Sie sind auf Seite 1von 19

The current issue and full text archive of this journal is available at

www.emeraldinsight.com/1355-2511.htm

METHODOLOGY AND THEORY Blades condition


monitoring
Blades condition monitoring using
shaft torsional vibration signals
275
B.O. Al-Bedoor
Mechanical Engineering Department, Faculty of Engineering and Technology,
University of Jordan, Amman, Jordan, and
S. Aedwesi and Y. Al-Nassar
King Fahd University of Petroleum and Minerals, Dhahran, Saudi Arabia
Abstract
Purpose – The purpose of this paper is to validate mathematically the feasibility of extracting the
rotating blades vibration condition from the shaft torsional vibration measurement.
Design/methodology/approach – A mathematical model is developed and simulated for
extracting rotating blades vibration signatures from the shaft torsional vibration signals. The
model simulates n-blades attached to a rigid disk at setting angles and the shaft drives the disk is
flexible in torsion. The model is developed using the multi-body dynamics approach in conjunction
with the Lagrangian dynamics. A three-blade rotor system example is simulated for blades free and
forced vibration under stationary and rotating conditions. Frequency spectrums for the shaft torsional
and blades bending vibration are represented and studied for analysis verification purposes.
Findings – The torsional vibration frequency spectrums showed blades free and forced vibration
signatures. The blade setting angle is shown to reduce the sensitivity of torsional vibration signal to
blades vibration signatures as it increases. The torsional vibration signals captured the variation in blades
properties and produced broadband frequency components for mistuned system. The shaft torsional
rigidity is shown to reduce the sensitivity of torsional vibration signal to blades vibration if increased to
extremely high values (approaching rigid shaft). The rotor inertia is shown to have less effect on the
torsional vibration signals sensitivity. The method of torsional vibration as a tool for rotating blades
vibration measurement, based on the proposed mathematical model and its simulation, is feasible.
Practical implications – There is a growing need for reliable predictive maintenance programs
that in turn requires continuous development in methods for machinery health monitoring through
vibration data collection and analysis. Turbo machinery and bladed assemblies like fans, marine
propellers and wind turbine systems usually suffer from the problem of blades high vibration that is
difficult to measure. The proposed new method for blades vibration measurement depends on the shaft
torsional vibration signals and can be used also for verifying the signals from other types of bearings
sensors for possible blades vibration condition monitoring.
Originality/value – This paper presents a unique mathematical model and simulation results for the
rotating blades vibration monitoring. The developed model can be simulated for studying coupled
blades vibration problems in the design stage as well as for condition monitoring in maintenance
applications.
Keywords Condition monitoring, Vibration, Maintenance
Paper type Research paper

The authors acknowledge the support of King Fahd University of Petroleum. This work is Journal of Quality in Maintenance
Engineering
funded by KFUPM Research Office under project number ME/BLADE-VIBRATION/215. The Vol. 12 No. 3, 2006
first author acknowledges the support of the University of Jordan. pp. 275-293
q Emerald Group Publishing Limited
B.O. Al-Bedoor is on leave from King Fahd University of Petroleum and Minerals, Dhahran, 1355-2511
Saudi Arabia. DOI 10.1108/13552510610685110
JQME 1. Introduction
12,3 Rotating bladed mechanical systems are widely used as basic parts in machinery.
Bladed rotors can range from systems with hundred of blades as in turbomachinery
to systems with three to four blades as in fans, marine propellers and wind
turbines. Regardless of the number of rotating blades difficulties were faced in
measuring the rotating blades vibration due to the fact they are rotating and
276 interacting with the working environment. Recent laboratory experiments on rotors
with limited number of blades and field measurements on fans (Maynard and
Trethewey, 1999, 2001; Maynard et al., 2000) showed that rotating blade vibration
frequencies could be extracted from the shaft torsional vibration signal. However,
due to the experimental nature of these studies and the importance of the problem
of rotating blades vibration measurement, a mathematical model that enables
verifying the approach and opens the road for more quantified studies in this
direction is extremely needed.
Vibration measurement is known as a powerful tool in machinery condition
monitoring, Laws and Muszynska (1987) and Vance (1988), that put growing
demands on developing reliable vibration measuring systems. These systems are
required to sense and represent closely the machinery particular component
vibration with minimum interference from other vibration sources. When blades
vibration information is required, the task becomes very complicated, Simmons and
Smalley (1990), as blades are rotating and interacting with the working
environment. To directly monitor blade vibration, strain gages were used in
many laboratory-testing studies (Kaufman and Kershisnik, 1984; Srinivasan and
Cuts, 1984; Scalzo et al., 1986; Fan et al., 1994) that showed the practical technique
limitations in terms of sensor survival due to the harsh operational conditions.
Other techniques for blades vibration measurement were proposed; among these
techniques is the use of Laser-Doppler and optical methods (Cookson and
Bandyopadhyay, 1980; Nava et al., 1994; Reihardt et al., 1995), with some problems
and limitations on the speed of rotation. Detailed discussion of the available
methods for blade vibration measurement is reported by Al-Bedoor (2002). Wherein,
he classified methods for blades vibration measurement into two categories: the
direct and the indirect methods.
The approach of extracting blades vibration frequencies from the shaft torsional
vibration was investigated experimentally by Maynard and Trethewey (1999, 2001)
and Maynard et al. (2000). They explained the superiority of this approach to the
lateral vibration approach, mainly, as torsional vibration is less affected by the
boundary conditions than lateral vibration. In addition, Muszynska et al. (1992) used
the torsional vibration to identify rotor crack. They came up with an explanation for
the sensitivity of torsional vibration that damping in torsional vibration modes is
minimal. Al-Bedoor (1999) proposed a mathematical model for the coupled blade
bending and shaft torsional deformations in rotors using the finite element method.
In addition, Al-Bedoor and Al-Nassar (2001) used the ANSYS finite element package
to model the coupled multi-bladed disk-shaft flexible system. The ANSYS model
was able to simulate the system rotating condition for natural frequencies and mode
shape analysis. The package accounts for the effect of rotation in the geometrical
stiffness matrix and no coupling between the rigid body and flexible modes was
affordable. Due to the dimensionality of the finite element model of reference Blades condition
(Al-Bedoor, 1999) and the limitation of the ANSYS model (Al-Bedoor and Al-Nassar, monitoring
2001) to account for the coupled dynamics, Al-Bedoor (2001) developed a reduced
order model for the coupled shaft torsional and blade bending vibrations using the
assumed modes method. This model simulated the shaft-disk-blade system driven
by a torque that enabled monitoring the system vibration in its start-up. The model
simulations showed strong dynamic coupling between blade bending and shaft 277
torsional vibrations. However, the model was not oriented to account for steady
state running condition and no provision was given to account for the blade setting
angle and multi-bladed systems.
In this study a mathematical model for a shaft-disk-blades system that is
running at constant rotating speed is developed and simulated. The shaft is
considered to be flexible in torsion and the disk is assumed to be rigid. The blades
are modeled as attached to the rigid disk with a setting angle that allows
monitoring the tangential (lead-lag) and axial (flapping) blades vibration modes.
Moreover, the model allows simulating the effect of differences in blades properties
that is known as the mistuning effects. The model is developed using the
Lagrangian dynamics in conjunction with the assumed modes method to descritize
blade deformation. The model is simulated for various transient and steady state
blades vibrations and the shaft torsional vibration is monitored in time and
frequency domains.

2. Model development
The schematic diagram of the motor-shaft-disk-blades system is shown in Figure 1.
The model is developed having the following assumptions in mind:
.
The system is rotating at constant running speed v that represents steady state
running condition.
.
The disk is rigid with radius Rd, mass md and mass moment of inertia Jd.
.
The shaft is flexible in torsion and no lateral shaft deflections are considered.
.
The blades are simple uniform inextensible beams and the Euler-Bernoulli beam
theory is adopted. The assumed modes method (AMM) and the cantilevered

Figure 1.
Shaft-disk-blade unit
driven by electrical motor
through flexible coupling
JQME modes shapes are used. The blades are attached radially to the rigid disk with
12,3 setting angle bi for each blade.
.
The shaft torsional and blade bending deformations are small.
.
The effect of axial dynamics known as blade stiffening is accounted for using the
added potential energy that results from the inertial forces and blades
shortening.
278
The coordinate system used in developing the model is shown in Figure 2 that shows a
model with three blades and the typical deflected configuration of one of the blades.
The used axes are the inertial reference frame XYZ and the blade body coordinate
system x b y b z b .

2.1. Kinetic energy expressions


The kinetic energy of the shaft-disk-blades unit is constituted of the motor kinetic
energy U m , the disk kinetic energy U d and the blades kinetic energy U b . The shaft
inertia is lumped into the disk and motor inertia. The motor rotor and the disk are
modeled as rigid inertia with mass moment of inertia J m and J d about the axis of
rotation (Z), respectively. As the system is rotating at a constant angular velocity v, the
motor and disk kinetic energy expressions can be written, respectively, as:

1
Um ¼ J mv 2
2 ð1Þ
1 ! "2
U d ¼ J d v þ c_
2

Figure 2.
Blades coordinate systems
and typical blade-deflected
configuration
where c is the small shaft torsional deformation angle measured with respect to the Blades condition
motor shaft. monitoring
To develop the kinetic energy expression for the blades, the deformed configuration
of the shaft-disk-blade(s) system, shown in Figure 2, is used. The global position vector
of a material point P, on a typical blade, can written as:
( )
Rd cos u # $ 279
RP ¼ þ AðuÞ r bP ð2Þ
Rd sin u

where the angle u ¼ vt þ c describes the rotation of the blade coordinate system
x b y b z b with respect to the inertial reference frame XYZ due to the system rigid body
rotation at constant speed v and the shaft torsional deformation angle c. r bP is the
position vector of a material point on the deformed blade principal axis and can be
written as:
8 9
>
> x >
>
< =
r bP ¼ uðx; tÞ cos b ð3Þ
>
> >
: uðx; tÞ sin b >
;

where uðx; tÞ is the blade bending


# deflection
$ and b is the blade setting angle. The
rotational transformation matrix AðuÞ can be written in the following form, based on
the assumption of small torsional deformation:
2 3
cos vt 2 c sin vt 2c cos vt 2 sin vt 0
# $ 6 7
AðuÞ ¼ 6
4 sin vt þ c cos vt 2c sin vt þ cos vt 07
5 ð4Þ
0 0 1

The velocity vector of the material point in the inertial reference frame can be obtained
by differentiating equation (2) that after manipulation can be represented as follows:
8 9
>
> 2L1 sin vt 2 L2 cos vt >
>
< =
R_ P ¼ 2L2 sin vt þ L1 cos vt ð5Þ
>
> >
>
: u_ sin b ;

where:
! "
L1 ¼ v Rd þ x 2 cu cos b þ c_ðx þ Rd Þ þ u_ cos b
! " ð6Þ
L2 ¼ Rd cv þ v xc þ u cos b þ c_u cos b þ cu_ cos b

The kinetic energy of blade i can be found by using the velocity vector of equation (5)
into the integral:
Z
JQME 1 L
T
U bi ¼ rR_ P R_ P dx ð7Þ
12,3 2 0

where r is the ith blade mass per unit length and L is its length.
Now the total kinetic energy expression of the system can be written as follows:
280
X
n
U ¼ Um þ Ud þ U bi ð8Þ
i¼1

where n is the number of blades.

2.2. Potential energy expressions


The system potential energy is constituted of the blades bending strain energy, V b , the
torsional strain energy, V T and the potential energy of the axial shortening due to
transverse deformations and the motion generated inertial forces, V A . The ith
blade-bending elastic strain energy that has flexural rigidity EI ðxÞ is given by:

Z L
1 ›2 u
V bi ¼ EI ðxÞ dx; ð9Þ
2 0 ›x 2

the torsional elastic potential energy stored in the flexible coupling is given by:

1
VT ¼ kT c 2 ð10Þ
2

The axial shortening due to transverse deformations in conjunction with the radial
inertial forces contributes to the system elastic potential energy by the known
(Al-Bedoor, 2001), axial shortening potential energy. The system potential energy is:

X
n
V ¼ VT þ ðV bi þ V Ai Þ ð11Þ
i¼1

where n is the number of blades.

2.3. Equations of motion


Now using the discretized form of the system kinetic and potential energy expressions
into the Lagrange’s equation, performing the required differentiation and
mathematical manipulation for the system degrees of freedom (c, fqgi , i ¼ 1 ! n), n
is the number of blades, the system equations of motion can be represented in the
following compact matrix form:
2 38 € 9 2 38 _ 9
mcc mcq1 mcq2 ··· mcqn >
>
>
c > >
>
ccc ccq1 ccq2 ··· ccqn >
>
>
c > >
>
Blades condition
6 7>
> >
> 6 7>
> >
>
6
6 mq1 q1 0 ··· 0 7> 7> >
> fq€ g1 >
>
6
6 cq1 q1 0 ··· 0 7>7>
> 1>
fq_ g >
> monitoring
6 7>
> >
> 6 >
7> >
>
6 7>
< >
= 6 >
7< >
=
6 mq2 q2 ··· 0 7 fq€ g2 6 cq2 q2 ··· 0 7 fq_ g2
6 7 þ6 7
6 7>
> >
> 6 >
7> >
>
6 .. 7> > 6 .. 7> >
6 Symm 0 7 > ... >
> > 6 Symm 0 7 > ... >
> >
6 . 7>
>
>
>
>
> 6
. >
7> >
>
4 5>
> >
> 4 5>
>
>
>
>
>
>
mqn qn : fq€ gn ;
> > >
cqn qn : fq_ gn ; 281
8 9
2 38 9 > Fc > ð12Þ
kcc 0 0 ··· 0 > c >
> > > >% & >
> >
>
6 7>>
>
>
> > > F >
>
6 kq1 q1 0 ··· 0 7 > fqg >
> > >
> > q 1>
> >
>
6 7>> 1>
> >
> >
>
6 7>
7<> >
>
= < > %F & >
> >
6
6 kq2 q2 ··· 0 7 fqg2 q 2=
þ6 7 ¼ ' ( >:
6 7> > >
6 .. 7>> .. >
> > >
> > .. >
> >
>
6 Symm 0 7 > . > > . >
>
6 . 7>>
>
>
>
> > >
>
>
>
4 5>> >
> > >
> > > >% & > >
kqn qn : fqgn ; > > Fq >
: >
;n

The entries of equations (12) are as follows:

X
n X
n
mcc ¼ J d þ J bi þ cos 2 bi fqgi ½I &fqgi
i¼1 i¼1
h # $i
mcqi ¼ cos bi ½a&i þRd b i þ c cos 2 bi fqgi ½ I &
! "
mqi qi ¼ ½ I & 1 þ c 2 cos 2 bi
X n
ccc ¼ 2 cos 2 bi fqgTi ½ I &fq_ gi
i¼1
# $ ð13Þ
ccqi ¼ 2vc cos bi ½a&i þRd ½b&i þ v cos 2 bi fqgi
cq q ¼ 2cc_ cos 2 bi ½I &
i i

X
n X
n
kcc ¼ kT 2 v 2 J bi 2 v 2 cos 2 bi fqgi ½I &fqgi þ2v cos 2 bi fqgi ½I &fq_ gi
i¼1 i¼1
EI i # $ # $
kqi qi ¼ 4
k þ v 2 ksi 2 v 2 ð1 þ c 2 Þ cos 2 bi ½I & 2 2vc_ cos 2 bi ð1 þ cÞ½I &
ri L i

where fqgi is the ith blade modal vector that has a size equal to the number of modes N
& bi is the setting angle for the blade, #F c$ is the torsional
considered for this% blade,
excitation torque, F q i is the ith blade modal excitation vector,
# $ ksi is the ith blade
stiffening matrix due to system rotation, the vectors ½a&i , b i are one dimensional
vectors that represent the coupling between the ith blade modal degrees of freedom and
the torsional degree of freedom of the system.
The size of dynamic system, equations (12), depends on the number of blades n in the
system and the number of modes N considered for each blade, which result in a system
of ð1 þ n £ N Þ £ ð1 þ n £ N Þ size. The first matrix in equations (12) is the coupled
inertia matrix, wherein its first entry is the torsional degree of freedom inertia that is
JQME function of the disk inertia J d , the blades inertia J b , the setting angles bi and nonlinear
12,3 function of the blades elastic modal deflections fqgi . The rest of the first row entries of the
inertia matrix are the coupling vectors between the torsional degree of freedom and the
blades modal degrees # $ of freedom that are functions of the setting angle bi , the coefficient
vectors ½a&i and b i and the torsional deflection c multiplied by the blades modal
degrees of freedom fqgi . The diagonal entries are the modal inertia matrices for each
282 blade. The second matrix is a matrix that developed from the formulation, not damping
matrix, and can be called the Coriolis matrix. The entries of this matrix are nonlinear
functions of the torsional deflection and velocity, the blades modal deflections and
velocities, the setting angles and the speed of rotation. Structural damping can be added
to the entries of the second matrix at corresponding blades entries. The third matrix is
the equivalent stiffness matrix with the first entry corresponding to the torsional degree
of freedom stiffness. As shown, in addition to the shaft torsional stiffness kT , the
equivalent torsional stiffness is softened as a result of speed of rotation v, the blades
inertia J b , the setting angles and the blades modal deflections. It is stiffened as function
of the blades modal deflections and velocities that are multiplied by the speed of rotation.
The diagonal entries of the stiffness matrix are the blades modal stiffness matrix that are
softened and stiffened as function of the speed of rotation, the setting angles and the
torsional
# deflection.
$ Moreover, the inclusion of the stiffening effect in the form of the
matrix ks i has produced coupling # $ in the stiffness matrix between the individual modes
of each blade; as the matrix ks i is non-diagonal.

3. Numerical simulations and discussion


To simulate the dynamic model of equations (12), the data of the experimental set-up
available in the Advanced Mechanics Laboratory (AML) at KFUPM are used. The
properties of this three-blades-disk-shaft system, shown in Figure 2, are given in
Table I. Structural damping is accounted for in the simulation by adding the Rayleigh
proportional damping to the system with damping ratio j ¼ 0:01 for the shaft torsional
and blade bending modes as all are made of steel. The blades undamped natural
frequencies, for the non-rotating and uncoupled system, are calculated as shown in
Table II. In all simulations, the system basic parameters are kept the same except the
parameter that is explicitly mentioned to be changing.

Property Value

Blade material Steel (E ¼ 200 GPA, r^ ¼ 7; 850 kg=m3 )


Blade length L 0.125 m
Blade cross section 2:54 £ 0:1 cm
Blade mass per unit length, r 0.2 kg/m
Blade flexural rigidity, EI 0:4233 N:m2
Disk radius, RD 0.05 m
Disk mass Md 3.08 –kg
Disk moment of inertia, JD 0:01233125 kg:m2
Steel shaft G ¼ 80 GPA
Shaft length 0.5 m
Table I. Shaft diameter 1 cm
Blade-disk-shaft data Torsional stiffness kT 157.08 N.m/Rad
The dynamic model, equations (12) has a size of 16 £ 16 for the considered three-blades Blades condition
model with five modes for each blade, is solved using the ODE15S solver of the monitoring
MALTAB package and the system free and forced vibration response is monitored.
The system response is represented as direct time history and power spectral density
(PSD). The PSD gives the amplitude at each frequency component squared and divided
by the frequency bandwidth (Hz) to eliminate the spectrum amplitude dependence on
the frequency bandwidth (Thompson and Dahleh, 1998). The simulations are 283
conducted for free and forced vibration of the blades.

3.1. Free vibration simulations


For free vibration, selected blades are given static tip deflections d and the individual
corresponding modal values are calculated based on the integration of assumed static
deflection curve that is physically acceptable. After integration, the modal initial
conditions for the first five modes as function of the blade i tip deflection are found in
the form:
pffiffiffiffiffiffiffiffi
di ri L i # $
qi ð0Þ ¼ 0:9707 20:0247 0:0046 20:0008 0:0003 : ð14Þ
2
For the non-rotating three-blades unit with properties given in Table I and setting
angles b ¼ 08 for all blades, a free vibration simulation is conducted. For the purpose
of introducing mistuning, blade 2 length is made 0.9 of other blades length and blades 1
and 2 are given tip deflections of 2 0.012 m. The system response is shown as time
history for blades 1-3 and shaft torsional vibration in Figure 3. The difference in the
period of oscillation between Figures 3a and 3b, for blades 1 and 2, shows the effect of
blade 2 having shorter length. Blade 3 was not given any initial tip deflection, however
its time history, Figure 3c, shows more sophisticated time history with less decaying
rate. The torsional vibration signal, Figure 3d, shows also similar time history as blade
3. The corresponding frequency spectrums are shown in Figure 4. Figure 4a shows that
blade 1 is vibrating at its own first mode natural frequency of about 52 Hz. Blade 2
frequency spectrum, Figure 4b, contains its first mode natural frequency, which is
about 60 Hz due to the decrease in its length by 10 percent. Although blade 3 is not
given any initial excitation, it vibrates with three different frequencies as shown in the
spectrum of Figure 4c. The 52 Hz and 60 Hz frequencies are blades 1 and 2 vibration
frequencies, respectively. The third frequency component is 325 Hz, which is blade 3
second mode free vibration frequency due to its base excitation by the shaft torsional
vibration. This means that blades 1 and 2 free vibration excited the shaft torsional
vibration and thus excited blade 3 vibration. Figure 4d shows the torsional vibration
frequency spectrum associated with blades free vibration. The spectrum shows not

Degree of freedom Natural frequency (Hz)

Blade 1st bending mode 52.1


Blade 2nd bending mode 326.25
Blade 3rd bending mode 914.271 Table II.
Blade 4th bending mode 1,791.6 Uncoupled blades natural
Blade 5th bending mode 2,961.6 frequencies
JQME
12,3

284

Figure 3.
Shaft-disk-blade free
vibration response time
history, blades 1 and 2 are
given initial tip deflection
of 20.01m
Blades condition
monitoring

285

Figure 3.
JQME
12,3

286

Figure 4.
Shaft-disk-blade free
vibration response
frequenct spectrums
Blades condition
monitoring

287

Figure 4.
JQME only blades 1 and 2 free vibration frequency components but also captures blade-3
12,3 second mode free vibration frequency, which is excited by the shaft torsional vibration.
In addition, one can observe that the difference in blades 1 and 2 vibration frequency is
reflected in the torsional vibration spectrum. This directs the attention towards the
broadness of torsional vibration spectrum as indication of mistuning in the blades.

288 3.2. Forced vibration simulations


To simulate operational condition the three-blades disk shaft model is operated at
100 Hz rotating speed and the shaft torsional vibration signals are monitored for
different excitations and system properties. Figure 5 shows the torsional vibration
spectrum when all blades have setting angles b ¼ 08 and blades 1, 2 and 3 are excited
by sinusoidal force at their tips with frequencies 1 £ (at running speed), 2 £ (twice
running speed) and 3 £ (three times running speed), respectively. As shown the
torsional vibration spectrum captures the vibration of the three blades at their
respective external excitation frequencies. In addition, the spectrum captures the
blades free vibration frequency of about 70 Hz that is increased in this case due to
stiffening effect. For the 100 Hz speed of rotation, the system is simulated when the
setting angels are b1 ¼ 08 , b2 ¼ 458 and b3 ¼ 858 under 1 £ , 2 £ and 3 £ excitation
frequency for blades 1, 2 and 3, respectively. The torsional vibration spectrum,
Figure 6, reflects the effect of setting angle by showing less sensitivity to blade 3
vibrations as b3 ¼ 858 . The effect of shaft torsional rigidity on the sensitivity of the
system torsional vibration signal to blades vibration is investigated by increasing the
shaft torsional stiffness by 103 and 106 times its original kT . The torsional vibration
spectrums, Figures 7a and 7b, show that the torsional vibration signal carries the

Figure 5.
Shaft torsional vibration
frequency spectrum for a
rotor rotating at speed
v ¼ 100 Hz
Blades condition
monitoring

289

Figure 6.
Shaft torsional vibration
frequency spectrum for a
rotor rotating at speed
v ¼ 100 Hz

three-blades vibration signatures but with less amplitude as the shaft torsional rigidity
is increased to extreme values approaching rigid shaft. The effect of increasing the
drive mass moment of inertia is investigated by increasing the disk mass md by 100
and 1,000 times. The resulting torsional vibration spectrums are shown in Figures 8a
and 8b, for 100md and 1,000md disk mass, respectively. As shown the spectrums
capture the blades forced vibration signatures effectively.
To this end one can observe that for all forced vibration simulations of the rotating
system, the torsional vibration signal captured the blades vibration signatures with
varying sensitivity depending on blade(s) setting angle, the amplitude of excitation, the
shaft torsional rigidity and the drive mass moment of inertia. The least sensitivity is
shown when the setting angle approaches 908 and the shaft torsional rigidity is
infinitely high (approaching rigid).

4. Conclusions
To investigate the feasibility of extracting rotating blades vibration from the shaft
torsional vibration measurement, a mathematical model is developed and simulated in
this study. The model considers n-blades attached radially to a rigid disk with setting
angles, the disk is driven by shaft that is flexible in torsion and the shaft-disk blades
system is rotating at constant speed. The Lagrangian dynamics is employed in
deriving the equations of motion in conjunction with the assumed modes method to
descritize blades flexibility. The model accounted for the blades axial shortening due to
blades bending deflection in the form of the associated stiffening effect. The model is
nonlinear and coupled second order ordinary differential equations with a size
JQME
12,3

290

Figure 7.
Shaft torsional vibration
frequency spectrum for a
rotor rotating at speed
v ¼ 100 Hz
Blades condition
monitoring

291

Figure 8.
Shaft torsional vibration
frequency spectrum for a
rotor rotating at speed
v ¼ 100 Hz
JQME depending on the number of blades and the number of bending modes for each blade
12,3 considered.
The blades free vibration simulation results showed that the shaft torsional
vibration signal carries the blades vibration signatures at their respective natural
frequencies. In addition the effect in blades natural frequencies due to difference in
properties was captured by the torsional vibration spectrum. Less sensitivity is shown
292 for blades with setting angle approaching 908. The forced vibration simulations of the
rotating system showed that the torsional vibration spectrum contains the individual
blades forced vibration frequencies with magnitude depending on the setting angle, the
amplitude of blades’ vibration, the shaft torsional rigidity and the drive inertia.
The proposed mathematical model and simulation results showed that the approach
of measuring the rotating blades vibration using the shaft torsional vibration
measurement is feasible provided that sensitive and reliable torsional vibration pickup
is used.

References
Al-Bedoor, B.O. (1999), “Dynamic model of coupled shaft torsional and blade bending in rotors”,
Computer Methods in Applied Mechanics and Engineering, Vol. 169, pp. 177-90.
Al-Bedoor, B.O. (2001), “Reduced-order nonlinear dynamic model of coupled shaft-torsional and
blade-bending vibrations in rotors”, ASME Journal of Engineering for Gas Turbines and
Power, Vol. 123 No. 1, pp. 82-8.
Al-Bedoor, B.O. (2002), “Blade vibration measurement in turbo-machinery: the current status”,
The Shock and Vibration Digest, Vol. 34 6, November, pp. 455-61.
Al-Bedoor, B.O. and Al-Nassar, Y.N. (2001), “ANSYS finite element analysis of multi-bladed
rotors with flexible shafts”, King Fahd University of Petroleum and Minerals, unpublished
work.
Cookson, R.A. and Bandyopadhyay, P. (1980), “A fibre-optic Laser-Doppler probe for vibration
analysis of rotating machines”, Transactions of ASME, Journal of Power, Vol. 102, July,
pp. 607-12.
Fan, Y.C., Ju, M.S. and Tsuei, Y.G. (1994), “Experimental study on vibration of a rotating blade”,
Transactions of the ASME, Journal of Engineering for Gas Turbines and Power, Vol. 116,
July, pp. 672-7.
Kaufman, P. and Kershisnik, M. (1984), “Case history – a high speed turbine wheel”, Mechanical
Engineering, July, pp. 38-45.
Laws, W.C. and Muszynska, A. (1987), “Periodic and continuous vibration monitoring for
preventive/predictive maintenance of rotating machinery”, Transactions of the ASME,
Journal of Engineering for Gas Turbine and Power, Vol. 109, April, pp. 159-67.
Maynard, K.P. and Trethewey, M. (1999), “On the feasibility of blade crack detection through
torsional vibration measurements”, Proceedings of the 53rd Meeting of the Society for
Machinery Failure Prevention Technology, Virginia Beach, VA, April 19-22, pp. 451-9.
Maynard, K.P. and Trethewey, M. (2001), “Application of torsional vibration measurement to
blade and shaft crack detection in operating machinery”, Proceedings of the Maintenance
and Reliability Conference Gatlinburg, Tennessee, May 6-9.
Maynard, K.P., Lebold, M., Groover, C. and Trethewey, M. (2000), “Application of double
re-sampling to shaft torsional vibration measurement for the detection of blade natural
frequencies”, Proceedings of the 54th Meeting of the Society for Machinery Failure
Prevention Technology, Virginia Beach, VA, May 1-4, pp. 87-94.
Muszynska, A., Goldman, P. and Bently, D.E. (1992), “Torsional/lateral vibration cross-coupled Blades condition
responses due to shaft anisotropy: a new tool in shaft crack detection”, Proceedings of the
IMechE Conference on Vibrations in Rotating Machinery, C 432-090, Bath, UK, pp. 257-62. monitoring
Nava, P., Paone, N., Rossi, G.L. and Tomasinin, E.P. (1994), “Design and experimental
characterisation of a non-intrusive measurement system of rotating blade vibration”,
Transactions of the ASME, Journal of Engineering for Gas Turbines and Power, Vol. 116,
July, pp. 657-62.
293
Reihardt, A.K., Kadambi, J.R. and Quinn, R.D. (1995), “Laser vibrometery measurements of
rotating blade vibrations”, Transactions of the ASME, Journal of Engineering for Gas
Turbines and Power, Vol. 117, July, pp. 484-8.
Scalzo, A.J., Allen, J.M. and Antos, R.J. (1986), “Analysis and solution of a non-synchronous
vibration problem in the last row turbine blade of a large industrial turbine”, Transaction
of the ASME, Journal of Engineering for Gas Turbines and Power, Vol. 108, October,
pp. 591-8.
Simmons, H.R. and Smalley, A.J. (1990), “Effective tools for diagnosing elusive turbo-machinery
dynamics problems in the field”, Transactions of the ASME, Journal of Engineering for
Gas Turbines and Power, Vol. 112, October, pp. 470-6.
Srinivasan, A.V. and Cuts, D.G. (1984), “Measurement of the relative vibratory motion at the
shroud interfaces of a fan”, ASME Journal of Vibration, Acoustics, Stress and Reliability in
Design, Vol. 106, April, pp. 189-97.
Thompson, W.T. and Dahleh, M.D. (1998), Theory of Vibration with Applications, 5th ed.,
Prentice-Hall, Englewood Cliffs, NJ.
Vance, J.M. (1988), Rotor Dynamics of Turbo-machinery, John Wiley & Sons, Chichester.

Corresponding author
B.O. Al-Bedoor can be contacted at: albedoor@ju.edu.jo or bobedoor@kfupm.edu.sa

To purchase reprints of this article please e-mail: reprints@emeraldinsight.com


Or visit our web site for further details: www.emeraldinsight.com/reprints

Das könnte Ihnen auch gefallen