Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Active Vibration Control of A High Speed Rotor Using

PZT Patches on the Shaft Surface


H.-G. HORST

AND

H. P. WOLFEL

Department of Mechanical Engineering, Darmstadt University of Technology, Petersenstr.30, 64287 Darmstadt, Germany
ABSTRACT: This paper describes the development of a structural model of a high speed
rotor for the examination of active vibration control in rotor dynamics. Suppression of lateral
bending vibrations of the elastic shaft is realized by means of surface-bonded piezoceramic
actuator patches on the shaft surface. Models for actuator implementation are derived.
Simulations demonstrate the effectiveness of this approach. To validate the simulation, a rotor
test-rig was built. The characteristics of the system model with implemented actuators are
compared to experimental tests. Both results show good agreement.
Key Words: rotor dynamics, shaft vibrations, finite element model, modal state space model,
PZT actuator, optimal control

the actuators mass distribution. This technique only


allows the reduction of unbalance-induced synchronous
vibrations; non-synchronous or transient vibrations
cannot be suppressed.
Direct active vibration control approaches use actuators to generate lateral forces or displacements counteracting the rotor vibration. These techniques are of course
very similar to the vibration control of nonrotating
structures except that additional phenomena of rotor
dynamics have to be taken into account. Actuation
forces are mostly applied at the rotor bearings either
directly (noncontacting) or via the bearing housings.
For force generation magnetic, piezoelectric or hydraulic physical effects can be employed (Ulbrich, 1998).
Furthermore active tilting pad journal bearings can be
used to actively tune bearing stiffness and damping
characteristics. Active magnetic bearings are applied in
high-speed rotors to reduce friction and mechanical
wear but can also be used to tune stiffness and damping
and to generate contact free lateral forces for vibration
isolation and unbalance compensation (Schweitzer,
1998; Nordmann et al., 2000). The drawbacks of magnetic bearings are the technical effort and the continuous
power consumption to maintain bearing functionality. If
friction and wear of rolling element bearings are not an
issue it may be more straightforward to use conventional bearings with attached actuators. High stiffness
and fast dynamic response are the main advantages
of piezoelectric stack actuators making them suitable
for vibration control purposes. Palazzolo et al. investigated rotor-bearing systems with piezoelectric pushers
acting on the bearing housing (Barrett et al., 1995; Tang
et al., 1995). They developed simulation models and test
rigs demonstrating the effectiveness of their approach.

INTRODUCTION
TRUCTURAL vibrations in rotating machinery are
a limiting factor concerning the productivity in
many important industrial branches such as the machine
tools, textile paper and printing industries. Particularly
lateral bending vibrations of elastic shafts which are in
most cases very lightly damped are problematic: either
it is not possible to pass through critical speeds or
vibrations during operation affect bearing lifetime and
product quality negatively. The rotor vibrations can be
caused by unbalances, process forces or system instabilities. The excitation forces often cannot be further
reduced (e.g. by balancing) so that vibration reduction
or increased system stability is achieved through
additional damping. Damping forces can be introduced
for example by damped bearing supports or squeezefilm dampers in the bearings (Gasch et al., 2002). To
increase productivity and improve product quality in
many cases, the performance of passive damping devices
alone is no longer sufficient. Therefore new solutions
using active control are sought to effectively reduce
vibration amplitudes. Active vibration control techniques for rotor systems can be categorized into two
major groups: direct active vibration control and active
balancing (Zhou and Shi, 2001).
Active balancing aims to eliminate the system eccentricity by using a mass redistribution actuator mounted
on the rotor which is able to change the center of mass.
The controller estimates the system unbalance from
vibration measurements and calculates a signal to adapt

*Author to whom correspondence should be addressed.


E-mail. horst@fmd.tu-darmstadt.de

JOURNAL

OF INTELLIGENT

MATERIAL SYSTEMS

AND

STRUCTURES, Vol. 15September/October 2004

1045-389X/04/9/10 07218 $10.00/0


DOI: 10.1177/1045389X04041938
2004 Sage Publications

721

722

H.-G. HORST

The presented work is part of the integrated project


AVR Rotor which is funded by the German
government. The aim of this project is the active
reduction of shaft vibrations for rotational speeds of
up to 12,000 rpm and by means of functional materials
and the development of smart adaptive rotor systems.
Several solution strategies were developed and are now
being investigated. One such solution, based on surfacebonded PZT actuators, will be presented in this article.

AND

H. P. WOLFEL

has two nodes with four lateral degrees of freedom each


(y, z, roty, rotz). Since only lateral vibrations are of
interest, axial and torsional degrees of freedom are not
implemented. In this case the model was divided into 33
elements leading to 136 degrees of freedom.
The element matrices are derived using the principle
of virtual displacements and local trial functions.
In order to include shear deformations the work
integrals are based on Timoshenko beam theory. The
element mass and stiffness matrices are of the form:
2

FINITE ELEMENT SYSTEM MODEL


In order to design actuators and model-based
controllers for active vibration control, a mathematical
model of the system under investigation is needed.
In this work a discrete system model is built using the
finite element method. Rather than employing models
generated by commercial FE-software, the elements
were formed with self-made code and modeling
was completely done in MATLAB . Several model
parameters were then updated to fit experimental
measurements. This approach offers the advantage of
transparent and easy-to-use models. Moreover, special
rotor dynamic features like e.g. gyroscopic effects can
be easily integrated which is not possible with
most commercial codes that also contain piezoelectric
elements. The modeled structure is depicted in Figure 1.
It is a simple model of a textile machine and consists
of a mass (approximately 14 kg) on a cantilevered
rotor shaft ( 25 mm) supported by elastic bearings.
Additionally an intermittent shaft lying in between the
motor drive and the main shaft is modeled. The two
shafts are connected by flexible couplings. Active
vibration control has to be achieved in the frequency
range from 0 to 200 Hz.

6
6
6
6
6
6
6
6
6
Meli 6
6
6
6
6
6
6
6
6
4

m1

m2

m1

m2

m1 m2

m1 m2

m3

m2 m4
m3

m2

m4
m2

m1
sym:

m1 m2
m3

3
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
5

m3
2
6
6
6
6
6
6
6
6
6
el
Ki 6
6
6
6
6
6
6
6
6
4

k1

k2

k1

k2

k1 k2

k1 k2

k3

k2 k4
k3

k2

k4
k2

k1
sym:

k1 k2
k3

3
7
7
7
7
7
7
7
7
7
7:
7
7
7
7
7
7
7
7
5

k3

System Matrices
The structure is divided into pipe elements of different
lengths and outer and inner diameters. Every element

A detailed derivation and description of the coefficients mj and kj can be found in (Kramer, 1993) and

Figure 1. Rotor model.

723

Active Vibration Control in Rotor Dynamics

(Marguerre and Wolfel, 1979). The equation of motion


of the whole system in stationary coordinates is formed
by transforming the local element coordinates to global
degrees of freedom and summing up the resulting
matrices:
n
X

!
TTi Meli Ti

i1

n
X

!
TTi Keli Ti

q Mq Kq f:

i1

2
Now the elastic boundary conditions of the bearings
and flexible couplings can be integrated by adding
their stiffness values at the corresponding degrees of
freedom in the stiffness matrix. Additionally the inertia
of bearings and housings is added as discrete masses in
the mass matrix.
Gyroscopic Effects
When the shaft is rotating with rotational speed
, lateral displacements and rotations will induce
gyroscopic moments in a perpendicular direction. For
a rigid mass with polar moment of inertia P the
gyroscopic moment is easily obtained by Equation (3).
MY P _ Z
MZ P _ Y :

In the case of the cantilevered mass the rotation about


the y- and z-axis can become quite large and therefore
the gyroscopic effects cannot be neglected especially
since the rotational speeds are going to be high. For a
pipe element the gyroscopic moments can be implemented by a skew-symmetric matrix (Kramer, 1993)
2
6
6
6
6
6
el
Gi  6
6
6
6
4

g1
0

g2
0

skew  sym:

g1
g2
g4
0

g2

g1
g2
0 g1
0

g3
g2
0

3
g2 7
7
g3 7
7
7
7:
7
g2 7
7
5
g4
0
4

According to the method used in Equation (2) a


gyroscopic matrix G for the whole structure can be
obtained leading to the equation of motion:
Mq G_q Kq f

Obviously the system behavior is now depending on


rotational speed leading to certain phenomena, e.g. the
change of natural frequencies. Moreover there now

exists a coupling between the motions in the y- and


z-directions.
External and Internal Damping
Damping forces can arise both in stationary parts of
the system like bearings or supports or in rotating parts
like the shaft. Since they have quite different results
on the system behavior they are separated in external
and internal damping forces respectively. The effects
of external damping can be modeled as in any other
stationary structure. Since very little discrete damping
parameters are known, modal damping ratios are
assumed which can be experimentally identified by
means of modal testing. The damping ratios Di are then
employed to calculate a Damping matrix Dex by using
the natural frequencies !i , the generalized masses mgi
and the matrix of eigenvectors, U of the undamped
system:
 1
Dex UT diag2D1 mg1 !1 , . . . , 2Di mgi !i , . . . ,
2Dn mgn !n U1 :

Internal damping results from material damping of


the shaft and much more important from damping
caused by assembly components like shrink fits,
couplings or viscous layers on the shaft (Gasch et al.,
2002). It is more difficult to determine and therefore
is assumed to be a fraction of the viscous external
damping:
Dint cD Dex

The parameter cD is not known a priori but may be


tuned to fit experimental observations. To integrate
internal damping in the model, the equation of motion
has to be transformed into a complex rotating coordinate frame. In rotating coordinates the viscous damping
is implemented like external damping in a stationary
system. The transformation back to stationary coordinates causes an additional term Dik in the equation of
motion which is proportional to the displacements q
(Gasch et al., 2002):




Mqq Dex Dint G q_ K Dik q f

The coefficients of Dik are partly negative resulting


in system instability at certain rotational speeds. Hence
to avoid stability problems, internal damping has to be
minimized and external damping has to be increased.
For anisotropic bearings the frequencies at which the

724

H.-G. HORST

lim
i

H. P. WOLFEL

excitation is a static load with respect to rotating parts


the reactive power consumption is close to zero.

system is becoming unstable can be obtained by:


s

2 
2
Dex

L
i
1 int
!i
int
Di
Di

AND

Modal State Space Representation and


Model Reduction
Especially for the design of model-based controllers,
a modal state space representation of the system is
very useful. Because of gyroscopic effects and internal
damping the system is no longer proportionally damped.
Therefore for modal decomposition a bimodal approach
using left and right eigenvectors is necessary (Kramer,
1993). The resulting eigenvalues and eigenvectors are
complex but can be transformed to real values using a
similarity transformation (Nordmann, 2001).
For real time computations and controller implementation, the model order still becomes too large. The
modal representation enables to reduce the model
order by modal reduction neglecting the higher modes.
The resulting error in the interesting lower frequency
range can be compensated by a static correction term
(Horst and Kronig, 2001).

Principle of Operation and Modeling


The bending vibrations of the rotor occur both in the
yx-plane and in the zx-plane. Therefore two orthogonal
pairs of PZT actuators are used to generate actuator
forces. The actuators of one pair are positioned opposite
to each other and are driven out of phase, so that one is
extending while the other is contracting. Consequently
a bending moment around the y- or the z-axis is
induced in the structure (see Figure 2(a) for the zx-plane
actuators).
The effects of surface-bonded actuators can thus
be separated in passive mechanical effects caused by
the actuators mass and stiffness and the active electromechanic effect, which is the actuation strain caused
by an electric voltage applied to the actuator surfaces.
The effects of the actuators mass and stiffness are taken
into account by building additional pipe elements. The
mass and stiffness matrices of those actuator elements
are added to the main matrices at the degrees of freedom
where the piezo elements are positioned.
The actuation strain induced in the shaft at point A
(Figure 2(b)) by a pair of actuators is a fraction of the
unconstrained piezoelectric strain due to a voltage VP:

SURFACE-BONDED PZT ACTUATORS


In the field of active vibration control the approach of
controlling structural vibrations of beams by means of
surface-bonded PZT actuators is well-known (Hagood
et al., 1990; Horst and Kronig, 2001). The contribution
of this work is to transfer the method to a rotor dynamic
system by applying PZT plate actuators on the surface
of an elastic shaft. The actuation forces can thus be
directly applied on the shaft where controllability of
vibration modes is optimal. Hence very little additional space is required on the shaft and at the bearings.
Moreover the placement on the rotating shaft leads to
reduced power consumptions when suppressing synchronous vibrations. Due to the fact that synchronous

"S kd31

VP
,
tP

10

d31 and tP being the piezoelectric strain constant and the


thickness of piezoelectric layer respectively. The factor k
can be obtained by using the force equilibrium and
the geometric constraints in the shaft and the actuators
(Lin and Chu, 1994). Finally a control moment can be
derived which is proportional to the applied voltage:
4 EP d31 br0 tP 3 r30 c sin 
VP
31 tP
ES IS =EP IP :

MP

Figure 2. Pipe element with attached actuator pair.

with
11

725

Active Vibration Control in Rotor Dynamics

In Equation (11) only a linear one-dimensional stress


distribution is assumed. Therefore the actuator model
was compared to results from a three-dimensional FEmodel of the pipe-actuator system in ABAQUS . For
the selected parameters of PZT thickness and angle of
enclosure both results showed good agreement, differing
by about 8%.
Actuator Placement and Sizing
For modal structural control, the controllability of
each mode to be influenced has to be optimized. Hence
the piezo actuators have to be properly sized and
positioned. According to Figure 2(a) the control
moments act on the right and left edges of the PZT
elements. Thus, to maximize controllability of the ith
mode, the generalized control moment


g
MP,
i r, i  l, i MP

12

must be maximized. Since the moment is acting on


rotational degrees of freedom, Equation (12) is proportional to the integral of the ith physical mode shapes
mean curvature in the actuator region. The actuators
themselves will change the mode shapes because of their
mass and stiffness. Therefore an iterative optimization
procedure was chosen by first implementing actuators
on a certain position then calculating the mode shapes
and the generalized control moment. By repeating the
procedure for varying positions the optimal placement
for each mode can be found. In this case the first two
mode shapes of the rotor need to be controlled. Since
the optimal position for the first and the second mode is
different, a compromise has to be found. The unbalance
excitation being highest in the second mode, the
actuators are positioned near the optimal place for
the second mode. Because of the material effort and the
limits of the supplier, the actuators length is chosen to
be 140 m (two 70-mm patches). This is much shorter
than the optimal length for the first two modes though
still providing sufficient actuation force.

CONTROLLER DESIGN
For the active vibration control of the rotor system,
an optimal feedback controller is designed. The optimal

LQ-regulator can be calculated by minimizing the


following cost functional:
Z1
min J, and J
KLQ


pT tQpt uT tRut dt:

13

In Equation (13) the modal states p rather than the


physical states are used in the integral. The resulting
controller KLQ thus is a modal controller. This enables
to weigh each vibration mode separately and to shape
the control behavior in the frequency domain by using
the weighing matrix Q. The control outputs p are
weighed by R to limit the resulting voltages to levels that
can be used with the actuators ( 500 V). The controller
scheme relies on full state feedback but since only four
states are measured, the other states have to be
estimated by an observer. The observer being used is
a Luenberger observer containing a modally reduced
system model. The LQ-regulator is only a first step
towards robust control and is used for feasibility
studies and simulations, one major drawback being the
assumption of a time invariant system. From Equation
(8) it is clear that the system behavior is dependent on
rotational speed, therefore controller adaptation is
needed. Additionally arbitrary changes in the system
may arise from different operation conditions in a real
world application. Hence future work will concentrate
on the development of robust adaptive control for real
time implementation.

SIMULATION RESULTS
The material data used for all the simulations is listed
in Table 1. First free vibrations of the rotor system,
especially natural frequencies and mode shapes were
simulated. Figure 3 shows the dependence of natural
frequencies on the rotational speed. The first two
frequencies (15.4 Hz, 15.7 Hz) belonging to the first
mode in y- and z-direction change by about 14% from
0 to 12,000 rpm. The next two frequencies change only
by about 2% since the rotor mass is concentrated near a
node of the second mode shape. Besides the frequencies
also the eigenvectors will change with increasing speeds.
Since unbalances are the most common excitation
mechanism in rotor dynamics, unbalance excitation

Table 1. Material properties.


Actuator material
Strain constant d31
PZT thickness
Angle of enclosure 2
Actuator length

PIC 255
180  1012 C/N
1 mm
90
140 mm

Youngs modulus of the shaft ES


Youngs modulus of PZT EP
Max. actuator voltages VP,max
Modal damping ratio D1
Modal damping ratio D2

210 GPa
62 GPa
 500 V
0.18%
0.27%

726

H.-G. HORST

AND

H. P. WOLFEL

Figure 3. Campbell diagram: change of natural frequencies with respect to rotational speed.

Figure 4. Amplitude response for unbalance excitation: with and without control q1: z-displacement at the rotor mass; q2: z-displacement in
the middle between the bearings.

is used in the simulation as a reference for actuator


design and assessment of controllers. Two unbalance
forces due to an eccentricity of 2 mm are assumed, one
at the rotor mass, the other in the middle between
the bearings. This unbalance is equivalent to a rotor
balanced with a quality of G 2.5 (according to DIN ISO
1940) being common for turbo machinery. The resulting
displacement amplitudes for the uncontrolled system
depicted in Figure 4 are rather large especially in the

second resonance (peak value 1.4 mm). The sharp


resonance peaks do not show in the controlled system
anymore since the controller was designed to reduce
vibration in these modes.
The vibration suppression shown in Figure 4 is 96%
in the first and 98% in the second resonance. The
control effort can be judged by the amplitude of the
actuator voltage output by the controller. Its maximum
value is about 400 V in the second resonance leading

Active Vibration Control in Rotor Dynamics

to the conclusion that the actuators are sized properly


to compensate the assumed excitation.

EXPERIMENTAL RESULTS
For experimental investigations a rotor test-rig was
built. Since curved piezo ceramics are very difficult to
manufacture, common PZT plate elements are used
instead. Therefore the shaft surface had to be milled
to get an octagonal shape. The piezo elements are
attached to the shaft using a two-component epoxy. The
transmission of actuator voltages to the rotating shaft
is done by a high speed slip-ring with gold brushes.
Metal bellow couplings are used to connect the shafts
and the motor. To measure the shaft vibrations four
eddy current sensors are used (two in y- and two in
z-direction). Additionally the vibrations at the bearings
are measured using piezoelectric accelerometers. From
modal testing the modal damping ratios (see Table 1)
were identified. The stiffness of the bearings and
couplings in the model were tuned to match the natural
frequencies and mode shapes of the test-rig. Further
experiments concentrated on the system response to
PZT actuator excitation in order to validate the actuator
model.
Figure 6(a) shows the response for harmonic excitation at 30 Hz. The displacement amplitude is roughly
linear dependent on the applied voltage as expected from
the model (Equation 11). Simulation and measurement
agree quite well considering the different shape of the

727

actuator patches in the test-rig. The measured frequency


response in Figure 6(b) shows also good agreement with
simulation results. Again the level of the displacement
amplitude is well predicted. The quality of the structural model can be seen in the agreement of resonance
and anti-resonance positions. The additional peaks in
the measurements around 120 and 155 Hz are due
to unconsidered dynamics of the rotor foundation.

CONCLUSIONS
A structural FE-model of a rotor system with
attached piezo actuators on the shaft surface was

Figure 5. Experimental setup.

Figure 6. Comparison of simulated and measured actuator response for  0. (a) response for harmonic PZT excitation with 30 Hz and varying
voltages and (b) frequency response for PZT noise-excitation at two shaft locations S1 (near rotor mass) and S2 (between the bearings).

728

H.-G. HORST

developed and validated by means of experimental


testing. The actuator response and the system behavior
are well predicted by the model although the dynamics
of the foundation are not modeled yet. An optimal
feedback controller was designed to simulate the closedloop response and demonstrate the potential of active
vibration control. In the next steps robust adaptive
controllers will be developed and implemented in the
experimental setup. Further experimental tests and
simulations need to be done especially concerning the
rotating system, excitation mechanisms and closed-loop
performance.

ACKNOWLEDGMENTS
This work and the underlying project AVR Rotor
(Reference No.: 02PP2282) was sponsored by the
German federal department for education and research
(BMBF). The authors are responsible for the contents of
this publication.

REFERENCES
Barrett, T.S., Palazzolo, A. and Kascak, A. 1995. Active Vibration
Control of Rotating Machinery Using Piezoelectric Actuators
Incorporating Flexible Casing Effects, Journal of Engineering
for Gas Turbines and Power, 117(1):176187.
Gasch, R., Nordmann, R. and Pfutzner, H. 2002. Rotordynamik, 2.
Auflage. Springer Verlag, Berlin, Heidelberg.

AND

H. P. WOLFEL
Hagood, N.W., Chung, W.H. and von Flotow, A. 1990. Modelling
of Piezoelectric Actuator Dynamics for Active Structural
Control, Journal of Intelligent Material Systems and Structures,
1:327354.
Horst, H.-G. and Kronig, K. 2001. Aktive Schwingungsminderung
mit piezokeramischen Aktuatoren an einer elastischen Balkenstruktur, in: VDI-Berichte 1606: Schwingungen in Anlagen und
Maschinen, VDI Verlag GmbH, Dusseldorf, 1(1):143162.
Kramer, E. 1993. Dynamics of Rotors and Foundations. Springer
Verlag, Berlin Heidelberg.
Lin, Y.-H. and Chu, C.-L. 1994. Comments on Active Modal Control
of Vortex-Induced Vibrations of a Flexible Cylinder, Journal
of Sound and Vibration, 175(1):135137.
Marguerre, K. and Wolfel, H.P. 1979. Mechanics of Vibration. Sijthoff
& Noordhoff, Alphen aan den Rijn.
Nordmann, R. 2001. Mechatronische Systeme im Maschinenbau.
Shaker Verlag, Aachen.
Nordmann, R., Aenis, M., Knopf, E. and Straburger, S. 2000.
Active Magnetic Bearings A Step towards Smart Rotating
Machinery, Transactions of the 7th International Conference on
Vibrations in Rotating Machinery IMechE, July 2000,
Nottingham, UK.
Schweitzer, G. 1998. Magnetic Bearings as a Component of Smart
Rotating Machinery, In: Proceedings of the 5th International
Conference on Rotor Dynamics IFToMM, September 710, 1998,
Darmstadt, Germany.
Tang, P., Palazzolo, A., Kascak, A., Montagne, G. and Li, W.
1995. Combined Piezoelectric-Hydraulic Actuator Based Active
Vibration Control for Rotordynamic Systems, Journal of Vibration and Acoustics, 117(1):285293.
Ulbrich, H. 1998. Active Vibration Control of Rotors, In:
Proceedings of the 5th International Conference on Rotor
Dynamics IFToMM, September 710, 1998, Darmstadt, Germany.
Zhou, S. and Shi, J. 2001. Active Balancing and Vibration Control of
Rotating Machinery: A Survey, The Shock and Vibration Digest,
33(5):361371.

You might also like