You are on page 1of 17

Original Article

Dynamic characterization of the


rectangular piston seal in a disk-caliper
braking system using analytical and
experimental methods

Proc IMechE Part D:


J Automobile Engineering
226(12) 16131629
IMechE 2012
Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/0954407012450118
pid.sagepub.com

Jared Liette, Jason Dreyer and Rajendra Singh

Abstract
A new simplified, yet representative experiment of a floating dual-piston disk-caliper braking system is designed to isolate
the rectangular seal and the pistonbore chamber from the complexities of a braking system. The physical sources of
the stiffness and damping mechanisms associated with the seal during an applied pressure event are identified and quantified under harmonic excitation. A tractable analytical model of the experiment that incorporates the identified dynamic
seal properties is proposed. This linear time-invariant model describes the governing equations of both the hydraulic
brake system components and the mechanical caliper components and provides some insights into a seemingly nonlinear
system. For a range of pressure amplitudes and brake configurations, excellent agreement between predictions and measurements is obtained for the peak-to-peak values of the pistonbore chamber pressure, the force transmitted by the
pistons, and the caliper displacement. The proposed model and experiment could be utilized in brake control, vibration,
and pedal feel studies.

Keywords
Rectangular piston seal, linear system analysis, parameter identification, time domain analysis

Date received: 22 December 2011; accepted: 27 April 2012

Introduction
A dynamic model of a floating or fixed (single- or dualpiston) disk-caliper braking system is critical in the
design of braking control systems, the development of
lighter and more fuel-efficient vehicle components, and
the reduction in brake-induced vibration by tuning the
dynamic properties of the brake system.1 The brake
pads, rotor disk, brake caliper, and piston(s) represent
a mechanical system, and the pistonbore chamber and
hydraulic lines represent a hydraulic system. Together
with the hydraulic pressure within the pistonbore
chamber, the rectangular seal around each piston is an
important interfacial element since it couples the
mechanical and hydraulic components of the system.1
Several computational models based on commercial
multiphysics codes have been utilized to investigate
braking issues including brake-induced vibration, pedal
feel, brake control, and brake drag.24 However, the
seal models in such studies24 lacked a clear mathematical representation. Another approach was to represent
the entire brake system by an analogous discrete

mechanical model.57 For instance, Kim et al.5 and


Leslie6 represented the hydraulic system by a linear
time-invariant stiffness which was then determined
from compression testing of the entire hydraulic system, but no physical sources were identified. Also, both
studies5,6 neglected significant damping in the brake
pads, the rectangular seal(s), and the pistonbore
chamber. Kang and Choi7 included an empirical system
level damping coefficient but did not specifically model
the hydraulic system. Again, the physical damping
sources were not clear as the hydraulic system and viscoelastic brake pad material damping seemed to be
lumped together.

Acoustics and Dynamics Laboratory, Department of Mechanical


Engineering, The Ohio State University, Columbus, OH, USA
Corresponding author:
Rajendra Singh, Acoustics and Dynamics Laboratory, Department of
Mechanical Engineering, The Ohio State University, Columbus, OH
43210, USA.
Email: singh.3@osu.edu

1614
Previous experimental or computational studies on
the rectangular seal component in disk-caliper braking
systems focused on the static properties of the seal, specifically its springback force or its ability to retract the
piston away from the rotor during an off-brake
event;3,8,9 however, the properties of the seal during a
dynamic braking event were not studied. The significance of the interfacial friction in elastic seals and linear actuators1013 has been studied using experimental
or stochastic methods,1012 and frictional forces have
also been empirically estimated on the basis of the seal
groove design and the actuation pressure.12,13 Such
studies focused on the measurement of friction but did
not directly calculate the effective damping or stiffness
parameters of the rectangular seal. Additionally, the
elastohydrodynamic lubrication issues of rectangular
seals have been investigated experimentally14 and
numerically,1416 with a focus on the reciprocating piston rods in high-pressure hydraulic actuators (often
found in aircrafts). Such analyses did not yield the
macroscopic component properties required by the
dynamic models.

Proc IMechE Part D: J Automobile Engineering 226(12)


designed to study the keh and ceh mechanisms. In this
experiment, a disk-caliper braking system should be
simplified, so as to isolate the pistonbore chamber
from the rest of the brake caliper structure (rotor and
pads). The system could be excited at a single excitation
frequency ve (rad/s) in order to measure the hysteresis
loop, which is critical to the parameter identification.
Using ramp or other excitations would be of interest,
particularly for nonlinear models; however, this paper
focuses on initial work in characterizing these seals for
use in linear time-invariant models. For brevity, examination of nonlinear effects has been omitted and is left
to future work.
The specific objectives are as follows:
1.

2.

Problem formulation
To the best of our knowledge, a comprehensive analytical and experimental study into the dynamic characterization of the rectangular seal in a disk-caliper braking
system has not yet been reported. The primary goal of
this article is to fill this void in the literature. An analogous linear time-invariant mechanical model (of dimension one) is used to represent the pistonbore chamber
for both fixed and floating (single- or dual-piston) caliper designs. A discrete mass element Mp represents the
piston(s), and effective discrete stiffness keh and viscous
damping ceh elements represent the hydraulic system.
Each is quantified as a linear time-invariant parameter.
Since no dynamic models are found in the literature for
piston seals, a linear time-invariant system model of an
experiment (discussed in the section on the design of a
new experiment) is proposed and used to estimate the
contributions of different physical mechanisms to keh
and ceh , as well as to assess the need for more complex
nonlinear models of piston seals in future studies. For
a dual-piston design, it is assumed that the two pistons
act in parallel, and thus the discrete elements add
together to form a one-dimensional representation. The
pressure in the pistonbore chamber is defined as pb(t)
where t is the time, and the effective piston surface area
to which pb(t) is applied is defined as Aep , while the displacement of the piston(s) is denoted by xp(t).
The source of the stiffness keh in the pistonbore
chamber is the bending of the rectangular seal(s) due to
the pb(t) excitation. Damping can be manifest in several
different forms, such as viscous, structural, and
Coulomb forms, all of which can be represented by
equivalent viscous damping coefficients ceh under harmonic excitation. A simple experiment must be

3.

4.

to understand and describe mathematically the


physical keh and ceh mechanisms of the rectangular
seal using linear time-invariant parameters, as well
as to identify and rank their contributions;
to propose a tractable linear time-invariant analytical model of a simplified, yet representative
experiment to gain a better understanding of the
underlying physics in the system, utilizing analytical relationships or approximations to calculate
the model parameters (which may be useful in the
design of such systems);
to predict the pressures, forces, and displacements
in the time domain and to compare the results with
experiments;
to identify possible sources of error and to suggest
some refinements to the proposed model.

Normalized values are analyzed throughout this


paper to emphasize the methods used, as opposed to
design-specific parameters.

Design of a new experiment


A floating dual-piston disk-caliper braking system is
first simplified by removing the rotor disk and cutting
the caliper in half to include only the pistonbore chamber side, better isolating its components. Furthermore,
the brake pad is replaced with a steel block to remove
the complex behavior of the brake pad material. The
hydraulic system is also reduced in complexity from
that contained on a vehicle, and only the piping for a
single disk-caliper braking system is retained; the air
booster and anti-lock brake system are also excluded. A
schematic diagram of the new experiment is shown in
Figure 1 with the individual components and sensors
labeled.
A large amount of heat is generated during a normal
braking event, which can affect the rotor surface profile, the caliper bore dimensions, the brake fluid viscosity, and the material properties of the rubber piston
seals. This initial study does not include the effects of
this heat generation and assumes that the brake components are at a constant (ambient) temperature. Future

Liette et al.

1615

Figure 1. Schematic diagram of the new experiment with focus on the rectangular seals: (a) front view; (b) side view.

studies should investigate the effects that elevated temperatures as well as thermal gradients have on the
system.
The hydraulic system is actuated by application of a
displacement to a brake master cylinder by a pneumatic
cylinder. The hydraulic pressure is increased by adjusting the air cylinder pressure, and the air supply
(6.89 MPa maximum) is pulsed using a three-way solenoid valve (100 ms response time, normally closed).
From the master cylinder, the hydraulic fluid flows
through a hard pipe, a needle valve, and an elastic hose
into the pistonbore chamber; both the hard pipe and
the elastic hose are similar to those found in a vehicle
brake system. Two pressure sensors (Honeywell model
MLH01KPSB06A; 06.89 MPa range; 1.7225 MPa/V
sensitivity17) are used in the hydraulic system: one at
the master cylinder and one in the pistonbore chamber. The first sensor serves as the measured input into
the hydraulic system, denoted as the master cylinder
pressure pi(t), and the second serves as the measured
input into the mechanical system, which is pb(t). A
baseline behavior of the hydraulic system is measured
when the needle valve is fully open. Varying the restriction of the needle valve changes the total resistance to
fluid flow and offers an alternative hydraulic system
configuration, allowing verification of the hydraulic
model and further investigation into the hydraulic system dynamics.
The mechanical components in the experiment in
Figure 1 include a cut caliper, two rectangular seals,
two pistons, a rigid steel block that the pistons push
down on to, three support columns on to which the cut
caliper is bolted, and washers which allow variation in
the leakage path length and the dead volume. The leakage path is defined as the thin fluid path between the
side of the piston and the bore; the dead volume is
defined as the volume of fluid in the pistonbore

chamber prior to excitation of the system. Three load


cells (Omega model LC307-5K; 0 22.24 kN range; 13.6
kN/(mV/V) sensitivity18) are used to measure the steel
block force Fk(t) exerted through the pistons, and a displacement transducer (Omega model LD500-2.5;
62.5 mm range; 12.8 mm/(mV/V) sensitivity18) measures the caliper displacement xc(t). These allow estimations of the relative seal displacement xs(t), which is
used to characterize the seal parameters. They also
allow an estimation of ceh via a hysteresis loop, which
requires both dynamic displacement and force measurements. The displacement xc(t) is defined from the center
point of the pistonbore chamber, equal to half the
measured displacement at the front edge. The displacement is measured at the front edge to avoid any distortion caused by bulging of the pistonbore chamber
directly above the pistons.
A National Instruments Compact DAQ data acquisition system is used. The pi(t), pb(t), and xc(t) time histories are recorded using an NI-9239 module,19 while
the Fk(t) and air solenoid time histories are recorded
using an NI-9237 module and an NI-9485 module19
respectively. All signals are then sent to a computer
using an NI-cDAQ-9172 chassis,19 where commercial
LabVIEW software19 is used for triggering and userinterface purposes. For this study, all signals are
processed with a third-order zero-phase Butterworth
low-pass filter with a cutoff frequency of 30 Hz for initial signal conditioning, creating a consistent frequency
range for the different sensors used. This filter is chosen
as it is near the upper bound of the bandwidth for
hydraulic components within a brake system (as suggested from common simulation step intervals for typical brake system physical effects1).
An initial study is carried out to examine the nature
of pb(t), as well as to characterize ceh and keh . This study
applies a 5 Hz waveform into the hydraulic system via

1616
pulsations of the air solenoid, with one washer at each
support column. A 5 Hz excitation is representative of
the response time of a brake booster with a master
cylinder,1 with other components having slower
response times.1 The pressure is ramped up to 1.38 MPa
(200 lbf/in2) before the solenoid is engaged, and the
resulting pb(t) signal has a mean pressure pm
b =
0.67 MPa and a peak-to-peak pressure p^b = 0.61 MPa,
as shown in Figure 2. All analyses conducted consider
the steady state portion of pb(t) only, and ideally the
pb
steady state response should be pb(t) = pm
b + 0:5^
sin(10pt). However, the resulting waveform resembles a
5 Hz sine wave with some distortion that is caused by
the physics of the actuation system. When taking the
fast Fourier transform of pb(t), 5 Hz is indeed the fundamental frequency, although some superharmonics
(10 Hz, 15 Hz, etc.) are present. The spectral energy
density at the fundamental frequency, however,
accounts for roughly 99% of the total spectral energy
density in the signal. Thus, pb(t) is essentially a 5 Hz sine
wave, as desired.
The fast Fourier transform of the corresponding
pi(t), Fk(t), and xc(t) time histories are also taken, and
the resulting spectral energy densities are shown in
Table 1. Here, five separate experimental runs are averaged together in the frequency domain, and only the
first five harmonics are considered (owing to the lowpass filter used for initial signal conditioning). All the
signals have about 99% of the energy spectral density
at the first harmonic; however, there is some amplification of the second and third harmonics from input to
output, particularly in xc(t). While this indicates some
nonlinear effects, the energy spectral density of each
harmonics is less than 1% combined and thus is
neglected in this study.
A harmonic excitation allows an estimation of the
total damping present in the system using a hysteresis
loop. With no brake pads present, the only significant

Proc IMechE Part D: J Automobile Engineering 226(12)


damping is due to the rectangular seals and piston
bore chamber. A hysteresis loop is constructed from
the measured dynamic Fdk (t) and xdc (t) time histories.
Both are passed through a third-order zero-phase
Butterworth bandpass filter, with cutoffs at 4 Hz and
6 Hz (for parameter estimation only). This yields signals at only the fundamental frequency, removing the
effect of the harmonic distortion on the hysteresis loop
shape and allowing estimation of an effective linear
time-invariant viscous damping parameter. Then, ceh is
calculated as
 d d
Fk dxc
DUh
e
ch =
1
 2 =
 2
pve Xdc
pve Xdc
from the energy dissipated DUh and the caliper displacement amplitude Xdc under harmonic excitation at
ve;20 the integral is numerically evaluated.
To characterize keh , it must first be determined
whether the seal slips on the piston surface. Thus, the
seal material properties and percentage squeeze (i.e. seal
strain es) are used to estimate whether the static friction
threshold force Ffs is overcome during a dynamic braking event. The seal is made from ethylene propylene
diene methylene with an assumed Youngs modulus Es
9.2 MPa (as reported in the literature21), and it is
assumed that the percentage squeeze is 15% (es 0.15)
based on the design guidelines for a dynamic seal 3 mm
thick.22,23 The force Ffs is defined as


2a
Ffs = mfs Ns = mfs Es es pdp ws
Es es = ss =

Ns
Ns
=
As
pdp ws

2b

where Ns is the normal force at the sealpiston interface, ss is the stress on the seal, As is the inner surface
area of the seal, ws is the thickness of the seal, dp is the
diameter of the piston, and mfs 0.30 is the static

Figure 2. Measured pb(t) for a 5 Hz waveform input with a 1.38 MPa maximum pressure.

Liette et al.

1617

Table 1. Percentages of energy spectral density at the first five harmonics of pi(t), pb(t), fk(t), and xc(t). The results are for a 5 Hz
waveform input with a 1.38 MPa maximum pressure.
Frequency (Hz)

5
10
15
20
25

Energy spectral density (%) of the following at the first five harmonics
Master cylinder
pressure pi(t)

Pistonbore chamber
pressure pb(t)

Steel block
force Fk(t)

Caliper
displacement xc(t)

99.0
0.17
0.38
0.07
0.40

99.6
0.10
0.16
0.08
0.01

99.4
0.11
0.40
0.02
0.02

99.1
0.41
0.47
0.01
0.02

Figure 3. Schematic diagram of a single piston with the seal modeled as a shear beam.

coefficient of friction assumed for a lubricated steel


elastomer interface. Experimental results from a compression test machine support the Ffs value calculated in
equation (2a). This compression test is performed with
the cut caliper and no hydraulic fluid (i.e. the caliper is
empty). With the pistons initially in the fully extended
position, the pistons are then slowly pushed in while
the prescribed xc(t) and resulting Fk(t) time signals are
measured.
In order to determine whether the seal slips, the static seal stiffness ks is also needed. An analytical approximation is obtained by assuming that the seal acts as a
shear beam with one edge constrained by the seal
groove and a force applied at the other edge by the
moving piston. The length of the shear beam is slightly
larger than the thickness of the leakage path, taking
into account the chamfer in the seal groove design; the
aspect ratio of the length to the width is roughly 1 to
30. A schematic diagram of a single piston is shown in
Figure 3 for clarity, including a close-up of the seal
modeled as a shear beam. Here, w is the leakage path
thickness, is the leakage path length, db is the bore

diameter, s is the seal length, Gs is the calculated seal


shear modulus, ns is the assumed Poissons ratio of the
seal, Fs is the force applied to the seal by the moving
piston, and my is the assumed dynamic viscosity of the
hydraulic fluid. The stiffness ks is defined as
Gs As
s
1
Es
pdp ws
=
s 2 1 + n s

ks =

by assuming an isotropic seal with ns 0.49, typical


for elastomers subjected to small strains. Experimental
results from the aforementioned compression test also
confirm the ks value calculated in equation (3).
The threshold displacement cs=Ffs /ks required
to overcome Ffs is now known. It is compared with
xs(t)=xp(t) xc(t)=Fk(t)/kk xc(t) to determine whether
the seal slips, where kk is the assumed boundary stiffness of the steel block. Note that the load cells are the
most compliant component between the piston and the
grounded steel block and thus dictate kk. The three

1618

Proc IMechE Part D: J Automobile Engineering 226(12)

Figure 4. Dynamic seal displacement versus dynamic seal resistive force.

stiffness elements act in parallel and each is found from


the manufacturers specifications. Experiments show
that the peak-to-peak seal displacement x^s = 2Xds . cs ,
where Xds is the seal displacement amplitude. Therefore,
the seal does indeed slip during the dynamic excitation
of the system. The seal stiffness is assumed to be negligible while slipping, and as a result it does not participate
throughout the entire period of oscillation. A plot of
the displacement xds (t) of the dynamic seal versus the
resistive force Fds (t) of the dynamic seal is shown in
Figure 4. Note that the typical asymmetry of the seal
groove designs is ignored, and the same static stiffness
is used regardless of the directionality of the piston
motion. For simplicity, it is assumed that the static coefficient mfs of friction and the kinetic coefficient of friction are the same.
When considering the entire oscillation period, the
effective seal stiffness kes is the peak-to-peak force over
the peak-to-peak displacement, linearized about the
current operating condition. This is shown in Figure 4,
and it is the linear time-invariant approximation of the
seal stiffness under harmonic excitation. This results in
a softer seal than a seal that does notslip, and the normalized effective seal stiffness kes = kes ks is found to be
about 0.29. Two seals act in parallel when considering
the effective hydraulic stiffness keh for the dual-piston
caliper design, which is given by
keh

= 2kes


Ffs
Ffs
=2
=
2Xds
Xds

Use of this effective seal stiffness parameter may not be


sufficient for a brake model where the input is
unknown, for a brake model with a ramp, or for a stochastic brake model. Future work involving nonlinear
models would probably be needed to handle such

inputs. Nevertheless, the proposed effective seal stiffness is useful for brake vibration models, where the
rotor surface input can be measured and approximated
as a combination of known harmonic inputs into the
system.

Examination of damping mechanisms


Physical damping mechanisms such as Coulomb friction at the pistonseal interface (when the seal is slipping), structural damping of the seal material, viscous
damping associated with the hydraulic fluid flowing
through the leakage path, and Coulomb friction at the
pistonbore interface must be identified and mathematically described. For comparison purposes, all damping mechanisms must be considered under the same
basis, and thus all are estimated with equivalent viscous
damping coefficients (as is the measured damping ceh ).
Depending on the physics of each mechanism, different
procedures may be needed. First, consider the viscous
damping coefficient cfs associated with the Coulomb
friction at the pistonseal interface. It is defined as24
DUs
 2
pve Xds
 d d
Fs dxs
=
 2
pve Xds


Ffs 2Xds  cs
=
 2
pve Xds

cfs =

from the plot shown in Figure 4 (similar to the ceh calculation). This is an estimation, as the theory is derived
on the assumption of a pure sinusoidal input, resulting
in an elliptical hysteresis loop.20 For a comparison with

Liette et al.

1619

the measured damping,


 a normalized viscous damping
coefficient 2
cfs = 2cfs ceh is used, taking into account the
dual pistons. Here, 2
cfs = 0.97, indicating that the measured damping is accounted for almost entirely by the
Coulomb friction between the seal and the piston.
The structural damping associated with the seal
material is estimated as a viscous damping coefficient
css by assuming a typical seal loss factor (hs 0.10)
given in the literature,25 and an assumed Es. Since the
seal stiffness does not participate in the entire oscillation period, neither does the damping coefficient. It is
thus scaled by the percentage of seal displacement corresponding to the seal sticking, equal to cs =2Xds . The
coefficient css is defined as
css =

cs k s hs
2Xds ve

6a

Substituting cs = Ffs /ks into equation (6a) gives


css =

Ffs hs
2Xds ve

6b


A normalized coefficient 2
css = 2css ceh = 0:11 is calculated. This is an order of magnitude less than 2
cfs and
thus is less significant.
The viscous damping coefficient cv associated with
the fluid flowing through the leakage path is defined as
#
"
#"
2
2
2
2
d
d

d
6pm

d

w

p
p
y
b

b

 w
cv =
4
4 2db  w
w3
7

which assumes laminar flow and utilizes capillary tube


theory;26 the relevant parameters are shown in Figure 3.
A normalized coefficient 2
cv = 2cv ceh = 0:0001 is calculated. This is negligible when compared with 2
cfs . It
might increase when the flow is turbulent, but it would
still be much smaller than the other mechanisms.
Slight scarring is found on the piston surface after
running several tests, indicating that there is contact
between the piston and the bore and that Coulomb
friction is probably present at this interface. When the
pressure is applied in the pistonbore chamber, the top
portion of the cut caliper body rotates slightly and initiates contact between the piston and the bore at two
locations, cocking the piston in the chamber. Consider
the side-view schematic diagram of the piston and caliper shown in Figure 5. The normal force Nb(t) is estimated as
Nb t = Aeb pb t sinuc t
2xc t
= Aeb pb t
c

Figure 5. Side view of a single piston with pistonbore contact.

motion of the center of the caliper, and the transducer


location is equal to twice that. The effective force area
opposite the piston(s), denoted as the effective piston
bore area Aeb , must account for the additional chamber
manifold or cover area. For instance, Aeb 1.15Aep in
this example case. The frictional force Ffb (t) per piston
bore is calculated as
Ffb t = 2mfb Nb t
= 4mfb Aeb pb t

from the measured xc(t) and pb(t), assuming that the


caliper rotates as a rigid body at an angle of uc(t) and is
fixed at the back support. The distance from the back
edge of the caliper to the displacement transducer location is denoted c . Recall that xc(t) is defined as the

8b

by assuming a static coefficient of friction mfb 0.20,


typical for lubricated steel on steel, and accounting for
2Nb(t) per pistonbore.
An equivalent viscous damping coefficient cfb is estimated from Ffb (t) by considering the relative velocity at
the interface, equal to x_ ds t: To estimate this from the
measured data, it is assumed that xds t = Xds sinve t,
and x_ ds t = ve Xds cosve t is found by differentiation.
Then, cfb is calculated as
 f 
F t
f

cb = db t
x_ t
=

8a

xc t
c

s
f e
4mb Ab
c ve Xds

hpb txc tit


hjcosve tjit

by taking the time-averaged



values of xc(t), pb(t), and
the rectified velocity x_ ds t (denoted by the h it operator); the rectified velocity is used since cfb is positive
regardless of the velocity direction. A normalized coefficient 2
cfb = 2cfb ceh = 0:04 is calculated. This is much
css .
lower than both 2
cfs and 2

1620

Proc IMechE Part D: J Automobile Engineering 226(12)

Table 2. Alternative damping sources and their equivalent viscous damping coefficients at 5 Hz.
Damping mechanism

Location

Normalized equivalent
viscous damping coefficient

Percentage of the total estimated


viscous damping (%)

Coulomb friction
Structural
Coulomb friction
Viscous

Pistonseal interface
Seal material
Pistonbore interface
Leakage path

2cfs = 0:97
2css = 0:11
2cfb = 0:04
2cv = 0:0001

86.9
9.6
3.5
0.0

All the normalized equivalent viscous damping coefficients for alternative damping mechanisms discussed
are summarized in Table 2. The individual mechanisms
are also ranked based on the percentages of the total
estimated viscous damping 2cfs + 2csc + 2cfb + 2cv that
they account for, and the Coulomb friction at the
pistonseal interface is the most dominant source. In
order to quantify whether the identified physical
damping mechanisms account for all the damping
measured from the system hysteresis loop, an index x
is defined as
x=

2cfs + 2css + 2cfb


ceh

10

where cv is excluded since it is negligible. For this initial


experimental study, x=1.12, indicating that slightly more
damping is calculated than is measured. However, a good
estimate is still obtained (within 12%). Possible sources
of error include the estimation of cfs using a nonsinusoidal
hysteresis loop and linear theory, the estimation of css
based on the material participation, the estimation of cfb
based on time averages, the estimation of ceh using a hysteresis loop (which ignores nonlinearities in principle),
and errors associated with measuring ceh .

Effect of the actuation pressure and


washers on parameters
Further experiments are necessary to validate the characterization of keh and ceh , examining a range of conditions, such as pb(t), that are likely to affect these
parameters. A higher p^b will increase Xds , causing the
seal to slide for a longer time in each oscillation period.
According to equation (4), keh will decrease as p^b
increases, since Xds is larger and Ffs remains unchanged.
The effect that pb(t) has on the hydraulic damping is
not clear, as there might be competing mechanisms.
For instance, a larger Xds will either increase or decrease
cfs based on equation (5), as the cfs versus Xds curve is
parabolic, and it will reduce css based on equation (6b).
A larger Xds will also reduce cfb , but a larger pm
b will
increase cfb , both based on equation (9). Consequently,
calculations will have to be made for several cases to
determine the net effect.
For the limiting case where the seal does not slip, the
seal parameters would be described as keh = 2ks, cfs = 0,
and css = kshs/ve. There would be no effect on cfb . To
study this case, a small peak-to-peak pressure would

have to be applied. The hardware configuration used in


this initial study is not capable of producing sufficiently
small peak-to-peak pressures to validate this limiting
case effectively.
Since pb(t) affects both keh and ceh , it is the primary
variable of interest in the experiments, which are conducted using maximum pressure levels of 1.38 MPa
(200 lbf/in2), 2.07 MPa (300 lbf/in2), and 2.76 MPa
(400 lbf/in2). It is found that a higher maximum pres^b pm
sure increases both p^b and pm
b while reducing p
b in
the experiment. This provides a variety of pb(t) signals. The amount of dead volume in the pistonbore
chamber is another pertinent factor. By increasing the
number of washers used at the support columns, the
dead volume is increased while reducing the length of
the leakage path. Two extreme cases are tested: one
washer as the initial setup, and five washers as the
limiting case. Five washers roughly double the dead
volume with one washer. The alternative hydraulic
system setup using the restricted needle valve is also
investigated in the experiments, all of which are carried out with the same bleed on the system within an
8 h period; the system is allowed to reach a steady state
before signals are acquired and passed through a lowpass filter (as discussed in the section on the design of a
new experiment). The full set of experiments is split into
two subsets for analysis purposes. Experiments A and B
involve tests with boundary conditions of one washer
and five washers respectively.
For each experiment, keh , ceh , cfs , css , cfb , and x are calculated via equations (4), (1), (5), (6b), (9), and (10)
respectively. The resulting dimensionless parameters
keh , heh , 2
cfs , 2
css , 2
cfb , and x are summarized in Table 3,
e
e
where kh = 0:5kh ks (accounting for two seals) and
heh is the effective hydraulic system loss factor defined
as
heh =

ceh ve
keh

11

A repeatability study is conducted to quantify the error


in the measurement of ceh , and a standard error of 61%
with a 95% confidence interval is calculated.
Trends show that keh decreases as the maximum
pressure (and thus the peak-to-peak pressure)
becomes larger within each subset, as expected from
equation (4). For a given pressure, it is also observed
that keh generally increases when the needle valve is
restricted, caused by a decrease in p^b , which in turn

Liette et al.

1621

Table 3. Summary of keh , heh , 2cfs , 2css , 2cfb , andx for all experiments where A denotes experiments with one washer and B denotes
experiments with five washers.
Experiment

Maximum
pressure (MPa)

Peak-to-peak
pressure (MPa)

ke
h

heh

2cfs

2css

2cfb

1A
2A
3A
4A (restricted needle valve)
5A (restricted needle valve)
6A (restricted needle valve)
1B
2B
3B
4B (restricted needle valve)
5B (restricted needle valve)
6B (restricted needle valve)

1.38
2.07
2.76
1.38
2.07
2.76
1.38
2.07
2.76
1.38
2.07
2.76

0.59
0.70
0.81
0.40
0.64
0.74
0.57
0.71
0.88
0.52
0.72
0.86

0.29
0.23
0.19
0.44
0.27
0.21
0.40
0.28
0.20
0.44
0.27
0.20

0.93
1.04
1.36
0.60
0.95
1.28
1.09
1.28
1.92
1.08
1.39
1.93

0.97
0.94
0.76
1.18
0.99
0.78
0.71
0.72
0.53
0.66
0.67
0.53

0.11
0.10
0.07
0.17
0.11
0.08
0.09
0.08
0.05
0.09
0.07
0.05

0.04
0.11
0.17
0.06
0.11
0.18
0.03
0.08
0.12
0.03
0.08
0.11

1.12
1.14
1.00
1.41
1.21
1.04
0.83
0.88
0.70
0.79
0.81
0.69

reduces Xds . Furthermore, keh also increases when


going from a single-washer to a five-washer boundary
condition, probably related to a reduction in Xds . The
loss factor heh follows a trend opposite to that of keh
for a given washer amount, as suggested by equation
(11). Little change is observed in ceh as the maximum
pressure increases, limiting its effect on heh for a given
washer amount. However, a larger ceh is calculated for
the five-washer cases when compared with the onewasher cases, inducing an increase in heh for the
five-washer cases. This change is not reflected in the
damping mechanisms, causing x \ 1 for all cases in
experiment B, which indicates an underestimation
cfs remains domiof ceh in calculations. The coefficient 2
f
nant for all cases; however, 2
cb becomes more significant than 2
css as the maximum pressure increases. One
final observation noted is that, for a given hydraulic
system setup, x tends to decrease as the maximum
pressure becomes higher. When considering all the
experiments, 0.69 4 x 4 1.41, indicating that the
damping mechanisms estimate ceh to within 641%.
This does not seem to have a significant effect on the
peak-to-peak predictive capability of the proposed
analytical model (described in the next section).

Development of an analytical model


A linear time-invariant analytical model of the experiment is developed next using the schematic diagram in
Figure 6. The mechanical system is represented by discrete elements, and the hydraulic system is represented
by the main components and their pertinent properties.
The analogous mechanical damping and stiffness elements associated with the hydraulic system are represented by the ceh and keh parameters (derived in the
section on the design of a new experiment); additional
stiffness elements include the stiffness kk of the steel
block, the stiffness kc of the cut caliper, and the stiffness
ki of the master cylinder. The discrete mass elements
include the mass Mp of the piston, the mass Mc of the
cut caliper, and the mass Mi of the master cylinder. The

master cylinder chamber, hard pipe, needle valve, elastic hose, and pistonbore chamber are the hydraulic
components included (denoted by the subscripts i, r, v,
t, and b respectively); the component properties include
the area, length, diameter, wall thickness, and Youngs
modulus (denoted by A, , d, w, and E respectively).
The time-varying state variables include pi(t), pb(t),
xp(t), xc(t), the master cylinder displacement xi(t), and
the hydraulic system flow rate Qh(t). The excitation
force for the system is the air cylinder pressure pa(t)
multiplied by the air cylinder area Aa, as applied in the
experiment.
All the discrete masses are determined by weighing
the individual components. For the discrete stiffness
elements, kk is determined from the three load cells in
parallel (as previously discussed in the section on the
design of a new experiment) and ki is determined from
the provided master cylinder specifications for its
return spring. The stiffness kc is determined from a step
response of the experiment (with one washer) at three
pressure amplitudes (1.38 MPa, 2.07 MPa, and
2.76 MPa), denoted as Pb. Since the seal slips during
the excitation, it offers little resistance to the caliper
motion. It is thus assumed that the stiffness of the caliper is the dominant resistance to the movement of the
caliper body, caused by the pressure excitation on the
top surface of the pistonbore chamber. The stiffness is
the slope of the applied force Aeb Pb versus the caliper
displacement amplitude Xc for the three Pb values
considered.
A constant mean flow rate is assumed in the hydraulic system, neglecting the effects of the hard pipe and
elastic hose compliances on the flow rate. This assumption simplifies the analysis and avoids some numerical
issues. As a result, the hard pipe and elastic hose have
only resistance and inertance terms. A linear timeinvariant fluid resistance R is calculated for both,
assuming laminar flow in the experiment. The Reynolds
number is calculated at a maximum pressure of
2.76 MPa, which produces the highest Qh(t) of all the
cases considered in the experiments. The Reynolds

1622

Proc IMechE Part D: J Automobile Engineering 226(12)

Figure 6. Analytical model schematic diagrams: (a) fluid-mechanical model; (b) mechanical model.

number is defined for the hard pipe (denoted Rer) and


for the elastic hose (denoted Ret) as26
jQh tjry dr
Rer t =
my Ar

It =
12a

and
Ret t =

jQh tjry dt
my At

12b

respectively, where ry is the hydraulic fluid density.


Both are much lower than 2000; thus the flow can be
considered laminar.26 The fluid resistance is calculated
for the hard pipe (denoted Rr) and for the elastic hose
(denoted Rt) as26
Rr =

128my r
pd4r

13a

128my t
pd4t

13b

and
Rt =

respectively. The fluid inertance is also calculated as a


linear time-invariant parameter, but the end effects
must be considered for pipe flow. An empirical factor
of 43 is assumed for laminar flow, and the inertance is
calculated for the hard pipe (denoted Ir) and for the
elastic hose (denoted It) as26
4 r y r
Ir =
3 Ar

and

14a

4 ry t
3 At

14b

respectively. The compliance and inertance associated


with the needle valve are assumed to be negligible. The
resistance Rv, however, is significant both when the
valve is fully open and when the valve is restricted.
Specifications for the valve provide a volume coefficient BV defined as
BV = Qh

r
gy
Dpo

15

where g y is the hydraulic fluid specific gravity and Dpo


is the operating pressure difference. By linearizing the
first derivative of Qh about Dpo,26 Rv is derived according to


r
gy
dQh
1
BV
=
=
16a
Rv
dDp Dp = Dpo
2
Dpo
s
2
Dpo
Rv =
16b
BV
gy
The value of BV calculated in equation (15) is used
when the valve is fully open. When it is restricted, BV is
approximated by


Ddv 2
BV } 1 
dv

17

Liette et al.

1623

based on the flow rate relationships,26 where dv is the


current needle valve diameter and Ddv is the deviation
from the geometric (or fully open) needle valve
diameter.
The effective hydraulic system bulk modulus beh is
calculated as26
!1
1  ja
ja
dr
dv
dt
e
bh =
+
+
+
+
by
ba
wr Er
wv Ev
wt Et
18

It takes into account the bulk modulus by of the


hydraulic fluid, the bulk modulus ba of the entrapped
air, the volume fraction ja of entrapped air in the
hydraulic fluid, and the component properties of the
hard pipe, needle valve, and elastic hose. It is assumed
that ja 0.001 for all experiments and that ba = pm
i ,
is
the
mean
value
of
the
master
cylinder
where pm
i
pressure.26
The master cylinder chamber and pistonbore chamber both have compliances associated with the contained fluid, denoted Ci and Cb respectively. These are
calculated as26
Vi
beh
Vb
Cb = e
bh

19a

Ci =

19b

where Vi is the volume of the master cylinder chamber


and Vb is the volume of the pistonbore chamber. Both
chamber volumes, however, depend on the number a
of washers present at the caliper boundaries and are
defined as
Vi = Ai i  Aep a  1ww

20a

Vb = Aeb b

20b

+ Aep a

 1ww

The master cylinder chamber length i and pistonbore


chamber length b correspond to the one-washer
boundary condition, and additional washers increase
the total length of the pistonbore chamber while
decreasing the total length of the master cylinder chamber by each washer thickness ww. It is assumed that the
initial volume of the pistonbore chamber relates to
Aeb , and that the additional volume relates to Aep as the
pistons expand out of the chamber.
For the hydraulic system in Figure 6(a), the governing equations are derived by considering the continuity
equation as well as conservation of momentum, and
for the mechanical system in Figure 6(b), the equations
of motion are derived by using Newtons second law.
The resulting equations are
Mi xi t + ki xi t = Aa pa t  Ai pi t
Ci p_ i t =  Qh t + Ai x_ i t

21a

21b
pi t  pb t = Rr + Rv + Rt Qh t + Ir + It Q_ h t
21c



Qh t  Aep x_ p t  x_ c t = Cb p_ b t


Mp xp t + ceh x_ p t + keh + kk xp t
ceh x_ c t  keh xc t = Aep pb t


Mc xc t + ceh x_ c t + keh + kc xc t  ceh x_ p t
keh xp t =  Aeb pb t

21d

21e

21f

These equations are numerically solved using commercial MATLAB Simulink software27 (using the ode15s
solver with the 0.001 s maximum step size). There are
six differential equations, corresponding to six state
variables and one input pi(t). The measured pi(t) signal
is directly fed into the simulation (with the corresponding t vector) as an input, allowing for a comparison of
the measured and predicted responses. The six state
variables are pa(t), xi(t), Qh(t), pb(t), xp(t), and xc(t);
equations (21a) to (21f) are rearranged to calculate
these with pi(t) as the input.

Results and discussion


Results for experiments A (with one washer)
All measured signals are acquired for the same time
duration and taken only under steady state conditions.
Additionally, all measured signals are sent through a
low-pass filter (as discussed in the section on the design
of a new experiment), including the pi(t) signal used as
the input into the Simulink model; no signal processing
is done on the model outputs. Only a small segment of
the response is examined to observe the dynamic
response better, and the main goal of the model is to
predict the peak-to-peak values of the measured pb(t),
Fk(t), and xc(t) time signals accurately with minimal
time offset between the predicted and the measured signals. The time offset t is quantified at the mean value
of the measured signal. Since the measured and predicted signals have slightly different shapes, the time
offset during the falling portion t1 of the response is
different from the rising portion t2 of the response. An
average of the two, t = 0.5(t 1 + t 2), is thus taken.
Typical measured and predicted xc t time histories
for experiment
are shown in Figure 7, where
 1A
is the caliper displacement normalxc t = xc t xm
c
ized by its absolute mean value. The time offset components are labeled accordingly with a sample t of about
8.0 ms; t . 0 indicates that the predicted signal passes
the mean value before the measured signal. The deviation in the time offset suggests a deviation in the estimation of ceh based on various assumptions made for
the linear time-invariant model formulation. A more
complex nonlinear model, such as a model that utilizes
friction elements, may provide an improvement in this
time offset. However, as will be shown later in this section, the current linear model provides good agreement
with the peak-to-peak estimation of the measured data.
For experiments 4A to 6A, Rv is scaled accordingly
on the basis of equations (16b) and (17). The diameters

1624

Proc IMechE Part D: J Automobile Engineering 226(12)


pressure increases, x decreases and moves from overestimating to correctly estimating ceh (i.e. x ! 1+ ). This
should counteract the trend seen in the time offsets.
Finally, it should be noted that ceh has little effect on
tb/T.

Results for experiments B (with five washers)

Figure 7. Time offset between the measured and the


predicted xc t values with one washer for experiment 1A: ,
measured; - - - -, predicted.

dv are not exactly the same for experiments 4A to 6A


since the needle valve is adjusted by hand. As a result,
dv is tuned differently for each individual case, although
it tends to be about half the geometric dv value. The
predicted-to-measured peak-to-peak amplitude ratios
and the t/T ratios (all in per cent) for pb(t), Fk(t), and
xc(t) are listed in Table 4 for experiments 1A to 6A.
Here, T=200 ms is the oscillation period, and the sign
of t/T reflects whether or not t . 0. Sample graphical
results of normalized time signals pb t, Fk t, and xc t
for experiments 2A and 5A are shown in Figure 8,
where each is normalized by the corresponding absolute
mean value.
An excellent agreement between the predicted and
the measured results is obtained for the peak-to-peak
values in all three experiments, within 612.4%. The
agreement between the t/T ratios is also good, as all
the cases are under 4%. A trend is present when looking at tk/T and tc/T versus the maximum pressure; as
the pressure increases, t k/T and tc/T decrease. They
usually move from positive to negative, which might
suggest a move from underestimating to overestimating
the effective damping. Better results may be obtained if
the damping coefficients associated with the identified
damping mechanisms were to be used in the model
instead of ceh . As evident from Table 3, when the

The resistance Rv is scaled accordingly for experiments


4B to 6B. The predicted-to-measured peak-to-peak
ratios and the t/T ratios (all in per cent) for pb(t), Fk(t),
and xc(t) are listed in Table 5 for experiments 1B to
6B. Sample results of normalized time signals
pb t, Fk t, and xc t for experiments 2B and 5B are
shown in Figure 9. An excellent agreement between the
predicted and the measured results is obtained for p^b in
all six experiments, within 67.5%. A good agreement
is also found for F^k , within 16.4%, where all the predictions are under-predictions. However, poor results are
obtained for x^c , over-predicting the measured values.
When the aforementioned step response test is repeated
for the five-washer boundary condition, the same kc
parameter is calculated. However, the force versus displacement curve shifts to the left.
The mechanism that causes this shift in the force versus deflection curve is unclear. One possible explanation
could be that a mechanical advantage exists owing to
an eccentric loading of the pistons, caused by a reaction
force at the steel block. Using the three load cells in the
steel block, the center of force location is calculated
with respect to the center of the pistons, and the two do
not coincide. This eccentric loading could cause a reaction at the support columns, altering the total force Fc
acting on the caliper body. Figure 10 displays a simplified side view schematic diagram of the piston and cut
caliper, where boundary conditions are assumed to
make the problem statically determinate for ease of
explanation. Here, Fc includes the applied pressure
force and the reactions at the two front support columns (all assumed to act in the same plane); Fp is the
pressure force applied to the pistons. The moment arms
for the piston, steel block, pistonbore chamber, and
caliper forces are denoted bp, bk, bb, and bc respectively,
and the lengths p and b are for the one-washer boundary condition. The reaction force at the back support

Table 4. Predicted-to-measured peak-to-peak amplitude ratios and time offsets for experiments A (with one washer).
Experiment

Predicted-tomeasured
^pb ratio (%)

Predicted-tomeasured
^Fk ratio (%)

Predicted-tomeasured
^xc ratio (%)

t b/T (%)

t k/T (%)

t c/T (%)

1A
2A
3A
4A (restricted needle valve)
5A (restricted needle valve)
6A (restricted needle valve)

89.1
96.3
97.2
102.7
100.5
103.4

89.4
92.8
90.8
100.2
96.0
95.2

93.6
102.6
97.2
112.4
109.3
102.7

1.2
0.7
21.0
20.7
0.5
1.3

0.8
22.0
23.5
1.5
1.0
21.0

4.0
0.2
21.2
3.7
2.0
1.5

Liette et al.

1625

Figure 8. Measured versus predicted pb t, Fk t, and xc t values with one washer for (a) experiment 2A and (b) experiment 5A:
, measured; - - - -, predicted.

column is denoted F1, and the seal forces are neglected


from the analysis since they are an order of magnitude
less than the applied pressure forces.
The relationship between Nb and Fp is found to be
Nb =

bk
Fp
bp  p  a  1ww

22a

by summing the moments about point O and considering the force balance in the xp direction (Fp = Fk).

Similarly, the relationship between Nb and Fc is found


to be
Nb =

0:5db + bc
Fc
bb  b  a  1ww

22b

by summing the moments about point O and considering the force balance in the xc direction (Fc=F1).
Substituting equation (22a) into equation (22b) and
rearranging gives

1626

Proc IMechE Part D: J Automobile Engineering 226(12)

Table 5. Predicted-to-measured peak-to-peak amplitude ratios and time offsets for experiments B (with five washers).
Experiment

Predicted-tomeasured
^pb ratio (%)

Predicted-tomeasured
^Fk ratio (%)

Predicted-tomeasured
^xc ratio (%)

t b/T (%)

t k/T (%)

t c/T (%)

1B
2B
3B
4B (restricted needle valve)
5B (restricted needle valve)
6B (restricted needle valve)

101.2
102.8
98.8
100.0
94.7
92.5

88.4
91.9
89.7
83.6
84.1
85.1

138.6
130.4
113.2
136.8
119.8
107.1

2 2.8
2 3.5
2 4.5
2 1.5
2 2.8
2 3.3

2 5.8
2 7.5
2 9.0
2 4.2
2 6.2
2 7.5

2 2.7
2 4.5
2 6.2
2 1.3
2 3.5
2 4.2

Figure 9. Measured versus predicted pb t, Fk t, and xc t values with five washers for (a) experiment 2B and (b) experiment 5B:
, measured; - - - -, predicted.

Liette et al.

1627

Figure 10. Simplified static analysis of the piston and cut caliper (side view).

Fc
bk
bb  b  a  1ww
=
Fp
0:5db + bc bp  p  a  1ww

22c

Only a (the number of washers) changes in equation


(22c) when the washer amount is altered. Using a rough
estimate for the dimensions, it is found that Fc/Fp
decreases by 4% when going from a=1 to a=5, thus
reducing the effective force acting on the cut caliper for
the same pressure input, which in turn should reduce
the displacement of the caliper. Note that this simplified static analysis shows the correct trend but does not
predict a large reduction in Fc/Fp. Thus, a better understanding of the underlying physics would require an
investigation in the future.
In general, t b/T, tk/T, and t c/T are worse for experiments B than for experiments A. However, all the
ratios still fall under 9%. Similar to experiments A, the
t/T ratios in experiments B decrease as the pressure
increases, but t \ 0 for all cases. This might suggest an
overestimation of the effective damping. Recall that
x \ 1 for all cases in experiments B, underestimating
ceh . Again it appears that the damping coefficients associated with the identified damping mechanisms may
produce better results if they were to be used in the
model instead of ceh .

Conclusion
Two key contributions emerge from this study. First,
an experiment to characterize dynamically the rectangular seal component found in a floating or fixed (single- or dual-piston) disk-caliper braking system is
conducted, and physical sources of stiffness and damping associated with the seal (and pistonbore chamber)
are both identified and quantified. The stiffness source
is attributed to the static stiffness of the rectangular
seal (modeled as a shear beam) and the duration of the
response that the seal is slipping on the piston surface.

This is defined by an effective linear time-invariant discrete stiffness element. Significant damping mechanisms are identified, all of which are estimated by linear
time-invariant damping coefficients. These mechanisms
include Coulomb friction at the pistonseal interface,
structural damping in the seal material, and Coulomb
friction at the pistonbore interface; the friction at the
pistonseal interface is the most significant source.
These damping mechanisms are found to account for
the measured effective viscous damping coefficient to
within 641%, but this deviation does not seem to have
a significant effect on the peak-to-peak predictive capabilities of the analytical model.
A second contribution is the development of a tractable analytical model of the experiment (using linear
time-invariant parameters), with which a good agreement between the predicted and the measured peak-topeak values is obtained with minimal time offsets
between the two. This model incorporates the identified
dynamic rectangular seal parameters, giving confidence
in the proposed linear time-invariant representation of
the seal. The model can be extended for use in investigating a range of brake system issues, including brakeinduced vibration, pedal feel, brake control, and brake
drag. Finally, it provides a linearized insight into a seemingly nonlinear system.
A better understanding of the physics behind the
possible mechanical advantage phenomenon observed
at the five-washer boundary condition is left to future
work. An extension of the proposed linear model to
include nonlinear effects such as Coulomb friction
instead of equivalent viscous damping coefficients
should also be proposed in the future. A multivalued
seal stiffness could also be incorporated into the model,
with the static stiffness being used when the seal sticks
and a near-zero stiffness being used when the seal slips;
the asymmetry of the seal groove design should be considered for this type of model. The effects of elevated

1628
temperatures and temperature gradients on the system
are omitted in this paper, and future work should
investigate these thoroughly. Finally, experimental and
analytical methods should be extended to a wider frequency range (say 525 Hz, common for brake judder1)
and to more complex wave forms associated with transient brake apply or control behaviors.

Proc IMechE Part D: J Automobile Engineering 226(12)

12.

13.
14.

Funding
This work was supported by Honda R&D Americas,
Inc. (grant nos 60020014 and 60030553) and the Honda
Partnership Program and Transportation Research
Endorsement Program at The Ohio State University
(fund no. 291742).
Acknowledgements
The authors would like to thank Honda R&D
Americas, Inc. and the Honda Partnership Program
and Transportation Research Endorsement Program at
The Ohio State University for guidance that allowed
this research to be possible.
References
1. Breuer B and Bill KH. Brake technology handbook. Warrendale, PA: SAE International, 2008.
2. Alirand M, Lebrun M, and Richards CW. Front wheel
vibrations: a hydraulic point of view models and first
results. SAE paper 2001-01-0490, 2001.
3. Day A, Ho HP, Hussain K and Johnstone A. Brake system simulation to predict brake pedal feel in a passenger
car. SAE paper 2009-01-3043, 2009.
4. Liette J. A coupled, multi-physics model of the automotive
brake system with focus on dynamic torque prediction. BS
Thesis, The Ohio State University, Columbus, OH,
USA, 2009.
5. Kim SH, Han EJ, Kang SW and Cho SS. Investigation
of influential factors of a brake corner system to reduce
brake torque variation. J Automot Technol 2008; 9(2):
233247.
6. Leslie AC. Mathematical model of brake caliper to determine brake torque variation associated with disc thickness variation (DTV) input. SAE paper 2004-01-2777,
2004.
7. Kang J and Choi S. Brake dynamometer model predicting brake torque variation due to disc thickness variation. Proc IMechE Part D: J Automobile Engineering
2007; 221(1): 4955.
8. Cai H and Anwana O. Seal/groove performance analysis
models. SAE paper 2002-01-2588, 2002.
9. Backstrom A. An experimental investigation of brake
rotor DTV under laboratory conditions Part 3. SAE
paper 2010-01-1690, 2010.
10. Schroeder L and Singh R. Experimental study of friction
in a pneumatic actuator at constant velocity. Trans
ASME, J Dynamic Systems, Measmt Control 1993; 115:
575577.
11. Kwak BJ, Yagle AE and Levitt JA. Nonlinear system
identification of hydraulic actuator friction dynamics
using a Hammerstein model. In: 1998 IEEE international

15.

16.

17.

18.

19.

20.
21.

22.
23.

24.
25.
26.
27.

conference on acoustics, speech and signal processing, Seattle, WA, USA, 1215 May 1998, vol. 11, pp. 19331936.
New York: IEEE.
Belforte G, Manuello A and Mazza L. Optimization of
the cross section of an elastomeric seal for pneumatic
cylinders. J Tribol 2006; 128: 406413.
Martini LJ. Practical seal design. New York: Marcel Dekker, 1984.
Prati E and Strozzi A. A study of the elastohydrodynamic
problem in rectangular elastomeric seals. J Tribol 1984;
105: 505512.
Nikas GK and Sayles R S. Nonlinear elasticity of rectangular elastomeric seals and its effect on elastohydrodynamic numerical analysis. Tribol Int 2004; 37: 651660.
Ruskell LE. A rapidly converging theoretical solution of
the elastohydrodynamic problem for rectangular rubber
seals. Proc IMechE, J Mechanical Engineering Science
1980; 22(1): 916.
Honeywell. Sensing and Control, MLH01KPSB06A,
Honeywell International Inc., Morristown, New Jersey,
USA,
http://sensing.honeywell.com/product%20page?
pr_id=31500 (2011, accessed 22 July 2011).
Omega. Sensors, thermocouple, PLC, operator interface,
data acquisition, RTD. Omega Engineering Inc., Stamford, CT, USA, http://www.omega.com/ (2011, accessed
22 July 2011).
National Instruments. Test, measurement, and embedded
systems. National Instruments Corporation, Austin, TX,
USA, http://www.ni.com/ (2011, accessed 22 July 2011).
Meirovitch L. Fundamentals of vibrations. New York:
McGraw-Hill, 2001.
ASi, AcuSeal Inc. Material Specs, AcuSeal Inc., Clinton
Township, MI, USA, http://www.acuseal.com/material_
specifications.php (2003, accessed 12 April 2011).
Apple Rubber Products Inc. Seal design guide. Lancaster,
New York: Apple Rubber Products, 2000.
Oberg E, Jones FD, Horton HL and Ryffel HH. Machinerys handbook, 26th edition. New York: Industrial Press,
2000.
Gaillard CL and Singh R. Dynamic analysis of automotive clutch dampers. Appl Acoust 2000; 60: 399424.
Jones DIG. Handbook of viscoelastic vibration damping.
New York: John Wiley, 2001.
Doebelin O. System dynamics modeling and response.
New York: Marcel Dekker, 2006.
MathWorks. Simulink, simulation and model-based
design, MathWorks, Natick, MA, USA, http://www.
mathworks.co.uk/products/simulink/ (2011, accessed 9
June 2011).

Appendix
Notation
A
b
BV
c
C
d
E
F
G

area
moment arm
volume coefficient
viscous damping coefficient
compliance
diameter
Youngs modulus
force
shear modulus

Liette et al.

1629

I
k

M
N
p(t)
P
Q(t)
R
Re
t
T
U
V
w
x(t)
X

inertance
stiffness
length
mass
normal force
pressure
pressure amplitude
flow rate
resistance
Reynolds number
time
oscillation period
energy
volume
thickness
displacement
displacement amplitude

a
b
g
D
e
h
u(t)
m
mf
n
j

number of washers
bulk modulus
specific gravity
change in a quantity
strain
loss factor
angle
dynamic viscosity
static coefficient of friction
Poissons ratio
volume fraction of entrapped air to
hydraulic fluid
density
stress
time offset
viscous damping index
threshold displacement
circular frequency

r
s
t
x
c
v

Subscripts
a
b
c
d
e
h
i
k

o
p
r
s
t
v
w
y

air
pistonbore chamber
caliper
brake pads
excitation
hydraulic system
master cylinder
steel block
leakage path
operating point
piston
hard pipe
seal
elastic hose
needle valve
washer
hydraulic fluid

Superscripts
d
e
f
m
s
v
_

dynamic
effective
friction
mean
structural
viscous
d( )/dt
d2( )/dt2
normalized value
peak-to-peak value

Operator
hit

time average

You might also like