Gearbox Vibration Control Dissertation
Gearbox Vibration Control Dissertation
Mingfeng Li
I, _________________________________________________________,
hereby submit this work as part of the requirements for the degree of:
Doctorate of Philosophy
in:
Mechanical Engineering
It is entitled:
Active Vibration Control of A Gearbox System
With Emphasis On Gear Whine Reduction
DOCTOR OF PHILOSOPHY
2005
by
Mingfeng Li
The gear whine problem that widely occurs in power transmission systems is typically
characterized by one or more high amplitude tonal acoustic signals. The whine response
originates from the vibration of the gear pair system caused by transmission error excitation that
arises from tooth profile errors, misalignment and tooth deflections. In very severe cases, the
gear vibrations can also reduce life and performance of the power transmission systems. The
fundamental gear whine problem has been of interest to many researchers for a very long time.
However, most of the work have focused on shaping the gear tooth profiles, adding vibration
isolation to the gearbox, and other passive control methods. Effort involving active gearbox
vibration control is still quite limited. In fact, other than those papers on the active control of the
vibration transmitted through support struts of gearboxes, only a few isolated studies that directly
deal with control of gearbox internal components have been seen. Therefore, addressing this
In this research, a dynamic finite element model is first developed for a power re-
circulating gear train system for use to examine the feasibility of applying an active shaft
transverse vibration control system near the gear pair in the effort to tackle the gear excitation
problem more directly. For this particular actuator setup and gearbox system, a complete set of
controller which is based on the FXLMS algorithm combined with improved frequency
estimation technique is designed. A series of computational studies are performed to refine the
algorithm design and to determine the most suitable parameters. Finally, a series of experimental
iii
studies are performed on an actual power re-circulating gearbox system that is equipped with the
simultaneously. Furthermore, the vibrations at other locations on gear housing and gear whine
are also monitored and analyzed. The inadequacies of system are also quantified, and
iv
ACKNOWLEDGEMENT
I would like to thank Dr. Teik C. Lim, who is serving as the chair of my academic
committee, advisor and research supervisor, for his support and instructions that is important to
this research and my graduate study. I would also like to thank Drs. Jay H. Kim and David F.
Thompson for serving as my academic committee members. Furthermore, I would like to thank
Dr. W. Steve Shepard, Dr. Y.H. Guan, Mr. Chris Moon and Mr. Jim Parker, who collaborated
with me on this same project, for their assistance, suggestions and insights.
The material is based upon work supported by the U.S. Army Research Laboratory and
the U.S. Army Research Office under contract/grant DAAD19-00-1-0158 (Project No. P40942
And, finally my gratitude goes to my wife, Li Feng, and my parents for their patience and
supports.
v
TABLE OF CONTENTS
ABSTRACT................................................................................................................................... iii
ACKNOWLEDGMENTS ...............................................................................................................v
TABLE OF CONTENTS............................................................................................................... vi
LIST OF TABLES......................................................................................................................... ix
CHAPTER Page
1. INTRODUCTION .................................................................................................................1
1.1 Introduction....................................................................................................................1
1.5 Summary........................................................................................................................6
vi
2.1.5 Summary .............................................................................................................14
vii
4.5.3 Performance of Proposed Frequency Estimator .................................................65
7.1 Conclusions................................................................................................................146
REFERENCES ......................................................................................................................151
viii
LIST OF TABLES
Table Page
6.1: List of hardware for the active vibration control system .................................................. 95
6.3: Vibration levels at four locations along gear mesh line-of-action (dB ref. 1g) .............. 137
6.4: Detailed Values of four cross-point transfer functions at four harmonics...................... 140
6.5: Average Phase difference between other locations vibrations and Accy ....................... 140
ix
LIST OF FIGURES
Figure Page
2.3 Effect of constant load on the maximum stroke of PZT stack actuator............................ 18
2.4 Effect of dynamic load on the maximum stroke of PZT stack actuator ........................... 19
2.5 Effect of structure load on the maximum stroke of PZT stack actuator ........................... 20
3.1: A two-degrees-of-freedom torsional dynamic model of a spur gear pair [24] ................. 24
3.3: A 2-noded beam element with transverse u z parallel to the gear mesh line-of-action,
bending rotation x and torsion y coordinates used to construct the input and output
3.5: Dynamic mesh force spectrum of the test gear pair of interest due to 30m of
3.9: Active control applied directly at the housing bearings (concept 3) ................................ 40
x
3.11: Housing response due to a unit magnitude of actuation force .......................................... 42
4.2: The adaptive linear combiner: (a) general form (b) transversal filter............................... 49
4.9: The structure of the proposed FXLMS algorithm with online secondary path modeling
feature ............................................................................................................................... 59
4.13: Result of frequency estimation for linearly time-varying frequency in the presence of
xi
4.17: Comparison of estimated result for with filter and without filter case ............................. 72
4.18: Comparison of estimated result for with filter and without filter case ............................. 73
4.19: Result of frequency estimation applied to measured gearbox vibration data compared to
5.2: Control result using FXLMS algorithm with offline secondary path modeling............... 77
5.3: The finite impulse response of the offline identified secondary path............................... 78
5.4: The frequency response of the offline identified secondary path ..................................... 78
5.6: Amplitude and phase responses of estimated filter of secondary path ............................. 81
5.8: Effect of amplitude ratios of the harmonic components in reference signal on the control
results ................................................................................................................................ 85
5.9: Control result of the frequency jump phenomenon jumping from 300Hz to 400Hz: (a) the
5.10: Control result of the frequency jump phenomenon jumping from 300Hz to 320Hz: (a) the
6.1: Schematic diagram of the closed-loop, power recirculation gearbox setup equipped with
6.3: Closed-up view of: (a) actuation setup, and (b) PZT actuator.......................................... 94
xii
6.4: Enhanced FXLMS algorithm with secondary path identification and frequency estimation
modules ............................................................................................................................. 97
6.6: Comparison of the frequency response functions of the secondary path for modal test
......................................................................................................................................... 107
6.8: Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft rotational
speed of 200 rpm. (Keys: , control off; , 3rd mesh harmonic control;
6.9: Time domain error signal Accy amplitude at the 3rd ( ) and 7th ( ) mesh
harmonics for active control with shaft rotational speed of 200 rpm ............................. 110
6.10: Time domain response of controller output signal at the 3rd ( ) and ( ) 7th
mesh harmonics with shaft rotational speed of 200 rpm ................................................ 110
6.11: Identified secondary path transfer functions. (Keys: , 200 rpm; , 300 rpm)
......................................................................................................................................... 112
6.12: Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft rotational
speed of 300 rpm. (Keys: , control off; , 2nd mesh harmonic control;
6.13: Frequency domain active control result of Accy (dB re. 9.8m/s2) at the 2nd mesh
harmonic with shaft rotational speed of 700rpm (a) and 800 rpm (b). (Keys: ,
xiii
6.14: Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft rotational
speed of 250 rpm. Simultaneous control targets are the 4th thru 7th mesh harmonics with
6.15: Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft rotational
speed of 250 rpm. Simultaneous control targets are 4th thru 7th mesh harmonics with the
6.16: Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft rotational
speed of 230 rpm. Simultaneous control targets are 3rd thru 8th mesh harmonics.
6.17: Vibration response (dB re. 9.8m/s2) for simultaneous control of multiple harmonics: (a)
4th mesh harmonic; (b) 5th mesh harmonic; (c) 6th mesh harmonic; and (d) 7th mesh
6.18: Vibration Spectrum before control ( : low torque; : high torque) ............. 122
6.19: Identified secondary path response while system is operating at 250rpm harmonic
6.20: Frequency domain control result for higher torque case and 3rd harmonic. ( ,
6.21: Frequency domain control result for higher torque case and 5th harmonic ( ,
xiv
6.22: Frequency domain active control result at the 4th mesh harmonic with shaft rotational
speed of 250 rpm. (Keys: , control off; , ALE off and control on; ,
6.23: Frequency domain of first element of FIR controller weights (Keys: , ALE off;
6.24: Frequency spectrum of the controller output (dB re. 1 volt) for shaft rotational speed of
6.25: Frequency domain active control result of Accy at the 5th mesh harmonic with shaft
6.26: Frequency domain active control result of Accy at the 6th mesh harmonic for different
step sizes with shaft rotational speed of 250 rpm. (Keys: , control off; ,
6.27: Time domain amplitude response of the targeted 6th mesh harmonic of housing vibration
6.28: Mean value of FIR controller weights for different step sizes . (Keys: , =0.005;
6.30: Frequency domain of identified secondary path ( N=64; N=128) ............. 134
6.31: Frequency domain of control result for order 128 secondary path FIR model............... 134
6.32: Frequency domain of control result for order 64 secondary path FIR model................. 135
6.33: Frequency domain of control results( ,control off; , control on)............... 136
xv
6.34: Phase difference between Accy and Acc2y. ( ), 4th harmonic; ( ), 5th
6.35: Cross-point transfer functions of the paths from actuator force output to Accy and Acc2y
......................................................................................................................................... 139
6.36: Cross-point transfer functions of the paths from actuator force output to Acc3y and
6.37: Spectrum of gear whine ( ) and gear housing vibration Accy ( ) ........... 142
xvi
LIST OF SYMBOLS
e
K beam : beam element stiffness matrix
e
M beam : beam element mass matrix
I: identity matrix
Hjk: transfer function between k-th reference coordinate and j-th physical
respond coordinate
or secondary path
s: Laplace variable
xvii
x j: displacement of j-th coordinate
L: element length
Id : load inertia
xviii
Ip, Ig: pinion and gear inertia
E: modulus of elasticity
G: shear modulus
t: time
xix
Subscripts
p: pinion
g: gear
h: housing
r: actuator
c: coupling
s: shaft
Superscripts
T: transpose vector
e: element
xx
CHAPTER 1
INTRODUCTION
1.1 Introduction
Gears are essential parts of many precision power and torque transmitting machines.
However, they do generate tonal noise (whine) that is often considered highly undesirable and
annoying. In very severe cases, the gear vibrations can reduce life and performance of the power
transmission systems. Typical gearbox vibration and noise measurements contain several
narrowband tonal signals that are mixed in with a broadband response. The tonal signals are
basically the fundamental gear mesh frequency and its corresponding harmonics generated from
the perturbation in the gear meshing action, which are the primary cause of the annoyance
problem. On the other hand, the broadband signatures are background machine acoustic noise
that is usually not troublesome unless it becomes extremely high. Addressing the gear whine
concern is the main purpose of this research. Although the fundamental gear whine problem has
been of interest to many researchers for a very long time [1-4], the work on the active gearbox
vibration control are still quite limited. Hence, the intent here is to develop and test a new, more
effective method for reducing the annoying tonal gear mesh response by employing a unique
active vibration control approach that will be described further in latter part of this dissertation
proposal. In most active noise and vibration control applications, there are several major
approaches available, such as vibration transmission path control, excitation source control and
receiver treatment. The first two concepts have been of interest in gear vibration control
1
2
applications. Since the vibration control at the source can potentially achieve global vibration
reduction, it is potentially the best choice to use among the above three methods. This
dissertation research will attempt to demonstrate the ability to control the gearbox vibration by
applying internal active actuation scheme directly at the vibration source. An advantage of the
source control approach compared to the previous attempts is the increased potential of reducing
both the structure-borne vibration and air-borne radiated sound pressure responses. The success
of this method largely depends significantly on gaining an understanding of the physics of the
gearbox excitation and vibration transmission mechanisms, and developing suitable actuators,
sensors and active control laws. Thus, a good system model including gear pair system and
actuator should be developed for analyzing the physics of gearbox dynamics. Furthermore, a
suitable actuation concept will be designed from the analysis of the dynamics of gear pair
system. As a result, the gear pair system model and control strategy such as adaptive algorithm
Although numerous applications of smart material systems for vibration control exist [5,
6, 7, 8], many of them only deal with controlling the response of individual modes that are often
less than a few hundred Hertz of simple structures, such as a beam or a plate. However, for
many machinery structures, the modal densities are often quite high especially in the mid and
high-frequency ranges. In these frequency ranges, the individual modes can be difficult to
identify and control. These difficulties lead to the lack of applications of vibration control in
practical industrial system. However, for a gearbox system, unlike the system described above,
the response spectrum is dominated by the mesh frequency (the product of the number of the
3
teeth and shaft speed) and the well-separated harmonics that can span over a wide frequency
range in spite of the higher modal density characteristic. Furthermore, even though the gearbox
system can sometimes exhibit nonlinear and time-varying characteristics under light load
conditions [9], when the transmitted load is sufficiently high, the gear drives often behave quite
linearly in most applications. The discrete nature of the spectral peaks makes this complex
system a potentially suitable candidate for active control using smart material systems.
In 1999, Rebbechi, Howard and Hansen [10] presented a method for performing active
control of a gearbox by actively isolating the vibration between the shaft and housing via a pair
demonstrated that the response of the fundamental gear mesh frequency and its first two
More recently, Chen and Brennan [11] proposed a control scheme that used three
magnetostrictive actuators mounted directly on the gear to produce circumferential forces for
suppressing the torsional vibrations. The experimental results showed about 7 dB of reduction in
gear angular vibrations at the tooth meshing frequencies between 150 and 350 Hz. Other than
these studies, no other recent work on the active vibration control of internal gearbox
components is found in the public domain literature. There are several related studies that
concentrated on active vibration control applied to the structural support system outside of the
gearbox [13, 14, 15] such as the gearbox strut. From these few available studies, it can be
concluded that active gearbox vibration control has not been studied thoroughly. In fact, the
application is very much at its infancy. Seeking for improved understanding to this problem and
From the viewpoint of active vibration control laws for suppressing periodic disturbance
in gear pair systems, there are fundamentally two available control laws. One is the filtered-x
LMS algorithm that is also used by Rebbechi, Howard and Hansen [10], and the other is a
frequency domain adaptive harmonic controller employed earlier by Chen and Brennan [16]. The
filtered-x LMS algorithm is in fact one of the most widely used control algorithms in real world
applications, but in this context, it requires a reference signal that is highly coherent with the gear
mesh frequencies. As a matter of fact, Rebbechi, Howard and Hansen [10] were unable to control
the response at higher-order gear mesh harmonics due to this problem. Fundamentally, the
problem is mainly due to the presence of frequency estimation errors that increase as the order of
the harmonic rises. Hence, the correlations between the reference signals of the higher order
harmonics generated by the multiplier and the true mesh frequencies are quite low. These errors
tend to impede the ability of the controller to attenuate vibration efficiently. To address this
limitation, the authors suggested using a finer measurement scale in combination with a divider
to generate more accurate reference signals at the higher order harmonics. Nevertheless, their
approach still relies on producing highly precise reference signals of the fundamental gear mesh
frequencies.
Frequency domain adaptive harmonic controller that is especially designed for nonlinear
applications [16] was also used to suppress the vibration of a gear pair system [11, 17]. This
adaptive harmonic controller only works for constant mesh frequency cases and requires the
identification of plant sensitivity arrays at the critical mesh harmonics [16]. The algorithm also
relies on having accurate gear mesh frequency data like in the case of the filtered-x LMS
algorithm noted above, and may be further limited by the fact that the reference signal and
5
sensitivity arrays are time-varying in practice due to either variation in gear speed and/or mesh
In spite of this lack of active gear vibration control investigations, there exist several
algorithms to control periodic disturbances for generic mechanical plants. Some of these
techniques include classical feedback control [18], LMS-based adaptive feed-forward control
[19], higher harmonic control [20], linear quadratic gaussian (LQG) and H? [21], and model
reference-based adaptive control (MRAC) [22]. However, these existing algorithms either
require a detailed model of the whole control system, or require exact information of the gear
rotation speed or the fundamental gear mesh frequency. A detailed discussion of the control
Because many conventional actuators, such as electromagnetic and hydraulic shakers, are
massive and have a limited frequency bandwidth, actuators made from smart materials have
undergone rapid development recently [23]. With recent technological advances in material
science and control theory, many forms of smart materials have emerged for use as advanced
sensors and actuators. These materials are able to overcome some of these drawbacks of
conventional sensors and actuators, such as size, that are not convenient to implement. Typical
memory alloys, electrorheological fluids, and fiber optics, which have been successfully applied
in a variety of fields.
have a large frequency bandwidth and are able to generate a large dynamic force. As a result,
these two kinds of actuators are often applied in active vibration control of simple structure, such
6
as beam and plates [5, 6, 7, 8]. As noted above, magnetostrictive actuators have been used in the
gear vibration control [10, 16]. However, magnetostrictive actuators are typically bulkier than
piezoelectric actuators although the former can generate larger forces. One the other hand, other
types of actuator made from shape memory alloys or electrorheological fluids have a relatively
low frequency bandwidth. Consequently, piezoelectric actuators are selected for this research. A
1.5 Summary
As mentioned above, the use of active vibration control in gearbox systems is very
limited. The lack of applications may be due to the facts that: (a) the physical mechanisms of the
gear mesh excitation and vibration transmission control are not well established [2, 24,]; (b) the
bandwidth of the gearbox system dynamics is large, up to several thousand Hertz; (c) it is not
easy to find an actuator setup which has enough force and bandwidth; (d) the model of the smart
material actuators is not well defined [25]; (e) the gear system dynamics can be a time varying
due to variation of rotation speed and gear mesh stiffness; and (f) the control algorithm has to
compensate between the control performances and DSP hardware computational limitations [19].
These current limitations provide a need for further development of an active gear vibration
control system. In the following chapters, a simple model of piezoelectric stack actuator, proper
actuation concepts, proper control laws and the proposed experimental validations are discussed.
Note that the research discussed in this proposal has potential applications in the rotorcraft
Chapter 2 presents a detailed overview and applications of smart materials, such as the
piezoelectric actuator is also presented in this chapter. This model will be used to guide the
Chapter 3 presents a dynamic finite element (FE) model of the gear pair system used to
understand the vibration generation and transmission mechanisms, and also for demonstrating
the impact of the proposed active vibration control method. The four potential actuation
concepts for the gearbox system will be introduced there. To select a proper actuator, the power
requirements of actuator and amplifier based on the proposed FE gear pair system model and
piezoelectric actuator dynamic model are predicted. Through this analysis, the active shaft
transverse vibration control method is adopted for further investigation in this dissertation
research.
Chapter 4 presents a filtered-x LMS based control law for the proposed active shaft
transverse vibration control. The filtered-x LMS algorithm is one of the most common control
algorithms in real applications. However, in this context, the identification of the secondary path
has to be sufficiently precise and reference signals that are coherent with the gear mesh
frequencies are readily available. To overcome these two drawbacks, two enhancements are
applied. The first enhancement uses an offline/online secondary path identification using
additive random noise method. The second one applies an online frequency estimator for
estimating the gear mesh frequency directly while adding a reference signal synthesizer to the
control algorithm.
In Chapter 5, a series of computer simulation studies are performed in order to guide the
experimental work.
Chapter 6 discusses the result of experimental study where the designed active control
Chapter 7 gives the conclusion and suggestion for future potential research work.
CHAPTER 2
In this chapter, the existing smart materials and actuators that potentially can be used for
gearbox applications are reviewed and discussed. There are several typical smart materials, such
fluids [26]. In the following sections, general properties as well as applications of relevant smart
material systems are briefly discussed. Then, one piezoelectric stack actuator is considered in
detail and modeled as a lumped mass - spring system for the mechanical part and a lumped
Magnetostrictive materials are the substances that can generate strains upon application
of magnetic field. Conversely, a magnetic field is generated when the material is stressed. This
field, however, is proportional to the materials rate of strain [27, 28]. Although
magnetostrictive materials have been extensively studied, few practical applications exist
because these materials do not generally produce large strains and forces like piezoelectric and
electrostrictive materials for the same level of power consumption. This fundamental
disadvantage has been recently changed with the development of the so-called giant
magnetostrictors, namely Terfenol-D. This material is an alloy of iron, and the rare earth
elements of Terbium and Dysprosium. These magnetostrictive materials are able to offer strains
9
10
an order of magnitude more than conventional piezoceramics with comparable force outputs.
Therefore, one of the primary advantages of a magnetostrictive actuator is that it can generate
very large strains in comparison with those created by piezoelectric actuators. However, the
constitutive relationship between applied magnetic field strength and generated strains is highly
nonlinear and exhibits significant hysteresis. Furthermore, magnetostrictive actuators are larger
micropositioning, damage mitigation and shape control. Preliminary results show that simple
analog control systems consisting of Terfenol-D actuators can significantly reduce flexural
vibrations in a rotating machine tool shaft [29]. Recently, three magnetostrictive inertial
actuators were clamped to a helicopter gearbox support strut to minimize the kinetic energy of
The piezoelectric effect is that certain crystalline material produces an electrical charge
on its surface when it is subject to a stress field, which was first discovered in 1880 by Pierre and
Currie [30]. It was also demonstrated that the converse effect, when an electric field is applied to
a piezoelectric material, its shape and size change, is also true. Piezoelectric materials generally
used in smart structures come in two forms, namely piezoceramics and piezoelectric polymers.
The piezos ability to actuate the structure is a function of its stiffness, electromechanical
coupling coefficients, flexibility, and applied voltage limitations. Piezoceramics, such as Lead-
zirconate-titanate (PZT), have large stiffnesses and mechanical coupling coefficients. Thus, they
are used as actuators. And, polymer types are usually applied as sensors because of their high
(PZT) can generate a large force up to 10 KN, offers very good response time (microsecond time
Piezo-actuators are applied as either distributed form or stacked form. For distributed
form, two configurations are basically used: surface bonded to the host structure, and embedded
in the structure. For both configurations, the position of the actuator, the orientation of the patch
polarization, and the polarity of the input determine whether the induced loading is pure bending
or axial extensional / compressional loading. As an example, if only one surface of the host
structure is bonded with a piezo-actuator (i.e. a unimorph configuration), then it will exhibit a
bending effect. The unimorph configuration has been used in speakers for noise production and
cancellation applications. If both sides of host surface are bonded using a pair of piezo-actuators
with opposing polarity (a bimorph configuration), then the structure will experience axial
extensional or compressional loading effect. On the other hand, embedding the actuator allows
the designer to place it exactly where it is needed to get a desired effect. However, this design
can limit the service temperature of the structure, interrupt the load path, and prevent service or
repair work. As for the piezo-stack, they contain multiple layers of piezoceramics, which are
coupled mechanically in series and connected electrically in parallel. The stack are capable of
producing kilo-Newtons of force with only microns of expansion when connected with an
inertial mass or contained in a stiff structure. Their response time is around ten microseconds.
Piezoelectric elements have been used as actuators and sensors in a wide range of
precision mechanics, vibration and noise control, and even life science and biology [23]. They
have been successfully embedded [31] in composites, and surface mounted on both composites
12
and other substrates [32]. Typical composite applications include vibration damping [33],
vibration isolation [34], and acoustic control [35] due to the adaptability and lightweight
characteristics. In addition, the ability to be easily integrated into structures makes the
piezoelectric elements very attractive for structural control since no moving parts are needed
Shape memory alloys are a special class of adaptive materials that convert thermal energy
directly into mechanical work. A variety of alloys, when properly mixed, exhibit this conversion
effect by heat treatments. Such an effect is accomplished through a phase change and can impart
large strains. The heat may be applied by numerous means including electrical resistive heating.
However, the frequency bandwidth of the actuator is limited by the speed at which it can be
Common SMA forms are wire, ribbon, sheets and springs. They can also be conditioned
to go between two preset shapes to produce two-way shape memory effect. The advantages of
SMAs are: a) high power density (ratio of actuator output to actuator volume); b) superelastic
property of Nitinol(the most common type of shape memory alloys), which allows recovery from
strains of up to 6 percent without being heated; and c) excellent cycling performances in repeated
loading. As the result, SMAs have been used for applications like eyeglass frames, coffeepot
thermostats, satellite release devices, electrical couplings, robot, position control and vibration
control. However, SMA-based actuators have some disadvantages including low frequency
bandwidth, large non-linear hysteresis, and difficulty in developing good dynamic models.
13
Electrorheological fluids, which were first investigated by Winslow in 1949 [23], exhibit
a coupling between their fluidic and electrical properties. Typical ER fluid consists of
alumina powder and silica gel have all been used as particles in ER fluids. Without an applied
electric field, the ER fluid behaves like a Newtonian fluid. However, once a sufficient electric
field is applied, the fluid undergoes a dramatic change, inducing a Bingham fluid-like behavior.
The most significant change is associated with complex shear modulus of the material.
Most ER material applications can be categorized into two classes [26]: controllable
devices and adaptive structures. Controllable devices applying ER studied include valves,
clutches, engine mounts, brakes, dampers or shock absorbers, and damping of helicopter blades
[25]. On the other hand, adaptive types possess tunable properties by incorporating controllable
ER material [36] or damper [37]. Yao [38] studied a disk type ER damper to actively suppress
the vibration of a rotor system theoretically and experimentally. He showed that the ER damper
can suppress large vibration amplitudes around the critical speed, dampen any sudden
unbalanced response, shorten the transient process, and reduce the amplitude of the steady-state
response. In spite of these successes, one of the main drawbacks of ER fluids, which keep them
from being used in many practical applications, is the need for very high voltage levels.
2.1.5 Summary
Among the above four types of smart materials, the piezoelectric elements and
magnetostrictive materials have the fastest response time, and their effective frequency range can
reach up to 10 kHz. Furthermore, they can provide large static and dynamic forces (up to 10 kN).
In contrast, SMA possesses the slowest response time. Although ER fluids have reasonably high
response time, they are difficult to integrate into a gear system due to lack of structural strength.
Since gear vibrations are dominated by high frequency tones, piezoelectric elements and
magnetostrictive materials appear to be the most suitable candidates for active gear vibration
control application. Accordingly, the model of piezoelectric element is discussed further next.
Piezoelectric elements have been studied and applied for active control applications
extensively [29, 31, 32, 33, 34, 39]. However, there are relatively few papers addressing the
modeling of piezoelectric stack actuator. A simple lumped parameter linear model for the
piezoelectric stack actuator based on the impedance model from Flint and Liang [40] is
presented. It is known that if the applied electrical field density is small, then the piezoelectric
element can be treated as a linear system. However, if the electrical field density is too high,
The relationship between the applied force and resultant response of a piezoelectric
material depends upon a number of parameters, such as shape, composition, properties, electrode
position, poling axis location (that requires a set of suitable coordinate system be defined), and
the applied field polarity. Figure 2.1 shows an element of piezoelectric material and a coordinate
system [23]. Three axes are used to identify the directions of the piezoelectric element, namely
15
3(z)
Applied
+ Field E 2(y)
V p
1(x)
Under an applied electric field that is sufficiently small, the mechanical strain of the
piezoelectric element reacts proportionally to the electric field density to produce a constant
mechanical stress. This is characterized by the piezoelectric strain coefficient denoted by dkp
(see [41]), which is defined as the ratio of free strain to the applied electric field E. The first
subscript in the strain constant gives the direction of the electric field associated with the voltage
applied or the charge produced, while the second subscript gives the direction of the mechanical
strain. The constant dkp can also be viewed as a relationship between the charge on the electrode
and the applied mechanical stress. In particular, the strain constants d33, d31 and d32 are of
primary interest. In fact d31 = d32 = -d33, where v is Poisson ratio, since the piezoelectric
A piezoelectric stack actuator shown in Figure 2.2 is considered here. The actuator is
Each layer has a thickness of t and the total height of the actuator is L. The polarization direction
is parallel to direction 3.
16
Due to the rod-type configuration of the actuator shown in Figure 2.2, only the
extensional stress along the length of rod, that is direction 3, exists. Thus, the constitutive
D3 = d 33T33 + 33
T
E3 , (2.1b)
where S33 is the strain along the rod, T33 is the stress, is the complex compliance at zero electric
and D3 is the electric displacement which is given by the electric charge q over cross area A.
But, this set of equations is used in the micro-crystal form and not easy to use directly in the
model. However, if the working frequency of the actuator is far less than the first resonant
frequency of the actuator [40, 42, 43], the strain of each PZT layer can be assumed to be same.
Applying equation (2.1a) and (2.1b) to each layer and summing them up can yield the macro
equation of the piezoelectric stack actuator. Thus, for externally applied force: F = AT3 , the
transformed into
F = k p x d 33 k pV , (2.2a)
q = d 33 F + C ps V , (2.2b)
Note that the second term of right-hand-side of equation (2.2a) can be defined as block force of
the actuator and also can be approximated as the multiplication of its stiffness and maximum
stroke
Fb = d 33 k pV k p l , (2.3)
where l is the maximum stroke of actuator. Actually, the block force Fb is the maximum output
force that the actuator can generate when it acts on a rigid structure with infinite impedance.
valuable to view the effect of external loads on the maximum stroke of the actuator can achieve.
The simplest external load is applying a mass on top of the actuator. For this case,
solving equation (2.2a) will get the instantaneous displacement of the actuator
x = d 33V + F / k p , (2.4)
Note that the second term of right-hand-side F / k p is constant, so the maximum stroke of the
actuator, L0 , is the same as the zero-applied force case. Compared with the free zero-applied
Figure 2.3. Effect of constant load on the maximum stroke of PZT stack actuator.
2) Spring load
When the external load is not constant or a mass, such as a spring load, then the
externally applied force is related to the instantaneous displacement of the actuator. Assume a
spring element is connected to the actuator as shown in Figure 2.4, and the spring stiffness is
F = k ext x , (2.5)
kp
x = d 33V , (2.6)
k p + k ext
As seen in equation (2.6), the maximum stroke of the actuator will be reduced. For a
stiffer external spring, the maximum stroke of the actuator will be smaller. This effect is shown
in Figure 2.4. If the external spring is infinitely rigid, then, the actuator cannot present any
stroke at all and will output the block force to external structure. If actuator is free of end, it will
Figure 2.4 Effect of dynamic load on the maximum stroke of PZT stack actuator.
For a more complex external load, such as when the actuator is acted on by a complex
host structure, the equivalent dynamic stiffness can be used. For example, when the actuator is
applied on a shaft structure, as shown in Figure 2.5, then the driving point stiffness can be used
as the equivalent dynamic stiffness. Note that the equivalent dynamic stiffness might be a
complex value, which means the displacement response will have a different phase relative to the
force excitation. The equivalent dynamic stiffness can be calculated from experimental results or
a finite element modeling. One can still use equation (2.5) to calculate the external applied force
to the actuator. Consequently equation (2.6) can be used to calculate the displacement response
of the actuator. Substitution of equation (2.6) into equation (2.5) yields the effective force
k ext k ext
F = k p d 33V = Fb , (2.7)
k ext + k p k ext + k p
where block force Fb is defined in equation (2.3). From the above equation, the effective force
applied on the host structure depends on the stiffness of the host structure and the effective
piezo-stack stiffness. The stiffer the host structure is, the more force the actuator can generate.
20
Shaft structure
kext
kext
Piezo-
actuator x kp
Fb
Figure 2.5 Effect of structure load on the maximum stroke of PZT stack actuator.
Furthermore, the electrical charge applied to the actuator can be calculated by equation
k ext
q = C ps [1 k 2 ]V , (2.8)
k ext + k p
where k 2 = d 332 / 33
T E
s33 is the piezo coupling coefficient, which is always less than unity. From
k ext
C p = C ps [1 k 2 ], (2.9)
k ext + k p
Taking a time derivative of equation (2.8) and substituting equation (2.9), the electrical
I = i C p V , (2.10)
where i is unit imaginary number and is the angular frequency of the applied voltage. It
should be known that piezoelectric elements can only bear compressive force and not tensile
force. Thus, in order to apply the piezo-stack actuators into a dynamic application, the actuator
must be preloaded.
21
This actuator model of complex structure load will be used to calculate the required
voltage and current for active gear housing vibration control in Chapter 3.
CHAPTER 3
This chapter discusses the development of the gear pair system model and the ensueing
dynamic analysis performed. The dynamic model is used to assist in the design of the
experimental setup, to gain a better understanding of the nature of gear response due to
transmission error excitation, and to help in the design of the control algorithm. The simulation
conducted is intended to determine the effectiveness of the proposed active control concept and
to compute the required actuator parameters. First, a brief review of existing gear dynamic
models is given. Second, a finite element (FE) model is presented. Third, the proposed actuator
configuration for the gearbox system is introduced. At last, the electrical power requirements
based on gear system model and piezo-stack actuator model are presented.
Analytical work on the dynamics of gear pair systems is abundant. This includes a series
of mathematical dynamical models that have been developed for different purposes. Reference
[24] groups the existing models into 5 different classifications based on the specific purpose of
the models as well as the dynamic characteristics of the primary elements in the system. The
classifications are (a) simple dynamic factor models; (b) models only with tooth compliance; (c)
gear dynamics models that include the flexibilities of other elements as well as the tooth
compliance; (d) geared rotor dynamics models in which the transverse vibrations of gear-
carrying shafts are considered in two mutually perpendicular directions, and thus allowing the
22
23
shaft to whirl; and (e) torsional vibrations models in which the flexibility of gear teeth is
The classifications (a)-(d) are based on the work by Tuplin [45] in the 1950s, which are
of interest to the present study. In his work, Tuplin introduced a simple spring-mass gear
dynamic model that can be used to estimate dynamic factors at frequencies well below the gear
mesh resonance mode. His theory forms the basis for modern gear dynamic models used
extensively today. Figure 3.1 shows a typical model comprising of a gear mesh stiffness element
that forms the coupling between the 2 rigid gear bodies. Using this concept, Ozguven and
Houser [46] developed a torsional model to analyze the dynamic behavior of spur gear pairs due
to loaded static transmission error excitation. In addition, Amabili and Rivola [47] applied a
similar model with time-varying gear mesh stiffness characteristic to compute the steady-state
response and to examine the stability of a spur gear pair. The calculations are made assuming
viscous damping that is proportional to the mesh stiffness. In a subsequent paper, Amabili and
Fregolent [48] reported the use of this model to identify critical modal parameters and gear errors
from gear vibrations. Other more recent studies that applied the same class of models are by
Zhang and Rescigno [49] who used a four degrees-of-freedom dynamic model assuming rigid
bearings to estimate the impact load on gear tooth due to torque pulsations. Cricenti et al. [50]
used a six degrees-of-freedom model that included both translation and torsion coordinates to
perform both static and dynamic transmission error calculations, and Lim and Li [51] developed
a general multi-mesh counter-shaft transmission model for analyzing coupled torsional and
translational response. Using essentially the same type of models, a team led by Houser [52, 53]
yg
Km
yp
Cm e(t)
~
Figure 3.1. A two-degrees-of-freedom torsional dynamic model of a spur gear pair [24].
Steyer and Lim [54] also formulated a theory to relate the dynamic mesh stiffness to the
driver and driven compliances at the mesh. The concept enables one to determine the controlling
factor and compute the sensitivity of mesh force to drivetrain system dynamics. Using this
formulation, Donley, Lim, and Steyer [55] developed a unique scheme to compute the effect of
Kahraman et al. [56] also extended the basic gear modeling concept to devise a more complex
finite element model that considers the effect of rotary inertia of the shafts, bearing flexibility,
and gear mesh stiffness to study the dynamic coupling between torsional and transversal motions
of the gears. Other more sophisticated finite element models have been proposed by Parker et al.
[57], Yakhou et al. [58, 59] and a team of researchers led by Velex [60-63]. Some of these finite
element models utilized more sophisticated mesh models compared to the one shown in
Figure 3.1. For example, the one proposed by Parker et al. [57] actually simulates the detail gear
engagement process by combining finite element and contact mechanics. On the other hand, the
one used by Velex relies on a discretization of the mesh stiffness along the line of contact.
Another class of geared system model is generated, by applying the finite element
method. In this research, a finite element model with lumped parameters of gear mesh model is
developed to model the single-stage geared rotor system. The widely used gear mesh modeling
25
concept, applying a unidirectional lumped stiffness element, is adopted. In this study, the finite
element approach will be employed. The formulation of the dynamical equations of motion is
presented next.
A cross-section of the power re-circulating gear system considered in this study is shown
in Figure 3.2. This system consists of two pair of gears, a belt driving motor, a torque
transducer, 3 shaft-to-shaft couplings, two additional support bearings at two sides of flywheel
on driven shaft, and a series of driver and driven shaft segments. The main noise and vibration
source of interest to this study is the gear transmission error at the mesh of test gear set. It is
commonly known that the transmission error excitation interacts with the system dynamic
characteristics to generate a dynamic mesh force that in turn excites the rest of the gearbox
structure. The resulting vibrational energy is transmitted through the shafts and bearings and
into the housing where some of this energy is ultimately radiated as high-frequency gear whine
noise. Since the frequency of the gear tones are significantly higher than the shaft rotational
speed, they can be easily monitored and identified. This frequency identification ability is
Housing
Motor
Torque
transducer
Figure 3.2. The picture and model of power re-circulating gear system.
A dynamic FE model of the gear pair system is first developed for use in examining the
proposed actuation concept. The model, which is constructed from beam elements and lumped
masses and springs, provides an analytical representation of the critical gear rotational and
translational vibration characteristics. The shafts are assumed flexible in the transverse and
torsional directions. The longitudinal degree of freedom is neglected since no significant axial
excitation is expected from the spur gears and to avoid unnecessary complexity in the analysis
[64]. Moreover, the natural frequency associated with the longitudinal vibration is much higher
than the frequency range of interest. Accordingly, each nodal point on the discretized shaft is
only defined by three degrees-of-freedom (DOF) that include transverse, bending rotation, and
torsion coordinates as shown in Figure 3.3. The beam element stiffness matrix formulation can
be shown to be
27
12 K 1 6 K1 L 0 12 K 1 0
6 K1 L
6K L 4 K1 L 2
0 6 K1 L 2 K1 L 0
2
1
0 0 Ks 0 0 Ks
e
K beam = , (3.1)
12 K 1 6 K1 L 0 12 K 1 6 K 1 L 0
6 K1 L 2 K 1 L2 0 6 K 1 L 4 K 1 L2 0
0 0 Ks 0 0 K s
3
where L is the element length, K s = GJ L , K 1 = EI L , E is the modulus of elasticity, I and J
are the area and polar moments of inertia, and G is the shear modulus. The corresponding mass
matrix is
where m is the element mass and R is the shaft radius. These matrices correspond to the
coordinate vector {u z1 , x1 , y1 , u z 2 , x 2 , y 2 }T .
z
uz
y
z y
2
uz
y x x
y
1
x x
Figure 3.3. A 2-noded beam element with transverse u z parallel to the gear mesh line-of-
action, bending rotation x and torsion y coordinates used to construct the input and output
shaft models. The nodes are labeled as points 1 and 2.
28
The gear mesh kinematics is modeled using a concept originally proposed by Tuplin [46],
which has been widely used by many gear researchers [24, 46, 50]. The model consists of an
infinitesimal spring-damper element positioned in series with the transmission error excitation
e(t) at the mesh point as shown in Figure 3.1. The mesh model couples the translational
coordinates of the gear and pinion centroids along the line-of-action. Additionally, the bending
rotation and torsion motions of the gears are considered. Hence, each gear body is represented as
a lumped mass element whose instantaneous position is defined by one translational and two
rotational coordinates expressed as uz, x, and y. The proposed gear mesh model is shown in
Figure 3.4. For the combined gear-pinion system, the position vector is
Km 0 K m Rg Km 0 K m Rp
0 0 0 0 0 0
K m Rg 0 K m Rg2 K m Rg 0 K m Rg R p
K e
= , (3.3)
Km
mesh
0 K m Rg Km 0 K m Rp
0 0 0 0 0 0
K m R p 0 K m R p Rg K m Rp 0 K m R p2
m g 0 0 0 0 0
0 I xg 0 0 0 0
0 0 I yg 0 0 0
M emesh = , (3.4)
0 0 0 mp 0 0
0 0 0 0 I xp 0
0 0 0 0 0 I yp
where Km is the gear mesh stiffness, and (Rg, Rp), (mg, mp), (Ixg , Ixp), and (Iyg , Iyp) are the radii,
masses, and mass moment of inertias of the gear and the pinion about the x and y directions,
29
respectively. Note that Equations (3.3) and (3.4) are similar to the one proposed by Lim and
Singh [65]. Here, our model has been extended to include bending rotation.
yg
uzg xg Km
yp
Rg uzp
Cm e(t) xp
~
Rp
Figure 3.4. Used infinitesimal spring-damper gear mesh model.
The belt driving motor is modeled as lumped mass moments of inertia, while the bearings
are formulated using lumped spring elements having the ability to resist motion in the transverse
(parallel to the line-of-action) and bending rotation directions. The housing structure is
represented as a 1-DOF rigid body (lumped mass element) connected to the gear train at the 4
bearing locations via a set of bearing stiffnesses. This rigid body housing approximation does
not pose significant limitation to the simulation work, as the focus of this study is to control the
overall housing vibrations via the internal gear-shaft dynamics. In fact, this simplified housing
model eases simulation efforts. The assembled system model contains 51 finite elements
including 12 lumped masses, 10 lumped springs, 27 beam elements, and two gear mesh stiffness
elements, and correspondingly twenty-nine 3-DOF nodes and two 1-DOF nodes for the housing
of slave and test gear setups. This gives a total of N = 89 DOFs for the complete system
coordinate set.
The system mass M and stiffness K matrices, each of dimensions N? N, can then be used
&& + CX
MX & + KX = F , (3.5)
where X is the response vector, F is the forcing vector containing the transmission error term, C
is the damping matrix assumed to be equivalent to the typical 5% uniform modal damping
& and X
observed for gearing systems, and X && signify the first and second time derivatives of X.
Here, M and K are the summations of the expanded version of each element matrix derived
using a standard FE formulation [64]. Note that Equation (3.5) is linear and its coefficient
matrices are time-invariant. An output vector, Y = DX , is defined for a subset of the n response
coordinates of interest, where D is the output coefficient matrix of dimension n? N. Applying the
resulting in the identity matrix. The resulting dynamical equations in terms of these uncoupled
q&1 0 1 0 0 ... q1 0
q&& 2 2 1 1 0 0
... q&1 F
1 1
1
T
q& 2 = 0 0 0 1 ... q 2 + 0 , (3.6)
q&& 0 0 2
2 2 2
... q& 2 F
2 2
2
... ... ... ... ...
... ... ...
where Fj, (j=1,2N) is the element of the physical force vector F, i is the i-th natural
frequency, and i is the corresponding modal damping coefficient. In terms of the modal
Y = D Q . (3.7)
31
Solving Equation (3.6) for qi and transforming them back into the physical coordinates yields the
output response of interest. Accordingly, the transfer function between any k-th reference
xj N ji ki
H jk = = , (3.8)
Fk i =1 s 2 + 2 i i s + i2
where s is the Laplace variable. This FE model and the corresponding transfer function results
shown in Equation (3.8) are employed in the subsequent active vibration control simulations to
examine the true capabilities and limitations of the four proposed actuation concepts to minimize
For the power re-circulating gear system shown in Figure 3.2, the design parameters of
each component are listed in Table 3.1. Using these parameters, the above FE system model can
be obtained.
Since the dynamics of the gear pair system is mainly excited by the transmission error at
the gear meshing point, the resulting dynamic mesh force is normally used to describe the
severity of the dynamics of the geared rotor system. For this power re-circulating gear system,
the transmission error on the slave gear mesh point can be neglected, because its smaller than
the transmission error on the test gear mesh point. Hence the formulated force term is really a
function of the magnitude of transmission error, mesh stiffness, and the pitch radii. As shown in
Figure 3.4, the transmission error excitation leads to a pair of translational dynamic force and
33
torque acting on the gear and pinion. This dynamic force vector F in Equation (3.6) is expressed
as
F = [0 L 0 K m 0 K m Rg 0 L 0 K m 0 K m R p 0 L 0 ]T e , (3.9)
u zg yg u zp yp
yp are the gear and pinion translation and rotation coordinates. Here, E=30 is assumed.
Omitting the damping term, the dynamic mesh force can be formulated as [45-46]
Fmesh = K m (u zg u zp Rg yg R p yp e) . (3.10)
The predicted dynamic mesh force for the power re-circulating gear system of interest (see Table
3.1 for actual design parameters) is shown in Figure 3.5. The trend shows a general increase in
Fmesh as frequency increases. And, there is a peak around 3 kHz. This is due to the contribution
from the fundamental mesh mode (out-of-phase rotation of the gear pair) at around 3 kHz for this
9
10
Amplitude (N)
8
10
7
10
2 3 4
10 10 10
Frequency (Hz)
Figure 3.5. Dynamic mesh force spectrum of the test gear pair of interest due to 30m of
transmission error excitation.
The vibratory response of the gear housing due to the unit magnitude of the dynamic
force at the gear mesh coordinate, expressed by Equation (3.11), can be calculated using the
modal superposition-based transfer function theory given by Equation (3.8). By scaling the
equation linearly to reflect the total mesh force and summing the contributions from the resultant
dynamic forces and torques on the gear and pinion, the net displacement response of the housing
4 N
hi 4
y h ,TE = H hk Fk = [ Fk ] , (3.11)
s + 2 i i s + i2
2 ki
k =1 i =1 k =1
where Fk, k=1 to 4, are the four non-zero forcing elements of Equation (3.9). By taking the
obtained, which is shown in Figure 3.6. From this figure, one can see that the responses higher
35
than 300 Hz tend to be almost same except the valley around 350 Hz. Higher response means
1
10
0
10
Amplitude (m/s )
2
-1
10
-2
10
-3
10
0 200 400 600 800 1000
Frequency (Hz)
The dynamic mesh stiffness is the frequency response function relating mesh force to
Fmesh (i )
Dmesh (i ) = , (3.12)
e (i )
where Fmesh (i ) and e(i ) are Fourier Transform of mesh force and transmission error,
respectively. From Equation (3.12), the dynamic mesh stiffness should be nearly the same shape
curve as the mesh force shown in Figure 3.5. The dynamic mesh stiffness is also used to analyze
the response of the geared rotor system. The larger dynamic mesh stiffness leads to higher
response of system.
36
Since the actuator is one of the primary elements and its setup may significantly affect
the performance of the active vibration control system, it is highly desirable to be able to identify
an effective configuration. There are four candidate active control concepts available. They are:
(1) active inertial actuators positioned tangentially on the gear body, (2) semi-active gear-shaft
torsional isolation coupling, (3) direct active bearing vibration control, and (4) active shaft
transverse vibration control. These approaches are designed to either impede the transmission of
vibration energy into the housing or reduce the effect of the dynamic mesh force excitation.
Here, these concepts are analyzed to determine the levels of required actuation efforts and the
corresponding driver electrical power requirements. Note that while the overall goal is to reduce
the housing response, each concept attempts to control the vibrations at different structural points
inside the gearbox system. Therefore, their control requirements may be inherently different.
The first approach uses three inertial actuators, which can be made of magnetostrictive or
piezoelectric stacks, directly mounted tangentially onto the side of the gear body at equal angular
intervals, as shown in Figure 3.7. As the smart material expands and contracts at the selected
excitation frequency, the acceleration of the inertial mass mounted on one end of the actuator
will provide a reaction force that acts on the gear body. Due to the circumferential arrangement,
these actuators can produce a net torque fluctuation as well as a translation force parallel to the
mesh load line-of-action. The relative amount of these two actuation loads can be controlled by
appropriately phasing the drivers. The net dynamic torque and force can then be applied in a
controlled manner to reduce gear torsional and translational vibrations. Since this actuation
37
concept directly acts near the excitation source, it is expected to be able to suppress vibrations
transmitted into the shaft and bearings, and also reduce dynamic mesh force production.
Additionally, the overall design includes three miniature accelerometers, located near the
actuators, to measure gear vibrations and provide feedback signals to the controller.
To keep the actuation design as simple as possible, only the driven gear (or pinion) is
treated. Since one of the purposes is to minimize the generation of dynamic mesh force, which
involves lowering dynamic transmission error dictated by the relative motion at the mesh, it is
only necessary to actuate one of the two gears. The disadvantage here is the lack of vibration
transmissibility control through the shaft of the untreated gear. Note that a related design was
used by Chen and Brennan [11] as indicated earlier. The fundamental difference is that the
control object of Reference [11] was the gear torsional vibration, while the goal in this work is to
Mounting flange
Pinion Gear
Slip ring Fp
Shaft
Magnetostrictive or
piezoelectric actuator Accelerometer
Inertial mass (mounted on gear body)
Other considerations include the design of the fundamental resonant frequency of the
actuator, which ideally needs to be near or less than the excitation frequency, and the fact that we
have a set of rotating actuators requiring the use of slip rings. In the former issue, either an
38
configuration may be needed to lower the first resonant frequency below the working frequency.
This is especially critical when a piezoelectric type actuator is selected since its resonant
frequency is inherently high. In the latter problem, suitable slip rings with proper power ratings
must be used to ensure sufficient supply of power to the rotating actuators. Of course, the slip
The second proposed concept is aimed at modifying the torsional vibration transmission
path between the gear and shaft. In this setup, the gear body and corresponding shaft are
connected via several piezoelectric actuators, as shown in Figure 3.8 for one set of gear-shaft
coupling setup. The piezoelectric actuators serve two purposes. They transmit the mean torque
in the gear-shaft load path and simultaneously generate reactive dynamic forces to minimize
transmitted perturbations. Like the first concept, slip rings are needed to provide input power to
the actuators. Note that the translation force of the gear is transmitted via a rolling element
This particular actuation concept is similar to the application of an active dynamic force
directly on the shaft (the receiver) with a reaction against a gear body (the mass), as described in
Reference [8]. Therefore, the dynamic force required to isolate the torsional vibration
transmitted from the gear into the shaft is inversely proportional to the excitation frequency.
Accordingly, at relatively high frequency, only a small dynamic force is needed. However, when
the frequency is lower, a large dynamic actuation force is required. Note that this scheme may
also provide some level of passive torsional vibration isolation in the lower frequency range due
F
Accelerometer (mounted on gear body)
F
Piezoelectric actuator
Since most gear whine problems are the result of structure-borne vibration transmission
from the geared rotor system through the bearings and into the housing, a natural path control
scheme inside the gearbox is to setup choke points at the support bearings. One way to
accomplish this is to use two pairs of piezoelectric stack actuators placed in two orthogonal
radial directions between the bearing raceway and the housing support structure, as shown in
Figure 3.9. One pair of opposing actuators is oriented parallel to the gear mesh line-of-action,
while the second pair is oriented perpendicular to the line-of-action. Note that although only one
pair of actuators oriented parallel to tooth load line-of-action is needed theoretically, this double
set configuration is recommended in practice to achieve a more robust ability to control general
transverse plane motions of the shaft-bearing structure. This is especially useful when non-
negligible transverse vibration orthogonal to the load line-of-action caused by misalignment and
friction force excitation exists. In this work, the analyses of concepts 2 and 3 only incorporate
Gear
F pair
Housing
Pinion Gear bearing
F
Piezoelectric
actuators
Housing
Figure 3.9. Active control applied directly at the housing bearings (concept 3).
This specific actuation concept is designed to directly add active dynamic forces to the
housing support structure by reacting against the bearing raceway, which may effectively
suppress the housing vibration. Unlike the previous two concepts, the actuators here are
stationary. As a result, slip rings are not needed in this design. Nevertheless, the housing
supports must be modified substantially to accommodate the actuators, which may not always be
feasible.
In this configuration, two pairs of actuators are attached to the rotating shaft using an
additional set of bearing as shown in Figure 3.10. The actuators are acting against an additional
redundant bearing along the line-of-action of gear transmission error excitation. Although the
dynamic force is still being introduced between the shaft and housing, the added bearing
component does not carry the required static load. Therefore, this bearing can accommodate a
large range of motion. A similar idea was examined by Rebbechi, Howard, and Hansen [10],
where magnetostrictive actuators were used. The proposed system here applies piezoelectric
stack actuators along with an improved control algorithm to be described later in this proposal.
41
For the present analysis, it is assumed that the actuation position is located at one-third of the
length from the gear position to the housing support location solely based on packaging
considerations. Furthermore, due to the possible run-out and vibration deformation of the shaft,
the actuators may see some lateral motions that may be harmful to the actuators. Hence, the
setup must be carefully designed such that the actuators are not directly connected to the bearing
raceway. Instead a thin stinger rod as shown in Figure 3.10 is used. This stinger rod provides
some degree of lateral flexibility, thereby reducing the potentially harmful lateral forces in the
actuator.
Gear
Additional pair
bearing Housing
Pinion Gear bearing
F Stinger rod
Piezoelectric
actuators
Housing
Figure 3.10. Active shaft transverse vibration control (concept 4).
3.4.5. Summary
Based on the work in Reference [12], the most feasible active vibration control concept
for suppressing the housing response is the active shaft transverse vibration control one. For this
proposed actuation setup, the housing response due to actuation force can be easily calculated
using the above system model. The result of this analysis is shown in Figure 3.11. This figure
can be combined with Figure 3.6 to determine the force requirement for the actuator [12]. The
42
force requirement for the actuator at each frequency can be determined by dividing the curve in
Figure 3.6 by the curve in Figure 3.8 at each frequency point. After having the force
requirement, the suitable model of actuator, which can provide enough force to suppress the
housing vibration of gearbox system, can be determined for the experimental setup. Actually,
the force requirement is not the only requirement. The electrical power requirement must be
considered as well as the force requirement in next subsection. Note that these results are for TE
excitation with amplitude of 30m. In fact Figure 3.11 is the amplitude response function of the
-2
10
Amplitude (m/s )
2
-3
10
When the pizeo-actuator model, gear system model and actuation are determined, it is
necessary to select proper actuator and its power amplifier for this gear vibration control system
by estimating the electrical power requirements of the system. The structure of gear system can
43
be treated as an external stiffness as described in Chapter 2. Note that this external stiffness is
It has been showed that a piezoelectric stack actuator can be modeled as a spring element
of stiffness Kp when the working frequency of the actuator is far less than the first resonant
frequency of the actuator [40, 42, 43]. Thus the mechanical impedance of the piezoelectric stack
is Z p = K p / i . The block force Fb, defined as the maximum output force when the actuator is
acting on a rigid structure with infinite impedance, is shown in Equation (2.3). If the actuator is
acting against a flexible structure, such as the bearing housing, the maximum possible output
k ext Z ext
Fr = Fb = Fb , (3.13)
k ext + k p Z ext + Z p
where Zext is the external driving-point mechanical impedance, which can be modeled
analytically or measured experimentally using a force hammer. Notice that there might exists a
phase difference between impedances Zext and Zp. If these two impedances are in phase, then the
output force Fr will be less than block force Fb. Otherwise, Fr might be large than Fb depending
the relative phase between Zext and Zp. One advantage of the above analysis is that it doesnt
need to consider the existence of the actuator when estimating the required active force, because
Assumed the gear system is linear and the loaded gear transmission error will not change
after the actuator is installed, it is relatively easy to determine the required active force for
completely canceling the gear housing vibration with the absence of the actuator. Using the
superposition principle, the nominal required active force Fn can be directly estimated by
Z ph
Fn = Ah , (3.14)
i
44
where Ah is the gear housing acceleration and Zph is the cross-point impedance from the actuator
point to the gear housing measurement point. Notice that Zph can be directly measured from the
cross-point frequency response function (FRF) or estimated by the above FE model. Notice that
maximum possible output force Fr must be larger or equal to nominal required active force Fn.
Combining equations (2.3, 3.13-3.14), it is easy to determine the required driving voltage for a
On the other hand, the piezoelectric stack actuator can also be modeled as an electrical
capacitor. Due to the inherent electrical-mechanical coupling, the dynamic electrical admittance
Y is given by
I k 2 k ext k 2 Z ext
Y= = iC 0 [1 ] = i C 0 [1 ], (3.15)
V k ext + k p Z ext + Z p
where k2 is the piezoelectric coupling coefficient, which is assumed to be 0.3, C0 is the zero-
stress electrical capacitance of the actuator, and I is the applied current [14]. It is assume here
that C0 is insensitive to temperature, in spite of the heat generated as the stack expands and
contracts. Hence, given the required driving voltage V, the operating current I can be calculated
using equation (3.15). Notice that this operating current I can not be large than the peak current
2.1 and the estimated driving-point FRF and cross-point FRF at actuator installation location, the
resultant required current and voltage are shown in Figure 3.12 and 3.13. From the estimation
results shown in Figure 3.12 and 3.13, mesh frequency between 200 Hz and 400 Hz will be
control frequency range of interest, because in this area, the required current is less than 100mA,
90
80
70
Required Voltage (V)
60
50
40
30
20
10
0
0 200 400 600 800 1000
Frequency (Hz)
250
200
Required Current (mA)
150
100
50
0
0 200 400 600 800 1000
Frequency (Hz)
The adaptive controller is one of the most important components of the active control
system. This component can be implemented in a DSP board. It calculates the control force
(actuator input) from the dynamic signals picked by the sensors (reference and error sensors)
provide sufficient reduction to achieve the desired response under certain conditions. In this
section, the proposed control structure and its corresponding filtered-x LMS (FXLMS) algorithm
will be given. Also, a frequency estimation technique is introduced and used to construct the
In the applications of active noise and vibration control [66], two major types of control
methods have been widely used: (i) feedforward control and (ii) feedback control. Some
researchers have also combined these two types of control methods. The first one that is the
feedforward control method requires both reference and error signals, and can be used to reduce
broadband error signal as well as narrowband ones. The measured reference signal that is
correlated to the primary disturbance (error signal without the effect of control input) is used to
derive the control input. The correlation between the reference and error signals due to the
primary excitation source is critical to the reduction in the performance of the controller. If the
correlation between the reference and error signals is perfect, it is theoretically possible to force
46
47
the error signal to become zero, which can be a very attractive feature. However, if the
correlation is only partial, the proposed active vibration control system can only suppress the
primary response frequency that is correlated with the reference signal. Due to its inherently
stable and robust characteristics, the proposed approach has in fact been used in many practical
applications.
The second method mentioned earlier is the feedback control that requires only the error
posing difficulty in obtaining suitable references and active control systems characterized by
narrowband disturbances, typically employ the feedback control approach in order to avoid the
problems associated with obtaining a good reference signal for use in the feedforward
configuration. It is well known that a certain threshold level of vibration reduction can be
achieved within a limited bandwidth and the smaller the error signal is driven to, the higher the
control gain becomes. Hence, the less stable the system becomes as well.
be attained using a feedforward control method only when a clean reference signal is available.
However, for the gearbox system, it is very difficult to acquire a suitable set of reference signal
for the feedforward FXLMS scheme. Accordingly, a proposed feedback controller will be used
in this study to reduce the gearbox housing response as shown in Figure 4.1. Unlike the standard
feedback controller, this one uses a signal generator equipped with frequency estimation feature
that synthesizes the reference signal. This effectively converts the feedback scheme into a
feedforward type. The resulting scheme is more stable than the standard feedback structure.
48
Gearbox system
TE x(n) HhTE(z)
+
Hhr(z) +
error input
Adaptive Controller
output
Signal generator with
reference frequency estimation
input
Figure 4.1. Proposed active vibration control system applied to a gearbox plant.
In Figure 4.1, HhTE(z) and Hhr(z) represent the real transfer functions of the primary and
secondary paths, respectively. The primary path relates the primary excitation source to the error
sensor position where vibration is being controlled. On the other hand, the secondary path
connects the actuation force to the error sensor. In this analysis, the error sensor is mounted on
the gear housing as described in the earlier chapter. Thus, the housing vibration is the main
There exist many algorithms in adaptive system. In this section, two algorithms
Although LMS algorithm is more restricted in its use than some other algorithms because
of using a special estimate of gradient is valid for the adaptive linear combiner shown as Figure
49
4.2, its still important for its simplicity, ease of computation and not requiring off-line gradient
Input
x0k w0k dk
w1k Error
x1k k
w1k
x1k
(a)
Input -1 -1 -1
xk z z z
Error
k
(b)
Figure 4.2. The adaptive linear combiner: (a) general form (b) transversal filter
For the controller shown in Figure 4.2, the error signal can be expressed as
k = d k X Tk Wk (4.1)
where dk is the desired signal which will be cancelled during adaptive process, X k is the vector
of input samples in either of two configurations in Figure 4.2, Wk is the weights vector. In LMS
algorithm, the gradient of performance surface, which expresses the expectation of square of
error signal = E [ k2 ] for different weights of control filter, can be derived as,
50
k2 k
w w
0
= M = 2 M = 2 X
0
k 2 k k k (4.2)
k k
wL wL
With this simple estimate of gradient, the LMS algorithm to adapt the weights value is derived
as,
= W + 2 X
Wk +1 = Wk (4.3)
k k k k
The above standard LMS algorithm has been widely used to adaptively update the
weights of the FIR filter in numerous applications. However, in many active noise and vibration
control applications, the standard LMS algorithm must be modified to improve stability. The
path between the adaptive controller output to the housing response point (error signal) is
generally known as the secondary path in contrast to the primary path that relates the external
excitation to the same response point. The existence of this secondary path transfer function in
the closed-loop system may cause the standard LMS algorithm used for active vibration control
applications to become unstable. This is because the error signal is not properly aligned in time
with the reference signal [19]. To address this problem, many solutions to compensate for the
effect of the secondary path have been proposed in the past [67-69]. One of the most effective
schemes is the well-known FXLMS algorithm as shown in Figure 4.3. The standard FXLMS
d(n) + yh(n)
x(n) HhTE(z)
+
r(n) y(n)
W(z) Hhr(z)
Shr(z)
LMS
r(n)
Controller
Description Algorithm
Initialization W (0) = 0
Update y ( n) = W ( n) r ( n)
(at each iteration) X (n) = [r (n) r (n 1) L r (n L + 1)]
X ' (n) = [r ' (n) r ' (n 1) L r ' (n L + 1)]
r ' (n) = S hr (n) r (n)
W (n + 1) = W (n) + 1 X ' (n) y h (n)
The FXLMS algorithm is well studied and has been used in many active vibration control
applications. In fact, both of its Wiener and non-Wiener behaviors have been studied by many
researchers [70-72]. In Reference [71], Bermudez and Bershad used the orthogonal subspace
approach to analyze the non-Weiner solution of filtered LMS algorithm for noiseless reference
sinewave signal. In their research, the proposed FLMS algorithm is shown in Figure 4.4.
52
d(n) + e(n)
y (n)
X(n) y(n)
W(z) S2(z)
S1(z)
e(n)
LMS S3(z)
X (n)
The difference between their filtered LMS algorithm with the proposed FXLMS
algorithm for gearbox vibration control is S3(z). Their non-Weiner solution for FLMS algorithm
is
Px Pd
wi (n) = Re{( 1 + 2 ) exp[ j[( x + d )n + 2 d i x ]]
M2 , (4.4)
+ (( 1* *2 ) exp[ j[( x d )n 2 d i x ]]}
where i=1,2,N (N is the length of adaptive FIR filter) is the weights index, Re{ } stands for the
= M 1 M 2 M 3 , (4.5)
T = e j ( + 2 ) d 2 cos( x )e j ( +1) d + e j d
. (4.8)
+ NPx cos[(1 + 1 ) x ]e j d NPx cos( 1 x )
Here, Px and Pd are the powers of the sinusoidal signal x(n) and d(n), respectively. Furthermore,
x and d are the digital frequencies of x(n) and d(n), respectively. The transfer functions S1(z),
53
S2(z) and S3(z) can be expressed by amplitudes M1, M2, M3, and phase delay 1, 2, 3 (where
=2+3), respectively, at frequency x or d. From Equation (4.4), it can be seen that weights
will oscillate at frequencies x+d and x-d if x? d. Only if x=d, the weights will converge
to fixed values.
Also, they gave the limitation of learning rate , which is the minimum value of
c
sin( ) + sin[( + 1) ] = 0, 2k x , (4.9)
b
c
cos( ) + cos[( + 1) ]
b 1
= , (4.10)
a + 2 cos b
From the discussion on filtered-x LMS algorithm in section 4.2, one know that the
weights of controller will oscillate at frequencies x+d and x-d if x? d. Only if x=d, the
weights will converge to a set of fixed values. To mimic the cases where only one harmonic is
being controlled with other uncorrelated harmonics in the unwanted vibration signals present,
one can assume that there are two harmonics 1 and 2 in unwanted signal d(n), and one
harmonic 1 in reference signal r(n). Thus, one can expect that the weights will oscillate at
frequencies 1+2 and 1-2 with a mean value because there is one frequency value contained
in d(n), which is at the same as the frequency of r(n). These weights can be expressed as
where i=0,1,N-1 (N is the length of adaptive FIR filter), and n is time index. The output of
N 1
y ( n ) = W ( n ) r ( n ) = wi ( n ) X i ( n ) , (4.12)
i =0
separately. Note that the first weighting value is constant. Thus, the multiplier of the first term
and the reference signal is a sine wave with frequency 1. Based on the discussion in section
4.2, one can see that the amplitude a1i is independent of index i. That means all a1i are the same.
Similarly, all a2i are also the same. But their phases depend on index i ( i = i = i1 ). The
a
i =0
1i ( e j [(1 +2 ) n +i ] + e j [(1 +2 ) n +i ] ) b ( e j ( n i )1 + e j ( n i )1 )
N 1 N 1
= c1 ( e j ( 2 + 21 ) n e j (i i1 ) + e j (2 + 21 ) n e j (i i1 )
i =0 i =0
N 1 N 1
+ e jn2 e j (i +i1 ) + e jn 2 e j (i +i1 ) ) (4.14)
i =0 i =0
N 1 N 1 N 1 N 1
= c1 ( e j ( 2 + 21 ) n e j ( 2i1 ) + e j ( 2 + 21 ) n e j ( 2i1 ) + e jn 2 1 + e jn2 1)
i =0 i =0 i =0 i =0
j 2 N1 j 2 N1
1 e 1 e N ( N 1) jn 2
= c1 e j ( 2 + 21 ) n j 21
+ e j ( 2 + 21 ) n j 21
+ (e + e jn2 )
1 e 1 e 2
This term will generate two frequency components, 21+2 and 2 where 2 N1 = m . If
2 N1 m , only one frequency component with frequency 2 will be generated by the above
term. Similarly, the multiplier of third term and reference signal can be derived as
55
a
i =0
2i ( e j [(1 2 ) n + i ] + e j [(1 2 ) n + i ] ) b ( e j ( n i )1 + e j ( n i )1 )
N 1 N 1
= c 2 ( e j ( 21 2 ) n e j (i i1 ) + e j ( 21 2 ) n e j (i i1 )
i =0 i =0
N 1 N 1
+ e jn 2 e j (i +i1 ) + e jn 2 e j (i +i1 ) ) (4.16)
i =0 i =0
N 1 N 1 N 1 N 1
= c 2 ( e j ( 21 2 ) n e j ( 2i1 ) + e j ( 21 2 ) n e j ( 2i1 ) + e jn 2 1 + e jn2 1)
i =0 i =0 i =0 i =0
j 2 N1 j 2 N1
1 e 1 e N ( N 1) jn 2
= c 2 e j ( 21 2 ) n j 21
+ e j ( 21 2 ) n j 21
+ (e + e jn2 )
1 e 1 e 2
Again, this term will also generate two frequency components, 21-2 and 2 where
generated by above term. Hence, the controller output, y(n), includes 4 harmonics given by 1,
Figure 4.5 shows the frequency domain of first element of FIR controller. Figure 4.6
shows frequency domain of output of the controller. In this case, there are two frequencies that
exist in the unwanted signal d(n), namely 128Hz and 512Hz. The reference signal is a sine wave
with 128Hz, the sampling rate is 4096Hz, and the length of FIR controller is 48, which satisfies
the relation 2 N1 = m . Hence, the weights of FIR controller should include the mean value
and the two frequency components, 384Hz and 640Hz. And, in the output of controller, it should
contain two frequency components, 128Hz and 512Hz as shown in Figures 4.5 and 4.6. For the
case of two frequencies, 100Hz and 200Hz, in y(n), and reference signal at 100Hz, the results are
given in Figures 4.7 and 4.8. Furthermore, the length of FIR controller is 10. That implies that
56
the weights of FIR controller should include the mean value and the two frequency components,
100Hz and 300Hz. Furthermore, the output of controller should contain the three frequency
1
10
Mean Value
0
10
-1
10
384Hz
Magnitude
640Hz
-2
10
-3
10
-4
10
-5
10
0 500 1000 1500
Frequency (Hz)
Figure 4.5. Frequency domain of first element of FIR controller weights
4
10
128Hz
3
10
2
10
Magnitude
512Hz
1
10
0
10
-1
10
-2
10
0 500 1000 1500
Frequency (Hz)
Figure 4.6. Frequency domain of output of controller
57
0.6
mean value
0.5
0.4
Magnitude
0.3
0.2
100Hz
0.1 300Hz
0
0 500 1000 1500
Frequency (Hz)
1.5
100Hz
1
Magnitude
0.5
200Hz
400Hz
0
0 500 1000 1500
Frequency (Hz)
As stated above, the FXLMS algorithm requires knowledge of the transfer function of the
secondary path. If the system is time-invariant but unknown, an offline modeling technique can
be used to estimate it during the initial stage. After the initial stage, the estimated model of the
secondary path is essentially fixed during the operation of adaptive controller. However, the
offline technique has a disadvantage of not tracking the changes in the secondary path. Because
the transfer function of the secondary path may be continuously changing in a real system, the
online modeling technique is desirable. It is well known that as long as the phase modeling error
is within ? /2, the adaptive noise canceler will be stable [73]. For the case when the reference
signal is narrowband but the disturbance is broadband, numerical results show that the modeling
error with a phase error of 40? hardly affects the convergence speed of the algorithm [19].
Hence, in real life application, the secondary path is modeled in the offline mode. During the
operation of the controller, if there is significant change in the transfer function of the secondary
path tested, an online modeling technique is needed. The resultant algorithm is examined here to
determine its suitability in performing active vibration control of geared rotor systems.
Figure 4.9 is an illustration of the FXLMS algorithm with the online secondary path
modeling feature in Figure 4.1. The offline modeling technique used here is similar except that
the controlling filter does not update in real time, and the gearbox system is not running, i.e., y(n)
and x(n) are zeros. This online identification is defined as an additive white noise type since it
injects a small amount of broadband power into the system [70]. In this figure, HhTE(z) and
Hhr(z) represent the real transfer functions of the primary and secondary paths, respectively.
Furthermore, W(z) is the adaptive controller whose weights are updated by the FXLMS
feeding a white noise u(n) unrelated to y(n) into the actuator, and updated using the LMS
algorithm. Since the interest here is to control the dynamic response due to transmission error
excitation e(n), a frequency estimator scheme is employed to extract the required fundamental
mesh frequency information from the measured vibration signal and at the same time generate a
reference signal r(n) that is correlated with the primary excitation source e(n).
HhTE(z) d(n) +
x(n)
yh(n)
+
Controller
Figure 4.9. The structure of the proposed FXLMS algorithm with online secondary path
modeling feature.
Frequency estimation is a critical component of the proposed control system. From the
pervious sections, one can see that the control result depends on the reference signal of the
FXLMS algorithm. In this research work, the reference signal is generated through a frequency
60
estimator and sinewave generator. Thus, the performance of the frequency estimator will
In applying the filtered-x LMS algorithm, the reference signal is one of the most
important parameters needed. For the gearbox system investigated here, the aim of controller is
to suppress a set of multi-tone vibration and ultimately the unwanted gear whine signals. The
dominant multi-tone frequency components in the housing vibration response spectrum typically
contain the fundamental mesh frequency and its higher order harmonics. Due to the
configuration of the gearbox dynamic system, it is quite hard to obtain the reference signal
accurately. Hence, the inaccuracy of the estimated frequency may affect stability and vibration
reduction performance. To deal with this problem, numerous frequency estimation approaches
have been previously developed. These include methods such as spectral estimation, adaptive
containing multiple sinusoids is essential. Numerous techniques currently exist. One of them
uses a spectral estimation scheme. In 1994, Gough [74] presented an iterative spectrum
estimation technique based on the discrete Fast Fourier Transform (FFT) algorithm. In another
study, Kay and Shaw [75] proposed a principal component autoregressive spectral estimation
technique that avoids the computation of the eigenvalues and eigenvectors by filtering them out.
Similarly, to avoid the cost of computing the eigenvectors, Karhunen and Joutsensalo [76]
proposed an efficient Fourier transform-based approach for approximating the signal subspace
used to estimate the sinusoidal frequency of the signal. In all of these techniques cited, the
frequencies are found by locating the maximum or minimum values of the computed spectrum,
61
which usually incur high computational cost even though they produce fairly high precision
results. Due to the high computational cost needed for these techniques, they are basically
An alternate technique developed for frequency estimation is the adaptive notch filter
(ANF), which is inherently stable, fast and efficiency. Classically, the ANF technique is
parameterized by the polynomial coefficients of its transfer function. These coefficients are a
function of the notch frequencies with zero or nearly zero gain. In this technique, the frequencies
are computed from the estimated transfer function coefficients. It has been a popular method for
estimating the fundamental frequency of a dynamic signal, and well studied by many researchers
in the last several decades. In 1991 Gegalia [77] proposed an improved lattice-based adaptive
infinite impulse response (IIR) notch filter. The proposed algorithm possesses good convergence
and an improved tracking ability compared to earlier designs. More recently, Li [78] proposed a
cascaded notch filter, which is purely parameterized by the conventional notch filter. This makes
the algorithm simpler, more efficient, and more robust. Kim and Park [66] applied the adaptive
notch filter to an active multi-tone noise control problem. In their approach, a reference signal
generator using frequency estimator is designed for the filtered-x LMS algorithm.
There is yet another class of frequency estimator, which uses the concept of the phase-
locked loop. In 1997 Bodson and Douglas [79] presented an algorithm in continuous time
domain form. The new formulation is shown to produce a good local convergence property.
This property is very helpful for some geared rotor system with an almost stationary rotation
speed of the transmission such as in helicopter systems. For this class of system, numerous peaks
exist in the spectrum. Having a good local convergence property eases tracking and locking
fundamental frequency. However, it needs an initial rough estimation of the frequency to initiate
62
the calculation process. Later, Bodson [80] also implemented the algorithm in discrete time
domain form. He showed that the algorithm could be used to effectively suppress the sinusoidal
signal in the presence of high noise through simulation results. From his results, the estimation
error is not only determined by the signal noise present in the sinusoidal signal, but also decided
by the transfer function of the compensator. Therefore, a good choice of compensators will give
a small estimation error. This does help in the design of the algorithm. This particular algorithm
Figure 4.1 shows the overall control scheme proposed for the gearbox plant of interest.
The adaptive filter shown is updated by the filtered-x LMS algorithm. The error input of
FXLMS algorithm is essentially the housing vibration and the reference input is generated by the
proposed frequency estimator. The subsequent output is used to drive the actuator for
suppressing the housing vibration. The primary path is denoted by HhTE that relates the
transmission error excitation to housing vibration, while the secondary path Hhr is from actuator
to housing vibration as well. In the figure, the frequency estimator is a critical part of the
controller.
The proposed frequency estimator is shown in Figure 4.10. This algorithm is similar to
additive G matrix that is used to compensate for the primary path. Here, the G matrix is defined
as the identity matrix. Thus, it is not shown in Figure 4.10. In the schematic shown, yh is the
input signal whose frequency must be estimated, and n is the additive signal noise. The signal r,
which is used as the reference signal for the filtered-x LMS algorithm shown in Figure 4.3, is the
63
output of the estimator. Finally, C(z) is the compensator of the estimator designed to stabilize
+
Normalization
yh(n) +
u(n)
r(n)
2 y1
Shr(z)
?
? cos() C(z)
1 1 y2
?
y(n) z 1
-sin()
From Reference [80], for the proposed frequency estimator, the control law is given by
C1 ( z )
1 z 1 y1
= C ( z ) , (4.14)
2 2 y2
z 1
where C1(z) and C2(z) are the compensators of the two channels, as
C1 ( z ) = g1 , (4.15)
z za
C2 ( z) = g 2 . (4.16)
z zb
In the Equations (4.14)-(4.16), g1 , g 2 , z a , and z b are the design parameters. There are in fact
two simple methods available for the design of these parameters. The first uses
3(1 z d ) 2 z +2
g1 = 1 z d , g 2 = , za = d , z b = 3z d 2 , (4.17)
d1 3
64
where z d is some desirable location to place the closed-loop poles, and d 1 is the roughly
estimated amplitude of the input signal yh. The other option employs
2(1 z d ) z +1
g2 = , za = d , zb = 0 , (4.18)
d1 2
in which the pole of the compensator for the second channel is placed at the origin, thus allowing
one to neglect its effect entirely. Note that the design parameters mentioned above are critical to
the frequency estimator. Poor selections of these parameters can make the system unstable
and/or ineffective.
From the parametric design formula, it can be seen that the parameter is quite sensitive to
the amplitude of the input signal. In real application, the error signal is reduced by the
adaptation of the controller. This will lead to the instability of the frequency estimator to track
the frequency. In order to address this problem, Bodson [81] also proposed a method, which
subtracts the effect of the control signal feeding into the actuator from the error signal. Using
this technique, the amplitude of signal feeding into the frequency estimator remains constant if
the external excitation is steady. This technique can be used if the secondary path is modeled.
This implies that the offline modeling technique should be used first. However, this kind of
method is applicable to the case when the secondary path remains constant. For the time-varying
system, the secondary path must be identified online. Thus, this particular method cannot be
applied. Also, the system response is different for different frequency point. Under this
condition, the parameters must be updated constantly, which is not very convenient. In this
research, the method is improved by dividing the input signal of frequency estimator by its RMS
value. Thus, the signal is almost normalized to a unit amplitude. This improved the
The performance of the frequency estimator is studied next to provide guidance in the
future experimental work. The sensitivity of the performance of the frequency estimator toward
critical design parameters is studied. This includes the initial values of the integrators used in the
algorithm, tracking ability of the algorithm due to the changing frequency signal, and the signal
There are a total of four integrators in the algorithm. Two of these are very sensitive to
the initial values. They are amplitude and frequency integrators. Here, the amplitude integrator
uses its output as the estimated amplitude of input signal. Similarly, the frequency integrator
also uses its output as the estimated frequency of the input signal. The other two integrators
initial values are based on the estimated initial phase of the signal. For sinusoidal signal, the
Figure 4.11 shows the effect of various initial value of frequency integrator on the
performance of frequency estimation. In this case, the true frequency value is 500Hz. From the
figure, it can be seen that the initial guess of 510Hz needs about 0.1 second to converge to the
true value. When the initial guess is at 520Hz, about 0.14 second is needed, and for a 530 Hz
guess value, a much longer time (about 0.24 second) is needed. Hence, the less accurate the
initial frequency value is, the longer it takes to converge to the true value, as expected.
66
550
540
Estimated Frequency (Hz)
530
520
510
500
490
480
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (sec)
Figure 4.11. Effect of initial frequency guess on convergence rates. (Initial guess values
of: , 510Hz; , 520Hz; , 530Hz)
Figure 4.12 shows the effect of the various initial value of the amplitude integrator. Here, the
amplitude of input signal yh is 10. After normalization, the signals amplitude is almost 1. In this
figure, three cases are studied. The initial values of the amplitude integrator are 1 (solid line), 10
(dotted line), and 15 (dash line), respectively. From the figure, it can be seen that the algorithm
cannot converge to the true value when the initial value is 15. In general, the farther away its
initial amplitude guess is from the true amplitude, the worse the result becomes.
67
200
150
Estimated Frequency (Hz)
100
50
-50
-100
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (sec)
Figure 4.12. Comparison of results for different initial amplitude guess. (Initial guess values
of: , 1; , 10; , 15)
In practical applications, the frequency of the excitation source may be changing with
time. In order to further the study of the performance of the proposed frequency estimator,
example cases in which the signals frequency is changing with time are studied and results are
In Figure 4.13, it is shown that the tracking ability of the frequency estimator in the case
of the frequency of signal varies linearly with time from 500Hz to 900Hz within 9 secs. In this
simulation, the effect of measurement error, in the form of white noise, is added into the input
signal. From this figure, it is observed that some initial time period is needed to converge to the
true value. Once the estimator reaches the true value, though, the estimator is then able to track
950
900
800
750
700
650
600
550
0 1 2 3 4 5 6 7 8 9
Time (sec)
Figure 4.13. Result of frequency estimation for linearly time-varying frequency in the presence
of measurement signal noise. The initial frequency guess is 600 Hz and the true start frequency
is 500Hz.
Figure 4.14 shows the tracking ability of frequency estimator for the frequency jump
phenomenon. In this case, the frequency of signal jumps from 100Hz to 150Hz at 1 sec. From
the figure, it can be seen that the proposed frequency estimator shows a good tracking ability for
frequency jump case. For some real gearbox systems, the mesh frequency can be kept almost
constant. From the result, this frequency estimator can track any small change in frequency
values.
69
200
180
140
120
100
80
60
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (sec)
Figure 4.14. Result of frequency estimation for frequency jump phenomenon.
its performance relative to the adaptive notch filter technique in Reference [66] by tracking
simulated signals with two strong mesh harmonics. Two cases are examined with both involving
a steady-state dynamic signal made up of a fundamental mesh frequency at 100 Hz and its
second harmonic at 200 Hz. In the first case, the amplitude of the fundamental mesh frequency
is twice that of the second harmonic, which is characteristic of geared rotor systems. The
frequency identification results of both the proposed and adaptive notch filter techniques are
shown in Figure 4.15. The comparison results show both techniques converge to the true
fundamental frequency. However, the proposed algorithm converges faster than the Kims
adaptive notch filter technique. Also, the transient signals of the proposed frequency estimation
technique are more bounded while the adaptive notch filter signals go through a larger variation.
70
However, the proposed algorithm must rely on an initial guess as described earlier, while the
adaptive notch filter is not so sensitive to the initial value, which is actually easier to implement.
In a second case, the amplitude of the fundamental mesh is assumed to be half of the amplitude
of the second harmonic. The simulation processes are repeated again using both techniques and
results are shown in Figure 4.16. The proposed technique is still able to converge to the
fundamental mesh frequency solution as long as the initial guess is not too far off. On the other
hand, the adaptive notch filter zooms in on the second harmonic and completely missed the
fundamental frequency.
600
500
400
Frequency (Hz)
300
200
100
0
0 0.5 1 1.5 2
Time (sec)
Figure 4.15. Comparison of results predicted by the proposed frequency estimator ( )and
Kims adaptive notch filter technique( ).
71
600
500
400
Frequency (Hz)
300
200
100
0
0 0.5 1 1.5 2
Time (sec)
Figure 4.16. Comparison of results predicted by the proposed frequency estimator( ) and
Kims adaptive notch filter technique( ).
From the above results, it can be seen that when signal noise is present, the estimated
result also contains signal noise. In order to get a more accurate result, a low-pass filter is added
to the algorithm to filter out the signal noise in the estimated result. In the following analysis,
the effect of the filter on the estimated results is examined. In this case, two low-pass
butterworth filters both with 6 orders are used. Their cut-off frequencies are 20Hz and 100Hz,
respectively (Sampling frequency is 20kHz). In Figure 4.17, the result for cut-off frequency of
100Hz is shown. From this figure, residual signal noise can be seen in the upper plot, while the
bottom plot is much cleaner. Also, in Figure 4.18, the bottom plot converges to the true value
(100Hz) better than the ones in Figure 4.17. This is because it has a lower cut-off frequency than
the first case. However, notice that the low-pass filter results do introduce a phase delay in the
72
algorithm. Thus, the response without filter converges in 0.05 sec (upper plot in Figures 4.17
and 4.18). For the 100Hz cut-off frequency of low-pass filter, the response converges in 0.1 sec
(lower plot in Figure 4.17), and for the 20Hz cut-off frequency, it converges in 0.3 sec (lower
plot in Figure 4.18). Therefore, a suitable low-pass filter is required depending on the nature of
200
(a)
150
100
50
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Estimated Frequency (Hz)
200
(b)
150
100
50
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time (sec)
Figure 4.17. Comparison of estimated result for with filter and without filter case. Filter cut-
off frequency is 100Hz: (a): without filter; (b): with filter.
73
(a)
150
100
50
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
200
(b)
150
100
50
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time (sec)
Figure 4.18. Comparison of estimated result for with filter and without filter case. Filter cut-
off frequency is 20Hz: (a): without filter; (b): with filter.
To further test the capability of the proposed frequency estimator technique, a set of
measured vibration data acquired from an actual gearbox setup is used. In this experiment, the
gearbox is continuously ramped up in speed and then ramped down. The estimated result of the
fundamental gear mesh frequency is shown in Figure 4.19 in comparison to the true gear mesh
frequency measured using an optical speed sensor. In general, the results obtained are fairly
good even though it took a brief time for the algorithm to converge. It can be expected that if the
shaft speed is almost constant, the frequency estimator can lock into the gear mesh frequency
much quicker.
74
1000
900
800
700
Frequency (Hz)
600
500
400
300
200
100
0
0 20 40 60 80 100 120
Time (sec)
Figure 4.19. Result of frequency estimation applied to measured gearbox vibration data
compared to the true mesh frequency signal. ( , Estimated frequency; ,True
frequency)
CHAPTER 5
In this section, the simulation study of active gear casing vibration control based on the
proposed algorithm, which is described in the previous chapter is performed. The simulation
results are used to guide the experimental design of the active vibration control system in the
In this subsection, simulation studies are performed on the active vibration control of the
gear pair system. The system model used in simulation is a state-space model whose state space
matrix is generated as described in Chapter 3. Figure 5.1 shows the MATLAB model with
offline secondary path modeling feature used in this simulation study. The model includes one
FIR adaptive filter with 64-order that is updated by the FXLMS algorithm. The Sw block shown
in Figure 5.1 is the FIR filter represented by a vector length of 1000. It represents the offline
identified secondary path. The bandpass filter used is an eight-order Butterworth filter with low
and high cut-off frequencies of 100 Hz and 4 kHz, respectively. The switch shown is used to turn
the ANC on and off. The time is controlled by a clock. The signal compensator is used to
compensate the effect of the control force on the error signal in order to maintain an almost
constant error signal, which is crucial in achieving a well working frequency estimator. The sine
wave generator is used to produce a reference signal for the FXLMS algorithm using the
estimated frequency. It can be easily controlled to generate a suitable signal with desired
75
76
harmonics. The signal and spectrum displays are used to monitor the control results. The
0
Sw
Switch
Constant
Signal
Out In Out
Sw
Frequency Signal
Estimator Compensator
B-FFT
Spectrum
Dispaly
Figure 5.1. MATLAB model with offline secondary path modeling feature.
Figure 5.2 shows the simplest control case applying the proposed FXLMS algorithm with
an offline secondary path identification method. In this simulation, the sinusoidal TE excitation
occurs at 450 Hz with amplitude of 30m. The active controller is switched on at time equals to
3 sec. From the figure, it needs about 0.3 sec to converge. Note that the converging time is
dependent on the learning rate of the FXLMS algorithm. The larger the learning rate of the
FXLMS algorithm, the sooner it converges. However, it does have a limitation as described in
77
Chapter 4. The finite impulse response of the offline identified secondary path is shown in Figure
5.3. Figure 5.4 illustrates the resultant frequency response function. The upper plot is the phase
response and the bottom one is the amplitude response. From this figure, the poles and zeros of
the system can be easily identified. The main pole of the system is around 420 Hz. Here, only the
2.5
1.5
Housing Vibration (m/s )
2
0.5
-0.5
-1
-1.5
-2
-2.5
2.9 2.95 3 3.05 3.1 3.15 3.2 3.25 3.3 3.35 3.4
Time (sec)
Figure 5.2. Control result using FXLMS algorithm with offline secondary path modeling
(450Hz).
78
-4
x 10
4
-2
-4
0 100 200 300 400 500 600 700 800 900 1000
Tap Index (n)
Figure 5.3. The finite impulse response of the offline identified secondary path.
Phase (Degree)
100
0
-100
100 200 300 400 500 600 700 800 900 1000
Amplitude (m/s )
-2
2
10
-3
10
100 200 300 400 500 600 700 800 900 1000
Frequency (Hz)
Figure 5.4. The frequency response of the offline identified secondary path.
79
Figure 5.5 shows the control result using online identification method. It shows the RMS
value of housing vibration. The active controller is switched on at time equals 0.3 sec. The
secondary path modeling begins at time equals 0.3 sec. The TE excitation signal is at 400Hz.
From this figure, it can be seen that a much longer time is needed to achieve convergence. Also,
in Figure 5.5, the amplitude of housing vibration behaves complicated. The reason is clearly
seen in Figure 5.6, which gives the plot of amplitude response functions of filter S (z ) and
illustrates the plot of phase response functions of filter S (z ) . The solid line is the offline
modeling result, which can be thought of as the true response of the system. The dashed line,
dotted line, and dashed-dotted line are online modeling results at time equals 1.2 sec, 1.8 sec, and
5 sec, respectively. From the phase response, we can see that at time equals 1.2 sec and 5 sec,
the phase of the estimated transfer function of the secondary path is around 0 at 400Hz, the true
phase is also around 0 radian from offline result. This is the reason that leads to the decrease in
housing vibration. At 1.8 sec, the phase of the estimated transfer function of the secondary path
is around 2 radians at 400Hz, and the phase error of the estimated filter is larger than /2 radians.
This shows that around 1.8 sec, the error signal increases. Also, from Figure 5.5, it can be seen
that the additive noise remains in the residual signal after 4 sec. From Figures 5.6(a), one
observed that the amplitude response of S (z ) is not good. It can be concluded that phase
response of S (z ) is more critical than amplitude response for the control system. In all of the
above control cases, the reference signal is the same as the TE excitation signal. This is done
8
RMS Value of Error Signal (m/s )
2
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (sec)
-1
-3
10
-4
10
-5
10
-6
10
0 200 400 600 800 1000
Frequency (Hz)
4
Phase of Response of Secondary Path (rad)
-1
-2
-3
-4
0 200 400 600 800 1000
Frequency (Hz)
Figure 5.6. Amplitude and phase responses of estimated filter of secondary path.
( , offline result; , online at 1.2 sec; , online at 1.8 sec; , online at
5 sec).
82
Figure 5.7 examines the case of attempting to control three sinusoidal components, i.e.,
the fundamental one at 200Hz, the second and third harmonics. The solid line is the housing
vibration spectrum without control. The dotted line is the housing vibration spectrum with
control turned on. The amplitudes of these three sinusoidal components in TE excitation signal
are all 30m. Also, a small amount of white noise filtered by a low pass filter with cut-off
frequency of 2 kHz as background noise is fed into the system. The result shows excellent
performance by the controller after convergence. In this case, the three harmonic components in
1
10
0
10
Housing Vibration (m/s2)
-1
10
-2
10
-3
10
Figure 5.7. Control results of response with three sinusoidal components. ( , without
control; , with control)
However, its far known that the converging rates at those different harmonics are
different and associated with the amplitude response of secondary path for this time domain
FXLMS algorithm. But for frequency domain FXLMS algorithm, the converging rate can be
83
adjusted by setting different learning rate value to different frequency component. Here,
simulation results show that one can control the converging rate of each harmonics by setting the
In order to get a suitable amplitude ratio between the various harmonics in the reference
signal, several cases are studied next. Those results are shown in Figure 5.8. In all cases shown
in Figure 5.8, the amplitudes of three harmonics of TE excitation are 30m. In all cases, the
learning rates are the same, and the frequency response function of the error signals shown in
Figure 5.8 are taken at the same time. In plot (a), the amplitudes of the three harmonics in the
reference signals are 2.49, 1, and 4.48. These values are taken from the amplitude response of
the identified secondary path. The amplitude response function of the identified secondary path
at 230Hz, 460Hz, and 690Hz are 5.6? 10-3 m/s2, 1.4 ? 10-2 m/s2, and 3.1 ? 10-3 m/s2, respectively.
The amplitude response of 460Hz is about 2.49 times that of the 230Hz, and 4.48 times that of
the 690Hz. From this plot, it can be seen that all of the harmonics can be reduced by almost the
same amount simultaneously. In plot (b), the amplitudes of the three harmonics in the reference
signals are 1, 1, and 1. From this plot, it can be seen that the first and third harmonics cannot be
reduced by much simultaneously, but the second harmonic frequency response can be reduced
significantly. In plot (c), the amplitude ratios of the three harmonics in the reference signals are
square of the ones in first case. Here, the second harmonic cannot be reduced more than plot (b).
From these three cases, it can be seen that in order to reduce all harmonics simultaneously, the
ratio of the amplitude of all harmonics should be inversely proportional to the amplitude
response of the secondary path. To further demonstrate this effect, it is assumed the secondary
path is a pure delay, which means that the frequency response amplitude of this secondary path
at each harmonic is the same. Moreover, the amplitudes of the three harmonics in the reference
84
signal are set to unity. The control result of this case is shown in plot (d), which gives the same
vibration reduction in all three harmonics. From the control law of the FXLMS algorithm shown
in Table 4.1, the updating of the weights of the control filter depends on multiplication of the
learning rate , filtered reference signal X ' (n) , and error signal y h (n) . Since the learning rate
for each frequency component and the amplitude of each frequency component of the error
signal are set to be same, the convergence rate of each frequency component will depend on the
amplitude of that frequency component in the reference signal that is filtered by the estimated
secondary path. If the amplitude ratio between the various harmonics in the reference signal is
set to be inversely proportional to the amplitude response of the identified secondary path, the
amplitudes of each harmonic in the filtered reference signal must be almost the same amplitude.
Thus, the convergence speed and ultimately the vibration reduction of each harmonic component
turn out to be almost the same. From the above analysis, it can be concluded that the amplitude
ratio of the various harmonics is closely related to the amplitude response of the secondary path.
85
(a) (b)
0 0
10 10
Housing Vibration (m/s2)
-2 -2
10 10
0 200 400 600 800 1000 0 200 400 600 800 1000
Frequency (Hz) Frequency (Hz)
(c) (d)
0 0
10 10
Housing Vibration (m/s2)
-1 -1
10 10
-2 -2
10 10
0 200 400 600 800 1000 0 200 400 600 800 1000
Frequency (Hz) Frequency (Hz)
Figure 5.8. Effect of amplitude ratios of the harmonic components in reference signal on the
control results. ( , without control; , with control): (a) amplitude ratios between
three harmonics is 2.49:1:4.48; (b) amplitude ratios between three harmonics is 1:1:1; (c)
amplitude ratios between three harmonics is square of case (a); (d) amplitude ratios between
three harmonics is 1:1:1, and Hhr(z) is a pure delay.
86
(a)
2
-1
-2
-3
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Time (sec)
550
(b)
500
Estimated Frequency (Hz)
450
400
350
300
250
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Time (sec)
Figure 5.9. Control result of the frequency jump phenomenon jumping from 450Hz to 350Hz:
(a) the time series of the error signal, (b) estimated frequency.
87
Figure 5.9 shows the control result of the frequency jump phenomenon. In this case, the
TE excitation frequency jumps from 450Hz to 350Hz at 0.3 sec. Plot (a) is the time series of the
error signal (housing vibration). The ANC is switched on at time equals 0.1 sec. From plot (a),
it can be seen that it takes about 0.1 sec for the controller to drive the housing vibration to zero
for the 450Hz signal. The controller cannot work well at 350Hz immediately when the
frequency of excitation signal jumps from 450Hz to 350Hz at 0.3 sec. It takes an additional 0.3
sec to converge again. This is because of the delay in the frequency estimator that cannot
converge to 350Hz immediately as seen from plot (b), which shows the estimated frequency.
One can see that it takes almost 0.25sec to converge to the true frequency 350Hz after the
Figure 5.10 shows the control result of a second frequency jump case. In this case, the
TE excitation frequency jumps from 450Hz to 400Hz at 0.3 sec. Plot (a) is the time series of the
error signal. The ANC is also switched on at time equals 0.1 sec. From plot (a), it can be seen
that it takes about 0.1 sec for the controller to attenuate the housing vibration to nearly zero.
When the frequency of the excitation signal jumps from 450Hz to 400Hz at 0.3 sec, the
controller needs to adapt again to the new frequency. Note that it takes almost 0.15 sec to
converge again. This is because the frequency estimator cannot converge to 400Hz immediately.
This can be easily seen from plot (b) depicting the estimation result of frequency estimator.
Here, one can see that it takes almost 0.05 sec to converge to the true frequency of 320Hz after
the jump point. From these two cases, it can be seen that the smaller the frequency change
occurs, the sooner the proposed controller converges after the jump point. Thus, for a relative
3 (a)
-1
-2
-3
-4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time (sec)
550
(b)
Estimated Frequency (Hz)
500
450
400
350
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Time (sec)
Figure 5.10. Control result of the frequency jump phenomenon jumping from 450Hz to 400Hz:
(a) the time series of the error signal, (b) estimated frequency.
89
Figure 5.11 shows the control case in which the excitation signal is based on measured
data from a real gearbox system. In this simulation, the error signal due to TE excitation is
replaced by the measured data. This new error signal will be added into the error signal due to
the actuation force yielding a total error signal as the control error signal of the FXLMS
algorithm. The secondary path (from actuator excitation to housing response) is a measured one,
which is shown in Figure 5.12. In this case, the first two harmonics are attenuated greatly.
-20
-25
Power Spectrum Magnitude (dB)
-30
-35
-40
-45
-50
0 100 200 300 400 500 600 700 800
Frequency (Hz)
0.03
0.01
-0.01
-0.02
-0.03
-0.04
0 20 40 60 80 100 120
Tap Index (n)
Figure 5.12 Measured finite impulse response of the secondary path.
CHAPTER 6
As discussed in previous sections, the proposed active gear control system with
developed algorithm which is enhanced filtered-x LMS algorithm with frequency estimation
feature is tested in simulation study. In this section, the propose control system is implemented
on a power re-circulating gear system. And experimental studies are performed to test effect of
The cross-sectional view illustrating the inner details of the closed-loop, power re-
circulating, gearbox setup is shown in Figure 6.1. And, the real system is shown in Figure 6.2.
Figure 6.2(a) shows the gear system and used equipments for active control study. Figure 6.2(b)
shows detailed information of the target gearbox system. The basic experimental setup consists
of a belt driver, a torque transducer, a speed sensor, a clutch, a flywheel, one slave gearbox and
one test gearbox. The test gearbox, which is being treated with the active vibration control
system, contains a spur gear pair with a 1:1 tooth ratio. Each gear has identically 35 teeth, 25.4
mm (one-inch) face width and 152.4 mm (6-inch) pitch diameter. The pressure angle is about 20
degrees. The torque transducer is used to record the operating load. The speed sensor provides a
rotation pulse train that consists of 60 pulses per shaft revolution for computing mesh
frequencies. To simulate the high power density without the need for a high power driver, the
91
92
gears can be preloaded by adding torque to one of the shafts coupling the two gearboxes. Of
course, this will also create a matching static torque in the other shaft. The flywheel and
couplings are used to isolate vibrations in the slave gearbox from the test gearbox.
Added Needle
Bearing
Belt Driver
Acc Acc4
Pinion Coupling Housing
Motor
Bearing
Speed Sensor
Stinger Rod
Actuator
Figure 6.1 Schematic diagram of the closed-loop, power recirculation gearbox setup equipped
with an active shaft transverse vibration control system.
93
Drive Motor
Actuator Support
Test Gearbox
(a)
5 6
10 11
8
7
3
Accy
Accx
9
4
2
1
(b)
Detail information of the active gearbox vibration control system: (1) test gearbox; (2) slave
gearbox; (3) belt driver; (4) clutch and flywheel; (5) torque transducer; (6) speed sensor; (7)
piezoelectric stack actuator; (8) adjustable preload nut; (9) accelerometer; (10) power amplifier.
20mm
(a) (b)
Figure 6.3 Closed-up view of: (a) actuation setup, and (b) PZT actuator.
In order to monitor the vibration levels and used as error signal to be controlled in the
experiment, one tri-axial accelerometer is attached to the gearbox housing just above the driven
shaft bearing location facing the slave gearbox (labeled as Acc). The x and y axes of the
accelerometer are perpendicular and parallel to the gear mesh line-of-action, respectively. The
angle between the y direction and vertical is adjusted to 20 degrees, which is identical to the
pressure angle. The y-axis signal from the accelerometer, denoted as Accy, is used as the error
signal for the controller. Other than this tri-axial accelerometer, to monitor the vibrations at
other locations on gearbox housing, three similar tri-axial accelerometers are attached to gearbox
housing right above other three support bearings as shown in Figure 6.1. Among these three
accelerometers, the tri-axial accelerometer labeled as Acc2 is at same side housing panel with
Acc. And, other two accelerometers, Acc3 and Acc4, are located at another side panel. The
acquired vibration signals are processed by a PowerPC board located in a host PC. The actuation
method includes a piezoelectric stack actuator, which is shown in Figure 6.3, connected to the
driven shaft via a needle bearing. The actuator is driven by a high voltage amplifier that receives
command signals from the controller. Note that the actuation direction is parallel to the gear
95
mesh line-of-action, such that the active control force is better able to disrupt mesh force
generation and vibration transmissibility. Also, in the mounting of the actuator, a proper preload
level is set to avoid problems associated with bearing clearances as well as gear eccentricity.
Furthermore, to compensate for possible run-out and excessive shaft deflection that can generate
harmful lateral forces to the actuator, a thin stinger rod that is stiff in the longitudinal direction
and relatively flexible in the transverse direction is used. The other end of the force sensor is
Table 6.1 lists all hardware and equipment used in the experimental study.
Table 6.1 List of hardware for the active vibration control system.
As noted above, the proposed control scheme involves a modified filtered-x least mean
square algorithm with frequency estimation. The filtered-x least mean square (FXLMS) control
algorithm, which is an extended version of the LMS algorithm typically used for dynamic
systems with phase delay secondary paths, is well studied and has been widely applied to many
active vibration and noise control applications. This controller requires detailed information
about the secondary path transfer function. For the gearbox system being studied here, the
secondary path transfer function describes the part of the system from the output of the adaptive
controller to the measured housing response (referred to earlier as the error signal). The
controller also requires a clean reference signal that is coherent with the primary disturbance.
The first issue, concerning the need for a secondary path transfer function, can be handled by
applying a system identification method. Because of the periodic nature of the gear vibration
signal, a predictive error filter (PEF) can be used to increase the accuracy of the secondary path
modeling. The PEF filter, which was proposed by Kuo and his coworker [82], reduces the effect
of a periodic primary signal on the results of the secondary path identification. This approach is
illustrated in the overall control model illustrated in Figure 6.4. Although the overall approach
has already been mentioned, the diagram will be described in more detail below.
The coherence between the reference signal and gear transmission error excitation at the
target mesh frequencies greatly affects the achievable vibration attenuation performance. To
ensure an adequate coherence level between the reference and excitation signals, the exact
fundamental gear mesh frequency must be obtained in advance. For most geared rotor systems,
the shaft rotational speeds vary to a certain degree even for steady-state operating condition,
which makes it difficult to obtain the instantaneous shaft rotation speed precisely and thus the
97
gear mesh frequency as well. Furthermore, the estimation error tends to increase at higher
harmonics due to the multiplying nature of the calculation. The estimation errors encountered at
the higher harmonics can deteriorate the vibration reduction performance especially for orders
above the third mesh harmonic, as discovered in the previous study [10]. To alleviate this
problem, a speed sensor is located inside the torque transducer. This sensor has a higher tooth
count (60) than the test gear (35) employed in the experiments. The acquired pulses from the
speed sensor are used in a frequency estimator that uses the phase-lock loop concept [79] to
estimate in real time the instantaneous shaft rotation frequency and subsequently the fundamental
gear mesh frequency and its harmonics. The performance of the proposed frequency estimator in
terms of tracking the signal frequency change was discussed in detail and tested in the previous
sections. Using this proposed technique, the reference signal r(n) at the target mesh frequencies
TE Excitation d(n)
HhTE(z) +
Rotation te(n) yh(n)
Pulse Train Gearbox System +
Figure 6.4. Enhanced FXLMS algorithm with secondary path identification and frequency
estimation modules.
98
In the control schematic diagram shown in Figure 6.4, HhTE(z) represents the transfer
function between the gear transmission error (TE) signal te(n) and the untreated gear housing
vibration response d(n). The secondary path transfer function is given by H2(z), which relates the
controller output to the gearbox housing vibration at the control point. The housing vibration
response along the gear line-of-action, Accy, is used as the error signal yh(n) in the filtered-x
LMS algorithm for adjusting the adaptive controller W(z). The estimated secondary path transfer
function h2 ( z ) is identified by injecting a small amount of white noise v(n) through the actuator.
By using a secondary disturbance that is uncorrelated with the controller output signal u(n), the
secondary path can be updated using the standard LMS algorithm. Note that the overall
performance of the control system was simulated and refined using Matlab/Simulink software
The active control system described above has been implemented on a dSPACE DS1104
PowerPC board. The processing chip is a Motorola PowerPC 603e model that operates at the
speed of 250 MHz. The Simulink real-time workshop software can be used to generate the
required C codes and also to transfer the resultant objective codes to the processor. Although
this step is very convenient, some of the auto-generated C-codes are not efficient enough to allow
for real-time execution of the control algorithm. This problem is more acute when both the
controller and secondary path transfer function require the use of high order FIR models, such as
those larger than 64. In order to improve the computational efficiency of the algorithm, several
C-coded S-functions were programmed manually to implement some of the critical controller
blocks, including the control law updates and the secondary path identification part.
Furthermore, in order to communicate directly with the PowerPC processor, such as for adjusting
critical control parameters, and monitoring and downloading important signals, a graphical user
99
interface control desktop was designed in the Matlab environment using MLib, which is part of
the dSPACE system. The designed UGI of gear active control system is shown in Figure 6.5.
As shown in Figure 6.5, some control parameters of algorithm such as step size, order of FIR
filter, probe signal strength, etc., can be adjusted manually in the designed GUI. The identified
Now that the test system hardware and the active controller have been described in detail,
experimental results associated with measuring the performance of the active vibration control
system will be described in the next sections. The following sections also discuss the range of
experiments performed to evaluate the ability of the proposed active system to control various
After the actuation system is set up, its system dynamics can be varied by comparing the
experimental study and analytical result. The system dynamics analysis including piezoelectric
actuator model is introduced in previous chapters. At that time, the system dynamics is based on
the FEA model. In this section, only measured system transfer functions by modal test are used
for analysis.
In the analysis, the gearbox system is assumed to be a linear time-invariant plant and the
transmission error excitation is the only excitation source to be controlled. To suppress the gear
housing response resulting from the gear mesh excitation, an actuation force, Fr, exerted by the
piezoelectric actuator is applied to the gear shaft using the hardware configuration described
above. The net housing response along the gear mesh line-of-action direction, yh, is
y h = y h ,TE + H hr Fr . (6.1)
Here, yh,TE is the housing vibration due to transmission error excitation and Hhr is the transfer
function relating the actuator force to the housing response, which is the cross-point frequency
response function measured using a standard modal testing technique. In order to completely
suppress the housing vibration, set yh = 0 and solve for the required actuation force as
Fr = y h ,TE / H hr . (6.2)
Notice that the needed active force in equation (6.2) is an ideal control force for a perfectly
modeled actuator and gear system. When the control force is generated by a real actuator, the
dynamics of the actuator must be considered as part of the analysis as well. Furthermore, before
predicting the required electrical requirement for suppressing gear housing mesh response, an
analytical description of the stack actuator is needed. A model developed for those purpose is
now discussed.
101
From the chapter 2, the block force, i.e. the maximum possible output force that the
actuator can generate when it acts on a rigid structure with an infinite impedance, generated by
Fb = d 33 k pV
. (6.3)
The maximum block force is actually equal to kp*lmax , where lmax is the maximum stroke
of actuator. The maximum free stroke of the actuator is equal to d 33Vmax , where Vmax is the
largest operating voltage. When an external load is applied to the actuator, then the maximum
stroke of the actuator will decrease. Thus, the effective output force will also reduce.
When the actuator acts against a complex host structure, such as the bearing supported
shaft structure in our system, the effective force applied by the actuator on the host structure can
be derived as
where Fb is block force, kp is the effective piezo-stack stiffness and kext is the equivalent
stiffness of that host structure. The driving point stiffness can be used as the equivalent stiffness
of the host structure along with equation (6.4) to calculate the applied force. Note that the
equivalent stiffness might be a complex value (i.e. dynamic stiffness), which means that the
displacement response will have a phase difference relative to the force excitation. From the
above equation, it can be seen that the effective force applied on the host structure depends
strongly on the stiffness of the host structure and the effective piezo-stack stiffness. The same
actuator can generate a higher dynamic force when acting on a stiffer host structure. Note that
since kext may vary with frequency, the force applied to the structure may also vary with
frequency, not only with applied voltage. When selecting a proper actuator, the maximum
102
applied force determined using equation (6.4) should be larger than the required control force
determined using equation (6.2). Once the actual vibration level yh,TE , transfer function Hhr and
kext have been identified, equations (6.2) and (6.4) can be used to guide the selection of a suitable
piezoelectric actuator that is typically rated by the operating voltage range and maximum free
stroke.
The transfer function relating the actuator supplied voltage to the actuator force output
F k ext l k ext
Hf = = k p d 33 = k p max
V k p + k ext Vmax k p + k ext
. (6.5)
By using this transfer function, the secondary path transfer function from the active
controller digital to analog converter (D/A) output to the gear housing vibration, can be
synthesized by
H 2 = g * H f H hr H o
, (6.6)
where H hr , H o and g are the measured cross-point frequency response function, the low-
pass filter transfer function between the D/A output and the piezoelectric amplifier, and the gain
As stated in Chapter 2, to determine the required actuator current, one can first determine
k ext
q = C ps [1 k 2 ]V = C pV
k ext + k p
, (6.7)
where k 2 = d 332 / 33
T E
s33 is the piezo coupling coefficient, which is always less than unity,
and Cp is the effective capacitance of the entire piezo-actuator stack. By assuming a harmonic
time dependence, e jt , the electrical current I applied to the actuator can be found to be
103
I = i C p V
, (6.8)
where i is the unit imaginary number and is the driving angular frequency of the voltage.
Since the required electrical current found using equation (6.8) should be less than the maximum
output current of the piezoelectric amplifier, equation (6.8) can be used in selecting an amplifier
after the piezoelectric actuator has been identified. It should be noted that heat generated by
continued operation of the actuator will increase the actuator electrical capacitance, Cp, as well as
the amplifier requirements. As a result, a slightly oversized amplifier should be selected. For
example, if the calculated maximum current value from equation (6.8) is 80 mA, an amplifier
with larger output, such as 100 mA, should be considered for the system.
With the analytical tools described thus far, it is now possible to select components for
the experimental active gearbox vibration system described earlier. The following section will
discuss the components selected for the particular experimental effort as well as provide a
Before the results are analyzed, though, one must first determine the effective stiffness of
the structure. To obtain the effective stiffness of the host structure as noted above, the driving-
point frequency response function is needed. Furthermore, to estimate the required actuation
force, the cross-point frequency response function is needed. These two frequency response
functions can be determined by performing a standard modal test without the actuator connected
to the host structure. The hammer impact location is the point where the actuator connects to the
structure. For the cross point frequency response function measurement, the accelerometer is
placed on the side panel of the gearbox housing right above the support bearing, as denoted by
Acc in Figure 6.1. Based on the measured frequency response functions and the operating
gearbox vibration spectrum, the analyses presented above were applied to select a commercially
104
available piezoelectric stack actuator. Detailed specifications for this actuator are provided in
Table 2.1.
For this specific experimental gearbox system, the housing vibration is determined to be
at most 0.2 m/s2 at the mesh frequency between 550 Hz to 1,400 Hz. To suppress this level of
vibration using the actuator listed in Table 1, the power amplifier should be able to supply about
70 mA and a little less than 60 volts. In fact, the acceleration is actually much lower in some
cases. Based on this requirement, the most suitable amplifier is the E-420.00 model by PI (see
Table 6.2 for detail specifications). This amplifier also matches the stack actuator rating.
As described above, the actual force applied on the structure depends on both the
effective stiffness of the host structure as well as the actuator. Of course, when an actuator is
installed in the gearbox system, additional hardware including a stinger rod and support structure
is also needed. Here, only the stiffness of the stinger rod is considered because the other support
parts are significantly more rigid when compared to the gear structure. Therefore, the effective
k sr k gs
k ext = , (6.9)
k sr + k gs
105
where ksr is the effective stiffness of the stinger rod along its axial direction, which is 43 N/m
for a 0.32 cm (1/8 inch) in diameter and 3.81 cm (1.5 inches) in length stinger rod. The effective
F 2 2
k gs = = = , (6.10)
x &x& / F H rr
There are two basic approaches to finding the secondary path transfer function, H2. The
first approach involves using equations (6.6) and (6.9-10) to calculate the H2 before the actuator
is installed. If the actuator has already been installed, then the alternate control system approach
described earlier is utilized. The secondary path transfer functions obtained using both of these
approaches are shown in Figure 6.6. Overall, the two results are found to be quite close to each
other. The small differences may be due to the fact that the modal test setup contains no actuator
preload, unlike the identification technique using the control system where the installed actuator
has a preload that acts along the gear mesh line-of action direction. Furthermore, when the
modal testing was conducted, the stationary torque in the system is held constant at 74 Nm.
When the system is running during the system identification process, the torque level may be
different and even fluctuate. This change in torque may also have contributed to the differences
To further verify the models, the predicted voltages for powering the actuator can be
compared to measurements taken while the closed-loop control is on. Under real operating
conditions, the predicted voltages for the actuator required can be calculated from equation (6.4),
measured cross-point FRF and the measured gear housing vibration level. The predicted voltage
and measured results for the piezoelectric stack actuator needed are shown in Figure 6.7. The
comparison demonstrates that the predicted values are in general close to the measured ones.
106
Some differences can be seen as expected. From Figure 6.6, one can see that identified H2
response is lower than predicted result below 800 Hz or so. Thus, one can expect that the
experimental voltages for the actuator below 800 Hz are larger than the predicted values. This
effect is evident in Figure 6.7. Similar conclusion for the relation between measured and
predicted values can be obtained for other frequency points. Another reason can be the non-
exactness of the driving point Hrr and cross-point H hr frequency response functions. Those two
frequency response functions, measured using a standard modal testing setup, are obtained while
the geared rotor system is stationary with a certain level of locked torque. The actual system
response at a specific running speed may be different from those of the stationary geared rotor
system, since the overall alignment may be different. In addition, the sensitivity of the driving
point transfer function Hrr to accelerometer location may also affect the predicted force result.
Unfortunately, the electrical current provided by the amplifier cannot be measured due to
hardware limitations. Therefore, equation (6.8) cannot be directly verified. In spite of these
0
10
-2
10
-3
10
-4
10
0 500 1000 1500
Frequency (Hz)
Figure 6.6 Comparison of the frequency response functions of the secondary path for modal test
result combined with piezo-electric model ( ), and identification result applying the
developed control system ( ).
2
10
1
10
Voltage (Volts)
0
10
-1
10
500 600 700 800 900 1000 1100 1200 1300
Frequency (Hz)
Figure 6.7 Comparison of voltages at the actuator input: ( ), predicted; ( ),
measured.
108
After the implementation of the control algorithm and the design of the Matlab-based
GUI control desktop, a series of experiments were performed to verify the effectiveness of the
new control system. In this section, general control results for single harmonic control and
multi-harmonic control will be described at first. Then, detail results discussion will be
The purpose of these experiments was to gauge the performance of the active control
system for a range of mesh harmonic targets and operating conditions in a high power density
gearbox system. Prior to describing the results obtained, there are a few system parameters that
should be defined first. During the experimental study, the FIR length of the controller W(z) can
be chosen from four options: 4, 16, 32 and 64. All of these options can be set in the control
desktop software interface described earlier. The length of the identified secondary path impulse
response h2 ( z) was set to 128. The controller parameter update rate was set to 4,096 Hz since the
control frequency of interest is below 1,200 Hz. All signals fed into the control board were
filtered by a low-pass filter that has a 1,400 Hz cut-off frequency. As described earlier, the
signal Accy along the gear mesh line-of-action (y-axis of accelerometer Acc) was taken as the
error signal. All vibration results shown below are given in decibels (dB) referenced to a 9.8
m/s2 (1g) acceleration amplitude. Also, it may be noted that in these experiments, fluctuations in
the shaft rotational speeds were observed. These speed fluctuations increase the difficulty of
achieving satisfactory active vibration control results. In spite of this difficulty, it will be seen
that the control results are quite reasonable. First the results for controlling a single mesh
harmonic will be presented. Then, the effect of two important control system features on the
109
ability to control the vibrations will be discussed. Finally, the ability of the system to
simultaneously control the response at multiple mesh harmonics will be evaluated and discussed.
In the first series of experiments, active vibration control of a single gear mesh harmonic
up to the 7th order for various operating shaft rotational speeds was studied. The results were
observed to be quite successful for the 2nd harmonic and higher cases. However, for the
fundamental gear mesh harmonic, the proposed active control system did not perform as well.
Because the actuator used is not capable of generating a force large enough to control the
gearbox system dynamics at the fundamental frequency, significant control could not be
achieved at this frequency. Therefore, results for this case are not presented. In the discussion
below, experimental results for controlling the 2nd and higher harmonics are presented.
-35
7 x mesh
Target frequencies
-40
3 x mesh
-45
Acc y (dB)
-50
-55
-60
-65
300 400 500 600 700 800 900 1000
Frequency (Hz)
Figure 6.8 Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft rotational
speed of 200 rpm. (Keys: , control off; , 3rd mesh harmonic control; , 7th
mesh harmonic control)
110
-3
x 10
10
9
Control On
3
2 4 6 8 10 12 14
Time (sec)
Figure 6.9 Time domain error signal Accy amplitude at the 3rd ( ) and 7th ( ) mesh
harmonics for active control with shaft rotational speed of 200 rpm.
0.4
0.35
0.3
Output of Controller (v)
0.25
0.2
0.15
0.1
Control On
0.05
0
2 4 6 8 10 12 14
Time (sec)
Figure 6.10 Time domain response of controller output signal at the 3rd ( ) and ( )
7th mesh harmonics with shaft rotational speed of 200 rpm.
111
Figure 6.8 shows the typical active control results for the 3rd and 7th gear mesh harmonic
control cases with 26.0 N-m (230 in-lbf) of nominal pre-load torque and a 200 rpm shaft
rotational speed. Due to the dynamics of the system, the transmitted torque actually fluctuated
between 23.7 N-m (210 in-lbf) and 28.3 N-m (250 in-lbf). For discussion sake, the mean value of
26.0 N-m (230 in-lbf) is used as a reference. In the figure, the treated spectrum of gear housing
vibration at the 3rd harmonic is shown by the dash-dotted line, while the 7th harmonic case is
given by the dotted line. The comparisons of the controlled versus baseline response functions
reveal about 10 dB of reduction in the gear housing vibration at these two mesh frequencies. The
time series of the amplitudes for both of these harmonics are shown in Figure 6.9. Here, only the
error signals (i.e. housing vibration response along the line-of-action at the denoted frequency)
are shown. From the plots, it is clear that the response amplitudes at the target frequencies show
substantial reduction going from before to after control is activated. For the case of the 3rd mesh
harmonic, the gearbox housing vibration signal at this frequency decreased sharply from about
0.0617m/s2 (0.0063 g) to 0.0392m/s2 (0.004 g), which is about a four dB drop, and reached a
relatively stable magnitude after only about 1 second into the control mode. Similarly, the 7th
mesh harmonic of gearbox housing vibration dropped from about 0.0813m/s2 (0.0083 g) to
0.0333m/s2 (0.0034 g), which is approximately an eight dB reduction. Although not shown,
similar performances were also observed for the other gear mesh harmonics. The associated
controller output voltages for the piezoelectric amplifier for these two mesh harmonics are
shown in Figure 6.10. The required control voltages were found to be about 0.35 and 0.02 volts
for the 3rd and 7th mesh harmonics, respectively. These voltage strengths will later be compared
to levels computed using the secondary path transfer function and the uncontrolled response
200
Phase (deg)
0
-200
0 200 400 600 800 1000 1200
0
10
Magnitude (g/v)
-1
10
-2
10
-3
10
0 200 400 600 800 1000 1200
Frequency (Hz)
Figure 6.11 Identified secondary path transfer functions. (Keys: , 200 rpm; , 300
rpm)
Figure 6.11 shows the identified secondary path transfer function, which describes the
system from the controller output to the gear housing vibration. The two identified results are
for 200 rpm and 300 rpm shaft speeds, respectively. From the results, the identified path
functions for the two speeds are seen to be quite close to each other. The similarity occurs
because the system setup was essentially the same, the pre-loading torque was set to a similar
value, and the shaft speeds did not vary significantly (i.e. only 100 rpm deviation). Although not
shown, the identified secondary path transfer functions for much higher speeds, such as at 800
rpm, or at substantially different pre-loading torque levels, are different from the functions
shown in Figure 6.11. By using the transfer function shown in the figure for the current test
conditions, the voltages needed to completely suppress the desired harmonics can be estimated as
follows. From the solid line curve in Figure 6.11, the vibration response levels due to 1 volt of
113
input at the 3rd and 7th mesh harmonics for 200 rpm shaft rotational speed are approximately
0.049m/s2 (0.005g) and 3.332m/s2 (0.34g), respectively. By comparing these levels to the peak
amplitudes for the uncontrolled case shown in Figure 6.8, the controller output voltages for the
piezoelectric amplifier needed to suppress the vibration response at the 3rd and 7th harmonics
completely can be computed to be 1.26 volts and 0.024 volts, respectively. The actual controller
output voltage for the 7th mesh harmonic, as shown previously in Figure 6.10, is quite close to
the 0.0024 volts predicted from the secondary path transfer function. However, for the 3rd mesh
harmonic, the estimated and actual controller output voltages are quite different. The reason for
this voltage discrepancy is that the identified secondary path in the low frequency region is not as
accurate as in the high frequency region. Several factors actually contributed to this inaccuracy.
These factors include very low response sensitivity levels in the low frequency range, large
-30
-40
Acc y (dB)
-45
-50
-55
-60
300 400 500 600 700 800 900 1000 1100
Frequency (Hz)
Figure 6.12. Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft
rotational speed of 300 rpm. (Keys: , control off; , 2nd mesh harmonic control;
th
, 4 mesh harmonic control).
114
To further demonstrate the performance of the proposed active control system, Figure
6.12 shows individual control results at the 2nd (dash-dotted line) and 4th (dotted line) gear mesh
harmonics. In these two cases, the preload torque in the system was again set to about 26 N-m
(230 in-lbf) but the shaft rotational speed was increased to a nominal value of 300 rpm. The
results show about 5 and 10 dB of reduction in the gear housing vibrations at the corresponding
2nd and 4th mesh harmonics, respectively. Although this figure demonstrates reductions at the
target harmonics, the impact at the other harmonics due to this control is also worth examining.
From the two cases shown in Figures 6.8 and 6.12, it is evident that the 3rd mesh
harmonic for the 200 rpm case and the 2nd mesh harmonic for the 300 rpm case are each reduced
when they are the control object. However, the response levels at some of the higher harmonics
for those two cases actually increase. These unexpected increases, which are often referred to as
out-of-band overshoot, can be traced back to the wide dynamic range of the secondary path
transfer function shown in Figure 6.11. The magnitudes of the secondary path transfer function
between 800 Hz and 1,100 Hz are much larger than those at frequencies below 500 Hz. Due to
the non-ideal filter response of the controller, even a small controller output component between
800 and 1,100 Hz will be greatly amplified by the secondary path transfer function. This
amplification can lead to large responses in these other frequency ranges. Furthermore, to
control the housing vibration below 500 Hz, a large controller output for the amplifier is
required. This large control signal increases the possibility of the housing vibration
amplification between 800 and 1,100 Hz, as shown by dash-dotted line in Figures 6.8 and 6.12.
On the other hand, when controlling the housing vibration between 800 and 1,100 Hz, a much
smaller controller output is needed. Hence, the housing response below 500 Hz is not as easily
115
excited, and therefore, the unexpected increases do not occur in the results shown by dotted line
Apart from large controller output requirements and the out-of-band overshoot problem,
the lack of accuracy in the identified secondary path transfer function below 500 Hz will
deteriorate the potential active control performance in that frequency range. As a result, while
more than 10 dB of vibration reduction can be easily achieved in the 800 Hz to 1,100 Hz
frequency range, it is more difficult to obtain similar performance below 500 Hz. Work is still in
progress to improve the estimation of the dynamics in the secondary path at these lower
frequencies.
Since the control cases shown in Figures 6.8 and 6.12 were for relatively low shaft
rotational speeds of 200 rpm and 300 rpm, it is insightful to consider higher rotational speeds.
Figure 6.13 shows the active vibration control results at the 2nd mesh harmonic for a shaft
rotational speed of 700rpm and 800 rpm. In this case, the preloading torque of the system
remained at 26 N-m (230 in-lbf). For this case, about 12dB and 8 dB of reduction in the gear
housing vibration at the 2nd mesh harmonic was achieved for 700rpm and 800rpm, respectively.
Note that there is a very slight increase in some of the amplitudes at other frequencies very near
However, no reductions on mesh response can be obtained. Its found that the reason is the
preload nut lost its tight during high speed running. The loosing of tight resulted in the response
of secondary path becoming much lower than normal one. This means that system needs much
higher voltage output from controller to control same level vibrations. To overcome this
problem, the preload nut should be modified to prevent the loosing of tight.
116
-15
-20
(a) Target frequency
-25
-30
Acc y (dB)
-35
-40
-45
-50
-55
200 400 600 800 1000 1200 1400
Frequency (Hz)
-10
Target frequency
-15 (b)
-20
-25
Acc y (dB)
-30
-35
-40
-45
-50
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.13. Frequency domain active control result of Accy (dB re. 9.8m/s2) at the 2nd mesh
harmonic with shaft rotational speed of 700rpm (a) and 800 rpm (b). (Keys: , control off;
, control on)
117
In the previous simulation study, it was concluded that the amplitude ratio of the
reference signal at each gear mesh harmonic should be inversely proportional to the amplitude
response of the secondary path at those harmonics. To experimentally verify this conclusion, the
case of a 250 rpm shaft speed was tested. Here, the control goal is to simultaneously suppress
the vibrations at the 4th, 5th, 6th and 7th gear mesh harmonics. The amplitude ratios for this case
were set to 5.5:1.04:0.5:1, which is inversely proportional to amplitude response of the secondary
path at these four harmonics with the 7th harmonic level assumed to be unity. The results in
Figure 6.14 show that all of these four harmonics are simultaneously reduced by at least 6 dB.
For the 5th mesh harmonic, the reduction is actually closer to 11 dB.
-30
7 x mesh
5 x mesh
-35
6 x mesh
7 dB
-40
11 dB
10 dB
Acc y (dB)
-45 4 x mesh
6 dB
-50
-55
-60
-65
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.14. Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft
rotational speed of 250 rpm. Simultaneous control targets are the 4th thru 7th mesh harmonics
with amplitude ratios of targeted mesh harmonics in reference signal given by 5.5:1.04:0.5:1.
(Keys: , control off; , control on)
118
The active control result when the amplitude ratios were set to 1:1:1:1 is shown in Figure
6.15. While response at the 6th mesh harmonic was reduced by almost 14 dB, no apparent
reduction is seen for the response at the 4th mesh harmonic. Besides demonstrating
simultaneous control, this experiment also confirms the theory reported in Chapter 5. That is, the
amplitude ratio should be inversely proportional to the amplitude response of the secondary path.
-30
-35 7 x mesh
5 x mesh
6 x mesh
-40
9 dB
8 dB
14 dB
4 x mesh
Acc y (dB)
-45
~ 0 dB
-50
-55
-60
-65
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.15. Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft
rotational speed of 250 rpm. Simultaneous control targets are 4th thru 7th mesh harmonics with
the amplitude ratios of targeted mesh harmonics in reference signal given by 1:1:1:1.
(Keys: , control off; , control on)
Figure 6.16 shows another active control result targeting the vibrations at the 3rd thru 8th
mesh harmonics when the gearbox ran at 230 rpm with the same torque used earlier. Using the
same reference amplitude ratio strategy at each harmonic indicated above (the ratios applied
were 26:49:18:5.5:0.8:1 for the 3rd to 8th mesh harmonic), reasonably good active control results
were again achieved at all target harmonics. Specifically, about 9, 7, 4, 6, 4 and 2 dB reductions
119
were obtained for the 3rd to 8th mesh harmonics, respectively. From these results, it is clear that
this active control system can handle up to 6 mesh harmonics simultaneously for a constant
operating speed. Of course, the number of simultaneously controlled harmonics depends on the
system dynamics as well as the actuation ability. It is also of interest to consider the
simultaneous control of multiple harmonics over a range of operating speeds as discussed next.
-35
1 x mesh
Controlled Harmonics
-40
3 x mesh
4 dB
-45
6 dB
7 dB
8 x mesh
4 dB
9 dB
Acc y (dB)
2 dB
-50
-55
-60
-65
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.16. Frequency domain active control result of Accy (dB re. 9.8m/s2) with shaft
rotational speed of 230 rpm. Simultaneous control targets are 3rd thru 8th mesh harmonics.
(Keys: , control off; , control on)
The simultaneous active vibration control of the steady-state housing response at the 4th
thru 7th mesh harmonics were then tested for a speed range of 160 rpm to 260 rpm at 10 rpm
increments. Because the sampling rate was set to 4,096 Hz, the highest frequency that can be
controlled is limited to below 1,200 Hz. As a result, the highest shaft rotational speed is limited
to 260 rpm. Figure 6.17 shows the active vibration control result for the Accy signal.
120
Experimental results for the 4th thru 7th mesh harmonics indicate that an average of 5 dB
reduction with the largest reduction found to be 14 dB. The reductions that can be attained, to
some degree, depend on the secondary path and how large the target harmonic amplitudes are
when compared to background noise levels. Also, as shown in Figure 6.17, the housing
vibration response at most of the mesh harmonics did not improved much at the lower shaft
rotational speeds. In fact, most of the vibration reductions are seen in higher speeds between 230
-30
(a)
4 x mesh
-40
-50
-60
-30
(b)
5 x mesh
-40
-50
Acc y (dB)
-60
-30
(c)
6 x mesh
-40
-50
-60
-30
7 x mesh
-40 (d)
-50
-60
160 180 200 220 240 260
Shaft Speed (RPM)
Figure 6.17. Vibration response (dB re. 9.8m/s2) for simultaneous control of multiple harmonics:
(a) 4th mesh harmonic; (b) 5th mesh harmonic; (c) 6th mesh harmonic; and (d) 7th mesh harmonic.
(Keys: , without control; , with control)
121
In all above cases, the torque loads are set to 26 N-m (230 in-lbf). This value is not so
high. To test the performance of the active control system working under relative high torque
preload condition, in the following set of cases, the torque load is set to 180 N-m (1600 in-lbf).
In Figure 6.18, the vibration spectrums of Accy before the control system is activated for
low torque and high torque preloads are shown. From the figure, the vibration spectrum of
higher torque (solid line) is somehow different from the low torque case (dotted line). Most of
harmonics for large torque case are higher than small torque case. Especially for 3rd harmonic, it
is hardly to be seen for low torque case which can be observed from the solid line in Figure 6.18.
However, its strong in when large torque (dotted line) is applied to system. The difference of
vibration spectrum results from the changes of dynamics of gear system due to different torque
preloads. In the figure, the frequencies of corresponding mesh harmonics for large torque case
are a little low than small torque case. The reason is that the speed of gear system might not be
very stable although the speed is set to 250rpm through control box. There exists small
-20
-30
-40
Acc y (dB)
-50
-60
-70
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.18 Vibration Spectrum before control ( : low torque; : high torque)
Because of the change of system dynamics resulting from different torque load, the
response of secondary path should also be different from each other. In Figure 6.19, the
responses of secondary path for different torque applications are shown. From the result, one
can see that the response of secondary path for larger torque case is lower than small torque case.
This means that system needs more voltage to cancel same level of vibration.
123
200
Phase (deg)
100
-100
-200
0 500 1000 1500
0
10
Magnitude (g/volt)
-1
10
-2
10
-3
10
0 500 1000 1500
Frequency (Hz)
Figure 6.19 Identified secondary path response while system is operating at 250rpm harmonic
( : low torque; : high torque)
Figure 6.20 and 6.21 shows the control results for the 3rd and 5th gear mesh harmonic
when gear system is operating at 250rpm, respectively. From Figure 6.20, one can see that the
target mesh frequency, 3rd harmonic, is reduced from -36.2 dB (0.016 g) to -42.2 dB (0.0078 g).
About 6 dB reduction is obtained. For the case shown in Figure 6.21, the 5th harmonic is
reduced from -29.4 dB (0.034 g) to -39.6dB (0.011 g). About 10.2 dB reduction is achieved. For
other harmonics, such as 6th, 7th harmonics, they are also tried in the experiment. Similar
reductions are obtained for other harmonics except 1st and 2nd harmonics because of systems
ability.
124
-20
-30
3rd
-40
Acc y (dB)
-50
-60
-70
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.20 Frequency domain control result for higher torque case and 3rd harmonic. ( ,
control off; , control on)
-20
5th
-30
-40
Acc y (dB)
-50
-60
-70
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.21 Frequency domain control result for higher torque case and 5th harmonic ( ,
control off; , control on)
125
As we can see in the previous section, when the harmonic is controlled, the higher
harmonic response may be increased by a little bit. The reason might be discussed in section 4.3.
In that section, some simulation works and theoretical formula derivation are performed to prove
The increase in the response at the higher-order mesh harmonics when a lower-order
mesh harmonic is controlled can be attenuated by adding an adaptive linear enhancer to the
active control system. The adaptive linear enhancer (ALE), when used on the error signal before
supplying it to the filtered-x LMS algorithm, can keep the magnitude and phase information of
the error signal at the target frequency while filter out all other frequency components. As a
result, the ALE actually improves the signal-noise-ratio in the error signal for a wide frequency
range. The unexpected increase in gearbox housing vibration can therefore be alleviated as
demonstrated in Figures 6.22-6.24 where the 4th harmonic is the target frequency for a shaft
rotational speed of 250 rpm. Figure 6.22 shows the spectrum of gear housing vibration Accy
before and after the controller is activated. The control results both with and without the ALE
are compared to the results without control. Although the 4th mesh harmonic is the control
target, one can clearly see that the 6th and 7th mesh harmonics are increased by about 2 or 3 dB
when the adaptive linear enhancer is not used. However, the increase at these non-targeted
frequencies is not seen when the adaptive linear enhancer is applied. The reason for this
behavior can be understood by examining Figure 6.23 and 6.24, which shows a comparison of
the spectra of first element of FIR filter weights and controller output spectra with and without
the adaptive linear enhancer. As shown in Figure 6.23, the harmonic components in the filter
weights are reduced a lot when ALE is used. From the discussion at section 4.3, this means that
126
unexpected frequencies in controller output should be reduced, which is shown in Figure 6.24.
In Figure 6.24, there are clearly many harmonics components besides those at the target
frequency (in this case, the 4th mesh harmonic). The unwanted harmonics at the 5th, 6th and 7th
mesh harmonics can potentially induce large vibration response when the resultant housing
vibration is in phase with the original response. This behavior occurs because the response of the
secondary path transfer function is quite large around 800 Hz to 1,200 Hz, as previously shown
in Figure 6.11. With the introduction of the ALE, those unwanted harmonics, especially at the
6th and 7th mesh harmonic, are reduced by more than 13 dB. Thus, even though they are
magnified by the secondary path dynamics, these components are not as dominant in the gear
housing vibration response spectrum when the adaptive linear enhancer is applied.
-30
6 x mesh 7 x mesh
-35
Target frequency
-40
Acc y (dB)
-45 4 x mesh
-50
-55
-60
600 700 800 900 1000 1100 1200
Frequency (Hz)
Figure 6.22. Frequency domain active control result at the 4th mesh harmonic with shaft
rotational speed of 250 rpm. (Keys: , control off; , ALE off and control on;
, ALE on and control on)
127
-3
10
-4
10
Magnitude
-5
10
-6
10
-7
10
0 200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.23 Frequency domain of first element of FIR controller weights (Keys: , ALE off;
, ALE on)
-10
4 x mesh
-20
6 x mesh
7 x mesh
-30
u (dB)
-40
-50
-60
-70
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.24. Frequency spectrum of the controller output (dB re. 1 volt) for shaft rotational
speed of 250 rpm. (Keys: , ALE off; , ALE on)
Although not mentioned earlier, the order of the adaptive linear enhancer may also affect
the performance of controller. This effect is demonstrated in Figure 6.25, where only the
128
spectral content between 700 Hz and 1,000 Hz is plotted to highlight the effect more clearly. In
this figure, three cases are compared: no control, control with 4 order ALE, and control with 64
order ALE. Actually, four different orders (4, 16, 32 and 64 order) are tested in the experiment.
For brevity, only two different orders are shown here because the results for the other orders
follow the same trend as shown in the figure. The results show that the order of the adaptive
linear enhancer should be set as small as possible, since a large order will increase the phase
delay of the error signal and thus reduce the possible vibration reduction performance.
-40
Target frequency 5 x mesh
-45
-50
Acc y (dB)
-55
-60
-65
700 750 800 850 900 950 1000
Frequency (Hz)
Figure 6.25. Frequency domain active control result of Accy at the 5th mesh harmonic with shaft
rotational speed of 250 rpm. (Keys: , control off; , 4 order ALE; , 64 order
ALE)
129
Besides considering the effect of the ALE order size on the active control system
performance, the effect of different step size of the controller adaptation on the active control
results is also of interest. In Figure 6.26, the frequency domain active control results for several
cases, each with a different step size, are shown. In these cases, all of the other parameters were
kept the same, and the gearbox was running at 250 rpm. Only the 6th mesh harmonic, which is at
990 Hz for shaft rotational speed of 250 rpm, was being controlled. The different reductions
achieved for the spectrum of Accy at the 6th mesh harmonic using different step sizes are clearly
shown here. The corresponding time domain response functions for the amplitude of the 6th
harmonic are shown in Figure 6.27. Note that the 6th harmonic amplitude in the original gear
housing vibration response before control was almost the same for all different step sizes. The
adaptive controller was then activated at about the 10 seconds point. All results, even for
different step sizes, converged to a different stable value in less than 1 second from the time
when the controller was activated. However, they did not all converge to the same global
minimum value. As described in section 4.3, the weights of adaptive controller will converge to
non-Weiner solution expressed as Equation (4.4). As shown in Equation (4.4), one knows that
the solution also depends on the step size when the step size is small. Within certain limitations,
a larger step size generally improves the vibration control performance (i.e. smaller residual
housing vibration). The differences in the active control results due to the different step size
used can be understood more in-depth by examining the steady-state FIR controller weights
shown in Figure 6.28. It is noted that the phase of the adaptive controller for different step sizes
is almost the same, but the amplitude is quite different. The controller output signal is a result of
the convolution of the reference signal and the above FIR filter. The controller output signal is
130
itself convoluted with the same secondary path transfer function, and the resultant housing
vibration due to the controller output signal will have to be out-of-phase with the original
that all the above FIR controller weights will be in-phase with each other such that they can all
attenuate the original housing vibration, although with different attenuation levels as shown in
-30
6 x mesh
Target frequency
-40
Acc y (dB)
-50
-60
800 850 900 950 1000 1050 1100 1150 1200
Frequency (Hz)
Figure 6.26. Frequency domain active control result of Accy at the 6th mesh harmonic for
different step sizes with shaft rotational speed of 250 rpm. (Keys: , control off; ,
=0.005; , =0.01; , =0.02; , =0.03)
131
0.025
0.02
= 0.03
0.015
= 0.02
Acc y (g)
0.01 = 0.01
= 0.005
0.005
0
2 4 6 8 10 12 14
Time (sec)
Figure 6.27. Time domain amplitude response of the targeted 6th mesh harmonic of housing
vibration for different step sizes . (Keys: , =0.005; , =0.01; ,
=0.02; , =0.03)
-4
x 10
3
= 0.01 = 0.03
2
1
Mean Weight
-1
-2
= 0.005 = 0.02
-3
5 10 15 20 25 30
Tap Delay (n)
Figure 6.28. Mean value of FIR controller weights for different step sizes . (Keys: ,
=0.005; , =0.01; , =0.02; , =0.03)
132
In specifying the control system parameters, the order of the estimated secondary path
FIR order N and controller FIR order are studied. Since the control frequency of interest is
below 1400 Hz and the controller parameter update rate is set to 4.096 kHz, it is natural to select
a 128-order FIR type controller. The offline identified secondary path (from controller output to
the gear housing vibration) impulse response is shown as dotted line in Figure 6.29. Apparently,
the identified impulse response weights are rather small for tap number greater than 65. The
offline identification process was performed again for N=64 case, as shown as solid line in
Figure 6.29. It is clear that the two identified impulse responses are almost the same before 64-
tap. A little difference is from the different system condition, such as torque preload setting.
The frequency domain responses of the two identified impulse responses are given in Figure
6.30. The two frequency performances are also similar to each other. Especially for phase part,
they are very close for N=64 and N=128 cases for the frequency range from 400Hz to 1400 Hz.
As we discussed in previous chapter, the phase information is very important for the control
system stability. From the identified result of secondary path response, one can conclude that 64
order of FIR filter for modeling secondary path is enough. The effect of large and small FIR
order N on the control performance is shown in Figures 6.31 and 6.32, respectively. Figure 6.31
shows the steady-state frequency domain gear housing vibration parallel to the gear line-of-
action (denoted as Accy) when the control is turned off and on (the step size is 0.001 and the
leakage coefficient is 0.99) for the N=128 case. Figure 6.32 shows a similar performance for
the N=64 case (the step size is 0.001 and the leakage coefficient is 0.99). Comparing these
two figures, it is seen that the vibration reductions at controlled harmonics for the N=128 case
are almost as good as the N=64 case. Since using 64 order FIR filter to model secondary path
133
response will reduce the computation burden of DSP board, it has advantage over using 128
order FIR filter for some cases in which the computation efficiency is highly desired. In this
study, for most of control case, 128 order FIR filter is used because DSP board has enough
computation capability. In this experimental study, up to 6 harmonics for 230rpm shaft speed as
shown in Figure 6.16 are controlled simultaneously. Its found that the DSP will exceed its
ability while controlling more than 7 harmonics simultaneously and using 128 order FIR
secondary path model. If using 64 order secondary path model, more harmonics can be
-3
x 10
5
Estimated Impulse Response of Secondary Path
-1
-2
-3
-4
-5
0 20 40 60 80 100 120
Taps (n)
Figure 6.29 Identified secondary path impulse responses ( N=64; N=128)
134
200
100
Phase (deg)
0
-100
-200
0 500 1000 1500
0.4
Magnitude (g/volt)
128
0.3 64
0.2
0.1
0
0 500 1000 1500
Frequency (Hz)
Figure 6.30 Frequency domain of identified secondary path ( N=64; N=128)
-30
-35
-40
Acc y (dB)
-45
-50
-55
-60
-65
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.31 Frequency domain of control result for order 128 secondary path FIR model
135
-30
-35
-45
-50
-55
-60
-65
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.32 Frequency domain of control result for order 64 secondary path FIR model
136
In the designed control system, only the vibration along gear mesh line-of action at one
location is controlled. Vibrations at other locations on gear housing are also important for gear
whine generation. To monitor the vibration at other locations, three more tri-axial
accelerometers are mounted on the face of gear housing at right above other three supporting
bearings as shown in Figure 6.1. In this section, the vibrations of other locations are discussed.
-30 -30
-35
-35
-40
-40
Acc2 y (dB)
Acc y (dB)
-45
-45
-50
-50
-55
-55
-60
-65 -60
200 400 600 800 1000 1200 1400 200 400 600 800 1000 1200 1400
Frequency (Hz) Frequency (Hz)
-20 -20
-25 -25
-30 -30
-35 -35
Acc3 y (dB)
Acc4 y (dB)
-40 -40
-45 -45
-50 -50
-55 -55
-60 -60
200 400 600 800 1000 1200 1400 200 400 600 800 1000 1200 1400
Frequency (Hz) Frequency (Hz)
Figure 6.33 Frequency domain of control results( ,control off; , control on).
Figure 6.33 show the frequency domain of control result. In this case, the system is
running at 250rpm. And from 4th thru 7th harmonics are controlled simultaneously. The error
signal is the vibration along the gear mesh line-of-action, Accy. Also, the vibrations at other
locations, Acc2y, Acc3y and Acc4y, are monitored. The results are shown in Figure 6.33 (a)-(d),
137
respectively. From the results, one can see that Accy are reduced a lot at the four target
frequencies. This is because Accy is treated as error signal. The aim of controller is to reduce
error signal as possible as it can. Furthermore, Acc2y can also show some reductions, especially
at 5th harmonic. However, for Acc3y and Acc4y, all target frequencies only show a little bit of
reductions. The detailed vibrations levels before and after the actuator is activated are listed in
Table 6.3.
Table 6.3 Vibration levels at four locations along gear mesh line-of-action (dB ref. 1g)
The reason can be seen in Figures 6.34-6.36. Figure 6.34 shows the phase difference
between Accy and Acc2y caused by gear transmission error excitation, i.e., the controller is off.
From that figure, one can see that phase differences of 5th, 6th and 7th harmonics are almost
stable, however, the phase difference of 4th harmonic is changing a lot with time. The reason is
that the 4th harmonic is not so significant. The phase information at that harmonic might not be
accurate because of the mask effect of broadband noise. Figure 6.35 shows the cross-point
transfer functions of two paths which are from actuator force output to housing vibration Accy
and to housing vibration Acc2y. Figure 6.36 shows the cross-point transfer functions of two paths
which are from actuator force output to housing vibration Acc3y and to housing vibration Acc4y.
138
Their detailed phase and magnitude information at four harmonics are listed in Table 6.4. From
the results shown in Figure 6.35 and Table 6.4, one can see that the phases of two transfer
functions are very close. This means that housing vibrations Accy and Acc2y due to the control
force are almost in same phase. However, the total resultant vibration also depends on the gear
transmission error excitation. Even Accy and Acc2y due to control force are in same phase,
Acc2y can not be cancelled if Acc2y due to gear transmission error is out of phase with Accy due
to gear transmission error because Accy should be cancelled finally. From Figure 6.34, the phase
difference of the shown 4 harmonics of housing vibrations Accy and Acc2y due to gear
transmission error excitation are within 90 degrees, except 4th harmonic. In average, the phase
difference of 4th harmonic also can be looked as within 90 degree. That means that if Accy can
be reduced, then Acc2y also can be reduced at these four harmonics. Furthermore, from the
Figure 6.34 and Table 6.5, the phase difference of 5th harmonic is smallest among four
harmonics and magnitudes are almost same. And, from Table 6.3, the vibrations levels for Accy
and Acc2y at 5th harmonic are -33.5 dB and -31.1 dB before the controller is switched on. Thus,
its reasonable to get 9.5 dB reductions at 5th harmonic for Acc2y when about 10.1 dB reductions
for Accy are obtained. This explanation can also explain why the housing vibration at other two
180
90
Phase (deg)
-90
-180
0 5 10 15 20
Time (sec)
Figure 6.34 Phase difference between Accy and Acc2y. ( ), 4th harmonic; ( ), 5th
harmonic; ( ), 6th harmonic; ( ), 7th harmonic.
200
Phase (Deg)
-200
200 400 600 800 1000 1200 1400
Frequency Response (g/N)
-3
10
-4
10
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.35 Cross-point transfer functions of the paths from actuator force output to Accy and
Acc2y
140
200
Phase (Deg)
0
-200
200 400 600 800 1000 1200 1400
Frequency Response (g/N)
-3
10
-4
10
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.36 Cross-point transfer functions of the paths from actuator force output to Acc3y and
Acc4y.
Table 6.4 Detailed Values of four cross-point transfer functions at four harmonics
Accy Acc2y Acc3y Acc4y
n Frequency Phase Amplitude Phase Amplitude Phase Amplitude Phase Amplitude
(Hz) (deg) (10-3g/N) (deg) (10-3g/N) (deg) (10-3g/N) (deg) (10-3g/N)
Table 6.5 Average Phase difference between other locations vibrations and Accy
Furthermore, we know that gear whine is from the vibrations of the whole gear system.
From the above discussion, we also know that the gear housing cannot be treated as a rigid body
in real application. Hence, the vibrations of one location at gear housing cannot represent whole
gear housing vibrations. In this section, experimental studies are performed to decide which
vibration from four bearing locations is more suitable for the sake of gear whine reduction
consideration.
To study the effect on gear whine when the vibrations at other locations are treated as
error signal, several cases are studied and both vibrations and sound pressure level are monitored
in the tests. An example of both gear housing mesh response and gear whine response are
plotted in one plot for comparison as shown in Figure 6.37. In Figure 6.37, the solid line gives
the spectrum of the gear noise radiated from running gear system which is operated at 250rpm.
The dotted-line is the spectrum of gear housing vibration Accy. From the figure, one can see that
the frequencies of some significant harmonics below 600Hz are not matched between gear whine
and gear vibration. The reason is that the monitored gear noise is not only from test gearbox, but
also from slave test gear and motor. To screen the effects of other potential noise sources, the
gear system should be enclosed in the further study. Due to the limitation of experimental setup,
in this study, the sound meter is placed nearby the test gearbox to minimize the effects of other
sources.
142
75 -30
70 5 x mesh -35
Acc y (dB)
60 4 x mesh -45
55 -50
50 -55
45 -60
40 -65
200 400 600 800 1000 1200 1400
Frequency (Hz)
Figure 6.37 Spectrum of gear whine ( ) and gear housing vibration Accy ( ).
Tables 6.6-6.9 list the reductions of gear noise and housing vibrations for different error
signal. Table 6.6 shows the housing vibration reductions at four locations and sound pressure
level reductions when housing vibration Accy is treated as error signal. Tables 6.7-6.9 show the
reductions of vibration and gear whine when Acc2y, Acc3y and Acc4y are treated as error signal,
respectively. From the results shown above, we can see that the phases of cross-point transfer
functions from control force to housing vibration Accy and Acc2y are close, and the phases cross-
point transfer functions from control force to housing vibration Acc3y and Acc4y are close, but
they are different for different housing panels. However, their magnitudes are different. The
control results should be analyzed for particular cases. For example, when Acc3y is the error
signal and 5th harmonic is target frequency, as Table 6.8 shown, Acc3y got 5.6 dB reductions at
5th harmonic. However, Acc4y did not show any reductions. When Acc4y is treated as error
signal, Acc4y got 5.8 dB reductions and Acc3y got 14.8 dB increases. The reasons can be found
143
in Table 6.4 and 6.5. From Table 6.5, we can see that the phase of Acc3y due to TE is almost out
of phase of Acc4y due to TE. And, the phase of Acc3y due to control force is almost in phase
with Acc4y due to control force. Thus, while controlling Acc3y, Acc4y suppose to increase.
However, the magnitude of Acc3y due to unit control force is much larger than the magnitude of
Acc4y due to unit control force. Hence, Acc4y due to control force might not affect total
vibration too much. However, if Acc4y is error signal, the control force will make big impact on
the resultant vibration just as the shown results. Note that the measured transfer functions might
be different from the real transfer functions because system might change when actuator is
installed and preload is added on the system. However, the explanation can also be made in such
way.
From the above control results listed in Table 6.6-6.9, one can see that the best error
signal choice from four bearing location vibrations is Accy. The worst choice is Acc4y. From
Table 6.9, we can see that gear whine at all four harmonics are increased. And, the vibrations at
other locations are increased for from 4th thru 7th harmonics. All treated harmonics in gear whine
signal are increased compared with SPL before actuator is activated. When Accy is the error
signal, most of harmonic vibrations at other locations are decreased. However, for the gear
whine, only 5th and 6th harmonics show some reductions. Of course, Accy might not represent
the global gear housing vibration because multi-modes existed in the system. To find a suitable
point which represents whole gear housing vibrations, a more detailed system model should be
145
analyzed. In that model, gear housing cannot be treated as a rigid body which is used in the
research. Alternatively, to control the global vibration of gear housing and then control gear
whine, several actuators with Multi-Input Multi-Output (MIMO) control system should be used.
CHAPTER 7
7.1 Conclusions
The focus of this dissertation research project is to tackle the gear whine problem through
an active vibration control means. Gear whine is assumed due to the transmission error
excitation at the mesh arising from tooth profile errors, misalignment and tooth deflections. This
frequently high level of tonal noise and vibration can be quite annoying, and also can induce
structural fatigue. Furthermore, rotorcraft interior noise and vibration requirement often limits
the achievable power density and efficiency of the propulsion system. However, current passive
and active techniques have not been able to reduce the gear noise and vibrations to the desired
level. The existence of this gap is the focus of this research work.
To address the desired goal, a finite element model is first developed for a power re-
circulating gear system for use to examine the feasibility of applying an active shaft transverse
vibration control method to tackle the classical gear whine problem. For this particular actuator
setup and gearbox system, a complete set of controller is designed. The proposed control system
is based on the FXLMS algorithm combined with improved frequency estimation technique.
Using the frequency estimation technique, the feedback control structure is naturally converted
into a feedforward structure that is more efficient and robust. From the simulation results, the
active shaft transverse vibration control system is shown to work well in attenuating the gear
vibration response at the fundamental mesh frequency and its harmonics. Also, using the
146
147
flexibility of the sine wave generator, a suitable reference signal can be obtained that enables
most of the harmonics to be controlled simultaneously. The value of the initial guess is critical
to the frequency estimator. The results show the closer initial guesses to the true frequency and
true amplitude of the signal lead to quicker convergence of frequency estimator. The farther
away its initial guesses from the true values, the worse the result becomes. Also, the results
illustrate the proposed frequency estimator can track any acceptable change in frequency values.
Furthermore, when signal noise is present, a low-pass filter can give a much cleaner result.
However, this filter introduces a phase delay in the algorithm. In Chapter 5, the computer
simulation study results show the performance of the FXLMS algorithm with frequency
estimator applied to the gearbox system vibration control. The effects of offline and online
secondary path modeling on the control results are compared in this study. The results show that
offline modeling gives a more satisfied control result. However, the offline modeling can only
amplitude ratios between the various harmonics in the reference signals illustrates that these
ratios depend on the amplitude response function of identified secondary path. All simulation
results show that the proposed active control system is feasible for the reduction of gear vibration
and whine noise. Its results also yielded insights that were applied to guidance the design of
experimental study.
Other than the above algorithm design and implementation, the actuator and power
amplifier selection is also critical to the approach. The actuator and amplifier should be chosen
to have enough capacity to provide enough dynamic force to reduce the gear housing vibration.
In order to select suitable actuator and power amplifier, a PZT stack actuator model under
structural load is proposed. This model combined with measured or predicted frequency
148
response functions (FRF) can be used to calculate the requirement of actuator and amplifier. The
results can be used to select suitable actuator. The experimental work shows good match with
the predicted results. In this study, a simplified linear piezoelectric stack actuator model is
developed for use in selecting the actuator for an active gearbox vibration control system. This
model enables the selection of a suitable piezoelectric stack actuator and its corresponding power
amplifier before the active control system is constructed. The secondary path transfer function
synthesized from the model matches quite well with the directly measured one using the control
systems online identification feature. The proposed methodology will ensure that the selected
actuator and amplifier can meet the requirements of the practical active control system without
the usual costly trial and error process. In the Chapter 6, this methodology is verified through
experimental study.
In Chapter 6, experimental studies are performed for the target power re-circulating
gearbox system which is equipped with the designed active control system. The housing
vibrations along the gear mesh line-of-action were suppressed at gear mesh harmonics for a
range of gear rotational speeds. In some cases, up to 14 dB of reduction in the vibration response
can be achieved. Furthermore, it is shown that the use of adaptive line enhancer with the proper
order can reduce the out-of-band overshoot and improve the system control performance. The
experiments also reveal some limitations in the active control system due to the inherent dynamic
characteristics of the gearbox system. Furthermore, the controller adaptation step size can affect
the active control results significantly. For the cases considered here, a larger step size yields
better reductions. However, the performance gains achieved by increasing the step size do have
an upper limit because a very large step size can cause the controller to become unstable. The
experimental work also reveals similar performance of the designed control system on different
149
system loading condition. Other location vibrations are also monitored and examined. It was
observed that only one actuator and one sensor (Single-Input Single-Output) can not achieve
global gear housing vibration control. And then, the gear whine can not be controlled as error
signal does. To achieve global vibration and gear whine control, Multi-Input Multi-Output
system should be used. In spite of the inadequacies in the prototype system, the results yielded
by the current active vibration control system are quite promising, and the underlying concept is
Even though the active gear vibration control system has been developed and
experimentally verified, there are other fundamental issues that should ultimately be addressed in
the future.
(a) The selection of cost function and error signal is important to whole control
system. Since the gear dynamics at different frequencies and different locations
play an important factor in actuation system design for global control goal, a more
detail gear housing dynamic analysis is needed to guide the selection of more
suitable error signal location and cost function to meet global control goals.
(b) More actuators might be considered for better control the gear housing vibrations.
Currently, only one actuator has been considered. Theoretically, one actuator has
only one degree of freedom to control one vibration response. To achieve gear
housing global vibration control with only one actuator is likely an over-
determine problem. There is no global optimal solution for such case. However,
as it has been shown in Chapter 6, at most frequencies, the four housing vibrations
along the gear line-of-action are not all independent each other. A detail modal
150
analysis of the gear housing might be needed in order to pick up the independent
gear housing vibration modes. After the number of independent vibration modes
within the frequency range of interest has been selected, the same number of
actuators might be needed in order to control the gear housing vibrations globally
(c) Eccentricity and misalignment of gear system reduce the efficiency of actuator
system. For better performance, the gearbox setup should be assembled with care
(d) The actuator mounting method and mounting position are also important to the
mounting position cannot be changed easily and the gear system cannot operate at
much higher speeds. In fact, the load transmitted through actuator may
deteriorate for speeds above 900 rpm. Thus the actuator will fail to generate the
required dynamic force for some cases. Alternate mounting configurations should
(e) The nonlinear effects of gear system and hysteresis of the piezoelectric stack
actuator might need to be considered for cases where the required force is larger
(f) To monitor and study the gear whine only from test gearbox, the test gear housing
should be enclosed in order to screen the effects of other potential noise sources,
such as belt driver and slave gearbox sources, leaving only the test gear housing
noise.
REFERENCES
1. G. CORNEJO and A. K. PAKHIT 1996 Proceedings of the 7th International Power Transmission
and Gearing Conference, San Diego, 195-204, Identification of spiking vibrations in gears
and their effect on the performance of gearbox in high speed applications.
2. J. D. SMITH 1999 Gear Noise and Vibration, New York: Marcel Dekker.
4. D. HOUSER 1986 Power Transmission Design, 27-30, The root of gear noise: Transmission
error.
5. Y.-S. LEE and S. J. ELLIOTT 2001 Journal of Sound and Vibration 242(5), 767-791. Active
position control of a flexible smart beam using internal model control.
6. C. H. PARK and A. BAZ 1999 Journal of Sound and Vibration 227(4), 711-734. Vibration
control of bending modes of plates using active constrained layer damping.
7. S. IKAI, K. OHSAWA, K. NAGAYA and H. KASHIMOTO 2000 Journal of Sound and Vibration
231(2), 393-409. Electromagnetic actuator and stacked piezoelectric sensors for controlling
vibrations of a motor on a flexible structure.
8. C. R. FULLER, S. J. ELLIOTT and P. A. NELSON 1996 Active Control of Vibration. San Diego,
CA: Academic Press Inc.
9. A. KAHRAMAN and R. SINGH 1990 Journal of Sound and Vibration 142(1), 49-75. Non-linear
dynamics of a spur gear pair.
10. B. REBBECHI, C. HOWARD and C. HANSEN 1999 Proceedings of the Active Control of Sound
and Vibration Conference, Fort Lauderdale, 295-304. Active control of gearbox vibration.
11. M. H. CHEN and M. J. BRENNAN 2000 Smart Materials and Structures 9(3), 342-350. Active
control of gear vibration using specially configured sensors and actuators.
151
152
California. Comparison of actuator designs for active vibration control of a gear pair
system.
13. R. MAIER, M. PUCHER and W. GEMBLER 1999 Proceedings of the Active Control of Sound and
Vibration Conference, Fort Lauderdale, 189-198. Helicopter interior noise reduction by
active vibration isolation with smart gearbox struts.
14. T. J. SUTTON, S. J. ELLOITT and M. J. BRENNAN 1997 Journal of Sound and Vibration 205(1),
81-101. Active isolation of multiple structure waves on a helicopter gearbox support strut.
15. I. PELINESCU and B. BALACHANDRAN 2000 Smart Structures and Materials 2000: Smart
Structures and Integrated Systems, 3985, 76-85. Analytical and experimental investigations
into active control of wave transmission through gearbox struts.
16. T. J. SUTTON and S. J. ELLIOTT 1995 Journal of Vibration and Acoustics, Transactions of the
ASME 117(3(A)), 355-362. Active attenuation of periodic vibration in nonlinear systems
using an adaptive harmonic controller.
17. M.H.CHEN 1999 Combining the Active Control of Gear Vibration with Condition
Monitoring, Ph.D. dissertation, University of Southampton.
18. K. B. SCRIBNER, L. A. SIEVERS and A. H. VON FLOTOW 1993 Journal of Sound and Vibration
167(1), 17-40. Active narrow-band vibration isolation of machinery noise from resonant
substructures.
19. S. M. KUO and D. R. MORGAN 1999 Proceedings of the IEEE 87(6), 943-973. Active noise
control: a tutorial review.
20. S. R. HALL and N. M. WERELEY 1993 Journal of Guidance Control and Dynamics 16(4),
793-797. Performance of Higher Harmonic Control Algorithms for Helicopter Vibration
Reduction.
21. A. J. CONNOLLY, M. GREEN, J. F. CHICHARO and R. R. BITMEAD 1995 Proceedings of the IEEE
Conference on Decision and Control 3, 2982-2987. Design of LQG & H infinity controllers
for use in active vibration control & narrow band disturbance rejection.
22. K. J. ASTROM and B. WITTENMARK 1995 Adaptive Control. New York: Addison-Wesley
Publishing Company, Inc.
23. H. T. BANKS, R. C. SMITH and Y. WANG 1996 Smart Material Structures Modeling,
Estimation And Control. John Wiley & Sons.
24. H. N. OZGUVEN and D. R. HOUSER 1988 Journal of Sound and Vibration 121(3), 383-411.
Mathematical models used in gear dynamics - A review.
25. I. CHOPRA 2000 Journal of the American Helicopter Society 45(4), 228-252. Status of
application of smart structures technology to rotorcraft systems.
153
26. C. SHAKERI, M. N. NOORI and Z. HOU 1996 Proceedings of the Fourth Materials Engineering
Conference Washington D.C., 2, 863-876. Smart materials and structures; a review.
27. L. D. JONES and E. GARCIA, 1994, Adaptive Structures and Composite Materials : Analysis
and Application : Presented at ASME 1994 International Mechanical Engineering Congress
and Exposition, Chicago, Illinois, 155-165, Application of self-sensing principle to a
magnetostrictive structure element for vibration suppression.
28. K. S. KANNAN and A. DASGUPTA, 1994, Adaptive Structures and Composite Materials :
Analysis and Application : Presented at ASME 1994 International Mechanical Engineering
Congress and Exposition, Chicago, Illinois, 21-28, Finite element modeling of multi-
functional composites with embedded magnetostrictive devices.
29. F. T. CALKINS, A. B. ZROSTLIK, A. B. FLATAU and J. M. VOGEL 1994 Adaptive Structures and
Composite Materials: Analysis and Application, ASME 1994 International Mechanical
Engineering Congress and Exposition Chicago, Illinois, 267-173. Terfenol-D vibration
control of a rotating shaft.
30. H. T. BANKS, R. C. SMITH and Y. WANG 1996 Smart Material Structures: Modeling,
Estimation and Control, Masson: Wiley.
32. E. F. CRAWLEY and J. D. LUIS 1987 AIAA Journal 25(10), 1373 - 1385. Use of piezoelectric
actuators as elements of intelligent structures.
33. N. W. HAGOOD and A. V. FLOTOW 1991 Journal of Sound and Vibration 146(2), 243 - 268.
Damping of structural vibrations with piezoelectric materials and passive electrical
networks.
34. F. LALANDE 1995 Modeling of the Induced Strain Actuation of Shell Structures Ph.D.
Dissertation, Virginia Polytechnic Institute.
35. C. A. GENTRY, C. GUIGOU and C. R. FULLER 1997 Journal of the Acoustical Society of
America 101(4), 1771-1778. Smart foam for applications in passive-active noise radiation
control.
36. M. YALCINTAS and J. P. COULTER 1994 Adaptive Structures and Composite Materials:
Analysis and Application, ASME 1994 International Mechanical Engineering Congress and
Exposition Chicago, Illinois, 101-112. Analytical modeling of electrorheological material
based adaptive beams.
37. S. B. CHOI, D. W. PARK, M. S. CHO and C. C. CHEONG 1999 Proceedings of SPIE - The
International Society for Optical Engineering, 3668, 487-494. Position control of a cylinder
system using ER valve actuators.
154
38. G. Z. YAO, Y. QIU, G. MENG, T. FANG and Y. B. FAN 1999 Journal of Sound and Vibration
219(1), 175-188. Vibration control of a rotor system by disk type electrorheological damper.
39. S. V. GOPINATHAN, V. V. VARADAN and V. K. VARADAN 2000 Smart Materials and Structures
9(1), 24-48. Review and critique of theories for piezoelectric laminates.
40. E. FLINT, C. LIANG and C. A. ROGERS 1995 Journal of Intelligent Material Systems and
Structures 6(1), 117-124. Electromechanical analysis of piezoelectric stack active member
power consumption.
42. C. LIANG, F. SUN and C. A. ROGERS 1996 Smart Materials and Structures 5(2), 171-186.
Electro-mechanical impedance modeling of active material systems.
43. S. ZHOU, C. LIANG and C. A. ROGERS 1995 Journal of Intelligent Material Systems and
Structures 6(6), 733-743. Integration and design of piezoceramic elements in intelligent
structures.
44. Piezomechanik GMBHP 2000 Piezoelectrical and electrostrictive stack actuator catalog.
45. W. A. TUPLIN 1950 Proceedings of the Institution of Mechanical Engineers 60, 162-167.
Gear tooth stresses at high speed.
46. H. N. OZGUVEN and D. R. HOUSER 1988 Journal of Sound and Vibration 125(1), 71-83.
Dynamic analysis of high speed gears by using loaded static transmission error.
47. M. AMABILI and A. RIVOLA 1997 Mechanical Systems and Signal Processing 11(3), 375-390.
Dynamic analysis of spur gear pairs: steady-state response and stability of the SDOF model
with time-varying meshing damping.
48. M. AMABILI and A. FREGOLENT 1998 Journal of Sound and Vibration 214(2), 339-357. A
method to identify modal parameters and gear errors by vibrations of a spur gear pair.
49. Q. J. ZHANG and J. RESCIGNO 2000 Proceedings of the ASME Power Transmission and
Gearing Conference, Baltimore. Study of transient response of a gear train under pulsating
torque.
50. F. CRICENTI, C. H. LANG and D. PARADISO 2000 Proceedings of the ASME Power
Transmission and Gearing Conference, Baltimore. Development of an analytical model for
the static and dynamic transmission error calculation: validation on a single vehicle gear
vibration test bench.
51. T. C. LIM and J. LI 1999 Journal of Sound and Vibration 219(5), 905-919. Dynamic analysis
of multi-mesh counter-shaft transmission.
52. M. VAISHYA and D. R. HOUSER 2000 Proceedings of the ASME Power Transmission and
Gearing Conference, Baltimore. Modeling and analysis of sliding friction in gear dynamics.
155
53. D. HOCHMANN and D. R. HOUSER 2000 Proceedings of the ASME Power Transmission and
Gearing Conference, Baltimore. Friction forces as a dynamic excitation source in involute
spur and helical gearing.
54. G. STEYER and T. LIM 1992 Proceedings of 9th International Modal Analysis Conference.
System dynamics in quiet gear design.
55. M. G. DONLEY, T. C. LIM and G. C. STEYER 1992 Journal of Passenger Cars, SAE
Transaction 101(6), 958-968. Dynamic analysis of automotive gearing systems.
56. A. KAHRAMAN, H. N. OZGUVEN and D. R. HOUSER 1989 Proceedings of the ASME Power
Transmission and Gearing Conference, 375-382. Dynamic analysis of geared rotors by
finite elements.
57. R. G. PARKER, S. M. VIJAYAKAR and T. IMAJOU 2000 Proceedings of the ASME Power
Transmission and Gearing Conference, Baltimore. Modeling the nonlinear vibration of a
spur gear pair.
58. K. YAKHOU, A. BOURDON and D. PLAY 2000 Proceedings of the ASME Power Transmission
and Gearing Conference, Baltimore. Prediction of automotive gearboxes dynamic behavior.
Part A: experimental validation of a global numerical model.
59. A. BOURDON, K. YAKHOU and D. PLAY 2000 Proceedings of the ASME Power Transmission
and Gearing Conference, Baltimore. Prediction of automotive gearboxes dynamic behavior.
Part B: numerical simulations for determination of new mechanical design trends.
60. P. VELEX, M. MAATAR and J. P. RACLOT 1996 Proceedings of the ASME Power Transmission
and Gearing Conference, 29-38. Some numerical methods for the simulation of geared
transmission dynamic behavior formulation and assessment.
61. P. VELEX and M. MAATAR 1996 Journal of Sound and Vibration 191(5), 629-660. A
mathematical model for analyzing the influence of shape deviations and mounting errors on
gear dynamic behavior.
62. J. P. RACLOT and P. VELEX 1999 Journal of Sound and Vibration 220(5), 861-903. Simulation
of the dynamic behavior of single and multi-stage geared systems with shape deviations and
mounting errors by using a special method.
63. P. VELEX and V. CAHOUET 2000 Proceedings of the ASME Power Transmission and Gearing
Conference, Baltimore. Experimental and numerical investigation on the influence of tooth
friction in spur and helical gear dynamics.
64. S. S. RAO 1999 The Finite Element Method in Engineering. Boston: Butterworth-
Heinemann.
65. T. C. LIM and R. SINGH 1991 Journal of Sound and Vibration 151(1), 31-54. Vibration
transmission through rolling element bearings, Part III: Geared rotor system studies.
156
66. S. KIM and Y. PARK 2001 Journal of Sound and Vibration 241(4), 681-691. On-line
fundamental frequency tracking method for harmonic signal and application to ANC.
67. B. WIDROW and S. D. STEARNS 1985 Adaptive Signal Processing. New Jersey: Prentice-
Hall.
68. D. R. MORGAN 1980 IEEE Transactions on Acoustics, Speech and Signal Processing ASSP-
28, 454-467. An analysis of multiple correlation cancellation loops with a filter in the
auxiliary path.
69. B. WIDROW, D. SHUR and S. SHAFFER 1981 Proceedings of the 15th ASILOMAR Conference
on Circuits, Systems and Computers, 185-195. On adaptive inverse control.
70. S. M. KUO and D. R. MORGAN 1996 Active Noise Control Systems: Algorithms and DSP
Implementation. New York: Wiley.
71. J. C. M. BERMUDEZ and N. J. BERSHAD 1999 IEEE Transactions on Circuits and Systems II:
Analog and Digital Signal Processing 46(8), 1110-1113. Non-Wiener behavior of the
filtered LMS algorithm.
72. N. J. BERSHAD and J. C. M. BERMUDEZ 1998 IEEE Transactions on Signal Processing 46(5),
1298-1313. Sinusoidal interference rejection analysis of an LMS adaptive feedforward
controller with a noisy periodic reference.
73. X. KONG and S. M. KUO 1999 IEEE Transactions on Speech and Audio Processing 7, No. 5,
587-591, Analysis of asymmetric out-of-band overshoot in narrowband active noise control
system.
74. P. T. GOUGH 1994 IEEE Transactions on Signal Processing 42, 1317-1322. A fast spectral
estimation algorithm based on the FFT.
75. S. M. KAY and A. K. SHAW 1988 IEEE Transactions on Acoustics, Speech, and Signal
Processing 36, 95-101. Frequency estimation by principal component AR spectral
estimation method without eigendecomposition.
76. J. T. KARHUNEN and J. JOUTSENSALO 1992 IEEE Transactions on Signal Processing 40(12),
2961-2972. Sinusoidal frequency estimation by signal subspace approximation.
78. G. LI 1997 IEEE Transactions on Signal Processing 45(8), 2001-2009. Stable and efficient
adaptive notch filter for direct frequency estimation.
79. M. BODSON and S. C. DOUGLAS 1997 Automatica 33(12), 2213-2221. Adaptive algorithms for
the rejection of sinusoidal disturbances with unknown frequency.
157
80. M. BODSON 2001 Automatica 37, 1133-1140. Performance of an adaptive algorithm for
sinusoidal disturbance rejection in high noise.
81. M. BODSON, 1998 Proceeding of the ISMA23, Leuven, Belgium, 3, 1279-1286, Algorithms
for feedback active noise control and narrowband disturbances.
82. S. M. KUO and D. VIJAYAN 1997 IEEE Transactions on Speech and Audio Processing 5, 374-
377. A secondary path modeling technique for active noise control systems.