Вы находитесь на странице: 1из 113

Rev B SAWE Paper No.

3521
Category Number 31.0

MASS PROPERTIES AND AUTOMOTIVE VERTICAL DYNAMICS


By

Brian Paul Wiegand, P.E.

For Presentation at the


70th Annual International Conference
of the
Society of Allied Weight Engineers, Inc.,
Houston, TX, 14-19 May 2011
(Rev A 4/13/2014)

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
Rev B SAWE Paper No. 3521
Category Number 31.0

TABLE OF CONTENTS

Chapter: Page:

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


Abstract …………………………………………………………………………………………. ii
1 - Introduction ……………………………….………………………….……………………… 1
2 – Some One DOF Models, and a Two DOF Model .……..……………………………..….… 2
3 - Damping …………………………………….….…………………………………................10
4 – Road Shock …..……………………………………………...….…........................……… 17
5 – Road Contact ………………………………………..………………….………….….......... 22
6 – Road Vibration ……………………………………………………………………………... 28
7 – Gyroscopic Reactions .……….…………………………………………………………….. 49
8 – More Two DOF Models ……...…………………...……….……………………….…....… 54
9 – The Principal Modes ………………………………………………….……………………. 57
10 – The Spring Center …..…………………………………..……….…………….………..… 61
11 – The Conjugate Centers of Percussion ..……….....…..…….………………………....…… 62
12 – 1958 Jaguar XK150S, 1980 Ford Fiesta S .………..……………………………………… 64
13 – Olley’s Rules ….………………………………………………………………………….. 69
14 - Conclusions ……………………………….………….…………………………………… 73
References…………………………………………….…………………………..…..…........... 81
Author’s Biographical Sketch……………………….…………………...…………..…....…… 83
Appendices………………………………………………………………….…………….…......84
A - Symbolism………………………………………………….…………………...……..85
B –Rolling Radius and Road Shock …………………………...………………..……..…93
C - Physiological Shock and Vibration Limits …………………...……………………...106

i
Rev B SAWE Paper No. 3521
Category Number 31.0
ABSTRACT

The mass properties of a vehicle affect its motion in all directions, translational and
rotational. Previously this author has dealt with how mass properties affect automotive
longitudinal dynamics 1 and automotive lateral dynamics 2. Now a consideration is in order of
how mass properties affect automotive vertical dynamics. Of course, lateral or longitudinal
inputs can lead to vertical responses; every aspect of a vehicle’s dynamics is interconnected with
every other aspect, but it is convenient to divide up automotive dynamics as if the subject were
truly a matter of independent motions in the longitudinal, lateral, and vertical directions.

Initially, this paper will investigate the significance of mass properties with regard to
automotive ride (transmission of road shock & vibration) and road-holding (maintaining contact
at the tire/road interface) through the use of simple, undamped, 1-DOF models. Later, the full
story of how mass properties influence the bounce and pitch motions of the sprung mass will
necessitate recourse to more complex 2-DOF models. The mass properties of greatest relevance
to this investigation will prove to be the “sprung mass”, the “unsprung masses”, the “sprung
mass distribution” (longitudinal, lateral, and vertical c.g.), the rotational inertias of the
rotating portions of the “unsprung masses”, and the “sprung mass” longitudinal and lateral
mass moments of inertia.

The basic intent of this paper is to counter the commonly held simplistic concept of the
role mass properties play in determining ride and road-contact. For those that have never
undertaken any study of the matter, the general presumption seems to be that all that is required
to achieve optimum performance is to minimize the weight and to obtain a balanced mass
distribution. The reality is that there are many aspects to automotive performance, and what
constitutes an optimum mass properties condition is generally a very complex matter which
often necessitates difficult compromises. Tailoring some mass property parameters so as to
achieve a desirable level of behavior with regard to one performance criterion will often
adversely affect other performance criteria.

Although this paper is restricted to mass properties issues related to performance


resulting from motion in the vertical direction, occasional reference will be made to those mass
properties requirements necessitated by performance considerations associated with the
longitudinal (acceleration, braking) and lateral (maneuver, roll-over, and directional stability)
directions, as revealed in the previous investigations noted earlier. To do otherwise would be to
work in a vacuum; the nature of reality tends to be such that all things are ultimately interrelated.
To the fullest extent possible, the greater intent herein is to approach reality through the totality
of the papers and articles written by this author on the subject of mass properties and
automotive performance.

1
Reference [20].
2
Reference [19].

ii
Rev B SAWE Paper No. 3521
Category Number 31.0
1 - INTRODUCTION

That mass properties play a significant role regarding automotive vertical motion is, of
course, obvious. What may not be obvious is the degree of complexity involved. Vertical motion
is the province of the suspension system, which involves the physics of what may be regarded as
a series of interacting mass-spring systems. The most complete model of an automotive
suspension system generally used may be illustrated as follows:

Figure 1.1 – TEN DOF SUSPENSION MODEL

This model has far too many DOF (degrees of freedom) for analysis by classical means.
Modern computer analysis can handle such a 10-DOF model 3 very easily and precisely, but such
easy precision is too often obtained at the expense of sound human understanding. By bringing
the model into the realm of human understanding through appropriate simplifications the matter
becomes amenable to those most powerful of human traits: thought and imagination. There is
also the added benefit of being able to perform a simple “sanity check” on a computer analysis of
the detailed model. This is not an argument against computer analysis; which should be carried
out with as much detail as the state of the art allows, but the basic design direction should always
be derived through human understanding.

3
The “sprung mass” is credited with 3 translations and 3 rotations, and each of the four “unsprung masses” is
credited with one translation (vertical). If each of the “unsprung masses” were credited with 3 translations and 3
rotations, then the total model would be 30-DOF, which would represent more complexity than is ever really
needed.

1
Rev B SAWE Paper No. 3521
Category Number 31.0

2 – SOME ONE DOF MODELS, AND A TWO DOF MODEL

Fortunately, suspension system behavior is such that it allows for a good deal of
simplification without much loss in accuracy. The automotive “sprung mass” vibratory motions
are generally of the order of 1 cps, whereas the major “unsprung mass” motions may be a
number of times that; the mode known as “wheel hop” may be around 10 times greater. This
wide separation in behavior indicates that the “sprung” and “unsprung” motions can be treated as
being independent. Hence, some insights into suspension behavior can be obtained by
construction of various mass models simplified by treating “sprung” and “unsprung” behaviors
as independent and reducing possible motion to just 1-DOF.

The resulting 1-DOF models, which are referred to as “quarter-car” models, are the
simplest suspension models possible. Three of these models are as depicted in Figure 2.1. The
first model considers only the relatively slow motion of the quarter-car sprung mass “ms”
(which is a “quarter-car” portion of the total vehicle sprung mass “Ms”) bouncing gently on the
suspension. The second model considers the somewhat more abrupt motion of the unsprung
mass “mus”, under the influence of the suspension spring constant “ks”, in response to road dips.
The third model considers the even higher frequency motion of the “unsprung mass”, mostly
under the influence of the much stiffer tire spring constant “kt”, as stimulated by road vibration.

Figure 2.1 – ONE DOF SPRUNG & UNSPRUNG MODELS

Ignoring damping (coefficients “cs” and “ct”) for now, the natural frequencies of the
sprung (“fs”) and unsprung (“fus”) mass systems can readily be determined by simple equations
corresponding to the models shown:

𝟏
𝒇𝒔 =
𝟐𝝅
�(𝒌𝒔 ⁄𝒎𝒔 ) (Eq. 2.1a)

2
Rev B SAWE Paper No. 3521
Category Number 31.0
𝟏
𝒇𝒖𝒔 =
𝟐𝝅
�(𝒌𝒔 ⁄𝒎𝒖𝒔 ) (Eq. 2.2a)

𝟏
𝒇𝒉𝒐𝒑
𝒖𝒔
=
𝟐𝝅
�(𝒌𝒕 ⁄𝒎𝒖𝒔 ) (Eq. 2.3a)

The first equation, Eq. 2.1a, represents the resonance or natural frequency of motion of
either a front or rear quarter of the sprung mass on a suspension spring as if the vehicle’s front
and rear were independent (which is a situation which can come close to realization under
special conditions); this is useful in the study of disturbance transmission from road to passenger.
The first unsprung frequency equation, Eq. 2.2a, quantifies the behavior of the unsprung mass
on the suspension spring, and is the appropriate model for the investigation of wheel reaction
(maintenance of contact) to road dip. In the second unsprung frequency equation, Eq. 2.3a, the
spring quality of the tire replaces the suspension spring, which results in the appropriate model
for the investigation of “wheel hop” or resonance frequency of the unsprung mass. There is a
huge difference between “fus” and “fhopus”; these are fundamentally different, whereas all the “fs”
bounce equations to be introduced herein will just represent variations in precision. Two of the
three equations can be readily expressed as intermediate equations of a little more complexity
and precision:

𝟏
𝒇𝒔 =
𝟐𝝅
�(𝒌𝒄𝒔 ⁄𝒎𝒔 ) (Eq. 2.1b)

𝟏
𝒇𝒉𝒐𝒑
𝒖𝒔
=
𝟐𝝅
�𝒌𝒄𝒑 ⁄𝒎𝒖𝒔 (Eq. 2.3b)

Note that for the sprung mass both the “springs”, suspension and tire, act in series giving
rise to the combined spring rate “kcs”, which produces a somewhat “softer” system 4:

𝒌𝒄𝒔 = (𝒌𝒔 × 𝒌𝒕 )⁄(𝒌𝒔 + 𝒌𝒕 ) (Eq. 2.4)


For the unsprung mass the springs are considered to be in parallel, with the combination
resulting in the additive spring rate “kcp”, producing a somewhat “harder” system 5:

𝒌𝒄𝒑 = 𝒌𝒔 + 𝒌𝒕 (Eq. 2.5)

There are also exact equations, but only the exact equation for the sprung mass need be
considered 6. To obtain an exact equation for the sprung mass means moving away from a view
of the sprung and unsprung masses as vibrating independently of each other, and toward a
more realistic view of their movements affecting one another. Since we now have two

4
Reference [6], page 20. If present, stabilizer bars, a.k.a. anti-roll bars, have an effect on spring rate for single
wheel motion, but this is neglected herein. For info on determining the effect see Reference [2], pages 70-72.
5
Ibid, page 19.
6
The exact equation for unsprung “hop” (resonance) is complicated, and produces results only slightly different
from the intermediate equation. However, if the reader must satisfy an obsession for achieving the greatest
accuracy possible, then the reader can see Reference [1], pages 13-14.

3
Rev B SAWE Paper No. 3521
Category Number 31.0
movements in our model, although both are vertical, the model is now actually of the 2-DOF
type:

Figure 2.2 – TWO DOF EXACT SPRUNG MODEL

The exact sprung mass equation comes about from the recognition that the motion of the
sprung and unsprung masses are linked, and that link is established by a quantity “x” which is
a value such that “0<x<1”. Start with the equation for the sprung mass system with springs
combined in series:

𝒌𝒔 = �𝒌𝒔 × 𝒌𝒕 (𝟏 − 𝒙)���𝒌𝒔 + 𝒌𝒕 (𝟏 − 𝒙)�

And correspondingly the unsprung system spring coefficient is:

𝒌𝒖𝒔 = 𝒌𝒕 𝒙
Where “x” is in effect dividing up the effect of the tire spring between the sprung and
unsprung systems, producing the equations:

𝟏
𝒇𝒔 = �(�𝒌𝒔 × 𝒌𝒕 (𝟏 − 𝒙)���𝒌𝒔 + 𝒌𝒕 (𝟏 − 𝒙)�)⁄𝒎𝒔
𝟐𝝅

𝟏
𝒇𝒖𝒔 =
𝟐𝝅
�(𝒌𝒕 𝒙)⁄𝒎𝒖𝒔

Equating the two linked frequencies results in:

(�𝒌𝒔 × 𝒌𝒕 (𝟏 − 𝒙)���𝒌𝒔 + 𝒌𝒕 (𝟏 − 𝒙)�)⁄𝒎𝒔 = (𝒌𝒕 𝒙)⁄𝒎𝒖𝒔

4
Rev B SAWE Paper No. 3521
Category Number 31.0
So then the quantity “ktx/mus” can be substituted for “((ks×kt(1-x))/(ks+kt(1-x)))/ms”
under the radical sign in the expression for “fs” and the resulting exact equation is:

𝟏
𝒇𝒔 =
𝟐𝝅
�(𝒌𝒕 𝒙⁄𝒎𝒖𝒔 ) (Eq. 2.1c)

Where the interrelating factor “x” is obtained from an iterative solution of:

𝒌𝒕 𝒙 = (𝒎𝒖𝒔 ⁄𝒎𝒔 )��𝒌𝒔 × 𝒌𝒕 (𝟏 − 𝒙)���𝒌𝒔 + 𝒌𝒕 (𝟏 − 𝒙)�� (Eq. 2.6)

Note the appearance of the unsprung-to-sprung mass ratio “mus/ms”; this is a ratio of
profound importance in automotive suspension design, and will be encountered again and again.

Anyway, all of the preceding mathematical manipulation was directed at obtaining a


solution method that is easy to use; “x” can now be found by successive iterations of the “ktx”
equation (Eq. 2.6) and then substituted into the “fs” equation (Eq. 2.1c) to find the exact sprung
mass system frequency.

That the simple equations are generally good enough, at least for most conceptual work,
will be illustrated by full progression of calculations from simple to exact for a particular set of
vehicle parameter values. The same parameter set will also be used to concretely show how
widely sprung mass system behavior differs from unsprung. For this demonstration parameter
values based upon the 1958 Jaguar XK150S will be used; this is a matter of convenience as the
XK150S was previously studied by this author in the course of developing a computer simulation
model 7. However, those values will be “averaged” to represent a “special case” of equal spring
rates and mass distribution all around, for which case the quarter car model is most appropriate.

Specifically, the values to be used are: ms = 1.907 lb-sec2/in (736.35 lb, 333.95 kg) 8, ks =
117.6 lb/in (21.00 kg/cm) 9, mus = 0.336 lb-sec2/in (129.90 lb, 58.91 kg) 10, and kt = 2266.5 lb/in
(404.68 kg/cm) 11. The sprung mass equations are to be used to solve for the natural frequency
with increasing levels of accuracy. First, the simple equation for the sprung “bounce”
(resonance) Eq. 2.1a produces:

𝟏 𝟏
𝒇𝒔 =
𝟐𝝅
�(𝒌𝒔 ⁄𝒎𝒔 ) = 𝟐𝝅 �𝟏𝟏𝟕. 𝟔⁄𝟏. 𝟗𝟎𝟕

= 𝟏. 𝟐𝟒𝟗𝟖 𝒄𝒑𝒔, 𝒐𝒓 𝟕𝟒. 𝟗𝟗 𝒄𝒑𝒎

7
Reference [20], pages 39-40.
8
This represents a sprung curb weight of 2770.4 lbs plus a 175 lb driver aboard; a simple sprung curb weight
would not represent a configuration of interest as vehicles seldom operate with no one on board. The curb weight
value includes a full fuel tank of 14 Imperial Gallons of gasoline; some early U.K. publications give a “kerb
weight” with five gallons or less onboard (so be careful, know your sources!).
9
This is an average of Jaguar front (106.4 lb/in) and rear (128.8 lb/in) spring rates “at the wheel” (i.e., “effective”).
10
This is an average of the Jaguar front (106.6 lb) and rear (153.2 lb) unsprung weights per wheel.
11
This is an average of the Jaguar front (2133 lb/in @ 40 psi) and rear (2400 lb/in @ 45 psi) tire spring rates.

5
Rev B SAWE Paper No. 3521
Category Number 31.0
Next, for the intermediate equation the combined spring constant must be found via Eq.
2.4:
𝒌𝒄𝒔 = (𝒌𝒔 × 𝒌𝒕 )⁄(𝒌𝒔 + 𝒌𝒕 )

= (𝟏𝟏𝟕. 𝟔 × 𝟐𝟐𝟔𝟔. 𝟓)/(𝟏𝟏𝟕. 𝟔 + 𝟐𝟐𝟔𝟔. 𝟓)

= 𝟏𝟏𝟏. 𝟖 𝐥𝐛/𝐢𝐧, 𝐨𝐫 𝟏𝟗. 𝟗𝟔 𝐤𝐠/𝐜𝐦

Then the intermediate equation Eq. 2.1b produces:

𝟏
𝒇𝒔 =
𝟐𝝅
�𝒌𝒄𝒔 ⁄𝒎𝒔

𝟏
= �𝟏𝟏𝟏. 𝟖⁄𝟏. 𝟗𝟎𝟕
𝟐𝝅

= 𝟏. 𝟐𝟏𝟖𝟔 𝐜𝐩𝐬, 𝐨𝐫 𝟕𝟑. 𝟏𝟐 𝐜𝐩𝐦

Last, going to the even higher level of the exact equation, there is the matter of the
interconnecting relation Equation 2.6:

𝒌𝒕 𝒙 = (𝒎𝒖𝒔 ⁄𝒎𝒔 )��𝒌𝒔 × 𝒌𝒕 (𝟏 − 𝒙)���𝒌𝒔 + 𝒌𝒕 (𝟏 − 𝒙)��

𝟐𝟐𝟔𝟔. 𝟓𝐱 = (𝟎. 𝟑𝟑𝟔/𝟏. 𝟗𝟎𝟕)((𝟏𝟏𝟕. 𝟔 × 𝟐𝟐𝟔𝟔. 𝟓(𝟏 − 𝐱))/(𝟏𝟏𝟕. 𝟔 + 𝟐𝟐𝟔𝟔. 𝟓(𝟏 − 𝐱)))

The iterative solution for “x” is obtained:

𝐱 = 𝟎. 𝟎𝟎𝟖𝟔𝟖𝟕

The substitution of that “x” value into the exact sprung mass frequency equation Eq. 2.1c
produces:
𝟏 𝟏
𝒇𝒔 =
𝟐𝝅
�(𝒌𝒕 𝒙⁄𝒎𝒖𝒔 ) = 𝟐𝝅 �(𝟐𝟐𝟔𝟔. 𝟓 × 𝟎. 𝟎𝟎𝟖𝟔𝟖𝟕⁄𝟎. 𝟑𝟑𝟔)

= 𝟏. 𝟐𝟏𝟗𝟏 𝐜𝐩𝐬, or 𝟕𝟑. 𝟏𝟓 𝐜𝐩𝐦

Using the same parameter values as for the sprung mass “bounce”, the simple (and only)
equation for the unsprung “dip” frequency, Eq. 2.2a, results in 2.9754 cps or 178.52 cpm (about
2.4 times the sprung “bounce”), while the simple unsprung “hop” (resonance) Equation 2.3a
results in 13.0622 cps or 783.73 cpm, which compares well with the intermediate “hop”
Equation 2.3b results of 13.3967 cps or 803.80 cpm. The exact “hop” equation, which exists but
has not been expounded upon in this paper (see footnote number 6), would give results around
13.4020 cps or 804.12 cpm (about 10.99 times the exact “bounce”). Therefore, this Jaguar
XK150S example substantiates what was said at the start of this chapter about the wide

6
Rev B SAWE Paper No. 3521
Category Number 31.0
separation (factor of ten) in the behavior of the natural sprung (1.2191 cps vs. 13.4020 cps exact;
1.2498 cps vs. 13.0622 simple) and unsprung motions which allows them to be treated as being
independent (simple) without much loss in accuracy.

If instead of using “averaged” values, representing a special case of equal spring rates
and mass distribution all around, actual conditions as existing at particular wheels can be taken
into account. This means the normal static load which bears down on each wheel and the actual
(“effective”) spring rate at each wheel will be used. If that is done for the Jaguar example of
curb weight plus a 175 lb driver the results are as per the following Table 2.1; note the average
of the front and rear values is about the same as the previous “special case” results of equal
(“averaged”) spring rates and mass distribution all around.

Table 2.1 – QUARTER-CAR UNDAMPED SPRUNG & UNSPRUNG MODEL


FREQUENCIES, ACTUAL MASS DISTRIBUTION & SPRING RATES

7
Rev B SAWE Paper No. 3521
Category Number 31.0
Note that when the distinction is made between fore and aft suspensions for the 1958
Jaguar XK150S the rear sprung mass suspension frequency is revealed to be significantly
greater (stiffer) than the front suspension frequency (e.g.: 77.41 cpm > 68.73 cpm). This is not by
happenstance; this is the engineered result of utilizing the real world interaction between front
and rear suspensions to minimize unwanted pitching motion. A higher rear frequency tends to
minimize the magnitude and to diminish the persistence of such pitching motion, producing what
famed automotive dynamicist Maurice Olley called a “flat ride” 12. Why this is the case will be
explained when this paper addresses “Olley’s Rules” (Chapter 13).

Everything so far, and throughout this paper, is predicated on the “effective” spring rate,
and not the actual rates of the springs themselves. Quite simply, this is because a spring as
installed in the suspension system of a car generally works at a mechanical disadvantage, and
tends to be become “softer” at the wheel, as per the following illustration:

Figure 2.3 – EFFECTIVE SPRING RATE

Summation of the moments about “o” gives the relationship of force “F2” to force “F1”,
and similar triangles gives the relationship of deflection “d2” to deflection “d1”:

𝒍𝟏 +𝒍𝟐
𝑭𝟐 = 𝑭𝟏
𝒍𝟏

𝒍𝟏
𝒅𝟐 = 𝒅𝟏
𝒍𝟏 +𝒍𝟐

As can be seen from the relationships, an actual spring rate as measured at the spring
(“F2/d2”) will result in a very different “effective” spring rate as measured at the wheel (“F1/d1”).
Since the force “F1” is less than “F2” and the deflection “d1” is greater than “d2”, the effective
12
Reference [10], pages 611-619.

8
Rev B SAWE Paper No. 3521
Category Number 31.0
spring rate at the wheel “F1/d1” will be a good deal less than the actual spring rate “F2/d2” at the
spring. In fact, if a spring of known spring rate “K2” is installed as shown in Figure 2.3, then the
effective spring rate at the wheel “K1” can be calculated as per the following example:

𝒍𝟏 𝒍
𝑭𝟏 𝑭𝟐 �𝒍𝟏 + 𝒍𝟐 � � 𝟏 � 𝒍𝟏 𝒍𝟏 𝒍𝟏
𝟐
𝒍𝟏 + 𝒍𝟐
𝑲𝟏 = = = 𝑲𝟐 = 𝑲𝟐 � �� � = 𝑲𝟐 � �
𝒅𝟏 𝒍 + 𝒍𝟐 𝒍 + 𝒍𝟐 𝒍𝟏 + 𝒍𝟐 𝒍𝟏 + 𝒍𝟐 𝒍𝟏 + 𝒍𝟐
𝒅𝟐 � 𝟏 � �𝟏 �
𝒍𝟏 𝒍𝟏

The quantity “𝒍𝟏 ⁄ (𝒍𝟏 + 𝒍𝟐 )” could be called the “moment ratio” or the “motion ratio”,
but in either case is symbolized by “MR” and is used to define the relationship between the
suspension spring rate and the suspension spring rate at the wheel:

𝑲𝟏 = 𝑲𝟐 (𝑴𝑹)𝟐 (EQ. 2.7)

There is still the further complication that occurs if the suspension spring is not installed
in an upright position as shown in Figure 2.3, but in an inclined position as is commonly the
case:

Figure 2.4 – EFFECTIVE INCLINED SPRING RATE

If the actual spring rate is “Ko”, then the effective rate “K2” (which can be substituted
into Equation 2.7 above for the rate at the wheel) is:

𝑲𝟐 = 𝑲𝟎 (𝐬𝐢𝐧 𝜽)𝟐 = 𝑲𝟎 (𝐜𝐨𝐬 ∅)𝟐 (EQ. 2.8)

9
Rev B SAWE Paper No. 3521
Category Number 31.0
3 - DAMPING

Damping is a necessary evil; it increases the level of road shock transmission, but without
damping the automotive spring-mass system would be “conservative”. That is, any motion
imparted the vehicle by road surface irregularity would continue unabated, perhaps to be added
to, or made more complex by, successive disturbances.

Damping conventionally comes in four types. First, there is “constant” resistance that
behaves according to the relation:

𝑭=𝑵𝝁
This is damping resulting from friction 13, where “F” is the force resisting motion, “N” is
the normal load pressing the friction surfaces together, and “μ” is the applicable (static or
dynamic) coefficient of friction. The transition between static (“μs”) and dynamic (“μd”) friction
can result in a “sticking” action, which is one of the reasons why friction dampers are not the
modern choice for suspension control.

Then there is resistance that behaves in accord with the relation:

𝑭=𝒄𝑽
This is resistance “F” proportional via the damping coefficient “c” to the flow velocity
“V”, which is typical of viscous fluid behavior. Viscous 14 damping in the form of hydraulic
“shock absorbers” 15 (dampers) is what is generally used on vehicles today, and is the most
significant form of damping for modern vehicles. Complicating matters is that such dampers are
set with much more damping in extension (higher “c”) than in compression (lower “c”) to
minimize shock transference to the sprung mass on bumps. While such settings do minimize the
adverse effect of dampers on shock transference, they also can have a detrimental effect on
maintaining road contact as shall be seen in Chapter 5.

Also, there is resistance that is proportional to the square of the flow velocity as resulting
from gas flow:

𝑭 = 𝒄 𝑽𝟐

13
Also known as “Coulomb damping”, after the great French physicist Charles-Augustin Coulomb (1736-1806)
who got involved with the physics of friction as a byproduct of his study of electrostatic behavior.
14
There is also a form of damping termed “visco-elastic” which is modeled for tires as a damper (visco) “c” and
spring (elastic) “Kc” in series, both of which are in parallel with another spring “Kp Pi”; the “Kc” represents the
uninflated tire carcass stiffness while “Kp” is the stiffness/pressure coefficient: “Kv = Kp Pi + Kc”.
15
“Shock absorber” is a popular misnomer in the US for what is correctly called a “damper” in the UK and
elsewhere.

10
Rev B SAWE Paper No. 3521
Category Number 31.0
This type of damping is not generally useful for automotive purposes, despite the
popularity of so-called “gas shocks” 16. Due to the generally much reduced density of gas with
respect to liquid density at standard temperature and pressure it would take truly enormous gas
dampers to be effective in an automotive application. However, gas damping plays a significant
role in the design of MEMS (Micro-Electro-Mechanical Systems) switches, and the effect of
atmospheric damping should be accounted for in certain empirical measurements such as the
“swinging” of an aerosurface such as a flap or elevator in order to ascertain its rotational
inertia.

Finally, there is the damping caused by internal energy loss though the flexing of
material, i.e., hysteresis damping. This is the second most significant form of damping for
modern vehicles, which results from being equipped with pneumatic tires. Although tire
hysteresis resulting in vibration damping is often neglected for automotive ride calculations, it is
critical in the form of “rolling resistance” for calculations involving acceleration and fuel
economy. When the tire damping is not ignored for automotive ride calculations it is usually
expressed as an equivalent viscous damping coefficient.

Throughout the entire previous chapter damping was neglected. However, that proves to
be a viable simplification as the damped natural frequency of the sprung mass system is not
very different from the undamped natural frequency, at least for conventional automotive design,
as calculated previously per Equation 2.1b:

𝟏
𝒇𝒖𝒏𝒅𝒂𝒎𝒑𝒆𝒅 = �𝒌𝒔 ⁄𝒎𝒔
𝟐𝝅

Which, when the values used are: ms = 1.907 lb-sec2/in (736.35 lb, 333.95 kg) and ks =
117.60 lb/in (21.00 kg/cm), becomes as shown in the previous chapter:

𝟏 𝟏
𝒇𝒖𝒏𝒅𝒂𝒎𝒑𝒆𝒅 =
𝟐𝝅
�(𝒌𝒔 ⁄𝒎𝒔 ) = 𝟐𝝅 �(𝟏𝟏𝟕. 𝟔⁄𝟏. 𝟗𝟎𝟕)

= 𝟏. 𝟐𝟒𝟗𝟖 𝐜𝐩𝐬, 𝐨𝐫 𝟕𝟒. 𝟗𝟗 𝐜𝐩𝐦

Now to get the frequency when damping is included, we simply have to multiply the
undamped frequency by a quantity “�𝟏 − 𝝃𝟐 ”:

𝟏
𝒇𝒅𝒂𝒎𝒑𝒆𝒅 =
𝟐𝝅
�(𝒌𝒔 ⁄𝒎𝒔 ) �𝟏 − 𝝃𝟐 (Eq. 3.1)

16
“Gas shocks” are conventional viscous “shock absorbers” that have been pressurized with a charge of nitrogen
gas to inhibit foaming and cavitation of the working fluid.

11
Rev B SAWE Paper No. 3521
Category Number 31.0

Where “ξ” is called the “damping ratio” (which is dimensionless, no units). The
“damping ratio” just is the ratio of the damping coefficient “c” to a spring-mass system
characteristic “�𝟒 𝒌⁄𝒎”, which is a characteristic of the system called the “critical damping”:

𝝃 = 𝒄⁄�𝟒 𝒌⁄𝒎 (Eq. 3.2)

The effect of various damping ratios on sprung mass system behavior is illustrated by
the following diagram:

Figure 3.1 – DAMPING RATIO EFFECT ON OSCILLATORY MOTION

As an example of the damping effect on automotive suspension frequencies, for the


previous simple sprung mass bounce case assume a “ξ” of 0.3 (conventional automotive
damping ratios tend to range around 0.2 to 0.4); the damped natural frequency would be:

𝟏
𝒇𝒅𝒂𝒎𝒑𝒆𝒅 = �(𝟏𝟏𝟕. 𝟔⁄𝟏. 𝟗𝟎𝟕) �𝟏 − 𝟎. 𝟑𝟐
𝟐𝝅

= 𝟏. 𝟏𝟗𝟐𝟐 𝐜𝐩𝐬, 𝐨𝐫 𝟕𝟏. 𝟓𝟑 𝐜𝐩𝐦


Note that the simple damped frequency is about 95.39% of the previous simple
undamped frequency. Even with a damping ratio of 0.4 the damped frequency would still be
91.65% of the undamped. If the undamped bounce frequencies calculated by the intermediate
equation Eq. 2.1b and the exact equation Eq. 2.1c (back in Chapter 2) are modified by the
quantity “√1-ξ2” with “ξ” of 0.3 the resulting damped frequency values are 1.1625 cps (69.75
cpm) and 1.1621 cps (69.72 cpm), respectively. So, for conventional automotive damping the

12
Rev B SAWE Paper No. 3521
Category Number 31.0
effect on the sprung mass natural frequency is not great, and may be neglected for initial design
studies.

Having discussed the damping ratio “ξ”, a word or two about the damping coefficients
“cs” and “ct” are in order. First, the damping coefficients are dimensionally in “force-
time/length” units. In the English Units System “cs” might have a value of 5.99 lb-sec/in (1049.0
N-sec/m in the metric MKS units system), and “ct” might have a value of 0.56 lb-sec/in (98.0 N-
sec/m) 17. Just as the two spring values, “ks” and “kt”, can be combined to obtain a single spring
rate, so can the two damping values in the quarter car model be combined to form a single
damping coefficient. For the sprung mass both the “dampers”, suspension and tire, act in series
giving rise to the combined damping rate “ccs”, which produces a somewhat “softer” system 18:

𝒄𝒄𝒔 = (𝒄𝒔 × 𝒄𝒕 )⁄(𝒄𝒔 + 𝒄𝒕 ) (Eq. 3.3)


For the unsprung mass the dampers are considered to be in parallel, with the
combination resulting in the combined damping coefficient “ccp”, producing a somewhat
“harder” system 19:
𝒄𝒄𝒑 = 𝒄𝒔 + 𝒄𝒕 (Eq. 3.4)

To illustrate the magnitude of the difference that occurs when the tire damping is
included the sprung mass “bounce” intermediate equation will be employed. In that equation
the combined spring constant (series) is used, so it seems appropriate to use the combined
damping coefficient (series):
𝒄𝒄𝒔 = (𝒄𝒔 × 𝒄𝒕 )⁄(𝒄𝒔 + 𝒄𝒕 )

ccs = (5.99×0.56)/(5.99+0.56)

= 1.9527 lb-sec/in, or 341.97 N-sec/m

With that as the combined damping coefficient, the damping ratio becomes:

ξ = 𝐜𝐜𝐬 ⁄�𝟒 𝐤 𝐬 ⁄𝐦𝐬

= 1.9527 / √4(111.80)/1.907

= 0.1275

This means that when plugged into the intermediate sprung mass “bounce” equation as
“√1- ξ ” the resulting resonance frequency will be 99.18% of the former value, or 1.2086 cps
2

17
Reference [18], page 4. This source states that “ct” can be realistically approximated as being about one-tenth the
value of “cs”, and uses a “ct” of 98.0 N-sec/m (0.56 lb-sec/in) in a ride response simulation.
18
Reference [6], page 32.
19
Ibid, page 30.

13
Rev B SAWE Paper No. 3521
Category Number 31.0
combined damping vs. 1.2186 cps undamped vs. 1.1625 cps when “shock absorber” only
damped.

As was noted earlier, damping tends to increase the magnitude of road shock. The need to
attenuate oscillatory motions conflicts with the need to provide a “soft” ride, which traditionally
resulted in a compromise setting of the damping ratio 20. The matter has traditionally been put in
terms of “soft ride” verses “sporting ride”. The problem of selecting a damping level that will
provide an acceptable “ride” with respect to road shocks while controlling vehicle oscillations to
an extent commensurate with “sporting” performance is compounded by road vibration.
The question of damping involves not only road shocks and the consequent large scale
sprung mass oscillations, but small scale road induced vibrations as well. Colin Campbell, in
one of his books, has a rather neat, though limited, graphic summary of the damping compromise
conundrum regarding vibration 21:

Figure 3.2 – DAMPING EFFECT ON SPRUNG MASS ACCELERATION DUE TO


ROAD VIBRATORY INPUT

Note that when damping is nonexistent, “undamped” or “ξ=0”, the sprung mass
acceleration levels due to road vibration input are low except when that input frequency is near
the natural sprung mass resonance frequency; at that point things go drastically out of control. A
modest amount of damping, “under damped”, brings that acceleration spike down, but it still
may be objectionable; also note that the general level of acceleration at higher frequencies has
risen. A “well damped” system is one that brings the resonance acceleration down to tolerable
levels, but doesn’t bring the general level of acceleration at higher frequencies up too much
except at those higher road speed levels where a little shaking up is not only expected but may
serve a cautionary function. The best performance at resonance is when the system is “over

20
Reference [3], page 212.
21
Reference [3], page 211.

14
Rev B SAWE Paper No. 3521
Category Number 31.0
damped”, “ξ >1”, but that causes too high an increase most everywhere else. As noted, this
summary is limited in that it doesn’t even address what happens around the unsprung mass
resonance frequency, “fr/fs ≅ 10”; that adds even more complication and will be dealt with in
Chapter 6, but this does serve to further the awareness of the complexity of compromise in
selecting a suspension damping level for a traditional passive system.

A solution to the necessity of compromise began to appear with the dawn of adjustable
“shocks”. However, as shall be made even clearer in Chapter 6, the matter is more complex than
just increasing the damping coefficient when driving fast and decreasing it when just cruising.
An early improvement resulted in “shock absorbers” with adjustable damping coefficients. A
dial on the outside of these dampers allowed for the adjustment of the size of the fluid flow ports.
Such adjustable “shock absorbers”, of which Koni and Bilstein are famous examples, can be set
just before a particular use as being more to the “comfort” or to the “control” side of things, but
after that the driver is stuck with those settings until the “shocks” can be readjusted again.

A more advanced system could be called “on-the-fly adjustable” which allows for the
adjustment of the damping coefficient from the cockpit while in motion. A 1930’s example of
this type of damper control would be the André Telecontrol, known as the Hartford Telecontrol
system in the US 22. This was a friction damper based system, and initially a Bowden cable
allowed the driver to adjust the pressure between the friction disks from the dashboard (later
versions would use hydraulics to make the adjustment) which modified the damping coefficient.

A more recent example of “on-the-fly adjustable” would be the EVSA (“Electronic


Variable Shock Absorber”) system introduced on the 1982 Mazda 626. This system had an
electric solenoid mounted at the top of each hydraulic damper; the solenoid rotated a control rod
which varied the size of the fluid ports in the damper (much like the “old” hand adjustable). The
solenoid is controlled essentially from an electric switch in the cockpit which has three settings:
“normal”, “auto”, and “sporty”. “Normal” and “sporty” are self-explanatory; on “auto” the
dampers are switched from “normal” to “sporty” and back automatically depending on whether
or not the vehicle is exceeding 50 mph (80.5 kph).

Admittedly, that automatic switching between “normal” and “sporty” settings is a step
toward the next stage in damper control, which is the “adaptive” 23 system. Such a system
continually varies the damper settings in accordance with a complex computer algorithm that
uses a number of parameters as input for optimum matching of damping to all conditions. Toyota
came close to fully adaptive damping control with its TEMS (“Toyota Electronic Modulated
Suspension”) system in 1983. The TEMS is similar but more sophisticated than the Mazda
EVSA in that it doesn’t just use vehicle speed as it only control input parameter; the Toyota
TEMS also utilizes throttle opening, steering wheel angle, and braking. However, in “auto”
mode it uses all that data to still only switch back and forth between two damper settings

22
Reference [2], page 120. The 1934 Bugatti Type 57 and the 1935 Jaguar SS-90 both had some version of the
André Telecontrol system.
23
“Adaptive” is a term that is often confused with “active”, but generally “active” designates an even higher degree
of suspension sophistication wherein both damper and spring functions are under computer control. This may mean
that the conventional springs and dampers are completely replaced by actuators in a full active suspension
implementation.

15
Rev B SAWE Paper No. 3521
Category Number 31.0
(“normal” and “sport”) with a 0.1 second cycle time. A true fully adaptive system would have
infinitely variable damping with a much shorter cycle time and also use road roughness
frequencies as input to the controller.

An early true full adaptive suspension system was developed by Automotive Products
Limited for the Rover Group PLC 24 circa 1980. The AP system used gas spring/hydraulic
damper units at each wheel controlled by hydraulic actuators. The actuators were powered by a
hydraulic pump that consumed 13.4 hp (10 kW) of the engine power at maximum output to
produce high volume flow at 2000 to 2500 psi (13789 to 17137 kPa). To decrease the need for
full output occurrences, during low demand times the pump could store hydraulic energy in a
large accumulator that was an essential part of the system 25.

The whole system of hydraulics, sensors, and computer was very quick acting and
effective in controlling vehicle ride height, roll, dive, and squat. When fitted to a Rover model
3500 sedan the vehicle was able to negotiate a slalom at 54.6 mph (87.9 km), as opposed to the
stock Rover 3500 speed of 48.5 mph (78.1 km). Nevertheless, because of cost, weight, and
power concerns the system was not deemed ready for production 26.

Many other manufacturers have developed some sort of system for computerized control
of suspension damping, often combined with control of other suspension aspects. The day of
compromise in suspension settings, and much else, may be largely coming to an end, mainly due
to advances in electronics. However, what has been holding back this dawning of a new age has
been cost, reliability, weight penalty, and energy consumption. Still, progress is continually
being made.

For instance, the most recent way for varying the damping coefficient is to use an electro-
rheological (ER) working fluid in the “shock absorber”. An electro-rheological fluid is one that
changes its physical characteristics in accord with an electrical stimulus. Modifying the damping
in this way is simpler (decreased cost, greater reliability, and less weight penalty), much less
energy consumptive (only uses a few watts of electric power), and allows for a faster response
time. An even more recent take on this matter has been concepts that involve using electrical
generation of energy via suspension motion to not only dampen oscillation but to recapture
energy so as to improve vehicle efficiency (very important for hybrid and electric vehicle
design).

So far, the fundamental significance of mass in a spring-mass system has not been a
revelation, but the fact that elements of the system have variance in behaviors great enough to
allow for essentially independent modeling, with consequent simplification in analysis, should
have been interesting. And, in moving away from the simplest independent models toward the
more complex interrelated reality, there has been a glimpse of the role the “unsprung-to-sprung
mass ratio” plays as a moderator between element models, and there is a great deal more mass
properties involvement to come…

24
Private Limited Company, a British legal classification for certain businesses.
25
Reference [2], pages 181-182.
26
Ibid, page 183.

16
Rev B SAWE Paper No. 3521
Category Number 31.0
4 – ROAD SHOCK

The primary function of any suspension system is the “isolation” of the sprung mass from
road shock. To consider the role of mass properties with respect to shock attenuation note the
“quarter-car model” as it approaches a 2 inch step in the road at some velocity “V”:

Figure 4.1 – QUARTER CAR MODEL & ROAD STEP

Encountering the 2 inch step will cause the wheel to move upward 2 inches, compressing
the spring 2 inches from its static position. This extra compression of the spring will result in a
net force inflicting an upward acceleration on the sprung mass. The sprung mass will
accelerate upward until the spring recovers the 2 inches of extra compression, but then, due to
the momentum established, the sprung mass will continue with decreasing acceleration as it
overshoots the original static position with respect to the wheel by another 2 inches. It will then
accelerate downward 4 inches, and then reverse its motion once again, establishing a sinusoidal
pattern of motion called Simple Harmonic Motion (SHM) with, in this case, an amplitude “A” of
2 inches. Note that this not only neglects damping or any effects from the unsprung-mass / tire-
spring system, but also that the sprung mass is assumed to have no vertical motion prior to the
encounter with the step (“damped” out) 27.

The total vertical movement of the sprung mass as set into motion by the step is 4
inches, but the undamped SHM amplitude (maximum extension or compression of the spring
from the static or rest position) “A” is 2 inches. The equations of SHM 28 state that the “period”,
or time for one complete oscillation, “τ” is 29:

27
Hence the significance of rapidly damping out any motion before the next bump adds to the disturbance of the
suspension caused by the previous bump.
28
Reference [15], pages 256-267.
29
Shock due to a bump in the road is generally a single wheel at a time event. This means that the anti-role bar, if
the vehicle is so equipped, will add to the suspension spring effective stiffness per wheel “ks” in accord with
Reference [2], pages 70-72 & 209-210. It is that combined spring stiffness that should be used here and for any other
single wheel motion.

17
Rev B SAWE Paper No. 3521
Category Number 31.0

𝝉 = 𝟐𝝅�𝒎𝒔 ⁄𝒌𝒔 (Eq. 4.1)

This period is a characteristic of the free oscillation of the spring/sprung-mass system,


of which the system’s natural frequency is the reciprocal. This brings us back to simple sprung
mass frequency equation Eq. 2.1a from Chapter 2:

𝟏
𝒇𝒔 = 𝟏⁄𝝉 = �𝒌𝒔 ⁄𝒎𝒔
𝟐𝝅
The maximum acceleration “amax” that the sprung mass will endure is expressed as 30:

𝒂𝒎𝒂𝒙 = (𝟐𝝅 𝒇𝒔 )𝟐 𝑨 (Eq. 4.2)

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

𝒂𝒎𝒂𝒙 = (𝒌𝒔 ⁄𝒎𝒔 )𝑨 (Eq. 4.3)

So, using the example Jaguar XK150S again, the effective spring constant (rate) at the
wheel will be taken as 117.60 lb/in, and the sprung mass is taken as 1.907 lb-sec2/in (736.35 lb,
333.95 kg). Therefore, the system characteristics are:
τ = 2π √1.907/117.6 = 0.8001 sec

fs = 1/τ = 1/0.8001 = 1.2498 cps, or 74.99 cpm

Note the system period is indicative of how fast the response time of any automatic
damping control system must be in order to be effective, and here a maximum of 0.08 seconds
(one tenth the period) cycle time is indicated. Anyway, the maximum acceleration inflicted on
the sprung mass is:

amax = (ks/ms) A = (117.6/1.907) 2 = 123.33 in/sec2, or 0.32 g’s 31

Colin Campbell, in his book The Sports Car, works through a similar example problem,
and obtains values of 0.67 sec, 90.0 cpm, and 0.45 g’s, respectively 32. However, Mr. Campbell
takes a slightly different tack in his discussion of road shock as he has a different point to make.
His point is that a “softer” (decreasing “ks” and/or increasing “ms” 33) mass/spring system will
decrease road shock, but will also increase the static deflection “ds” (total wheel travel must be
physically allowed for, and generally is around twice “ds”) and decrease the stiffness

30
SAE 670e.
31
The value of gravity is taken as 386.088 in/sec2 for calculations in the English IPS units system.
32
Reference [3], page 190. If only the period for Campbell’s example were given, the equivalent shock in “g’s” for
Campbell’s suspension can quickly be determined if “τ1 > τ2” by “(τ1/τ2)2 a1 = a2”: (0.80/0.67)2 0.32 g’s = 0.45 g’s.
33
Of course, increasing the sprung mass “ms” just to get a gentle ride is not a good idea, as has already been noted.

18
Rev B SAWE Paper No. 3521
Category Number 31.0
(proportional to “fn”) to levels that ultimately will not be sustainable. He makes his point via the
following table (he shows only columns 3, 4, 5, and 7):

Table 4.1 – SPRING CONSTANT, STATIC DEFLECTION, & ROAD SHOCK

Campbell states that an initial deflection greater than 6 inches is likely to result in more
suspension travel than could be accommodated on a low riding sports car, and that a frequency
lower than 77 cpm would likely result in excessive roll in cornering and a generally “spongy”
handling. For sports cars he recommends a frequency of about 80 cpm front and 90 to 100 cpm
rear 34.
It has been noted that increasing the sprung mass “ms” will cause the same effect on
shock as decreasing the spring constant “k2” per the simple formula of Eq. 4.2. However, what
about the effect of increasing unsprung mass “mus” on shock transmission? To see the effect of
“mus” on shock transmission to the sprung mass the exact sprung mass frequency equation Eq.
2.1c from Chapter 2 must be employed, as that equation (or model, see Figure 2.2) includes the
interaction between the sprung and unsprung masses. Using Eq. 2.1c, and holding all else
constant while varying “mus”, we modify Table 4.1 to obtain 35:

Figure 4.2 – UNSPRUNG MASS & ROAD SHOCK

34
However, remember the 1958 Jaguar XK150S actually had about 69 cpm front and 77 cpm rear!
35
Increasing unsprung mass will cause increasing deflection of the tire spring “kt”, but that deflection and all
effects that might result therefrom have been considered negligible.

19
Rev B SAWE Paper No. 3521
Category Number 31.0
The unsprung mass (weight) was increased by 175% to cause an indicated decrease in
shock to the sprung mass of about 0.08%! This indicates that increasing the unsprung mass
may decrease shock by virtue of the increased inertia acting to create a time lag, but the effect is
too slight to be of any use. And, contrary to this indication, if we look at the suspension model in
other situations (contact in road dip, vibration, gyroscopic reactions, etc.) it becomes obvious
that reducing unsprung weight is always the desirable goal.
Any discussion of road shock would be incomplete without some discussion of the oldest
method of all for reducing the shock incurred by road irregularities, which is simply to make the
wheels as large as possible in overall diameter. Donald Bastow 36, in his book Car Suspension
and Handling, presents an empirically derived plot 37 of longitudinal distance vs. vertical distance
for two different rigid wheels, both encountering a 1 in (0.025 m) step function at a constant
speed. The wheels are of 0.25 m (9.8 in) and 0.37 m (14.57 in) in radius and, since traveling at
constant speed makes the longitudinal distance axis also an implicit time axis, it is apparent that
the larger radius wheel imparts a lower average vertical acceleration.
From such scanty information this author attempted to extract much more info by
constructing two computer simulation programs of rigid (“RIGIDWHL.BAS”) and pneumo-
elastic (“FLEXIWHL.BAS”) wheels; the Bastow plot data being used as “validation” for the
rigid wheel simulation. Said programs are of admittedly questionable nature as they are based
mainly on the kinematics of the situation and not on a proper dynamic modeling. However, the
rigid wheel program did almost exactly reproduce the Bastow plot at an input velocity of 10 mph
(16.1 kph), and both programs produce various other data plots that seem quite reasonable; for
the interested more details concerning these programs are given in Appendix B.

Figure 4.3 – COMPUTER DUPLICATION OF BASTOW EMPIRICAL DATA

36
Donald Bastow is an engineer of some renown who worked with Maurice Olley at Rolls Royce in the late 1920’s.
37
Reference [1], page 4.

20
Rev B SAWE Paper No. 3521
Category Number 31.0
The point to all this is that the program “RIGIDWHL.BAS” was used to generate data to
illustrate how the vertical acceleration, i.e. the “shock”, of a rigid wheel encountering a step
function bump input of constant height varied with variation in wheel radius. While the output
may be suspect as to its absolute validity, there is cause to believe that it is reasonably indicative
of the physical reality. From consecutive runs of the program for rigid wheels of radius 2 in (0.05
m) to 60 in (1.52 m) and velocities from 10 mph (16.1 kph) to 40 mph (64.4 kph), which is about
the absolute maximum for horse drawn wheeled transport, the following figure was generated:

Figure 4.4 – ROAD SHOCK vs. RIGID WHEEL RADIUS


It’s easy to see why early horse drawn coaches of nonexistent to rudimentary suspension
utilized large wheels, generally 30 in (0.76 m) to 60 in (1.52 m) in diameter, although at times
exceeding 80 in (2.03 m). The chief limit to wheel size was the ability to withstand side loads,
although there is also a “law of diminishing returns” as decreasing shock levels come to be offset
by increasing weight and rotational inertia, but that would come to play more of a role in the
modern era.

21
Rev B SAWE Paper No. 3521
Category Number 31.0
5 - ROAD CONTACT

Another important function of the suspension is to keep the tires in firm contact with the
ground; in this regard there is an oft-encountered condition of a dip in the road surface as
depicted in the following illustration:

Figure 5.1 – ROAD CONTACT IN DIP

It’s easy to see that if the unsprung-mass/spring 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
may be lost. The road surfaces of the real world are often a series of such undulations,
interspersed with bumps and other irregularities. The question for the 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 spring-unsprung mass
system the following relationships may be derived 38:

𝒍 = (𝑽𝝉⁄𝟐) �𝒎𝒖𝒔 ⁄𝒎𝒔 (Eq. 5.1)

𝒅 = 𝒅𝒔 (𝟏 + (𝒎𝒖𝒔 ⁄𝒎𝒔 ) ) (Eq. 5.2)

Note again the appearance of the unsprung-to-sprung mass ratio “χ” (“mus/ms”); as
promised this is a key parameter that will appear frequently throughout any study of suspension
performance. Generally the smaller this value is then the better the performance is. Essentially,
the situation is that previously we have calculated parameters for the “ks-ms” system, but now
we are interested in the behavior of the “mus”, and since the “ks-ms” and “ks-mus” systems share
the same spring the parametrics vary only by the ratio “mus/ms”.

38
Reference [3], page 193.

22
Rev B SAWE Paper No. 3521
Category Number 31.0
As an example of how the previous formulae might be utilized, let’s assume a situation
where a vehicle of 920 lbs (417.3 kg) sprung and 145 lbs (65.8 kg) unsprung mass approaches a
dip at 30 mph (44 ft/sec, 48.3 kph, 13.4 m/sec). This model is the same as that initially utilized
for the road shock investigation, with an effective spring stiffness of 193.3 lb/in (34.5 kg/cm) at
the wheel, resulting in a period of oscillation of 0.6976 seconds and a static deflection of 4.76 in
(0.121 m). Therefore, in order for the wheel to keep in contact with the road, the length of the dip
can be no narrower than:

l = (44 ft/sec 0.6976 sec/2) √145/920 = 6.1 ft, or 73.1 in (185.7 cm)

And the depth must be no greater than:

d = 4.76 in (1 + 145/920) = 5.51 in (14.0 cm)

Now compare this with the results for a corresponding vehicle for which everything is the
same except the unsprung mass is one-half of previous:

l = (44 ft/sec 0.6976 sec/2) √72.5/920 = 4.3 ft, or 51.7 in (131.3 cm)

d = 4.76 in (1 + 72.5/920) = 5.13 in (13.0 cm)

While these results are basically good, the length and the depth results are moving in
opposite directions: there is a 29.3% improvement in “l” accompanied by a 6.7% worsening in
“d”. This conflict indicates a possible area of design compromise; the minimization of “l” runs
contra the maximization of “d”. A better understanding of how this works may be obtained by a
consideration of the original quarter-car model parameters as the unsprung weight is varied
about the nominal:
Road Contact Parameters as Unsprung Weight Varies Range Range
Quarter Model, General SHM Equations, at 30 mph: 4.3 1.13
1 2 4 3 5 6 7 8

Ws Wus chg k ds fn T min l max d


lb lb %Wus lb/in in cpm sec (ft) (in)
920 73 -50% 193.3 4.8 86.04 0.697 4.3 5.13
920 80 -45% 193.3 4.8 86.04 0.697 4.5 5.17
920 93 -36% 193.3 4.8 86.04 0.697 4.9 5.24
920 106 -27% 193.3 4.8 86.04 0.697 5.2 5.31
920 117 -19% 193.3 4.8 86.04 0.697 5.5 5.37
920 145 0% 193.3 4.8 86.04 0.697 6.1 5.51
920 173 19% 193.3 4.8 86.04 0.697 6.6 5.65
920 184 27% 193.3 4.8 86.04 0.697 6.9 5.71
920 197 36% 193.3 4.8 86.04 0.697 7.1 5.78
920 210 45% 193.3 4.8 86.04 0.697 7.3 5.85
920 290 100% 193.3 4.8 86.04 0.697 8.6 6.26

Table 5.1 - VIABLE DIP LENGTH & DEPTH vs. UNSPRUNG WEIGHT

23
Rev B SAWE Paper No. 3521
Category Number 31.0
Note that the best “l” and the best “d” occur at opposite ends of the variation. The effect
can be even better understood visually:

Figure 5.2 – VIABLE DIP LENGTH & DEPTH vs. UNSPRUNG WEIGHT

It is easily seen that decreasing the unsprung weight has a beneficial effect on “l” while
having a detrimental effect on “d”. However, the tradeoff is not proportional; the “l” benefit is
greater than the “d” detriment. The decline in “d” has a shallow slope, and the overall verdict is
that the decreasing of the unsprung weight is generally beneficial for maintaining road contact.

It should be remembered that firm tire to road contact is essentially assured only up to the
extent of the static deflection, which in this case is still 4.76 in (0.121 m). Any contact beyond
that is due to the dynamic “overshoot” of the fully extended spring position by the unsprung
mass (and some recovery of the normal deflection of the tire), and such contact is likely to be
much more tenuous.

Now that we have seen how road contact parameters “l” and “d” vary with unsprung
mass “mus”, how those parameters vary with spring stiffness “ks” becomes the next matter of
interest. How stiffness variation affects road contact may be explored from the following
manipulation of the original quarter-car model parameters wherein we vary the spring stiffness
about the nominal:

24
Rev B SAWE Paper No. 3521
Category Number 31.0

Table 5.2 – VIABLE DIP LENGTH & DEPTH vs. SPRING STIFFNESS

Again, what is happening is best understood visually:

Figure 5.3 - VIABLE DIP LENGTH & DEPTH vs. SPRING STIFFNESS

Note that once again the effects on the road contact parameters are counteracting, only
now “d” is more affected than “l”. Now let’s return to the previous modification of the baseline
model’s parameters wherein we had reduced the unsprung weight by half to get a big
improvement in “l” at 4.3 ft (down from 6.1 ft) and a small deterioration in “d” at 5.13 in (down
from 5.51 in). To compensate for this slight loss in firm tire to road contact when encountering
dips of depth greater than 5.13 in let’s “tweak” the spring rate “ks” by – 6.9 %:

l = (44 ft/sec 0.7229 sec / 2) √72.5/920 = 4.5 ft, or 54.0 in (137.16 cm)

d = 5.11 in (1 + 72.5/920) = 5.51 in (13.99 cm)

25
Rev B SAWE Paper No. 3521
Category Number 31.0
Note that now “d” is as good as it was in the original design at 5.51 inches, and “l” is
only slightly worse (+4.6%) than it was before tweaking the spring stiffness by -6.9%, but still
much better (-26.2%) than the original road contact parameter “l”. The chief concern about going
to a “softer” spring is excessive wheel travel and body motion, but if necessary those conditions
can be mitigated by “rising rate” springs, although this adds to the expense and complication.
The chief argument for rising rate springs is that they tend to keep carefully chosen frequencies
constant despite the considerable variation in sprung mass from “one-up” to full load 39. Anti-
roll bars, which admittedly do nothing for pitch and yaw control, can be used to mitigate
excessive roll due to “soft” springing, but are mainly used to significantly affect “weight
transfer”/directional stability in roll. However, there are yet further design resources to which
one may appeal (“active” suspensions, etc.), but always at the cost of increased complication &
expense.

So far the following has been shown: reducing the spring stiffness reduces the sprung
mass acceleration in bump (good), while increasing both “d” (good) and “l” (bad) in dip.
Reducing the unsprung mass reduces both “d” (bad) and “l” (good) in dip. Complicating matters
is the fact that generally there is only so much reduction of the unsprung mass possible; it is
easier to “tweak” the spring rates to achieve desired road contact performance. And, of course,
the “firmness” of the “shock absorbers” in extension gets involved in this too; shocks that are too
firm in extension will inhibit the maintenance of contact with the ground 40. Exactly what “l” and
“d” values to seek, and thereby the values of all other involved vehicle parameters, ultimately
depends on the vehicle concept; what sort of vehicle is it supposed to be and over what sort of
roads at what speeds is the intended use.

Speaking of the effect on sprung mass acceleration of the unsprung mass in bump
naturally begs the question of what is the effect in dip (although shock to the sprung mass in dip
is not the concern that it is in bump). The sprung and unsprung masses are linked together by
the spring (ignoring the damper for now), and any force exerted on or by the spring is equally
exerted in both directions. So, when the unsprung mass is free to accelerate downward due to a
sudden dip in the road:
Fspring = mus aus

Fspring = ms as
Therefore:
mus aus= ms as
Or:
as = aus (mus/ms)
39
This variation tends to be more serious for a light car than for a heavy one. For very small “economy” class
vehicles compensating for load variation can be difficult; see Reference [14].
40
Again, including the damping values will have only a minor effect on the dip parameters “l” and “d”, but those
values can be included for exactitude. The important note here is that adjusting the damper balance so the
greatest damping is on extension is good for ride, but for road holding the greatest amount of damping should
be on compression. This is another way, other than just simple magnitude variation, that damping can be
actively controlled.

26
Rev B SAWE Paper No. 3521
Category Number 31.0
However, in dip for particular spring stiffness, “aus” is solely driven by the magnitude of
“mus” so a reduction in “mus” will be exactly canceled out by a corresponding increase in “aus”;
therefore, in dip the benefit of a low unsprung mass with regard to sprung mass acceleration is
nil.

However, it has been that shown reducing unsprung mass is good for maintaining road
contact and, as we shall see, good in number of other ways as well. Those beneficial effects are
maximized if a reduction in unsprung mass “mus” can come about from just a reduction of the
rolling weight portion of the unsprung mass due to considerations of acceleration and braking
(reduction in rotational inertias means less “effective weight”), and also from the standpoint of
reduction of gyroscopic reactions (Chapter 7).

27
Rev B SAWE Paper No. 3521
Category Number 31.0
6 – ROAD VIBRATION

So far, the discussion of the suspension’s primary function in attenuation of road


disturbance input has been limited to the relatively simple matter of road shock. This simplicity
is due to the fact that road shock tends to consist of a large scale discrete event, but there is
another far more complex type of road disturbance: road vibration. Road induced vibration
results from the general roughness of the road surface which presents itself to a moving vehicle
as an infinite series of small scale events which have an adverse effect on ride quality.

Human perception of ride is mainly tactile, which involves vibrations in the range of 0-25
Hz. Above that range vibration is perceived as noise; human aural sensitivity involves vibrations
in the 25–20,000 Hz range. A vehicle that does not adequately attenuate road vibration input may
not only inflict unpleasant sensation, blur vision, and create noise, but in extreme cases can
inflict damage, if not to the passenger then to the vehicle and/or its cargo. Extreme cases tend to
occur on those occasions when there is a coincidence of natural frequency and frequency of
excitation resulting in a condition known as “resonance”.

Although the general road surface roughness is the major cause of ride vibrations there
are other possible causes: drive train rotational imbalance/dimensional & stiffness variations/
misalignment, engine torque flux, and aerodynamic buffeting. The main objective of the
suspension system is still to attenuate the effect of such disturbances on the sprung mass, but in
the case of vibration this goal is principally achieved by avoiding resonance. Road vibration
input must be modeled as an almost infinite number of continuous sine wave “mini-shock”
functions all varying in amplitude, frequency, and phase; so the matter can be dealt with only by
statistical means.

Road surface excitation is treated as a broad band random input to the vehicle suspension
system best described by its PSD (Power Spectral Density). Road surface PSD is determined
by measuring road surface roughness, then modeling that raw surface data through
transformation by a Fourier Transform process into an immense set of overlapping sine wave
functions varying in amplitude, frequency, and phase. PSD is the plot of Spectral Density
(in2/cycle/ft, a measure of power which corresponds with amplitude “A”) vs. Spatial Frequency
(cycle/ft, “Ω”), which is also known as “wave number” (the inverse of wavelength: ft/cycle,
“λ”). Each type of road surface (concrete, asphalt, etc.) has a unique PSD. All road surfaces
show a decreasing Spectral Density (power, amplitude) with increasing Spatial Frequency (wave
number, decreasing wavelength). This is because small amplitude waves tend to have small
wavelengths, and vice versa.

28
Rev B SAWE Paper No. 3521
Category Number 31.0
The following figure shows specific measured PSD’s for three actual roads, and averaged
PSD’s for two types of road (concrete and asphalt). Note that the units for the roughness spectral
density are in units of vertical variation squared per spatial frequency. This has to do with the
fact that the measured variation of a road is a random plus or minus flux for which a simple
arithmetic sum would tend toward a total of zero with increasing sample size, erroneously
indicating perfect smoothness. So, squared roughness (vertical variation) values are used to get a
true measure of the road roughness, but also are convenient for determinations of the relative
power (length-force per time) at each frequency.

Fig. 6.1 – ROAD ROUGHNESS POWER SPECTRAL DENSITIES

Note also the average concrete road has more power in the high wave number (short
wavelength, high frequency due to “V = f λ”) whereas the average asphalt road has the greater
power in the low wave number (long wavelength, low frequency). These average curves are
often modeled as “A = 1/f2” or “A = 1/f4”. The plots/curves change slightly from road to road,
rising with increasing roughness and dropping as the road in question is smoother, but the basic
shape(s) tends to stay the same.

29
Rev B SAWE Paper No. 3521
Category Number 31.0
The PSD for the average concrete or asphalt road can be approximated by 41:

Gz(Ω) = G0 [1+(Ω0/Ω)2] / (2πΩ)2 (Eq. 6.1)


Where:

Gz(Ω) = PSD amplitude (ft2/cycle/ft) as a function of “Ω”.

Ω = Spatial frequency, a.k.a. wave number (cycle/ft).

G0 = Roughness magnitude (ft2/cycle/ft):


1.25 x 105 (smooth) to 1.25 x 106 (rough).

Ω0 = Limit spatial frequency, a.k.a. cut-off wave number:


.05 cycle/ft (asphalt), .02 cycle/ft (concrete) 42.

A random number generator for “Ω” in conjuction with the above equation will generate
a PSD typical of certain roads. Once a PSD has been generated, there can be a PSD conversion
from the spatial domain to the more useful time domain. The time domain is more useful because
the most significant measure of ride quality is the level of vertical accelerations (d2z/dt2)
experianced by the vehicle passengers. Also, as a vehicle moves with velocity “V” over a road
surface the roughness constitutes a “traveling wave” (w.r.t. the vehicle), causing a vertical
motion which is more of a “standing wave” (w.r.t. the vehicle) that can best be dealt with as a
function of time.

To make this conversion to the time domain a particular vehicle velocity (“V”) of interest
must be identified. Then the velocity, here expressed in units of “feet/sec”, is used as a multiplier
times every PSD point to convert the Spatial Frequency (cycles/ft) to Temporal Frequency
(cycles/sec):

feet/sec x cycles/ft = cycles/sec = Hz


And to convert the Spectral Density (amplitude, power) in units of “in2/cycle/ft” into
vertical distance (squared) per temporal frequency (cycles/sec) a similar multiplication is
performed:

feet/sec x in2/cycles/ft = in2/cycles/sec = in2/Hz


Once this is done the resulting transformed PSD is essentially amplitude (elevation)
squared per frequency vs. frequency. This PSD can be differentiated (d/dt) once to obtain
41
Those involved in the measurement of road surfaces often utilize the standard ISO 8608 “Mechanical Vibration –
Road Surface Profiles – Reporting of Measured Data”, which is very much like the UK standard BS 7853. Such
standards play a key role in the establishment of quantitative values such as those limit values given for “G0” and
“Ω0”.
42
Reference [7], p. 129. This same reference also gives the extreme range of spatial frequency as being from “0.5
cycle/foot” to “0.005 cycle/foot” (p. 128)!

30
Rev B SAWE Paper No. 3521
Category Number 31.0
essentially what is velocity (vertical) per frequency vs. frequency, and differentiated (d/dt)
again (thus obtaining “d2/dt2”) 43 to essentially get road acceleration (vertical) per frequency vs.
frequency:

Fig. 6.2 – ROAD VERTICAL ACCELERATION INPUT SPECTRAL DENSITY

Now that the road vertical acceleration input for each frequency at a particular vehicle
velocity (“V”) has been obtained this process may be repeated at suitable velocity intervals to
cover the entire vehicle velocity range. These accelerations can be multiplied by the appropriate
transmissibility factor for a particular model to get the resulting sprung mass acceleration. Note
that there will be as many mass-spring models, and corresponding transmissibility factors, as
there are load conditions from one-up to GVWR. Also note that the greatest accelerations occur
at the greatest frequencies, and, as we shall see, at the greatest vehicle speeds. It is one of the

43
“Noisy” data such as depicted in Fig. 6.2 may be differentiated by any one of a number of numerical
differentiation methods of varying sophistication. The simplest method is that of “finite differencing”, but that
method is reputed to often “amplify” the noise to the extent that the result is useless. The exact differentiation
method used to get an acceleration PSD such as is shown is not specified by Reference [7]. Reference [6] assumes
an underlying sinusoidal relation and explicitly differentiates that relation to arrive at a result essentially the same as
Eq. 6.2. The author assumes (the reference doesn’t so state) that, given a particular “V” and “t”, values from such as
Fig. 6.1 for “Ω” and “A” can be “plugged into” Eq. 6.2 and plotted, resulting in something like Fig.6.2.

31
Rev B SAWE Paper No. 3521
Category Number 31.0
suspension system’s main purposes to attenuate the high frequency input (the other main purpose
is to attenuate shock); it does this by having its resonance point at the low frequency end of the
spectrum.

Since the above process has to be completed for a number of velocity levels, the question
arises what is the general effect on the road roughness acceleration levels as vehicle velocity
increases. The effect is that at any particular frequency (Hz) the acceleration will increase with
the velocity squared. This can be confirmed by considering a simple sine wave model of road
roughness:

Z = A sin(2πΩX)

Where:

Z = Road roughness height.


A = Sine wave amplitude.
Ω = Wave number (cycles/foot).
X = Distance traveled on road.

Remember “X = V t” (distance traveled equals the vehicle velocity times the time
elapsed); substitute for “X” into the original expression and get:

Z = A sin(2πΩVt)
We may differentiate this once (d/dt) to obtain the vertical velocity:

dZ/dt = (2πΩV) A sin(2πΩVt)


And differentiating again obtains the vertical acceleration:

d2Z/dt2 = -(2πΩV)2 A sin(2πΩVt) (Eq. 6.2)

Therefore, as the vehicle velocity increases the “Acceleration” PSD (g2/Hz) vs.
Frequency (Hz) plot line will shift upward (but also slightly to the left because each wave
number in the original PSD spatial domain plot will now represent a greater frequency in the
time domain plots); this shift in the acceleration plot will look something like this (exaggerated):

32
Rev B SAWE Paper No. 3521
Category Number 31.0

Fig. 6.3 – ROAD VERTICAL ACCELERATION INPUT CHANGE (EXAGGERATED)


WITH INCREASING VEHICLE VELOCITY

It should be remembered that road roughness not only causes vertical motions and pitch
motions, but roll motions as well 44. All of these motions: vertical, pitch, and roll; translate into
X, Y, Z accelerations at head level which are the most detrimental results and so must be
minimized. Automotive vehicles generally have less response to roll excitation except in the off-
road case when roll motions require special consideration.

Other than road roughness, other causes of what might be considered poor ride include
imbalance (static and/or dynamic)/misalignment/excessive tolerances/off-center mounting/
elastic deflection/eccentricity of rotating components, dimensional or stiffness variations of tire
/wheel assemblies, and engine torque/inertial pulsations. These are generally quality issues,
with the possible exception of engine excitations which also brings matters of engine and engine
mounting design (exhaust pulse excitation also a matter of design and mounting) into account.

The purpose of a suspension system is to lessen the effect on the sprung mass of the
shock and vibration of the road roughness. The situation is somewhat like that of a radio system,

44
Roll motions are not dealt with in this paper, but the treatment is analogous to that of pitch motions.

33
Rev B SAWE Paper No. 3521
Category Number 31.0
only in reverse. The radio receives a weak input signal which the system must amplify in order to
be played aloud, i.e., to be sufficiently strong to stimulate the human senses. The ratio of output
to input is the “gain” of the system, which for a radio system must be significantly greater than
unity (1). For an automotive suspension system the gain or transmissibility had better be a good
deal less than unity, else it would not be doing a very good job of “isolating” the sprung mass
from road shock and vibration, and keeping the human senses from unpleasant stimulation.

The input has been identified as a function of frequency just previously; the “gain” of
most interest would be ratio of the sprung response at a frequency of interest to the road input at
that frequency. The matter is phrased in that way because there other “gains” or ratios of interest:
the sprung response to axle input, or the sprung response to input directly applied by way of
driveline or aero buffeting excitation. In most cases the gain is a dimensionless number, but not
always.

To determine this relationship of input to output we must return to the 2-DOF quarter car
model wherein the sprung and unsprung masses did not move independently of each other but
were linked. From the following sprung mass free body diagram….

Fig. 6.4 – SPRUNG MASS FREE BODY DIAGRAM

…the sprung mass “dynamic equilibrium” 45 equation is obtained…

𝒎𝒔 𝒛̈ 𝒔 − 𝒄𝒔 (𝒛̇ 𝒖𝒔 − 𝒛̇ 𝒔 ) − 𝒌𝒔 (𝒛𝒖𝒔 − 𝒛𝒔 ) = 𝑭𝒔

…and then rearranged into the proper format for solution as a second order differential equation:

𝒎𝒔 𝒛̈ 𝒔 + 𝒄𝒔 𝒛̇ 𝒔 + 𝒌𝒔 𝒛𝒔 = 𝒄𝒔 𝒛̇ 𝒖𝒔 + 𝒌𝒔 𝒛𝒖𝒔 + 𝑭𝒔 (Eq. 6.3)


From the unsprung mass free body diagram…

45
Although Newton’s (1643-1727) Second Law is the basis for the solution of the problem, the concept of
“Dynamic Equilibrium” results from the work of Jean le Rond d’Alembert (1717-1783), and is sometimes referred
to as “D’Alembert’s Principle”.

34
Rev B SAWE Paper No. 3521
Category Number 31.0

Fig. 6.5 – UNSPRUNG MASS FREE BODY DIAGRAM

…the unsprung mass dynamic equilibrium equation is obtained…

𝒎𝒖𝒔 𝒛̈ 𝒖𝒔 + 𝒄𝒔 (𝒛̇ 𝒖𝒔 − 𝒛̇ 𝒔 ) + 𝒌𝒔 (𝒛𝒖𝒔 − 𝒛𝒔 ) = 𝒄𝒕 (𝒛̇ 𝒓 − 𝒛̇ 𝒖𝒔 ) + 𝒌𝒕 (𝒛𝒓 − 𝒛𝒖𝒔 ) + 𝑭𝒖𝒔


…and then also rearranged into the proper format for solution as a second order differential
equation…

𝒎𝒖𝒔 𝒛̈ 𝒖𝒔 + (𝒄𝒔 + 𝒄𝒕 )𝒛̇ 𝒖𝒔 + (𝒌𝒔 + 𝒌𝒕 )𝒛𝒖𝒔 = 𝒄𝒔 𝒛̇ 𝒔 + 𝒄𝒕 𝒛̇ 𝒓 + 𝒌𝒔 𝒛𝒔 + 𝒌𝒕 𝒛𝒓 + 𝑭𝒖𝒔


(Eq. 6.4)

These two second order differential equations describing the interlocking motions of the
sprung and unsprung masses employ the following symbolism:

ms = the quarter car model sprung mass.


mus = the quarter car model unsprung mass.
cs = the damping coefficient for the suspension damper (“shock absorber”).
ct = the damping coefficient for the damping inherent in the tire (hysteresis).
ks = the spring rate for the suspension spring.
kt = the spring rate for the tire spring quality (at a particular inflation pressure).
zs = the vertical displacement of the sprung mass.
zus = the vertical displacement of the unsprung mass.
zr = the vertical displacement input due to road surface roughness.
Fs = a force directly inflicted on the sprung mass (aero buffeting, engine vibration, etc.).
Fus = a force directly inflicted on the unsprung mass (wheel imbalance, out of round,
etc.).

35
Rev B SAWE Paper No. 3521
Category Number 31.0
The two differential equations are transformed into frequency functions and solved by the
following displacement and force substitutions (second order differential equations of such a
type as the sprung and unsprung dynamic equilibrium equations are typically amenable to such a
solution approach):

zs = Zs ejωt
zus = Zus ejωt
zr = Zr ejωt
Fs = Fs ejωt
Fus = Fus ejωt

The solutions to the two differential equations are complicated but determinable by
classic methods, although customarily the matter is first simplified by dropping the tire damping
terms 46; this practice has lead to a wide-spread lack of appreciation of the significance of the tire
damping 47. Anyway, the traditional solutions for sprung mass response to road input, sprung
mass response to axle input, and sprung mass response to direct sprung mass input are:

𝒛̈ 𝒔 𝑲𝟏 𝑲𝟐 + 𝒋[𝑲𝟏 𝑪𝝎]
=
𝒛̈ 𝒓 [𝝌𝝎𝟒 − (𝑲𝟏 + 𝑲𝟐 χ + 𝑲𝟐 )𝝎𝟐 + 𝑲𝟏 𝑲𝟐 ] + 𝒋[𝑲𝟏 𝑪𝝎 − (𝟏 + 𝝌)𝑪𝝎𝟑 ]

(Eq. 6.5)

𝒛̈ 𝒔 𝑲𝟐 𝝎𝟐 + 𝒋[𝑪𝝎𝟑 ]
=
𝑭𝒖𝒔 ⁄𝒎𝒔 [𝝌𝝎𝟒 − (𝑲𝟏 + 𝑲𝟐 χ + 𝑲𝟐 )𝝎𝟐 + 𝑲𝟏 𝑲𝟐 ] + 𝒋[𝑲𝟏 𝑪𝝎 − (𝟏 + 𝝌)𝑪𝝎𝟑 ]

(Eq. 6.6)

𝒛̈ 𝒔 [𝝌𝝎𝟒 − (𝑲𝟏 + 𝑲𝟐 )𝝎𝟐 ] + 𝒋[𝑪𝝎𝟑 ]


=
𝑭𝒔 ⁄𝒎𝒔 [𝝌𝝎𝟒 − (𝑲𝟏 + 𝑲𝟐 χ + 𝑲𝟐 )𝝎𝟐 + 𝑲𝟏 𝑲𝟐 ] + 𝒋[𝑲𝟏 𝑪𝝎 − (𝟏 + 𝝌)𝑪𝝎𝟑 ]

(Eq. 6.7)
Note that in the two later equations the input force is divided by the sprung mass; thus
the units of acceleration (l/t2) are obtained in the denominator resulting in dimensionless gain
values for all three equations. Also, in these solution equations a certain consolidation of
symbolism was undertaken in the interest of practicality. The consolidated symbolism is:

46
References [6] and [7] drop the tire damping (involving the “ct” parameter) terms to ease the solution process.
47
Reference [18], page 1.

36
Rev B SAWE Paper No. 3521
Category Number 31.0

χ = mus/ms
C = cs/ms
K1 = ks/ms
K2 = kt/ms
j = √-1

When these gain functions are plotted against the frequency in Hz (Hz = cps = ω/2π) the
result tends to look as follows:

Fig. 6.6 – SPRUNG MASS GAIN vs. FREQUENCY FOR ROAD, BODY, & AXLE
INPUTS

Note that the sprung mass gain for road input is the greatest around the sprung mass
resonance frequency of 1.2191 Hz as determined by the exact sprung mass “bounce” Eq. 2.1c
(undamped) 48. The sprung mass gain for input directly on the sprung mass is at its greatest at a
slightly higher input frequency such as that sprung mass resonance frequency of 1.2498 Hz as
determined by the simple sprung mass “bounce” Eq. 2.1a (undamped). The sprung mass gain
for input from the unsprung mass is greatest around the unsprung mass resonance frequency of
13.4020 Hz as determined by the exact “wheel hop” equation (undamped).
48
Or at least the greatest gain would be at 1.2191 Hz if all the parameters were the same for the generation of Fig.
6.6 as was used for Eq. 2.1c. The parameter values utilized to produce Figure 6.6 were close in value to, but not
exactly the same as, those used in Chapters 2 and 3.

37
Rev B SAWE Paper No. 3521
Category Number 31.0
The road acceleration inputs of Fig. 6.3 for a particular velocity (Fig. 6.2), when
multiplied by the appropriate gain factor from Figure 6.6, will generate the sprung mass vertical
acceleration response spectrum at that velocity as per the equation:

𝑮𝒛𝒔 (𝒇) = [𝑯𝒛𝒔 (𝒇)]𝟐 𝑮𝒛𝒓 (𝒇) (Eq. 6.8)


Where:

Gzs(f) = Sprung (“s”) mass vertical (“z”) acceleration PSD response (Fig. 6.7).

Hzs(f) = Sprung (“s”) mass vertical (“z”) gain or transmissibility (Fig. 6.6).

Gzr(f) = Road (“r”) vertical (“z”) acceleration PSD input at a velocity (Fig. 6.2).

When plotted, the resulting sprung mass vertical acceleration response spectrum at that
velocity (multiple velocities are not shown for clarity, but the plot would “move” with vehicle
velocity in a fashion similar to that of Fig. 6.3) would be is as follows:

Fig. 6.7 – SPRUNG MASS VERTICAL ACCELERATION RESPONSE TO ROAD


INPUT AT A PARTICULAR VELOCITY

38
Rev B SAWE Paper No. 3521
Category Number 31.0
The sprung mass response to road input is the transmissibility function of the greatest
interest, and it is now readily determinable to see how this transmissibility varies with mass,
stiffness, and damping. First, let’s see how the sprung mass response gain varies with the
unsprung mass via the data plots of Figure 6.8:

Fig. 6.8 – UNSPRUNG MASS EFFECT ON SPRUNG MASS RESPONSE GAIN TO


ROAD INPUT

As would be expected, decreasing the unsprung mass is beneficial; the response gain
decreases with decreasing unsprung mass. This because the lighter the unsprung mass the
easier it is for the suspension system to control, the less unsprung mass energy (or momentum)
there is to pass on to the sprung mass.

Now, let’s look at the variation in mass in a more generalized way, in terms of the
familiar unsprung to sprung mass ratio “χ”; how the sprung mass response gain varies with
variation of the “χ” ratio is plotted in Figure 6.9 (which looks very much like Figure 6.8, and
would look even more so if the dependent axis had not been made non-linear in the interest of
visual clarity):

39
Rev B SAWE Paper No. 3521
Category Number 31.0

Fig. 6.9 – MASS RATIO EFFECT ON SPRUNG MASS RESPONSE GAIN TO ROAD
INPUT

If the unsprung mass “mus” decreases, then the “χ” ratio decreases, and it can be seen
from Figure 6.9 that then the sprung mass response gain “Hzs(f)” decreases as well. This is
exactly the same insight that was obtained from Figure 6.8, as expected. But, now it can also be
seen that as the sprung mass “ms” increases, then again the “χ” ratio decreases and the sprung
mass response gain “Hzs(f)” decreases. This is because as the sprung mass (inertia) increases it
becomes less affected by disturbance input (or any input, for that matter, including steering
input). This is why a fully loaded vehicle will ride “softer” than when lightly loaded, but as a
matter of design philosophy it’s not generally considered desirable to increase sprung mass just
to get a smooth ride (at the expense of acceleration, braking, maneuverability, and fuel
economy).

Note that all of this variation occurs after the sprung mass resonance frequency. Up to
that point it would seem that variation in the “χ” ratio has no effect. That is because the sprung
mass in this particular plot was held constant; only the unsprung mass was actually varying 49
(if the sprung mass were varied the sprung resonance point would have varied as per the
unsprung resonance). However, this additional fact does not invalidate the previous conclusions!

Now let’s see the effect of suspension spring rate variation on the sprung mass response
(not the gain as previous, which is why the dependent axis value is zero at zero input and not

49
Reference [6], page 268. This raises the question as to why the reference authors did not create the plot as just the
unsprung mass effect on the sprung mass response gain to road input; perhaps it was thought that use of the mass
ratio allowed for some information regarding sprung mass variation effect on gain to be gleaned from the plot;
hence its inclusion in this paper.

40
Rev B SAWE Paper No. 3521
Category Number 31.0
unity) to road vibration input. Remember that the sprung mass natural (resonance) frequency
𝟏
“fs” varies with the spring rate “ks” as per the simple Equation 2.1a (“fs = 𝟐𝛑 √(ks/ms)”) from
Chapter 2. Even though the following figure is sprung mass acceleration response (“g’s”) vs.
road input frequency, instead of sprung mass response gain (dimensionless) vs. road input
frequency, the general look of the plot is very similar:

Fig. 6.10 – SPRUNG MASS NATURAL FREQUENCY EFFECT ON SPRUNG MASS


RESPONSE TO ROAD INPUT

Note in Figure 6.10 that as the sprung mass natural frequency decreases, the sprung
mass acceleration response decreases dramatically. Putting this in another, more informative,
way: given Equation 2.1a (or 2.1b, or 2.1c), this means as the spring rate decreases, or as the
sprung mass increases, the natural frequency of the sprung mass decreases and therefore so
does the sprung mass acceleration response. So, from just a ride perspective, it is desirable to
have “soft” springs or a huge unsprung mass, and the latter possibility is not desirable from
many other standpoints as has already been pointed out.

Another way to look at the effect of spring rates on sprung mass response to road input
is to observe the effect of variation in the spring rate ratio “kt/ks”. As before, variation of spring
rate(s) means a variation in natural frequency, only now we are constrained to considering just
the intermediate and the exact sprung mass frequency equations as the simple equation doesn’t
involve “kt”. However, since even just the intermediate Equation 2.1b becomes somewhat
𝟏
complicated as “fs = 𝟐𝛑 √[((ks×kt)/(ks+kt))/ms]” when combined with the Equation 2.4 for the
suspension spring “ks” and the tire spring “kt” in series together, the information imparted
becomes harder to fully digest. In any case, the plot of sprung mass response gain vs. spring rate
ratio “kt/ks” is as per Figure 6.11 (again with a non-linear dependent axis):

41
Rev B SAWE Paper No. 3521
Category Number 31.0

Fig. 6.11 – SPRING RATE RATIO EFFECT ON SPRUNG MASS RESPONSE GAIN TO
ROAD INPUT

The sprung mass response gain seems to increase somewhat with increasing “kt/ks” up
to the input frequency point(s) of sprung mass resonance, and then decrease (!) at higher
frequencies up to the input frequency point(s) of unsprung mass resonance, after which the
situation reverses itself yet again! This may seem confusing and, barring the possibility that the
source documentation may be faulty, it may help to remember the conclusions drawn from
Figure 6.10, in particular that reducing the stiffness of the suspension spring is an unalloyed
benefit as far as ride is concerned (ignoring roll, sway, and pitch considerations). If the
suspension spring stiffness is considered constant, then what Figure 6.11 seems to indicate is that
the sprung mass response gain increases with increasing “kt” up to the point of sprung mass
resonance, and then decreases the response with increasing “kt” up to the point of unsprung
mass resonance, after which the sprung mass response gain again increases with increasing
“kt”. This seems counter-intuitive; one would expect that the gain would increase with increasing
tire stiffness “kt” throughout, so the area between the two resonance inflection points might seem
suspect.

However, for this plot the tire stiffness was actually held constant, so only the suspension
stiffness was really varying 50. This is why the sprung mass resonance point varied, but the
unsprung mass resonance point seemed not to vary, as the unsprung mass resonance point is
predominantly a matter of tire stiffness variation (remember the simple unsprung mass
1
resonance Equation 2.3a: “fhopus = 2π √(kt/mus)”). This fact renders the previous conclusion with
regard to the effect of tire stiffness on sprung mass response gain questionable (it may be
50
Reference [6], page 269. Again, it is assumed the source authors used the spring rate ratio in lieu of just the
suspension spring stiffness “ks” because more information was imparted that way, or at least got the reader to give
some consideration to the role “kt” might play.

42
Rev B SAWE Paper No. 3521
Category Number 31.0
beneficial across the board with respect to vibration and shock transmissibility to reduce tire
stiffness, but generally softer tires means more rolling resistance and reduced fuel economy); in
an actual physical case this question can easily be resolved empirically by measurement of
sprung response at constant road input frequency while varying tire pressure (stiffness).

In any case, the sprung mass gain (with respect to road input) variation would not seem
to be as dramatic with “kt” as with “ks”, although it is wise to remember the difference in the
dependent axis scales between Fig. 6.10 and Fig. 6.11, and that in Fig. 6.11 any gain value under
unity is pretty good.

Having investigated the spring ratio effect on the sprung mass gain (with the
reservations as noted), now let’s see the damping ratio effect on the sprung mass response gain:

Fig. 6.12 – DAMPING RATIO EFFECT ON SPRUNG MASS RESPONSE GAIN TO


ROAD INPUT

Again, the non-linearity of the vertical axis in the above plot is there to allow us to see
more clearly what happens in the low gain/high input frequency plot region. What would
otherwise be obscured is that, although generally the lower the damping ratio the lower the gain
over a wide frequency range, the lower damping ratio also generates higher gains at the sprung
(circa 7.0 gain shown) and unsprung resonance (circa 0.1 gain shown) frequencies. It would
seem that the 1 Hz gain might be a problem. However, sprung resonance is around 1 Hz (cps) by
design as that frequency is particularly tolerable by humans, so a gain spike there might not be as
serious a problem as it would be at a different frequency, but a gain in the range of 1.5 to 2.0
would be about the limit. Of course, such a 1 Hz gain spike would not be a problem at all as long
as the vehicle speed and road surface do not generate 1 Hz road input, but avoiding certain
speeds on certain roads is a bit of an onerous restriction to place on a driver.

43
Rev B SAWE Paper No. 3521
Category Number 31.0
A similar plot of sprung mass response gain vs. damping ratio variation, but over a
wider range of variation, and this time with a linear dependent axis (which does obscure what is
happening after the unsprung resonance frequency, but that is largely irrelevant as the gain
magnitude is so small), is presented:

Fig. 6.13 – DAMPING RATIO EFFECT ON SPRUNG MASS RESPONSE GAIN TO


ROAD INPUT

Note that in Figure 6.13 the maximum damping ratio is going all the way up to 200% (ξ
= 2.0) percent of the critical value, as opposed to just the 30% (ξ = 0.3) of the previous figure. At
200% the “shock absorber” generates a very high resistance to motion, which in the above plot is
peaks around 3 to 4 Hz. This is because the suspension is essentially “locked” by the “stiff”
damper and the sprung mass is just bouncing on the tire, for which the resonance, using some
“typical” parameter values from Chapter 2, would be around 51:

fξ=2 = 1/(2π) √2133/1.907 = 5.3 cps

Consequently, between the very high gain at around the sprung resonance frequency for
10% damping, and the generally high gain over a wider frequency range which peaks at the
“locked” suspension frequency for 200% damping, a damping ratio value of about 30% to 40%
has traditionally been taken for automotive systems as a good compromise 52.

All of the preceding has concentrated on the damping of the “shock absorber”, totally
neglecting the damping of the tire. Consider again Equation 3.3 for the damping ratio:

ξ = 𝐜𝐬 ⁄�𝟒𝐤 𝐬 ⁄𝐦𝐬

51
Remember that the parameter values used to generate the plots were not quite the same as those used in Chapter 2.
52
Reference [7], page 156.

44
Rev B SAWE Paper No. 3521
Category Number 31.0
The damping ratio “ξ” could be made to include the effect of the tire damping coefficient
“ct”, and not just the “shock absorber” damping coefficient “cs”, by using a combined damping
coefficient as discussed in Chapter 3. However, remember that the tire damping coefficient “ct”
terms were dropped when the equations for sprung mass gain due to road input (Eq. 6.5), due to
unsprung input (Eq. 6.6), and due to sprung input (Eq. 6.7) were derived; any effects due to tire
damping variation are simply not present. As the gain equations were derived by dropping all
“ct” terms, it is not certain that including “ct” in the damping ratio determinations would
represent any improvement in the plots.

The tire damping was included in the derivation of the equation for sprung mass gain
due to road input in a study of active suspensions by authors Türkay and Akҫay 53. They ran their
quarter-car model simulation using that derived gain equation and a value for “ct” of zero and got
the sprung mass acceleration response over a wide range of road vibration frequency input.
Then they took a value for “ct” of about 10% of “cs” as being a realistic estimation, which given
their model parameter values resulted in about 0.56 lb-sec/in (98 N-sec/m) 54, and again ran their
simulation. Comparison of the sprung mass acceleration results indicated a general reduction in
acceleration of about 3 percent for a traditional passive suspension system (for the “locked”
suspension vibration case just previously discussed the difference would be a less than 1 percent
reduction). However, for the active suspension system that was the principal object of Türkay’s
and Akҫay’s concern, inclusion of reasonable values for “ct” in the control algorithms resulted in
a general reduction of 6.8 percent 55.

What is revealed by all this is that, for conventional passive suspension systems,
choosing the damping coefficient to minimize response gain is a matter of compromise for
reasons even more complicated than those that are mentioned in Chapter 3. The algorithm for a
fully active damping system would have to be complex and involve many input parameters like
the vehicle velocity, lateral acceleration, wheel acceleration, load, steering input, etc., in order to
exercise optimum control over the damping.

Having considered the basic aspects of the sprung mass response to road input, we now
move on to a consideration of the sprung mass response to “body” 56 input. The sprung mass
response to body input is quite different from its response to road input, not just because the gain
function is different, but because the input is fundamentally different. Input at the body level
results from such things as power plant torque/inertial fluctuations, gear mesh imperfections,
driveline imbalance and/or misalignment, and exhaust system pulsation. Unlike the random
nature of road input, such inputs are much more regular (periodic) in nature, with their frequency
and power varying in direct relation to power plant rpm. The only possible exception to this
characterization of body inputs would be the input due to aerodynamic buffeting.

How the mass ratio affects the sprung mass response gain to body input may be seen
from the following plot:

53
Reference [18], pages 1-2.
54
Ibid, page 4.
55
Ibid, page 4.
56
“Body”, a.k.a. sprung mass, input is any input originating from within the sprung mass itself.

45
Rev B SAWE Paper No. 3521
Category Number 31.0

Fig. 6.14 – EFFECT OF MASS RATIO ON SPRUNG MASS RESPONSE GAIN TO


BODY INPUT

Since the sprung mass resonance is unchanging, the variation in unsprung to sprung
mass ratio “χ” must be coming about from variation of the unsprung mass “mus” alone.
Because the input to the sprung mass is not coming from the road and working its suspension,
but instead is coming from directly within the sprung mass itself, it is not expected that varying
“mus” would have any effect on sprung mass response. The above plot simply confirms that
expectation.

Now consider the effect of the spring rate ratio on the sprung mass response gain to
body input:

Fig. 6.15 – EFFECT OF SPRING RATE RATIO ON SPRUNG MASS RESPONSE


GAIN TO BODY INPUT
46
Rev B SAWE Paper No. 3521
Category Number 31.0
Again, since the unsprung mass really isn’t involved, neither is the tire’s spring
constant. What Figure 6.15 is simply telling us is that as the spring ratio increases, i.e., as the
suspension spring constant “ks” gets smaller, the sprung mass resonance frequency (naturally)
and response gain decreases.

Lastly, there is the effect of the damping ratio variation on the sprung mass response
gain to body input:

Fig. 6.16 – EFFECT OF DAMPING RATIO ON SPRUNG MASS RESPONSE GAIN TO


BODY INPUT

Here the effect of the damping ratio on the sprung mass response gain is quite simple,
unlike the damping ratio effect on the sprung mass response gain to road input. The effect on
response gain to road input was complicated and indicated a basic necessity for compromise in
damping ratio selection. Here the case seemly requires no compromise in the design choice for
damping ratio, but remember that any choice of “shock absorber” damping ratio must be
governed by the effect on road input transmissibility, not body. The suspension’s raison d’être is
to mitigate road input; there are other ways to handle body input (better engine mounts, etc.).

Lastly, the sprung mass response to unsprung mass (a.k.a. “axle” in the literature) input
is yet another story. Input at the “axle” level is the result of tire/wheel/hub/brake imbalance,
misalignment, out-of-round conditions, and radial stiffness variations. Such inputs are like body
inputs in that they are more regular in nature than road input, but the inputs vary in frequency
and power in direct relation to vehicle speed, not engine speed.

As an illustrative example, suppose that a vehicle has tires with a rolling radius of 14.09
inches (0.3579 m), and there is a simple imbalance of a rotating assembly (wheel, tire, brake
disk, etc.) of 1 lb (0.00259 lb-sec2/in, or 0.4536 kg) at 12 inches (0.3048 m) radius from the hub.

47
Rev B SAWE Paper No. 3521
Category Number 31.0
As the vehicle accelerates the imbalance causes, at a correspondingly increasing frequency, an
axle level input to the sprung mass of ever increasing magnitude:

𝐅 = 𝐦 𝐫 𝛚𝟐 = 𝟎. 𝟎𝟎𝟐𝟓𝟗 × 𝟏𝟐 × 𝛚𝟐

This force amplitude translates into an acceleration amplitude (a = F/mus) for a


corresponding oscillating acceleration input into the sprung mass. The corresponding sprung
𝒛̈
mass response is the input acceleration times the appropriate gain value (“ 𝑭 ⁄𝒔𝒎 ”) from Figure
𝒖𝒔 𝒔
6.6 over the range of frequencies considered. When plotted, the results should look something
like this:

Fig. 6.17 – SPRUNG MASS VERTICAL ACCELERATION RESPONSE TO AXLE


INPUT AT VARYING VEHICLE VELOCITY / FREQUENCY

Note that the acceleration response gets worse in accord with increasing velocity/
frequency and with the gain function (at that frequency) until the resonance frequency of the
unsprung mass is neared; the response then increases at a decreasing rate, levels off, and begins
to decrease, which all makes sense as the unsprung mass is the source of the disturbance input.

Of course, it is very unlikely that there ever would be a rotating imbalance as large as this
example. What is much more likely is that a large number of imbalances, out-of-round
conditions, and stiffness variations will all be making various inputs which, as the number of
these minor disturbances grows, will tend to offset each other. Furthermore, even when they
don’t offset each other, the resulting situation will never be as simple as in this illustrative
example.

48
Rev B SAWE Paper No. 3521
Category Number 31.0
7 - GYROSCOPIC REACTIONS

The matter of gyroscopic reactions is an area often neglected, but very real. Gyroscopic
torque reactions result from the inevitable changes in rolling axis orientation (precession) due to
camber and steering angle changes as the wheel moves with respect to the body due to bumps,
dips, and vehicle maneuver. Such unwanted reactions can cause “shimmy” and “tramp”, and it is
best to reduce those reactions as much as possible in the interest of control and wear.

For a quantitative determination of gyroscopic effects 57, we turn to the equation:

𝑻 = 𝑰 𝝎𝒓 𝝎𝒑 (Eq. 7.1)

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/sec).

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

Radially decreasing rolling mass size increases angular speed proportionately, velocity
being held as a constant:
V = r 𝝎r

However, the roll inertia “I” varies with the radius to the fourth power:

𝐖
𝐈= r4
𝐠

Let’s say a vehicle is traveling at a velocity of 60 mph (96.6 kph) when the right front
wheel encounters a 2 in (5.1 cm) high bump forcing the IFS to go through 1.5 degrees of camber
change (determined from the suspension kinematics). The rolling inertia of 28.53 lb-ft2 (1.25
kg-m2) corresponds to a rolling weight of approximately 52.1 lb (23.63 kg) with a radius of
gyration of 8.93 inches (0.74 feet, or 0.23 m). The suspension is equipped with a 152/92R16 tire
of about 12.63 in (1.05 ft, or 32.1 cm) rolling radius; it is assumed that the time it takes for the
tire to “engulf” the bump/rise to bump height will be approximately equal to the time necessary

57
Reference [4], pages 255-257

49
Rev B SAWE Paper No. 3521
Category Number 31.0
for the vehicle to travel a distance equal to the tire cross-section height (aspect ratio × section
width):

t = (((0.92 × 152 mm) / 25.4 mm/in) / 12 in/ft) / ((60 mph × 5280


ft/mile) / 3600 sec/hr)) = 0.0052 seconds

This allows for the determination of “ωp” as degrees camber change per time (in radians
per second):

ωp = 1.5 degrees / 0.0052 seconds = 288.5 deg/sec, or 45.9 rad/sec

As for the determination of “ωr” it would be equal to the number of tire rotations per time
at speed (in radians per second):

ωr = 60 mph × (5280 ft/m)/(3600 sec/hr)/(2 π × 1.05 feet) = 13.3 cps,


or 83.8 rad/sec

So finally we can determine the reaction torque (converting weight units to mass units:
mass = weight / 32.174 ft/sec2):

T = (28.53 lb-ft2 / 32.174 ft/sec2) x 45.9 rad/sec × 83.8 rad/sec

= 3410.7 lb-ft, or 347.8 kg-m

Note that halving the wheel/tire size (0.526 ft rolling radius) would double “ωr” (167.6
rad/sec), but a half-size “wheel” has one 1/16th the inertia, so the reaction torque would be about
1/8th (426.3 lb-ft). Therefore, the smaller the wheel would have 87.5% less of a tendency toward
any disturbance or wear as induced by gyroscopic reactions.

Elimination of shimmy and tramp was one of the reasons for adoption of independent
front suspension. With a beam axle a precession motion at one side was mirrored at the other
side, causing a double impact. This effect was to become especially significant at the front axle
as that was the steering axle. At the rear axle the lack of steering complications helped slow the
advance to independent suspension; retention of the simpler and cheaper beam was to continue to
be a viable option up to the present day.

Early vehicles were generally satisfactory with a beam axle front suspension as such
vehicles tended to have lighter weight tire/wheel combinations (though admittedly of somewhat
50
Rev B SAWE Paper No. 3521
Category Number 31.0
larger overall diameter) and to travel at slower speeds than later vehicles. Also, it wasn’t until
around 1915 that it became customary to have brake mechanisms included at the front wheels
(previously braking was generally rear wheel only). As roads improved and engines became
more powerful, there was a consequent increase in speed. With heavier wheel/tire designs that
included the mass of front brakes, the use of beam axle front suspensions became less desirable.
Also the placement of the engine, although still in the front of the vehicle, needed to be
somewhat more forward in response to a need to modify vehicle mass properties (c.g., “J”) in
an effort to control pitch motions; but then the engine and the beam axle tended to vie for the
same space. The introduction of IFS incurred more expense, but meant the engine could sit
between the right and left front suspensions, which not only affected a change in mass
properties but tended to allow for shorter hood lines, while the problems of shimmy and tramp,
pitch motion, and rough ride could all be alleviated.

Just as gyroscopic reactions can have undesirable effects, they can also have beneficial
effects. This is possible if the direction of gyroscopic reactions is understood, and if the
orientation of the rotating masses generating those reactions is adjusted accordingly. As the
determination of reaction direction is complex and not intuitive, the whole matter of automotive
gyroscopic reactions is generally ignored in the literature.

Two exceptions in the literature come to mind: Colin Campbell and Erskine Crossley.
Both these authors give us mutually corroborating statements concerning directional
determination of the gyroscopic effects of the major automotive rotating masses. Campbell
states 58:

“A factor which is always present when cornering and which is often


ignored is the gyroscopic reaction of the rotating mass of the engine
components…With the normal direction 59 of rotation a left hand turn
produces a tendency for the nose to dip…a right hand turn throws more
weight on the rear wheels.”

This is a statement which Crossley echoes and expands upon 60:

“…consider the action of the crankshaft and flywheel of an automobile as


you drive around a corner. These usually revolve clockwise when viewed
from the front, so that on turning a corner to the right, the torque applied
to the shaft through its bearings will cause it to attempt to precess by
raising its forward end and depressing its rear end. Thus some weight will
be lifted off the front wheels of the car.”

Crossley goes on to address the next largest automotive rotating masses, the wheels 61:

58
Reference [3], page 180.
59
Clockwise (“CW”) when viewed from the front of the engine as installed in a conventional front engine
configuration of longitudinal orientation.
60
Reference [4], page 255.
61
Ibid.

51
Rev B SAWE Paper No. 3521
Category Number 31.0
“…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.”

Furthermore 62:

“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…”

To illustrate the engine situation, and something called the “Right Hand Rule” 63 which
allows for determination of gyroscopic torque reactions, the following illustration is provided:

Figure 7.1 – ENGINE GYROSCOPIC TORQUE REACTION & THE RIGHT HAND
RULE

Note that when the fingers of the right hand curl in the direction of the rotation being
considered, then the thumb points in the direction of the spin vector, and attempts to redirect that
vector generate a reaction torque as illustrated. For a conventional longitudinal engine
installation, with the engine rotation being clockwise (CW) as viewed from the front, turning
right or left generates a not very helpful nose-up or nose-down reaction, respectfully. The fact
that the vehicle may behave differently (“oversteering” vs. “understeering”?) in a left hand turn
as opposed to a right hand turn is not conducive to safe, predictable handling.

62
Ibid, pages 256-567.
63
There is also a “three finger rule” for determining the orientation of all the vectors concerned, including the
reaction torque vector, but that is more complication than warranted given the scope of this paper.

52
Rev B SAWE Paper No. 3521
Category Number 31.0
If the engine were mounted transversely, as is also shown, then the reaction torques
from turning right or left could generate a helpful right-side or left-side counter to the centrifugal
force reaction (not shown). Of course, this would only matter if the engine were massive enough
and high-revving enough, and if the vehicle were in need of every possible advantage in getting
through the turn….

The rotating portion of the unsprung mass: wheels, tires, brakes, shafts, etc.; also
conventionally generate unhelpful gyroscopic torque reactions but, unlike the engine, the
palliative of re-orientation of the rotation axis is not available. The only course of mitigation
open is to reduce the rotational inertia of these rotating assemblies as much as possible.

53
Rev B SAWE Paper No. 3521
Category Number 31.0
8– MORE TWO DOF MODELS

In accord with the movement toward greater realism, remember that there is a sprung
mass rotational (pitch) oscillation, in addition to the translational (bounce), oscillation. The
consideration of this new motion necessitates a change from the previous quarter-car models to
the 2-DOF half-car model as shown 64:

Figure 8.1 – TWO DOF SPRUNG MASS BOUNCE & PITCH MODEL

This is a special model in that “lf = lr” and “kf = kr”, which is generally not the case.
However, this special model is useful in introducing the study of automotive bounce and pitch
motions because in this special case those motions are said to be “uncoupled”. That means the
sprung mass can bounce vertically without necessarily incurring any rotational motion about the
c.g., or the sprung mass can rotationally oscillate about the c.g. without necessarily incurring
any translational motion at the c.g. The simple uncoupled equation for the pitch motion is
analogous to the simple sprung mass bounce equation Eq. 2.1a of Chapter 2:

𝟏
𝒇𝜽 = ���𝒌𝒇 𝒍𝒇 𝟐 + 𝒌𝒓 𝒍𝒓 𝟐 ��𝑴𝒔 𝑲𝟐 � (Eq. 8.1)
𝟐𝝅

The quantity “K” is the pitch radius of gyration of the sprung mass. The quantity “Ms
2
K ” is equivalent to the sprung mass pitch moment of inertia “J”. When the pitch frequency is
equal to the bounce frequency the following equivalency of the above simple sprung pitch
equation to the simple sprung bounce equation can be made 65:

𝟏 𝟏
���𝒌𝒇 + 𝒌𝒓 �⁄𝑴𝒔 � = ���𝒌𝒇 𝒍𝒇 𝟐 + 𝒌𝒓 𝒍𝒓 𝟐 ��𝑴𝒔 𝑲𝟐 �
𝟐𝝅 𝟐𝝅

64
Note that “Ms” is the vehicle total sprung mass, so “kf” and “kr” are double the value that would have been used
in the corresponding quarter-car models (ignoring anti-roll bar complications).
65
Reference [3], page 197.

54
Rev B SAWE Paper No. 3521
Category Number 31.0
This may be further reduced:

(kf+kr) = (kflf2 +krlr2)/ K2

K2 (kf+kr)/(kflf2 +krlr2) = 1

𝑲𝟐 ��𝒍𝒇 × 𝒍𝒓 � = 𝟏
2
The quantity “K /(lf x lr)” 66, known as the Dynamic Index (DI), is significant as an
indicator of when the pitch frequency will be the same as the bounce frequency, which
eliminates the possibility of “heterodyning”. Heterodyning is when two different frequencies
interact to produce a “beat” frequency and can be depicted:

Figure 8.2 – FREQUENCY INTERACTION BEAT

For two such frequencies of equal amplitude “A” the beat “motion” can be written as 67:

𝝎𝟏 −𝝎𝟐 𝝎𝟏 +𝝎𝟐
𝒛 = [𝟐 𝑨 𝐬𝐢𝐧( 𝒕)] 𝐬𝐢𝐧( 𝒕) (Eq. 8.2)
𝟐 𝟐

Note that resulting amplitude is now an oscillating amplitude, and that the resulting beat
frequency is equal to the difference in the interacting frequencies because the sine function will
reach an absolute maximum twice every cycle, and “2 x [(𝝎1 – 𝝎2)/2]” is equal to “𝝎1 – 𝝎2”. So
a frequency of 14 cps and a frequency of 13.5 cps could interact to form a rather languid beat
frequency of 0.5 cps.

The possible heterodyning of pitch and bounce frequencies can produce a motion much
like being at sea in a small boat, so it should not be surprising that, in vehicles with poorly
designed suspensions, passengers may be able to enjoy all the sensations of “mal de mare” while
still on dry land. Note that the long slow cycle of the “beat” is the problem, and that ordinary

66
In the literature this quantity is also often symbolized as “k2/ab” (SAE J670e).
67
Reference 15, page 521.

55
Rev B SAWE Paper No. 3521
Category Number 31.0
long slow vibrations under 1 cps that have nothing to do with heterodyning can have the same
physiological effect.

However, the reality is such that the Dynamic Index is seldom exactly equal to “1”; the
automotive designer is left with getting as close to “1” as possible. For vintage sports cars of
conventional design (four wheels, front engine, etc.) where the engine was situated well aft of a
front beam axle, and there was no significant body overhang of the wheelbase (i.e., “K” tended
2
to be relatively small), the value of “K /(lf x lr)” tended to be about 0.6 (circa 1954) with
consequent significant beat motions. Many later sports cars tended to be closer to 0.8 (circa
1969) 68. Even more modern designs may be around 0.9 (circa 1980) 69. This slow progression
toward unity has been explained as the result of a more forward location of the engine (often
right on the front axle line) made possible by the adoption of IFS, and other mass
redistributions such as to produce higher values of “K” (which used to involve greater body
overhangs of the wheelbase, especially to the rear, although that is no longer a favored method).

In the case of the 1958 Jaguar XK150S, using estimated values for the pitch radius of
gyration and the sprung mass longitudinal CG location, the value of the Dynamic Index
comes out to 0.65, which would seem to be right in line with the historical trend. In passing, it
should be noted that the vehicle radius of gyration in pitch can’t be significantly increased
without also increasing the radius of gyration in yaw, which is not regarded as a good thing, at
least for maneuvering performance 70 (another area of necessary compromise).

68
Reference [3], page 197.
69
Reference [1], page 29.
70
Reference [1], pages 31-34. The chief adverse effect occurs during the transient stage of a turning maneuver.
However, large yaw inertia does improve directional stability, but to acquire stability that way is like acquiring a
smooth ride by increasing the sprung mass (not good design practice).

56
Rev B SAWE Paper No. 3521
Category Number 31.0
9 – THE PRINCIPAL MODES

So far the most complete, but still simplified, automotive vibration sprung mass model is
the undamped 2-DOF, which therefore must have 2 principal modes of motion. One mode is
“translational” and the other mode is “rotational”, each with its own natural frequency unless the
Dynamic Index is equal to unity. The more general 2-DOF model, as opposed to the previous
special 2-DOF model (where the c.g. was centrally located and the suspension spring rates equal
front to rear), may be depicted as:

Figure 9.1 – GENERAL TWO DOF SPRUNG MASS MODEL

Said model is readily amenable to simple, classical, non-computerized analysis. Taking


the equilibrium attitude of the sprung mass and its c.g. position as the reference, the dynamic
equilibrium equations of the two small amplitude (thus eliminating non-linear effects) free
vibratory motions, one linear and one rotary, are:

F = ma => −k f (Z − lf θ) − k r (Z + lr θ) = Ms Z̈

T = Jα => k f (Z − lf θ)lf − k r (Z + lr θ)lr = Jθ̈


Where “J” is the pitch mass moment of inertia at the c.g., which relates to our old
friend “K”, the pitch radius of gyration: “J = Ms K2”. These equations of motion may be
rearranged to express the accelerations:

1 1
Z̈ = (k f + k r )Z + (k l − k f lf )θ
Ms Ms r r

1 1
θ̈ = (k f l2f + k r l2r )θ + (k r lr − k f lf )Z
J J

57
Rev B SAWE Paper No. 3521
Category Number 31.0
These equations, and the motions they represent, are linked by the one term they have in
common: “(krlr-kflf)/Ms”, which is called the “Coupling Coefficient” (CC). When this term is
zero, i.e., “krlr“ equals “kflf”, the motions of the sprung mass become uncoupled, and the sprung
mass can rotate about the c.g. without the c.g. bouncing, and the c.g. can bounce without the
mass rotating, as illustrated by the previous (Fig. 8.1) special case:

Figure 9.2 – UNCOUPLED ROTATION (PITCH) & TRANSLATION (BOUNCE)

The way this is explained is that the pitch node is at the c.g. while the “bounce” is really
just pitch about some node at an infinite distance away from the c.g.. Of course, in reality there
will always be some coupling of bounce and rotation of an automotive sprung mass because the
uncoupled situation requires an exact point in an infinite spectrum of possibilities. If for no other
reason, this will be because the location of the sprung mass c.g. is always varying as the vehicle
is operated (fuel is consumed, loads are varied). Vehicle behavior always has to be examined
throughout the full range of its possible loading (minimum operational weight, maximum
operational weight, most forward operational c.g., most aft operational c.g., etc.).

Although it is never possible to have completely decoupled motions, save perhaps for
some fleeting point in time during the vehicle’s operation, it is often deemed beneficial to design
an automobile so as to reach as close to an uncoupled state as possible (i.e., get “krlr” as close to
equality with “kflf” as possible), for as much of the operational envelope as possible. Note that in
a decoupled state a force applied at the c.g produces only a vertical displacement, and a torque
applied to the sprung mass produces only a rotation about the c.g.

Since the aim is for an uncoupled system, but it is certain that some small degree of
coupling will almost always be present, it is important to be able to find the natural frequencies
and normal modes of vibration for both uncoupled and coupled cases. The appropriate equations
are derived from the small amplitude equations of motion 71, and for the special uncoupled (krlr =
kflf) condition the equations are:

𝟏 𝟏
𝒇𝒛 = ���𝒌𝒓 + 𝒌𝒇 �⁄𝑴𝒔 �, 𝒐𝒓 √𝒂 𝒄𝒑𝒔 (Eq. 9.1)
𝟐𝝅 𝟐𝝅

𝟏 𝟏
𝒇𝜽 = ���𝒌𝒓 𝒍𝒓 𝟐 + 𝒌𝒇 𝒍𝒇 𝟐 ��𝑱�, 𝒐𝒓 √𝒄 𝒄𝒑𝒔 (Eq. 9.2)
𝟐𝝅 𝟐𝝅

71
Reference [17], pages 180-181.

58
Rev B SAWE Paper No. 3521
Category Number 31.0
For the usual coupled (“krlr ≠ kflf”) condition the general equations of the two angular
frequencies (each a combined bounce and pitch) for the two principle modes of vibration are 72:

𝟏
𝒇𝟏 = ��𝒂 − �𝒃𝟐 ⁄�𝑲𝟐 (𝒄 − 𝒂)��� 𝒄𝒑𝒔 (Eq. 9.3)
𝟐𝝅

𝟏
𝒇𝟐 = ��𝒄 + �𝒃𝟐 ⁄�𝑲𝟐 (𝒄 − 𝒂)��� 𝒄𝒑𝒔 (Eq. 9.4)
𝟐𝝅

To simplify calculations, certain parametric expressions were condensed into various


equation coefficients. These equation coefficients are:

𝒂 = �𝒌𝒓 + 𝒌𝒇 �⁄𝑴𝒔 (Eq. 9.5)

𝒃 = 𝑪𝒐𝒖𝒑𝒍𝒊𝒏𝒈 𝑪𝒐𝒆𝒇𝒇𝒊𝒄𝒊𝒆𝒏𝒕 = �𝒌𝒓 𝒍𝒓 − 𝒌𝒇 𝒍𝒇 �⁄𝑴𝒔 (Eq. 9.6)

𝒄 = �𝒌𝒓 𝒍𝒓 𝟐 + 𝒌𝒇 𝒍𝒇 𝟐 ��𝑱 (Eq. 9.7)

𝑫𝑰 = 𝑫𝒚𝒏𝒂𝒎𝒊𝒄 𝑰𝒏𝒅𝒆𝒙 = 𝑲𝟐 ��𝒍𝒇 𝒍𝒓 � (Eq. 9.8)

The introduction of “a”, “b”, and “c” is just a matter of making the equations
manageable. Note that “b” is our old friend the Coupling Coefficient and that when this
coefficient is equal to zero (“krlr = kflf”) then the general equations of coupled motion (Eq. 9.3 &
Eq. 9.4) degenerate to the special case equations of uncoupled motion (Eq. 9.1 & Eq. 9.2). Also,
when the Dynamic Index (more specifically, the Dynamic Index in Pitch) is equal to one the
bounce and pitch frequencies are equal (heterodyning is not a problem).

The node points for the uncoupled frequencies (“fz” & “fθ”) are, as noted, at infinity and
at the c.g. respectively, but what about the node points for the more general coupled case?
Something called the “amplitude ratio equation” is used for finding the node point locations for
the coupled frequencies (“f1” & “f2”) of the Principal Modes of vibration 73:

𝑿 = 𝒃⁄(𝝎𝟐 − 𝒂) (Eq. 9.9)

72
Colin Campbell consistently uses Eq. 9.3 and Eq. 9.4, plus Eq. 9.2, to determine the main free motions (thereby
forming 3, not 2, principal motions of a 2-DOF suspension system model!); this author is at a loss to explain this.
73
SAE J6a, “Ride and Vibration Data Manual”, refers to “bounce” as being motion associated with a node outside
the wheelbase, and to “pitch” as being motion associated with a node within the wheelbase. Gillespie simply terms
the node furthest from the c.g. the “bounce” node, and the closest node the “pitch” node (Reference [7], pages 179-
180).

59
Rev B SAWE Paper No. 3521
Category Number 31.0

Where the frequency “𝜔” is just “f1” or “f2” but in radians per second (drop the “1/2π”
conversion factor). How to use the equations presented so far for determining the Principal
Modes of Vibration frequencies and the corresponding node point locations will be concretely
illustrated in Chapter 13.

60
Rev B SAWE Paper No. 3521
Category Number 31.0
10 – THE SPRING CENTER

Along with the c.g. and the Principal Modes node points, there are a number of other
special points located along the 2-DOF bounce/pitch model. For instance, if the two vibration
modes are in a coupled state (krlr ≠ kflf), there will always be some point “SC” (the “Spring
Center”) somewhere along the longitudinal axis which will seem to be a node of uncoupled
motion as a force gradually applied at that point will produce only vertical motion, no rotation.
Yet when said force is suddenly removed the resulting vibratory motion will be both vertical and
rotational, due to the reaction moment about “SC” caused by the inertia at the c.g. times the arm
from the c.g. to the “SC”. Thus when the motions are uncoupled at the c.g. the situation is called
“dynamic uncoupling”, but when the motions at uncoupled at some point “SC” the situation is
called “static uncoupling” (or “dynamic coupling”). The point “SC” is where “krβ = kfα”.

Figure 10.1 – THE SPRING CENTER

Since “α + β = l” (“l” is the wheelbase) it’s easy to solve for the SC locating dimensions
“α” and “β” (two unknowns, two equations); the results are:

𝜶 = 𝒌𝒓 𝒍⁄�𝒌𝒇 + 𝒌𝒓 � (Eq. 10.1)

𝜷 = 𝒌𝒇 𝒍⁄�𝒌𝒇 + 𝒌𝒓 � (Eq. 10.2)

61
Rev B SAWE Paper No. 3521
Category Number 31.0
11 – THE CONJUGATE CENTERS OF PERCUSSION

However, there are special points other than the SC, such as the “Conjugate Centers of
Percussion”, also known as “Double Conjugate Points”. There are always a set of two such
points, and the special thing about them is that a force applied at one will produce only a rotation
at the other. When the Dynamic Index is equal to one the conjugate points are located exactly at
the front and rear wheel centers. Consequently, when the conjugate points are so located, the
bounce frequencies front and rear can be found by reduction to the simple corresponding 1 DOF
quarter-car suspension models. This phenomenon was first expounded upon by the noted
automotive dynamicist J.J. Guest in his paper “The Main Free Vibrations of an Autocar” in
1926 74. The general equations and procedure for locating these points, whether they are at the
wheel centers or not (although if they’re not close to the wheel centers they are of little interest),
and determining the associated frequencies are as follows 75:

1. Find the “Spring Center” locating dimensions “α” and “β” (Eq. 10.1 and Eq. 10.2).

2. Find the quantity “c2”:

𝒄𝟐 = 𝜶𝜷 (Eq. 11.1)
3. Find the quantity “e”:

𝒆 = 𝜶 − 𝒍𝒇 (Eq. 11.2)

4. Solve for “r” (locates point “H” forward of the c.g.):

𝒓 = (𝑲𝟐 − 𝒆𝟐 − 𝒄𝟐 )⁄𝟐𝒆 + �((𝑲𝟐 − 𝒆𝟐 − 𝒄𝟐 )𝟐 + 𝟒𝑲𝟐 𝒆𝟐 )⁄𝟐𝒆 (Eq. 11.3)

5. Solve for “s” (locates point “J” aftward of the c.g.):

𝒔 = 𝑲𝟐 ⁄ 𝒓 (Eq. 11.4)
6. Solve for the frequencies about points “H” and “J”:

𝐟𝐡 = 𝟏⁄𝟐𝛑 �(𝟐𝐤 𝐟 (𝛂 − 𝐞 − 𝐫)𝟐 + 𝟐𝐤 𝐫 (𝛃 + 𝐞 + 𝐫)𝟐 )⁄�𝐌𝐬 (𝐊 𝟐 + 𝐫 𝟐 )� (Eq. 11.5)

𝐟𝐣 = 𝟏⁄𝟐𝛑 �(𝟐𝐤 𝐟 (𝛂 − 𝐞 + 𝐬)𝟐 + 𝟐𝐤 𝐫 (𝐬 − 𝛃 − 𝐞)𝟐 )⁄�𝐌𝐬 (𝐊 𝟐 + 𝐬 𝟐 )� (Eq. 11.6)

74
Prof. Guest initially submitted a simplified paper before the British Association for the Advancement of Science,
and in March 1926 submitted his full paper to the Institution of Automobile Engineers (English, 1906-1947).
75
Reference [2], pp. 49-50.

62
Rev B SAWE Paper No. 3521
Category Number 31.0
It should be noted that regardless of the coordinate system used (origin at CG, origin at
“SC”, etc.) the physical system characteristics, i.e., the natural frequencies and normal modes of
vibration, are always be the same. However, by selecting a particular coordinate system the
analysis can be made difficult or easy. In general, it’s best to take the origin of the coordinate
system at the c.g. as is done throughout this paper.

Figure 11.1 – THE CONJUGATE CENTERS OF PERCUSSION

63
Rev B SAWE Paper No. 3521
Category Number 31.0
12 – 1958 JAGUAR XK150S, 1980 FORD FIESTA S

Fully understanding the Principal Modes of Vibration and the Conjugate Centers of
Percussion for the general 2-DOF model requires the use of illustrative examples. 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 (1532.75 kg-m2),
so the pitch radius of gyration squared is 11.6654 ft2 (“K” = 3.42 ft, or 1.0424 m). The spring
constants at the axles 76 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:

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

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


The principal modes of vibration frequencies are:

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


= 9.15 rad/sec, or 1.46 cps

f2 = √a-(b2/(K2(c-a))) = √53.72-(17.1852/(11.6654(29.23))
= 7.27 rad/sec, or 1.16 cps

To find the node points these principal mode frequencies are input (as “radians/sec”) into:

X = b/(𝜔2-a)

X1 = 17.185/((9.15)2-53.72) = 0.57 ft (0.1737 m)

X2 = 17.185/((7.27)2-53.72) = -19.84 ft (-6.0472 m)

76
The spring constants at an axle are double the spring constants at a wheel (in bounce there are no anti-roll bar
complications). This is a necessary consequence of using a “Half-Car Model” with a total car sprung mass value as
is the custom (actually more like a “Full-Car Model”).

64
Rev B SAWE Paper No. 3521
Category Number 31.0
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 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)

β = kfl / (kf + kr) = 2364.0 (8.5) / (2364.0+2841.6) = 3.86 ft (1.1765 m)

The Conjugate Point “H” (fwd) and “J” (rr) 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

= (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)

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

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

= ((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))

= ((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

65
Rev B SAWE Paper No. 3521
Category Number 31.0
When plotted against a side elevation of the Jaguar, all these points would look as follows:

Figure 12.1 – 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. The
Coupling Coefficient was not zero, so the two motions are linked to a certain extent, which is
not desirable 77. Also, the Dynamic Index is 0.65, which is appropriate for 1958, but means that
some heterodyning can occur, and that the Conjugate Points of Percussion will not play any
role in this design.

To better understand the use of the Centers of Percussion in the design of a vehicle
suspension, we’ll have to consider the case of the Ford Fiesta S. The Fiesta was an economical
European car circa 1980, which has been described as “…the ride is very good for such a small
vehicle” 78. Colin Campbell, in his book New Directions in Suspension Design, gives the vehicle
parameters:

Condition: One-Up (Driver, No Passengers)


Sprung Mass: 1602.7 lb (727.0 kg)
Wheelbase: 7.5 ft (2.29 m)
lf: 2.78 ft (0.85 m), lr: 4.72 ft (1.44 m), K: 3.62 ft (1.10 m)
kf : 1486.8 lb/ft (2212.6 kg/m), kr: 1713.6 lb/ft (2550.1 kg/m) 79
Now to use the general equations of coupled motion (which, again, will devolve into the
special case of uncoupled motion if that is the case):

a = ((1713.6 + 1486.8)) / (1602.7/32.174) = 64.25

77
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 (see Chapter 13 “Olley’s Rules”).
78
Reference [2], page 51.
79
Ibid.

66
Rev B SAWE Paper No. 3521
Category Number 31.0

b = Coupling Coefficient = ((1713.6x4.72)-(1486.8x2.78))


/(1602.7/32.174) = 79.39 80

c = ((1713.6×4.722) + (1486.8×2.782))/((1602.7/32.172)
× 3.622) = 76.1

DI = Dynamic Index = 3.622/ (4.72×2.78) = 1.0 (percussion points @ frt


& rr axles!)

f1 = √c+(b2/(K2(c-a)))=√76.1+(79.392/(3.622(76.1-64.25))) =
10.80 rad/sec, or 1.72 cps

f2 = √a-(b2/(K2(c-a))) = √64.25-(79.392/(3.622(76.1-64.25))) =
4.86 rad/sec, or 0.77 cps

X = b/(f2-a)

X1 = 79.39/(10.802-64.25) =1.51 ft (0.46 m)

X2 = 79.39/(4.862-64.25) = -1.96 ft (-0.60 m)

The Spring Center “SC” is located:

α = krl/(kf+kr) = 1713.6x7.5/(1486.8+1713.6) = 4.02 ft, or 48.2 in (1.22 m)

β = kfl/(kf+kr) = 1486.8x7.5/(1486.8+1713.6) = 3.48 ft, or 41.8 in (1.06 m)

Since the Dynamic Index is equal to one, the Conjugate Percussion Points are over the
axles (as noted) and the associated frequencies are calculated:

c2 = αβ = 4.02 x 3.48 = 13.99 ft2 (1.29 m2)

e = α-lf = 4.02-2.78 = 1.24 ft, or 14.9 in (0.38 m)

r = (K2-e2-c2)/2e + √((K2-e2-c2)2 + 4K2e2) /2e = 2.77 ft, or 33.3 in (0.84 m)

s = K2/r = 3.622/2.77 = 4.73 ft, or 56.7 in (1.44 m)

80
Unlike the Jaguar, the Fiesta DI is equal to one, and the Coupling Coefficient is much greater than was the case
for the Jaguar XK150S.

67
Rev B SAWE Paper No. 3521
Category Number 31.0

fh = 1/2π √2kf(α-e-r)2 + 2kr(β+e+r)2/(Ms(K2+r2)) = 2.17 cps, or 130.2 cpm

fj = 1/2π √2kf(α-e+s)2 + 2kr(s-β-e)2/(Ms(K2+s2)) = 1.55 cps, or 93.0 cpm

When these results are plotted against a side elevation of the Fiesta S it looks as follows:

Figure 12.2 – 1980 FORD FIESTA S KEY VIBRATION POINTS

What we have learned here is that the Fiesta’s vibrations (conjugate) are even stiffer at
1.55 cps and 2.17 cps, but not because it’s a sports car; this is because it’s a small (light weight,
low “Ms”) economy car which can have a large variation in gross weight depending on loading.
The Coupling Coefficient was far from zero, so the two principal motions are linked to a certain
extent, which usually is not desirable. However, the Dynamic Index is 1.0, which is great not
just because no heterodyning can occur, but because the Conjugate Points of Percussion and
associated frequencies are the intended dominant factors in this design.

The Conjugate Points, “H” and “J”, are located right over the front and rear
suspensions, respectively. A bump at the front suspension “H” will cause no displacement at “J”,
just an oscillation about “J” of 1.55 cps causing the front end “H” to bounce at that speed. This
is about the same frequency as the front quarter-car model calculation would produce. Likewise,
a bump at the rear suspension “J” will cause no displacement at “H”, just an oscillation about
“H” of 2.17 cps, causing the rear to bounce as the rear quarter-car model would indicate. In
short, it is this behavior that dominates, and the consequent front-to-rear bounce frequencies of
1.55-to-2.17 cps have that fore-aft relationship which Maurice Olley would characterize as being
necessary for what he called a “flat ride” (see next chapter).

68
Rev B SAWE Paper No. 3521
Category Number 31.0
13 - OLLEY’S RULES

What is needed now are some general rules to interpret the results of such an analysis. Such
rules were provided by that pioneer of automotive dynamics, Maurice Olley 81, in the 1930’s.
Olley’s rules have been incorporated in, and superseded by, the SAE “Ride and Vibration Data
Manual” J6a 82; but often it is informative to go back and review such things in their seminal
stage 83. Olley’s approach was not just theoretical but empirical; he built a test vehicle known as
the “K2-rig” (1931) which allowed him to modify the mass properties and suspension
characteristics while studying the results. It was from such studies that Olley came up with the
“Criteria for General Suspension Cases” (not verbatim) 84:

1. The front suspension should have a 30% lower ride rate than the rear suspension, or the
Spring Center (SC) should be at least 6.5% of the wheelbase behind the c.g. (the idea
being to generally ensure that the rear bounce frequency is greater than the front).
2. The bounce frequency should be less than 1.2 times the pitch frequency to diminish the
possibility of severe interaction between the two frequencies (i.e., the Dynamic Index
“DI” should be as close to 1 as possible to prevent heterodyning).
3. Neither bounce nor pitch frequencies should be greater than 1.3 Hz (which means static
deflections would probably need to be greater than 6 inches) (See Figure 6.10 for
indication as to why this would be desirable).
4. The roll frequency should be in the same range as the pitch and bounce frequencies (the
treatment of roll frequencies is analogous to the treatment of pitch frequencies).

There was also recognition of the possibility of “Special Suspension Cases” and rules for
those cases (also not verbatim) 85:

1. When the Dynamic Index = 1, the “Conjugate Points of Percussion” are located at the
Front & Rear Axles. Many modern automobiles have a mass distribution such that the DI
is close to unity, which puts the centers of motion at the axles; this is a desirable
condition if the general criteria are also satisfied.
2. When the Spring Center “SC” is at the c.g. (the Coupling Coefficient “CC” is equal to 0),
then the principal nodes are at the c.g. and at infinity; the pitch and bounce motions are
totally independent and can result in a very irregular total motion. Some coupling tends to
even out the ride.
3. When the Dynamic Index > 1, the “bounce” center is beyond the front axle and the c.g.,
and the pitch center is between the c.g. and the rear axle; then the pitch frequency is less
than the bounce frequency and a “flat ride” will result if the Spring Center “SC” is

81
Maurice Olley left Rolls Royce to work for Cadillac in 1930. By 1952, as head of Chevrolet R&D, he went to
work on the chassis design for the 1953 Corvette, and hired Zora Arkus-Duntov, destined to become the “father” of
the modern Corvette.
82
Reference [7], page 413-470. This is the “Rev. A” of the document as revised from the original 1936 version by
the SAE Vehicle Dynamics Committee dated October 1965. The current revision may be obtained directly from the
SAE.
83
Aristotle is reputed to have said that “He who considers things in their growth and origin…will obtain the clearest
view of them”.
84
Reference [7], pages 176-177.
85
Ibid, page 178.

69
Rev B SAWE Paper No. 3521
Category Number 31.0
located far enough aft of the c.g. (front suspension cps sufficiently less than the rear
suspension cps).
4. When the Dynamic Index (DI) = 1 and the Coupling Coefficient (CC) = 0, then the pitch
and bounce frequencies are equal and the ride is adversely affected as the overall
response to road input is irregular and unpredictable.

Olley was a proponent of what he called the “flat ride”, which has been alluded to elsewhere
in this paper and could constitute Olley’s ultimate “rule”. What he meant by “flat ride” is
indicated by his famous quote (verbatim):

“I think the solution is that for a flat ride at, say, 40 mph, you use a
front end much softer than the rear. For a flat ride at 100 mph you use
almost equal front and rear deflections” 16 July 1959

“Flat ride” means minimizing pitch motion initiated by the front wheels encountering a
bump in the road. As soon as contact is made with the bump the front begins to rise in accord
with its spring-mass system natural frequency causing a growing pitch angle. Depending on
vehicle speed “V” and wheelbase “l” the rear wheels will encounter the same bump at some time
lag “tθ” after the front wheel contact. The time lag determination is:

𝒕𝜽 = 𝑽⁄𝒍 (Eq. 13.1)


The time lag will determine the phase relation between the front suspension response and
the rear suspension response. If the phase and frequency relations are just right then the initial
“pitch up” motion at the front will largely be cancelled out by the subsequent “pitch down”
motion at the rear. The resulting motion will be of a rather limited pitch movement and more like
a simple bouncing motion that will quickly diminish and die out in accord with the suspension
damping.

That the maintenance of a “flat ride” is dependent upon the velocity means that a target
velocity must chosen in accord with the character of the vehicle considered, a relatively lower
target velocity for an economy car and a relatively higher target velocity for a sports car. For a
conventional passive suspension the optimum flat ride condition will only be attained at the
target velocity, and will get further away from that optimum as velocity varies with respect to
that design speed. So, it would seem that the best target choice would be near midpoint of the
subject vehicle speed range.

Of course, the matter is also dependent on the suspension frequencies and wheelbase, but
the variation of those parameters is circumscribed by other concerns as demonstrated in this
paper. The frequencies are important factors in the mitigation of shock and vibration, the
maintenance of road contact, etc., and the wheelbase is an important factor in vehicle packaging,
directional stability, etc. So, for a conventional passive suspension, variation of these parameters
to achieve “flat ride” requires judgment and compromise 86.

86
See Reference [2], page 165 (“Interconnected and No-Roll Suspensions”) for other ways to obtain “flat ride”.

70
Rev B SAWE Paper No. 3521
Category Number 31.0
To illustrate the situation, let’s consider the familiar case of the 1958 Jaguar XK150S.
The Jaguar had a top speed of about 133 mph, so it might seem reasonable to set the flat ride
target speed at 66.5 mph (107.0 kph) (since no one is ever likely to drive at top speed, the
argument can be made for adjusting this target downward to be midpoint of a range more
representative of real use). The equation for damped simple harmonic motion used to give the
mass position “z” as a function of time “t” is 87:

𝒛 = 𝒆−ξ𝝎𝒏 𝒕 𝑨 𝐬𝐢𝐧(�𝟏 − ξ𝟐 𝝎𝒏 𝒕) (Eq. 13.2)

The front to rear spring rate relationship was varied 88 so the frequency relationship went
from “fr < ff” to “fr = ff” and finally to “fr > ff” (1.29 > 1.14, which is the actual XK150S case),
while plotting the consequent body height differential “zf – zr” and pitch angle “θ” changes:

𝒛𝒇 −𝒛𝒓
𝛉 = 𝐬𝐢𝐧−𝟏 ( ) (Eq. 13.3)
𝒍

The results clearly show the benefit of Maurice Olley’s “flat ride” (“fr > ff”) when
plotted using the characteristic parameters of a 1958 Jaguar XK150S 89, with a vehicle velocity of
66.5 mph (107.0 kph), encountering a bump of 2 inches (5.1 cm) in height. Increasing the
wheelbase “l” would increase the lag time “tθ” (Eq. 13.1), but would also increase the pitch
inertia “J” (the effect of which would not be seen here as only quarter-car models were used for
analysis 90) and decrease the pitch angle “θ” (Eq. 13.3). Overall, increasing the wheelbase, like
increasing the rolling radius, makes for a smoother riding vehicle. This is a tendency that does
much to explain the enormous size of many early luxury type vehicles (Bugatti Royal, etc.).

One last note on this matter: it is important to remember that even though a “flat ride”
situation might be obtained for the “one-up” (driver only) operating condition (as in this
example), as a vehicle’s operating condition approaches GVWR (Gross Vehicle Weight
Rating) the general tendency is for the longitudinal c.g. to shift aftward which can cause a
reversal of the left to right pattern illustrated in Figure 13.1 (which is a good example of why a
design must be evaluated across the entire envelope of operating speeds and loading conditions
in order to have a complete analysis):

87
Reference [1], pg. 26.
88
Although the usual discussion is to talk of spring rates, it is important to remember that the frequency relationship
front-to-rear can be just as affected by a sprung mass c.g. shift.
89
The ’58 Jaguar front to rear suspension frequencies of 1.14:1.29 cps compare well with General Motors 1980’s
design paradigms of 1.00:1.20 cps for passenger cars and 1.20:1.40 cps for sports cars.
90
It would be an interesting, and possibly informative, exercise to conduct the same analysis with a half-car model,
not two quarter-car models.

71
Rev B SAWE Paper No. 3521
Category Number 31.0

Figure 13.1 – MAURICE OLLEY’S “FLAT RIDE”

72
Rev B SAWE Paper No. 3521
Category Number 31.0
14 - CONCLUSIONS

Although it is not a mass properties conclusion, this paper has attempted to demonstrate
that, although the vehicle suspension is a complicated dynamic system with many degrees of
freedom, there is much that can be learned from simplified models. The key to using simplified
models successfully is to know the limitations of each model and which is the most appropriate
for the intended purpose. Despite the current trend to use ever more expensive and complicated
modeling, such as a FEM (Finite Element Model), to determine vehicle dynamic behavior, there
still is some opportunity for the small company or individual, short on bucks but long on brains,
to operate successfully in this field. To quote from an authoritative SAE text 91:

“………building and running the finite element model is complicated,


time consuming, and costly. In a ride quality analysis…..approximate
solutions are enough to determine vehicle parameters such as………..
damping & stiffness. Thus, simplified models will satisfy requirements.”

It is the use of appropriate simplified models that brings the subject most fully within the
realm of human comprehension, allowing for a true understanding of the role played by all the
parameters and, in this paper, especially the mass property parameters. The efficacy of choosing
the appropriate simplified model may be illustrated by the following example 92:

Fig. 14.1 – EFFECT OF “BOUNCE” MODEL ON SPRUNG MASS RESPONSE GAIN


TO ROAD INPUT

Note that the 1-DOF model gives a slightly less accurate value for the response gain at
sprung mass resonance, but it completely misses the response gain at the unsprung mass
“wheel hop” resonance. This latter omission is to be expected, as the 1-DOF model does not
91
Reference [6], Page 261.
92
Ibid, Page 270.

73
Rev B SAWE Paper No. 3521
Category Number 31.0
include the motion of the unsprung mass. Of course, computer programs as utilized by major
automotive manufacturers do not involve such obvious gross simplifications, but any model
utilized in a specific purpose built simulation program (as opposed to a general purpose program
like NASTRAN or other FEM program) must be subject to some simplification and thus fall
short of reality. The degree of the “short fall” is generally negligible, as long as the simulation is
run within a certain designed-in range of parameter variation, but that is hardly conducive to
innovative, ground-breaking design. The whole thrust of any engineering effort should be to go
beyond what has been done before, to make something faster, cheaper to produce, more
economical to run…

Ultimately, computer analysis is a “tool”, and the use of that tool should be guided by an
aware human intelligence. However, that human awareness is often missing. The “easy”
computerized answer is often unquestioningly seized upon, especially when under pressure. To
quote from yet another SAE text 93:

“………problems are now routinely solved with the use of computers,


running….analysis programs…. (which) are complex…requiring the
user to place a certain amount of trust in the program...it has become
obvious to us that the use of canned software has not improved the
general understanding…….in fact the opposite has occurred. “Blind
trust” in the computer has led to some amusing (and time consuming)
design flaws…”

An inherent reason that “canned software” may generate misleading results is often the
result of a confluence of a multitude of small “errors”, each the result of a reasonable
simplification for which the effect is negligible by itself. Such errors can combine and become
significant when the design envelope is being enlarged beyond the paradigm upon which the
program was based. In this connection it is an interesting observation to note that the summation
of a vehicle’s unsprung and sprung masses may not necessarily equal the total vehicle mass.
This might seem perplexing, but technically only ⅓rd of the mass of a coil spring may be
considered as part of the unsprung, or sprung, mass. This could cause confusion as taking ⅓rd
of the spring as unsprung mass and ⅓rd as sprung mass leaves ⅓rd of each spring unaccounted
for 94. For simplicity’s sake the author advocates taking ⅓rd of the suspension spring as
unsprung mass, and taking the remaining ⅔rd as sprung mass (where such an error is of less
significance).

The usefulness of that small quibble aside, the major mass properties observations are,
by chapter:

Chapter 1 – The complete model of automotive vertical accelerations, usually 10 DOF,


would generally only be amenable to analysis by computer. However, there is much to be gained
by bringing the matter into the realm of human understanding through a series of model
simplifications.

93
Reference [10], page 448.
94
Reference [15], pages 266-267. The derivation conducted in the referenced source is for a coil spring, but the
author assumes most other types of suspension springs could behave in a roughly similar fashion.

74
Rev B SAWE Paper No. 3521
Category Number 31.0

Chapter 2 - The second chapter demonstrated the significance of the sprung mass “ms”
and the unsprung mass “mus” via simplified 1-DOF quarter-car models. Such simplification is
made possible by the wide separation between sprung and unsprung mass system behavior.
Stepping slightly closer to reality, a 2-DOF quarter-car model is introduced and the first hint is
given of the significance of the unsprung-to-sprung mass ratio “χ” in linking sprung and
unsprung motions.

The exposition was advanced though the use of example calculations using “averaged”
parametric values, as if the example vehicle, a Jaguar XK150S, had the same masses, spring
rates, etc., all around, front and rear. The chapter concluded by moving away from that simplicity
of equal “ms”, “mus”, “ks”, etc. all around; the vehicle longitudinal c.g. was used to apportion the
vehicle sprung mass between front and rear, and all other parameters were set as appropriate for
the Jaguar XK150S to achieve more realistic quarter-car model results.

Chapter 3 – The second chapter takes further steps toward realism by introducing the
role damping plays in automotive vertical accelerations; in all previous analysis the damping was
taken as negligible. Mass now played a new role, in determining the critical damping level
“�4k⁄m”, which in turn affects the frequency-modifying (“fdamped = ξ fundamped”) damping ratio
“ξ”:
ξ = 𝐜 / �𝟒𝐤⁄𝐦

As noted at the end of the chapter, the fact that mass properties constitute major factors
in determining the behavior of spring-mass systems is no revelation, but that mass properties
can do so in complex and often subtle ways is already being hinted at: the mass in a spring-mass
system affects the behavior of the system through the natural frequency, but then “double dips”
in further affecting the behavior through the damping.

Chapter 4 – The “shock” or maximum acceleration “amax” experienced by the vehicle’s


sprung mass “ms” as a result of an encounter with a road surface bump of height “A” is
determined using a simple (1-DOF) sprung mass-suspension spring model:

𝐚𝐦𝐚𝐱 = (𝐤 𝐬 ⁄𝐦𝐬 ) 𝐀

Increasing the sprung mass “ms” will obviously significantly decrease the road shock,
which is why a fully laden car will often seem to ride smoother, but to increase vehicle sprung
mass for the sake of ride is like adding a ton of ballast to balance an aircraft, not good design
practice; increased vehicle weight has adverse effects on acceleration, maneuverability, fuel
economy, etc.

However, increasing “ms” does reduce the unsprung-to-sprung mass ratio “χ”, so the
question naturally arose: “How about the unsprung mass; can changing “mus” be any help?”.
Using the exact (2 DOF) sprung mass-suspension spring quarter-car model, which includes
the effect of the unsprung mass, it was determined that increasing the unsprung mass has only
the most negligible effect in decreasing shock transmission to the sprung mass, so this is yet

75
Rev B SAWE Paper No. 3521
Category Number 31.0
another design path not to be taken. Still, given the emphasis in the conventional wisdom on
reducing the unsprung-to-sprung mass ratio, it is a bit of a surprise that taking an action that
would increase that ratio would have any positive effect at all 95!

Chapter 5 – Using the simple (2-DOF) unsprung mass-suspension spring quarter-car


model, two resulting road contact quantifiers are presented, “l” and “d”:

𝐕𝛕
𝐥=� � �𝐦𝐮𝐬 ⁄𝐦𝐬
𝟐
𝐦𝐮𝐬
𝐝 = 𝐝𝐬 (𝟏 + )
𝐦𝐬

For maintaining good road contact, “l” must be minimized and “d” must be maximized.
However, reducing “mus” improves “l” while worsening “d”. Fortunately, the effects are not
proportionate; the improvement in “l” is more than 4 times the degeneration in “d”; so a decrease
in “mus” is indicated, but again is not exactly the unalloyed virtue of conventional wisdom. A
pattern seems to be emerging, indicating that reducing the sprung-to-unsprung weight may be
generally beneficial as per the conventional wisdom, but in ways more complex than that
“wisdom” usually imparts.

Chapter 6 – For minimizing road vibration transmission to the sprung mass, reducing
the unsprung mass has a definite, but not overwhelming, beneficial effect. For minimizing
engine, transmission, etc. vibration effect on the sprung mass, reducing (or increasing) the
unsprung mass has virtually no effect at all. However, for minimizing wheel, tire, brake, etc.,
vibration transmission to the sprung mass no simple rule can be established with respect to the
unsprung mass because the situation is so dependent on the exact nature of the disturbance,
whether it is due to imbalance, misalignment, out-of-round, or radial stiffness variation 96.

Chapter 7 - Gyroscopic torque reactions “T” due to forced precession of a rotating


mass is an area that often seems to be overlooked, but can be significant given sufficient
magnitudes of inertia “I” and angular velocities “𝜔r” and “𝜔p”:

T = I 𝜔r 𝜔p

There are generally two main sources of automotive gyroscopic reactions: the rotating
portion of the engine/drive train, and the rotating portion of the unsprung masses. The
orientation of the engine can be arranged so as to take a positive advantage of its torque
reactions, but this is not possible with the rotating portion of the unsprung masses 97. The best
that can be done is to minimize the adverse gyroscopic effects on maneuverability, ride, and
95
Maurice Olley took issue with the idea that reduction of unsprung mass was an unalloyed virtue, and speculated
that increasing unsprung mass might to some extent reduce shock transmission to the sprung mass (Reference
[10], page 314). For a complete understanding of his complex thoughts, see the source cited.
96
However, generally it would seem a reduction in rolling radius size would be beneficial.
97
Vehicles don’t function well unless the wheels point fore-and-aft.

76
Rev B SAWE Paper No. 3521
Category Number 31.0
control by minimizing the rotating portion of the unsprung masses (which will also benefit
acceleration, braking, fuel economy).

Chapter 8 – The introduction of the 2-DOF “Bounce & Pitch” half-car model brings to
the fore the significance of another mass property, the sprung mass pitch radius of gyration
(the longitudinal c.g. of the sprung mass is also significant, but that was established back at the
end of Chapter 1). The pitch radius of “K” is significant in determining pitch oscillation and in
calculation of a key quantity known as the “Dynamic Index” (“DI”) (and the “lf” and “lr”
distances testify to the significance of the c.g.):

DI = K2 / (lf x lr)
The significance of the Dynamic Index is as an indicator of the possibility of interaction
between the pitch and bounce frequencies such as to give rise to a “beat” frequency, which is
something to be avoided due to its adverse effect on human comfort. Designers traditionally have
attempted to get the DI as close to one as possible, but modifying “K” to do so affects the radii
of gyration in the other two rotational directions: roll and yaw. Therefore, modifying “K” may
be circumscribed by performance considerations other than “ride” (e.g.: maneuverability,
directional stability).

Chapter 9 – The half-car model, having two degrees of freedom, has two principal
modes of sprung mass motion. In determining these two principal modes of motion another
mass property is high-lighted: the sprung mass moment of inertia in pitch “J” (a reflection of
mass properties previously introduced: “J = Ms K2”). The equations for the principal modes of
motion are derived from the sprung mass free body dynamic equilibrium equations and form
two distinct sets. One set is the general equations of motion for the two principal modes 98:

𝟏
f1 = √[a-(b2/(K2(c-a)))] cps
𝟐𝝅

𝟏
f2 = √[c+(b2/(K2(c-a)))] cps
𝟐𝝅

Each of these modes constitutes a mix of translational and rotational motion. When the
“Coupling Coefficient” (“CC”) is equal to zero…

(krlr-kflf)/Ms = 0

…then the general equations for the principal modes of motion devolve into the special
set of two uncoupled motions, where one oscillatory motion “fz” is at a node of infinite distance
away thus producing a pure translation (“bounce”) of the sprung mass, and the other motion
“fθ” is a pure rotation (“pitch”) about the c.g. The resulting special uncoupled equation set is:

98
For definition of the “a”, “b”, and “c” coefficients return to Chapter 9, page 59.

77
Rev B SAWE Paper No. 3521
Category Number 31.0
𝟏 𝟏
fz = √[(kr+kf)/Ms] , or √a cps
𝟐𝝅 𝟐𝝅

𝟏 𝟏
fθ = √[(krlr2+kflf2)/J] , or √c cps
𝟐𝝅 𝟐𝝅

From inspection of the Coupling Coefficient equation it can be seen that it is a special
placement of the sprung c.g. (“krlr-kflf = 0”) that brings this uncoupled case about. The
frequencies of the motions, whether general or special case, are determined by the mass
property parameters “Ms”,” lr”,”lf”, and “J”.

Chapter 10 – The sprung mass c.g., and the node locations of the Principal Modes, all
constitute special points along the sprung mass longitudinal axis, but there are other special
points which warrant attention. The “Spring Center” (“SC”) is one of these other points. Unlike
the other special points discussed in this paper, the location of the “SC” is not determined or
influenced by any mass properties parameter:

α = krl/(kf+kr)

β = kfl/(kf+kr)

The “α” and “β” are the “SC” locating distances with respect to the front and rear axle-
lines, respectively. The “SC” seems to be a node of uncoupled motion when acted upon by a
force (“static uncoupling”), but when released from that force will respond with a bounce plus
pitch motion (“dynamic coupling”). Perhaps the true significance of the “SC” is that it is
important to the calculation of those other special node points known as the “Conjugate Centers
of Percussion”.

Chapter 11 – The “Conjugate Centers of Percussion” are also known as the “Double
Conjugate Points”. When the DI is equal to one then these points are located directly over the
wheel centers. When this is the case, then a disturbance at the front suspension will generate a
pitching motion about the rear point, and a disturbance at the rear suspension will generate a
pitching about the front point. The corresponding bouncing motions at the front and rear will be
exactly in accord with quarter-car model predictions.

Placing the Conjugate Centers of Percussion over the wheels is a good stratagem for
obtaining a good ride in a small car design; when the conjugate centers are not over the wheels in
a design then they are of no interest. The location values “r” and “s” for the conjugate centers are
determined by the “SC” locators “α” and “β” , plus the mass property parameters “Ms”,” lr”,”lf”,
and “J” 99:
r = (K2-e2-c2)/2e + √((K2-e2-c2)2 + 4K2e2) /2e
s = K2/r

99
For definition of “e” and “c” coefficients return to Chapter 11, page 61.

78
Rev B SAWE Paper No. 3521
Category Number 31.0
Chapter 12 – A full 2-DOF half-car model analysis is provided of the 1958 Jaguar
XK150S and the 1980 Ford Fiesta S as illustrative examples of how to use the equations
presented in previous chapters. The Jaguar proved to be fairly stiffly sprung, as befits a
traditional sports car, with a “flat ride” relationship set for high speed (over 65.5 mph / 107.0
kph):
Front: 1.20 cps Rear: 1.54 cps

The Jaguar had a Dynamic Index of 0.72, which is quite good for the 1958 time frame,
and so the Conjugate Centers of Percussion were not in-line with the wheels and therefore of no
interest. The Coupling Coefficient was 43.7, so the Jaguar did not represent that special case
where the sprung mass rotated about the c.g. and/or bounced about some center infinitely distant.

The Ford Fiesta S also proved to be fairly stiffly sprung, the consequence of being a little
economy car whose sprung mass could increase tremendously with load (Curb Wt = 730 kg,
GVWR = 1160 kg), with a “flat ride” relationship 100 seemingly set for a target speed of around
80.0 mph (128.7 kph), which is surprising for a car that could barely manage 89.0 mph (143.0
kph) full out 101:

Front: 1.55 cps Rear: 2.17 cps

The Ford had a Dynamic Index of 1.00, which was not only perfect, but meant the
Conjugate Centers of Percussion were in-line with the wheels, and therefore it was the conjugate
center frequencies that were of interest (and shown above). The Coupling Coefficient was 158.8,
so the Ford also did not represent that special case where the unsprung mass rotated about the
c.g. and/or bounced up and down about some center infinitely distant.

Chapter 13 – “Olley’s Rules” are provided for interpreting the results of any suspension
analysis, or for guiding the design of any suspension. The rules are too lengthy and involved to
summarize here, but the matter of “flat ride” merits some comment. A “flat ride” will occur only
at one precise point in the vehicle operating envelope; that is to say, at one particular weight,
c.g., and speed. If that point is well chosen, then a condition near to “flat ride” can be attained
throughout much of the operational envelope for a conventional passive suspension system.

Chapter 14 - In conclusion, the significance of certain mass properties has been


demonstrated throughout this paper with regard to every aspect of automotive vertical
accelerations: road shock, road contact, road vibration, and ride motions; much of which relates
to what automotive designers call NVH (Noise, Vibration, and Harshness). The particular
mass properties concerned have proven to be: the unsprung mass “Ms”, the sprung mass(es)
“mus” (two parameters which tend to get reduced to the mass ratio “χ”), the sprung mass
longitudinal c.g. (generally located by “a” and “b” dimensions), and the sprung mass pitch

100
The 1980 Ford Fiesta S (1.1L) “flat ride” calculations are not included in this paper.
101
The spring rates and the target speed for the “flat ride” suggest that the Fiesta was really intended to operationally
be a car with a lot of “junk in the trunk”, whereas the Jaguar would be at full capacity with two people and a couple
of golf bags. These were two vehicles designed for very different markets, and the results reflect that dichotomy.

79
Rev B SAWE Paper No. 3521
Category Number 31.0
inertia represented by “J” or “Ms K2” 102. What is important to remember is not only the
significance of those mass properties, but the fact that most of those properties will vary in
operation, and that variance must be provided for. An analysis must not just be conducted for the
“one up” condition, but for every conceivable loading condition at various speed points
throughout the operational envelope. That results in a lot of analysis, and a lot of compromise
and “juggling” of design parameters, especially when other performance aspects, cost, reliability,
safety, etc., are brought into consideration. However, with sufficient knowledge and appropriate
analysis tools, it can be amazing what is achievable by human thought processes alone, before
ever resorting to a computer.

102
Still putting aside the matter of oscillations in roll and the significance of lateral weight distribution (usually
close to even) and the roll inertia, etc.

80
Rev B SAWE Paper No. 3521
Category Number 31.0
REFERENCES

[1] Bastow, Donald; Car Suspension and Handling, Plymouth, Devon; Pentech Press
Ltd., 1980.

[2] Campbell, Colin, New Directions in Suspension Design, Cambridge, MA; Robert
Bentley Inc., 1981.

[3] Campbell, Colin, The Sports Car, Cambridge, MA; Robert Bentley Inc., 1969.

[4] Crossley, F.R. Erskine; Dynamics in Machines, New York, NY; The Ronald Press
Co., 1954.

[5] Dixon, J.C.; Suspension Geometry and Computation, Hoboken, NJ; John Wiley &
Sons Ltd, 2009.

[6] Dukkipatti, Rao V.; Jian Pang, Mohamad S. Qatu, Gang Sheng, and Zuo Shuguang;
Road Vehicle Dynamics, Warrendale, PA; SAE R-366, 2008.

[7] Gillespie, Thomas D.; Fundamentals of Vehicle Dynamics, Warrendale, PA; SAE R-
114, 1992.

[8] Gillespie, Thomas D.; “Technical Considerations in the Worldwide Standardization


of Road Roughness Measurement”, Highway Safety Research Institute, The
University of Michigan, Report #UM-HSRI-81-28, 1981.

[9] Judge, Arthur W.; The Mechanism of the Car, Cambridge, MA; Robert Bentley Inc.,
1966.

[10] Milliken, William F.; and Douglas L. Milliken, Chassis Design, Principles and
Analysis (Based on…Notes by Maurice Olley), Warrendale, PA; SAE R-206, 2002.

[11] Milliken, William F.; and Douglas L. Milliken, Race Car Vehicle Dynamics,
Warrendale, PA; SAE R-146, 1995.

[12] Newton, K.; W. Steeds, T.K. Garrett; The Motor Vehicle, Oxford, Butterworth-
Heinemann Linacre House, SAE R-298, 1996.

[13] Reinpell, J.; H. Stoll, and J. W. Betzler, The Automotive Chassis, Oxford, Elsevier
Linacre House, SAE R-300, 2001.

[14] Riley, Robert Q.; “Automobile Ride, Handling, and Suspension Design; With
Implications for Low-Mass Vehicles”, Phoenix, AZ; Robert Q. Riley Enterprises,
1999.

81
Rev B SAWE Paper No. 3521
Category Number 31.0
[15] Sears, Francis Weston; and Mark W. Zemansky, University Physics, Reading, MA;
Addison-Wesley Publishing Co., 1967.

[16] Steeds, W.; Mechanics of Road Vehicles, London, Iliffe & Sons Ltd., 1960.

[17] Thomson, William T.; Vibration Theory and Applications, Englewood Cliffs, NJ;
Prentice-Hall Inc., 1965.

[18] Türkay, Semiha; and Hüseyin Akҫay, “Influence of Tire Damping on the Ride
Potential of Quarter-Car Active Suspensions”, Eskişehir, Turkey; Anadolu
University, 2009.

[19] Wiegand, B.P.; “The Weight and C.G. Implications of Obtaining Maximum
Automotive Lateral Acceleration Levels”, SAWE Journal Weight Engineering, pages
10-12, Winter ’82-’83.

[20] Wiegand, B.P.; “Mass Properties and Automotive Longitudinal Acceleration”, Los
Angeles, CA; SAWE #1634, 1984.

[21] Wiegand, B.P.; “The Basic Algorithms of Mass Properties Analysis & Control”, Los
Angeles, CA; SAWE #2067, 1992.

[22] Wiegand, B.P.; “Automotive Mass Properties Estimation”, Los Angeles, CA; SAWE
#3490, 2010.

82
Rev B SAWE Paper No. 3521
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 and of the Society of Automotive
Engineers. He has presented three SAWE papers: “Mass Properties and Automotive
Longitudinal Acceleration” (SAWE #1634, 1984), “The Basic Algorithms of Mass Properties
Analysis & Control” (SAWE #2067, 1992), and “Automotive Mass Properties Estimation”
(SAWE #3490, 2010). He has also published two articles: “The Weight and C.G. Implications of
Obtaining Maximum Automotive Lateral Acceleration Levels” (SAWE Journal Weight
Engineering, Winter 1982-’83), and “The Mystery of Automotive POI Values” (SAWE Journal
Weight Engineering, Spring 2011). Recently he reaffirmed his long-standing interest in
automotive engineering by attending the SAE Seminar “Vehicle Dynamics for Passenger Cars
and Light Trucks” (Troy, MI; August 11-13, 2009)

83
Rev B SAWE Paper No. 3521
Category Number 31.0

APPENDICES

84
Rev B SAWE Paper No. 3521
Category Number 31.0
APPENDIX A - SYMBOLISM

A The amplitude of a mass in Simple Harmonic Motion (one cycle of


motion could be from +A to –A and back to +A). Also the height of a road
shock step input.

a A pitch motion principal frequency equation coefficient meaning


“(kr+kf)/ms”, or a general acceleration as in “F = ma”, or the distance
from the front axle-plane to the c.g. (SAE J670e).

amax The maximum acceleration of a mass in SHM (Simple Harmonic Motion).

as The acceleration of the sprung mass.

aus The acceleration of the unsprung mass.

α The Greek lower case letter “alpha” used to symbolize the angular
acceleration of a rotating mass, or the distance from the front axle to the
Spring Center (“SC”) (SAE 670e).

BOF Stands for “Body on Frame”, as opposed to unitary or monocoque


construction.

b A pitch/bounce motion principal frequency equation coefficient, called the


“Coupling Coefficient”, meaning “(krlr-kflf)/ms”, or the distance from the
rear axle-plane to the c.g. (SAE J670e).

β The Greek lower case letter “beta” used to symbolize the distance from
the Spring Center (“SC”) to the rear axle (SAE J670e).

C The consolidated symbolism in the sprung mass vibration response gain


Equations 6.5 through 6.7 meaning “cs/ms”

CC The pitch/bounce motion principal frequency equation coefficient, called


the “Coupling Coefficient”, meaning: “(krlr-kflf)/ms”. When “krlr” equals
“kflf” then the sprung mass bounce and pitch motions are uncoupled (see
Fig. 7.4).

CG The center of gravity of the total vehicle.

cg The center of gravity of the vehicle sprung mass.

c A pitch/bounce motion principal frequency equation coefficient meaning


“(krlr2+kflf2)/Js”, or a damping coefficient (lb-sec/in or N-sec/m units).

85
Rev B SAWE Paper No. 3521
Category Number 31.0
c2 The conjugate centers of percussion, a.k.a. “double conjugate points”,
coefficient meaning: αβ.

cf, cr The effective damping coefficients for the front and rear suspensions,
respectively (lb-sec/in or N-sec/m units).

cs The effective damping coefficient of the “shock absorber” in the


suspension (lb-sec/in or N-sec/m units).

ct The effective damping coefficient accounting for the hysteresis damping


effect in a tire (lb-sec/in or N-sec/m units).

DI The “Dynamic Index” of the sprung mass bounce/pitch oscillations and


equal to “ky2/(lf x lr)”; it is an indicator of possible heterodyning action.

DOF The degrees of freedom in a dynamic situation, which is the number of


translation motions plus the number of rotational motions necessary to
fully describe the situation.

d The maximum depth of a dip in the road surface that can be traversed
without the tire losing contact with the road surface; see Fig. 5.1.

dt A time differential as in “dx/dt”.

d1, d2 Distances from one point to another point.

dt The static deflection of the tire resulting from bearing its portion of the
sprung weight and unsprung weight (mass).

ds The static deflection of a suspension spring resulting from bearing its


portion of the sprung weight (mass).

e The distance to the Spring Center from the sprung mass Center of
Gravity. Also represents the natural logarithmic base (Euler’s constant).

ξ The Greek lower case letter “xi” used to symbolize the damping ratio of a
sprung mass system; see Eq. 3.2.

F A force, as in “F = m a”.

F1, F2 Specific forces applied at points “1” and “2” respectively.

Fspring The force inflicted by or on a spring.

Fs, Fus A force input directly inflicted upon the sprung mass and the unsprung
mass, respectively.

86
Rev B SAWE Paper No. 3521
Category Number 31.0

f A frequency of translational oscillation, usually measured as the number


of cycles per unit of time.

fθ A frequency of rotational oscillation, usually measured as the number of


radians per unit of time.

f1, f2 The frequencies of the two principal modes of vibration of the sprung
mass.

fh, fj The frequencies of the of vibration at the conjugate points “H” and “J”,
respectively.

fn A natural, a.k.a. resonance, frequency which is a fundamental


characteristic of a spring-mass system.

fdamped The natural frequency of a spring-mass system with damping being taken
into account.

fundampted The natural frequency of a spring-mass system without damping being


taken into account.

fs, fus The natural, or resonance, frequencies of the sprung mass and the
unsprung mass systems, respectively.

fhopus The natural, or resonance, frequency of the unsprung mass and tire
“spring” system.

f z, f θ The bounce and pitch frequencies for the 2 DOF automotive model.

GVWR This stands for “Gross Vehicle Weight Rating”.

Go The roughness magnitude for a road, usually 1.25x105 ft2/cycle/ft to


1.25x106 ft2/cycle/ft; see Eq. 6.1.

Gz(Ω) The PSD amplitude (ft2/cycle/ft) as a function of the spacial (spatial)


frequency (Ω).

Gzr(f) The road vertical acceleration PSD input at a velocity as a function of


frequency.

Gzs(f) The sprung mass vertical acceleration PSD response at a velocity as a


function of frequency.

g The gravitational constant in the English FPS System of Units 32.174


ft/sec2, or IPS System of Units 386.088 in/sec2 (standard values).

87
Rev B SAWE Paper No. 3521
Category Number 31.0

H The forward conjugate point.

Hz The “Hertz” or cycles per second; named after Heinrich Rudolf Hertz
(1857-1894).

Hzs(f) The sprung mass vertical gain or transmissibility factor at a velocity as a


function of frequency.

ISO The International Standards Organization (actually, the “International


Organization for Standardization”; apparently there is a need for a
standard in forming an acronym) is a network of national standards
institutions from 160 countries with headquarters in Geneva, Switzerland.

Ix, Iy, Iz The mass moments of inertia about the “X”, “Y”, and “Z” axes
respectively.

I, Ip A rotational or polar moment of inertia, units: lb-ft2 or kg-m2.

IFS Stands for “Independent Front Suspension”.

IRS Stands for “Independent Rear Suspension”.

J The vehicle mass moment of inertia about the pitch axis (“J” = “Iy”), or
the aftward conjugate point.

j The imaginary number operator “√-1”.

K The pitch radius of gyration of the sprung mass.

K1, K2 A consolidated symbolism in the sprung mass vibration response gain


Equations 6.5 through 6.7 meaning the quantities “ks/ms” and “kt/ms”,
respectively.

kcs The combined spring rate of the suspension effective spring rates in a
series combination (Eq. 2.4).

kcp The combined spring rate of the suspension effective spring rates in a
parallel combination (Eq. 2.5).

kf , kr The effective spring rates per axle at the front and rear suspensions,
respectively; units lb/in or N/m.

ks The effective suspension spring rate per wheel in a quarter-car model,


units lb/in or N/m.

88
Rev B SAWE Paper No. 3521
Category Number 31.0
kt The effective tire spring rate at a particular inflation pressure, units lb/in or
N/m.

l Length in general, a wheelbase, or the minimum length of a dip in the road


surface traversable without loss of firm tire/road contact.

l1, l2 General lengths between points.

lf, lr The longitudinal distances from the front axle to the c.g. and from the c.g.
to the rear axle, respectively.

λ The Greek lower case letter “lambda”, used to symbolize wave length
(distance/cycle).

Ms The total sprung mass of a vehicle.

m A mass, as in “F = ma”.

ms, mus The quarter-car model sprung mass (“ms = Ms/4” assuming equal mass
distribution all around), and unsprung mass, respectively.

mus/ms The unsprung to sprung mass ratio at a wheel.

N The normal load, as in “F = μ N” (Coulomb’s friction law).

o Signifies a point of origin, or the point about which moments are


calculated.

θ The Greek lower case letter “theta” symbolizing an angle (radians), often
the pitch angle of the sprung mass.

θ̇ Angular velocity (radians/second), often of the sprung mass in pitch


motion.

θ̈ Angular acceleration, same as “α” (radians/second2), often of the sprung


mass in pitch motion.

PSD This stands for “Power Spectral Density”, which is a plot of measured
road roughness power (in2/cycle/ft) across a range of spatial frequencies
(cycles/foot) for a road which, for a particular vehicle velocity (ft/sec) of
interest, allows for conversion into the time domain (cycles/ft x ft/sec =
cycles/sec).

π The Greek lower case letter “pi”, used to symbolize the standard rounded
off value of “3.14159”.

89
Rev B SAWE Paper No. 3521
Category Number 31.0
r The radius of a rotating mass, or the distance from the sprung mass center
of gravity to the forward center of percussion or conjugate point.

SAE The Society of Automotive Engineers International is a global


association of over 128,000 engineers in the aerospace and automotive
industries.

SAWE The Society of Allied Weight Engineers in an international organization


of 22 chapters with members from the United States, Europe, United
Kingdom, and Canada. Its purpose is to promote that specialized branch of
engineering known as “mass properties analysis and control” in the
aerospace, marine, and automotive industries.

SC The “Spring Center” of the sprung mass where “krβ=kfα”.

SHM This stands for “Simple Harmonic Motion”, or the free motion of a
spring/mass system.

s The distance from the sprung mass center of gravity to the aftward
center of percussion or conjugate point.

T A torque, as in “T = I α”, or “T = I ωsωp (see Eq. 7.1).

t Time in general, or an elapsed quantity of time.

tθ The time differential between front and rear wheel inputs, usually in
seconds, and equal to “V/ l”.

τ The Greek lower case letter “tau”, used to symbolize the “period”, i.e., the
elapsed time for an oscillation to complete one cycle (seconds).

μ The Greek lower case letter “mu” symbolizing a coefficient of friction.

μs,μd The static and the dynamic coefficients of friction.

V The vehicle velocity, mph or kph.

Ω The Greek upper case letter “omega” used to represent the wave number
(cycles/distance), which is the inverse of wave length (distance/cycle),
a.k.a. “spatial frequency”.

Ωο The limiting spatial frequency for road input: 0.05 cycle/ft (asphalt) to
0.02 cycle/ft (concrete).

W A weight, generally in lb or kg.

90
Rev B SAWE Paper No. 3521
Category Number 31.0
Wf, Wr The vehicle weight on the front axle and the vehicle weight on the rear
axle, respectively.

Ws The total vehicle sprung weight.

Wt The total vehicle curb weight.

Wus The total vehicle unsprung weight.

Wusf, Wusr The total unsprung weight at the front axle and the total unsprung
weight at the rear axle, respectively.

wusf, wusr The unsprung weight at a front wheel and the unsprung weight at a rear
wheel, respectively.

ws, wus The sprung weight at a wheel and the unsprung weight at a wheel,
respectively.

ω The Greek lower case letter “omega” symbolizes angular velocity


(radians/second): ω = 2πf.

ωp The Greek lower case letter “omega” with a subscript “p” symbolizes the
angular velocity of the precession of a rotating mass.

ωr The Greek lower case letter “omega” with a subscript “r” symbolizes the
angular velocity of a rotating mass.

X The longitudinal distance traveled by a vehicle along a road in a certain


amount of time, or a longitudinal location.

x A factor that divides up the effect of the tire spring quality between the
automotive sprung mass and the unsprung mass systems.

χ The Greek lower case letter “chi” which is used to symbolize the
unsprung mass to sprung mass ratio: “mus/ms”.

Z The road roughness amplitude (height); see Eq. 6.2. Also the vertical
location or direction.

zf, zr The vertical coordinate or displacement of the sprung mass at the front or
rear quarter-car model respectively.

zr, zs, zus The vertical displacement of the road, the sprung mass, and the
unsprung mass; respectively.

91
Rev B SAWE Paper No. 3521
Category Number 31.0
ż s , ż us The vertical velocity of the sprung and the unsprung masses,
respectively.

z̈ s , z̈ us The vertical acceleration of the sprung and the unsprung masses,


respectively.

92
Rev B SAWE Paper No. 3521
Category Number 31.0
APPENDIX B – ROLLING RADIUS AND ROAD SHOCK

An analytical consideration of what happens when a wheel encounters a bump in the road
gives insight into the mechanics of diminishing the consequent road shock. In general, the
greater the size, and/or the greater the compliance, of the wheel then the less shock is transmitted
to the vehicle. The method of mounting of wheel to vehicle also has a great effect on the
transmission of road shock, but that is a matter of suspension design, which is to be treated as a
separate and very complex subject in its own right. Herein, size and elasticity will be considered
only as it pertains to the design of the wheel itself, and all analysis is limited to that subject 103.

To facilitate the investigation of wheel design for the minimization of road shock two
closely related computer programs, named “FLEXIWHL.BAS” and “RIGIDWHL.BAS”, were
created. This two-fold approach was necessitated by the fact that elastic wheel, i.e. a tire
mounted on a rim, behavior is determined by the mechanics of deformation, whereas rigid wheel
behavior is mainly a matter of geometry. It should be possible to merge these disparate
approaches into a unified program, but that will require a future effort. In either case, various
parametric values associated with a simplified wheel-bump model were input to see what the
effect on road shock would be; this paper serves as documentation for those programs and as a
summary of the knowledge thereby obtained regarding wheel design.

The Parameters

For this analysis it is assumed that a wheel of rolling radius “RR”, moving with a
velocity at the hub “V”, encounters a step in the road of height “H”. The use of a step function as
a paradigm of a road disturbance forcing function input may be questionable. However, such
functions do occur on a fairly frequent basis in reality, are easy to model, and the results obtained
are useful for comparative evaluations of wheel performance; wheels that can handle a step input
will do even better with regard to half-sinusoidal, ramp, or other inputs. The chief effect of a
“sharp edged” input, such as a step function, is that it reduces the effective spring constant of a
tire; it’s a matter of the relatively limited contact area involved. To crudely account for the effect
of varying road bump types a form factor “FF” was introduced into the mathematical model, but
the effect of “FF” parameter variation was not studied as there was no associated empirical data
available for comparison.

Other parameters required for this analysis include “N”, the normal load which the wheel
supports, and “WEFF”, which is the effective weight of the vehicle. For an automobile this last
would be dependent upon the gear ratio employed at the moment of impact, which in turn would
largely be dependent on the vehicle velocity. For a simple cart the effective weight would simply
be the weight plus the rotational inertia contribution of the wheels and any other rotating parts.

That brings up a significant aspect of this analysis, which is that the original program
“ROADSTEP.BAS”, from which everything else developed, was designed to be able to handle
the input of parameters varying all the way from that representative of the most primitive cart to

103
It is curious to note that on a road of sufficient smoothness a rigid wheel will have less rolling resistance than a
tire/wheel (hence the efficiency of trains). However, the rigid wheel does little to attenuate vibration or provide
traction, and is definitely far inferior when road irregularities of significant size are encountered.

93
Rev B SAWE Paper No. 3521
Category Number 31.0
that representative of the modern automobile (this proved unmanageable later). The given
parametric input is:

P = Tire inflation pressure, PSI (for non-pneumatic models this may be input as “0”).

K = Wheel (tire) spring rate, Lbs/Inch (for pneumatic models this is related to “P”).

Do = Deflection constant, Inches (generally for pneumatic model, otherwise very small
or zero).

Ri = Tire inflated radius, Inches (for non-pneumatic models this is the no-load radius).

N = Normal load, Pounds (this the load supported by the wheel in question).

Vo = Initial (at contact) velocity of vehicle, MPH (the wheel hub is assumed to move at
this value).

Weff = Effective weight of vehicle, Pounds (for an automobile this is a complex


determination).

H = Road bump height above the general ground plane, Inches.

FF = Form factor associated with the road bump, dimensionless (for a step function
bump the “FF” is taken as 2).

From the above set of given input parameters certain explicitly calculated (closed form)
parameters are directly generated:

dT = Appropriate time increment size, Seconds (based on “V”, required for simulation).

Rr = Rolling radius, Inches (this is the no-load radius less the deflection under normal
load).

Zo = Initial (at contact) height of wheel hub above ground plane, Inches.

Xo = Initial (at contact) longitudinal distance of wheel hub from road step at contact,
Inches.

ANGo = Initial (at contact) angle between the resultant acceleration vector and vertical,
Radians.

From the above total (given & calculated) set of input parameters the following implicitly
calculated, time-step driven computer simulation output parameters are obtained:

I = Computer simulation time-step counter, Iterations.

94
Rev B SAWE Paper No. 3521
Category Number 31.0
T = Elapsed time since contact, Seconds.

V = Velocity of vehicle since contact, MPH (the wheel hub is assumed to move at this
value).

X = Longitudinal post initial contact distance of wheel hub from “Xo”, Inches.

Z = Vertical distance of wheel hub from ground plane since contact, Inches.

Ax = Longitudinal acceleration at wheel hub, G’s.

Az = Vertical acceleration at wheel hub, G’s.

The Physical Model

The analysis began with a consideration of a conventional automobile wheel/tire moving


toward a bump in the road, modeled as a step. Upon contact with the step the pneumo-elastic
nature of the tire is such that the road surface irregularity tends to be engulfed by the tire,
accompanied by a consequent compression of the tire and a climbing motion; there are both
longitudinal and vertical compressions and accelerations generated. For a visualization of the
situation see the following:

This initial condition of both longitudinal and vertical accelerations occurring as the
wheel attempts to climb the step, which is from first contact to when the step is surmounted (as
per the illustration), is termed “Phase 1”. After the step is completely surmounted comes a
“Phase 2” during which there is no longitudinal deceleration (other than roll resistance, which is
neglected for this analysis); there is only some vertical acceleration as a result of residual
compression. There may also be a “Phase 3” and a “Phase 4”, but those are neglected herein as
not significantly worthwhile to model, given the limited scope of this analysis.

Phase 2 terminates with the expansion of the wheel back to its initial rolling radius (or
close approximation). Phase 3, if modeled, would have represented a further expansion, or
“overshoot”, beyond this equilibrium position, while Phase 4 would have represented a return to
equilibrium position. As Phase 3/Phase 4 behavior would involve consideration of vehicle/

95
Rev B SAWE Paper No. 3521
Category Number 31.0
suspension characteristics it would be outside the scope of a consideration limited to only wheel
behavior.

The previous elastic situation, redrawn to represent a rigid wheel, may be illustrated as
follows:

The Program(s)

When the wheel first makes contact with the step the initial conditions are:

T = 0, Fx = 0, Ax = 0, Fz = 0, Az = 0

Other initial parameter values, other than those given, are established as follows. The
rolling radius is determined from the wheel’s unloaded radius, spring rate, and deflection
constant, plus the normal load supported by the wheel:

Rr = Ri – (N / K + Do)

The initial vertical coordinate, “Zo”, of the wheel hub is equal to “Rr”. The longitudinal
distance of the wheel hub to the step face at initial contact, “Xo”, is dependent simply on
geometry; in particular it is dependent on the dimensions “Ri”, “Rr”, and “H”:

Xo = √Ri2 – (Rr – H)2

There is also the initial angle the resultant acceleration vector makes with the vertical
reference axis, which is determined by:

ANGo = arcsin ((Rr-H)/Ri)

This leaves only the time increment size, “dT”, the appropriate value of which depends
inversely on the magnitude of “V” (if the vehicle is traveling very fast an inappropriately large

96
Rev B SAWE Paper No. 3521
Category Number 31.0
value of “dT” will result in output insufficiently precise to be of value, and if the vehicle is
traveling very slow then an inappropriately small value of “dT” will result in an unnecessarily
slow running simulation). The value of “dT” is determined by the reciprocal of the velocity
transformed from miles per hour to units of inches per second:

dT = 1 / (V x (5280/3600) x 12)

Other than the above starting parameter calculations, the elastic program is dependent
upon two simple algorithms, one “longitudinal compression/motion” and one “vertical
compression/motion”, for its continuing method. Initially in the construction of this program
consideration was given to the idea of utilizing the ideal gas law relationship, “p v = n R T”
(pressure times volume equals the moles of a gas times the gas constant times temperature) for
the method basis, but this would only be valid for wheels with pneumatic tires. Instead, by
basing the program on spring rates, the result could have validity for wheels of all types: “rigid”
(e.g.: wood with steel rim), elastic (e.g.: wood or steel with a solid rubber “tire”), or modern
pneumatic (e.g.: steel or other metal with a pneumatic “rubber” tire). The core result is
essentially as follows (in GW Basic, a complete program listing follows later) for Phase 1:

781 FWD=((V*12*5280/3600)*DT:REM VELOCITY DISTANCE INCREMENT


782 IF I=1 THEN AX=(FWD*K)/(WEFF*2):REM INITIAL LONGITUDINAL
DECELERATION
783 AFT=-(AX*32.2*12*DT^2/2):REM LONGITUDINAL DECELERATION
DISTANCE INCREMENT
784 DX=FWD+AFT:REM NET LONGITUDINAL DISTANCE INCREMENT
785 UP=((DX/X0)*H):REM VERTICAL COMPRESSION INCREMENT
786 IF I=1 THEN AZ=(UP*K)/(WEFF*2):REM INITIAL VERTICAL
ACCELERATION
787 DWN=-(AZ*32.2*12*DT^2/2):REM VERTICAL ACCELERATION DISTANCE
INCREMENT
788 DZ=UP+DWN:REM NET VERTICAL COMPRESSION INCREMENT
789 SZ=SZ+DZ:REM CUMULATIVE NET ELAPSED TIME VERTICAL
COMPRESSION
790 X=X+DX:REM LONGITUDINAL COORDINATE FROM X0
863 Z=Z-DWN:REM VERTICAL COORDINATE FROM GROUND PLANE
864 FX=(DX*(X0-X)/X1)*K*(H/RR):REM LONGITUDINAL FORCE
866 AX=FX/WEFF:REM LONGITUDINAL ACCELERATION
867 FZ=SZ*K:REM VERTICAL FORCE
868 AZ=FZ/N:REM VERTICAL ACCELERATION
873 V=V-(AX*32.2*DT*3600/5280):REM VELOCITY
880 I=I+1:REM PROGRAM ITERATION COUNT
889 T=T+DT:REM SIMULATION ELAPSED TIME COUNT
890 IF X<X0 THEN GOTO 781:REM CHECK ON PHASE I TERMINATION

Note that the key algorithms are the tying of the vertical compression increment to the net
longitudinal distance increment (line 785), the recognition that the change in vertical coordinate
is due only to the vertical acceleration (the tendency is for the wheel to compress as the step is

97
Rev B SAWE Paper No. 3521
Category Number 31.0
surmounted with the hub vertical location constant, line 863), and the tying of the longitudinal
compression to the longitudinal coordinate (line 864). The termination of Phase I (line 890)
results in a “drop down” into a program area that consists of a simplified version of the above; in
that simplified version there are no longitudinal compression, force, or acceleration
determinations, and the vertical force/acceleration “slowly” declines as the residual compression
depletes. The attainment of a new vertical coordinate at the hub of “Zo + H” (or very nearly so)
signals the end of Phase 2. As Phases 3 and 4 were not modeled, the conclusion of Phase 2
invokes the program run termination.

Results

Initial program “validation and verification” (not in any strict sense of the words) runs
were made using info for a vintage Dunlop RS4 6.00x16 tire as mounted on a 1958 Jaguar
XK150S. The “given” input, and consequent calculated input data, was:

The output data was as shown in the following excerpt (the complete output is listed in
Appendix II):

<I><TIME><VEL><X-COORD>< Z-COORD ><X-ACCEL><Z-ACCEL>


PHASE 1 - CONTACT, COMPRESSION, AND CLIMBING
1. 0.00 10.00 0.0000 13.42 0.0000 0.0000
2. 0.01 9.96 0.9989 13.43 0.3065 0.3342
3. 0.01 9.93 1.9931 13.43 0.2515 0.6646
4. 0.02 9.91 2.9846 13.43 0.1975 0.9893
5. 0.02 9.89 3.9740 13.44 0.1439 1.3088
6. 0.03 9.88 4.9618 13.45 0.0908 1.6234
7. 0.03 9.87 5.9489 13.46 0.0379 1.9333
PHASE 2 - VERTICAL EXPANSION, FORWARD VELOCITY
8. 0.04 9.87 6.9359 13.47 0.0000 2.2388
9. 0.05 9.87 7.9230 13.48 0.0000 2.1134
10. 0.05 9.87 8.9102 13.50 0.0000 2.0838

98
Rev B SAWE Paper No. 3521
Category Number 31.0
11. 0.06 9.87 9.8974 13.51 0.0000 2.0546
12. 0.06 9.87 10.8846 13.52 0.0000 2.0259
Etc.

The numeric output looks promising, but a surer sense of the program’s validity (or lack
of) results from a charting of the output parameters versus time:

Note that upon impact with the step the vehicle slows down a little bit, about 0.13 mph in
about 0.02 seconds. This does not seem unreasonable.

The loss in velocity is so small that it doesn’t even noticeably affect the plot of
longitudinal position versus time.

99
Rev B SAWE Paper No. 3521
Category Number 31.0

The plot of the change in the wheel hub’s vertical location with time seems to be smooth
and gentle curve; it goes from the rolling radius height of 13.4 inches to the rolling radius plus
step height (naturally) of 14.4 inches in a duration of a little over one and a half seconds.

As noted when considering the velocity versus time plot, there is some deceleration
incurred by impacting the road step. This deceleration peaks at about 0.31 g’s in approximately
0.01 seconds, and then drops to 0.00 g’s in a further 0.01 seconds, forming a typical “spike”
function. This would seem to indicate a loss in forward velocity of 0.54 mph, which is
significantly more than (4×) the 0.13 mph loss indicated by the previous “Velocity” plot (?).

100
Rev B SAWE Paper No. 3521
Category Number 31.0

The vertical acceleration versus time is the plot of most interest, as the vertical
acceleration constitutes the “road shock” so desirable to minimize. Note that the vertical
acceleration seems to maximize at about 2.24 g’s in approximately 0.04 seconds, and then take a
little more than 1.5 seconds to smoothly diminish to zero. This is behavior that would seem
expected of a modern automobile wheel/tire.

For comparison some rigid wheels were modeled with the following parametric values
which were input to the “RIGIDWHL.BAS” program:

The above data is for two sizes of rigid wheel, both with an estimated spring constant of
52,500 lbs/in104. One wheel has a radius of 9.8 inches, and the other wheel has a radius of 14.57

104
Hull, Wang, and Moore in their paper “An Empirical Method for Determining the Radial Force-Deflection
Behavior of Off-Road Bicycle Tires” noted a wheel (no tire) spring coefficient of 11,420.3 lb/in. In a crude attempt
to apply their methodology to a rigid carriage wheel this author got a value of 173,588.6 lb/in, but 52,500.0 lb/in
seemed to work better.

101
Rev B SAWE Paper No. 3521
Category Number 31.0
inches. The 14.57 inch radius rigid wheel is approximately the same size as the 14.095 inflated
radius pneumatic tired wheel investigated previously, thus allowing a meaningful comparison
between rigid and pneumatic. The choice of 14.57 inch radius, and the choice of 9.8 inch radius,
as the rigid wheel input sizes also allows for a program-validating comparison with empirically
obtained data published in Donald Bastow’s Car Suspension and Handling (page 4) 105.

Program

10 CLS:SCREEN 0:COLOR 7,1,4:BEEP:KEY OFF:S$=".DAT":F$=".OUT":S1$="


":PI=3.14159:CLS
20 PRINT"**********************************< FLEXIWHL.BAS
>******************************"
30 PRINT" A WHEEL VS. ROAD STEP PROGRAM"
40 PRINT""
50 PRINT" V.1 BRIAN WIEGAND 2009 IBM PC (GW-BASIC)"
60 PRINT" "
70 PRINT"NOTE: THIS PROGRAM ASSUMES OPERATION WITHIN THE RANGE OF
LINEAR / ELASTIC"
71 PRINT"BEHAVIOR; IT CAN'T IDENTIFY A SITUATION OF TIRE RUPTURE / WHEEL
FAILURE"
80 PRINT""
81 PRINT"************************************* 01.20.09
*********************************"
90 PRINT""
100 PRINT"INPUT: TIRE PRESSURE (P, psi), TIRE SPRING CONSTANT (K, lb/in), TIRE
DEFLECTION CONSTANT (D, in), INFLATED RADIUS (RI, in), NORMAL LOAD (N, lb),
VEHICLE VELOCITY (V, mph), VEHICLE EFFECTIVE WEIGHT (WEFF, lb), ROAD STEP
HEIGHT (H, in)."
130 PRINT""
140 PRINT"****************************(hit space bar to
continue)*************************"
150 K$=INKEY$:IF K$="" THEN GOTO 150
160 CLS
210 BEEP
219 PRINT L$
220 PRINT"IS INPUT FROM KEYBOARD (K) OR DATA FILE (D)?"
230 D$=INKEY$:IF D$="" THEN GOTO 230
240 PRINT D$
250 LN=1:IF D$<>"K" AND D$<>"D" AND D$<>"k" AND D$<>"d" THEN GOTO 1994
260 BEEP
270 PRINT"WHAT KIND OF HARDCOPY DO YOU WANT ?"
280 PRINT""
290 PRINT" 1 - NONE."

105
The data presented in Bastow’s book required translation from metric to English units. Also, the data required
conversion from the spatial domain to the time domain, which was a simple matter as the longitudinal velocity was
apparently kept constant (or very nearly so) by the test apparatus.

102
Rev B SAWE Paper No. 3521
Category Number 31.0
300 PRINT" 2 - OUTPUT ONLY."
310 PRINT" 3 - INPUT & OUTPUT."
320 H$=INKEY$:IF H$="" THEN GOTO 320
330 HC=VAL(H$)
340 LN=2:IF HC<>1 AND HC<>2 AND HC<>3 THEN GOTO 1994
360 IF HC>1 THEN PRINT L$:PRINT "WHAT IS FILENAME FOR OUTPUT DATA FILE
(FILE TYPE IS '.OUT')":LN=3:BEEP
361 IF HC>1 THEN INPUT N$:OPEN "O",#2,N$+F$:ON ERROR GOTO 1994
370 IF D$="D" OR D$="d" THEN GOTO 410
380 PRINT""
392 PRINT"ENTER THE INPUT VALUES: TIRE PRESSURE, SPRING CONSTANT,
DEFLECTION CONSTANT, INFLATED RADIUS, NORMAL LOAD, VELOCITY,
EFFECTIVE WEIGHT, STEP HEIGHT:"
400 GOTO 500
410 BEEP
420 PRINT L$:PRINT"WHAT IS THE NAME OF THE INPUT DATA FILE TO BE USED
(FILE TYPE IS '.DAT'):"
430 LN=4:INPUT M$
440 OPEN "I",#1,M$+S$
450 ON ERROR GOTO 1994
490 IF D$="D" OR D$="d" THEN GOTO 550
500 BEEP
510 INPUT P,K,D,RI,N,V,WEFF,H:ON ERROR GOTO 1994
520 GOTO 560
550 INPUT#1,P,K,D,RI,N,V,WEFF,H
560 IF D$="D" OR D$="d" THEN CLOSE #1
570 PRINT L$:PRINT"INPUT DATA:":PRINT L$
580PRINT"TIRE PRESSURE =",P,"SPRING CONSTANT =",K,"DEFLECTION
CONSTANT=",D,"INFLATED RADIUS =",RI,"NORMAL LOAD=",N,"VELOCITY
=",V,"EFFECTIVE WEIGHT=",WEFF,"STEP HEIGHT=",H:PRINT L$
750 IF HC=3 THEN PRINT#2,"<<<<INPUT DATA>>>>":PRINT#2,""
751 IF HC=3 THEN PRINT#2,"TIRE PRESSURE =",P,"SPRING CONSTANT
=",K,"DEFLECTION CONSTANT =",D,"INFLATED RADIUS =",RI,"NORMAL LOAD
=",N,"VELOCITY =",V,"EFFECTIVE WEIGHT=",WEFF,"STEP HEIGHT=",H:PRINT#2,L$
752 IF HC=2 OR HC=3 THEN PRINT#2,"<<<<OUTPUT DATA>>>>":PRINT#2,L$
753 IF HC=1 THEN PRINT"<<<<OUTPUT DATA>>>>":PRINT""
754 IF HC=1 THEN PRINT"<<N>><< TIME >><<VEL>><< X-COORD >><< Z-COORD
>><< X-ACCEL >><< Z-ACCEL >>"
755 IF HC>1 THEN PRINT#2," <<N>><< TIME >><<VEL>><< X-COORD >><< Z-
COORD >><< X-ACCEL >><< Z-ACCEL >>"
760 IF HC=2 OR HC=3 THEN PRINT"<<<<< FOR HARDCOPY OF OUTPUT DATA PRINT
APPROPRIATE '.OUT' FILE >>>>>"
761 RR=RI-((N/K)+D):Z0=RR:REM ROLLING RADIUS = HUB VERT HEIGHT AT
CONTACT (T = 0)
762 X1=SQR(RI^2-(RR-H)^2):REM LONGITUDINAL DISTANCE FROM HUB TO STEP
AT CONTACT (T = 0)

103
Rev B SAWE Paper No. 3521
Category Number 31.0
763 DT=(1/(V*5280*12/3600))*(2000/K)*(1/H):REM TIME INCREMENT SIZE BASED ON
VEHICLE VELOCITY
765I=1:T=0:X=0:Z=Z0:DX=0:DZ=0:SZ=0:FX=0:FZ=0:AX=0:AZ=0:FWD=0:AFT=0:UP=0:D
WN=0:REM INITIALIZATION OF VARIABLES COMPLETE
766 REM START OF PHASE I (COMPRESSION AND CLIMBING)
767 IF HC=1 THEN PRINT "PHASE 1 - CONTACT, COMPRESSION, AND CLIMBING"
768 IF HC=2 OR HC=3 THEN PRINT#2,"PHASE 1 - CONTACT, COMPRESSION, AND
CLIMBING"
770 IF HC=1 THEN GOSUB 1949
780 IF HC=2 OR HC=3 THEN GOSUB 1861
781 FWD=((V*5280*12/3600)*DT)
782 IF I=1 THEN AX=(FWD*K)/(WEFF*2)
783 AFT=-(AX*32.2*12*DT^2/2)
784 DX=FWD+AFT
785 UP=((DX/X1)*H)
786 IF I=1 THEN AZ=(UP*K)/(N*2)
787 DWN=-(AZ*32.2*12*DT^2/2)
788 DZ=UP+DWN
789 SZ=SZ+DZ
863 X=X+DX:Z=Z-DWN
864 FX=DX*((X1-X)/X1)*K*(H/RR)
865 IF X=X1 OR X>X1 THEN FX=0
866 AX=FX/WEFF
869 FZ=SZ*K
870 AZ=FZ/N
873 V=V-(AX*32.2*DT*3600/5280)
880 REM ACCELERATIONS WILL BE IN GRAVITY UNITS FOR OUTPUT
889 I=I+1:T=T+DT
890 IF X<X1 THEN GOTO 770
900 REM END OF PHASE 1, START OF PHASE 2 (VERTICAL EXPANSION, FORWARD
VELOCITY"
901 IF HC=1 THEN PRINT "PHASE 2 - VERTICAL EXPANSION, FORWARD VELOCITY"
902 IF HC=2 OR HC=3 THEN PRINT#2,"PHASE 2 - VERTICAL EXPANSION, FORWARD
VELOCITY"
910 IF HC=1 THEN GOSUB 1949
920 IF HC=2 OR HC=3 THEN GOSUB 1861
940 DX=((V*5280*12/3600)*DT)
950 DZ=(AZ*32.2*12*DT^2/2)
960 X=X+DX:Z=Z+DZ
961 FZ=((Z0+H)-Z)*K
970 AZ=(FZ/N)
975 I=I+1:T=T+DT
980 IF Z<(Z0+H)*.999 AND I<350*(.05/DT) THEN GOTO 910
981 GOTO 1999:REM THIS LINE SKIPS PHASE 3 AND PHASE 4
1000 REM END OF PHASE 2 (VERTICAL EXPANSION WITH FORWARD VELOCITY),
NOW PHASE 3 (OVERSHOOT)

104
Rev B SAWE Paper No. 3521
Category Number 31.0
1010 REM TBD
1040 REM END OF PHASE 3 (OVERSHOOT), NOW PHASE 4 (RETURN TO
EQUILIBREUM: Z = RR + H)
1050 REM TBD
1060 GOTO 1999
1861 REM HARDCOPY OF OUTPUT (HC=2 OR HC=3)
1862
PRINT#2,USING"&#####.&###.##&####.##&###.####&#######.##&######.####&#####.#
###";S1$;I;S1$;T;S1$;V;S1$;X;S1$;Z;S1$;AX;S1$;AZ
1940 RETURN
1949 REM SCREEN OUTPUT ONLY (HC=1)
1950 PRINT I,T,V,X,Z,AX,AZ
1993 RETURN
1994 REM ERROR ROUTINE
1995 PRINT"":ERCODE=ERR:ERLINE=ERL
1996 PRINT"THE ERROR CODE IS ",ERCODE," FOR ERROR AT LINE
",ERLINE,"PLEASE TRY AGAIN!"
1997 PRINT""
1998 ON LN GOTO 210,260,360,410
1999 CLOSE #2:KEY ON:END

105
Rev B SAWE Paper No. 3521
Category Number 31.0
APPENDIX C – PHYSIOLOGICAL SHOCK AND VIBRATION LIMITS

What the human body can endure, which can be surprisingly extreme, is very different
from what human’s consider comfortable with regard to shock and vibration tolerance. The
subject of comfort is a matter of a complex interplay between frequency, amplitude (which
relates to acceleration and power), directional orientation, and the exposure time. For a graphical
presentation of the discomfort zone in a frequency vs. amplitude plot the following “Figure 20”
borrowed from SAE J6a “Ride and Vibration Data Manual” sums up the situation:

106
Rev B SAWE Paper No. 3521
Category Number 31.0
With regard to orientation 106, one might feel that most, if not all, of the focus of this
paper has been on motions in the vertical direction, but with the discussion of pitch there has
been an implicit introduction to motions in the longitudinal direction (and, of course, there are
roll/lateral motions not considered herein). For instance, consider the effect of an intense
pitching motion about a node point in the following diagram:

The motion is exaggerated as drawn, but 1931 Cadillac V12 limos did on occasion
undergo motion like this to a lesser degree, and note that it is the patrician passenger, seated way
back over the rear axle (!), who is getting the worst of the ride. The chauffeur, who is seated near
the oscillation node point, is experiencing relatively moderate amplitude in his bouncing up and
down. His Lordship, however, is further away from the node and is experiencing a motion three
to four times that which his employee is enduring.

This situation illustrates the important role of node placement. It not just desirable to get
the frequency to be at a level within the range most conducive to human comfort, but to get the
node points of the most significant vibration near the driver position. This is what has
traditionally been done in practice, and explains why any tendency toward car sickness is usually
exhibited by the rear seat passengers.

Note also in the above illustration that the angular motion about the node translates into a
good deal of whipping back and forth at the head level. This is the longitudinal motion spoken
of, although a “pure” longitudinal vibration does arise through the suspension which is not a
byproduct of such rotary oscillation. This whipping motion is especially uncomfortable if it

106
The effect of orientation with respect to vibration is similar to the effect of orientation with respect to
acceleration tolerance.

107
Rev B SAWE Paper No. 3521
Category Number 31.0
occurs at frequencies in the 18 to 20 cps range 107, although any motion at sufficient amplitudes
can be bad. In any case, what happens at the head and neck is especially important and
appropriate care must be taken by the designer.

Whatever the cause of a longitudinal vibration, the human tolerance to longitudinal


vibration is summed up in a general way by a plot borrowed from Gillespie’s book Fundamentals
of Vehicle Dynamics 108:

Other than the general summation of vertical vibration tolerance in the SAE J6a chart,
and the general summation of longitudinal tolerance above, there are a few modes of vibration /
vibration effects that should receive special note, even when at very moderate amplitudes:

Raynaud’s Syndrome is associated with high frequency vibrations in the range of 20 to


200 cps. There are other causes for this circulatory disorder than vibration, but what is called
Secondary Raynaud’s Syndrome often comes about through use of power tools or machines that
generate the necessary stimulation. Severe enough exposure can damage the nerves and joints of

107
Reference [14], page 2.
108
Reference [7], page 184. “Lee” stands for R.A. Lee and F. Pradko, authors of “Analytical Analysis of Human
Vibration”, SAE Paper 680091, January 1968. “Parsons” stands for K.C. Parsons, et al, authors of “Six Axis
Vehicle Vibration and Its Effects on Comfort”, Ergonomics, Vol. 22, No. 2, 1979, pp. 211-225.

108
Rev B SAWE Paper No. 3521
Category Number 31.0
the hands and arms. Early vehicles often generated such vibration through the steering system,
and a long session of “white-knuckle” driving could result in a tingling feeling in the hands
accompanied by an otherwise general numbness. Modern vehicles generally do not transmit such
vibrations through the controls, except perhaps when damaged and/or in off-road use.

Motion sickness is associated with very low frequency vibrations in the 0.1 cps to 0.8
cps range 109. The exact location of where a passenger sits can have a significant effect (as His
Lordship can no doubt attest); generally the front seat of an automobile, the upper deck on a
boat, or wing seats on an aircraft constitute the best locations for minimizing the adverse effects.

Fatigue is associated with exposure to vibrations in the range of 4 to 8 cps. The


Technical Committee 108 of the International Organization for Standardization recommended
limiting RMS acceleration of 0.1 g at 4.0-to-8.0 cps to 1.5 hours, and increasing that limit to 4.0
hours at 1.0 cps. 110

Lastly it should be noted that the human perception of ride is not just tactile, but acoustic
as well. Human hearing is in the 20 to 20,000 cps range. However, even “sounds” that are
outside the range of human hearing, infrasound and ultrasound, can still have physiological
effects. An article in The New Scientist, Vol. 53-787, page 584, of 16 March 1972, reported that
considerable “noise” below 25 cps (infrasound) was generated in automobiles, especially with
the windows open. Visual recognition times were found to increase with sound levels in the 2 to
15 cps range at strengths greater than 105 dB, and visual tracking error increased in this range at
only 96 dB. To further complicate matters, prolonged exposure to such infrasound can generate
feelings of euphoria; all of which does not sound conducive to safe driving, but should sound
conducive to developing vehicle designs that not only bring shock and vibration to within
desirable levels, but sound as well.

109
Reference [14], page 2 (0.5 cps to 0.83 cps); Reference [2], page 22 (0.1 cps to 0.5 cps); Reference [1], page 5
(0.5 cps to 0.75 cps).
110
Reference [2], pages 22-23.

109
Rev B SAWE Paper No. 3521
Category Number 31.0

110

Вам также может понравиться