION: LATERAL

For the lateral case there is an old “traditional” relation between the normal load “N”
and the potential maximum lateral traction force which accounts for the decrease in traction
coefficient due to increasing contact pressure. However, in the lateral case the increasing
contact pressure results not just from increasing normal load, but also from decreasing contact
area due to the “curl up” noted earlier. The traditional equation is an ingenious but simple
looking formulation that obscures the underlying complexity of the situation by determining
the lateral traction coefficient in a manner that blends both effects indistinguishably
together 385:

μy = b – mN (EQ. F.15)

383
The rolling resistance moment “Fr × e” is affected by the shear stress distribution/contact area distortion
resulting from an application of acceleration/deceleration torque. The initial effect of the application of a braking
torque may be a minor decrease in rolling resistance coefficient; but generally both acceleration and deceleration
torques will cause significant increases in rolling resistance.
384
Reference [38], pg. 48. The consequences of detail structural variation within the overall tire structural
paradigm can be very significant, but the similarity in tire behavior regardless of “bias” or “radial” structural
detail allows for a general discussion such as in this paper. However, a more detailed study would of necessity
concentrate on the differences in behavior resulting from “bias” or “radial” structural differences.
385
Reference [42], pg. 127. Reference [43], pg. 27. A “b” value of 1.2 may be considered a good value for
modern passenger car tires; values up to 1.6 might be appropriate for racing tires. For the “m” parameter a
variation of 0.0002 to 0.0006 would be considered appropriate for passenger car tires.

271
SAWE Paper No. 3602
Category Number 31.0

“Typical” passenger car tire values for the coefficients “b” and “m” might be “1.2”
and “0.0004”, respectively. How the lateral traction coefficient varies with normal load for
such parametric values may be illustrated as follows 386:

Figure F.13 – TIRE LATERAL TRACTION COEFFICIENT vs. NORMAL LOAD

The “b” represents some base value for the tire traction coefficient as obtained by
extrapolation to zero normal load, a value which gets whittled away by the diminishing
coefficient effect in conjunction with the tread “curl up” phenomenon with the total effect
represented by the function slope “m”. If the lateral traction coefficient equation is “plugged”
into the Coulomb friction equation, then the resulting relation is:

Fy = (b-mN)N = bN – mN2 (EQ. F.16)

When plotted this presents a visual illustration of how tire lateral traction force
potential varies with normal load (for “b” equal to “1.2” and “m” equal to “0.0004”):

386
Reference [93], pg. 36 shows the “cornering coefficient” (lb/deg) as varying in a shallow curve with normal
load. However, what we have here is “lb/lb”; the “deg” refers to “slip angle” which is yet to be addressed.

272
SAWE Paper No. 3602
Category Number 31.0

Figure F.14 – TIRE LATERAL TRACTION FORCE “Fy” vs. NORMAL LOAD “N”

Even though the equation by which the above was obtained represents a tremendous
simplification of a complex underlying reality, the modeling produces useful results on the
macroscopic level of human experience, and “traditionally” that has been good enough.
Figure F.14 represents a plot of lateral traction force potential vs. normal load for a specific
tire at a specific inflation pressure, but for any tire there is an infinite family of such curves
over the permissible range of inflation pressures.

Logically, for such a “family” of curves, as inflation pressure “Pi” increases the
tire/road contact area “Ac” will decrease, and therefore the peak lateral traction values
(“Fymax”) will decrease. Also, the seeming slopes “ΔFy/ΔN” of the “near-linear” portions of
the family of the parabolic “Fy = bN – mN2” plot lines will increase (grow steeper). An over-
plot of lateral tire traction behavior for various inflation pressure values seems to bear these
conjectures out:

273
SAWE Paper No. 3602
Category Number 31.0

Figure F.15 – TIRE LATERAL TRACTION FORCE “Fy” vs. NORMAL LOAD “N”
for VARIOUS INFLATION PRESSURES “Pi”

Application of a lateral force on a tire not only causes a distortion of the tire carcass
diminishing the tire/road area contact, but there are other effects as well. When a tire moving
with velocity “Vo” encounters a side load there is a consequent sideways motion. The
resultant new net motion “V” is the combination of the original motion “Vo” and the motion
resulting from the side load. The angle “ψ” between this new direction “V” and the original
direction “Vo” is called the “slip angle”. This term is a misnomer as it gives an erroneous
impression 387; the tire is not necessarily slipping or sliding in the direction of the side load.
What is actually happening is that there are a series of small lateral movements “dy” of the tire
due to the cyclical distortion of portions the carcass as those portions come into contact with
the road as the tire rolls forward. The combination of the original forward rolling velocity and
all those infinitesimal “side steps” results in the new velocity direction “V”.

In the following illustration of the situation note that the lateral sheer stress
distribution is such that the resultant lateral traction force “Fy” acts at a distance “e”, called
the “pneumatic trail”, aft of the tire center “o”. This offset of the lateral force from the tire
center results in a moment “Mz”, called the “self-aligning torque”, about that center. The

387
Reference [4], pg. 73. Famed automotive engineer and author (Reference [4]) Donald Bastow has suggested
that a better term for “ψ” might be “drift angle”, an idea this author has enthusiastically endorsed but admittedly
only sporadically adhered to.

274
SAWE Paper No. 3602
Category Number 31.0

generation of the angle “ψ” and the moment “Mz” are two of the most significant aspects of
tire lateral behavior.

Figure F.16 – DIRECTIONAL DISTURBANCE PRODUCES “dy”, “V”, “ψ”, “Fy”, &
“Mz”

This scenario of a change in direction due a side load, possibly a wind gust or an
inertial load due to an uneven road, may be called the “directional disturbance” scenario. The
incidence of a disturbance force causing tire carcass distortion, “drift angle”, and consequent
change in direction represents only one particular order of “cause and effect”.

There is another “cause and effect” scenario wherein a “drift angle” is intentionally
created resulting in a traction force inflicting a turning motion on the vehicle; this scenario is
called the “turning scenario”. Unlike the random happenstance of directional disturbance, the
“turning scenario” begins with the intentional act of a driver cranking over the steering wheel,
thereby rotating the front tires through the steering angle “δ”. As the tires rotate through the
angle “δ” the portion of the tire in contact with the ground becomes distorted giving rise to a
lateral shear force distribution of which the resultant centripetal force “F” tends to drive the
tire towards the center “TC” of a turn of radius “R”.

The combination of the original direction “Vo” and this new direction along the turn
radius “R” produces a new resultant direction “V” operating at some “slip angle” “ψ” to the
longitudinal axis (apparent direction) of the tire. The operation of the centripetal force “F”
generates an equal and opposite centrifugal virtual “side load” force in accord with the
concept of dynamic equilibrium; this side load force is proportional to the mass supported by
the tire times the radial acceleration: “m V2/R”. Because “F” must lie along the turn radius

275
SAWE Paper No. 3602
Category Number 31.0

(just as “V” must be orthogonal to the radius) it is not truly lateral in orientation with respect
to the tire; it may be resolved into two components: “Fx” and “Fy”.

The “Fy” is the true lateral traction force, while “Fx” may be referred to as the “turning
drag force”. The existence of this drag force is why, even without any braking or easing up on
the throttle, there will be some loss of velocity for a vehicle undergoing a turning action.
Lastly it should be noted that the action of the “self-aligning torque” “Mz” is such that in this
case the tendency is to return the steered wheel back to its original direction “Vo”; this“self-
aligning torque” is an important factor in vehicle steering and control 388.

Figure F.17 – STEERING PRODUCES “δ”, “ψ”, “V”, “Fx”, “Fy”, & “Mz”

With regard to these two scenarios of lateral tire behavior the observation has often
been made as to how it is difficult to discern between cause and effect: does the lateral force
cause the formation of a slip angle, or does the formation of a slip angle lead to the creation of
a lateral force? It is hoped that the previous careful expositions regarding the two possible
scenarios will clarify the matter, but it doesn’t really matter 389. Regardless of whether the

388
The “self-aligning torque” varies with “ψ” (proportional to “Fy”) and “N”, and therefore typically has units of
approximately 0.033 ft-lb/lb/deg for bias tires, and 0.043 ft-lb/lb/deg for radial tires (Reference [28], pg. 360).
389
Reference [52], pg. 19: “…lateral…force may be thought of as the result of slip angle, or the slip angle as
the result of lateral force…”.

276
SAWE Paper No. 3602
Category Number 31.0

force creates the acceleration, or the acceleration creates the force, it is sufficient to know the
relation “F = m a”; the rest is philosophy.

However, it is of importance in constructing computer simulations of dynamic effects


to not only have a clear definition of cause and effect, but to have some estimation of any time
lag between the two: the time lag between when the driver cranks the front tires over to the
steering angle “δ” and when the tires settle into the steady state values of “ψ”, “V”, “F”, and
“Mz” is generally equal to the time it takes for the tire to undergo a rotation of about 180 (π
radians) to 360 degrees (2π radians) 390.

Since the drift angle/lateral force relationship is dependent upon quite a few
parameters (inflation pressure, normal load, etc.), it is common to look at functions which
constitute only a partial differential of the total relationship (for which no one has yet
established a complete definitive formulation based on physics 391) in order to achieve a
degree of understanding. If the tire drift-angle/lateral-force partial differential function is
plotted the result looks like Figure F.18:

Figure F.18 – TIRE LATERAL FORCE AND DRIFT ANGLE

The lateral force increases with drift angle (or vice versa) on a shallow curve,
reminiscent of stress-strain curves, up to a deformation limit. At that point the situation is very

390
Reference [28], pg. 349.
391
Hans Bastiaan Pacjeka, Professor Emeritus at Delft University of Technology in the Netherlands, has
developed a tire model called the “Magic Formula” because it is based on relatively little underlying physics;
instead it is mainly based on a regression analysis of reams of empirical tire data.

277
SAWE Paper No. 3602
Category Number 31.0

unstable and will either quickly return to the safety of the “useful traction region” or result in
100% slip at the tire/road contact patch (a “skid”).

If the appropriate portion of the useful region is linearized by the fitting of a straight
line to the original shallow curve, then a simplified lateral force/drift angle relationship (the
slope of the fitted line) can be obtained. This quantity is commonly called the “Cornering
Stiffness” and symbolized as “Cs” (sometimes it is also called the “Cornering Power” of the
tire, which is yet another misnomer like “shock absorber” or “slip angle”). Linearized tire
relationship values such as “Cs” are often used in directional stability determinations and for
various simulations and studies of automotive dynamic behavior 392.

The problem with all such simulations and studies based on linear models is that tire
behavior beyond the proportional limit is not taken into account. Although road going
passenger vehicles seldom venture into the nonlinear region between proportional limit and
deformation limit, racing vehicles must always do so. A simulation/study of race car behavior
would seem to require a non-linear curve fit to the entire “useful traction” (both “linear” and
“transition” portions) region 393. For passenger car tires the proportional limit tends to
correspond to drift angles of around 12 to 13 degrees; race car tires may go higher. The great
skill of racing drivers is the ability to “walk the tight rope” of the transition region without
“falling off” (going into a skid).

If matters were just as simple as Figure F.18 then understanding of tire behavior would
be very easy. However, the potential or maximum lateral force that a tire can supply, and the
drift angle associated with that force, is dependent on many parameters. The lateral force
potential is primarily influenced by394 normal load, camber angle (which can change with
roll), roll steer (which can be the result of normal load and camber change with roll, but toe
in/out can also change with roll), tire type (size, carcass type and material, rubber type, tread
design, aspect ratio), inflation pressure, wheel rim width, road material and surface (smooth,
rough, dusty, etc.), weather (rain, snow, ice), temperature (road surface, ambient, and of the
tire itself), and the speed of the vehicle (all basic tire coefficients of traction are somewhat
speed dependent; the same lateral force will produce a smaller drift angle at high speed than
at low speed 395). While all these factors are significant, only tire type, normal load, inflation

392
Reference [28], pp. 198 & 350 states that “Cs” (a.k.a. “C∝”) is the “…slope of the curve at zero slip angle…”
(even though that would seem to be a poor fit). Although cars turn left and right, implying positive and negative
“Cs” values, by SAE convention “Cs” is always positive. “Cs” has units such as “lb/deg”, but when normalized
by the normal load “N” the “Cornering Stiffness” becomes the “Cornering Coefficient” “Cc” in units of
“lb/deg/lb”.
393
Reference [52], page 126.
394
That the lateral traction force potential is influenced by longitudinal forces is a subject yet to be dealt with,
but one that would immediately be anticipated by Poisson.
395
Reference [12], page 164, exact quote: “…cornering power (sic) increases slightly with speed...”.

278
SAWE Paper No. 3602
Category Number 31.0

pressure, temperature, and speed are fundamental tire behavior and will be discussed herein;
the rest has to do with tire/suspension interaction.

To illustrate the effect of normal load on lateral resistance and drift angle for a specific
tire and inflation pressure a plot such as Figure F.19 may be used. Note that it is essentially
like Figure F.18 except that there are now a large set of “Fy, ψ” functions which serve to
represent an infinite variation; any change in normal load alters the “Fy, ψ” relation, but these
five example curves may suffice as the intermediate possibilities can be approximated by
interpolation:

Figure F.19 – LAT TRACTION vs. DRIFT ANGLE, VARIOUS NORMAL LOADS

Only the useful traction region is shown; as noted, 13 degrees is about the maximum
drift angle/deformation limit for passenger car tires, which in normal driving seldom go
beyond 5 degrees or so. Note that as normal load increases the lateral traction force necessary
for a certain drift angle increases as well, but at a decreasing rate just as was the case for
Equation F.16/Figure F.14. Actually, Figure F.19 is a poor way to illustrate this behavior,
which is better shown by the following actual data plot of lateral force vs. normal load for a
6.00×16 tire inflated to 28 psi (193 kPa) 396:

396
Reference [12], page 165.

279
SAWE Paper No. 3602
Category Number 31.0

Figure F.20 – LATERAL TRACTION vs. NORMAL LOAD AT VARIOUS DRIFT


ANGLES
Figure F.20 clearly shows how increasing normal load will increase the lateral force
necessary to cause the same amount of deformation (drift angle), but at a decreasing rate and
only up to a point; traction demands beyond that point are likely to result in an out-of-control
skid of the vehicle 397. Another way of looking at tire lateral traction force behavior is
presented in Figure F.21. This figure is very much like Figure F.19, only now the lateral
traction force “Fy” is normalized by “dividing out” the now constant normal load “N”, which
of course results in the effective lateral traction coefficient “μy” (“μy = Fy/N”).

397
Reference [4], pg. 76.

280
SAWE Paper No. 3602
Category Number 31.0

Also, while Figure F.19 was limited to Quadrant I, Figure F.21 is in all four quadrants,
although only Quadrant I (positive “ψ”, positive “μ”) and Quadrant II (negative “ψ”, negative
“μ”) 398 have any physical significance here. Although the normal load “N” is constant, there
is a plot variation due to the effect of a varying coincident longitudinal traction force,
represented by its corresponding slip values. The total effect is as if the “N = 1000” plot
line/tire of Figure F.19 were subjected to varying longitudinal load and the results replotted
with normalization and for both left and right turns:

Figure F.21 – NORMALIZED LAT TRACTION vs. DRIFT ANGLE, VARIOUS SLIPS

The total effect is one of generalization and broadening of content. The normalization
results in a value of “μ” that will be of value over a short range of “N” variation; and the
extension into four quadrants shows that tire behavior is the same for both right and left
turns 399. The only really new information provided is the indication of the effect of
simultaneous longitudinal force generation upon lateral traction force potential; as slip (and
therefore longitudinal traction force) increases the peak lateral traction coefficient (and hence

398
This would seem to be indicative of left-hand and right-hand turns respectively. The extent to which Figure
F.21 is simplified with respect to reality can be gauged by comparison of Quadrant I with Figure 18 of Reference
[34], page 90.
399
However, some tires of asymmetrical construction might prove an exception to this statement. Also, a tire’s
behavior on a test rig may be symmetrical, but when installed on a vehicle may perform differently left vs. right
hand turns due to suspension effects (camber, etc.). Figure F.21 is the result of data generated by a computerized
tire model; to see a plot resulting from actual empirical data refer to Reference [34], pg. 90.

281
SAWE Paper No. 3602
Category Number 31.0

the max lateral traction force available) decreases. Furthermore, if a certain level of lateral
traction force is attained, and then the slip (longitudinal traction force) is increased, the lateral
traction force will drop unless the slip angle is sufficiently increased.

TIRE TRACTION: LONGITUDINAL

The “traditional” tire lateral traction model accounts for change in normal load and
change in contact area (“curl up”) effect on the lateral traction coefficient, but the
“traditional” way of dealing with the longitudinal coefficient of traction has often been simply
to choose some seemingly appropriate constant value and make do with that. However, since
Equation F.10 gives the traction coefficient variation with contact pressure, it would seem that
the only info needed to relate the longitudinal coefficient of traction “μx” to normal load “N”
is the Dixon Equation relating contact area “Ac” to normal load which is:

𝑨𝒄 = 𝑳𝒄 𝒕𝒘 = 𝟏. 𝟐𝟒 𝑹𝒊 𝑪𝒐𝒔−𝟏 [(𝑹𝒊 − 𝒅)/𝑹𝒊 ] 𝒕𝒘 (EQ. F.17)


For this area determination the “d”” values need to be estimated by the Nokian
Formula (bringing “N” into play):

𝒅 = 𝑵�𝑲 + 𝒅𝟎 (EQ. F.18)


𝒁

And the tread width “tw” values need to be estimated by the Michelin Formula:

𝒕𝒘 = 𝟎. 𝟎𝟑𝟗𝟑𝟕 (−𝟎. 𝟎𝟎𝟒 + 𝟏. 𝟎𝟑)𝑺𝑵 (EQ. F.19)


This leaves only the matter of the tire vertical spring constant “KV” which can be
estimated by the Rhynes Equation:

𝑺𝑵 ×
𝑲𝑽 = 𝟎. 𝟎𝟎𝟎𝟐𝟖 𝑷𝒊 ��−𝟎. 𝟎𝟎𝟒 + 𝟏. 𝟎𝟑�𝑺𝑵 × � 𝟓𝟎
+ 𝑫𝑹 � + 𝟑. 𝟒𝟓 (EQ. F.20)

All the parametric information necessary to plug into these four equations for
determining a specific tire’s “contact area = f(normal load)” is contained in a tire’s “P-
Metric” designation as inscribed on the sidewall. Determination of the contact area “Ac”
under normal load “N” provides the contact pressure “Pc” (= “N/Ac”) so that the peak
longitudinal coefficient of traction can be obtained by the Koutný Formula:

μx = a Pcn (EQ. F.21)

282
SAWE Paper No. 3602
Category Number 31.0

In order to generate a realistic contact pressure variation the following exposition will
utilize an example tire of designation “P152/92R16”, “a” will be set to “15.7369” (English
psi units, for Metric kPa units use “58.2587”), and “n” will be set to “-0.67791” (a Koutný
value, presumed typical):

μx = 15.7369 (Pc)-0.67791 (EQ. F.22)

There are a number of somewhat rough “rules of thumb” that also attempt to define
the relationship between the peak longitudinal coefficient of traction and the normal load;
these may be useful in comparison with Equation F.21/22. One of these “rules of thumb” is
given by Prof. Gillespie: “…as load increases the peak and slide (traction) forces do not
increase proportionately…in the vicinity of a tire’s rated load 400… (the traction)
coefficients will decrease…0.01 for each 10% increase in load 401”. There is a similar “rule
of thumb” attributed to Formula 1 competitors (source unknown) which may be paraphrased
as: “…for every +5.82% increase in contact pressure there will be a -1.00% decrease in the
traction coefficient…”. The variation of the peak longitudinal coefficient of traction with
normal contact pressure as per the Koutný model (Equation F.21) and the “rules of thumb”
may be graphically presented as follows:

Figure F.22 – TIRE LONGITUDINAL TRACTION COEF vs. NORMAL LOAD

400
That is, within the working range.
401
Reference [28], pg. 57 and pg. 344. See also Figure 10.8 (pg. 345) of the reference, and compare with above.

283
SAWE Paper No. 3602
Category Number 31.0

Unlike the lateral traction coefficient model of Figure F.09 there are a number of
longitudinal traction coefficient models presented here, although only the Koutný model is of
any real concern in this paper. The first (upper) model is the constant longitudinal traction
coefficient model which, while totally unrepresentative of reality, is used fairly often for
vehicle acceleration/braking performance calculations. The “accuracy” of such calculations is
totally dependent on the shrewdness with which the traction coefficient value used is chosen.

The second model shown is the “Gillespie” model, which does show decreasing
coefficient with increasing normal load, but does so in a linear fashion and does not come
close to reality in magnitude 402. The third model, the one of unknown origin but associated
with Formula 1 racing, is near-linear and seems to represent an improvement; however, it is
still insufficient if the Koutný model is regarded as the “gold standard”.

The last two models utilize the Koutný Formula, but the plot labeled “Koutný Const
Area” shows how the traction coefficient would vary if the tire/road contact area stayed
constant as the normal load on the tire varied. That is reasonable for a simple rubber material
sample (see Figure F.07) but is not realistic for a tire; constant tire/road contact area results in
overvalued contact pressures with a consequent undervaluing of the longitudinal traction
coefficient. If Equations F.17-F.20 are used to determine the increasing contact area “Ac” as
normal load “N” increases, then the use of the resulting correct contact pressure “Pc” (equal to
“N/Ac”) in the Koutný Formula gives the most accurate variation of the longitudinal traction
coefficient as per the plot simply labeled “Koutný”.

When the “Koutný” longitudinal traction coefficient equation is “plugged” into the
Coulomb friction equation, the result is:

Fx = (a Pcn)N (EQ. F.23)

If the variation in “Ac”, and hence the variation in “Pc”, is correctly accounted for
(using Equations F.17 to F.20), then when plotted this equation presents a visual illustration
of how tire longitudinal traction force potential varies with normal load (for “a” equal to
“15.7364” and “n” equal to “-0.67791”):

402
Reference [28], pg. 345, Fig. 10.8. Prof. Gillespie illustrates his concept regarding the variation with normal
load of both the static “μs” (rolling) and dynamic “μd” (skidding) tire traction coefficients as straight parallel
lines, decreasing with increasing load, and with “μs > μd”.

284
SAWE Paper No. 3602
Category Number 31.0

Figure F.23 – TIRE LONG TRACTION FORCE “Fx” vs. NORMAL LOAD “N”
As was the case with the lateral traction force potential, the longitudinal traction force
potential also varies with inflation pressure, and will form a family of curves for a particular
tire. Furthermore, just as the increasing lateral traction force meant a corresponding increasing
tire deformation “ψ” commonly called the “slip angle”, increasing longitudinal traction force
means a corresponding increasing tire deformation “%S” (or “S”) called “percent slip” (or
just “slip”). And, just as was the case with “slip angle”, “slip” is a misnomer, but no
reasonable replacement has been advanced. The apparent “slip” of a rolling tire transmitting a
longitudinal traction force is due to a cyclical stretching or compression of “portions” of the
tire tread as those “portions” roll into contact with the road; when those “portions” roll out of
contact the longitudinal stress is relieved and an expansion or contraction back to original
length occurs (along with a release of energy as heat due to hysteresis).

285
SAWE Paper No. 3602
Category Number 31.0

Figure F.24 – TIRE LONGITUDINAL TRACTION FORCE “Fx” vs. NORMAL LOAD
“N” for VARIOUS INFLATION PRESSURES “Pi”

The tire deformation that accompanies longitudinal traction loadings may be


illustrated as per the following figure:

Figure F.25 – TRACTION AFFECTS “Fx”, “Fy”, “d”, “α”, “ω”, “a”, “V”

286
SAWE Paper No. 3602
Category Number 31.0

In Figure F.25, for the conditions of acceleration and braking, the vertical force “Fz”
(the resultant of the contact area vertical pressure distribution and equal to “N”) times the
offset arm “d” constitutes the rolling resistance 403. The presence of longitudinal traction
forces for acceleration (“Fx”) and braking (“-Fx”) 404 both increase rolling resistance, but not
to the exact same extent. Acceleration and braking generate different longitudinal shear stress
distributions (compression vs. tension), which interact with the vertical contact stress
distributions, which affects the rolling resistance.

Figure F.25a – ACCEL & BRAKE EFFECT ON ROLLING RESTANCE


As noted, as each tire tread segment rolls into contact with the ground, there is a
longitudinal stretching or compression of that segment, followed by a contraction or
expansion as the segment rolls up out of contact. It is this cyclical distortion of the tread that
gives the appearance of “slip”, which is to say the speed of rotation of the tire “ω” seems out
of synch with the velocity “V”. “Slip” may be represented as “%S”, or just “S”. The tire
segment in contact with the road and under traction stress is generally not in motion with
respect to the road (although some portions of the contact area may be); on the whole the tire
contact area may be regarded as actually being “static” with respect to the road.
The apparent “slip” in percent may be mathematically defined as 405:

%𝑺 = �𝑹𝝎�𝑽 − 𝟏� × 𝟏𝟎𝟎% (EQ. F.24)

403
Reference [93], pp. 18 and 25. Note that the vertical offset “d” is depicted as being smaller in acceleration.
404
SAE J670 2008 pg. 42 defines the acceleration force as positive and the braking force as negative.
405
Ibid, pg. 39. Full wheelspin “%S = ∞%”, free rolling “%S = 0%”, locked brakes “%S = -100%”.

287
SAWE Paper No. 3602
Category Number 31.0

For acceleration “V < Rω”, so Equation F.24 produces a positive number for “%S”.
For deceleration (braking) “V > Rω”, so Equation F.24 then produces a negative number for
“%S”. This variation in sign for acceleration and braking “slip” is in accord with the SAE
J670 2008 standard, although the matter is not clearly stated 406.
Just as the drift angle “ψ” was proportional to the lateral force “Fy”, the apparent slip
“%S” is proportional to the longitudinal force “Fx” the potential for which depends on the
normal load “N”. A representation of these relationships may look as per the following
figures. Figure F.27 shows the relationship between the longitudinal traction force “Fx” and
slip “%S” for some constant normal load “N”; note the similarity to the lateral traction force
“Fy” vs. drift angle “ψ” of Figure F.18.

Figure F.26 – TIRE LONGITUDINAL FORCE (Fx) AND SLIP (%S)


Figure F.27 shows how this relationship can vary for different normal loads. Figure
F.27 is based on a plot made by use of the BNP (Bakker-Nyborg-Pajecka) or “Magic” tire
model using parameters empirically obtained for a particular truck tire and inflation pressure;

406
Note that there are variations on the “slip” formulation depending on the use to which the result is to be put;
see Reference [52], pp. 39-41, and Reference [8], pg. 2. One of more common alternate formulations for “slip” is
such that “slip” is positive for both acceleration and deceleration.

288
SAWE Paper No. 3602
Category Number 31.0

for “N = 12350 lb” and “N = 3088 lb” the model results are over-plotted with some empirical
results to indicate the degree of model veracity 407. Note that the proportional limit is indicated
as being around 10% slip; normal roadway driving slip is generally under 3%.

Figure F.27 – LONGITUDINAL TRACTION vs. %SLIP AT NORMAL LOAD (N)


In a way analogous to the case of Figure F.20, where the lateral traction force “Fy”
variation with normal load “N” per drift angle “ψ” produced a family of curves, it should be
possible to plot longitudinal traction force “Fx” variation with normal load “N” per percent
slip “%S” to produce a similar family of curves. However, no such plot could be readily
found in the literature, possibly because the “%S = f(Fx, N)” function has been of less general
interest than “ψ = f(Fy, N)”. Therefore, the data inherent in Figure F.27 was replotted in an
attempt to construct such a figure, with the result being:

407
Reference [9], pg. 3.

289
SAWE Paper No. 3602
Category Number 31.0

Figure F.28 – LONGITUDINAL TRACTION vs. NORMAL LOAD AT VARIOUS


SLIPS (%S)

Figure F.28 would appear to be of limited usefulness due to the degree of overlap and
other issues, but at least it indicates that generally increasing normal load allows for the
generation of increasing longitudinal traction for the same percent of slip 408. Just as was done
earlier for the lateral traction force vs. drift angle functions at various normal loads, the
longitudinal counterpart to Figure F.19 can also be normalized and extended over four
quadrants:
408
This could be interpreted as saying that increased vertical stress increases longitudinal stiffness, which seems
to make sense.

290
SAWE Paper No. 3602
Category Number 31.0

Figure F.29 – NORMALIZED LONG TRACTION vs. SLIP, VARIOUS DRIFT


ANGLES (ψ)

Again an interaction between longitudinal and lateral (represented by drift angle “ψ”)
traction forces is indicated; to maintain a constant longitudinal traction as slip angle increases
there must be an increase in slip fraction (requiring more drive torque) 409. Also interesting is
the fact that the longitudinal force function is not identical for acceleration (positive “μx”,
positive “S”) and braking (negative “μx”, negative “S”) 410. This is due to the different contact
area stress distributions engendered by these two cases as discussed earlier. Otherwise, most
of the remarks made about lateral traction Figure F.21 are relevant here as well.

409
Reference [52], pg.42. The reference actually says “To reach peak traction/braking forces requires higher
and higher slip ratio as the slip angle is increased”, but the two statements are complementary in indicating the
influence of the lateral traction upon the longitudinal traction (and vice versa).
410
Admittedly, Figure F.29 doesn’t illustrate this fact very dramatically. Again, this is a plot of computer model
generated data, and is not the same model or input data that generated Figure F.27. Also, the sign convention is
exactly opposite that employed in Reference [52], pg. 42 (the reference uses data produced by a Prof. Sakai who
uses his own sign conventions).

291
SAWE Paper No. 3602
Category Number 31.0

TIRE TRACTION: LONGITUDINAL AND LATERAL TOGETHER

The question now presents itself: how do we “synchronize” the longitudinal and
lateral traction functions so that they coherently represent the same tire? This is not, or should
not, be a problem if all the necessary tire coefficients are properly determined by empirical
means for use in a unified tire model like the “Magic” or “LuGre” models, but it does become
a question if the attempt is made to construct a model for a theoretical tire using just the
simple relationships expounded on herein. Of course, the establishment of the maximum
longitudinal/lateral traction forces that a tire could generate for a particular inflation pressure,
normal load, etc., would go a long way toward that construction, but the result can only be
used for exposition and conceptual thinking. For realistic engineering determinations of what
performance levels a detailed design could achieve on a specific road course, only a model
such as the “Magic”, or perhaps the “LuGre”, can truly suffice 411.

This is because a tire may experience an essentially purely longitudinal or a purely


lateral traction loading under certain limited circumstances, such as a drag racing or skid pad
simulation; but generally a tire undergoes a simultaneous combination of lateral and
longitudinal loading. If the maximum loading a tire could undergo were equal in either
direction, then the maximum resultant combination of longitudinal and lateral traction forces
that a tire could generate would be obliged to fall within a “traction circle” 412, and for
simplicity’s sake the situation is often portrayed that way, but only for exposition. Because
tire traction behavior is anisotropic the situation is much more accurately modeled as a
“traction ellipse” 413:

When the lateral/longitudinal traction relationship is portrayed as a circle it allows for


some very simple determinations. For instance, note that the two orthogonal traction forces
“Fx” and “Fy” must always combine to form the resultant force “Fr” as per:

𝑭𝒓 = �𝑭𝒙 𝟐 + 𝑭𝒚 𝟐 (EQ. F.25)

When utilizing the circle model, this resultant force “Fr” can’t exceed the circle radius
or maximum traction force “R”, if a skid is not to set in. So, using this simple relation, if “R”

411
Of course, such tire models must be linked to an overall vehicle model such as might be constructed in a
multi-body dynamic simulation program such as ADAMS (ADAMS/Car).
412
Reference [13], pg. 21.
413
Reference [1], pg. 51.

292
SAWE Paper No. 3602
Category Number 31.0

is 560 lb (254.0 kg), and “Fy” is 300 lb (136.1 kg), then in order for the particular tire
considered to not go into a skid “Fx” can only go up to:

𝑭𝒙 = �𝑹𝟐 − 𝑭𝒚 𝟐 = √𝟓𝟔𝟎𝟐 − 𝟑𝟎𝟎𝟐 = 𝟒𝟕𝟐. 𝟗 𝒍𝒃, 𝒐𝒓 𝟐𝟏𝟒. 𝟓 𝒌𝒈

So simple, but far from realistic. In the quest to keep things as simple as possible, but
with a closer correspondence to reality, the ellipse model was developed. Although the ellipse
model is a step closer to a realistic longitudinal/lateral tire traction force distribution it is still
a gross simplification 414, but such a model is necessary for ease in exposition and conceptual
level calculations (any attempted use in a quantitative computer simulation would require
resorting to huge data sets and interpolation routines). An over-plot comparison of the circle
and ellipse models for the same tire would look as follows:

Figure F.30 – TIRE TRACTION CIRCLE AND ELLIPSE MODELS

414
Reference [9], pg. 1.

293
SAWE Paper No. 3602
Category Number 31.0

Ellipses have many wonderful properties, and have long been a subject of interest for
mathematicians and astronomers such as Menaechmus (c.380-c.320 BC), Euclid (c.325-c.265
BC), Apollonius (c.262-c.190 BC), Pappus (c.290-c.350 AD), Kepler (1571-1630 AD),
Newton (1642-1727 AD), and Halley (1656-1742 AD) 415. The major axis of the ellipse is
“2a” in length, while the minor axis is “2b”. In this tire traction model the “a” corresponds to
the maximum longitudinal traction available, and the “b” corresponds to the maximum lateral
traction available. Note that the ellipse properly represents the passenger car tire relation
between maximum longitudinal and lateral traction forces in that “a > b” 416:

a = Fxmax (EQ. F.26)

b = Fymax (EQ. F.27)

So, with “a” and “b” quantified, a property called the “eccentricity” of the ellipse can
be calculated:

𝒃 𝟐
𝒆 = �𝟏 − � � (EQ. F.28)
𝒂

The eccentricity “e” represents the ratio of the “c” dimension, which is the distance of
the “foci” from the center (origin) divided by the “a” dimension (one half the major axis):

𝒄
𝒆= (EQ. F.29)
𝒂

Given the above, it is now possible to find the location of the “foci” in our tire traction
ellipse. This is very important historically because the focus points were used to generate an

415
Generally, when using an ellipse to model the solar system, an astronomer places the sun at one of the foci,
making that focus the origin; luckily our purpose is best suited by placing the origin at the center.
416
Reference [93], pg.51. Reference [9], pg. 1, states that “racing…lateral vehicle accelerations can exceed
longitudinal accelerations” which means that “Fymax” can exceed “Fxmax” for at least some racing tire types.
Road car tire lateral traction coefficients go as high as 1.1 (2008), and Reference [52] on page 27 says that
Formula 1 lateral traction coefficients can go as high as 1.8 (1995).

294
SAWE Paper No. 3602
Category Number 31.0

ellipse by means of a string of which the ends were anchored at those points, and even though
such a “trammel construction” might not be used today the foci are still key points in ellipse
definition. Technically, a circle is also an ellipse, but an ellipse whose foci are both at the
center. That is to say, for a circle the “c” dimension will be zero, “e” will be zero, and “a” will
equal “b”.

The significance of so thoroughly defining the tire traction ellipse lies in the fact that
no combination of longitudinal and lateral tire traction forces, i.e. no resultant traction force,
should be so great that if plotted to scale it would project beyond the ellipse periphery. The
need for any traction force so great that it would tend to fall on the ellipse periphery is
indicative of a condition of impending skid. Any point on the periphery of an ellipse centered
on the origin with a major axis in the “X” direction must conform to the ellipse equation:

𝑭𝒙 𝟐 𝑭𝒚 𝟐
+ =𝟏 (EQ. F.30)
𝒂𝟐 𝒃𝟐

So, if “Fxmax” (“a”) = 629.4 lb (285.5 kg) and “Fymax” (“b”) = 498.5 lb (226.1 kg, all
scaled from Figure F.27), then when “Fy” = 300.0 lb (136.1 kg) the maximum amount of “Fx”
tolerated before a skid would ensue is:

𝑭𝒚 𝟐 𝟑𝟎𝟎. 𝟎𝟐

𝑭𝒙 = 𝒂 �𝟏 − 𝟐 � = �𝟔𝟐𝟗. 𝟒 �𝟏 −
𝟐 𝟐 � = 𝟓𝟎𝟐. 𝟕 𝒍𝒃, 𝒐𝒓 𝟐𝟐𝟖. 𝟎 𝒌𝒈
𝒃 𝟒𝟗𝟖. 𝟓𝟐

Because of the change in traction models, a tire that could only deliver 472.9 lb (214.5
kg) of braking force, while supplying 300.0 lb (136.1 kg) lateral force in a right hand turn, is
now credited with being able to deliver 502.7 lb (228.0 kg) braking force in the same turn.
Still pretty simple, but now closer to a quantitative reality…

Note that a traction ellipse such as Figure F.30 holds only for a particular tire on a
particular surface at specific normal load, inflation pressure, velocity, and temperature. These
limitations can be countered somewhat with by normalizing the longitudinal/lateral forces,
assuming constant inflation pressure and temperature (due to the attainment of thermal

295
SAWE Paper No. 3602
Category Number 31.0

equilibrium at operating temperature), and by putting the “secondary” matter of velocity


effect aside for the moment. Given those circumscriptions, a reasonably useful and realistic
representation of combined longitudinal/lateral tire behavior for a specific tire may be
constructed as per the following example 417:

Figure F.31 – TRACTION ELLIPSE WITH LINES OF CONSTANT “ψ” AND


CONSTANT “%S”

Although this traction ellipse is truly valid only for a normal load “N” of 661 lb (300
kg), the “normalization” of the “Fx” and “Fy” axes (“μx = Fx / N”, “μy = Fy / N”) allows for
this same graph to be used with reasonable accuracy over a short interval of normal load

417
Reference [28], pg. 351 states that a negative steering force is to the left, but that is probably in accord with
the “old’ SAE 670e “Z-down” standard. Figure F.31 indicates that lateral force behavior may be somewhat
different (higher for same slip angle) for braking than for acceleration, which per Reference [93] (pg. 49) is true,
especially for bias tires. Figure F.31 is an adaptation of a figure found in Reference [66] on page 139.

296
SAWE Paper No. 3602
Category Number 31.0

variation. Note that lines of constant drift angle have been superimposed over the traction
ellipse as is a common practice; this increases the usefulness of the plot. Constant “slip” value
lines have been superimposed as well, but that is less frequently done 418.

A three dimensional plot of this specific tire’s traction potential from “N = 0” to its
absolute maximum load capacity “N = Lcap” at its absolute maximum inflation pressure
“Pcap” could be constructed which would allow for use with reasonable accuracy over a long
interval of normal load variation. If so, then the result would look something like a
hemispheroid (or a grapefruit serving) as shown in Figure F.32:

Figure F.32 – TRACTION ELLIPSE IN THREE DIMENSIONS

If enough data were available for a specific tire to construct a volume as shown for “N
= 0” to “N = Lcap” for each “Pi” increment of, say, 5 psi throughout the working pressure
range, then this perhaps would constitute enough data for utilization in a full quantitative
automotive performance simulation. Such a simulation would have to contain interpolation
routines for the determination of appropriate data values that lie between the known “N” data

418
Reference [93], pp. 52-56 addresses this matter in the course of illustrating traction ellipse construction from
empirical data. Reference [34], pg. 89 explains this as “…an attempt to plot Fy versus Fx for constant values of
Sx would result in a less regular set of curves as a result of the greater imprecision encountered in measuring
Sx in comparison to Sy…”.

297
SAWE Paper No. 3602
Category Number 31.0

levels, and further routines to modify that interpolated data to account for the effects of
parameters such as temperature, velocity, and camber. All things considered, that’s usually
enough to make utilization of the “Magic” or other such complex formulations look
attractive 419.

The sort of information contained in tire data plots such as Figure F.31 and Figure
F.32 is always useful if one needs to know how the vertical/longitudinal/lateral forces interact,
i.e.-what the resultant values of “%S” and “ψ” will be given a certain set of those forces.
However, if one’s concern is on a less sophisticated level, i.e. - “is there enough traction to
complete a particular maneuver?”, then reference to an appropriate tire traction ellipse may
not be necessary; if the summation of the “����⃗
𝑭𝒙 ” and “����⃗
𝑭𝒚 ” force vectors (“�𝑭𝒙 𝟐 + 𝑭𝒚 𝟐” when the
forces are expressed as scalars) is less than “0.3 N” in value then the answer is “yes”. It’s only
when a vehicle is being driven hard, almost at the point of loss of control (skid), that the tire
traction ellipse data becomes extremely significant. Thus the significance of the ellipse for
racing vehicles, which tend to spend a lot of time at high “g” levels, and the general lack of
significance for normal “go to work, take the kids to school” driving, which rarely exceeds
that “0.3 N” level (which means that acceleration levels greater than 0.3 G’s are seldom
incurred) 420.

TEMPERATURE EFFECTS

The effect of temperature is often ignored; either it is considered inconsequential or a


benign condition of thermal equilibrium is assumed. However, there are cases when the blind
discounting of temperature can be disastrous. The variation with temperature of material
properties, mostly the properties of rubber, causes significant variation in tire behavior.
Traction, rolling resistance, and inflation pressure are all affected, which in turn causes other
effects (slip, slip angle, cornering stiffness, fuel economy, wear, vertical spring constant, etc.)

Tire temperature and inflation pressure are very closely interrelated as per the Ideal
Gas Law:

PV = n R T (EQ. F.31)

419
Useful simulations of transient maneuvering behavior have been accomplished without recourse to such
complex tire models; consider Reference [1]. This fine example of an automotive simulation written expressly
for microcomputer utilization can be obtained as a SAE Paper (870495), or as included in the SAE Publication
(SP-699) Automotive Crash Avoidance Research.
420
Reference [34], pg. 69.

298
SAWE Paper No. 3602
Category Number 31.0

Where gas pressure “P” (in “atmospheres”) and volume “V’ (in liters) are related to
the gas temperature “T” by the factors “n” (the amount of gas in “moles”) and “R” (the
Universal Gas Constant: 0.08207 liter-atm/mole-oK). Ignoring the small changes in tire
volume “V” with inflation pressure “P” (“V” is constant) means that “P = (nR/V) T” where
“nR/V” is a constant; tire pressure varies in a direct linear relation with tire temperature:

Figure F.33 - TIRE INFLATION PRESSURE vs. TEMPERATURE

And, since the tire vertical spring constant “Kv” varies with pressure in accord with
Equation F.20, the deflection under load will vary per Equation F.18, which in turn affects the
tire-road contact area per Equation F.17. Therefore, increased “T” means increased “Pi” and
“Kv”, which in turn leads to decreased “d” and “Ac”. And that ultimately means decreased
rolling resistance, which means improved fuel economy, but also less traction …

However, the tire mechanical effects caused by temperature variation aren’t the whole
story; there are material effects as well. Rubber energy dissipation (hysteresis) and traction
coefficient varies directly with temperature in a very non-linear fashion 421:

421
Reference [26], pg. 491. Current (2014) Pirelli Formula 1 tires are kept warm in electric blankets at a
temperature of 230 oF (110 oC) before the start of a race (the FIA may ban tire pre-race warming by 2015).

299
SAWE Paper No. 3602
Category Number 31.0

Figure F.34 – RUBBER ENERGY DISSIPATION & TRACTION vs.


TEMPERATURE

The total effect will be a combination of the mechanical and the material, which leads
to somewhat puzzling situations such as 422:

“…increase in energy dissipation that accompanies an


increased load causes the temperature of the tire to
rise…results in lower hysteretic loss coefficient…as a result
the coefficient of rolling resistance often decreases somewhat
with increasing load…”

There are forms of automotive endeavor in which temperature levels play a


significant, and complicated, role. For instance, Formula 1 and Indy car tires require operation
within a narrow temperature band for optimum performance; per an authoritative source 423:

“Modern race tire compounds have an optimum temperature for


maximum grip. If too cold, the tires are very slippery; if too hot
the tread rubber will “melt”; in between is the correct
temperature for operation.”
422
Ibid. Note that it says “the coefficient of rolling resistance”, not “the rolling resistance”, decreases with
load. The total effect of increasing load is one of increased rolling resistance, but at a decreasing rate of increase.
423
Reference [52], pg. 56.

300
SAWE Paper No. 3602
Category Number 31.0
Rev. C

For racing tires the increase in temperature with velocity and hard use is planned for,
and if laps have to be run at reduced speed under a safety flag, or during a rolling start, then it
is not unusual to see the cars alternately darting hard right and left as the drivers try to keep
their tires at optimum temperature as they wait for all out racing to recommence. Possibly one
of the worst scenarios that can occur in racing is to be caught with rain tires installed as the
track starts to dry out and there is no chance of a pit stop for tire replacement; the “softer”
compound rain tires are certain to overheat unless the driver commences driving far less
aggressively, which is a tactic not likely to place him on the podium.

INFLATION PRESSURE EFFECTS

Changes in tire inflation pressure will affect not just the vertical stiffness, as per
Equation F.20, but other measures of tire stiffness as well, such as the “Cornering
Stiffness” 424 or the lateral traction coefficient “m” (which is an inverse measure of stiffness).
As noted, this causes a cascade of other changes; changes in vertical stiffness will affect
vertical deflection under load (and therefore rolling radius 425) per Equation F.18, which in
turn will affect tire-ground contact area (and therefore traction) per Equation F.17, and that
leads to changes in rolling resistance (and therefore fuel economy), heat generation, and
temperature. And, of course, this tends to run in a full cycle, as temperature will, in turn,
affect the inflation pressure. Some graphical illustrations of inflation pressure effects are as
follows:

Figure F.35 – STUTTGART FORMULA ROLL RESIST COEFS vs. INFLATE PRESS

424
Reference [28], pg.354: “…increasing inflation pressure results in increasing cornering stiffness for
passenger car tires…tires at reduced inflation pressures arrive at lateral force saturation at substantially
higher values of slip angle…”
425
The pressure/deflection effect on rolling radius is large compared to a pressure/strain effect which is often
neglected, but which will be considered later.

301
SAWE Paper No. 3602
Category Number 31.0

These particular plots of the rolling resistance coefficients versus inflation pressure are
applicable only for a specific tire, but may be considered typical in their general variation, at
least for circa 1957 (or possibly earlier) bias ply type tire rolling on concrete such as the one
from which the plots were presumably developed 426.

The variation in tire-road contact area versus inflation pressure (note that without
temperature or normal load change the longitudinal tire traction may be considered as varying
in exact proportion to the change in area) may be illustrated as:

Figure F.36 – TIRE-ROAD CONTACT AREA vs. INFLATION PRESSURE

426
Reference [81], pg. 32. Reference [28], pg. 118. The latter reference makes the statement that the coefficients
were developed for tires “on a concrete surface”.

302
SAWE Paper No. 3602
Category Number 31.0

Increasing inflation pressure will also cause some expansion of the tire circumference,
although such expansion usually is very small 427. The situation may be depicted as in Figure
F.37:

As the inflation pressure “P”


increases the force “F” pushing the tire
semi-sections apart; the force is equal
to the pressure times the horizontal
plane area: “F = P A”. Expressed in
differential form this relation may be
expressed as:

𝒅𝑭 = 𝒅𝑷 𝑨

This causes corresponding


stress “dS” and strain “dε” differentials
in the tire periphery (tread):

Fig. F.37 – INFLATION PRESSURE 𝒅𝑺 = 𝑬 𝒅𝜺


EXPANSION

By definition “dL/L” may be substituted for “dε”, and the expression rearranged:

𝒅𝑳 = 𝑳 𝒅𝑺⁄𝑬

The stress “dS” is equal to “2 dF/2” (“dF”) over the tire tread cross-sectional areas “2
tt tw”, or “dS = dF/2 tt tw”, which allows for the following substitution and simplification:

𝒅𝑳 = 𝑳 (𝒅𝑭 / 𝟐 𝒕𝒕 𝒕𝒘 ) / 𝑬 = 𝑳 𝒅𝑭 / 𝟐 𝒕𝒕 𝒕𝒘 𝑬

Remember that “dF = dP A”, and note that “A = 2 R tw”; this allows for the following
substitution and simplification:

𝒅𝑳 = 𝑳 𝒅𝑷 𝑹 / 𝒕𝒕 𝑬

Since the tire circumference “C” (“2 π R”) is equal to “2 L”, the relation of the tire
radius “R” to “L” is “2 π R = 2 L” or “π R = L”. Therefore “L = π R” and “dL = π dR”;
substitute for “L” and “dL”:

427
This is a true expansion, not a decrease in the deflection under load. Increasing inflation pressure will increase
the vertical spring constant as per Equation F.20, thereby decreasing the deflection as per Equation F.18, and
resulting in a larger rolling radius; this is sometimes called an “expansion”.

303
SAWE Paper No. 3602
Category Number 31.0

𝒅𝑹 = 𝒅𝑷 𝑹𝟐 / 𝒕𝒕 𝑬

Remember that this is only for rough estimation as this simplified relationship was
made by ignoring the stiffness contribution of the sidewalls, etc. However, the final relation
for determining the difference in an inflated no-load tire radius due to an inflation pressure
change is:

𝒅𝑹𝒊𝒏𝒇𝒍𝒂𝒕𝒆 = (𝑷𝟐 − 𝑷𝟏 ) 𝑹𝒊 𝟐 / 𝒕𝒕 𝑬 (EQ. F.32)

The use of Equation F.32 produces an inflated no-load tire radius (upper plot line)
versus inflation pressure plot as per Figure F.38 428:

Figure F.38 – TIRE RADIUS vs. INFLATION PRESSURE

428
Reference [89], pg. 58.

304
SAWE Paper No. 3602
Category Number 31.0

VELOCITY EFFECTS

An authoritative source states “In preliminary performance calculations the effect of


speed may be ignored (with respect to rolling resistance)” 429. However, the rolling resistance
variation with speed (velocity) is explicitly known via the Institute of Technology in Stuttgart
formula of circa 1938 430:

CR = CS + 3.24 CD (V/100)2.5 (EQ. F.33)


Given coefficient values such as those of Figure F.35, but appropriate for the specific
tires concerned, the vehicle rolling resistance can be reasonably determined for a wide range
of velocity variation.

There are other authoritative sources which state “Velocity does not significantly
affect cornering stiffness of tires in the normal range of highway speeds” 431 (therefore this
is a reference to passenger car tires), and “To a first order, tire forces and moments are
independent of speed” 432. However, there are definite decreases in traction coefficients with
velocity which may not be well defined, but for which there is considerable empirical data,
such as that contained in the following table for longitudinal traction 433:

Table F.02 – EMPIRICAL TEST: TRACTION DECLINE WITH SPEED INCREASE

429
Reference [93], pg.17. This is probably meant to apply only to passenger cars.
430
Reference [81], pg. 34. Other rolling resistance formulations exist, but in this author’s opinion the Stuttgart
equation is probably the best.
431
Reference [28], pg. 355. Also, on page 199: “Speed does not strongly influence the cornering forces
produced by a tire”.
432
Reference [52], pg.41. Also, on page 27: “…lateral friction coefficient is approximately independent of
speed…”.
433
Reference [34], pg. 81.

305
SAWE Paper No. 3602
Category Number 31.0

The average reduction of dry static traction coefficient of about “-0.003/mph” (-


0.002/kph) over the 20 mph (32 kph) to 40 mph (64 kph) observed range is admittedly a
limited linearization of the actual relation, which raises the question as to what is the real
nature of that relation. An authoritative source graphically depicts the variation in the wet tire
static (rolling) and dynamic (skidding) traction coefficients as follows 434:

Figure F.39 – STATIC & DYNAMIC TRACTION COEF vs. VELOCITY1/2

This shows that both the static and dynamic coefficients of traction are functions of
velocity, which is contrary to the nature of Coulomb friction. However, because the
coefficients are plotted against the square root of the velocity the nature of the relationship is
not easily seen. A plot of the static (rolling) and dynamic (skid) coefficients from the same
source, for the same tire under the same conditions, but replotted as a direct function of
velocity, is as follows 435:

434
Reference [26], pg. 456. The info here is both static and dynamic for a “175R14” (possible 70 aspect ratio)
tire under a 350 kg load @ 1.9 bar inflation pressure.
435
Ibid, pg. 434. This is essentially the same as above, but with dynamic (“locked” wheel) info only.

306
SAWE Paper No. 3602
Category Number 31.0

Figure F.40 – STATIC & DYNAMIC COEFFS OF TRACTION vs. VELOCITY

The traction relations reveal themselves as gentle curves to which parabolic regression
lines fit very well. The average reduction of the static coefficient is about “-0.008/mph” (-
0.005/kph) over the 20 mph (32 kph) to 40 mph (64 kph) range, which is about 2.7 times that
observed previously; however, that is probably mainly the effect of this being wet traction.
How the variation in traction with velocity affects the longitudinal traction/longitudinal “slip”
relationship may be depicted in a general way as follows 436:

Figure F.41 – LONGITUDINAL TRACTION vs. %SLIP & VELOCITY

436
This figure is a “cartoon” adaptation of an illustration found in Reference [30].

307
SAWE Paper No. 3602
Category Number 31.0

In acceleration and deceleration the variation in both the normal load and velocity
affect the maximum longitudinal traction available. However, it is more important to account
for the effect of both “N” and “V” on “μx” in a braking simulation such as
“MAXDLONG.BAS” than it is in an acceleration simulation such as “MAXGLONG.BAS”.
A braking simulation commences at a high “V”, and the determination of the maximum
traction available for deceleration is very much dependent on that “V”. An acceleration
simulation commences at zero “V”, and as the velocity increases the need for determination of
the maximum traction available decreases because the propulsion force available for
acceleration is also decreasing; this is contrary to the braking situation wherein the brake
force available for deceleration may actually increase with time.

Therefore this author chose to develop an expression for the variation of “μx” (a.k.a.
“μpeak”, maximum longitudinal traction coefficient) with regard to both “N” (“Pc”) and “V”:
“μx = f(N, Pc)”. A number of reference documents were utilized to develop a suitable
expression for “μx = f(V)” 437, which when combined with the known “μx = f(Pc)” (Eq. F.22)
creates:

𝝁𝒙 = 𝟒𝟖. 𝟕𝟏𝟓𝟗𝟔 𝑷𝒄 −𝟎.𝟗𝟔𝟑𝟒𝟑𝟐𝟔𝟔 𝑽−𝟎.𝟏𝟎𝟏𝟎𝟖𝟖𝟕𝟏 (EQ. F.34)

While the form of this equation may be generally applicable, it really represents just
the specific tire (152/92R16 @ 45 psi) for which it was developed; a different version of this
equation must be developed to represent any other tire/inflation pressure. Also note that
technically this equation is not a function of the normal load “N”, but of the contact pressure
“Pc”. However, that is a quibble (“Pc = N/Ac”). This equation for determination of “μx” was
incorporated in the traction subroutine of the “MAXDLONG.BAS” program as used in
Chapter 5.

An indication of the usefulness of Equation F.34 may be gleaned by a consideration of


how the equation performs as the variables “N” and “V” are varied independently, with the
results plotted as follows:

437
Reference [14], Reference [20], Reference [30].

308
SAWE Paper No. 3602
Category Number 31.0

Figure F.41 – CONSTANT “Pc” & CONSTANT “V” PLOTS OF μx = f(Pc, V)

Note that the general form of the fitted line equation to the “μx = f(Pc, V) with V held
constant” plot is in agreement with the form of Equation F.22, while in the other “μx = f(Pc,
V) with N held constant” plot the linearized variation in “μx” over the range 20 mph (32
kph) to 40 mph (64 kph) range of “-0.00345/mph” agrees well with the information of Table
F.02. This isn’t proof of the validity of Equation F.34, but the results seem reasonable and the
equation therefore useful. However, it should be noted that this whole effort flies in the face
of an authority’s statement “Tire performance varies with speed…the effect is not consistent
enough to be generalized” 438.

Increased velocity means increased flexing of the tire tread per unit time, thereby
generating more heat flow and raising the tire temperature, as indicated per the following
table 439:

Table F.03 – TIRE TEMPERATURE and VELOCITY

438
Reference [52], pg. 56.
439
Reference [94], pg. 205.

309
SAWE Paper No. 3602
Category Number 31.0

The increased temperature through the material properties of rubber has its own effect
on traction as per Figure F.34, but complicating matters is the subsequent daisy chain of
structural consequences: higher inflation pressure (per Equation F.31), increased vertical
stiffness (per Equation F.20), decreased deflection under load (per Equation F.18), decreased
tire-road contact area (per Equation F.17), leading to a further decrease in traction (per
Equation F.22) 440. Given all this action and reaction it is understandable that tire researchers
have historically had difficulty in trying to separate such things as velocity effects from
pressure effects; on the tire testing machine the two effects go hand-in-hand; such hard to
separate parameter interactions have been the main reason for the slow progress in the
understanding of tire behavior.

Since an increase in tire temperature results from increased tread flexure, velocity is
not the only driving parameter behind that effect. Varying longitudinal/lateral accelerations
will also cause tread flexure resulting in temperature increase; generally, the harder a vehicle
is driven the higher tire temperatures will raise. It also follows that the more underinflated a
tire is, the higher its temperature will climb; for optimum tire life it is wise to maintain tire
inflation pressures in accord with manufacturer’s specifications.

Velocity also directly affects the vertical spring constant of the tires, and seemingly in
a manner that leads to all sorts of confusion. Taylor, Bashford, and Schrock have
demonstrated that the measured value of a tire vertical spring constant “Kv” can vary
considerably depending upon the method used to do the measuring. The most common test
method, which accounts for the vast majority of measured vertical spring constant data, is the
“Load-Deflection” (LD) method. This method, along with four other methods, was evaluated
by these researchers regarding the “Kv” of a 260/80R20 agricultural (!) tire 441:

1. Load-Deflection (LD).
2. Non-Rolling Vertical Free Vibration (NR-FV).
3. Non-Rolling Equilibrium Load Deflection (NR-LD).
4. Rolling Vertical Free Vibration (R-FV).
5. Rolling Equilibrium Load-Deflection (R-LD).

From this list it’s easy to see that the major distinctions between the methods involve
whether the test is static/quasi-static (response to load) or dynamic (response to
impact/vibration). This author believes this distinction is the key to understanding what seems

440
Actually, there is a further complication as the decreased contact area means less tire tread being flexed,
thereby ameliorating the original temperature increase tendency, but not completely offsetting it.
441
Reference [82], pg. 1415.

310
SAWE Paper No. 3602
Category Number 31.0

to be a lot of confusing and conflicting test results, beginning with the Taylor, Bashford, and
Schrock paper and many of their cited references, and ending with a later paper by Kasprzak
and Gentz whose main result with regard to Formula SAE off-road racing tire vertical spring
rate is in seeming direct contradiction to Taylor, Bashford, and Schrock 442.

This author’s personal resolution of the matter, based mainly on intuition unsupported
by any solid substantiation, is that there may be two types of vertical spring rate involved.
One type of stiffness may be that measured by static/quasi-static load response methods (LD);
the other type of stiffness may be that measured by dynamic impact/vibration response
methods (FV 443). The static/quasi-static stiffness decreases with increasing velocity
asymptotically up to a certain speed dependent on the particular tire/inflation pressure
concerned. The impact/vibration stiffness increases with increasing velocity (due to inertial
effects). To say more than that would require further study, but it would seem that the
static/quasi-static vertical stiffness would be suitable for use in determining the tire rolling
radius variation due to “weight transfer” as used in acceleration/braking simulations, while the
impact/vibration stiffness would be suitable for use in suspension shock/vibration studies.

While any change in tire rolling radius variation under load due to velocity effect on
“Kv” is problematic, the rolling radius variation with velocity due to centrifugal force is
determinate. A tire under load tends to “expand” back to its full no-load inflated radius
dimension (“Ri” or “Di / 2”) due to centrifugal force as the velocity increases; accompanying
this is a true expansion (stretching of the carcass) also due to centrifugal effect, but such
stretching tends to be relatively small in most cases.

An extreme example of tire rolling radius “expansion” would involve the huge racing
slicks (usually bias, a.k.a. cross-ply, without “belts”) that some “dragster” types use. As an
“AA” fuel dragster accelerates off the line the huge rear slicks, usually Hoosier brand, expand
until the rear of the dragster appears to be standing on “tip toes”. This “standing on tip-toes”
behavior has been witnessed and photographically documented innumerable times. This tire
behavior is by design; the enlarged rolling radius at high speed is intended to change the
overall drive ratio of the vehicle, compensating for the fact that such dragsters tend to be
direct drive or only two-speed.

442
On page 9 Reference [38] shows a rolling vertical tire stiffness of 797 lb/in (1395 N/cm) versus a non-rolling
vertical tire stiffness of 708 lb/in (1239 N/cm) and states “This trend was seen in all the tires tested”.
443
There are also forced vibration tests, non-rolling or rolling, which would seem to be included here.

311
SAWE Paper No. 3602
Category Number 31.0

Such tire “expansion” is less extreme and considerably less noticeable for passenger
444
cars . The following figure shows a measured increase in rolling radius with speed (to about
150 kph, or 93.2 mph) for a 5.60×5 cross-ply (0-0.43 in) and a 155SR15 radial (0-0.15 in),
both tires are at 22 psi (152 kPa) pressure and under 661 lb (299.8 kg) normal load 445:

Figure F.42 – EMPIRICAL TIRE EXPANSION WITH VELOCITY

Even though this figure is supposedly empirical (such is not actually stated), it is
curious in certain aspects. The 5.60×15 tire at 22 psi (0.15 mN/m2) would have a vertical
spring rate of around 782 lb/in (140 kg/cm), and therefore should have a static deflection of
about 0.85 in (2.16 cm) under the 661 lb (299.8 kg) load 446; using the same sort of calculation
the 155SR15 radial should have a static deflection of about 0.81 in (2.06 cm). That the bias

444
Reference [26], pg. 10 notes: “Higher speed rated tires may feature a full (tread) width nylon cap pr
plies…on top of the…belts to further restrict expansion from centrifugal forces…Nylon cap strips used in
some constructions …cover only the belt edges”.
445
Reference [4], pg. 113. Bastow’s “Fig. 3-15” was redrawn with the plot lines extended slightly. The radial tire
seems to reach its expansion limit sooner than the bias tire supposedly by virtue of the constraining action of its
peripheral belts (per Bastow).
446
A “5.60×15” is an “Imperial” or “Numeric” tire designation dating from the 1923-1967 period; such a tire
would have an aspect ratio of 82. That ratio plus the information given explicitly in the designation (“SN = 5.60
in, DR = 15 in”), and the stated inflation pressure (“Pi = 22 psi”), allows for the estimation of the tire vertical
spring constant per Equation F.20 as 782 lb/in (140 kg/cm). The “155SR15” is a “Euro-numeric” designation of
the same period; such a tire would also have an aspect ratio of 82; the estimated vertical spring constant is 818
lb/in (146 kg/cm). Given the normal load as being 661 lb (299.8 kg) for both tires it is now a trivial matter to
calculate the static deflections.

312
SAWE Paper No. 3602
Category Number 31.0

tire only recovered about 55% of its static deflection by 124.3 mph (200.0 kph) seems
reasonable, but that the radial tire would recover only about 22% of its static deflection upon
reaching that same speed seems questionable. The fact that the radial tire would have
peripheral tread belts, possibly of steel construction, would have bearing only on the case of
true expansion (strain) of the tire circumference, not on deflection recovery.

Such curious aspects notwithstanding, Figure F.42 was utilized for the development of
a “tire rolling radius” subroutine for use in the MAXGLONG.BAS automotive acceleration
program. A 1984 parameter dump during successive runs of that program revealed the
following with regard to this tire expansion subroutine function:

Figure F.43 – SIMULATED TIRE EXPANSION WITH VELOCITY

The expansion-velocity plots of Figure F.43 are both for the same 6.00×16 bias tire,
just at different inflation pressures. Here by 122.5 mph (197.1 kph) the tire inflated to 23 psi
(158.6 kPa) has recovered 100% of its static deflection, and the tire inflated to 45 psi (310.3
kPa) has recovered 108% of its static deflection 447. The exact values displayed for the
6.00×16 tire of Figure F.43 are less significant than the fact that the simulated general form of
the Figure F.43 seems to be in reasonable agreement with the empirical form of the 5.60×15
tire expansion shown in Figure F.42.
447
The static deflections and other tire parameters used for Figure F.43 are those of the 1984 computer
simulation runs (Reference [89]); the tire methodology used by this author has progressed significantly since
then. In 1984 the value for “Kv” @ 23 psi was 1227 lb/in, and for “Kv” @ 45 psi was 2400 lb/in. The 1984 “E”
value used was 1,000,000 psi.

313
SAWE Paper No. 3602
Category Number 31.0
Rev. D
A closer match is not to be expected considering that the tires are different, as are the
inflation pressures (22 psi vs. 23 psi) and the loads (661 lb vs. 951 lb), not to mention the fact
that much of the tire parameter estimation methodology/tire expansion model of Figure F.42
has been considerably revised since 1984 (and is now supposedly more accurate). Referring to
Figure F.44, the present methodology for the static deflection recovery is derived as follows…

The mass of the deflected portion


of the tire periphery “m1-3” is pushed
back against the normal force reaction
“N” by the centrifugal acceleration
“V2/R” giving rise to the radius recovery
increment “dRrecovery”:

𝒅𝑹𝒓𝒆𝒄𝒐𝒗𝒆𝒓𝒚 = 𝒎𝟏−𝟑 𝑽𝟐 / 𝑹𝒓 𝑲𝒗

Using weight as a measure of


mass requires the substitution:

𝒅𝑹𝒓𝒆𝒄𝒐𝒗𝒆𝒓𝒚 = 𝑾𝟏−𝟑 𝑽𝟐 / 𝑹𝒓 𝒈 𝑲𝒗

Fig. F.44 – DEFLECTION RECOVERY


EXPANSION
The weight “W1-3” is equal to the volume “tt tw Lc” times the density “δ”, where “Lc”
is equal to “1.24 Ri Cos-1((Ri-d)/Ri)” as indicated by Equation F.17; making the substitution:

𝒅𝑹𝒓𝒆𝒄𝒐𝒗𝒆𝒓𝒚 = �𝒕𝒕 𝒕𝒘 𝜹 𝟏. 𝟐𝟒 𝑹𝒊 𝐂𝐨𝐬 −𝟏 �(𝑹𝒊 − 𝒅)/𝑹𝒊 ��(𝑽𝟐 ⁄𝑹𝒓 𝒈 𝑲𝒗 )

Substitute “386.088 in/sec2” for “g”, add “mph to in/sec” conversion constant “17.6”,
and rearrange slightly:

𝒅𝑹𝒓𝒆𝒄 = ((𝟏𝟕. 𝟔 𝑽)𝟐 ⁄𝑹𝒓 𝟑𝟖𝟔. 𝟎𝟖𝟖 𝑲𝒗 )(𝒕𝒕 𝒕𝒘 𝜹 𝟏. 𝟐𝟒 𝑹𝒊 𝐂𝐨𝐬−𝟏 ((𝑹𝒊 − 𝒅)/𝑹𝒊 ))

(EQ. F.35)
Although the effect is usually minor, there is a real expansion (strain) of the tire
periphery that causes a further increase in the effective tire radius with increasing velocity.
This tire expansion radius increment “dRexpand” runs concurrent with the recovery increment,
at least until the static deflection is fully counteracted which is when “dRrecovery” ceases. The

314
SAWE Paper No. 3602
Category Number 31.0

expansion model is as per Figure F.45, which is a free body diagram of a half periphery of the
tire:

The stress in the periphery


resulting from the centrifugal force
“F” can be determined by a study of
this free body diagram. This study
begins with:

𝑪=𝑫𝝅=𝟐𝑹𝝅

The ½ periphery length “L” is


equal to half of “2 R π”, so:

𝑳=𝑹𝝅

Fig. F.45 – CENTRIFUGAL STRAIN


EXPANSION
Therefore “R” can be expressed as:

𝑹=𝑳/𝝅
And the radius expansion increment as:

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = 𝒅𝑳 / 𝝅

Since the strain “ε” is equal to “dL / L” by definition the substitution for “dL” may be
made:

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = 𝑳 𝜺 / 𝝅

Make use of the stress-strain relation “S = E ε” to substitute “S/E” for “ε”:

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = 𝑳 𝑺 / 𝑬 𝝅

Substitute “(F/2)/A” for “S”:

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = 𝑳 𝑭 / 𝟐 𝑨 𝑬 𝝅
� ω2 / g” for “F”:
Substitute “WL 𝐑

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = 𝑳 𝑾𝑳 �
𝑹 𝝎𝟐 / 𝒈 𝟐 𝑨 𝑬 𝝅

315
SAWE Paper No. 3602
Category Number 31.0
Rev. D
The weight “WL” of the tire tread sector “L” is equal to “L tw tt δ”, the CG coordinate

“𝐑” of that sector is equal to “2 R / π” 448, the angular velocity “ω” is equal to “V/Rr”, and the
transverse tread area of the tire cords is estimated as “tw tt f”449; making the corresponding
substitutions and simplifying results in:

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = 𝑳𝟐 𝜹 𝑹 𝑽𝟐 / 𝑹𝒓 𝟐 𝒈 𝒇 𝑬 𝝅𝟐

From earlier simple circular relations the length “L” is known to be equal to “R π”,
which is now substituted for “L”:

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = (𝑹𝝅)𝟐 𝜹 𝑹 𝑽𝟐 / 𝑹𝒓 𝟐 𝒈 𝒇 𝑬 𝝅𝟐

Simplify:

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = 𝑹𝟑 𝜹 𝑽𝟐 / 𝑹𝒓 𝟐 𝒈 𝒇 𝑬

Now some unit conversion factor (mph × 17.6 = in/sec) and constant value (g =
386.088 in/sec2) adjustments are required, and “R” becomes “Ri” as that radius better
represents the general condition along the periphery “C” than “Rr”:

𝒅𝑹𝒆𝒙𝒑𝒂𝒏𝒅 = 𝑹𝒊 𝟑 𝜹 (𝟏𝟕. 𝟔 𝑽)𝟐 / 𝑹𝒓 𝟐 𝟑𝟖𝟔. 𝟎𝟖𝟖 𝒇 𝑬 (EQ. F.36)


As the velocity “V” increases the deflection recovery (Equation F.43) and the
centrifugal expansion of the tread (Equation F.44) work together (are additive) in increasing
the rolling radius “Rr” until the point is reached when “Rr = Ri” which signifies total recovery
of the initial static deflection. Beyond that point the centrifugal expansion carries on alone
until the tire self-destructs. Of course, that seldom happens as velocities high enough to cause
tire destruction by centrifugal stresses are reached only by vehicles such as LSRs.

For passenger car tires true centrifugal expansion is very minor, especially if the tires
in question have steel (“E ~ 29,000,000 psi”) or other high Modulus of Elasticity material
belts. Most modern passenger car tires are radials and belted, but a bias tire dating from the
1945-1967 period might rely on only nylon and/or rayon cords for constraint (“E ~ 410000
psi or so”).

It should perhaps be noted at this point that the exact value of the Modulus of
Elasticity to be used for a calculation or simulation is not necessarily a “cut and dried” matter.
Even for the same tire, the modulus value will vary depending on the use to which that value

448
Reference [98], pg. 4.4.1.
449
Only a fraction “f” of the tread cross-section area is actually resisting the expansion, which is the area of the
carcass cord and tread belts. Admittedly, this is essentially a matter of “fudge factors”, but such factors are
significant in computer simulations of a complex reality as the judicious use of such factors allows for a “dialing
in” or “fine tuning” of the simulation. Initially “f = 0.12” may be used for just the carcass cord area, and if belts
are present then the use of “f = 0.25” is a possibility.

316
SAWE Paper No. 3602
Category Number 31.0
Rev. D
is to be put. For example, the Krylov and Gilbert equation for determining the critical velocity
for “standing wave” formation was derived using a model of a longitudinal tire tread segment
as a beam supported on springs. Even though a modern (c. 2003) passenger car tire (exact
type unspecified) parametric value set served as input for their example calculation, the
modulus value was only 171,304 psi (“3·107 N/m2”) 450! This is a far cry from the modulus
value for steel (29,000,000 psi), or any other conceivable tread belt material, but that is
because the belts reside near the neutral axis of the beam model, and thus have no role in this
particular calculation. The modulus value that Krylov and Gilbert used was some combination
of the carcass cord and tread materials (Ref.: Enylon ~ 410,000 psi, Erubber ~ 500 psi 451), as
was suitable for their purpose. If the calculation had instead been one of tire peripheral
expansion in response to centrifugal force, then the modulus used would have been much
higher, probably very close to the Modulus of Elasticity of the belt material 452.

Consider the result of a practical application of Equations F.35 and F.36 as presented
herein. The tire is the 1958 Dunlop RS4 6.00×16 bias tire (with inner tube) at 45 psi (310 kPa)
under a 1238 lb (561.5 kg) normal load, resulting in a 0.84 in (21.3 mm) static deflection. The
tire is presumed to be unbelted, and largely of nylon cord/rubber construction; the Modulus of
Elasticity is taken as 450,000 psi (3·109 N/m2). An iterative spreadsheet calculation of the
subject equation results looked as follows when plotted:

Figure F.46 – DUNLOP RS4 TIRE ROLLING RADIUS EXPANSION W/ VELOCITY

Note that the general shape of this plot is much like the 1984 plot of the same tire as
depicted in Figure F.43. More significantly, it is also similar to the empirical plots of Figure
450
Reference [41], pg. 4227.
451
As noted earlier, the Modulus of Elasticity for rubber exists only as an infinite number of “short interval”
values of limited applicability. However, for the sort of rubber used in tires it is never more than a few hundred
psi.
452
While in such a case the rubber contribution may be so minor as to be negligible, the Modulus of Elasticity of
some carcass cord materials is high enough that there may be justification for the use of a composite “E” value.

317
SAWE Paper No. 3602
Category Number 31.0
Rev. D
F.42. Of course, this does not constitute a validation of the methodology, but it does give
some reason for confidence. Of particular interest is the point of inflection in the plot at 172
mph (277 kph) where the axle height change reaches 0.84 in (2.1 cm), indicating full initial
static deflection recovery; from that point onward only true expansion of the tire periphery
(Equation F.36) occurs. Ultimately, if a high enough velocity can be reached, the tire
peripheral ultimate strength will be exceeded and it will suffer catastrophic destruction, if it is
not destroyed by other effects (temperature/pressure, “standing wave”, etc.) first.

The symbolism for Equations F.32, F.36, and F.37 is as follows:

d = Tire vertical deflection under a normal load (in).

dRexpand = Tire vertical expansion (in).

dRrec = Tire vertical deflection recovery (in).

δ = Tire tread cords/belts/rubber composite density (ex.: 0.0462 lb/in3)

E = Tire tread Modulus of Elasticity (Nylon 410K – Steel 29M lb/in2).

f = Tire tread cross-sectional area “E” adjust factor (0.12-0.25 dimensionless).

Kv = Tire vertical spring rate (lb/in).

P1 = Tire initial inflation pressure (lb/in2).

P2 = Tire new inflation pressure (lb/in2).

Ri = Tire inflated no-load radius (in).

Rr = Tire rolling radius (in).

tt = Tire tread thickness (ex.: 0.65 in).

tw = Tire tread width (in).

V = Vehicle velocity (mph).

(This concludes the “Notes on Tire Behavior”)

318
SAWE Paper No. 3602
Category Number 31.0

APPENDIX G – TRANSIENT STATE: CENTER OF OSCILLATION

When a force is suddenly applied to a body, the center of oscillation is a point on the
body which does not move translationally at that instant; instead the body rotates around that
“center of oscillation”. In the following automotive plan view a step steer increment has
suddenly produced the net traction force “Ff” at the front axle; this tends to angularly
accelerate the vehicle at some rate “α” about a point “o” (the Center of Oscillation) causing a
dynamic reaction at the vehicle CG of “m bα”, and an opposing traction force “Fr” at the rear
axle 453:

Figure G.01 – TRANSIENT STATE STEP STEER INITIATION OF A TURN

453
Since the location of the rotation point is a matter of conjecture at this time it is conveniently represented as
being on the rear axle, and since the rear axle reaction existence and direction can be arbitrarily assumed at this
point such is also subject to convenience.

319
SAWE Paper No. 3602
Category Number 31.0

The following analysis parallels that of Dr. Hugo S. Radt Jr. 454 as presented in Chassis
Design (Principles and Analysis) 455, but is somewhat different in the details of its approach.
However, this analysis is the same in that it begins with the dynamic summation of the forces
and moments. The dynamic summation of the forces in the lateral direction at the instant in
question is:

𝑭𝒇 = 𝑭𝒓 + 𝒎𝒃𝜶 (EQ. G.01)

The dynamic summation of the moments about “o” at that instant is:

𝑭𝒇 (𝒂 + 𝒃) = (𝑰𝒛 + 𝒎𝒃𝟐 )𝜶 (EQ. G.02)

The term “Iz + mb2” represents the transfer of the vehicle yaw inertia “Iz” about the
CG to the point “o”. The rest of this analysis is very simple mathematics. Equation 1 can be
used to eliminate “Ff” from the above by substitution:

(𝑭𝒓 + 𝒎𝒃𝜶)(𝒂 + 𝒃) = �𝑰𝒛 + 𝒎𝒃𝟐 �𝜶

(𝑰𝒛 + 𝒎𝒃𝟐 )𝜶
𝑭𝒓 + 𝒎𝒃𝜶 =
(𝒂 + 𝒃)

(𝑰𝒛 + 𝒎𝒃𝟐 )𝜶
𝑭𝒓 = − 𝒎𝒃𝜶
(𝒂 + 𝒃)

The “Iz” term is equal to the yaw radius of gyration of the vehicle squared “k2” times
the vehicle mass “m”, which means “k2m” may be substituted for “Iz”:

( 𝒌𝟐 𝒎 + 𝒎𝒃𝟐 )𝜶
𝑭𝒓 = − 𝒎𝒃𝜶
(𝒂 + 𝒃)

The result is then subjected to a little manipulation to get the expression into the
appropriate form:

( 𝒌𝟐 + 𝒃𝟐 )𝒎𝜶
𝑭𝒓 = − 𝒃𝒎𝜶
(𝒂 + 𝒃)

454
Dr. Radt is the former Head of the Land Vehicles Section, Vehicle Dynamics Department, Cornell
Aeronautical Laboratory (now Calspan Corporation).
455
Reference [53], pp. 247-249.

320
SAWE Paper No. 3602
Category Number 31.0

� 𝒌𝟐 + 𝒃𝟐 �
𝑭𝒓 = � − 𝒃� 𝒎𝜶
(𝒂 + 𝒃)

� 𝒌𝟐 + 𝒃𝟐 � 𝒃(𝒂 + 𝒃)
𝑭𝒓 = � − � 𝒎𝜶
(𝒂 + 𝒃) (𝒂 + 𝒃)

� 𝒌𝟐 + 𝒃𝟐 � �𝒂𝒃 + 𝒃𝟐 �
𝑭𝒓 = � − � 𝒎𝜶
(𝒂 + 𝒃) (𝒂 + 𝒃)

� 𝒌𝟐 + 𝒃𝟐 − 𝒂𝒃 − 𝒃𝟐 �
𝑭𝒓 = � � 𝒎𝜶
(𝒂 + 𝒃)

� 𝒌𝟐 − 𝒂𝒃 �
𝑭𝒓 = � � 𝒎𝜶
(𝒂 + 𝒃)

� 𝒌𝟐 − 𝒂𝒃 �⁄𝒂𝒃
𝑭𝒓 = � � 𝒎𝜶𝒂𝒃
(𝒂 + 𝒃)

The division of the numerator “k2-ab” by “ab” results in the final equation:

� 𝒌𝟐 ⁄𝒂𝒃 �−𝟏
𝑭𝒓 = � (𝒂+𝒃)
� 𝒎𝜶𝒂𝒃 (EQ. G.03)

Note that “k2/ab” is our old friend, the Dynamic Index in Yaw, or DIY. If “k2/ab = 1”
then the equation reduces to “Fr = 0”, which means that point “o” is indeed the Center of
Oscillation. If “k2/ab > 1” then “Fr” is greater than zero (positive) and therefore acts in the
direction shown in the plan view. If “k2/ab < 1” then “Fr” is less than zero (negative) and
therefore acts opposite the direction shown in the plan view.

321
SAWE Paper No. 3602
Category Number 31.0
Rev C

APPENDIX H – ROAD TO UNSPRUNG MASS TRANSMISSIBILITY

“…reducing the (road to) unsprung mass transmissibility provides…

better road holding ability and improved vehicle stability.” 456

Ahmadian and Pare, 2000

After this paper was originally presented the author became aware of yet another
useful vibration transmissibility (a.k.a. “gain”) equation. The equation for the transmissibility
factor of small scale road input to unsprung mass output was presented by authors
Mehmood, Khan, and Kahn in their paper “Vibration Analysis of Damping Suspension Using
Car Models” on page 205 457:

𝒁𝒖 𝒎𝒔 𝑨𝝎𝟒 +(𝒎𝒔 𝑩+𝒄𝒔 𝑨)𝝎𝟑 +(𝒎𝒔 𝑪+𝒌𝒔 𝑨+𝒄𝒔 𝑩)𝝎𝟐 +(𝒌𝒔 𝑩+𝒄𝒔 𝑪)𝝎+𝒌𝒔 𝑪
=
𝒁𝒓𝒐𝒂𝒅 𝒄𝒔 𝑬𝝎𝟓 +(𝒄𝒔 𝒄𝒕 𝑭+𝒌𝒔 𝑬)𝝎𝟒 +(𝒄𝒔 𝑮+𝒌𝒔 𝑭)𝝎𝟑 +(𝒄𝒔 𝑩+𝒌𝒔 𝑮)𝝎𝟐 +(𝒄𝒔 𝑪+𝒌𝒔 𝑩)𝝎+𝒌𝒔 𝑪

(EQ. H.1)
Where:

𝑨 = 𝒄𝒔 𝒄𝒕 , 𝑩 = 𝒄𝒔 𝒌𝒕 + 𝒄𝒔 𝒌𝒔 , 𝑪 = 𝒌𝒔 𝒌𝒕 , 𝑬 = 𝒎𝒔 𝒎𝒖 ,

𝑭 = 𝒎𝒔 𝒄𝒕 + 𝒎𝒔 𝒄𝒔 + 𝒎𝒖 𝒄𝒔 , 𝑮 = 𝒎𝒔 𝒌𝒕 + 𝒎𝒔 𝒌𝒔 + 𝒎𝒖 𝒌𝒔 + 𝒄𝒔 𝒄𝒕
So the only real parameters are:

ω = the angular frequency in radians per second (ω = 2π Hz).


ms = the sprung mass (lb-sec2/in).
mu = the unsprung mass (lb-sec2/in).
cs = the damping coefficient for the suspension damper (lb-sec/in).
ct = the damping coefficient for the tire “damper” (lb-sec/in).
Rev C

ks = the spring rate for the suspension spring (lb/in).

kt = the spring rate for the tire (lb/in).


456
Ahmadian, Mehdi; and Christopher A. Pare, “A Quarter-Car Experimental Analysis of Alternative Semi-
Active Control Methods”, Journal of Intelligent Material Systems and Structures, Vol. 11, No. 8, ISSN 1045-
389X, 2000, pp. 604-612.
457
Reference [48], pg. 5. Also see: Mehmood, Khan, and Kahn; “Vibration Analysis of Damping Suspension
Using Car Models”, Innovation Space of Scientific Research (ISSR) Journal, September 2014, pp. 202-211.

322
SAWE Paper No. 3602
Category Number 31.0
Rev C

For supposedly typical passenger car parameter values of “ms = 1.97 lb-sec2/in, mus =
0.1796 lb-sec2/in, cs = 17.1304 lb-sec/in, ct = 0.0000 lb-sec/in, ks = 314.1 lb/in, kt =1741.6
lb/in”, the small scale “road input to unsprung mass output gain” function supposedly should
look like: 458

Figure H.1 – ROAD INPUT-UNSPRUNG MASS OUTPUT FUNCTION (Hz = ω/2π)

However, in actual use for this paper Equation H.1 produced plots that looked as
follows when using the parameter values of Table 7.2:

Figure H.2 – BASELINE FRONT AXLE OUTPUT FUNCTION (Hz = ω/2π)

458
Mehmood, Khan, and Mehmood; “Dynamic Modeling of Vehicles and Damper”, International Organization
of Scientific Research (ISOR) Journal, April 2014, pp. 22-33.

323
SAWE Paper No. 3602
Category Number 31.0
Rev C

To investigate this matter of form further, recourse was made to J.Y. Wong’s book
Theory of Ground Vehicles which presents a different formulation for the small scale “road
input to unsprung mass output gain” function 459:

𝒁𝒖 �[𝒌𝒕 (𝒌𝒔 − 𝒎𝒔 𝝎𝟐 )]𝟐 + (𝒄𝒔 𝒌𝒕 𝝎)𝟐


=
𝒁𝒓𝒐𝒂𝒅 �[(𝒌𝒔 − 𝒎𝒔 𝝎𝟐 )(𝒌𝒕 − 𝒎𝒖 𝝎𝟐 ) − 𝒎𝒔 𝒌𝒔 𝝎𝟐 ]𝟐 + (𝒄𝒔 𝝎)𝟐 [𝒎𝒔 𝝎𝟐 + 𝒎𝒖 𝝎𝟐 − 𝒌𝒕 ]𝟐

(EQ. H.2)

This simpler formulation (note the fact that the tire damping coefficient “ct” is
neglected) produces gain vs. frequency plots that look a lot more like Figure H.1, but only at
elevated (Table 7.2 “cs” multiplied by 5) suspension damping values:

Figure H.3 – BASELINE FRONT AXLE OUTPUT FUNCTION (Hz = ω/2π)

The simpler formulation of Equation H.2 agreed with the formulation of Equation H.1
that the “advanced concept” would have significantly higher unsprung mass gain with
respect to the “baseline” vehicle, but in many other ways was at odds not only with the
Mehmood equation but with common sense (this last is a comment on a “gain sensitivity to
parameter variation” analysis conducted using the Wong formulation).

459
Reference [93], pg. 481. The formulation of interest in this reference is labeled “Equation (7.23)”.

324
SAWE Paper No. 3602
Category Number 31.0
Rev C

Therefore, the following results are based on Equation H.1 and the parameter input
values of Table 7.2. Based on the average obtained when the transmissibility is treated as a
linear function from “0 Hz” to “15 Hz”, the results show increased gain for the “advanced
concept” by about 7.36% to 10.37%:

Table H.1 – BASELINE & ADVANCED ROAD-UNSPRUNG MASS GAINS

Given the fact that the “advanced concept” has so many more individual tire-to-
ground contact patches than the “baseline”, as well as more overall ground contact area, the
indicated increase in road-to-unsprung mass transmissibility may not be significant.
However, if it were deemed necessary to diminish the increased transmissibility of the
“advanced concept”, then that could be accomplished by changing some of the relevant
parameters (although consideration must also be given to the affect on many of the other
performance aspects, such as acceleration, braking, ride, road shock, gyroscopic reaction,
etc.). To determine the road-to-unsprung mass gain sensitivity to such changes a variation of
+
/- 10% of each of the relevant parameters was conducted for one of the gain calculations and
the resultant gain effect noted:

Table H.2 – ROAD-UNSPRUNG MASS GAIN SENSITIVITY

325
SAWE Paper No. 3602
Category Number 31.0
Rev C

The first thing that may be noted is that the sensitivities to parameter change tend to
not be symmetric. That is, a parametric change of +10% may produce a greater or lesser
absolute gain variation than a parametric change of -10%. The next item of note is that the
road-to-unsprung mass vibration gain sensitivity is greatest for variations in tire stiffness
“kt”, which corresponds to the well known fact that road “harshness” is very amenable to
simple reductions in tire inflation pressure. A reduction in tire inflation pressure would also
increase tire damping (“ct”), which also would have a beneficial effect on the gain. However,
it should be remembered that inflation pressure reduction has an adverse effect on the load
carrying capacity “Lc” which has been shown to be critical for the “advanced concept”.

Also notable is that the road-to-unsprung mass vibration gain is almost as sensitive to
changes in unsprung mass as it is to tire stiffness; a 10% increase in unsprung mass (“mu”)
will decrease the gain, and thereby proportionately increase road contact, by 1.30%. This last
may be part of the reason certain automotive engineers (such as Peter E. Bryant, Engineering
VP of AMTECH Corporation?) seem to consider the reduction of unsprung mass to be a
mitigated benefit 460.

However, to increase unsprung mass to improve road contact is like increasing


sprung mass to improve ride; it may work with regard to small scale road induced vibration
but will be detrimental to many other performance aspects. So, if necessary, a reliance on
varying suspension damping (“cs”) may suffice; the “advanced concept” has shown itself
particularly amenable to such control due to its long cycle times, and such semi-active
suspension treatments are becoming commonplace anyway.

460
Bryant, Peter E; “Un-Sprung Weight, the Enemy that Became a Friend”, SAE Paper 2004-01-3534,
December 2004. Actually, Mr. Bryant’s paper seems to be in support of using “rebound control springs” to
modify the rate of unsprung mass rebound, and is not in favor of increasing unsprung mass (which would also
modify the rate of rebound). However, historically some engineers have voiced reservations about decreasing
unsprung mass “too much”, and in particular with regard to road shock attenuation. However, that particular
fallacy was dealt with in Reference [91], pp. 19-20.

326
SAWE Paper No. 3602
Category Number 31.0

(THIS PAGE INTENTIONALLY LEFT BLANK)

327

You might also like