You are on page 1of 300

Mircea Rade

Dynamics
of Machinery
III

2008
Preface

This textbook is based on the third part of the Dynamics of Machinery


lecture course given since 1993 to students of the English Stream in the
Department of Engineering Sciences (D.E.S.), now F.I.L.S., at the University
Politehnica of Bucharest. It grew in time from a postgraduate course taught in
Romanian between 1985 and 1990 at the Strength of Materials Chair and continued
within the master course Safety and Integrity of Machinery until 2007.
Dynamics of Machinery, as a stand alone subject, was first introduced in
the curricula of mechanical engineering at D.E.S. in 1993. To sustain it, we
published Dynamics of Machinery in 1995, followed by Dinamica sistemelor
rotor-lagre in 1996 and Rotating Machinery in 2003.
The course aims to: a) increase the knowledge of machinery vibrations; b)
further the understanding of dynamic phenomena in machines; c) provide the
necessary physical basis for the development of engineering solutions to machinery
problems; and d) make the students familiar with machine condition monitoring
techniques and fault diagnosis.
As a course taught for non-native speakers, it has been considered useful to
reproduce, as language patterns, full portions from English texts. For the students
of F.I.L.S., the specific English terminology is defined and illustrated in detail.
Basic rotor dynamics phenomena, simple rotors in rigid and flexible
bearings as well as the rotor dynamic analysis tools are presented in the first part.
Finite element modeling of rotor-bearing systems, hydrodynamic bearings, seals
and floating ring bearings are treated in the second part. This third part is devoted
to the analysis of rolling element bearings, gears, vibration measurement for
machine condition monitoring and fault diagnosis, standards and recommendations
for vibration limits, balancing of rotors as well as elements of the dynamic analysis
of reciprocating machines and piping systems. No reference is made to the
vibration of discs, impellers and blades.

May 2008 Mircea Rade


Prefa

Lucrarea se bazeaz pe partea a treia a cursului de Dinamica mainilor


predat din 1993 studenilor Filierei Engleze a Facultii de Inginerie n Limbi
Strine (F.I.L.S.) la Universitatea Politehnica Bucureti. Coninutul cursului s-a
lrgit n timp, pornind de la un curs postuniversitar organizat ntre 1985 i 1990 n
cadrul Catedrei de Rezistena materialelor i continuat pn n 2007 la cursurile de
masterat n specialitatea Sigurana i Integritatea Mainilor. Capitole din curs au
fost predate din 1995 la cursurile de studii aprofundate i masterat organizate la
Facultatea de Inginerie Mecanic i Mecatronic.
Dinamica mainilor a fost introdus n planul de nvmnt al F.I.L.S. n
1993. Pentru a susine cursul, am publicat Dynamics of Machinery la U. P. B. n
1995, urmat de Dinamica sistemelor rotor-lagre n 1996 i Rotating Machinery
n 2005, ultima coninnd materialul ilustrativ utilizat n cadrul cursului.
Cursul are un loc bine definit n planul de nvmnt, urmrind: a)
descrierea fenomenelor dinamice specifice mainilor; b) modelarea sistemelor
rotor-lagre i analiza acestora cu metoda elementelor finite; c) narmarea
studenilor cu baza fizic necesar n rezolvarea problemelor de vibraii ale
mainilor; i d) familiarizarea cu metodele de supraveghere a strii mainilor i
diagnosticare a defectelor.
Fiind un curs predat unor studeni a cror limb matern nu este limba
englez, au fost reproduse expresii i fraze din lucrri scrise de vorbitori nativi ai
acestei limbi. Pentru studenii F.I.L.S. s-a definit i ilustrat n detaliu terminologia
specific limbii engleze.
n prima parte se descriu fenomenele de baz din dinamica rotorilor,
rspunsul dinamic al rotorilor simpli n lagre rigide i lagre elastice, precum i
principalele etape ale unei analize de dinamica rotorilor. n partea a doua se
prezint modelarea cu elemente finite a sistemelor rotor-lagre, lagrele
hidrodinamice, etanrile i lagrele cu inel flotant. n aceast a treia parte se
trateaz lagrele cu rulmeni, echilibrarea rotoarelor, msurarea vibraiilor pentru
supravegherea funcionrii mainilor i diagnosticarea defectelor, standarde i
recomandri privind limitele admisibile ale vibraiilor mainilor, precum i
elemente de dinamica mainilor cu mecanism biel-manivel i vibraiile
conductelor aferente. Nu se trateaz vibraiile paletelor, discurilor paletate i ale
roilor centrifugale.

Mai 2008 Mircea Rade


Contents

Preface i
Contents iii
8. Rolling element bearings 1
8.1 Rolling-element radial bearings 1
8.2 Kinematics of rolling bearings 3
8.2.1 Basic assumptions 3
8.2.2 Simple kinematic relations for angular contact ball bearings 4
8.2.3 Primary rolling element bearing frequencies 6
8.2.4 Kinematic relations for tapered roller bearings 7
8.2.5 General kinematic relations 8
8.3 Structural frequencies 9
8.4 Bearing mechanical signature 10
8.5 Rolling element bearing damage 13
8.5.1 Primary damage 14
8.5.2 Secondary damage 14
8.5.3 Other damages 15
8.6 Time domain bearing diagnostic methods 16
8.6.1 Time-waveform indices 16
8.6.2 Crest factor 17
8.6.3 Amplitude probability density 18
8.6.4 Statistical moments 21
8.6.5 Kurtosis 22
8.7 Frequency domain bearing diagnostics methods 23
8.7.1 Band-pass analysis 24
8.7.2 Spike energy 25
8.7.3 Envelope detection 28
8.7.4 Shock Pulse Method 30
8.8 Cepstrum analysis 35
iv FINITE ELEMENT ANALYSIS

References 36

9. Gears 39
9.1 Gear types 39
9.2 Gear tooth action 40
9.3 Gear vibrations 45
9.3.1 Tooth engagement 45
9.3.2 Effect of tooth deflection 46
9.3.3 Effect of tooth wear 47
9.3.4 Ghost components 48
9.3.5 Modulation effects 48
9.3.6 Resonance effects 53
9.4 Gear errors 54
9.5 Gear faults 55
9.5.1 Wear effects 55
9.5.2 Effects of fatigue 56
9.5.3 Tooth fracture 58
9.6 Gear condition monitoring 58
9.6.1 Vibration signal processing 59
9.6.2 Condition indicators 61
9.6.3 Oil debris analysis 67
9.7 Cepstrum analysis 69
9.8 Time-frequency analysis 72
References 72

10. Vibration measurement 75


10.1 General considerations 75
10.2 Measurement locations 76
10.2.1 General criteria 76
10.2.2 Shaft precession 77
10.2.3 Casing vibrations 78
10.3 Measured parameters 79
10.3.1 Measurement of rotor precession 80
10.3.2 Measurement on bearings 81
10.3.3 Displacement, velocity or acceleration 81
CONTENTS v

10.3.4 Peak-to-peak vs. r.m.s. 82


10.4 Transducers and pickups 85
10.4.1 Transducer selection 85
10.4.2 Eddy current proximity transducers 88
10.4.3 Velocity pickups 91
10.4.4 Accelerometers 94
10.4.5 Summary about transducers 96
10.4.6 Placement of transducers 98
10.4.7 Instrumentation 100
10.5 Data reduction 101
10.5.1 Steady state vibration data 101
10.5.2 Transient vibration data 108
References 112

11 Condition monitoring and fault diagnostics 115


11.1 Machine deterioration 115
11.2 Machine condition monitoring 116
11.2.1 General considerations 116
11.2.2 Maintenance strategies 117
11.2.3 Factors influencing maintenance strategies 119
11.3 Diagnosis process 120
11.4 Fault diagnostics 121
11.4.1 Unbalance 121
11.4.2 Misalignment and radial preload 123
11.4.3 Fluid induced instabilities 127
11.4.4 Rotor-to-stator rubbing 130
11.4.5 Mechanical looseness 135
11.4.6 Cracked shafts 138
11.5 Problems of specific machines 141
11.5.1 Centrifugal equipment 141
11.5.2 Bladed machines 145
11.5.3 Electrical machines and gears 151
11.5.4 Reciprocating compressors 152
Annex 11.1 Shaft alignment 155
References 159
vi FINITE ELEMENT ANALYSIS

12 Vibration limits 163


12.1 Broadband vibration standards and guidelines 163
12.2 Vibration severity charts 164
12.3 Vibration limits for nonrotating parts 168
12.3.1 General guidelines 168
12.3.2 Steam turbine sets 169
12.3.3 Coupled industrial machines 170
12.3.4 Gas turbine sets 172
12.3.5 Hydraulic machines 172
12.3.6 Reciprocating machines 174
12.4 Vibration limits for rotating parts 176
12.4.1 General guidelines 176
12.4.2 Steam turbine sets 177
12.4.3 Coupled industrial machines 178
12.4.4 Gas turbine sets 180
12.4.5 Hydraulic machine sets 181
12.4.6 Selection of measurements 183
12.5 Gear units 185
12.6 API Standards 186
12.7 Industrial buildings 187
12.7.1 Vibration intensity 188
12.7.2 Limits based on vibration velocity 190
Annexes 192
References 199

13 Balancing of rotors 203


13.1 The mass unbalance 204
13.1.1 Definitions 204
13.1.2 Static unbalance 205
13.1.3 Couple unbalance 205
13.1.4 Quasi-static unbalance 206
13.1.5 Dynamic unbalance 207
13.1.6 Static vs dynamic unbalance 207
13.2 Single plane balancing 208
13.2.1 Vector balancing 208
13.2.2 Influence coefficient method 209
CONTENTS vii

13.2.3 Three-trial-mass method 215


13.3 Two-plane balancing 217
13.3.1 Influence coefficient method 217
13.3.2 Resolution into static and couple unbalance 223
13.4 Unbalance tolerances 225
13.4.1 Permissible residual unbalance 225
13.4.2 Balance quality grades 225
13.4.3 Classification of rigid rotors 226
13.5 Multiplane flexible rotor balancing 229
13.5.1 Balancing in N+2 planes 229
13.5.2 Modal balancing 232
13.5.3 General remarks 234
References 235

14 Reciprocating machines 237


14.1 Single cylinder engines 237
14.1.1 Gas pressure excitation 237
14.1.2 Inertia effects 239
14.1.3 Kinematics of crank mechanism 241
14.1.4 Connecting rod and equivalent two-mass system 242
14.1.5 Unbalance of a single cylinder engine 243
14.2 Multi cylinder engines 246
14.2.1 Unbalance forces and couples 246
14.2.2 Othe vibration sources 250
14.2.3 Fault diagnosis of a diesel engine 251
14.3 Reciprocating compressors and piping systems 256
14.3.1 Compressor-manifold system 256
14.3.2 Excitation forces 258
14.3.3 Pulsation analysis 261
14.3.4 Piping vibration 274
References 284
Index 287
8.
ROLLING ELEMENT BEARINGS

This chapter presents the vibration characteristics of rolling element


bearings, and techniques for detecting bearing damage.

8.1 Rolling element radial bearings

The four essential parts of a ball bearing are shown in Fig. 8.1. These are
the inner ring, the outer ring, the balls or rolling elements and the cage (separator,
retainer).

Fig. 8.1 (from [8.1])

The inner ring is mounted on the shaft and rotating with it. There is a track
for the rolling elements incorporated in this ring. For most applications, the outer
ring is mounted in a housing and usually fixed. It also contains a track for the
2 DYNAMICS OF MACHINERY

rolling elements. In some instances, both races rotate. The cage connects the rolling
elements and keeps an equal spacing between them. It rotates about the shaft. The
rolling elements are moving with the cage between the races.
Generally, rolling elements rotate around their axes and simultaneously
they orbit round the bearing axis. If pure rolling motion is considered, the absolute
motion can be seen as the sum of a transport motion with the cage and a relative
spinning motion with respect to the cage. In addition, a certain degree of sliding
occurs on the raceways, called skidding. In ball bearings with zero contact angle, a
ball may have a rotational sliding normal to the contact surface. At the same time,
the ball can have another kind of motion due to gyroscopic moments. If the roller
axis does not coincide with the rolling axis, a slight skew of the roller in roller
bearings may exist. Other motions may occur due to the misalignment of the two
raceways.
The kinematics of rolling bearings is influenced by structural parameters,
operating conditions, lubrication and manufacturing accuracy. Higher clearances
and lighter loading can cause internal sliding. Roller bearings used in aircraft
engines are sometimes assembled with out of round outer raceways to yield a
certain amount of preload in the radial direction in order to reduce skidding.

a b
Fig. 8.2 (from [8.2])

According to the shape of the rolling element, there are ball bearings and
roller bearings. Figure 8.2,a shows an angular-contact ball bearing while Fig. 8.2,b
illustrates a tapered roller bearing. For the latter, the inner ring is called the cone,
and the outer ring is called the cup.
8. ROLLING ELEMENT BEARINGS 3

8.2 Kinematics of rolling bearings

The main bearing elements have characteristic rotational frequencies at


which vibration energy is produced by the periodic impact of a defect. Theoretical
estimates of these frequencies can be determined assuming a perfect geometry.
This means: a) outer and inner bearing races are perfectly circular; b) all balls are
perfectly spherical and of equal diameter; c) perfect alignment of the inner and
outer races. In practice this is rarely the case and it is common to find additional
frequency components generated by errors such as lobing, ovality and ball diameter
differences.

8.2.1 Basic assumptions

In order to determine the angular speeds of the components of rolling


bearings, the following assumptions are made: a) bearing elements are rigid
(contact deformations are neglected); b) rolling elements have pure rolling motions
on raceways (sliding neglected) so that the linear velocities at the contact points of
a rolling element and a raceway are identical; c) radial clearances are neglected; d)
the effect of lubrication is ignored [8.3].

Fig. 8.3

Figure 8.3 shows an angular contact ball bearing. The index i is for the
inner ring, o for the outer ring, B for the ball, and m for the cage. Dm is the
pitch diameter, DB is the ball diameter, Di is the diameter of the inner contact
circle, and Do is the diameter of the outer contact circle, is the contact angle
4 DYNAMICS OF MACHINERY

( = 0 for radially loaded deep groove ball bearings). Symbols n i , no and nB


represent the rotational speeds of the inner ring, outer ring and ball. Clockwise
rotations are considered positive.

8.2.2 Simple kinematic relations for angular contact bearings

The linear velocity of the outer raceway at the contact points is

Do no Dm DB D
vo = o = + cos = no Dm 1 + B cos . (8.1)
2 30 2 2 60 Dm

The linear velocity of the inner raceway at the contact points is

Di n i Dm
D D
vi = i = B cos = n i Dm 1 B cos . (8.2)
2 30 2 2 60 Dm

The linear velocity at the center of rolling elements is equal to the mean
of the outer and inner raceway velocities at contact points (Fig. 8.3)

vo + vi D D
vm = = no Dm 1 + B cos + n i Dm 1 B cos . (8.3)
2 120 Dm 120 Dm

The linear velocity of the cage pitch circle is



vm = nm Dm . (8.4)
60
Equating the two equations, the rotational speed of the cage is derived as

1 DB D
nm = no 1 + cos + n i 1 B cos . (8.5)
2 Dm Dm

The rotational speed of the cage relative to the inner ring is equal to the
difference between the absolute rotational speed of the cage and that of the inner
ring
n D
nmi = nm n i = r 1 + B cos , (8.6)
2 Dm
where nr is the relative rotational speed between the outer and the inner races
nr = no n i . (8.7)
8. ROLLING ELEMENT BEARINGS 5

The rotational speed of the outer ring relative to the cage is


nr D
nom = no nm = 1 B cos . (8.8)
2 Dm

Fig. 8.4

The rotational speed of a rolling element around its own axis can be
obtained blocking the cage (nm = 0) . If vm = 0 , then

nm i = n i , no m = no . (8.9)

Equating the linear velocities vi = vo (Fig. 8.4) yields



vi = n i Di = vo = n B DB ,
60 60
so that
n m i Di = n B DB

and
Di
nB= n mi . (8.10)
DB
Similarly
Do
nB= n om . (8.11)
DB
The rotational speed of the rolling element is

1 Dm D D
nB = nr 1 B cos 1 + B cos ,
2 DB Dm Dm
6 DYNAMICS OF MACHINERY

nr Dm DB
2
nB = 1 cos . (8.12)
2 DB Dm

8.2.3 Primary rolling element bearing frequencies

Let Z be the number of rolling elements.


The impact rate for an inner race defect is equal to Z nmi , the number of
rolling elements passing a given point on the inner ring per minute

Z D
Z nm i = nr 1 + B cos . (8.13)
2 Dm

The impact rate for an outer race defect is equal to Z no m , the number of
rolling elements passing a given point on the outer ring per minute

Z D
Z no m = nr 1 B cos . (8.14)
2 Dm
The impact rate (per minute) for a ball defect is 2 nB , because the ball
defect strikes two surfaces (inner and outer races) in one revolution.
For a stationary outer ring, the impact rate for a cage defect is no m .

Expressing impact rates per second as frequencies f = n 60 [Hz], one


obtains
Z D
outer race ball pass frequency fo = f r 1 B cos ; (8.15)
2 Dm

Z D
inner race ball pass frequency fi = f r 1 + B cos ; (8.16)
2 Dm

D D
2
ball defect frequency f B = fr m 1 B cos ; (8.17)
DB Dm


cage defect frequency

1 no DB ni D
fc = 1 + cos + 1 B cos . (8.18)
2 60 Dm 60 Dm
8. ROLLING ELEMENT BEARINGS 7

Note that the above relations are approximate, assuming pure rolling
motion and neglecting sliding motions. For normal speeds, these defect frequencies
are usually less than 500 Hz. Amplitude modulations especially at the shaft
rotational frequency can produce sum and difference sidebands.

Example 8.1
A radial-thrust ball bearing type 46305, GOST 831-54 mounted on a shaft
with the rotational speed n i = 1000 rpm , has the following geometry:

ball diameter DB = 14.3 mm , pitch diameter Dm = 77.5 mm , contact


angle = 26o , number of balls Z = 10 [8.4].
From the formulas for bearing frequencies (8.15)-(8.18) we obtain:

f c = 6.99 Hz , f B = 30.72 Hz , f o = 69.9 Hz , fi = 97.1 Hz .

Example 8.2
A radial ball bearing type SKF6211, mounted on a shaft with the
rotational speed n i = 3000 rpm , has the following geometry:

ball diameter DB = 25 mm , pitch diameter Dm = 62 mm , contact angle


= 0 , number of balls Z = 10 .
The bearing frequencies (8.15)-(8.18) are:

f c = 20 Hz , f B = 260 Hz , f o = 205 Hz , fi = 295 Hz .

8.2.4 Kinematic relations for tapered roller bearings

Let be the taper angle and the contact angle. Denote

1
K1 = [ tan ( ) tan ] tan 1 ( ),
2
(8.19)
1
K 2 = [ tan ( ) + tan ] tan 1 ( ).
2
Dm - the pitch diameter and DR - the roller diameter.
When the two rings rotate in the same direction, we obtain the following
speeds
8 DYNAMICS OF MACHINERY

the cage speed nm = n i K1 + no K 2 , (8.20)

the cage speed relative to the inner ring ( )


nm i = no n i K 2 , (8.21)

the outer ring speed relative to the cage no m = ( no n i ) K1 , (8.22)

the roller spin speed (


nR = no n i ) 2DDm K1K 2 . (8.23)
R

Note that the relations for angular contact ball bearings can be obtained
from equations (8.20)-(8.23) by substituting DR = DB and

1 D 1 D
K1 = 1 B cos , K2 = 1 + B cos . (8.24)
2 2
Dm Dm

8.2.5 General kinematic relations

More accurate kinematic relations for rolling bearings can be obtained


taking into account the effect of Hertzian deformations, spinning and sliding ball
motions, radial clearances and elastohydrodynamic lubrication [8.5]. Rolling
element bearings are statically-indeterminate, nonlinear, elastic systems whose
motion is influenced by structural parameters, operating environment, lubrication
condition and manufacturing accuracy.
Mathematical models for the ball motion have been developed,
considering either three or five degrees of freedom [8.6]. Balls may have a
rotational sliding normal to the contact surface, called spinning, if its contact angle
is not zero. At the same time, balls have another type of motion due to gyroscopic
moments. In radial roller bearings a slight skew of rollers may exist, i.e. the roller
axis may not coincide with the rolling axis. Internal sliding is more serious in
rolling bearings with high clearances and relatively low external loads. Skidding is
sometimes reduced by intentional radial preload obtained with out-of-round outer
raceways.
Early quasi-static analyses of unlubricated roller bearings were based on
the assumption of Coulomb friction in the race contacts [8.7], [8.8]. The friction
forces resulting from interfacial slip at the ball-race contacts have been included in
the dynamic analysis of the elastically constrained bearing. Elastohydrodynamic
lubrication effects have been introduced later [8.9] and incorporated in more
accurate dynamic analyses [8.10].
More elaborate models have been developed to simulate distributed
defects such as off-size rolling elements, misaligned and out-of-round components
[8.11]. Their description is beyond the aim of this presentation.
8. ROLLING ELEMENT BEARINGS 9

8.3 Structural vibrations

The natural frequencies of the free bearing elements can be calculated


theoretically as [8.12]
race natural frequency

fn =
(
k k 2 1 )1 EI
[Hz] , (8.25)
2 2
2 k +1 a m

where k is the number of waves around circumference ( k = 2 , 3, 4 ) , a is the radius


to neutral axis, I is the moment of inertia of cross-section, E is Youngs modulus,
and m is the mass of race per linear length;
ball natural frequency

0.848 E
fBn = [Hz] , (8.26)
DB 2

where DB is the ball diameter and is the density of the ball material.
These are the free natural frequencies of individual elements. It is
difficult to estimate how these frequencies are affected by assembly into a full
bearing and mounted in a housing. However it is indicated that resonances are not
altered significantly. Resonance of the ball is usually far above the range of
vibration analysis and can be ignored.
The outer ring resonance can be excited by the rotating balls (rollers).
They deform the race into a flexural pattern (with a number of wavelengths equal
to the number of rolling elements) which rotates with the ball passing frequency. It
can also be produced by the waving motion of the balls around their theoretical
circumferential path.
In rolling bearings the external load is carried by a finite number of
rolling elements. Their number under load varies with the angular position of the
cage. The elastic deflection produced by the Hertzian contact under load varies
with the position of the rolling element relative to the line of load. This gives rise
to a periodical variation of the total stiffness of the bearing assembly and generates
the so-called varying compliance vibrations of the rotor [8.13].
Their fundamental frequency is equal to the ball (or roller) passage
frequency over the outer ring. Higher harmonics are also excited, to a degree
decreasing with their order, mainly due to deviations of the bearing parts from the
perfect geometric shape. The magnitude of shaft movements is a function of the
external load, number of rolling elements, radial clearance and the local stiffness
10 DYNAMICS OF MACHINERY

between rolling element and tracks, as given by the Hertzian theory for elastic
contacts (H. Hertz, 1881).
The parametrically excited vibrations of the rotor-bearing system, with
strongly coupled vertical and horizontal movements, are described by nonlinear
equations of motion with time varying coefficients. Variable contact compliance
vibrations are of importance only at frequencies in the neighborhood of the
rotational frequency of the bearing, and are generally of appreciable magnitude
only for rather high radial loads.
Structural resonances can also be excited by other distributed defects
such as race misalignment or eccentricity, lack of roundness, waviness of the
rolling surfaces and unequal ball diameters produced during the manufacturing
process. These distributed defects often give rise to excessive contact forcers
which in turn result in premature surface fatigue and ultimate failure.
Note that waviness defines relatively widely-spaced surface irregularities.
In principle, surface roughness is the same type of geometrical imperfection as
waviness. Their distinguishing characteristic is the spacing of irregularities, which
is finer for surface roughness. Waviness is used to imply irregularities up to an
order of 200 waves per circumference, while surface roughness contains waves of a
much higher order. Typical examples are the following: at a frequency of 300 Hz,
the inner ring has 16 to 17 waves per circumference, and the outer ring has 24 to
27. At a frequency of 1800 Hz, the inner ring has 94 to 101 waves per
circumference, and the outer ring has 147 to 166 [8.14].
Geometrical irregularities in the form of a waviness with a few cycles
around the circumference give rise to low frequency vibrations. The vibrations of
radially loaded bearings with stationary outer rings and positive radial clearances
are primarily related to the inner race waviness and varying roller diameter, rather
to other geometrical errors. The vibrations due to non-uniform roller diameters
occur at cage speed harmonics, while vibrations due to inner race waviness occur at
shaft speed harmonics with a side band spaced with the roller passage frequency
occurring at the high harmonics [8.15].

8.4 Bearing mechanical signature

The vibration signal produced by a rolling element bearing, as measured


by an accelerometer or other motion transducer, can be electronically broken into
its frequency components and their related amplitude levels. This plot of the
narrow-band spectrum of the vibration signal is called the mechanical signature
of the ball bearing, since it identifies the bearing and is unique to the unit selected.
Figures 8.5 and 8.6 are examples of mechanical signatures of two
different ball bearings. Many of the discrete frequencies contained in the
8. ROLLING ELEMENT BEARINGS 11

mechanical signature can be related to the specific mechanical defects within the
bearing. The amplitudes of these peaks are a measure of the energy transmitted by
impacts and, therefore, of the smoothness of the bearing operation. Peaks generated
by unbalance, misalignment and other sources have to be distinguished from
bearing generated peaks.

Fig. 8.5 (from [8.16])

Fig. 8.6 (from [8.16])


12 DYNAMICS OF MACHINERY

A comparison of the mechanical signatures of two ball bearings of the


same type would require data obtained at the same speed, since most of the
vibration frequencies are proportional to speed. Rather than trying to hold speed
constant, it was found better to have mechanical signatures independent of speed.
This is accomplished by normalizing all frequencies relative to the fundamental
rotational speed. The procedure is called order normalization. For stationary
outer ring, the fundamental frequency of rotation is that of the inner ring. The
spectra in Figs. 8.5 and 8.6 are plotted versus frequency orders.

Fig. 8.7 (from [8.16])

The mechanical signature of a good bearing is shown in Fig. 8.7. The


amplitude is calibrated for 90dB equal to 0.26 g. The noise floor is approximately
50dB or 0.0026 g. The first order is the only frequency evident in this spectrum.
The amplitude of the spectrum is plotted in log scale to provide the greatest vertical
magnification. This allows the detection of small defect frequencies in a
measurement containing a large frequency component. Otherwise the random noise
due to friction may dominate the spectrum making it difficult to locate frequencies
that can be correlated with bearing defects. A spectrum averaging technique can be
applied to enhance the signal-to-noise ratio of the periodic discrete frequencies
generated by the ball bearing.
A mechanical signature showing a ball defect is illustrated in Fig. 8.8.
The presence of two large orders (5.80 and 1.00) generates sum and difference
frequencies that can be identified at 5.80 1.00 and 5.80 2.00 . This bearing also
shows orders associated with inner race defects that can be explained by a non-
linear (N.L.) theory taking into account race waviness, eccentricity and large ball
diameter variations.
8. ROLLING ELEMENT BEARINGS 13

Fig. 8.8 (from [8.16])

Generally, the outer race geometrical imperfections produce a vibration


spectrum having peaks at the harmonics of the outer race defect frequency, with
side bands spaced with the cage frequency. The inner race surface irregularities
produce a spectrum having peaks at the harmonics of the inner race defect
frequency. The side bands are spaced with an interval related to the cage frequency
and the shaft running frequency.

8.5 Rolling element bearing damage

Each of the different causes of bearing failure inadequate or unsuitable


lubrication, careless handling, ineffective sealing, incorrect fits, etc. produces its
own characteristic damage. Such damage, known as primary damage, can be wear,
indentations, smearing, surface distress, corrosion and electric current damage.
Primary damage gives rise to secondary, failure-inducing damage flaking
and cracks. A failed bearing frequently displays a combination of primary and
secondary damage [8.17].
The local defects, including cracks, pits and spalls, give rise to impulsive
contacts between the bearing elements. These impulsive contacts produce
vibrations and noise, which can be monitored to detect the presence of a defect in
the bearing.
14 DYNAMICS OF MACHINERY

8.5.1 Primary damage

Wear
Wear may occur as a result of the ingress of foreign particles into the
bearing or when the lubrication is unsatisfactory. It may occur also in bearings
exposed to vibrations while not running, damage known as false brinelling.
Indentations
Indentations in raceways and rolling elements occur when the bearing,
while not running, is subjected to abnormally heavy loading in the form of impacts
or pressure. The distance between the dents is the same as the rolling element
spacing. Foreign particles in the bearing also cause indentations.
Smearing
When two inadequately lubricated surfaces slide against each other under
load, material is transferred from one surface to the other. This is known as
smearing and the surfaces concerned become ripped up and look scored. When
smearing occurs, the material is generally heated to such temperatures that
rehardening takes place. This produces localized stress concentrations that may
cause cracking or flaking.
Surface distress
If the lubricant film between raceways and rolling elements becomes too
thin, the peaks of the surface asperities will momentarily come in contact with each
other. Small cracks then form in the surfaces and this is known as surface distress.
These cracks must not be confused with the fatigue cracks that originate beneath
the surface and lead to flaking. These cracks may, however, hasten the formation of
sub-surface fatigue cracks and in that way shorten the bearing life.
Corrosion
Rust will form if water or corrosive agents get into the bearing in such
quantities that the lubricant cannot provide protection for the steel surfaces. This
process will soon lead to deep seated rust that can initiate flaking and cracks.
Fretting corrosion occurs when there is relative movement between bearing ring
and shaft or housing, on account of the fit being too loose.

8.5.2 Secondary damage

Flaking (Spalling)
Bearing life is determined by material fatigue. Fatigue is the result of shear
stresses cyclically appearing just below the load carrying surface. After a time
these stresses cause cracks which gradually extend up to the surface. As the rolling
elements pass over the cracks, fragments of material break away and this is known
8. ROLLING ELEMENT BEARINGS 15

as flaking or spalling. The flaking progressively increases in extent and eventually


makes the bearing unserviceable. The life of a rolling bearing is defined as the
number of revolutions the bearing can perform before incipient flaking occurs.
The causes of premature flaking may be heavier external loading than had
been anticipated, preloading on account of incorrect fits or excessive drive-up a
tapered seating, oval distortion owing to shaft or housing seating out-of-roundness,
axial compression as a result of thermal expansion, misalignment, etc. Flaking may
also be caused by other types of damage, such as indentations, deep seated rust,
electric current damage or smearing.
Cracks
Cracks may form in bearing rings for various reasons. The most common
cause is rough treatment when bearings are being mounted or dismounted (hammer
blows, excessive drive-up on tapered seatings, heating and mounting on shafts with
wrong tolerances). Flaking acts as a fracture notch and may lead to cracking of the
bearing ring.
Cage damage
Cage failures are due to vibrations, excessive speeds, wear and blockage
by flaked material wedged between the cage and a rolling element. Misaligned
rings produce oval ball paths that distort the cage once per revolution leading to
fatigue cracks. The cage is the first component to be affected when the lubrication
becomes inadequate. It is always made of softer material than the other
components of the bearing and consequently it wears comparatively quickly.
Two approaches have been used to study the vibration and acoustic
response of rolling element bearings due to defects in the bearings. One is to run
the bearings until they fail and monitor the changes in their vibration and acoustic
response. Usually the failure is accelerated by overloading, overspeeding, or
starving the bearings of lubricant. The other approach is to intentionally introduce
defects in the bearings by techniques such as acid etching, spark erosion,
scratching, or mechanical indentation. The vibration response of the bearings is
measured and compared with the responses of good bearings.

8.5.3 Other damages [8.14]

Denting is a defect in the raceway resulting from the introduction of


foreign particles which become pressed between the rolling elements and rings.
External debris is foreign matter introduced to the bearing from an external
source.
Glazing is a form of smearing whereby the affected area on the raceway
has a shiny appearance similar to the finish on a new ball. Metal flow has taken
place during this mode of failure.
16 DYNAMICS OF MACHINERY

Grooving shows as continuous circumferential indentation on balls


produced by the balls running on the retaining diameter of the counterbored
raceway.
Brinelling. The term applies to a bearing which has been statically loaded
to an extent such that the raceways and rolling elements are permanently deformed.
A brinelled bearing has indentations in the raceways and often has corresponding
flats on the rolling elements.
Fretting is a corrosive form of wear caused by very slight movement
between two metal surfaces under very high contact pressure. The formation of an
iron-oxide paste between two fretting steel members is not uncommon. It is often
seen between the inner ring and the shaft.
Creeping is a relative movement between the bearing inner ring and the
shaft, caused by inadequate interference fit for the applied load. Creeping is
evidenced by circumferential scoring on the bearing bore and shaft. It may be an
advanced stage of fretting.
Spinning is an advanced stage of creeping. The relative movement between
inner ring and shaft is much greater than in creeping and the sliding surfaces may
become polished. The iron-oxide from the fretting phase may still be present and
assist in further wear.
Discoloration due to temperature indicates operation of the bearing
elements with marginal lubrication or under excessive power conditions.

8.6 Time domain bearing diagnostics methods

The time-history of the vibration signal can be measured to detect defects


in rolling element bearings.

8.6.1 Time-waveform indices

Time-waveform indices are calculated based on the raw vibration signal


and used for trending and comparisons. Examples are the peak level (maximum
vibration amplitude within a given time signal), peak-to-peak amplitude (maximum
positive to maximum negative signal amplitudes), mean level (average vibration
amplitude), and root-mean-square (r.m.s.) level [8.12].
For a sample record x (t ) of duration T, the mean value and the root mean
square value have the following expressions:
8. ROLLING ELEMENT BEARINGS 17

T
1
mean value x=
T x (t )d t ;
0
(8.27)

T
1
root mean square value xr .m .s . = x
2
(t ) d t . (8.28)
T
0

Usual practice is to measure the r.m.s. velocity of the overall vibration


level at the bearing housing. Measured levels are compared with general standards
or with established reference values for each bearing. By plotting the measurement
results over time the trend in vibration can be followed and extrapolated to give a
prediction of when the bearing needs replacement. However, because the overall
vibration level often increases only in the final stages of failure, this method gives
late warnings of failure.
Two time-waveform indices used to get early warnings of the bearing
failure the Crest Factor and the Kurtosis are presented in the following.

8.6.2 Crest Factor

An early warning of bearing failure is obtained measuring the Crest Factor.


The Crest Factor is defined as the ratio of the peak level to the r.m.s. level
of a signal [8.18]

peak level
Crest Factor = . (8.29)
r .m.s . level

The curve in Fig. 8.9 shows a typical trend for the Crest Factor as the
bearing condition deteriorates.
Initially, for a bearing with no faults there is a relatively constant ratio of
about 3.0. As localized faults develop, the resulting impacts increase the peak level
substantially, but have little influence on the r.m.s. level. The peak level will
typically grow to a certain limit. As the bearing condition deteriorates, more spikes
will be generated per ball-pass, finally influencing the r.m.s. level, even though the
individual peak levels are not greater. Towards the end of the bearing life, the crest
factor may have fallen to its original value, even though both peak and RMS levels
have increased considerably.
The best way to trend the data is as illustrated in Fig. 8.9: peak and r.m.s.
levels on the same graph, with Crest Factor inferred as the difference between the
two curves (log scale).
18 DYNAMICS OF MACHINERY

Fig. 8.9 (from [8.19])

Measuring the overall vibration level over a wide frequency range (10 Hz
to 10000 Hz), the method is prone to interference from other vibration sources.

8.6.3 Amplitude probability density

A vibration signal taken near a rolling bearing can be analyzed as a


stationary random signal. Considering a sample record x (t ) of duration T, the
signal is described by the probability with which the signal will take values
between x and x + x (Fig. 8.10). It is equal to the time spent in the window x ,
equal to the sum t1 + t2 + .... + tn , divided by the averaging time T
n


ti
P ( x, x + x ) = . (8.30)
T
i =1

When x 0 and T , one obtains the amplitude probability density


p (x ) , giving the probability to have an amplitude x , plotted on the left of Fig.
8. ROLLING ELEMENT BEARINGS 19

8.10. The bell-shaped curve corresponds to the Gaussian (normal) distribution,


which describes signals occurring in practice with sufficient precision.

Fig. 8.10

Figure 8.11 shows the normalized probability density function


p ( x ) dx = 1

(8.31)

as a function of the dimensionless variable x , where is the r.m.s. value for


zero mean.

Fig. 8.11
20 DYNAMICS OF MACHINERY

It is found that 99.8% of all events occur in the range 3 . From that
follows approximately that the peak value is 3 , which, divided by the r.m.s. value
, gives for the Crest Factor (8.29) a value of 3.0 .
An obvious measure of bearing condition is obtained by observing changes
in the probability at particular amplitude levels, those above 3 providing most
significant information.

Fig. 8.12 (from [8.20])

A typical result for a bearing is shown in Fig. 8.12, where the vertical
logarithmic scale was chosen to enhance the changes at low probability which have
been found important in detection of bearing damage. Endurance tests have been
carried out at constant speed and twice the recommended load, to accelerate fatigue
failure. The overall acceleration level was measured in the frequency range
3Hz 5 kHz . The three curves correspond to increased test durations, expressed in
terms of the bearing life L10 = 50 h .

Note that L10 is defined as the rating life of a group of apparently identical
rolling element bearings, operating under identical loads and speeds, with a 90%
reliability before the first evidence of fatigue develops [8.21]. A fatigue spall of
specific size ( 6 mm 2 ) is usually considered (ISO 281, 2006).
In the early stages of the test, i.e. 0.067 L 10 ( 3.35 h ), when the bearing is
undamaged, the distribution curve is an inverted parabola which indicates a normal
(Gaussian) distribution. With incipient damage at 1.4 L10 ( 70 h ), pronounced
changes occur in the tail of the distribution curves. This is consistent with the
observation made on Fig. 8.9 that the measured peak acceleration level increases
8. ROLLING ELEMENT BEARINGS 21

but the r.m.s. level remains relatively unchanged. With increasing time, i.e. 1.6 L10
and advancing damage, the tail of the distribution curve initially broadens.

8.6.4 Statistical moments

Rather than examining the probability density function in detail, it is often


more informative to examine statistical moments of the data [8.22]. These are
defined by the general integral

Mn =
x p ( x ) dx , (n = 1, 2, 3, ...) .
n
(8.32)

The first two moments are



mean value x=
x p ( x ) dx ,

(8.33)

__

x p ( x ) dx .
2 2
mean square value x = (8.34)

The variance (dispersion) is

( x x ) p ( x ) dx ,
2 2
= (8.35)

where is the standard deviation (mean square error).


Odd moments, i.e. n = 1, 3, 5,, etc., relate information about the position
of the peak density relative to the median value. Even moments, n = 2 , 4, 6 ... , etc.,
indicate the spread in distribution.
Higher moments (n > 2 ) generally have the mean value removed and are
normalized by the standard deviation. The third moment yields

x p (x ) dx
3

Skewness skew (x ) =
, (8.36)
3

and the fourth moment yields


22 DYNAMICS OF MACHINERY

x p (x ) dx
4

Kurtosis kurt ( x ) =
. (8.37)
4
Skewness is a measure of symmetry, or more precisely, the lack of
symmetry. The skewness for a normal distribution is zero.

8.6.5 Kurtosis

The Kurtosis factor is the ratio of the fourth central moment of the
amplitude distribution to the second power of the second central moment.
Kurtosis characterizes the relative peakedness or flatness of a distribution
compared to the normal distribution (Karl Pearson in Biometrika, 1905).
A normal distribution has a Kurtosis of 3 and is called mesokurtic.
Indeed, for a Gaussian distribution

1 (x x ) 2
p (x ) = exp , (8.38)
2 2 2

the fourth statistical moment is


( x x ) 4 exp (x x2)
1 2

( x x ) p ( x ) dx =
4
M4 = dx .
2 2

Denoting
xx
y= , d x = 2 dy ,
2
we obtain

y exp ( y ) dy = 3
4 4 4 2 4
M4 = .

The second statistical moment is

y exp ( y ) dy =
2 2
(xx ) p ( x ) dx =
2 2 2 2
M2 = .


8. ROLLING ELEMENT BEARINGS 23

The Kurtosis results as

kurt ( x ) =
M4
=3.
(M 2 ) 2
A flat distribution with short tails has a Kurtosis value less than 3 and is
called platykurtic. A peaked distribution with longer tails has a Kurtosis value
greater than 3 and is said to be leptokurtic. Higher Kurtosis means that more of the
variance is due to infrequent extreme deviations, as opposed to frequent modestly-
sized deviations.
Kurtosis provides an early warning of surface damage (Dyer and Stewart,
1978). For a good bearing it equals 3. Bearing damage causes an increase in the
impulsive components of the vibration signal, due to impacting. The signals
become more spiky. A damaged bearing exhibits a non-Gaussian probability
distribution with dominant tails which increase the Kurtosis value.
The advantage of Kurtosis, as a parameter for detecting the condition of
rolling element bearings, lies in the finding that it remains close to 3 ( 8% ) for an
undamaged bearing and is insensitive to the load or speed of bearing. One
disadvantage is that the Kurtosis value comes down to the level of an undamaged
bearing (i.e. 3) when the damage is well advanced. Therefore, it has been suggested
to measure Kurtosis in selected frequency bands [8.23].
Experiments have shown that initial damage increases Kurtosis in the
lower frequency bands. As damage spreads, the Kurtosis value begins to decrease
in the first band (2.5 5 kHz ) , while increasing in the other bands. At the end of
the useful life of the bearing, the highest Kurtosis numbers are in the highest
frequency band (40 80 kHz ) [8.24].

8.7 Frequency domain bearing diagnostics methods

The time-domain bearing vibration signal is processed into the frequency


domain by applying a Fast Fourier Transform (FFT) algorithm. The principal
advantage of this format is that the repetitive nature of the vibration signal is
clearly displayed as peaks in the frequency spectrum at the frequencies where the
repetition takes place. This allows for faults, which usually generate specific
characteristic frequency responses, to be detected early, diagnosed accurately, and
trended over time as the condition deteriorates. However, the disadvantage of
frequency-domain analysis is that a significant amount of information (transients,
non-repetitive signal components) may be lost during the transformation process.
24 DYNAMICS OF MACHINERY

Particular emphasis is placed on changes in the frequency spectrum of the


vibration signals. Additional processing techniques are used as an aid to
interpretation of the spectrum, like synchronous time signal averaging and
cepstrum analysis (see Sections 9.6.1 and 9.7).

8.7.1 Band-pass analysis

Frequency spectra obtained from measurements made on bearings, referred


to as mechanical signatures in Section 8.4, are used in fault detection and
diagnosis. For fault detection, current spectra are compared with those obtained
over a period of time to detect changes in spectrum, which denote bearing
deterioration [8.25]-[8.27].

Fig. 8.13 (from [8.19])

Fig. 8.14 (from [8.28])


8. ROLLING ELEMENT BEARINGS 25

Figure 8.13 shows how a developing fault changes the spectrum as a


function of time. The frequency spectrum gives earlier warnings than monitoring of
overall vibration. The level of overall vibration only increases after an increasing
component has become the highest peak in the spectrum.
Whenever an increase of the baseline (reference) level is detected, a further
analysis is carried out for fault diagnosis. The frequency range in which the levels
are exceeded gives an indication of what type of faults to expect.
Band-pass analysis involves filtering the vibration signal above and/or
below specific frequencies in order to reduce the amount of information presented
in the spectrum to a set band of frequencies. These frequencies are typically where
fault characteristic responses are anticipated. Changes in the vibration signal
outside the frequency band of interest are not displayed.
Vibrations produced by machines with rolling element bearings occur in
three frequency regions (Fig. 8.14):
a) the rotor related region, in the range of 1 4 to 3 times the shaft
rotational speed. In the low frequency region, unbalance, misalignment, bent shaft,
mechanical looseness, oil whirl, hysteretic whirl, etc. will be found;
b) the prime spike (element passage) region, normally 1 to 7 times the
element passage rate. In the medium frequency region, indication on wear and
incipient faults in a gearbox will be found, as well as eccentricity, uneven
gearwheels, misaligned gearwheels, etc.;
c) the high frequency (spike energy) region, from 5 kHz to approximately
25 kHz . At very high frequencies, to the megahertz region, measured data may
contain information related to incipient faults in rolling-element bearings, rubs,
cavitation, valve noise, etc.
Guideline alarm threshold values are [8.1]:
7.7 mm s peak for region a,
2.5 to 3.8 mm s peak values for region b,
3 to 4 g peak values for region c.

8.7.2 Spike energy

Spike Energy is a measure of the intensity of energy generated by


repetitive mechanical impacts or pulses that occur as a result of surface flaws or
insufficient bearing lubrication. These impacts tend to excite the resonance
response of machine components. A signal measured near a rolling element bearing
appears as periodic spikes of high-frequency energy and can be measured by
accelerometers [8.29].
26 DYNAMICS OF MACHINERY

For a bearing with fixed outer ring, rotating inner ring and fixed load, Fig.
8.15 shows the signal produced by a defect in the fixed race

Fig. 8.15 (from [8.30])

For a defect in the rotating inner race, it is important to consider the load
distribution around the bearing circumference. This results in a modulation effect
illustrated in Fig. 8.16.

Fig. 8.16 (from [8.30])

When the load is not fixed in space, but rotating as for centrifugal forces,
modulations are also generated for a fixed outer race defect.
The intensity of impact energy is a function of pulse amplitude, pulse rate
and pulse duration. This signal is processed by a Spike Energy detector (IRD
Mechanalysis). A simplified flow chart of the Spike Energy signal processing is
shown in Fig. 8.17.
The vibration signal from an accelerometer is passed through a high
frequency bandpass filter. The purpose of filtering is to reject the normal rotational
8. ROLLING ELEMENT BEARINGS 27

vibration components, such as unbalance and misalignment, while allowing the


vibration generated by impacts to remain. The lower corner frequency, f c , can be
selected between 100 and 5000 Hz, and the upper corner frequency, f d , is
65 kHz .
The filtered vibration signal passes through a peak-to-peak detector with a
properly selected output time constant, which detects and holds the peak-to-peak
values. Then, it decays at the rate of the time constant until the next pulse occurs.
The instrument repeats this process.

Fig. 8.17 (from [8.31])

( )
It is customary to measure accelerations in g units 1 g = 9.81 m s 2 . The
acceleration measured to describe the energy produced by early bearing defects is
measured in gSE units (acceleration units of Spike Energy). These faults produce
a high frequency carrier and modulating sidebands. The carrier is the natural
frequency of the excited bearing component. The modulating sidebands are caused
by load and speed changes. The gSE reading is determined by the intensity of the
high-frequency peaks in the vibration signal. Pulses with large amplitude and high
repetition rate produce high overall gSE readings.
In addition to overall Spike Energy measurements, a Spike Energy
Spectrum can be obtained by fast Fourier transform (FFT) analysis of the signal
from the Spike Energy detector. It is different from the acceleration frequency
spectrum. The components in the gSE spectrum are modulation frequencies that are
related to the high frequency carrier, such as the resonance frequency of the
machine element.
28 DYNAMICS OF MACHINERY

Spike Energy readings can be affected by accelerometers and mounting


methods. The gSE readings can be different if different transducers are used unless
transducers have exactly the same frequency response characteristics. To ensure
the consistency of gSE data, it is necessary to always use the same accelerometer
and the same mounting conditions. Stud mounting is the best.
Spike Energy readings are highly dependent upon the machine size,
configuration and bearing details. Users must go through a learning phase, taking
periodic readings, observing trends, noting failed bearings and building up
historical background before accurate condition assessments are made. As an order
of magnitude, energy alarm levels of 0.5 gSE have been used in an application
with the dryer rolls on a paper machine.

8.7.3 Envelope analysis

Envelope analysis is essentially a signal processing technique that utilizes a


filter and rectification preprocessing of a standard accelerometer signal to reveal
the bearing defect at its fundamental frequency [8.32]. Sometimes it is referred to
as the high frequency resonance technique [8.33].

Fig. 8.18 (from [9.11])

The traditional method uses an analogue bandpass filter plus a rectifier and
a smoothing circuit (Fig. 8.18). The filter extracts the resonance excited by the
bearing fault from the frequency spectrum and the detector detects the envelope.
In modern signal analyzers, zooming around a resonance excited by the
bearing defect extracts the useful part of the frequency spectrum, and then the
Hilbert transform generates the envelope of the time signal. The spectrum of the
envelope is calculated to show the repetition frequency of the fault generated
pulses.
Envelope Detection or Amplitude Demodulation is the technique of
extracting the modulating signal from an amplitude-modulated signal. The result is
the time history of the modulating signal. This signal may be studied/interpreted as
it is in the time domain or it may be subjected to a subsequent frequency analysis.
8. ROLLING ELEMENT BEARINGS 29

Envelope Analysis is done on the FFT frequency spectrum of the modulating


signal.

Fig. 8.19 (from [8.34])

Envelope detection is detailed in Fig. 8.19. The time signal is filtered


around the frequency region where the increase is detected (in the kHz range). This
leaves the high frequency signal which contains the pulse-excited vibration of the
bearing housing without most of the contaminating signals.
This signal is then rectified and low-pass filtered, at a frequency
approximately one half the bandwidth of the bandpass filter. The signal now looks
somewhat like the original pulses from the bearing, but of most significance, we
have thus recreated the pulse frequency.
30 DYNAMICS OF MACHINERY

By analyzing this signal with an FFT analyzer, the pulse frequency can be
determined exactly. Since the impulse rate can be calculated, see equations (8.15)-
(8.18), the source of the fault can be pinpointed. Note that the real frequency will
be slightly lower than the calculated one due to sliding.

Fig. 8.20 (from [8.29])

If the fault is on the rotating race, then it is sometimes possible to see the
amplitude modulation from the varying load on the crack illustrated in Fig. 8.16.
This modulation effect will turn up as sidebands around the lines corresponding to
the pulse rate, spaced at the rotational speed (Fig. 8.20).

8.7.4 Shock Pulse Method

Shock pulse technology was developed by SKF AB, Gothenburg, in the


early 1970s [8.35]. It was prompted by difficulties encountered by techniques
based on the analysis of the repetitive components of the vibration signals from
rolling element bearings.
The method involves the analysis of the high-frequency (ultrasonic) shock
waves generated by metal-to-metal impacts in a rotating bearing, where most of the
information about bearing damage can be found.
8. ROLLING ELEMENT BEARINGS 31

Fig. 8.21 (from [8.14])

Empirical relationships were developed that provided both a measure of


the theoretical lubricant film thickness between the bearing surfaces, as well as the
overall condition of the bearing surface.
The shock pulse analyzer detects impacts of very short duration arising
from the presence of pits and spalls. Unlike conventional vibration analysis, that
monitors a broad vibration band with the objective of detecting discrete
frequencies, the shock pulse method (SPM) measures and evaluates the ultrasonic
frequency band centered around 36 kHz.
32 DYNAMICS OF MACHINERY

Shock (or stress) waves that result from metal-to-metal contact are short
duration bursts of energy that travel at the speed of sound through the material. As
the wave travels, it dissipates energy through the structure, thereby reducing the
wave pulse. The SPM is designed to detect the weak shock pulse signals using an
accelerometer with a natural frequency of about 36 kHz, ideally placed very
closely to the subject bearing. In fact, a patented design called Tandem-Piezo is
used, which enables the accelerometer to accurately measure both shock pulse and
vibration. To distinguish the shock pulses from vibration, a band pass filter around
de 36 kHz shock pulse signal is used. This helps isolate the shock pulse from other
interference created by machinery vibrations.
The last stage of signal processing is the conversion from a waveform to
analog pulses. This process provides a signal that then can be processed to
determine bearing condition.

a b c
Fig. 8.22 (from [8.36])

Figure 8.21 shows the block diagram of an early shock pulse meter [8.14].
The accelerometer output (Fig. 8.22, a) is passed through a high-gain amplifier
tuned at the resonant frequency of the accelerometer, the amplifier acting as a very
sharp band filter. The filtered and amplified shock pulse is shown in Fig. 8.22, b.

Fig. 8.23 (from [8.37])

The signal is rectified, averaged and then passed through a peak-sample-


and-held circuit. This measures the information and displays it on a counter which
records the number of peaks occurring above a defined peak amplitude;
8. ROLLING ELEMENT BEARINGS 33

alternatively, it presents the signal r.m.s. value. The amplitudes of analog shock
pulses are displayed as function of time in Fig. 8.22, c.
The bearing condition is defined by a string of pulses with varying
magnitudes (Fig. 8.23). A shock pulse analyzer measures the shock pulse
magnitude on a decibel scale, in dBsv (decibel shock value). It takes a sample
count of the shock pulses occurring over a period of time and displays: LR (Low
Rate of occurrence), the value for the relatively small number of strong shock
pulses, and HR (High Rate of occurrence), the value for the large number of weak
shock pulses in the pattern. The difference between LR and HR is called the delta
value, .

a b
Fig. 8.24 (after [8.38])
The strength of the individual pulses, and the ratio between stronger and
weaker pulses in the overall pattern, provide the row data for bearing condition
analysis. The magnitude of these pulses is dependent on the bearing surface
condition and the peripheral velocity of the bearing.
In undamaged bearings, the shock level varies with the thickness of the
lubricant film between the rolling elements and raceway. The relationship between
stronger and weaker pulses, however, is only slightly affected (Fig. 8.24, a).
34 DYNAMICS OF MACHINERY

Surface damage causes an increase of up to 1000 times (60 dB) in shock


pulse strength, combined with a distinct change in the ratio between stronger and
weaker pulses (Fig. 8.24, b).

Fig. 8.25 (from [8.37])

The shock pulse readings are evaluated and a code is displayed describing
the general bearing condition.
Code A is for a bearing in good condition. There is no detectable damage
to the surfaces of the load carrying parts, and no extreme lack of lubricant in the
rolling interface. Figure 8.25, a shows a typical shock pulse pattern from a good
bearing: a low shock level and a normal delta value.
Code B indicates a dry running condition, causing a high HR value and a
low delta value (Fig. 8.25, b). Code C is for reduced condition defined by an
increased shock pulse level with a large delta value (Fig. 8.25, c). This denotes
incipient surface damage. Code D is for bearing damage characterized by a high
shock level with a large delta value (Fig. 8.25, d). A contamination of the lubricant
by hard particles causes a similar pattern.

Fig. 8.26 (from [8.37])


8. ROLLING ELEMENT BEARINGS 35

Output data are displayed as in Fig. 8.26. The delta value = LR HR is


plotted as a function of HR. The fields marked A, B, C, D correspond to the
condition codes. The black point marks a shock pulse reading. For a bearing in
good condition it is in the field A.
Developing surface damage causes a marked increase of the delta value,
HR remains low while LR increases. The marker point moves upwards, from A
through field C towards D.
For poor lubrication, the condition code changes from A to B then to D as
damage develops and increases. The marker point moves to the right.
Shock pulse is not limited to determining the condition of rolling element
bearings. Any machine element with continuous metal-to-metal contact gives off
shock pulse signals. Equipment such as gearboxes, lobe compressors, screw
compressors and centrifuges can be monitored using SPM.

8.8 Cepstrum analysis


Cepstrum analysis is a post-processing technique involving a Fourier
transform of a logarithmic frequency spectrum (see definitions in Section 9.7). It is
used to detect and quantify families of uniformly spaced harmonics arising from
periodic added impulses generated by bearing faults.

Fig. 8.27 (after [8.39])


36 DYNAMICS OF MACHINERY

Figure 8.27 gives an example of the development of an outer race fault in a


ball bearing, showing the spectra on the left and the cepstra on the right [8.39].
The initial cepstrum has a single peak at the quefrency equal to the
rotational period. The second cepstrum (after 5 months) reveals the development of
a fault by a series of new rahmonics. The quefrency of the first rahmonic is 4.1
times lower than the shaft speed quefrency (r.p.m.). This means that its
corresponding frequency is 4.1 times the shaft speed. In this case it immediately
identified the source as corresponding to the impact rate of the outer race for a
particular bearing in the gearbox (which had 10 balls and a ball-diameter-to-pitch-
diameter of 0.18).
The cepstrum can only be used for bearing fault diagnosis when the fault
generates discrete harmonics in the spectrum. This is usually the case for high-
speed machines, where resonances excited by the fault represent a relatively low
harmonic order of the ballpass frequencies involved, but is often not the case for
slow-speed machines, where this order may be in the hundreds or even thousands,
and these high harmonics are often smeared together. It should be noted that
envelope analysis (see Section 8.7.3), where the envelope obtained by amplitude
demodulation of the band-pass filtered signal is frequency analyzed, can be used in
either case.

References

8.1. * Predictive maintenance through the monitoring and diagnostics of


rolling element bearings, Bently Nevada Application Note AN044, June 1988.
8.2. Li, C. J. and McKee, K., Bearing diagnostics, Encyclopedia of Vibration,
Braun, S., Ewins, D. and Rao, S.S., eds., Academic Press, London, 2002, p.143-
152.
8.3. Changsen, W., Analysis of rolling element bearings, Mechanical Engineering
Publications, Ltd., London, 1991.
8.4. Scheithe, W., Vibration measurement a method for early detection of rolling
element bearing failures, Practice of Vibration Analysis 13, Schenck C 1213e.
8.5. Hamrock, B. J. and Anderson, W. J., Rolling-Element Bearings, NASA
Reference Publication 1105, June 1983.
8.6. Jones, A. B., The mathematical theory of rolling element bearings, Mechanical
Design and Systems Handbook, H.A.Rothbart, ed., McGraw Hill, New York,
1964, p.13-1 to 13-76.
8.7. Jones, A. B., Ball motion and sliding friction in ball bearings, Journal of Basic
Engineering, Trans. ASME, vol.81, March 1959, p.1-12.
8. ROLLING ELEMENT BEARINGS 37

8.8. Jones, A. B., A general theory for elastically constrained ball and radial roller
bearings under arbitrary load and speed conditions, Journal of Basic
Engineering, Trans. ASME, vol.82, June 1960, p.309-320.
8.9. Harris, T. A., An analytical method to predict skidding in high speed roller
bearings, Trans. ASLE, vol.9, 1966, p.229-241.
8.10. Gupta, P. K., Dynamics of rolling element bearings, Journal of Lubrication
Technology, Trans.ASME, vol.101, no.3, 1979, p.293-326.
8.11. Meyer, L. D., Ahlgren, F. F. and Weichbrodt, B., An analytic model for ball
bearing vibrations to predict vibration response to distributed defects, Journal
of Mechanical Design, Trans. ASME, vol.102, no.2, April 1980, p.205-210.
8.12. Tandon, N. and Nakra, B. C., Vibration and acoustic monitoring techniques
for the detection of defects in rolling element bearings A review, Shock and
Vibration Digest, vol.24, no.3, March 1992, p.3-11.
8.13. Sunnersj, C. S., Varying compliance vibrations of rolling bearings, Journal
of Sound and Vibration, vol.58, no.3, 1978, p.363-373.
8.14. Collacott, R. A., Mechanical Fault Diagnosis, Chapmann and Hall, London,
1977.
8.15. Su, Y.-T., Lin, M.-H. and Lee, M.-S., The effects of surface irregularities on
roller bearing vibrations, Journal of Sound and Vibration, vol.165, no.3, 1993,
p.455-466.
8.16. Babkin, A. S. and Anderson, J. J., Mechanical signature analysis of ball
bearings by real time spectrum analysis, Nicolet Instruments Application Note
3, May 1972.
8.17. * Bearing failures and their causes, SKF Repro 19208.
8.18. Roos, C. H., Vibration signature analysis of bearings and electronic
packages, Paper SI-460, 41st Shock and Vibration Symposium, Colorado
Springs, Oct 1970.
8.19. * Detecting faulty rolling-element bearings, Brel & Kjaer Application
Note, BO 0210-11.
8.20. Dyer, D. and Stewart, R. M., Detection of rolling element bearing damage by
statistical vibration analysis, Journal of Mechanical Design, Trans. ASME,
vol.100, no.2, Apr 1978, p.229-235.
8.21. Lundberg, G. and Palmgren, A., Dynamic capacity of rolling bearings, Acta
Polytechnica, Mechanical Engineering Series, vol.1, no.3, Stockholm, 1947.
8.22. Martin, H. R., Statistical moment analysis as a means of surface damage
detection, Proc. 7th International Modal Analysis Conference, Schenectady,
New York, 1989, p.1016-1021.
38 DYNAMICS OF MACHINERY

8.23. Stewart, R. M., Application of signal processing techniques to machinery


health monitoring, Stewart Hughes Ltd., Southampton, U.K., 1981.
8.24. Volker, E. and Martin, H. R., Application of Kurtosis to damage mapping,
Proc. 4th International Modal Analysis Conf., Los Angeles, 1986, p.629-633.
8.25. Daadbin, A., and Wong, J. C. H., Different vibration monitoring techniques
and their application to rolling element bearings, International Journal of
Mechanical Engineering Education, vol.19, no.4, 1991, p.295-304.
8.26. Mathew, J. and Alfredson, R. J., The condition monitoring of rolling element
bearings using vibration analysis, Journal of Vibration, Acoustics, Stress and
Reliability in Design, Trans. ASME, vol.106, July 1984, p.447-453.
8.27. Taylor, J. I., Identification of bearing defects by spectral analysis, Journal of
Mechanical Design, Trans. ASME, vol.102, no.2, April 1980, p.199-204.
8.28. Angelo, M., Vibration monitoring of machines, Brel & Kjaer Technical
Review, no.1, 1987.
8.29. Xu, M., Spike Energy and its applications, Shock and Vibration Digest,
vol.27, no.3, May-June 1995, p.11-17.
8.30. Sidahmet, M. and Dalpiaz, G., Signal generation models for diagnostics,
Encyclopedia of Vibration, Braun, S., Ewins, D. and Rao, S.S., eds., Academic
Press, London, 2002, p.1184-1193.
8.31. Shea, J. M. and Taylor, J. K., Using Spike Energy for fault analysis and
machine-condition monitoring, IRD Mechanalysis Technical Report 11, 1990.
8.32. Mignano, F., Envelope detection, Shock and Vibration Digest, vol.29, no.3,
March 1997, p.18-23.
8.33. McFadden, P. D. and Smith, J. D., Vibration monitoring of rolling element
bearings by the high frequency resonance technique. A review, Tribology
International, vol.17, 1984, p.1-18.
8.34. Courrech, J. and Gaudet, M., Envelope analysis the key to rolling-element
bearing diagnosis, Brel & Kjaer Application Note No. BO0187-11.
8.35. Butler, D. E., The shock-pulse method for the detection of damaged roller
bearings, Non-Destructive Testing, April 1973, p.92-95.
8.36. Lee, G., What is shock pulse method?, www.reliabilityweb.com.
8.37. * Shock Pulse Analyzer A2011, Instruction Manual, SPM Instrument AB,
no.71416.B, Nov.1992.
8.38. Lundy, J., Detecting lubrication problems using shock pulse, Lubrication and
Fluid Power, Jan-Feb.2006, p.57-62.
8.39. Randall, R. B., Cepstrum analysis, Encyclopedia of Vibration, Braun, S.,
Ewins, D. and Rao, S.S., eds., Academic Press, London, 2002, p.216-227.
9.
GEARS

This chapter provides an overview of current vibration methods used for


gearbox diagnostics. It presents gear defects, gear errors and condition indicators
used for gear failure detection. Differences of various metrics are in the
characteristic frequencies that are included, excluded, or used as a reference.

9.1 Gear types

Four essential types of gear are shown in Fig. 9.1. Spur gears (Fig. 9.1, a)
are used to transmit rotary motion between parallel shafts. They are usually
cylindrical in shape, and the teeth are straight and parallel to the axis of rotation.

a b c d
Fig. 9.1 (from [9.1])

Helical gears, used to transmit motion between parallel shafts, are shown
in Fig. 9.1, b. The line of contact of helical-gear teeth is diagonal across the face of
the tooth, so that there is a gradual engagement of the teeth and a smooth transfer
of load from one tooth to another. Helical gears subject the shaft bearings to both
40 DYNAMICS OF MACHINERY

radial and thrust loads. Double helical gears (herringbone) are used for the
transmission of high torques at high speeds, and to cancel out the thrust load.
For power transfer between intersecting shafts, there are straight bevel
gears (Fig. 9.1, c). Spiral bevel gears (Fig. 9.1, d) are the bevel counterpart of the
helical gears. Their teeth are curved and oblique.

a b c
Fig. 9.2 (from [9.2])
Hypoid gears (Fig. 9.2, a) are like spiral bevel gears, but their pitch
surfaces are hyperboloids rather than cones, and their pitch axes do not intersect.
They operate more smoothly and quietly and are stronger for a given ratio. Crossed
helical gears (Fig. 9.2, b), also known as spiral gears, are ordinary helical gears
used in nonparallel shaft applications.
The worm gearset (Fig. 9.2, c) consists of a worm, which resembles a
screw, and a worm wheel, which is a helical gear, with the respective shafts 900
apart. They are quiet and vibration free, with lower Hertz contact stresses than the
crossed-helical gears.

9.2 Gear tooth action

For spur gears, the terminology of gear teeth is given in Fig. 9.3. Gear
calculations are based on the theoretical pitch circle. The operating pitch circles of
a pair of gears in mesh are tangent to each other. The clearance circle is tangent to
the addendum circle of the mating gear.
Additional terminology is shown in Fig. 9.4. Here the pinion rotates
clockwise and drives a gear in a counterclockwise direction. OP is the line of
centers, connecting the rotation axes of the meshing gears. The pitch circles are
tangent at P, the pitch point.
9. GEARS 41

The resultant force vector between a pair of meshing gears acts along the
pressure line (also called line of action or generating line). The pressure line is
tangent at points c and d to the base circles.

Fig. 9.3 (from [9.3])

The angle between the pressure line and the common tangent to the pitch
circles is the pressure angle, and it usually has values of 20 or 25 deg. The
operating diameters of the pitch circles depend on the center distance used in
mounting the gears, but the base circle diameters are constant and depend only on
how the tooth forms are generated, because they form the base of the starting point
on the involute profile.

Fig. 9.4 (from [9.3])


42 DYNAMICS OF MACHINERY

Point a is the initial point of contact, where the flank of the pinion driving
tooth just touches the tip of the driven tooth. This point is located at the intersection
of the addendum circle of the gear with the pressure line. Should point a occur on
the other side of point c on the pinion base circle, the pinion flank would be
undercut during the generation of the profile.
Point b is the final point of contact, when the tip of the driving tooth just
leaves the flank of the driven tooth. This point is located at the intersection of the
addendum circle of the pinion with the pressure line. For no undercutting of the
gear teeth, point b must be located between the pitch point P and the tangent point
d on the base circle of the gear.
Line aP represents the approach phase of tooth contact, while line Pb is
the recess phase. Tooth contact throughout the line of action ab is by both sliding
and rolling, except for an instant at P when the contact is pure rolling.
Sliding gives rise to friction forces that vary in magnitude and direction as
the teeth go through the meshing cycle. During the approach action, the flank of the
pinion tooth is sliding down the face of the gear tooth, producing a frictional force
oriented upwards in Fig. 9.4. During the recess action, the face of the pinion tooth
is sliding up the flank of the gear tooth, and the resulting friction force exerted by
the pinion against the gear is oriented in opposite direction (downwards in Fig.
9.4). Friction forces produce a characteristic type of gear wear.

Fig. 9.5 (from [9.4])

The zone of action of a pair of meshing gear teeth is shown in Fig. 9.5. The
arc of action AB is the sum of the arc of approach AP and the arc of recess PB.
In the unlikely situation in which the arc of action is exactly equal to the
circular pitch, when one pair of teeth are just beginning contact at a, the preceding
pair will be leaving contact at b. Thus, for this special condition, there is never
more or less than one pair of teeth in contact.
If the arc of action is greater than the circular pitch (their ratio is called the
contact ratio) but less than twice as much, then when a pair of teeth come into
contact at a, another pair of teeth will be still in contact somewhere along the line
9. GEARS 43

of action ab. Thus, for a short period of time, there will be two pairs of teeth in
contact, one near the vicinity of A and another near B. As the meshing proceeds,
the pair near B must cease contact, leaving only a single pair of contacting teeth,
until the procedure repeats itself. Gears are not generally designed having contact
ratios less than 1.20 because inaccuracies in mounting might reduce the contact
ratio even more, increasing the possibility of impact between the teeth as well as an
increase in the noise level. A contact ratio of 1.2 means 80 percent of the time
single tooth contact, and 20 percent of the time double tooth contact.

Fig. 9.6 (from [9.5])

The contact ratio is equal to the length of the line of action ab divided by
the base pitch. The base pitch is the distance, measured on the line of action, from
one involute to the next corresponding involute.
In Fig. 9.6, a the mating teeth of the meshing spur gears are in contact at
the pitch point. The number of tooth pairs in contact is shown in Fig. 9.6, b. The
transition from single to double tooth contact produces variations in the mesh
stiffness.
44 DYNAMICS OF MACHINERY

The tooth involute profiles are designed to produce a constant angular


velocity ratio during meshing. Ideally, when two gears are in mesh, their pitch
circles roll on one another without slipping. Denoting the pitch radii by r1 and r2 ,
and the angular velocities by 1 and 2 , the pitch line velocity is

v = r1 1 = r2 2 . (9.1)
Thus, the gear ratio is
2 r1
i= = . (9.2)
1 r2

Fig. 9.7 (after [9.4])

In order to transmit uniform rotary motion during meshing, a pair of gears


must meet the following requirements (Fig. 9.7):
a) the pitch point P must remain fixed on the line of centers O1O2 ;
b) the lines of action for every instantaneous point of contact e must pass
through the same point P;
c) the generating (pressure) line must be always tangent to the base circles
and normal to the involute profiles at the point of contact e.
Deviations from the above requirements produce transmission errors
giving rise to vibrations [9.6].
Changing the center distance, the above requirements are still satisfied,
because it has no effect on the base circles used to generate the tooth profiles.
Increasing the center distance increases the pressure angle and decreases the length
of the line of action, but the teeth are still conjugate, and the angular velocity ratio
9. GEARS 45

is not changed. This increase creates two new operating pitch circles having larger
pitch diameters but remaining tangent to each other at the pitch point.
Interference might be produced by the contact of portions of tooth profiles
which are not conjugate. It is eliminated by undercutting (which weakens the
teeth), by using a larger pressure angle, or by increasing the number of teeth, hence
increasing the pitch line velocity and making the gears noisier, which is an
unacceptable solution.

9.3 Gear vibrations

Rigid and geometrically perfect gears do not produce vibrations.


Deviations from the ideal tooth profile and gear geometry generate vibrations
whose measurement and analysis can help in diagnosing gearbox faults. The main
sources of such deviations are the tooth deflection under load, the wheel distortion
during heat treatment or gearbox assembly, and the geometrical errors in the profile
itself, resulting from the gear cutting process and wear.

9.3.1 Tooth engagement

Assuming the teeth to be perfectly formed, equally spaced, perfectly


smooth, and absolutely rigid, the meshing frequency, f m , is equal to the number of
teeth in the wheel, N, multiplied by the speed of the shaft on which the wheel is
mounted, f s , in rps

fm = N fs . (9.3)

For a pair of spur gears, if f s1 and f s 2 are the rotation frequencies of the
two shafts, and N 1 and N 2 are the corresponding number of teeth, the
fundamental meshing frequency is the same for both gears in mesh

f m = N1 f s1 = N 2 f s 2 . (9.4)
An epicyclic geartrain is shown in Fig. 9.8. It consists of three revolving
planet pinions that engage the central sun gear and the coaxial ring gear with
internal teeth, and a carrier in which the planet pinions are supported. For a
planetary gear system, the following relationships can be used:

the meshing frequency

f m = N s ( f s fc ) = Nr ( fc fr ) , (9.5)
46 DYNAMICS OF MACHINERY

the carrier frequency

Ns fs + Nr fr
fc = , (9.6)
Ns + Nr

where f r and N r are the speed (rps) and the number of teeth of the ring gear, and
f s and N s are the speed (rps) and the number of teeth of the sun gear.

Fig. 9.8 (from [9.2])

In most planetary gear systems one of the elements is attached to the frame
and has a zero input motion.
Profile errors identical on each tooth, or deflection effects which are the
same for each tooth mesh, produce vibrations with components at the tooth
meshing frequency and its harmonics.

9.3.2 Effect of tooth deflection

Consider a pair of gears whose teeth are not rigid, but equally spaced,
perfectly formed and at constant speed. Since the contact stiffness varies
periodically, as shown in the lower part of Fig. 9.6, with the number of teeth in
contact and with the contacting position on the tooth surface, vibration will be
excited at the tooth engagement frequency and its harmonics. A typical gearmesh
waveform is shown in the lower part of Fig. 9.9.
In Fig. 9.6, the segment ab on the line of action denotes the interval of
engagement of a pair of gears. At the point a, when the flank of the driving tooth A
just touches the tip of the driven tooth D, there are two pairs of teeth meshing, each
taking a share of the transmitted load. Tooth B will then be relieved of some of its
load and will tend to deflect towards its unstressed position, imparting a forward
acceleration to tooth E on the driven gear. At the termination of meshing of teeth B
and E, only teeth A and D are available to transmit the load, as a result of which
9. GEARS 47

tooth A is deflected back further and tooth D will momentarily lag. The final point
of contact b is where the addendum circle of the driver gear crosses the pressure
line.

Fig. 9.9 (from [9.7])

This tooth deflection is very load dependent. For condition monitoring


purposes it is necessary to make measurements always at the same loading, and this
loading should be sufficient to ensure that the teeth are permanently in contact, and
not able to move into the backlash.

9.3.3 Effect of tooth wear

During the motion of the compliant meshing gears, the wear produced by
sliding tends to give the kind of profile deviation indicated in exaggerated form in
Fig. 9.10.

Fig. 9.10 (from [9.7])

When the point of contact of the engaging teeth reaches the pitch point,
the direction of sliding reverses, causing a shock sometimes referred to as the
pitch-circle impulse which is perpendicular to the axes of rotation of the two
48 DYNAMICS OF MACHINERY

gears. The two shafts are then subjected to bending stress reversals at the rate of the
product of shaft speed and number of teeth.
When a new pair of teeth takes part in the transmission of load, the driven
gear retrieves its retardation by a renewed forward acceleration. It is subjected to
an engagement shock, the impulse acting in a tangential direction at a rate of the
product of rotational speed and number of teeth. These impulses cause the
transmitted torque to fluctuate about a mean level, with variations of the angular
velocity, producing a frequency modulation of the tooth-meshing frequency.
The pitch circle and engagement vibrations are transmitted through the
shaft and bearing housing causing casing vibrations. This vibration can be
measured using an accelerometer mounted on the casing.

9.3.4 Ghost components

Ghost (phantom) frequency components in the gear vibration signal are


due to periodic faults introduced into the gear by the machining process. They
normally correspond to the number of teeth on the index wheel driving the table of
the gear-cutting machine, and are due to errors in these teeth. Therefore, they
appear at a harmonic of the particular gear speed. Being a result of a fixed
geometrical error, they are not very load dependent and get smaller as a result of
gear wear. Hereditary excitations can also occur at frequencies determined by the
characteristics of the machine which made the gear-cutting machine.

9.3.5 Modulation effects

Gear defects alter the magnitude and phase of the meshing stiffness and
therefore produce changes in the amplitude and phase of the vibration at meshing
frequency and its harmonics as the teeth go through the meshing. In addition, these
changes introduce amplitude and phase modulation effects which create side-bands
around the meshing frequency and its harmonics. The spacing of these side-bands
is the rotating speed of the gear (Fig. 9.11).
Faults occurring in a gear system introduce time-varying torques. These
induce a multiplicative effect and obviously modulation effects. Distributed effects,
affecting all the teeth (imperfect tooth profile, wear, etc.) generate modulation at
the meshing frequency, f m . Localized defects (like spalling, cracks, and breakage)
generate repetitive impulses at the shaft rotation frequencies f s1 and f s 2 . This
gives rise to amplitude or phase modulation effects at these frequencies. Due to
imperfect profile and teeth surface quality, the gear vibration spectrum consists of
numerous harmonics, of frequencies
f (k , p , q ) = k f m p f s1 q f s 2 , k = 1, 2, .. , p , q = 0,1, 2, .. . (9.7)
9. GEARS 49

Fig. 9.11
The existence of complex phase and amplitude modulation may also be
interpreted as a nonlinear or cyclostationary phenomenon.

Amplitude modulation
When the excitation due to the tooth engagement occurs simultaneously
with excitations having a frequency of once or twice per gearwheel rotation,
amplitude modulation (multiplicative) effects are produced (Fig. 9.12).

Fig. 9.12
Typical once per revolution excitations are produced by: a) the
accumulative effect of the pitch error; b) an isolated error of the tooth form; c)
debris trapped in the teeth; d) eccentricity of mounting the gear wheel; e) load
variation and f) unbalance. Typical twice per revolution excitations are produced
by misalignment and wheel distortion (ovality).
50 DYNAMICS OF MACHINERY

Amplitude modulation of a carrier frequency by a lower frequency gives


rise to a pair of sidebands in the frequency spectrum, spaced on either side of the
carrier frequency by an amount equal to the modulating frequency.

Fig. 9.13 (from [9.7])

Fig. 9.13 shows the simple case of modulation of a cosine signal, uc (t ) ,


by a lower-frequency cosine (plus d-c component), um (t ) . U c ( f ) and U m ( f ) are
the corresponding Fourier spectra obtained by a forward Fourier transform. The
final resulting spectrum consists of the carrier frequency plus two sidebands spaced
at an amount equal to the modulating frequency. Indeed, transformation of the
product of cosines in a sum yields
1
cos c t cos m t = [ cos (c + m )t + cos (c m )t ] , (9.8)
2
where
cos c t =
2
e(
1 i c t
)
+ e i c t ,
and
c = 2 f c , m = 2 f m .
9. GEARS 51

Considering only amplitude modulation effects, Fig. 9.14 shows the


effect of fault distribution on the sideband pattern. A very localized fault, e.g., on
one tooth, would tend to give a modulation by a short pulse of length of the order
of the tooth mesh period, but repeated once every revolution. Figure 9.14, a shows
how this in the spectrum would result in the generation of a large number of
sidebands of almost uniform level.

Fig. 9.14 (from [9.7])

The effect of a more distributed fault is shown in Fig. 9.14, b. It is seen


that as the envelope of the fault in the time signal becomes wider, it makes the
corresponding envelope in the frequency domain narrower and higher. The
resulting modulation products become more obviously sidebands grouped around
the tooth meshing harmonics.

Frequency modulation
When the rotational speed of the gears is not constant, and the tooth
spacing is not perfectly uniform, a frequency modulation of the tooth meshing
frequency occurs. In fact, the same fluctuations in the tooth contact pressure which
give rise to amplitude modulation apply a fluctuating torque to the gears, producing
angular velocity fluctuations at the same frequency.
Frequency modulation, even by a single frequency f1 , gives rise to a
whole family of sidebands with a spacing equal to the modulating frequency, i.e.,
the same frequencies as produced by amplitude modulation by a distorted periodic
signal (Fig. 9.15). Since in gears the two effects are virtually inseparable, the
52 DYNAMICS OF MACHINERY

resulting spectrum is a combination of the sidebands produced by both amplitude


and frequency modulation.

Fig. 9.15

The instantaneous amplitude of a frequency modulated signal can be


represented by [9.7]
(
a = A cos 0 t + sin 1 t , ) (9.9)

where = 1 is the modulation index. It represents the maximum deviation


of phase from that of the unmodulated carrier.
The decomposition into cosine components, and further decomposition
into positive and negative frequency components, yields (giving details of the
positive frequency components alone)

{ C0 ( ) e 0 + C1 ( ) [ e 0 1 e
A i t i ( + ) t i ( 0 1 ) t
a= ]+
2 (9.10)
+ C2 ( ) [ e
i ( 0 + 2 1) t i ( 0 2 1) t
e ] + ...} + negative frequency terms...

The relative amplitude of the carrier frequency component is given by


C0 ( ) and that of the nth order sidebands by Cn ( ) .

Assuming in a gearbox that the carrier frequency f 0 is the tooth meshing


frequency and the modulating frequency f1 is the rotating speed of the gear, then

f f f0
= = =N , (9.11)
f1 f 0 f1

where is the relative speed fluctuation of the gear, and N is the number of teeth
on the gear.
9. GEARS 53

Fig. 9.16 (from [9.7])

It can be shown that, for << 1 only one pair of sidebands is required,
while for < 1 most information is contained in the first two pairs of sidebands
(Fig. 9.16).
Frequency modulation tends to modify the relative amplitude of the
sidebands produced by the amplitude modulation. Its additional effect is to increase
the number of sidebands somewhat and to make the sideband patterns
unsymmetrical by reinforcement/cancellation because of the different phase
relationships of the sidebands.

9.3.6 Resonance effects

The excitation frequency due to tooth engagement is a potential source of


vibration trouble through resonance if it coincides with a natural frequency of some
part of the structure such as the webs and walls of the gear cases, the discs of large
gear wheels and the blades of a turbine [9.8].
Gear pairs are part of torsional (and axial) vibrating systems so that their
location and characteristics have an important influence on the dynamic response
especially near resonances.
If the vibratory torque at the gears is greater than the mean transmitted
torque, tooth separation occurs due to the backlash and this can result in severe
impact loading (hammering) at the gear teeth upon re-engagement.
54 DYNAMICS OF MACHINERY

For main transmission gears, if the gear assembly is located at a node of


vibration and there are gear errors attempting to force a vibratory movement at that
point, very large dynamic loads will be produced. By placing the gears away from
a nodal point, the vibratory torque at the gears may be appreciably reduced.
In the case of auxiliary drives, it is usually preferable to take off power at a
node of free torsional vibration, since the vibration amplitude at a node is small and
is therefore less likely to cause noise and wear or interfere with the functioning of a
comparatively lightly loaded auxiliary system.
In heavily loaded gears, the use of tip and/or root relief is usually desirable
to accommodate tooth deflection under load.
With helical gearing, tooth errors can excite torsional and axial vibrations
or a combination of the two. In double-helical gearing, the type of vibration excited
depends upon the phasing of errors in the right- and left-hand helices. If errors are
in phase the tendency is to excite torsional vibrations. If the errors are 1800 out of
phase, the tendency is to excite axial vibrations.
In difficult cases of resonance, it might be desirable to change the number
of teeth in all gears whose meshing is responsible for the disturbance or,
alternatively, modify the structural elements which are responding to the
disturbance.

9.4 Gear errors

The motion between engaging gear teeth is likely to be subject to


variations from the ideal due to practical variations which are accentuated by the
teeth profile design. The main gear errors are listed below.
Pitch errors. Variations in spacing occur between corresponding sides of
adjacent teeth. Pitch errors create angular accelerations in the transmission motion,
with resultant torques and forces, which make a considerable contribution to gear
noise.
Radial run-out errors. These are eccentricities of the pitch cylinder
measured in the transverse plane, due to incorrect mounting of the gear on the shaft
or produced during gear manufacturing (ghost and hereditary errors). They
generate vibration and noise once per revolution. Amplitude modulation of the
tooth contact and its harmonics leads to side-band effects which extend over a
considerable frequency range.
Profile errors. Such errors are the deviations of an actual profile from the
theoretically correct involute profile in a transverse plane. They are primary factors
contributing to noise and vibration in gears.
9. GEARS 55

Similar effects are produced by variations in mesh stiffness, further


influenced by the bearing clearance, shaft and housing deflections and
manufacturing errors.
Teeth which have pitch or profile errors are prone to tooth separation,
high impact loads and failure. This occurs especially for lightly loaded gears where
the stress at the root of a tooth and the vibration may be excessive at critical
speeds.

9.5 Gear faults

Gear faults may be classified in terms of the effects of wear, fatigue and
breakage. The terminology of gear faults is given in the following [9.5].

9.5.1 Wear effects

Scoring, often referred to as scuffing, is evidenced by radial wear lines


superimposed on a roughened thin layer of melted material. The terms scuffing and
scoring are frequently interchanged.
Gear scuffing is characterized by material transfer between sliding tooth
surfaces. It occurs when inadequate lubrication film thickness permits metal-to-
metal contact between gear teeth. Without lubrication, direct metal contact removes
the protective oxide layer on the gear metal, and the excessive heat generated by
friction welds the surfaces at the contact points. As the gears separate, metal is torn
and transferred between the teeth.
The ASM Handbook Vol.18 defines scoring as the formation of severe
scratches in the direction of sliding. Scoring tends to progress steadily from the tip
of the driving pinion inwards, and correspondingly over the driven profiles.
Scoring can be inhibited by phosphate treatment or copper plating of the
tooth surfaces. Extreme-pressure additives inhibit scoring, without marked change
in the viscosity grade of the oil. Minor scoring may be considered as scratching.
Scuffing is most likely to occur in new gear sets during the running-in
period, because the gear teeth have not sufficient operating time to develop smooth
surfaces.
Galling is a form of contact welding that results in the transfer of material
from one gear number to another. It is quite infrequent in moderate to high-speed
gearing, but is often seen in low-speed and stop/start type operations.
56 DYNAMICS OF MACHINERY

Frosting applies to fully hardened gear teeth profiles defined by the


existence of small round or elliptical patches which (under high magnification)
exhibit the general appearance of minute scorings.
Destructive wear results in a corrosive change in the involute shape of the
gear tooth. It would be accompanied by extremely rough operation, non-uniform
motion and shock overloads, which would probably result in tooth breakage.
Corrosive wear generally occurs in long lasting operations in the presence
of chlorine or sodium based lubricants. This should not be confused with ordinary
oxidation corrosion which may occur during no operation intervals due to
improper preservation. It may be difficult to detect if the resultant pitting is fine.
Interference wear is the effect of the tip of one gear tooth tip in contacting
the fillet or root area of its mating tooth. When the initial contact occurs, the root of
the pinion tooth engages with the tip of the wheel tooth. Deflection occurs at the
point of maximum relative sliding speed, so that the tip may dig in and cause wear
of the pinion root. Conversely, during recession, the tip of the pinion tooth deflects
as it disengages from the wheel tooth causing wear of the wheel root. To overcome
this, the tips of both pinion and wheel teeth should be carefully and accurately
relieved.
Burning indicates surface tempering or softening of the tooth, most
probably accompanied by a total loss of lubricant. It is an advanced condition of
discoloring.
Discoloration is a term used to locate the existence of surface temper
colouring of the active profile (the top band) of a gear tooth. The condition
indicates a marginal lack of lubrication or an excessive power operation.
Misalignment wear arises from operation at skewed axes.

9.5.2 Effects of fatigue (surface contact)

Pitting is a fatigue failure due to the high contact stresses produced in


gears. Pitting occurs when small pieces of material break off from the gear surface,
producing pits on the contacting surfaces. The fatigue cracks are initiated on the
tooth surface or just below the surface, caused by the Hertzian contact due to low
lubricant film thickness. Cyclic action of applied and released excessive local
pressure combined with the sliding and rolling action between mating teeth is
believed to cause local fatigue failure which produces the pits (Fig. 9.17).
Pitting may also be produced by hydrogen embrittlement of metal due to
water contamination of the lubricant. Pitting is confined almost entirely to the
dedendum of both driving and driven gears, being most severe about the pitch line.
The occurrence of pitting is proportional to the tensile strength of the steel, it
increases with oil viscosity and is adversely affected by surface roughness.
9. GEARS 57

Micropitting occurs on surface-hardened gears and is characterized by


extremely tiny pits, approximately 10 m deep.
Spalling is a pitting condition whose origin can be physically detected at
the apex of the fan-shaped portion of the damaged area. This is a surface-initiated
type of fatigue, which has its origin in the surface tensile cracking which leads to
the gradual erosion and exfoliation of increasingly larger pieces of gear material as
the fan widens out in the direction of sliding action. The cracks will ultimately
undermine the entire case of case-hardened gear teeth as the spalling approaches
the extremities of the addendum.

Fig. 9.17 (from [9.9])

Arrested pitting describes very small shallow pits that are not propagating
into larger failure areas. It has been observed on spiral bevel gears and is frequently
associated with the waviness condition referred to as barber pole. This pitting is
often considered corrective in that it progresses immediately to the point of
relieving local compressive stress of overload.
Pitch line pitting belongs to the family of rolling contact fatigue and is
truly subsurface in origin. It is not generally associated with a condition of
lubrication distress but generally occurs at relatively high cycles of loading. In
fully hardened, properly designed gears, it is seldom seen in less than 100,000
cycles of operation.
Addendum pitting and dedendum pitting are terms which merely signify the
site of origin of one of the foregoing types of pitting or spalling.
Case crushing means shear failure of the core-case interface in case-
hardened gear teeth. It indicates insufficient case depth for the load magnitude.
58 DYNAMICS OF MACHINERY

Multiple cracking, often both transverse and longitudinal, is generally observed in


the tooth flank.
The Hertzian shear stress increases from zero at the surface to a maximum
at a depth depending upon: a) surface load concentration; b) contact length at the
surface, and c) relative curvature of the surfaces in contact. Case depth is the
distance from the surface to the position where the hardness reaches that of the core
material. For gear teeth, the maximum shear stress should be within the depth of
hard casing.

9.5.3 Tooth fracture

Bending stresses under heavy cyclic loads produce a fatigue crack at the
fillet of the root of the tooth and results in failure (Fig. 9.18). The crack progresses
inwards and slightly downwards, then rises again until the fracture is completed at
the opposite fillet. Other causes of high stress concentration and fatigue are the
incorrect fillet radii.

Fig. 9.18 (from [9.10])

High local temperatures arising from the grinding of hardened teeth


without adequate coolant leaves the surface layers in a state of tension. In severe
cases, grinding cracks may form at the root of tooth fillets. Such cracks act as
nuclei for tooth fracture.

9.6 Gear condition monitoring

Gear damage produces changes in the vibration signatures measured by


accelerometers installed on gearboxes. In practice, direct comparisons of current
vibration signatures with previous signatures are not effective, due to large
variations. Instead, more useful techniques that involve the extraction of features
from the vibration signature data are being used, based on some statistical
measurement of vibration energy.
9. GEARS 59

Feature extraction is the process of extracting condition indicators


(metrics) about the system input which are more informative than evaluating the
raw input itself. It is a parametrization process, which often reduces the data
volume. Feature extractors output only the information relevant for detecting the
failure modes to which the associated components are susceptible.
In vibration monitoring, damage detection techniques based on the
Discrete Fourier Transform (DFT) have been traditionally developed. The absolute
value of DFT contains an estimate of the signal power spectrum, which displays
different behavior between health state and damaged state signals. However, the
absolute DFT is insensitive to the shaft phase offset, which is random and thus a
source of unwanted variation in signal characteristics.
Given the geometry of rolling element bearings, it was possible to predict
which frequencies, i.e. DFT coefficients, are affected by different failure modes.
For gear systems, some fault-indicating signal characteristics are not well captured
by the DFT, but are better enhanced by other transforms. Several condition
indicators (figures of merit) have been introduced to detect localized damage in
gears [9.11-9.13]. Ideally, these features are more stable and well behaved than the
raw signature data itself.

9.6.1 Vibration signal processing

Before any feature can be calculated on the raw vibration data, the data
must be conditioned or preprocessed. Conditioning may range from signal
correction, based on the data acquisition unit and amplifiers used, and mean value
removal, to time-synchronous averaging and filtering. Different signal processing
techniques are used based on the condition indicator being implemented (Fig.
9.19).
Basic raw signal conditioning is used to calculate the root mean square
(r.m.s.), Kurtosis, Delta r.m.s., Crest factor, Enveloping and Demodulation, as for
rolling element bearings (see Chapter 8). The only preprocessing is removing the
mean of the signal. Conditioning is simply multiplying all of the data points by
some calibration constant that is based on the accelerometer and amplifier used.
Time synchronous averaging (TSA) is used to extract repetitive signals
from additive noise. This process requires an accurate knowledge of the repetitive
frequency of the desired signal or a tachometer signal that is synchronous with the
desired signal. The raw data is then divided up into segments of equal length
blocks related to the synchronous signal and averaged together. When sufficient
averages are taken, the random noise is canceled, leaving an improved estimate of
the desired signal. The TSA signal is used for calculation of the FM0 indicator
(Stewart, 1977).
60 DYNAMICS OF MACHINERY

Fig. 9.19 (from [9.11])

Further preprocessing calculates the residual signal, which consists of the


TSA signal with the primary meshing and driveshaft components along with their
harmonics removed. Good results are obtained by high passing the data about some
frequency and remove the meshing frequency and all harmonics. The cut-off
frequency of the high pass filter is system dependent, and is selected between d.c.
and the fundamental meshing frequency. Indicators NA4 (Zakrajsek, 1993) and
NA4* (Decker, 1994) are determined based on the residual signal.
The condition indicators FM4 (Stewart, 1977), M6A and M8A (Martin,
1989) are based on the difference signal, calculated by removing the regular
meshing (RM) components from the TSA signal. The RM components consist of
the driveshaft frequency and its harmonics, the primary meshing frequency and
harmonics, along with the first order sidebands. It turns out that the difference
signal can be obtained by removing the sidebands of the primary meshing
frequencies from the residual signal.
The NB4 condition indicator (Zakrajsek, 1994) is obtained from the band-
pass mesh (BPM) signal. The TSA signal is band-pass filtered around the primary
gear mesh frequency, including as many sidebands as possible. The Hilbert
transform is then applied to the filtered signal to produce a complex time series.
9. GEARS 61

The real part is the band-passed signal and the imaginary part is the Hilbert
transform of the signal. The envelope is the magnitude of this complex time signal
and represents an estimate of the amplitude modulation present in the signal due to
the sidebands.
Kurtosis and r.m.s. may be performed at different preprocessing levels,
while Demodulation and Enveloping may return multiple parameters.

9.6.2 Condition indicators

Some of the commonly used condition indicators are presented in the


following. The definitions assume that the input signal is finite. The primary
differences between the various condition indicators are in the signal based on
which the computations are made: the raw signal, the residual signal or the
difference signal, and in the signal used as a reference.

Root Mean Square


The root mean square (r.m.s.) is a measure of the power content of the
vibration signal. It is a general fault indicator, which provides no information on
which component is failing, and shows no appreciable changes in the early stages
of gear damage. Alone, it can be effective only in detecting major out-of-balance.
The r.m.s. of a digital signal defined by a data series xn over length N is defined as

N
1
r .m.s . =
N x
n =1
2
n . (9.12)

Delta r.m.s. is simply the difference between the current r.m.s. value and
the previous one. This parameter focuses on the trend of vibration signal and is
sensitive to changes in the vibration signal. Theoretically it allows selection of an
alarm level which is not sensitive to load, however in practice it came out that it is
sensitive to load change.

Kurtosis
Kurtosis is defined as the fourth moment of the distribution (about the
mean), normalized by the square of the variance. It measures the relative
peakedness or flatness of a distribution as compared to a normal distribution.
Kurtosis provides a measure of the size of the tails of distribution and is used as an
indicator of major peaks in a set of data. As a gear wears and eventually a tooth
breaks, this feature should signal a defect due to the increased level of vibration.
The equation for Kurtosis is given by
62 DYNAMICS OF MACHINERY

( x n x )4
N
1
N n =1
kurt = , (9.13)
4
where x is the mean of the data and 2 is the variance.
A more detailed presentation of the Kurtosis is given in Chapter 8 in
connection to the condition monitoring of rolling element bearings.

Crest Factor
The Crest Factor is defined as the ratio of the peak level to the r.m.s. level
of the signal [9.14]. In early stages of damage, there is no change in the r.m.s.
value, while the peak value increases, therefore the Crest Factor increases. As the
damage progresses, the r.m.s. value increases and the Crest Factor decreases. It is
used to detect changes in the signal pattern due to tooth breakage, but is not
considered a very sensitive indicator.
A presentation of the Crest Factor is given in Chapter 8.

Energy Operator
The Energy Operator is defined as the normalized Kurtosis of a signal in
which each point is computed as the difference of two squared neighborhood points
of the original signal

( s n s )4
N
1
N n =1
EO = 2
, (9.14)
1
( )
N
s n s 2
N n =1

where s is the mean value of signal s , sn = xn2+1 xn2 , and N is the number
of points in the dataset x. In the case of endpoints, the data is looped around [9.15].
Specifically, when calculating the first point, the last point is used and vice versa.

Enveloping
Enveloping is used to monitor the high-frequency response of a gearbox
to periodic impacts produced when a faulty tooth makes contact with the mating
tooth. These impacts usually excite a resonance in the system at a much higher
frequency than the vibrations generated by the other components. The
corresponding high frequency energy is usually concentrated in a narrow frequency
band. Tooth wear and breakage increase the amplitude of side bands near critical
frequencies such as the output shaft frequency.
9. GEARS 63

The enveloping technique, already presented in Chapter 8 for rolling


element bearings [9.11], consists of processing the structure resonance energy with
an envelope detector (Fig. 8.17).

Demodulation
When the teeth wear, the relative sliding results in a change of amplitude
or amplitude modulation of the vibrations at the gear meshing frequency f m and
its harmonics. Demodulation identifies the periodicity in the modulation of the
carrier.
The carriers are basically f m and 2 f m . Demodulation techniques detect
the amplitude modulation components induced by the gear wear at these
frequencies. This differs from enveloping which detects the combined effects over
a range of frequencies. The raw data is high-passed filtered at 0.85 f m and low-
passed filtered at 1.15 f m . The power spectral density of the filtered signal is
searched to obtain the actual carrier frequency f m . The actual carrier is used to
amplitude demodulate the filtered carrier signal. The power spectral density of the
resulting signal is searched within 5% of the output shaft frequency. The
condition indicators extracted from this technique are the frequency of the peak and
the magnitude squared amplitude.

FM 0
Major tooth faults typically result in an increase of the peak-to-peak
signal levels, but do not change the meshing frequency. The zero-order figure of
merit FM 0 is defined as the peak-to-peak level of the TSA signal divided by the
sum of the amplitudes at the gear-mesh frequency and its harmonics [9.16].
While the Crest Factor compares the peak value of the TSA signal to the
energy of the TSA signal, the FM 0 compares the peak value of the TSA signal to
the energy of the row signal.
The equation for FM 0 is

PPA
FM 0 = n
, (9.15)
ak
k =1

where PPA is the peak-to-peak amplitude of the TSA waveform and a k is the
amplitude of the kth mesh frequency harmonics.
64 DYNAMICS OF MACHINERY

FM 4
The indicator FM 4 was developed to detect changes in the vibration
pattern resulting from damage on a limited number of gear teeth [9.16]. FM 4 is
calculated as the Kurtosis of the difference signal divided by the square of the
variance of the difference signal

(d n d )
N
1 4
N n =1
FM 4 = 2
, (9.16)
1 2
(d n d )
N

N n =1
where d is the difference signal, d is the mean value of the difference signal, and
N is the total number of data points in the time record.

Fig. 9.20 (from [9.13])


9. GEARS 65

The difference signal is obtained by removing from the original signal the
gear meshing frequency, harmonics and first order sidebands. A flowchart for
calculating FM 4 is shown in Fig. 9.20.
It is assumed that a difference signal from a gear in good condition has a
Gaussian amplitude distribution, therefore resulting in a normalized Kurtosis value
of 3.0. As a defect develops in a tooth, such as a crack or pitting, peaks will grow
in the difference signal resulting in a less peaked amplitude distribution with a
Kurtosis value increasing beyond 3.0, typically larger than 7.0. If more than one
tooth is defective, the data distribution becomes flat and the Kurtosis value
decreases.

NA4
The NA4 indicator was developed to improve the behavior of the FM 4
indicator when more than one tooth is damaged [9.17]. It is determined by dividing
the fourth statistical moment of the residual signal by the current run time averaged
variance of the residual signal, raised to the second power.
The equation for NA4 is

( rn r )4
N
1
N n =1
NA4 = 2
, (9.17)
1 2
( r n ,m rm )
M N
1

M m =1 N n =1

where r is the residual signal, r is the mean value of the residual signal, N is the
total number of data points in the time record, and m is the current time record
number in the run ensemble.
NA4 was developed to detect the onset of damage and to continue to
react to this damage as it spreads and increases in magnitude [9.18]. If the gear
damage spreads from one tooth to another tooth, NA4 grows a) because the first
order sidebands increase, and b) the value of the average variance at the
denominator increases slower than the numerator.

NA4 *
NA4 * (or ENA4 ) was developed as an enhanced version of NA4 , and
was expected to be more robust when progressive damage occurs [9.19]. This
added robustness is obtained by normalizing the fourth statistical moment with the
residual signal variance for a gearbox in good condition, instead of the running
variance, which is used for NA4 . This overcomes the rapid increase of the
averaged variance at the denominator of equation (9.17) when the gear damage
progresses.
66 DYNAMICS OF MACHINERY

The equation for NA4 * is

( rn r )4
N
1
N n =1
NA4* = , (9.18)
( M~ 2 ) 2
~
where M 2 is the variance of the residual signal for a gearbox in good condition.

Energy Ratio
Heavy uniform wear can be detected by the Energy Ratio [9.14]. It
compares the energy contained in the difference signal, d, to the energy contained
in the regular meshing (RM) signal

d
ER = , (9.19)
RM

where denotes the standard deviation.


The basic idea is that the energy is transferred from the regular meshing
component to the rest of the signal as wear progresses.

M 6 A and M 8 A
The theory behind M 6 A and M 8 A is the same as that for FM 4 , except
that M 6 A and M 8 A are expected to be more sensitive to peaks in the difference
signal. The M 6 A indicator is determined by dividing the sixth statistical moment
about mean of the difference signal by the cube of variance of the difference signal.
The M 8 A indicator is obtained by dividing the eighth statistical moment about
mean of the difference signal by the fourth power of variance of the difference
signal [9.20].
The equations for M 6 A and M 8 A are as follows:

(d n d ) (d n d )
N N
1 6 1 8
N n =1 N n =1
M 6A = 3
, M 8A = 4
. (9.20)
1 2 1 2
( ) ( )
N N
dn d dn d
N n =1 N n =1

For a gear in good condition M 6 A = 15 and M 8 A = 105 . As a defect


develops in a tooth, M 6 A increases beyond 45 and M 8 A increases beyond 300.
9. GEARS 67

NB4
The NB4 indicator is similar to NA4 except that, instead of using the
residual signal, NB4 uses the envelope of a band-passed segment of the TSA
signal [9.21].
The idea behind this method is that a few damaged gear teeth will cause
transient load fluctuations that are different from the normal tooth load
fluctuations. The theory suggests that these fluctuations will be manifested in the
envelope of a signal which is band-pass filtered about the dominant meshing
frequency. The latter is either the primary meshing frequency or one of its
harmonics, whichever appears to give the most robust group of sidebands.
The envelope of the band-passed signal, s (t ) , is the magnitude of the
complex (i.e., analytic) signal, a (t ) + i H [ a (t )] , obtained by applying the Hilbert
transform

1 1
H [ a ( t )] = a ( ) t d (9.21)

to the band-passed signal a (t ) ,

s (t ) = (a (t ) ) 2 + H [a (t ) ] 2 . (9.22)

NB4 is then determined by dividing the fourth statistical moment about


mean of this envelope signal by the square of average variance of the envelope of
band-passed signals up to current time, with the following equation

( s n s )4
N
1
N n =1
NB 4 = 2
, (9.23)
1 2
( s n ,m sm )
M N
1
N
M m =1 n =1

where s is the envelope of the band-passed signal and s is its mean value.

9.6.3 Oil debris analysis

Oil debris analysis is a very reliable method for detecting gearing damage
in the early stages and allows estimation of the wear level [9.9]. During gearbox
operations, the mating surfaces of gearwheels are gradually abraded. Small pieces
of material break down from the contact surfaces and are carried away by the
68 DYNAMICS OF MACHINERY

lubricating oil. By detecting the number and size of particles in the oil one can
identify gear pitting damage in an early stage, which is unidentifiable by vibration
methods.
Oil debris sensors are usually based on a magnetic or an optical principle.
Magnetic sensors measure the change in magnetic field caused by metal particles in
a monitored sample of oil. A disadvantage of oil debris analysis is that it does not
localize the failure in complex gearboxes.
The oil debris sensor records counts of particles in bins set at different
particle size ranges. For each bin size range, an average particle size is first
determined. Then statistical distribution methods are applied to particles collected
from the lubrication system.
The mean particle size is calculated as

N
E ( ) =
i =1
i P [ i ] , (9.24)

where i is the average bin size, i the number of bins, and P [ i ] is the number
of particles per average bin size per reading divided by the total number of particles
per reading.

The Variance is

N
Variance = [
i =1
i E ( ) ] 2 P [ i ] . (9.25)

The Kurtosis is

N
Kurtosis = [
i =1
i E ( ) ] 4 P [ i ] . (9.26)

The relative Kurtosis is

Kurtosis
Re lative Kurtosis = . (9.27)
(Variance) 2
Laboratory experiments have shown that oil debris analysis is more
reliable than vibration analysis for detecting pitting fatigue failure of spur gears.
The increase in oil debris mass is related to damage progression, which is not
detected by some vibration based condition indicators.
9. GEARS 69

In order to extract an intelligent feature from the accumulated mass


measured by the oil debris sensor, a fuzzy logic analysis can be used. Integrating
oil debris analysis and vibration measurement results in a monitoring system with
improved damage detection and decision-making capabilities.

9.7 Cepstrum analysis

The frequency spectrum of the time signal measured on a gearbox is too


complex to be interpreted directly. Cepstrum analysis is used as a post-processing
technique to detect spectrum periodicity, i.e. the existence of sideband families.
Cepstrum is the spectrum of a logarithmic spectrum, hence a backward
transformation to the time domain. It is also a data reduction technique, effectively
reducing a whole family of sidebands into a single line and easing the monitoring
of changes in gearbox condition. Figure 9.21, b shows a typical cepstrum for a
gearbox, determined based on the frequency spectrum from Fig. 9.21, a.
Cepstrum is the inverse Fourier transform of a logarithmic spectrum

C ( ) = F 1 { log [ G ( f ) ] } , (9.28)

where G ( f ) is a frequency spectrum.


Thus, the cepstrum is a spectrum of a spectrum and for this reason, the
name cepstrum was coined from spectrum by reversing the first syllabe. Other
terms are coined in similar way, quefrency from frequency, rahmonic from
harmonic, gamnitude from magnitude, saphe from phase, quefrency alanysis
from frequency analysis, etc. [9.23].

a b
Fig. 9.21 (from [9.22])
70 DYNAMICS OF MACHINERY

When G ( f ) in equation (9.28) is a power spectrum Gxx ( f ) of the time


signal g x (t ) , i.e. G xx ( f ) = F { g x (t ) } and F { } represents the forward Fourier
2

Transform of the bracketed quantity, the resulting cepstrum is termed a power


cepstrum [9.24], defined by

C ( ) = F 1 { log [ G xx ( f )] } . (9.29)

When G ( f ) in equation (9.28) is a complex spectrum, i.e. the forward


Fourier Transform of a time signal g (t ) , the resulting cepstrum is termed a
complex cepstrum, defined by equation (9.28) but where

G ( f ) = A ( f ) e i ( f ) , (9.30)
and
ln [ G ( f ) ] = ln [ A ( f ) ] + i ( f ) . (9.31)
The independent variable, , of the cepstrum has the dimensions of time,
but is known as quefrency. A high quefrency represents rapid fluctuations in
the spectrum (small frequency spacings) and a low quefrency represents slow
changes with frequency (large frequency spacings).
When peaks in the cepstrum result from families of sidebands, the
quefrency of the peak represents the time period of the modulation. Its reciprocal is
the modulation frequency. Note that the quefrency says nothing about the absolute
frequency, only about the frequency spacings.

a b
Fig. 9.22 (from [9.22])

Figure 9.22 shows the results of this type of analysis for a gearbox.
The spectrum (Fig. 9.22, a) contains a large number of sidebands, but their
spacing is difficult to determine. Within the display range of the cepstrum (0 30
ms) the first three rahmonics of the 8.28 ms (120.75 Hz) component and only the
first rahmonic of the 20.1 ms (49.75 Hz) component are present (Fig. 9.22, b). The
periodicity is not apparent in the frequency spectrum since the mixture of the two
periodicities gives a quasi-periodic structure.
9. GEARS 71

Figure 9.23, a is a 400-line baseband spectrum from 0-20kHz of a gearbox


vibration signal containing at least the first three harmonics of the toothmeshing
frequency (4.3 kHz).
Figure 9.23, b is a 2000-line composite spectrum extending in frequency
from below the toothmesh frequency to above its third harmonic (3.5-13.5 kHz). It
excludes the low harmonics of both shaft speeds. This degree of resolution is
required to separate the individual sidebands with spacings equal to the shaft
speeds, but the eye cannot see the spectrum details.

a b

c d
Fig. 9.23 (from [9.22])

Figure 9.23, c shows the expanded 400-line section from 7500 to 9500 Hz.
The eye still cannot readily see the sideband families because of the mixture of
different spacings. The amplitude cepstrum (Fig. 9.23, d ) of the whole one-sided
spectrum (Fig. 9.23, a), reveals that all rahmonics come from one of two families,
corresponding to the speeds of the two gears in this particular gearbox (50 Hz and
85 Hz).
72 DYNAMICS OF MACHINERY

The cepstrum is considered to be an extremely useful tool for two tasks in


vibration monitoring and analysis [9.25]:
For fault detection: a) it is a sensitive measure of the growth of
harmonic/sideband families; b) the data is reduced to a single line per family; c) it
is insensitive to the location of the measurement point, to the phase combination of
amplitude and frequency, and to gearbox loading.
For fault diagnosis: a) it is an accurate measure of the spacing of frequency
components; b) it can be calculated from any section of a spectrum; c) it can be
used for the separation of different families of sidebands; and d) it is sensitive to
tooth and blade differences but not to uniform wear.

9.8 Time-frequency analysis

Local faults in gears produce impacts, hence transient modifications in


vibration signals. Therefore, vibration signals from gears are non-stationary.
However, most of the widely used signal processing techniques are based on the
assumption of stationarity. Thus they are not fully suitable for the detection of
short-duration dynamic phenomena and for the time localization of transient
events.
Application of time-frequency distribution techniques is suitable for the
detection and localization of cracks in gears. They show how the energy
distribution over frequencies changes from one instant to the next. Examples of
such distributions include wavelet transformations (J. Morlet, 1982), the short-time
Fourier transform (S. Gade and H. Herlufsen, 1987), the Wigner-Ville distributions
(E. Wigner, 1932, and J. Ville, 1948) and the exponential distribution (H. I. Choi
and W. J. Williams, 1989) [9.26-9.30]. Their study exceeds the frame of this book.

References

9.1. Sidahmed, M. and Dalpiaz, G., Signal generation models for diagnostics,
Encyclopedia of Vibration, Braun, S., Ewins, D. and Rao, S.S., eds., Academic
Press, London, 2002, p.1184-1193.
9.2. Coy, J. J., Townsend, D. P. and Zaretsky, E. V., Gearing, NASA/RP-1152,
1985.
9.3. Shigley, J. E. and Mischke, C. R., Gearing. A Mechanical Designers
Handbook, McGraw-Hill, New York, 1990.
9.4. Shigley, J. E., Mechanical Engineering Design, 2nd ed., McGraw-Hill
Kogakusha Ltd., Tokyo, 1972.
9. GEARS 73

9.5. Collacott, R. A., Gear faults diagnostics, U.K. Mechanical Health Monitoring
Group, Leicester Polytechnic, Nov. 1975.
9.6. Mark, W., Analysis of the vibratory excitation of gear systems: basic theory,
Journal of the Acoustical Society of America, vol.65, 1978, p.1409-1430.
9.7. Randall, R. B., A new method of modeling gear faults, Journal of Mechanical
Design, Trans. ASME, vol.104, April 1982, p.259-267.
9.8. Wilson, W. Ker, Practical Solution of Torsional Vibration Problems,
Chapman & Hall, London, 1956.
9.9. Dempsey, P. J., Integrating oil debris and vibration measurements for
intelligent machine health monitoring, NASA/TM-2003-211307.
9.10. Choi, S. and Li, C. J., Estimate gear tooth transverse crack size from
vibration by fusing selected gear condition indices, Measurement Science and
Technology, vol.17, 2006, p.1-6.
9.11. Lebold, M., McClintic, K., Campbell, R., Byington, C. and Maynard, K.,
Review of vibration analysis methods for gearbox diagnostics and prognostics,
Proc. 54th Meeting of the Society for Machinery Failure Prevention
Technology, Virginia Beach, VA, May 1-4, 2000, p.623-634.
9.12. Mosher, M. Pryor, A.H. and Huff, E.M., Evaluation of standard gear metrics
in helicopter flight operation, 56th Mechanical Failure Prevention Technology
Conference, Virginia Beach, VA, April 15-19, 2002.
9.13. Dempsey, P., Lewicki, D. G. and Le, Dy D., Investigation of current methods
to identify helicopter gear health, NASA/TM-2007-214664.
9.14. Swansson, N. S., Applications of vibration signal analysis techniques to
signal monitoring, Conf. on Friction and Wear in Engineering, Barton,
Australia, 1980.
9.15. Ma, J., Energy operator and other demodulation approaches to gear defect
detection, Proc. 49th Meeting of Soc. for Mechanical Failure Prevention
Technology, Virginia Beach, VA, April 1995.
9.16. Stewart, R. M., Some useful data analysis techniques for gearbox diagnostics,
Report MHM/R/10/77, Machine Health Monitoring Group, I.S.V.R., Univ. of
Southampton, July 1977.
9.17. Zakrajsek, J. J., An investigation of gear mesh failure prediction techniques,
NASA TM-102340, Nov.1989.
9.18. Zakrajsek, J. J., Townsend, D. P. and Decker, H. J., An analysis of gear fault
detection methods as applied to pitting fatigue failure data, NASA TM-105950,
April 1993.
9.19. Decker, H. J., Handschuh, R. F. and Zakrajsek, J. J., An enhancement to the
NA4 gear vibration diagnostic parameter, NASA TM-106553, June 1994.
74 DYNAMICS OF MACHINERY

9.20. Martin, H. R., Statistical moment analysis as a means of surface damage


detection, Proc. 7th International Modal Analysis Conference, Schenectady,
New York, Jan 1989, p.1016-1021.
9.21. Zakrajsek, J. J., Handschuh, R. F. and Decker, H. J., Application of fault
detection techniques to spiral bevel gear fatigue data, Proc. 48th Meeting of the
Society for Machinery Failure Prevention Technology, Wakefield, MA, April
1994.
9.22. Randall, R. B., Cepstrum analysis and gearbox fault diagnosis, Brel&Kjaer
Application Note No. 233-80.
9.23. Bogert, B. P., Healy, M. J. R. and Tukey, J. W., The quefrency alanysis of
time series for echoes: cepstrum, pseudo-autocovariance, cross-cepstrum, and
saphe cracking, Proc. Symp. Time Series Analysis, Rosenblatt, M., ed., Wiley,
New York, 1963, p.209-243.
9.24. Randall, R. B., Advanced machine diagnostics, Shock and Vibration Digest,
vol.29, no.6, 1997, p.6-26.
9.25. * Primer for Cepstrum analysis a powerful tool for simpler diagnosis of
REB and gear vibrations, Brel & Kjaer Application Note No. BAN0026 -EN-
11.
9.26. Cohen, L., The Time-Frequency Analysis, Prentice-Hall, New Jersey, 1995.
9.27. Kaiser, G., A Friendly Guide to Wavelets, Birkhuser, Boston, 1994.
9.28. Wang, W. J. and McFadden, P. D., Early detection of gear failure by
vibration analysis. Calculation of the time-frequency distribution, Mechanical
Systems and Signal Processing, vol.17, 1993, p.193-203.
9.29. Dalpiaz, G., Rivola, A. and Rubini, R., Effectiveness and sensitivity of
vibration processing techniques for local fault detection in gears, Mechanical
Systems and Signal Processing, vol.14, no.3, 2000, p.387-412.
9.30. Gade, S. and Gram-Hansen, K., Non-stationary signal analysis using Wavelet
Transform, Short-time Fourier Transform and Wigner-Ville distribution, Brel
& Kjaer Technical Review, no.2, 1996.
10.
VIBRATION MEASUREMENT

This chapter describes instrumentation and procedures used for the


measurement and evaluation of vibrations in machine condition monitoring and
fault diagnostics.

10.1 General considerations

The main steps in the evaluation of a particular machine are: a)


determining the most common types of malfunctions, b) determining how these
malfunctions will manifest themselves in terms of mechanical motion, and c)
measuring that motion which is both a reliable indicator of normal machine
performance and which is also most responsive to the primary malfunction
mechanisms.
The most probable malfunctions of a machine are a function of both the
machine design and the function the machine performs in a particular process.
Machines of identical design may have different prime malfunctions due to the
application of each machine in a different process. For example, a compressor may
exhibit unbalance (manifested as increased radial motion) as its prime malfunction
due to the erosion or fouling characteristic of the process gas, whereas an
identically designed compressor may exhibit thrust surge or erratic axial position
changes as its prime malfunction due to a history of uneven process gas flow
through the machine.
In machines with fluid film bearings, rotor-related malfunctions such as
unbalance, misalignment, thrust bearing failure, and rotor instability occur more
often than housing-related malfunctions and foundation problems. Journal
displacements relative to the bearing housing are measured with non-contacting
probes. Absolute shaft displacements are measured for machines with flexible
support structures.
For rolling element bearing machines, the velocity of bearing cap and
casing vibrations is measured using accelerometers or velocity pickups.
76 DYNAMICS OF MACHINERY

10.2 Measurement locations

In this section, criteria are presented for selecting which measurement to


make on a particular machine or for a particular purpose.

10.2.1 General criteria

In large machines with fluid film bearings, particularly those with flexible
rotors and relatively stiff casing, the most frequently occurring malfunctions
(unbalance, misalignment and rotor system instability) manifest themselves as a
change in the shaft motion relative to the housing. The displacement of the journal
relative to the bearing housing is a good indicator of the machine condition.
When the machine has a flexible support structure, the shaft absolute
displacement has to be measured.
Rolling element bearing machines exhibit significant housing motion so
that the absolute r.m.s. velocity of the bearing caps is measured. This is illustrated
in Fig. 10.1.

Fig. 10.1 (from [10.1])

When the machine has a relatively light rotor operating in a heavy stiff
casing (Fig. 10.1, a), most of the energy generated by the rotor is dissipated in
relative motion between the shaft and the bearing. On machines of this type (such
as high pressure centrifugal compressors) with casing to rotor weight ratios of 30:1
or more, the relative displacement between shaft and bearing measured with a
noncontacting probe is the best indicator of the machine condition.
10. VIBRATION MEASUREMENT 77

With the opposite configuration, a relatively heavy rotor running in stiff


bearings supported on flexible structure (Fig. 10.1, b), most of the energy
developed by the rotor is dissipated in the structural motion. On this type of
machine (fans, aircraft derivative gas turbines and machinery fitted with rolling
element bearings), the velocity of the casing vibration is the best measure of
condition.

10.2.2 Shaft precession

A typical noncontacting shaft displacement measuring system contains two


transducers mounted at each bearing spaced 90 degrees apart, as shown in Fig.
10.2.

Fig. 10.2 (from [10.2])

Displacement probes should be mounted in the same plane and facing in


the same direction at each bearing of a multi case machine string. Since it is often
impossible to mount a displacement probe in the horizontal plane, due to
interference from the horizontal splits of the bearing and casing, current practice is
to mount both probes in the upper half of the bearing, 45 degrees either side of the
vertical centerline. In this configuration, the probe on the right (looking from the
driver end) is arbitrarily called the horizontal probe and applied to the horizontal
axis of an oscilloscope to establish the correct orbital motion. The probe on the left
is called the vertical probe. One must of course remember that the oscilloscope
display is tilted 45 degrees from actual shaft motion.
Apart from the probes measuring the shaft radial displacement, a phase
reference probe is usually installed on each shaft. It is a standard displacement
probe, located so that it can observe an once-per-revolution mark on the shaft, such
as a keyway or a hole. The width of the mark should be at least twice the probe tip
78 DYNAMICS OF MACHINERY

diameter and a minimum of 3 mm deep. The probe should be gapped closer than a
standard displacement probe, to generate a high spike at the output of the oscillator
demodulator. The spike may then be applied to the Z axis of an oscilloscope to
produce a blank spot for a phase reference in the waveform and orbital
presentation.
The phase mark may be fed to a tachometer for speed indication. It is used
either as a reference for the horizontal axis of a spectrum plot to construct order
plots, or for phase measurement in balancing, or to calculate a correction for shaft
runout.

10.2.3 Casing vibrations

For many machines, measurements made on non-rotating parts are


sufficient to characterize adequately their running conditions with respect to
trouble-free operation. Measurements should be taken on the bearings, bearing
support housing, or other structural parts which significantly respond to the
dynamic forces and characterize the overall vibration of machine. Typical
measurement locations are shown in Figs. 10.3 to 10.8.

Fig. 10.3 (from [10.3]) Fig. 10.4 (from [10.3])

Figure 10.3 shows the recommended measuring points for pedestal


bearings, while Fig. 10.4 shows the measuring points for housing-type bearings.
Figure 10.5 shows the measuring points for small electrical machines. For
vertical in-line reciprocating engines, the measuring points are shown in Fig. 10.6,
where L and R define the left-hand and right-hand when facing the coupling flange,
1 the machine end of mounting, 2 the crankshaft level, 3 the top edge of
frame, .1 the coupling end, .2 the mid machine, and .3 the free end of
machine.
10. VIBRATION MEASUREMENT 79

Fig. 10.5 (from [10.3]) Fig. 10.6 (from [10.3])

Preferred measurement positions are specified in Fig. 10.7 for


multicylinder V-engines and in Fig. 10.8 for a horizontal opposed cylinder
machine.

Fig. 10.7 (from [10.4]) Fig. 10.8 (from [10.4])

The major precaution to be taken in making measurements of casing


vibrations is to ensure that the transducer mounting is solid and does not have a
natural frequency within the frequency range to be examined. In general,
cantilevered mounts should be avoided as well as mounting on inspection covers or
unsupported areas of bearing caps.

10.3 Measured parameters

The machine construction, the purpose of the measurement and the


frequency range of interest determine the measured variable (displacement,
80 DYNAMICS OF MACHINERY

velocity or acceleration) and value (zero-to-peak, peak-to-peak or r.m.s.), as well as


the transducer which must be used.

10.3.1 Measurement of rotor precession

In measurements made on rotating parts, of the possible measured values,


such as vibration displacement, velocity or acceleration, the vibration displacement
is selected as the most meaningful quantity.
For the complete determination of the shaft motion in a radial plane, two
transducers must be mounted in this plane, spaced 90 degrees apart as shown in
Fig. 10.2. If the motion contains only the fundamental frequency, the displacement
components x(t) and y(t) recorded along the two directions are harmonic, and the
rotor precession orbit is elliptical, as shown in Fig. 10.9. The ellipse major
semiaxis smax is a measure of the shaft vibration severity.

Fig. 10.9 (after [10.5]) Fig. 10.10 (after [10.5])

If the motion consists of the fundamental frequency and the first harmonic,
the displacement components x(t) and y(t) recorded along the two directions are
periodic, and the precession orbit is as shown in Fig. 10.10. The maximum
precession radius smax is a measure of the shaft vibration severity, as defined in the
recommendations VDI 2059 [10.5]. When measurements are made at bearings, this
value can be compared with the bearing clearance.
In the standard ISO 7919 [10.6], which superseeded VDI 2059, the shaft
vibration magnitude is defined as the higher value of the peak-to-peak
displacement measured in two selected orthogonal measurement directions,
10. VIBRATION MEASUREMENT 81

[ ]
max x pp , y pp . Peak to peak displacement amplitude has enjoyed success
because it allows calculation of percentage of bearing or seal clearance, a very
important correlation on nearly all rotating machinery [10.7].

Fig. 10.11

When the rotating assembly is five or more times heavier than the case of
the machine, the shaft absolute displacement is of interest. It can be measured in
two ways: a) electronically summing the signals of both an eddy current probe
measuring relative shaft motion with respect to the bearing, and an accelerometer
measuring case absolute displacement (integrated twice) (Fig. 10.11), and b) using
a shaft rider, which is a spring mounted device that physically rides on the surface
of the shaft, normally a velocity sensor mounted on top of the shaft rider whose
output is integrated electronically to displacement.

10.3.2 Measurements on bearings

In measurements made on non-rotating and, where applicable, non-


reciprocating parts of complete machines, it is common practice to consider the
root mean square (r.m.s.) value of broad-band vibration velocity, since this can be
related to the vibration energy. Moreover, in the range 600 12000 rpm it is
relatively independent of frequency, and thus yields a simple measure of severity
for a new operating machine.
The standard ISO 10816 [10.3] defines the vibration severity as the
highest value of the broadband r.m.s. value of the velocity amplitude in the
frequency range 10-1000 Hz, as evaluated on the structure at prescribed points,
under agreed machine support and operating conditions.
82 DYNAMICS OF MACHINERY

For most machine types, one value of vibration severity will characterize
the vibratory state of that machine. However, for some machines this approach may
be inadequate and the vibration severity should then be assessed independently for
measurement positions at a number of locations.

10.3.3 Displacement, velocity or acceleration

A common value used by the proponents of velocity measurement for a


machine operating with reasonable allowable vibration is 6.25 mm s zero to peak.
At a running speed of 3000 rpm, this yields a displacement of 40 m peak-to-peak
and an acceleration of 0.2 g zero to peak.
Now assume that a vibration is generated at one-third running speed
frequency with a displacement amplitude of 80 m peak-to-peak, twice the
displacement at running speed. The 1X component could be from unbalance while
the 1 3 X component could be the result of a slight looseness or rub condition, oil
whip, the excitation of a resonance, or several other malfunctions.
A displacement of 80 m peak-to-peak at 1000 rpm yields a velocity of
4.16 mm s zero to peak and an acceleration of 0.044 g zero to peak. If the two
signals were in-phase (a rare case) such that the separate amplitudes would be
additive, then the increase in vibration levels would be as follows: displacement,
40 m to 120 m or 200%; velocity, 6.25 mm s to 10.41 mm s or 68%; and
acceleration, 0.2 g to 0.244 g or 22% [10.8].
If the measurement being made is shaft precession, then this means that
the shaft is now 3 times closer to the internal clearances of the machine than it was
before the subsynchronous vibration occurred. In terms of velocity, the situation is
less than 2 times as bad, and the small increase in acceleration may even go
unnoticed by a casual operator.
If the measurement is being made on the machine support structure, the
evaluation of terms is even more critical. In most cases, the rotor amplitude will
have increased even more than was measured on the machine casing.
For diagnostic purposes in machines with fluid film bearings,
displacement is the most direct indicator of the relative severity of vibrations at
various frequencies. In the above example, if a tunable filter were used to sort the
various frequencies according to their respective amplitudes, the following analysis
would result. Displacement would show the 1 3 X component to be predominant
and 2 times the amplitude of the 1X component. Velocity would show the 1X
predominant with the 1 3 X component being 2 3 the amplitude of the 1X, and
10. VIBRATION MEASUREMENT 83

acceleration would indicate the 1 3 X component to be only about 1 5 the


amplitude of the 1X motion (Fig. 10.12).

Fig. 10.12

The shaft precession will more commonly generate these sub


synchronous frequencies, and thus the amount of shaft deflection (measured in
displacement), relative to the machine clearances, is the most important parameter
to evaluate in terms of vibration severity and the significance of the various
vibration frequency components.

10.3.4 Peak-to-peak vs. r.m.s.

Shaft displacement is usually expressed in terms of the peak-to-peak value


while casing vibration velocity is expressed in terms of the r.m.s. value.
The root-mean-square (r.m.s.) is an average of the composite waveform.
For a sine wave with unit amplitude, the r.m.s. value would be 0.707. If the sine
wave amplitude doubled to 2, the r.m.s. value would also double to 1.414.
However, the r.m.s. average responds in a linear manner for pure sine waves only.
If a waveform acquires an additional frequency component such that the total
amplitude increases, then the r.m.s. value could either increase or decrease as a
result of the new shape of the waveform. For example, the r.m.s. value of a square
wave is less than the peak value.
A harmonic vibration expressed in terms of velocity v (t ) = vo cos t is
defined by the amplitude vo and the circular frequency (Fig. 10.13, a). The
amplitude is sufficient to define the magnitude of harmonic vibrations.
In the case of periodic vibrations (obtained by the summation of several
harmonic components), the maximum value is called the peak value, v p (or zero-
to-peak value). In most cases it is simpler to measure the peak-to-peak value, v pp .
84 DYNAMICS OF MACHINERY

The root mean square (r.m.s.) value of velocity may be calculated as


follows
T
1
v (t )dt ,
2
vr .m .s . = (10.1)
T
0

where v (t ) is the instantaneous value, and T is the sampling time, which is longer
than the period of any of the major frequency components.
For non-periodic steady-state vibrations, the r.m.s. value is defined as

t
1
v (t )dt .
2
vr .m .s . = lim t (10.2)
t
0

For harmonic vibrations of velocity amplitude vo , the following relations


can be established

1
vr .m.s . = v0 = 0.707 v0 , v p = v0 , v pp = 2 v0 . (10.3)
2

a b
Fig. 10.13 (after [10.9])

Figure 10.13, b illustrates the addition of a harmonic component having the


same amplitude but a frequency 10 times higher than the initial harmonic vibration.
The resulting periodic vibration would have a peak value almost twice that of the
initial component, but an r.m.s. value only 1.4 times larger.
Figure 10.14 shows the influence of the relative phasing on the
compounding of two harmonic components. The higher harmonic has half the
amplitude of the fundamental component and a frequency 3 times higher.
10. VIBRATION MEASUREMENT 85

Though the peak values are different in the two cases, v2 p 1.4 v1 p , the
r.m.s. values are the same, v2 r.m.s. = v1 r.m.s. .

This means that the use of the r.m.s. vibration as a measure of the vibration
severity gives better results in comparisons with allowable limit values than in
detecting developing malfunctions by monitoring the change of broad-band
vibration magnitude.

Fig. 10.14 (after [10.9])

While the standard ISO 7919, based on VDI 2056, recommends the
measurement of the r.m.s. velocity on the bearing cap, some API standards and
[10.10] recommend the measurement of zero-to-peak velocity.

10.4 Transducers and pickups

The selection, placement and proper use of the correct transducer are
important steps in the implementation of a condition monitoring and fault
diagnostics program.

10.4.1 Transducer selection

When measuring vibrations, one of the most important factors in obtaining


accurate information involves selecting the proper vibration transducer. Important
basic considerations include: a) type of machinery to measure, b) frequency range
to be measured, c) environmental considerations, and d) permanent or portable data
collection.
Figure 10.15 illustrates typical frequency regions of operation for different
transducers. For constant velocity vibration amplitude across all frequencies, a
displacement transducer is more sensitive in the lower frequency range, while an
accelerometer is more sensitive at higher frequencies.
86 DYNAMICS OF MACHINERY

Fig. 10.15 (after [10.1])

For a signal with a velocity level of 6 mm/s, the displacement amplitude is


1m at about 1000 Hz and disappears into the background noise of most
commercially available measuring systems. In an extreme case, the 6 mm/s
velocity at 10 kHz corresponds to an acceleration level of 400 m s 2 , i.e.
approximately 40 g with a displacement of only 0.1m . Displacement is not an
effective means of measuring high-frequency vibration because large forces are
required at these frequencies to produce a measurable displacement and not
because the instrumentation system is limited to a maximum frequency. Below
10. VIBRATION MEASUREMENT 87

about 20 Hz, the displacement amplitude necessary to produce an easily


identifiable acceleration signal is so large that endangers the transducer mechanical
integrity.
Thus, it is recommended to use displacement pickups from 0 to about 1000
Hz, velocity pickups from 10 to 3000 Hz and accelerometers from 20 Hz to well
above 20 kHz, setting the lower limit for acceleration measurement at 0.4 m s 2 ,
and that for displacement at 2 m [10.1]. The trend is the extension of acceleration
measurement at lower frequencies.

Fig. 10.16 (from [10.11])

Once an accelerometer is selected for vibration measurement, the first


critical consideration is to make sure it has a frequency range that includes the
potential machine fault frequencies. The second key factor is to make sure the
accelerometer will perform in the environment intended to be used. Some of the
considered factors are: a) heat tolerance, b) moisture, c) chemical exposure, d)
electrical interference, e) intrinsically safe requirements, and f) shock limit.
Recommended ranges for displacement, velocity, and acceleration
transducers are specified in [10.10] as well as frequency ranges of operation.
Figure 10.16 illustrates the typical envelope ranges of velocity vs. frequency for
several type transducers.
Traditional vibration sensors fall into three main classes: a) noncontact
displacement transducers, b) velocity pickups, and c) piezoelectric accelerometers
[10.12].
88 DYNAMICS OF MACHINERY

10.4.2 Eddy current proximity transducers

Eddy current proximity devices are used in noncontacting vibration and


axial position monitoring systems [10.13, 10.14]. They consist of an eddy current
transducer (probe), a cable and an oscillator-demodulator (proximitor) (Fig. 10.17).

Fig. 10.17 (from [10.15])

The probe is a sensor that translates distance (gap) to voltage. It may be


used to measure dynamic motion and static gap. The eddy probe is a flat pancake
coil of wire molded into a universal threaded case. Usual probe tip diameter is 5
mm, with a body diameter of 8 mm and length of 25 mm.

Fig. 10.18

A typical eddy current transducer contains two coils: an active coil and a
balance coil. The active coil senses the presence of a nearby conductive object,
10. VIBRATION MEASUREMENT 89

while the balance coil is used for temperature compensation and to balance the
output bridge circuit. The lead wire is a single conductor shielded cable [10.16].
The eddy probe driver generates a high frequency signal to the eddy probe
and converts the return signal to a voltage which can then be displayed on a read-
out monitor and used for comparison of vibration levels with alarm set points.
When the appropriate voltage is supplied to the eddy probe driver, it
becomes an oscillator and generates a high frequency signal to the coil in the tip of
the eddy probe. The coil creates a small magnetic field that induces eddy currents
in metal targets (Fig. 10.18). These eddy currents absorb part of the energy and
change the sensors oscillation amplitude. As the gap narrows, more and more
energy is absorbed until finally the voltage output drops to zero at, or near, contact.
A typical response graph of gap vs. voltage (Fig. 10.19) shows the
sensitivity of 200 mV mil (8 mV m ) over the range of 100 mils (2.5 mm ) for a
standard supply voltage of minus 24V of direct current. Note that even though the
eddy probe has a response in excess of 2.5 mm , the linear response which ends the
useful range stops at about 2.5 mm .

Fig. 10.19 (from [10.15])

The direct current voltage output may be measured by a voltmeter, and by


referring to the probe calibration curve the exact gap between the eddy probe and
the observed surface can be determined. If the observed surface is moving, as with
a precessing rotating shaft, the signal is not constant but varies in proportion to the
amplitude of the movement. Therefore, both the negative d.c. voltage which gives
average gap distance, and the a.c. component which gives dynamic motion are
available.
90 DYNAMICS OF MACHINERY

This dynamic measurement provides not only the amplitude of the peak-to-
peak vibration, but also the frequency and waveform of the motion. This
information is of utmost significance in both monitoring and machinery
malfunction diagnostics.
The eddy current measurement is not disturbed by non-conductive material
in the gap between the probe and its observed surface, so that oil, steam and gases
do not adversely affect the measurement. The main disadvantage of proximity
probes is the sensitivity to shaft mechanical and electrical runout (glitch).
Mechanical runout is shaft eccentricity and depends on the manufacturing
tolerance. Electrical runout is a false indication of relative displacement due to
shaft anomalies (magnetization or internal stresses) and is indistinguishable from
actual displacement.
A single transducer mounted radially at one bearing provides the vibration
signal in only one plane. In order to obtain the shaft precession orbit, it is necessary
to mount two probes at 90 0 to each other (Fig. 10.20), at the same radius.

Fig. 10.20 (after [10.17])

In machine monitoring applications, an eddy current transducer is used as a


phase reference probe observing an once-per-turn discontinuity (notch, hole, pin,
keyway), to provide phase angle orientation information. This key-phase probe
provides both speed and phase references, and a timer pulse for use with peak-to-
peak eccentricity measurements.
10. VIBRATION MEASUREMENT 91

An oscilloscope is used with X vs. Y (orbit) display and Z-axis intensity.


The instrument must display voltage waveforms from vibration transducers in time
base and orbit format. The key-phase signal connected to the Z-axis intensity input
can be used to trigger the oscilloscope and to provide a reference point from which
to make phase angle measurements.
The horizontal probe is connected to the positive polarity input of the
horizontal amplifier of an oscilloscope (Fig. 10.20). The vertical probe output is
connected to the vertical amplifier jack. The key-phase probe output signal is
connected to the trigger jack of the oscilloscope for synchronizing the scope, but
more importantly, it is connected to the Z-axis input of the Cathode Ray Tube. This
connection must be a.c. coupled to the Z-input, and the normal ground shorting bar
must be removed. If the scope is a.c. coupled at this point, there is no problem. If it
is d.c. coupled, a capacitor is needed in series with the signal. Shielded coaxial
cables should be used, and only one earth ground is to be used for all equipment.
The key phase signal is superimposed on the time base and orbit traces
producing a bright/blank (or blank/bright) key phase mark. A notch-type keyphasor
will produce a voltage pulse which goes negative and then positive. As the notch
enters the probe face area (increase in gap) a boost in negative voltage (with
negative slope) is produced. When the notch trailing wall passes the probe face
(decrease in gap) a less negative burst (positive slope) of voltage occurs. As the
gap changes through the notch start and notch end, the time base waveforms and
the orbit trace are interrupted with a blank (break) mark, followed by a bright mark
[10.18].
If the shaft rotates clockwise, then the blank/bright sequence on the orbit
should also be clockwise if the shaft precession is forward. A bright/blank
sequence would indicate backward precession. The usual old convention was to
view, from outboard, the driving end facing the driven machine.
Eddy current proximity sets can be used to monitor rotor-to-stator
differential expansions and rotor positions relative to the thrust bearing.

10.4.3 Velocity pickups

Velocity sensing devices are either seismic pickups or fixed reference


instruments including electrodynamic transducers. In an electrodynamic
transducer, a coil moves through the magnetic field produced by a stationary
permanent magnet. The transducer can also be designed with a stationary coil and
the permanent magnet core moving within the coil. The principle of operation is
the same.
When the core moves, magnetic lines of the field created by the core cross
the turns. The electromotive force induced in the turns is proportional to the speed
of the core. The unit thus produces a signal directly proportional to vibration
92 DYNAMICS OF MACHINERY

velocity. It is self-generating and needs no conditioning electronics in order to


operate, and it has relatively low electrical output impedance making it fairly
insensitive to noise induction.
A vibration pickup consists of a seismic mass supported by two
membranes, so that part of the mass lies within the air gap of a magnetic circuit.
The velocity pickup is a seismic instrument fastened to a vibrating structure. At
frequencies above the resonance of the mass-spring system, the relative motion
between the mass and casing sensed by the transducer is essentially the same as the
motion of the structure under test. The seismic mass and the pickup casing vibrate
1800 out of phase. Relative to a fixed (inertial) reference frame, the mass remains
nearly stationary (becomes a fixed point) and the casing motion is measured with
respect to it. The amplitude of the e.m.f. induced into the measuring coil is
proportional to the velocity of the relative motion and hence to the vibration
velocity of the structure under test.

Fig. 10.21 (from [10.19])

The measuring coil, the damping cylinder and the additional damping coil
are supported in the air gap. The damping cylinder reduces the influence of the
transducers natural frequency on the measuring signal. The additional damping
10. VIBRATION MEASUREMENT 93

coil can be energized to compensate for possible reductions in damping at high


temperatures or to compensate for static sag if the transducer is used in a vertical
attitude.
A correction coil, wound round the magnetic flux source, i.e. permanent
magnet, eliminates the influence of eddy-current damping on the flux. Limit stops
are fitted to prevent excessive movement of the seismic mass.
The electrodynamic velocity pickup PR 9266 made by Philips is shown in
Figure 10.21 where: 1 permanent magnet, 2 correction coil, 3 measuring coil,
4 additional damping coil, 5 damping cylinder, 6 and 7 membranes, 8
casing, 9 output leads, 10 three-core screened cable, 11 and 12 limit stops.
The frequency range is 10 to 1000 Hz, for displacement amplitudes up to 1 mm and
accelerations up to 10 g. The undamped natural frequency is 12 Hz. The mass
without cable is about 0.5 kg. The sensitivity is 30 mVpp mm s at 110 Hz.

Another type of velocity transducer consists of an accelerometer with a


built-in electronic integrator. This unit is called a "velometer", and is by all
accounts superior to the classic seismic velocity probe

Fig. 10.22

In spite of these advantages, the velocity pickup has many disadvantages


that make it nearly obsolete for new installations, although there are many
thousands of them still in use today. It is relatively heavy and complex and thus
expensive, and it has poor frequency response, extending from about 10 Hz to 1000
Hz. The spring and the magnet make up a low-frequency resonant system with a
natural frequency of about several Hz (Fig. 10.22). This resonance needs to be
highly damped to avoid a large peak in the response at this frequency. The problem
is that the damping in any practical design is temperature sensitive, and this causes
the frequency response and phase response to be temperature dependent.
94 DYNAMICS OF MACHINERY

10.4.4 Accelerometers

Due to their advantages light-weight, ruggedness, wide frequency


response, good temperature resistance and moderate pricing piezoelectric
accelerometers are the most often used vibration sensing instruments. They are
made in several different configurations, but the compression-type, illustrated in
Fig. 10.23, serves to describe the principle of operation. This accelerometer is a
seismic pickup in which the sensing piezoelectric ceramic discs form the elastic
element of the spring-mass system.

Fig. 10.23 (after [10.20])

The seismic mass is clamped to the base by an axial bolt bearing down on a
circular spring. The piezoelectric element is squeezed between the mass and the
base. When the accelerometer is subjected to vibrations, the mass will exert a
variable force on the piezoelectric discs. The charge developed across the
piezoelectric discs is proportional to the applied force, which in turn is proportional
to the acceleration of the mass. For frequencies much lower than the resonance
frequency of the accelerometer assembly, the acceleration of the seismic mass is
equal to the acceleration of the whole pickup.
Accelerometers have a very large dynamic range. The smallest acceleration
levels they can sense are determined only by the electrical noise of the electronics,
and the highest levels are limited only by the destruction of the piezo element
itself. Acceleration levels can span an amplitude range of about 108 , which is 160
dB.
The frequency range of the accelerometer is very wide, extending from
very low frequencies in some units to several tens of kilohertz. The high-frequency
10. VIBRATION MEASUREMENT 95

response is limited by the resonance of the seismic mass coupled to the springiness
of the piezo element. This resonance produces a very high peak in the response at
the natural frequency of the transducer, and this is usually somewhere near 30 kHz
for commonly used accelerometers.
A rule of thumb is that an accelerometer is usable up to about 1/3 of its
natural frequency. Data above this frequency will be accentuated by the resonant
response, but may be used if the effect is taken into consideration. The lower limit
is determined by cable and preamplifier. The frequency response curve of an
accelerometer is presented in Fig. 10.24.

Fig. 10.24 (from [10.20])

Most accelerometers used in industry today are of the "ICP" type, meaning
they have in internal integrated circuit preamplifier. This preamp is powered by a
dc polarization of the signal lead itself, so no extra wiring is needed. The device the
accelerometer is connected to needs to have this d.c. power available to this type of
transducer. The ICP accelerometer will have a low-frequency roll-off due to the
amplifier itself, and this is usually at 1 Hz for most generally available ICP units.
There are some that are specially designed to go down to 0.1 Hz if very low
frequency data is required.
The resonant frequency of an accelerometer is strongly dependent on its
mounting. The best type of mounting is always the stud mount - anything else will
reduce the effective frequency range of the unit.
When mounting an accelerometer, it is important that the vibration path
from the source to the accelerometer is as short as possible, especially if rolling
element bearing vibration is being measured. If an accelerometer is mounted on a
surface that is being strained (bent), the output will be altered. This is known as
base strain, and thick accelerometer bases are used to minimize this effect. Shear-
type accelerometers are less sensitive because the piezoelectric crystals are
mounted to a center post and not to the base.
96 DYNAMICS OF MACHINERY

10.4.5 Summary about transducers

The recommended transducer types and their locations and directions for
various type machines are given in Table 10.1 ([10.21] and Annex A of [10.6]).

Table 10.1

Machine Evaluation Measurement


Transducer type Direction
type parameters locations
Noncontacting
Large
Relative transducer Radial
steam
displacement Shaft, at each 45 deg
turbine
or absolute Noncontacting and bearing or
generator
displacement seismic transducer X and Y
sets with
combination
fluid film
bearings
Velocity or Velocity transducer Each bearing Radial
acceleration or accelerometer housing X and Y
Power Shaft axial Noncontacting
generation transducer or Thrust collar Axial Z
displacement
axial probe
Phase Eddy
reference current/inductive/optical Shaft Radial
and rpm transducer

Medium Radial
Relative Noncontacting Shaft, at each 45 deg
and small
displacement transducer bearing or
industrial
steam X and Y
turbines Each bearing
with fluid Velocity or Velocity transducer housing and Radial
film acceleration or accelerometer turbine X and Y
bearings housing

Shaft axial Noncontacting


transducer or Thrust collar Axial Z
displacement
Mechanical axial probe
drive
Phase Eddy
reference current/inductive/optical Shaft Radial
and rpm transducer
10. VIBRATION MEASUREMENT 97

As a summary of the characteristics of the different types of transducers,


the following comparison is reproduced from [10.8].
Proximity probes
Advantages: a) measures directly the motion of the shaft (the origin of
most large machine vibrations), b) measures in terms of displacement (the most
meaningful engineering unit for fluid film bearing measurements), c) measurement
is noncontact (will not influence the measured vibratory motion because of
contact), d) solid-state with no moving parts, e) one sensor simultaneously
measures both dynamic motion and (average) position, f) system is modular with
the most inexpensive part, the probe, requiring only occasional replacement
(because of abuse), g) one extra transducer can be used as a rotor speed sensor and
a phase reference, h) excellent frequency response, i) small size, j) well-suited to
most machinery environments, k) ease of calibration, l) accurate low frequency
amplitude and phase angle information, and m) high level low impedance output.
Disadvantages: a) control of observed shaft surfaces desirable to avoid
excessive sensitivity to shaft mechanical and electrical runout, b) somewhat
sensitive to various shaft materials, c) requires an external power source, and d)
sometimes difficult to install.

Velocity pickups
Advantages: a) ease of installation due to external machine mounting, b)
strong signal in the mid-frequency range, c) some are suitable for relatively high
temperature environments, and d) no external power required.
Disadvantages: a) relatively large and heavy, b) manufactured as a unit so
that a transducer fault requires replacement of the entire pickup, c) sensitive to
input frequency (tendency to emphasize higher frequencies), d) relatively narrow
frequency response with amplitude and phase errors at low frequencies, e) has
moving parts and is expected to degrade under extended normal use, f) difficult to
calibrate, g) measures dynamic motion only (not static position), and h) can
respond with excessive cross-axis sensitivity at high amplitude levels.

Accelerometers
Advantages: a) ease of installation due to external machine mounting, b)
good frequency response (especially at high frequencies, although this could be a
disadvantage by increasing the noise level from various external vibrations), c)
small and light weight, d) some are suitable for relatively high temperatures, and e)
strong signal in the higher frequency ranges.
Disadvantages: a) most sensitive to input frequencies (although this can be
an advantage when measuring very high frequencies), b) difficult to locate on the
machine case for a meaningful measurement, c) very sensitive to the method of
attachment, d) output requires amplification, e) most sensitive to spurious
98 DYNAMICS OF MACHINERY

vibrations (confusing the acquired data and making exact mounting location
difficult), f) impedance matching (or charge amplifier) is needed, and g) normally
requires some filtering for monitoring applications.

10.4.6 Placement of transducers

Figure 10.25 shows a layout of a machine protection system that is


common in many oil and petrochemical plants, pumping stations, etc. It consists of
radial vibration, axial position, and speed monitoring of shafts, plus radial and axial
vibration of machine case and possibly of piping and foundation.

Fig. 10.25 (from [10.15])

This full-time system, as outlined, provides monitoring of: a) speed, b)


two-plane radial vibration of the shaft at each machine journal and each gear shaft,
plus case radial vibration on the gear, and c) shaft axial position (dual or single) on
all shafts (for protection against excessive thrust deflection).
In addition, the following are available from the permanently mounted
transducers for periodic monitoring and analysis information: a) shaft orbits, b)
gear case orbits, c) phase of the turbine, gear, and compressor shafts, d) axial
vibration, and e) eccentricity, or average position, of shafts. With the addition of a
roving velocity transducer, case, foundation and piping vibrations can also be
periodically monitored.
10. VIBRATION MEASUREMENT 99

Fig. 10.26 (from [10.19])

Figure 10.26 shows the layout of a comprehensive monitoring system for a


large turbine-driven compressor.

Fig. 10.27 (from [10.19])

Figure 10.27 shows the monitoring system for a turbo-generator.


100 DYNAMICS OF MACHINERY

Fig. 10.28 (from [10.2])

Figures 10.28 indicate the location of seismic pickups used for periodic
measurements on a motor driven fan using: a) elastic coupling and low shaft (Fig.
10.28, a), and b) belt-driven high shaft (Fig. 10.28, b).

10.4.7 Instrumentation

Transducer signals are processed by a wide variety of electronic


instruments. Conditioners, which include filters, analog integrators, and amplifiers,
are used to enhance data. Digital recorders, tape recorders and digital computers
are used for data recording, especially when transient vibration phenomena are
measured. Electronic data collectors are used for storing r.m.s. or peak data that
can be transferred to a digital computer; trends can then be established and reports
generated.
Data processing is carried out using tunable and swept-filter analyzers,
tracking filters and FFT spectrum analyzers. FFT analyzers acquire a block of data
over a designated frequency range during a period of time, digitize the data, and
perform a frequency analysis using the FFT algorithm. They contain buffers
capable of storing large quantities of data and can also produce spectrum cascade
(waterfall) diagrams, i.e. amplitude versus frequency for various times or speeds.
These analyzers can perform integration, r.m.s. band analysis, and compute power
spectral density. The magnitudes of r.m.s.-based bands can be displayed in a linear
or logarithmic format.
Data display instruments include monitors, oscilloscopes, strip chart
recorders, analog and digital plotters. Simpler instruments display r.m.s., peak, or
average values of measured vibration.
Apart from the mentioned transducers, optical (and magnetic) pickups are
used in torsional vibration measurements, and general speed and phase
measurements on rotating shafts. The optical pickup sends a voltage pulse to the
oscilloscope or analyzer when energized by light pulses from a reflective tape (or
10. VIBRATION MEASUREMENT 101

other marks) bonded on the shaft. The optical system includes a power supply and
an amplifier.

10.5 Data reduction

Vibration data are processed and reduced into interpretable formats to help
the malfunction identification process [10.22].

10.5.1 Steady state vibration data

Steady-state vibration data can be reduced in several useful formats.

10.5.1.1 Orbits and time base plots


Orbits and time base plots are useful for examining the magnitude,
frequency, phase angle, and shape of the shaft precession motion and its filtered
frequency components (Fig. 10.29).

Fig. 10.29 (from [10.22])

Their interpretation makes it possible to determine the precession


directivity (forward or backward), the bearing pre-loading (orbit distortion) and the
existence of sub-harmonic and supra-harmonic components.
Simple orbits are elliptical or Lissajous figures. The classical Lissajous
figures may be obtained by compounding two perpendicular harmonic motions
with two different frequencies.
In Fig. 10.30 it is shown how the orbit is built up from the two components

x = A cos t , y=
3
4
( )
A cos 2 t 450 .
102 DYNAMICS OF MACHINERY

The resulting orbit is known as a butterfly or rabbit ears. In the orbit


construction, the time steps are labelled 1, 2, 3, etc. Thus the direction of orbital
motion can easily be determined.

Fig. 10.30 (from [10.23])

While Lissajous figures result from harmonic motions with different


frequencies, actual rotor steady-state orbits result from two perpendicular
components which both are periodic motions (sums of harmonic motions), and
usually carry the same set of frequencies. Thus, if x contains 1X and 2X
components, then most probably y will have 1X and 2X components, though
with different amplitudes and phases.

Fig. 10.31 (from [10.24])


10. VIBRATION MEASUREMENT 103

Some machine faults generate periodic vibrations with sub- or


superharmonic components.
When the rotor vibration has a subharmonic component of order 1/N, the
complex vector of the precession radius (for zero phase angles) is of the form

z = R1 ei t + R1 N e i ( N)t
, (10.4)

where R1 is the amplitude of the synchronous component (due to unbalance) and


R1 N is the amplitude of the subsynchronous component. The plus sign is for
forward precession while the minus sign is for backward precession.
For N = 2 , and R1 R1 2 = 2 (dominant synchronous component), the orbit
is shown in Fig. 10.31, a for a forward subharmonic component, and in Fig. 10.31,
b for a backward subharmonic component. When the two components have
different phase angles, the orbits are no more symmetrical.

Fig. 10.32 (from [10.24])

The corresponding orbits for N = 2 , and R1 R1 2 = 1 2 (dominant


subsynchronous component) are shown in Fig. 10.32.
Similar conclusions result from the analysis of periodic vibrations with
superharmonic components, replacing N by 1/N in equation (10.4).
Generally, orbits have external loops when the so-called 2X component is
mainly due to shaft misalignment (including gear mesh and belt drives), coupling
misalignment (Fig. 10.33) and resulting radial preload. Orbits with internal loops
are mainly due to shaft asymmetry (such as with cracked shafts) together with
radial preload (from misalignment, gravity or fluid flow). The effect of the radial
preload on the shape of steady-state orbits is shown in Fig.11.9 (Chapter 11). With
increasing force, the initial elliptical orbit may become banana shaped, then
figure eight shaped.
104 DYNAMICS OF MACHINERY

10.5.1.2 Half spectrum plots


The direct (unfiltered) orbit and timebase plots of the vibration signals
measured at a particular location on a machine are quite complex. The orbit is far
from elliptical and the timebase traces are combinations of several harmonic
components (fig. 10.33, a). The half spectrum plot (Fig. 11.33, b) is a frequency
domain version of the timebase plot for the Y probe. It makes the identification of
frequencies and individual component amplitudes easier.

a b
Fig. 10.33 (from [10.25])

Most frequency analysis instruments display only the positive half of the
frequency spectrum, because the spectrum of a real-world signal is symmetric
around d.c. Thus, the negative frequency information is redundant.
When only one vibration measurement (e.g., the vertical component) is
made at a given point, the half spectrum plot is useful, e.g. for tracking changes in
the spectral content over a period of time, provided timebase plots are available to
check the vibration signal quality. Half spectrum plots reveal new frequency
components and changes in the magnitude of previous data at a particular
frequency. The phase information is lost.
The half spectrum has been intensively used as a machine signature for
assessing the machine condition, correlating the frequency and magnitude of the
peaks with specific machine faults.

10.5.1.3 Full spectrum plots


In the general spectral analysis, the two-sided spectrum shows both the
positive and negative frequency components of a signal. In machine diagnostics,
full spectrum plots use data from two orthogonal transducers converted into
information about the magnitude, frequency and phase of the directional (forward
and backward) response components. Forward components are shown in the
positive half of the full spectrum plot, and backward (reverse) components are
shown in the negative half. The relative magnitude of the forward and backward
components of equal frequency defines the direction of the precession.
10. VIBRATION MEASUREMENT 105

Fig. 10.34 (from [10.25])

The full spectrum plot for the example considered in the previous section is
shown in Fig. 10.34. Though the right hand side of the full spectrum plot appears to
be the same as a half spectrum plot for one probe, this is not true. Only forward
components are shown on the positive side of the full spectrum plot.

a b
Fig. 10.35 (from [10.26])

a b
Fig. 10.36 (from [10.26])
106 DYNAMICS OF MACHINERY

a b
Fig. 10.37 (from [10.26])

In the full spectrum plot, a forward circular orbit is represented by a


component present only on the right hand side. A backward circular orbit has a
component only on the left hand side.
When the components on the right and left hand sides are equal or of
different amplitudes, the orbit at that frequency is elliptical. The component with
the larger amplitude determines the direction of precession along the orbit. When
the forward component is larger (Fig. 10.35), the precession is forward. When the
backward component is larger (Fig. 10.36), the precession is backward. When the
amplitudes are equal (Fig. 10.37), the orbit degenerates into a straight line.

a b c
Fig. 10.38 (from [10.26])

The advantages of using a full spectrum plot are apparent in cases where
two different machinery malfunctions produce the same half spectrum (Fig. 10.38,
a). The full spectrum for oil whirl and whip is shown in Fig. 10.38, b, and that of a
rub is shown in Fig. 10.38, c. In both cases there is an X subharmonic
component, but for whirl/whip it is forward, while for the rub is backward.

10.5.1.4 Mode shape plots


Mode shape plots (Fig. 10.39) display the precession orbits at selected
sections along the rotor, and the dynamic deflected line at a given moment. They
provide estimates of the nodal points along the rotor and for the internal clearances
between the rotor and stator.
10. VIBRATION MEASUREMENT 107

Fig. 10.39 (from [10.5])

10.5.1.5 Trend plots


Trend plots are used for analyzing changes of observed data as a function
of time (Fig. 10.40) determined by modifications in the operating parameters of the
machine. They display both global level vibration data and other monitored
parameters useful in the machine condition diagnostics.

Fig. 10.40 (from [10.27])


108 DYNAMICS OF MACHINERY

10.5.2 Transient vibration data

Transient vibration data taken during start-up and shutdown can be reduced
in different formats.

10.5.2.1 Bod and polar plots


Bod plots (Fig. 10.41) and polar plots (Fig. 10.42) reveal the critical
speeds, the dynamic distorted shape and the mode shapes of the rotor, and the
amplification factor at the synchronous frequency of the rotor-bearing system. The
polar diagrams of the synchronous component 1X (filtered) are useful in the
multiplane field balancing and the detection of cracked shafts.

Fig. 10.41 (from [10.27])

Fig. 10.42 (from [10.27])


10. VIBRATION MEASUREMENT 109

10.5.2.2 Cascade half spectrum plots


Waterfall diagrams are used for examining the variation of the spectral
components (synchronous, sub- and supersynchronous) with speed. This permits
the detection of some instabilities (such as oil whirl/whip), of cracked shafts and
rubs. Figure 10.43 illustrates a cascade spectrum plot and orbits of a cracked shaft.

Fig. 10.43 (from [10.27])

10.5.2.3 Full spectrum cascade plots


Full spectrum cascade plots display the speed evolution of the directional
(forward and backward) components of the instantaneous precession radius.

Fig. 10.44 (from [10.28])


110 DYNAMICS OF MACHINERY

Figure 10.44 shows a normalized full spectrum cascade for a rotor in fluid
film bearings. The resonance peaks can be seen on the 1 lines (denoted 1X )
at the critical speed. The backward component is produced by the bearing
anisotropy. For a rotor in rolling element bearings this component is missing. At
high running speeds, above the onset speed of instability tresh , the unbalance
response is dominated by the component with frequency equal to the lateral natural
frequency of the rotor system.

Fig. 10.45 (from [10.29])

Figure 10.45 is the full spectrum cascade plot of a preloaded shaft. It


displays the known characteristics, i.e. forward precession, different amplitude
forward and backward components, denoting an elliptical orbit, and primary 1X
component for the entire machine startup.
Other phenomena, such as preload induced rubs, cracked shafts or fluid
produced instabilities are conveniently analyzed using such plots (see Chapter 11).

10.5.2.4 Journal center position diagrams


Diagrams of the shaft centerline position (Fig. 10.46) are useful for
observing changes in the steady-state position of the rotor in the bearings. They
give indications of bearing wear and major changes in the alignment state of the
machine, indicating rotor preload due to misalignment or thermal effects.
Shaft centerline plots warn about a journal operating near or above the
center of the bearing, a usual cause of instability.
10. VIBRATION MEASUREMENT 111

Fig. 10.46 (from [10.27])

10.5.2.5 Time variation diagrams


Graphs of the time variation of the broadband vibration amplitude (Fig.
10.47) permit to avoid the rotor thermal bow at start-up or when driven through the
turning gear (due to nonuniform heating). They are useful in the detection of severe
vibrations produced by rapid temporary variations of the steam temperature (due to
boiler malfunction) or by partial admission, which affects the journal average
position in the bearing, hence the stability of the precession motion.

Fig. 10.47 (from [10.30])

10.5.2.6 Acceptance region plots


Acceptance region plots are displays where filtered 1X (or 2X) vibration
vectors are shown as a trend in polar format (Fig. 10.48). A user-defined normal
operating range of the 1X vibration vector is determined within the polar plot to
form what is called an acceptance region. Deviation of the 1X vibration vector
tip from the acceptance region may be a vital warning of a shaft crack or other
rotor disturbances.
112 DYNAMICS OF MACHINERY

Fig. 10.48 (from [10.31])

Other display formats include a) d.c. gap voltage plots (for proximity
probes), b) axial thrust position plots, c) rpm versus time plots and d) multiple orbit
plots [10.32, 10.33].

References

10.1. Mitchell, J. S., An Introduction to Machinery Analysis and Monitoring, Penn


Well Books, Tulsa, 1993.
10.2. * Vorbeugende Maschineninstandhaltung, Schenck Seminar C 50, Nov
1989, p.70.
10.3. ISO 10816-1, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 1: General guidelines, 1995.
10.4. ISO 10816-6, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 6: Reciprocating machines with power
ratings above 100 kW, 1995.
10.5. VDI 2059 - Part 1, Shaft vibrations of turbosets. Principles for measurement
and evaluation, Nov 1981.
10.6. ISO 7919-1, Mechanical vibration of non-reciprocating machines
Measurement on rotating shafts and evaluation criteria, Part 1: General
guidelines, 1996.
10.7. Bently, D. E., Crude vibration amplitude measurements: Peak to peak versus
smax , Orbit, vol.15, no.3, Sept 1994, p.3.
10. VIBRATION MEASUREMENT 113

10.8. * Machinery protection systems for various types of rotating equipment,


Part 2, Bently Nevada Corporation, Application Note BNC-015, L0467-00,
June 1980.
10.9. Federn, K., Erfahrungswerte, Richtlinien und Gtemastbe fr die
Beurteilung von Maschinenschwingungen, Konstruktion, vol.10, no.8, 1958,
p.289-298.
10.10. Jackson, Ch., The Practical Vibration Primer, Gulf Publishing Company,
Houston, Texas, 1979.
10.11. ISO 13373-1, Condition monitoring and diagnostics of machines
Vibration condition monitoring Part 1: General procedures, 2002.
10.12. Khazan, A. D., Transducers and Their Elements: Design and Application,
Prentice Hall, 1994.
10.13. Bently, D. E., Proximity measurement for engine system protection and
malfunction diagnosis, Bently Nevada Corp. Publication BNC-1, from Diesel
and Gas Turbine Progress, March 1972.
10.14. Bently, D. E., Shaft motion and position Keys to planned machine
maintenance, Annual Meeting of the Technical Association of the Pulp and
Industry, Miami Beach, FL, 14-16 Jan 1974.
10.15. * Machine protection systems, Dymac Measurement and Control,
Application Note Dymac MPS-1, Dec 1977.
10.16. Harker, R. G., A new turbine supervisory instrumentation package, Bently
Nevada Corp. Publication BNC-3, Aug 1979.
10.17. * Bently Nevada Oscilloscope by Tektronix, Technical/Ordering
Information L6026, Jan 1990.
10.18. Jackson, Ch., Balance rotors by orbit analysis, Hydrocarbon Processing,
vol.50, no.1, Jan 1971, p.73-79.
10.19. * Machine Monitoring Systems, Equipment for electronic measurement
of mechanical quantities, Philips Catalogue 79/80, p.49.
10.20. * Accelerometer calibration for accurate vibration measurements, Brel
& Kjaer Application Note No. BR 0173.
10.21. Niemkiewicz, J., Standards for vibrations of machines and measurement
procedures, Encyclopedia of Vibration, Braun, S., Ewins, D. and Rao, S.S.,
eds., Academic Press, London, 2002, p.1224-1238.
10.22. * Data presentation techniques for trend analysis and malfunction
diagnosis, Bently Nevada Corporation, Application Note R7/79, July 1979.
10.23. Muszynska, A., Misalignment and shaft crack-related phase relationships
for 1X and 2X vibration components of rotor responses, Orbit, vol.10, no.2,
Sept.1989, p.4-8.
114 DYNAMICS OF MACHINERY

10.24. Tondl, A. ans Springer, H., Ein Beitrag zur Klassifizierung von
Rotorschwingungen und deren Ursachen, Schwingungen in rotierenden
Maschinen III, Irretier, H., Nordmann, R., Springer, H., eds., Vieweg,
Braunschweig, 1995, p.257-267.
10.25. Laws, B., When you use spectrum, dont use it halfway, Orbit, vol.19, no.2,
June 1998, p.23-26.
10.26. Southwick, D., Plus and minus spectrum, Orbit, vol.14, no.2, June 1993,
p.16-20.
10.27. Laws, W. C. and Muszynska, A., Periodic and continuous vibration
monitoring for preventive/predictive maintenance of rotating machinery,
Journal of Engineering for Gas Turbines and Power, vol.109, April 1987, p.159-
167.
10.28. Gasch, R., Nordmann, R. and Pftzner, H., Rotordynamik, 2nd ed., Springer,
Berlin, 2001.
10.29. Southwick, D., Using full spectrum plots, Orbit, vol.14, no.4, Dec 1993,
p.19-21 and vol.15, no.2, June 1994, p.11-15.
10.30. VDI 2059 - Part 2, Shaft vibrations of steam turbosets for power stations,
March 1983, p.6.
10.31. Bently, D. E. and Muszynska, A., Detection of rotor cracks, Proc. 15th
Texas A&M Turbomachinery Symposium, Corpus Cristi, Texas, 10-13 Nov
1986, p.129-139.
10.32. * ADRE 3, Bently Nevada Corporation, Technical/Ordering Information
L6024, Jan 1990.
10.33. Eshleman, R. L., Basic Machinery Vibration Analysis, Vibration Institute
Press, Clarendon Hills, IL, 1999.
11.
CONDITION MONITORING AND
FAULT DIAGNOSTICS

This chapter presents diagnostic techniques for determining machinery


performance and predicting mechanical problems. Methods are based on condition
monitoring, vibration measurement and analysis.

11.1 Machine deterioration

Figure 11.1 shows a typical bathtub machine deterioration time curve.

Fig. 11.1 (after [11.1])

Three periods can be distinguished: 1) the running-in period, 2) the


machine normal operation, and 3) the failure development period [11.2].
The beginning of a machine useful life is usually characterized by a
relatively high rate of failure. These wear-in failures are typically due to design
116 DYNAMICS OF MACHINERY

errors, manufacturing defects, assembly mistakes, installation problems and


commissioning errors. As the causes of these failures are found and corrected, the
frequency of failure decreases.
The machine then passes into a relatively long period of operation, during
which the frequency of failures occurring is relatively low. This period of a
machine life is called the normal wear period and usually makes up most of the
life of a machine. There should be a relatively low failure rate during the normal
wear period when operating within design specifications.
As a machine gradually reaches the end of its designed life, the frequency
of failures again increases. These failures are called wear-out failures. This
gradually increasing failure rate is primarily due to metal fatigue, wear mechanisms
between moving parts, corrosion, and obsolescence. The slope of the wear out part
of the bathtub is machine-dependent.

11.2 Machine condition monitoring

The objectives of condition monitoring of machinery include: a) control


of the machinery, especially for high power and dangerous machines; b)
optimizing the availability of machines by avoiding unexpected shutdowns,
especially for critical machines in a continuous production process, and c)
implementation of condition-based maintenance, for which the operations are
planned according to various constraints (cost, production, failure condition, etc.).

11.2.1 General considerations

The ultimate goal of machine condition monitoring is to get useful


information on the condition of equipment to the people who need it in a timely
manner. The personnel include operators, maintenance engineers and technicians,
managers, vendors, and suppliers. These groups will need different information at
different times. The task of the person or group in charge of condition monitoring
is to ensure that useful data is collected, that data is changed into information in a
form required by and useful to others, and that this information is provided to the
people who need it and when they need it. Useful references on this subject are the
books [11.2] to [11.16].
The focus of this chapter is on vibration-based data, but there are several
different types of data that can be useful in assessing the machine condition. These
include lubrication oil/grease analysis, wear particle monitoring and analysis,
noise, temperature, force, output (machine performance), product quality, odor, and
visual inspections.
11. CONDITION MONITORING 117

11.2.2 Maintenance strategies

Maintenance strategies can be divided into three main types: a) run-to-


failure, b) preventive, and c) predictive maintenance. Each of these different
strategies has distinct advantages and disadvantages. Specific situations within any
facility require the application of a different strategy. Therefore, no one strategy
should be considered as always superior or inferior to another.

11.2.2.1 Run-to-failure maintenance


Run-to-failure, or breakdown maintenance, is a strategy where repair
work or replacement is only performed when machinery has failed. In general, run-
to-failure maintenance is appropriate when the following situations exist: a) the
equipment is redundant, b) low cost spares are available, c) the process is
interruptible or there is stockpiled product, d) all known failure modes are safe, e)
there is a long mean time to failure (MTTF) or a long mean time between failure
(MTBF), f) there is a low cost associated with secondary damage, and g) quick
repair or replacement is possible [11.17].

Fig. 11.2 (from [11.17])

Figure 11.2 shows an illustration of the relationship between the machine


time in service, the load (or duty) placed on the machine, and the estimated
remaining capacity of the machine. Whenever the estimated capacity curve
intersects with (or drops below) the load curve, a failure will occur. At these times,
repair work must be carried out. If the situation that exists fits within the seven
rules outlined above, all related costs (repair work and downtime) will be
minimized when using run-to-failure maintenance.

11.2.2.2 Preventive maintenance


When specific maintenance tasks are performed at set time intervals (or
duty cycles) in order to maintain a significant margin between machine capacity
118 DYNAMICS OF MACHINERY

and actual duty, the type of maintenance is called preventive (or scheduled)
maintenance.
Preventive maintenance is most effective under the following
circumstances: a) data describing the statistical failure rate for the machinery is
available, b) the failure distribution is narrow, meaning that the MTBF is
accurately predictable, c) maintenance restores close to full integrity of the
machine, d) a single, known failure mode dominates, e) there is low cost associated
with regular overhaul/replacement of the equipment, f) unexpected interruptions to
production are expensive, g) low cost spares are available, and h) costly secondary
damage from failure is likely to occur [11.17].

Fig. 11.3 (from [11.17])

Figure 11.3 shows an illustration of the relationship between the machine


time in service, the load (or duty) placed on the machine, and the estimated
remaining capacity of the machine when preventive maintenance is being
practiced. Maintenance activities are scheduled at regular intervals in order to
restore machine capacity before a failure occurs. In this way, there is always a
margin between the estimated capacity and the actual load on the machine. If this
margin is always present, there should theoretically never be an unexpected failure,
which is the ultimate goal of the preventive maintenance.

11.2.2.3 Predictive maintenance


Predictive (on-condition) maintenance requires that some means of
assessing the actual condition of the machine is used in order to optimally schedule
maintenance, in order to achieve maximum production, and still avoid unexpected
catastrophic failures.
Condition based maintenance should be employed when the following
conditions apply: a) the machine is expensive or critical, b) a long lead-time is
necessary for replacement parts (no spares are readily available), c) the process is
uninterruptible, d) equipment overhaul is expensive and requires highly trained
11. CONDITION MONITORING 119

personnel, e) failures may be dangerous, f) secondary damage may be costly, and


g) failures are not indicated by degeneration of normal operating response [11.17].
Figure 11.4 shows an illustration of the relationship between the machine
time in service, the load (or duty) placed on the machine, and the estimated
remaining capacity of the machine when predictive maintenance is being practiced.
Note that the margin between duty and capacity is allowed to become quite small,
but the two lines never touch. This results in a longer time between maintenance
activities than for preventive maintenance. Maintenance tasks are scheduled just
before a failure is expected to occur. This requires the existence of a set of accurate
measures that can be used to assess the machine integrity.

Fig. 11.4 (from [11.17])

There are instances where a given machine will require different


maintenance strategies during its operational life, e.g. scheduling the maximum
time between overhauls during the early stages of the machine life, and increased
frequency of monitoring as the age of the machine increases, looking only for
unexpected failures.

11.2.3 Factors influencing maintenance strategies

While there are some general guidelines for choosing the most
appropriate maintenance strategy, each case must be evaluated individually.
Principal considerations will always be defined in economic terms. Sometimes, a
specific company policy, such as safety, will outweigh all other considerations.
The following eight factors should be taken into account when deciding
the best maintenance strategy for a given machine: a) classification (size, type) of
the machine, b) critical nature of the machine relative to production, c) cost of
replacement of the entire machine, d) lead-time for the replacement of the entire
machine, e) manufacturers recommendations, f) failure data (history), MTTF,
120 DYNAMICS OF MACHINERY

MTBF, failure modes, g) redundancy, and h) safety (plant personnel, community,


environment) [11.17].

11.3 Diagnosis process

The main steps of a diagnostic process used in condition monitoring are


presented in Fig. 11.5.

Fig. 11.5 (from [11.2])

1. Data measurement and validation


The information is taken from sensors and measurement systems that
have to be reliable. Bad measurements generate a wrong diagnosis so that
various techniques have been developed to detect invalid measurements.

2. Operating condition assessment


This step allows the definition of a reference signature (or base-line)
which is able to characterize the state of the machine (healthy or faulty), but can
also be related to various kinds of faults. This is an important preliminary step to
the diagnostic process. It makes use of the machine characteristics, the type of
11. CONDITION MONITORING 121

physical measurements and the effects of the faults, i.e. the symptoms. It obviously
requires the knowledge of faults that may occur in the machine, and their
criticality, to be able to define the most suitable signature.

3. Detection
Detection involves data gathering, comparison to standards and
recommendations (see Chapter 12), comparison to limits set in-plant for specific
equipment, and trending over time. The signature characterizing the healthy state is
compared to the one extracted from the real measurement. The fault is not defined.

4. Diagnosis
Diagnosis involves recognizing the types of fault developing and
determining the gravity of given faults once detected and diagnosed. Sometimes
it is referred to as fault isolation. When a signature is related to a specific fault,
steps 3 and 4 may be imbedded in one-step detection/diagnosis, which happens
often in vibration condition monitoring.

5. Decision
In this step, known as prognosis, the operator has to decide whether to
stop the machine for maintenance and repair, or to continue operating. Prognosis
involves estimating (forecasting) the expected time to failure, trending the
condition of the equipment being monitored, and planning the appropriate
maintenance timing. It may include recommendations for altering the operating
conditions, altering the monitoring strategy (frequency, type), or redesigning the
process or equipment. Sometimes it includes the root-cause failure analysis, and
involves some research-type laboratory and/or in situ investigations.

11.4 Fault diagnostics

In the following, a brief description is given for the major categories of


problems that can cause machine failures and how they may be recognized. An
attempt is also made to assess the effectiveness of the various methods of data
presentation and analysis in correlation with operational process data.

11.4.1 Unbalance

Unbalance (also referred to as imbalance) exists when the center of mass


of a rotating component is not coincident with the center of rotation. It is
practically impossible to fabricate a component that is perfectly balanced, and even
after balancing (see Chapter 13) some residual unbalance exists in rotors,
122 DYNAMICS OF MACHINERY

flywheels, fans, gears, etc. The causes of unbalance include excess mass on one
side of a rotor (lost blade, eroded or damaged parts), low tolerances during
fabrication (casting, machining, assembly), variation within materials (voids,
porosity, inclusions), non-symmetry of design, aerodynamic forces, and
temperature changes.
Unbalance results in a periodic vibration signal with the same amplitude at
each shaft rotation. The characteristic diagnostic symptom is a strong radial
vibration at the fundamental frequency, 1X (1 x rotational speed). If the rotor is
overhung, there will be also a strong axial vibration at 1X. The half spectrum has
the higher peak at 1X (Fig. 11.6, a), the orbit is generally elliptical (Fig. 11.6, b)
and the timebase waveforms have one key phase mark per shaft revolution (Fig.
11.6, c).

Fig. 11.6 (from [11.18])

The high amplitude may be in both horizontal and vertical directions or it


may be larger in one than the other. The latter is caused by asymmetrical radial
forces such as might be produced by a pressure dam bearing, torque reaction in
gears or orthotropic bearings (different horizontal and vertical support stiffnesses).
In each case, the higher force tends to suppress the coincident vibration.
A special form of unbalance is caused by thermal distortion of the rotor. It
is produced by thermally unstable shaft forgings which take a bow when they are
heated to an elevated temperature. This bow is a function of temperature and it will
11. CONDITION MONITORING 123

not straighten out with time unless the rotor is allowed to cool off. Turbines are
more affected than compressors and motors. To measure it, turbine rotor forgings
are rotated slowly in an oven while temperature is increased and decreased several
times, runout being recorded at several places along the rotor. Maximum allowable
runout, at 500 C above operating temperature, is usually 8 m m of bearing span.
Another form of unbalance is caused by bowed rotors, especially heavy
rotors that have been allowed to sit idle for a long time. Such rotors are difficult to
straighten, so they need to be balanced by adding counterweights. To avoid this
condition the rotor should be turned when the machine is not in use.

Fig. 11.7 (from [11.19])

Figure 11.7 shows the full spectrum cascade plot measured during the
startup of a machine with an unbalanced rotor. The 1X and -1X frequency
components have unequal peaks around 3000 rpm, denoting elliptical precession
orbits. The two neighboring peaks denote a split critical excited by unbalance.

11.4.2 Misalignment and radial preload

Following unbalance, the misalignment of machine train rotors is the


second most common malfunction of rotating machinery. One of the main effects
of misalignment between rotors in the machine train is the generation of rotor
preload in a radial direction. The misalignment causes a constant radial force which
pushes the rotor to the side. Gravitational preload on horizontal rotors, thermal
124 DYNAMICS OF MACHINERY

expansion preload, offset, or cocked bearing-related preload, and gear mesh forces
also belong to this category.
Misalignment between coupled machines can be caused by thermal
expansion of the casing support structure, by settling or thermal distortion of
foundation or baseplate, or by piping forces which deflect the casing and its
support. It can be produced by a strong radial component of the fluid flow in fluid-
handling machines, especially evident in single volute pumps, or in turbines during
partial steam admission on the first stage nozzles.
Due to the radial force, the rotor is displaced from the original position and
moved to higher eccentricity ranges inside the bearings and seals. It may also
become bowed, and rotate in a bow configuration. At these conditions the
nonlinear effects of the system become active. Due to nonlinearity, the unbalance
forced response of the rotor will contain not only the synchronous component 1X,
but also its higher harmonics 2X, 3X, etc.
There are two components of a coupling misalignment (Fig. 11.8, a): a)
parallel (offset), and b) angular (face). Parallel misalignment occurs when shaft
centerlines are parallel but offset from one another in a radial direction (Fig. 11.8,
b). Angular misalignment occurs when the shaft centerlines meet at an angle (Fig.
11.8, c). The intersection may be at the driver or driven end, between the coupled
units or behind one of the coupled units. A coupling misalignment is shown in Fig.
11.8, d.

Fig. 11.8 (after [11.20])

The typical effect of misalignment is a vibration with a predominant 2X


frequency component, as shown in Fig. 11.9. Shaft alignment techniques are
presented in Annex 11.1.
11. CONDITION MONITORING 125

Strong 1X vibrations with harmonics (usually up to the third, but


sometimes up to the sixth) in the frequency spectrum are the usual diagnostic
signatures. The harmonics allow misalignment to be distinguished from unbalance.
High horizontal relative to radial vibration amplitude ratios (greater than 3:1) may
also indicate misalignment.

Fig. 11.9 (from [11.18])

The rotor precession orbits show some distortion from the effect of a
misalignment load, as shown in Fig. 11.9, c. As the preload increases, the orbit will
progressively shift from an ellipse to a banana and finally, in extreme cases,
possibly to a figure eight (see Fig. 10.33).
In severe cases of misalignment it is not uncommon to have the low or
unloaded bearing become unstable due to the journal orbit location in the upper
half of the fluid film bearing (Fig. 11.10).
The analysis of the shaft centerline position can be used to diagnose
excessive preloads. A combination of shaft position and orbit representation gives a
clear indication of the position of the shaft in each bearing. In Fig. 11.10, preloads
on the shaft are forcing the shaft down in one bearing and up in the other. Note the
elliptical and banana nature of the orbits as a result of the preload and the bearing
constraint. The key phase dots on the orbits indicate that, although the bearings are
126 DYNAMICS OF MACHINERY

preloaded in opposite directions, the ends of the shaft vibrate in phase with each
other.

Fig. 11.10 (from [11.21])

Misalignment is the deviation of relative shaft position from a collinear


axis of rotation, measured at the points of power transmission when equipment is
running at normal operating conditions. For instance, if the points of power
transmission are 0.5 m apart and the maximum shaft centerline to projected shaft
centerline offset is 0.5 mm , the deviation is one mm per m of power transmission
distance. At 3000 rpm this deviation is acceptable. At 20,000 rpm the alignment
deviation is unacceptable.

Fig. 11.11 (from [11.20])


11. CONDITION MONITORING 127

A general guideline for alignment tolerances is shown in Fig. 11.11.


Acceptable amounts of misalignment must be tailored to suit each individual drive
train application. Gear-type couplings and universal joint drives must have small
amounts of misalignment for proper lubrication to occur. Staying within the
acceptable misalignment band usually fills this requirement. Diaphragm couplings,
on the other hand, should be aligned within the excellent range.

11.4.3 Fluid induced instabilities

Instability, or at least some tendency toward instability, is a relatively


common problem on high speed machinery equipped with fluid film bearings. As a
self excited phenomenon, instability causes the shaft to precess at a submultiple of
running speed and it is easy to be recognized as illustrated in Fig. 11.12.

Fig. 11.12 (from [11.18])

In a spectrum presentation (Fig. 11.12, a), instability shows up as a spectral


component between approximately 40% and 60% running speed. In its early stages
the subharmonic component generally fluctuates irregularly in amplitude. As
instability progresses, the subharmonic spectral component will increase in
amplitude and will remain at the higher amplitude for longer periods of time until it
finally dominates the spectrum. When the latter occurs, the amplitude fluctuations
generally stop and components at twice and higher multiples of the subharmonic
may appear in the spectrum.
128 DYNAMICS OF MACHINERY

In a timebase presentation (Fig. 11.12, b), instability causes the running


speed pattern to snake and bounce. The irregular bounce will likewise occur in an
orbital presentation (Fig. 11.12, c) but the key indicator in the final stages of
instability is two timing marks around the circumference of the orbit, indicating
that the shaft rotates twice in the time necessary to complete the orbit.
Instability is caused by a variety of factors. Oil whirl, where the shaft rides
a pressure wave in the oil film circulating at approximately one-half shaft speed is
perhaps the most common example. As a slight variation, the presence of oil whirl
close to a critical speed may cause the combined effect to latch in at the critical.
Other sources of instability can be hysteretic or frictional in nature, but all those
mentioned share a common cause, a force component perpendicular to the rotor
normal stabilizing force which, when resolved, produces a tangential whirling force
in the same direction as the shaft rotation.
Corrective measures for instability can range from minor changes in
bearing design such as decreased clearances, decreased bearing area in the lower
half to increase load or the addition of stiffness by substituting a pressure dam or
lobed bearing design for a plain cylindrical bearing. Next, one generally goes to a
much stiffer bearing design such as a tilting pad type with changes to the rotor
itself, such as decreasing the bearing span and/or adding diameter resorted to as a
final measure should all else fail.
In general, operating changes such as varying oil temperature are not
successful in eliminating instability although deliberate misalignment has been
used in the past to temporarily stabilize a bearing and permit continued operation.
A common type of malfunction in machines with fluid film bearings is
whirl and whip. Whirl and whip are rotor instabilities (self-excited vibrations)
generated by bearing, seal or main flow fluid dynamic forces. The malfunction is
characterized by rotor forward subsynchronous precession, often at destructive
levels, especially for whip conditions. Analysis of spectrum cascade plots and shaft
orbital motion can be used to diagnose whirl and whip.
Figure 11.13 shows a half spectrum cascade plot with orbits of an
unbalanced rotor supported in oil bearings. At a threshold of stability, below the
first resonance speed, the rotor undergoes whirl, as exhibited by the vibration
component with frequency just below X (orbit a).
When the machine speed increases and passes through the first resonance
(orbit b), the amplitude of 1X vibration, caused by unbalance increases, resulting in
higher oil bearing stiffness. It causes the suppression of the whirl. The bearing
stiffness increases significantly for higher journal eccentricities. This characteristic
is widely used for correction of oil whirl and whip malfunctions. A friendly
radial preload (provided for instance by a misalignment), holding the journal in an
eccentric position corrects effectively the oil whirl and whip in machines supported
by hydrodynamic cylindrical bearings.
11. CONDITION MONITORING 129

For higher running speeds, above the first resonance, when the shaft is
lightly loaded and 1X amplitude is reduced, the whirl appears again (Fig. 11.13).
It continues with the frequency just below X and then asymptotically
approaches to the rotor first natural frequency (orbit c). Oil whirl is replaced by oil
whip. The latter is much more violent and dangerous for the machine integrity
because the shaft vibrates at its resonant conditions. Therefore, relatively large
cyclical rotor bending stresses can be incurred, introducing a significant risk of
high cycle fatigue failure if the steady state tensile stresses in the rotor are high
enough.

Fig. 11.13 (from [11.22])

Vibration data taken during machine startup are conveniently displayed as


full spectrum cascade plots. Figure 11.14 shows such a plot of a machine with a
threshold of stability of approximately 2300 rpm. For speeds less than
approximately 4500 rpm, the vibration is composed almost entirely of high
amplitude forward vibration components. The absence of backward components
130 DYNAMICS OF MACHINERY

indicates that the shape of the precession orbits should be circular and the
precession is forward.
The instability vibration frequency is approximately 0.45X for the whirl
instability, and it begins to diverge when the system starts the transition into whip
instability as the rotor speed approaches 5000 rpm. For machine speeds above 4500
rpm, small backward vibration components exist and the orbit for the whip
instability is slightly elliptical (confirmed by measurements).

Fig. 11.14 (from [11.23])

Full cascade plots should be used together with orbit/timebase plots


[11.23]. The latter are not shown here for conciseness.

11.4.4 Rotor-to-stator rubbing

Rubbing between the rotor and a stationary part of the machine is a serious
malfunction that may lead to a catastrophic failure. Rubbing involves several
physical phenomena, such as friction, stiffening/coupling effect, impacting, and
may affect solid/fluid/thermal balance in the machine system. Rubbing always
occurs as a secondary effect of a primary malfunction, such as unbalance,
misalignment, or fluid-induced self-excited vibrations, which result in high lateral
vibration amplitudes and/or changes in the shaft centerline position.
There are two extreme cases of rotor radial rubs: a) a full annular rub,
when the rotor maintains contact with an obstacle (e.g., a seal) during the complete
11. CONDITION MONITORING 131

cycle ( 360 0 ) of its precession motion, and b) a partial rub, when the contact
occurs occasionally during a fraction of the period of precession.

Fig. 11.15 (from [11.22])

In the case of full annular rub, occurring mainly in seals, high friction
forces cause the change of the precession direction from forward to continuous
backward whirl (known as dry whirl). The waterfall plot of vertical vibrations of
a rotor rubbing inside the seal (Fig. 11.15) shows that in the lower speed range, the
rotor bounces inside the seal, producing multiple higher harmonics of 1X, while at
higher speed a full annular rub occurs.
In the case of short-lasting rotor/stator contact, the system becomes piece-
wise continuous with variable stiffness. The rub may be caused by a seal or other
non-rotating part acting as a bearing during part of the shaft revolution.
The periodic contact with an obstacle (Fig. 11.16, a), creating the effect of
a third bearing, produces a periodic variation of the rotor stiffness which
determines the self-excitation of the synchronous response and increases the
average spring constant to a higher value (Fig. 11.16, b). This tends to raise the
rotor critical speed (Fig. 11.16, c).
132 DYNAMICS OF MACHINERY

Systems with periodic stiffness variations exhibit parametric vibrations


described by Mathieu-type equations of motion. The solutions correspond to
submultiples of the running speed frequency. If the rotor resonance is less than 1/4,
1/3, 1/2 etc. of operating speed, has some unbalance and is lightly damped, the
resonance of the rotor system will be increased by the rub to exactly coincide with
the nearest higher fraction of running speed. The rotor will lock on this exact
submultiple.

a b

c d
Fig. 11.16 (from [11.24])

The total shaft motion orbit (Fig. 11.16, d) has two fixed once-per-turn
timer marks indicating that the rotor requires two full turns to complete one orbit.
The orbit of the filtered 1X and (1/2)X components show inverse precession
directions due to the shaft kicking back as it rubs. Generally, multiples of (1/2)X
are also produced by the nonlinearity of the normal/tight rub [11.25].
The partial rotor/stator rub, or rub in oversized or poorly lubricated
bearings, causes steady subharmonic vibrations of the frequency equal exactly to
half of the rotational speed. The range of the possible subharmonic vibrations
varies, however, with the rotational speed. When the rotor operational speed is
higher that 3 times its first natural frequency, the resulting steady subharmonic
vibrations can have the range (1 3) X (light rub) or (1 2) X (heavy rub). This
condition can be generalized to any value of the rotational speed. If it exceeds the
value i times the rotor first natural frequency, then the rotor response will consist
of the synchronous component 1X and one subsynchronous component with the
lowest frequency equal to (1/i)X, or (1/(i-1))X, or (1/3)X, or (1/2)X with
increasing rub force [11.26].
11. CONDITION MONITORING 133

b
Fig. 11.17 (from [11.26])
134 DYNAMICS OF MACHINERY

The half spectrum cascade plot in Fig. 11.17, a shows the subharmonic
vibrations in the case of light rub. An increase of the rotational speed causes the
change of the subharmonic order from higher to lower range. Figure 11.17, b
presents the case of a higher rub force. The subharmonic vibrations of the order 1/2
are steadily maintained while increasing the rub force and rotational speed. The
samples of the rotor precession orbits were taken at rotational speeds = 227,
404, 595, and 790 rad/s. The steady rotor response consists of two main harmonics:
a synchronous component 1X due to unbalance, and a subsynchronous component
(1/2)X, (1/3)X, (1/4)X or (1/5)X, only one at a time. Minor higher harmonics are
present in the frequency spectrum. The synchronous orbit is always reduced to a
straight line inclined to the left, i.e., the vertical and horizontal subsynchronous
components are 1800 out of phase [11.26].
At certain rotational speeds, the thermal effect of rubbing causes an ever
changing thermal bow-related unbalance of the shaft.

Fig. 11.18 (from [11.27])

An interesting feature of the full annular synchronous rub is illustrated by


Fig. 11.18. When the rotor speed is increased, the unbalance creates enough force
to cause the rotor to contact the stator before the resonance peak. The rotor gets
stuck on top of its critical. The phase lag is usually about 80 to 100 degrees. As
speed increases, the rub increases, so the system dynamic stiffness increases. This
raises the critical speed, thus establishing this lockup situation. If the rotational
speed is well above the critical, an impact on the shaft can remove the continuous
rub condition, and the machine operates at much higher efficiency.
11. CONDITION MONITORING 135

11.4.5 Mechanical looseness

Looseness related dynamic phenomena can be relatively easy to identify


and eventually corrected, as they cause very characteristic modifications of rotor
normal operational responses. The features of loose stationary parts, rotating parts
and oversize, poorly lubricated bearings are presented in the following.

11.4.5.1 Loose stationary parts

A common type of vibrations is produced in loose non-rotating machine


parts. Typical examples of these may be a bearing shell with excessive clearance
with respect to the bearing housing. Other examples may be a loose bearing
housing, loose pedestal support, loose grout or a frame set on the earth without tie
downs.
Symptoms of looseness of the casing on its supports can be observed
listening to the machine with a listening rod and feeling with the fingertips for
differential vibration at mating surfaces. It is always a good idea to check all bolts
in the support structure for tightness, including casing hold down bolts and
soleplate bolting. A thorough checkout should be made for any gaps between feet
and other mounting surfaces, using a feeler gage, giving special attention to
clearances under casing feet, cracks in the foundation, and clearances in guide
keys.
Based on experience, this type of problem produces a spectrum with a high
amplitude peak at running frequency, followed by a string of vibration components
at multiples and submultiples of running frequency (Fig. 11.19).

Fig. 11.19 (from [11.18])

The unbalance force carried by the rotor may occasionally exceed the
gravity force and/or other lateral forces applied to the rotor and pedestal. This
136 DYNAMICS OF MACHINERY

causes a periodic lifting of the pedestal, resulting in system stiffness softening, its
cyclic variability and impacting. As a result, the rotor may exhibit changes in the
synchronous response, and an appearance of fractional subsynchronous vibrations
((1/2)X, (1/3)X,) in some speed ranges. Most common is the occurrence of the
(1/2)X vibration component, often measured on rotating equipment.

11.4.5.2 Excessive rotor/bearing clearance

Specific dynamic phenomena are caused by increased looseness in


bearings (often referred to as dead band), usually due to poor lubrication.
Excessive clearances between journals and plain bearing bushes, as well as
between rolling element bearings and housing, produce periodic variations of the
stiffness of rotor/bearing system (Fig. 11.20), thus providing conditions for
parametric unbalance-related excitation which can lead to rotor instability.

Fig. 11.20 (from [11.24])

These phenomena are similar to those occurring during the rotor-to-stator


rubbing, namely variable stiffness, impacting and friction. The similarity is,
however, of the mirror image type. The rubbing system is described as normal-
tight, while the system with increased clearances is described as normal-loose
[11.25]. As shown in section 11.4.4, a rubbing rotor becomes periodically stiffer,
which leads to an increase of the average stiffness. In the rotor/bearing system with
excessive clearances, the average stiffness decreases. This tends to lower the rotor
critical speed. If the normal rotor resonance is greater than 1/2 the normal operating
speed and the system is lightly damped, the resonance of the rotor will be lowered
by the effective decreased stiffness to coincide with the nearest lower fraction of
running speed. The rotor will lock into this exact submultiple [11.25].
The diagnosis of excessive clearance, and distinguishing it from the
rubbing, should be based on the rotor centerline position and 1X data, frequency
spectrum and orbit analysis. While exhibiting similar spectra, the journal/bearing
contact is usually maintained during a longer fraction of the vibration period than
the rotor/stator rubbing contact, thus the orbits are substantially different from the
rub case. While maintaining the contact, the journal slides on the bearing surface,
and a part of the orbit follows the bearing clearance circle. The journal remains
close to the bearing surface even when the contact is broken. This is different from
rubbing when more impacting and unsteady transient motion occurs.
11. CONDITION MONITORING 137

Fig. 11.21 (after [11.21])

Figure 11.21 shows the half spectrum cascade recorded during start-up of a
journal rotating in a bearing with relatively large radial clearance in a brass
bushing. Subsynchronous vibrations of (1/2)X and (1/3)X, as well as self-excited
vibrations are present.

11.4.5.3 Loose rotating parts


Looseness may occur at discs or thrust collars mounted on rotating shafts
or at bearings untightened in bearing pedestals. A loose disc will still rotate, but at
a different speed than that of the rotating shaft. A loose bearing may start rotating,
dragged into rotation by the shaft. Their response is a function of clearances, the
friction conditions between the shaft and the loose part, as well as the tangential
external force applied to the loose part.
Depending on a particular machine, the drag force can drive the loose part
at a higher frequency than the rotational frequency (e.g.: a loose turbine disc) or
138 DYNAMICS OF MACHINERY

slow down the loose part. At steady-state conditions, the friction and fluid drag
may balance each other, and the loose part rotational frequency, l , becomes
constant. If it does not differ very much from the rotational speed, , the resulting
vibrations exhibit the characteristic of beat (Fig. 11.22).
Most often, however, the looseness of a rotating part leads to transient
conditions. The loose part related vibrations have most often a subsynchronous
frequency tending to the natural frequency of the rotor. These vibrations look
somewhat similar to fluid whirl/whip vibrations, and may sometimes be confused
with the latter.

Fig. 11.22 (from [11.21])

The time signal from a bearing that is loose on a shaft will also be
truncated (clipped). The extent and shape of the truncation depend on the physical
characteristics (stiffness, mass and damping) of the transmission path between the
rotor and stator. Spectral analysis of a truncated waveform yields a number of
discrete sum and difference frequencies.

11.4.6 Cracked shafts

There are two fundamental symptoms of a cracked shaft: a) changes in the


synchronous 1X response vector (amplitude and phase) and slow roll vector, and b)
the occurrence of the vibration component of twice rotational speed 2X,
occasionally at the operating speed, but especially on startup and shutdown.
The first symptom is caused by the shaft bowing (elastic unbalance
effect) which interferes with the original mass unbalance. The second symptom is
associated with the asymmetry of the shaft. The 2X component is due to a
combination of a transverse crack and a constant radial force. The 2X component is
especially dominant when the rotational speed is in the region of half of any rotor
system natural frequency.
11. CONDITION MONITORING 139

The changes in the synchronous 1X amplitude and phase, measured by


proximity probes, can be monitored under normal operating conditions to provide
alarming and early warning of a shaft crack. The polar plot with acceptance regions
(Fig. 10.48) is an excellent format for documenting these shifts. Deviation of the
1X vibration vector tip from an acceptance region may be a vital warning of a shaft
crack.
More effective is the shaft crack detection using transient data. Figure
10.43 shows a half spectrum cascade plot that documents 1X, 2X and other
vibration components from slow roll to maximum available speed.
A very useful diagnostic tool is the full spectrum cascade plot (Fig. 11.23).
It shows the resonance produced by a 2X excitation force when the operating speed
is near half the first critical speed.

Fig. 11.23 (from [11.28])

The usual format of a full spectrum cascade plot is shown in Fig. 11.24. It
clearly shows a peak in the 2X amplitude at about 1390 rpm. The first bending
resonance of the machine is approximately 2700 rpm. Note that there is a peak in
the 3X amplitude at about 900 rpm, and a peak in the 4X amplitude at about 700
rpm.
140 DYNAMICS OF MACHINERY

When operating at 1390 rpm, i.e. near half the first critical speed, the
journal orbit (Fig. 11.25) has an internal loop which is characteristic for signals
containing two vibration components with the same direction of precession [11.29].
A more detailed study (not presented here) implies analysis of orbit/timebase plots
of the filtered 1X and 2X components. For the examined case it was found [11.19]
that the 1X component is forward and slightly elliptical. The 2X component is
forward, more elliptical, and larger than the 1X component.

Fig. 11.24 (from [11.19])

Fig. 11.25 (from [11.19])

It is important to correlate the changes in the 1X vector with changes of the


system and process parameters as well as to trend the 2X vector, to determine
whether vector changes are caused by a rotor crack or other factors such as load,
field current, steam conditions, or other operating parameters.
11. CONDITION MONITORING 141

11.5 Problems of specific machines

Selection of a machine protection system depends on the machine design


and construction, support structure, service and type of operation, and response to
probable malfunctions. Specific problems related to different machine types are
presented in the following [11.35].

11.5.1 Centrifugal equipment

Centrifugal equipment generates vibration components at and near


rotational frequency. Additional prominent components generally are found at the
vane-passing frequency or frequencies (number of impeller vanes multiplied by
shaft speed), followed by a series of harmonics. Their amplitude may be related to
cavitation in pumps or surge in compressors and fans. Monitoring these vibration
characteristics may warn about incipient cavitation or surge.

11.5.1.1 Centrifugal pumps

Centrifugal pumps generally have flexible cantilevered bearing housings


(Fig. 11.26). With this type of construction, a large portion of the dynamic force
developed by the rotor is transmitted across the bearings with minimum relative
motion and is dissipated as structural vibration. Using vibration pickups, attached
to the bearing housings in the plane of least stiffness, usually provide the best
response and indication of mechanical condition.

Fig. 11.26 (from [11.30])


142 DYNAMICS OF MACHINERY

For optimum performance, a casing-monitoring system similar to that


shown in Fig. 11.27 is recommended. The signal obtained from the casing sensor is
divided, within the monitor, into bands related to specific mechanical components.
The first band, encompassing the frequencies around the running speed frequency,
is measured in terms of velocity and fitted with a low-pass filter to eliminate
interference from the impeller-vane passing frequency. If cavitation is likely, a
bandpass filter can be used to enclose the vane-passing frequency and one or two
of its multiples into a second monitored band. In centrifugal pumps with rolling
element bearings, high frequency resonant transducers are used to assess pulses of
energy above some adjustable threshold to provide early warning of an impending
failure.

Fig. 11.27 (from [11.31])

In high-head applications, where an appreciable amount of thrust force


may be developed if internal clearances are lost, it is advisable to include either an
axial-thrust position monitor or a thrust-bearing temperature indicator, plus an
alarm to warn of impending problems.

11.5.1.2 Centrifugal compressors

Modern centrifugal compressors, operating on hydrodynamic bearings,


generally have a relatively large casing-to-rotor weight ratio and a stiff support
structure (Fig. 11.28).
Most of the low-frequency energy developed by the rotor is dissipated by
relative motion between journal and bearing, within the bearing clearance. A
noncontact relative-motion displacement monitoring system, like that shown in
Fig. 11.29 (installed on a compressor of older design), has the fastest and most
easily recognizable response at small changes in mechanical condition.
11. CONDITION MONITORING 143

Fig. 11.28 (from [11.32])

Fig. 11.29 (from [11.31])

At higher frequencies, vane-passing and above, there may be significant


excitation generated by aerodynamic turbulence or an impeller resonance. This
excitation is relatively easy to detect as acceleration, but its displacement most
likely will be below the minimum detectable amplitude of a typical industrial
144 DYNAMICS OF MACHINERY

displacement-monitoring system. Fortunately, this type of problem occurs so


infrequently that the additional instrumentation required for protection is not
usually warranted on a permanent basis.
Thrust position monitoring capability should be included as a part of any
centrifugal compressor monitoring system. A typical position-monitoring system
consists of an axial-displacement sensor and an appropriate monitor. Two sensors
are recommended in high-head or critical applications. Thrust-temperature
monitoring is mandatory on high-differential-head compressors, where the failure
of a balance drum seal can overload the bearing to failure.

11.5.1.3 Centrifugal fans

Centrifugal fans and blowers used for forced- or induced-draft and


primary-air service generally have large diameter rotors operating from 500 to 900
rpm in pillow-block bearings, supported on structural steel or concrete foundations
(Fig. 11.30).

Fig. 11.30 (from [11.33])

Fig. 11.31 (from [11.31])


11. CONDITION MONITORING 145

As a rule, the major problem with fans is unbalance caused by: a) uneven
build-up or loss of deposited material, and b) misalignment. Both are characterized
by changes in vibration at or near the rotational frequency, which can be monitored
effectively with either a shaft-displacement or a casing system.
Selection of the monitoring system is dictated by the type of construction.
If the bearings are supported on stiff, reinforced-concrete pedestals, most of the
dynamic force developed by the rotor will be dissipated as relative motion within
the bearing clearance. A shaft-monitoring system is best suited for this construction
(Fig. 11.31). If bearings are supported on structural steel, the dynamic force
probably is dissipated as structural vibration, and a casing seismic monitoring
system, using sensors attached to the bearing housings, gives best results. For
optimum results, characteristics not specifically related to mechanical condition
should be eliminated by filtering the fan casing vibration signal to a bandpass
extending from approximately 50% of the running speed to three or four times it.

11.5.2 Bladed machines

Bladed machinery, such as axial compressors and steam and gas turbines,
usually produces more complex vibration characteristics, particularly in the higher
frequencies, than the centrifugal equipment discussed in section 11.5.1. Spectral
components at blade-passing frequencies (the number of blades multiplied by shaft
speed), as well as at their multiples and the sum and difference combinations,
usually are identifiable.
Blade characteristics can be observed in the vibration signatures obtained
from sensors mounted on bearing caps. But high frequencies are transmitted into
the casing by pressure pulses close to the point of origin, rather than across a
compliant oil film. Thus, blade frequencies are much stronger and easier to
recognize from accelerometers located at the middle of the casing.

11.5.2.1 Axial compressors

Axial compressors have cylindrical or conical rotors carrying successive


rows of moving blades with shaft extensions at both ends (Fig. 11.32). Many
experts consider that a casing monitoring system offers acceptable protection
against both low frequency problems and high frequency blade related problems. A
more conservative approach combines acceleration monitoring for the blade-
passing frequencies with a conventional shaft-displacement system. If blade
problems are anticipated, the frequency spectrum is monitored in three bands: a)
the low frequency band around the running speed, indicating unbalance and
misalignment, b) a band covering the blade fundamental resonance frequencies,
and c) a band incorporating the blade-passing frequencies and their harmonics (up
to the third or fourth multiple of the highest blade passing frequency).
146 DYNAMICS OF MACHINERY

A continuous thrust-monitoring system using a single axial-position sensor


generally provides sufficient protection.

Fig. 11.32 (from [11.34])

Usually axial compressors are driven by gas turbines so that the monitoring
system must be designed having in view the characteristics of both machines.

11.5.2.2 Steam turbines

Steam turbines designed for driving process equipment in petrochemical


plants, or boiler-feed pumps vary significantly in dynamic response from turbines
used for producing electric power. The former usually operate between 5000 and
12,000 rpm and deliver from 6000 to 30,000 hp, with inlet steam pressures up to
about 120 bar. Utility turbines operate much slower, at synchronous and half-
synchronous speed, generally are much larger in size, and may use steam at
pressures above 250 bar.
A relative-motion shaft-displacement system, like that shown in Fig. 11.33,
serves best on large process-drive and boiler-feed pump turbines, units with
moderate to high casing-to-rotor weight ratios and relatively stiff support
structures. It provides excellent data at the low frequencies near the rotational
speed, and where shaft instability might present problems. Further, it is the only
way to monitor shaft radial position.
On critical high-speed turbines, the monitoring system includes backup
accelerometers mounted at each bearing that monitor absolute shaft motion. They
are useful to avoid problems caused by inadvertent location of shaft sensors at
nodal points, in-phase motion of the bearing housing, forces caused by vibration at
high frequencies, or a large amount of opposing runout, when relative-motion
shaft-displacement systems may not exhibit abnormal changes in the machine
condition. Casing accelerometers have a wide frequency range, allowing them to
11. CONDITION MONITORING 147

observe both the low rotational frequencies (as a primary or backup means of
monitoring) and the high blade- and flow-related frequencies.

Fig. 11.33 (from [11.35])

Thermocouples imbedded in thrust pads are recommended for warning of a


thrust-bearing overload. Journal-bearing temperature, obtained from sensors
imbedded in these bearings, is a valuable indicator of bearing performance.

Fig. 11.34 (from [11.35])


148 DYNAMICS OF MACHINERY

Two sensors are installed for axial position measurement and thrust
monitoring, because some turbine conditions, such as blade fouling, can overload
the thrust bearing. Anyhow, steam turbines generally are less susceptible to thrust-
bearing problems than centrifugal problems, which depend on pressure balancing
to maintain thrust load within tolerable levels.
In addition to the measurements discussed for process-drive and boiler-
feed-pump turbines, the unique nature of large utility turbines dictates some
modifications. Figure 11.34 shows that the monitoring system for a thermal power
station turbine generally incorporates some means for obtaining absolute shaft
motion at each radial bearing (either a shaft-riding seismic sensor or a relative-
motion and a casing-absolute-motion sensor, electronically subtracted).
Rotor position indication, accomplished with a noncontact displacement
sensor located at the thrust bearing, should be provided on all turbines. For large
turbines with long bearing spans, it is also necessary to measure rotor eccentricity
while the unit is on turning gear, to warn of a thermal rotor bow, which could result
in packing rubs.
Other capabilities of the monitoring system should include phase reference
and speed measurements, made with noncontact sensors, and valve position
indication, accomplished with potentiometer or similar device.

Fig. 11.35 (from [11.36])


11. CONDITION MONITORING 149

Casing-expansion sensors are necessary to ensure that the sliding shoes are
free and functioning properly, to accommodate the large axial growths of high
temperature turbines. The differential rotor and casing expansion must also be
monitored to avoid rubbing between wheels and diaphragms. As shown in Fig.
11.34, differential expansion can be measured by using a noncontact axial position
sensor attached to the casing, and observing the rotor at the end opposite the thrust
bearing.

Fig. 11.36 (from [11.37])

Fig. 11.37 (from [11.38])


150 DYNAMICS OF MACHINERY

Examples of turbine-generator monitoring systems for different machine


train configurations are given in Figs. 11.35 to 11.37, where the usual conventions
for transducer numbering are shown, as well as in Fig. 10.27.

11.5.2.3 Gas turbines

Compared to other types of industrial rotating machinery, gas turbines


have: a) relatively low casing-to-rotor weight ratios, b) light, flexible casings, and
c) flexible support structures. Though speeds vary greatly, these machines
generally operate at moderate to high speeds. Aircraft derivatives are not
considered herein.

Fig. 11.38 (from [11.35])

Gas turbine vibration signatures, particularly those from units with two or
more independent rotors, contain a large number of spectral components, spanning
a wide frequency range. Along with several running frequencies, the signatures
also may contain components generated by power takeoffs, load and accessory
gearing, turbine- and base-plate-mounted auxiliaries, compressor and turbine
blades, as well as numerous harmonics and sum and difference combinations.
A casing system using accelerometers (Fig. 11.38) is suitable for gas
turbines because of its: a) ability to monitor the mechanical condition of several
components simultaneously, b) quick response to a variety of problems, c) ability
to withstand high temperatures, and d) ease of installation and replacement. Shaft-
vibration sensors are not favored because they cannot collect the data that define
blade and gear condition, and are ineffective on machines using rolling element
11. CONDITION MONITORING 151

bearings. However, they may be required in special applications involving rotor


stability.
Casing accelerometers should be mounted at each bearing and at midspan.
For maximum protection, axial position and journal bearing temperature sensors
should be included.

11.5.3 Electrical machines and gears

Electrical equipment usually has a relatively heavy rotor supported in


bearings mounted on a flexible structure. Thus, most of the dynamic force
developed by the rotor results in structural vibration rather than relative motion
between shaft and bearings. Aside from rotor bar passing frequencies, most of the
characteristics that define the mechanical condition are found in the low frequency
region, up to about four or five times the running speed.
Small electrical machines have rolling element bearings and are monitored
by casing-vibration systems. Thermal power station large generators are monitored
with shaft vibration systems, usually the same as those used for the turbine.

Fig. 11.39 (from [11.39])

High speed industrial gears have moderate bearing preloads and relatively
flexible casings, so that casing vibration monitoring systems are favored. One
sensor at the coupling end of the high speed shaft provides adequate protection on
small gears. On large gears, two accelerometers are usually attached to the gear
152 DYNAMICS OF MACHINERY

casing on, or adjacent to, the coupling end bearings of both the high- and low-
speed shafts. Generally, filters are used to divide a gear signature into manageable
segments. The first segment, containing the rotational frequencies of both shafts,
starts at about 50% of the lowest running speed and extends to the fourth or fifth
multiple of the high speed shaft. The second segment should include frequencies
around 1-2 kHz. The third band should enclose the gear mesh frequency and its
sidebands. The fourth band (if provided) will cover the very high frequencies
generated by pitting and spalling of gear teeth.
The layout of the monitoring system of a motor driven large compressor is
shown in Fig. 11.39. Similar information is given in Fig. 10.25 for a turbine driven
compressor.

11.5.4 Reciprocating compressors

Reciprocating compressors are monitored for problems such as rider band


(rings supporting the piston in the cylinder) wear, leaking valves, and excessive
vibration due to impact-type events or poor mounting/foundation, lack of
lubrication, piston ring wear, and excessive bearing wear. Trending the rod drop
reading provides an early indication of when the riders bands will fail, allowing the
engineers to schedule maintenance at a convenient time. Using both a vertical rod
drop probe and horizontal probe can provide additional valuable diagnostic
information.

Fig. 11.40 (from [11.40])

Since the main bearings of a reciprocating compressor are typically fluid


film bearings, problems related to bearing wear, crankshaft unbalance, or
misalignment can be detected with proximity probes in an X-Y configuration at
each main crankshaft bearing. Continuously measuring temperature at the main
11. CONDITION MONITORING 153

bearings and crosshead slipper provides an early indication of either overloading,


bearing failure, or insufficient lubrication.
A typical layout of the transducers used to monitor a horizontal
reciprocating compressor is illustrated in Fig. 11.40. Proximity probes mounted at
crankshaft are not shown.

Fig. 11.41 (from [11.41])

Figure 11.41 shows the overall layout and the transducer locations and
orientation of a large ammonia and CO2 compressor with a crankshaft with six
throws and rotational speed 330 rpm.
The velomitors (piezo-velocity sensors) are installed horizontally at
each end of the crankcase centerline. Six accelerometers are installed vertically on
the transition sections which connect the cylinders to the crosshead slipper guides.
They are intended to measure the high frequency signals generated by impacts
associated with piston rod looseness and knocking. Rod drop measurements are
made on all stages that have rider rings. At each monitored cylinder, a proximity
probe is mounted vertically on the crosshead oil wiper stuffing box, where it
measures the relative position of piston rod. A thermocouple or RTD mounted near
each valve measures the gas temperature.
Figure 11.42 shows the transducer locations on a large polyethylene
reciprocating compressor. The monitoring system collects and processes the
following data: a) valve temperature on 48 valves using resistive temperature
devices, b) rider band wear using proximity probes on 6 piston rods, c) crankcase
velocity with 4 piezo-velocity sensors per compressor, and d) crosshead
acceleration with 6 accelerometers.
154 DYNAMICS OF MACHINERY

Fig. 11.42 (from [11.42])

Vibration analysis of reciprocating compressors is better carried out in


time domain [11.43]. Velocity transducers are installed on the crankcase in the
horizontal plane parallel to the pistons to detect changes in running speed vibration
and in crankcase deflection. They can be mounted on the cylinder head, or on the
crosshead at a 450 angle in the plane of the piston motion, to detect both vertical
and cylinder vibration called stretch motion. Malfunctions can be detected
looking at timebase data in the region of 1X and 2X running speed
Proximity probes, which measure both the position and motion of the
piston rod, detect rider band wear. They can be used to measure crankshaft
displacement relative to the main bearings which, under normal conditions, follows
an elliptical orbit.
Accelerometers installed on the crosshead or distance piece of each
cylinder detect impact events. Accelerometers can also be used to confirm valve
problems when they are temporarily installed on the valve covers. Valve leaks can
be detected evaluating the timebase waveform relative to a reference that defines
top dead center and bottom dead center. The portion of the stroke in which the high
frequency vibration (produced by gas escaping through the valve) occurs can
isolate the problem to either a suction or discharge valve.
11. CONDITION MONITORING 155

Annex 11.1

Shaft alignment

By alignment, machine shafts are positioned to have their axes collinear


when the machine is running under normal load.
Coupling alignment measurements provide data for calculating the offsets
of each shaft centerline relative to the other, across the coupling distance, in the
horizontal and vertical directions. The maximum alignment deviation is the largest
from driver offset and driven offset values. For the machine operating speed, the
diagram from Fig. 11.11 shows if realignment is necessary.

Fig. A11.1 (from [11.20])

There are basically two cold alignment methods: a) measuring the axial
and radial displacement of one machine with respect to the other (Fig. A11.2, a),
and b) measuring the radial displacement of both machines (Fig. A11.2, b).

a b
Fig. A11.2

1. Rim and face method


In the first method, a bracket-mounted dial indicator is used to take
readings from outside diameter (rim) of the opposite coupling hub and axially on
the inside face of the coupling while one shaft is rotated (Fig. A11.3). The data is
reduced, plotted and corrections calculated.
156 DYNAMICS OF MACHINERY

This method is subject to several sources of error and is not as accurate as


the reverse dial indicator and laser methods, though in some situations it is
necessary to use this method [11.45].

Fig. A11.3 (from [11.44])

2. Reverse dial indicator method


The second method entails two dial indicators mounted on opposite shafts
which are read simultaneously (Fig. A11.4).

Fig. A11.4

The reverse dial indicator method is preferably used with the coupling in
place (Fig. A11.4). The indicators are mounted on brackets with extension arms, to
read the outside diameter of the opposite coupling hub. They are mounted at top-
dead-center on the hub and calibrated to read zero. Readings are taken every 900
as the shafts are rotated. These readings are reduced and plotted on graph paper.
Alignment corrections are measured directly from the graph. This is probably the
most accurate of the dial indicator-based methods. The method is used when laser
systems are unavailable or unsuitable for the machine.
11. CONDITION MONITORING 157

Fig. A11.4 (from [11.44])

The addition of special alignment bars has improved the reverse dial
indicator method, as both indicators can be read with ease as the shafts and
coupling rotate through 3600 .

Fig. A11.5 (from [11.46])


158 DYNAMICS OF MACHINERY

Hot alignment is achieved by replacing the dial indicators with alignment


bars and transducers (Fig. A11.5). The Dodd bar method [11.47] uses proximity
probes mounted on bars. The tubular bars are in a triangular arrangement with
supporting stiffeners between the tubes. One bar contains mounted probes which
are referenced to the other bar for the probe gap. Movement of the machines is
measured by the change in probe gap, and the relative alignment of the train is
calculated and plotted.
The alignment bars, representing the projected center lines of the two
machines, are attached to the inboard bearing housing of each unit. The
noncontacting transducer probes measure the relative movement between the bars,
indicating the thermally induced travel of the two shafts. Indicating blocks are
mounted on the bar fastened to the driven machine. Four proximity probes and two
probe brackets are mounted on the bar attached to the driver.
The probes and indicating blocks are positioned to measure the horizontal
and vertical movement at each coupling hub. The proximity probes measure the air
gap between the probe and indicating block. A proximitor amplifying unit
conditions the electrical energy supplied to the probe and linearizes the return
signal. The proximitor output signal is routed to readout meters calibrated to
display the different movements in displacement units.

Fig. A11.6 (from [11.44])


11. CONDITION MONITORING 159

3. Laser alignment
Laser alignment units consist of a laser fixed to the shaft on one side of the
coupling, behind the hub, and a prism affixed to the shaft on the opposite side of
the coupling. In Fig. A11.6, 1 is the laser support, 2 prism, a laser, b lens, c
focusing device, d filter, e lens, and f detector, 4 adapter, 5 driver, 6
driven machine.
The laser and prism are connected to a dedicated computer. As the shaft is
rotated, the computer records the alignment readings at multiple positions, typically
every 90 0 . The process is usually repeated two or three times to ensure accurate
results. Given machine dimensions, the computer will calculate the amount of
misalignment at the coupling and the corrections necessary at each machine foot to
achieve a correct static, or cold alignment.

4. Optical alignment
Optical alignment equipment generally consists of a precision jig transit or
sight level accurate to 1 arc-second ( 25.4 m over 5.2 meters), a portable
instrument stand, measurement scales, and tooling for mounting the scales on
machines. This method is very accurate and especially useful on long machine
trains. It directly shows the alignment of each rotor in the machine train, and the
catenary shape of the entire shaft system. This is done by placing scales directly on
the shaft and obtaining readings from the scales.
Optical alignment is used on machines that have rigid couplings or are not
easily measured using the previously presented methods. Examples are large
turbine-generator trains and hydroturbines.

References

11.1. * Notes on the use of vibration measurements for machinery condition


monitoring, Brel & Kjaer Application Note, 14-227.
11.2. Sidahmed, M., Diagnostics and condition monitoring, basic concepts,
Encyclopedia of Vibration, Braun, S., Ewins, D., and Rao, S.S., eds.,
Academic Press, London, 2002, p.376-380.
11.3. Collacott, R. A., Vibration Monitoring and Diagnostic, Wiley, New York,
1979.
11.4. Mitchell, J. S., Machinery Analysis and Monitoring, Penn Well Books, Tulsa,
1981.
11.5. Reeves, Ch., Vibration Monitoring Handbook, Coxmoor Publ. Comp., May
1998.
160 DYNAMICS OF MACHINERY

11.6. Beebe, R. S., Predictive Maintenance of Pumps Using Condition Monitoring,


Elsevier, Oxford, 2004.
11.7. Bloch, H. P., Practical Machinery Management for Process Plants, vol.1:
Improving Machinery Reliability, 3rd ed., Gulf Professional Publ., Oxford,
1998.
11.8. Eisenmann, R. C. Sr., and Eisenmann, R. C. Jr., Machinery Malfunction
Diagnosis and Correction, Hewllet Packard Professional Books, Prentice
Hall, Upper Saddle River, NJ, 1997.
11.9. Bloch, H. P. and Gleitner, F. K., Practical Machinery Management for
Process Plants, vol.4: Major Process Equipment Maintenance and Repair,
2nd ed., Gulf Publishing Comp., Houston, 1997.
11.10. Gleitner, F. K. and Bloch, H. P., Practical Machinery Management for
Process Plants, vol.5: Maximizing Machinery Uptime, Gulf Professional
Publ., Oxford, 2006.
11.11. Forsthoffer, W. E., Forsthoffers Rotating Equipment Handbooks, vol.5
Reliability Optimization through Component Condition Monitoring and Root
Cause Analysis, Elsevier, Amsterdam, 2005.
11.12. Macdonald, D., Practical Machinery Safety, Newness, Oxford, 2004.
11.13. Mobley, R. K., An Introduction to Preventive Maintenance, 2nd ed.,
Butterworth-Heinemann, Amsterdam, 2002.
11.14. Scheffer, C. and Girdhar, P., Practical Machinery Vibration Analysis and
Predictive Maintenance, Newnes, Oxford, 2004.
11.15. Barron, R., Engineering Condition Monitoring: Practice, Methods and
Applications, Addison Wesley Longman, London, 1996.
11.16. Adams, M. L., Rotating Machinery Vibration: From Analysis to
Troubleshooting, Marcel Dekker, New York, 2001.
11.17. Mechefske, C. K., Machine condition monitoring and fault diagnostics,
Chap.25 of C.R.C. Handbook, Taylor and Francis, 2005.
11.18. Mitchell, J. S., Bearing diagnostics: An overview, ASME Winter Ann. Mtg.,
10-15 Dec 1978, San Francisco, p.15-24, 1978.
11.19. Southwick, D., Using full spectrum plots, Part 2, Orbit, vol.15, no.2, June
1994, p.11-15.
11.20. Piotrowski, J., Shaft Alignment Handbook, 2nd ed., Marcel Dekker Inc.,
New York, 1995.
11.21. Muszynska, A., Vibrational diagnostics of rotating machinery malfunctions,
Course on Rotor Dynamics and Vibration in Turbomachinery, von
Karman Institute for Fluid Dynamics, Belgium, 21-25 Sept 1992.
11. CONDITION MONITORING 161

11.22. Laws, W. C. and Muszynska, A., Periodic and continuous vibration


monitoring for preventive/predictive maintenance of rotating machinery,
Journal of Engineering for Gas Turbines and Power, vol.109, April 1987,
p.159-167.
11.23. Southwick, D., Using full spectrum plots, Orbit, vol.14, no.4, Dec 1993,
p.19-21.
11.24. * Machinery protection and diagnostics topics, Bently Nevada
Application Note 003, Feb 1977.
11.25. Bently, D. E., Forced subrotative speed dynamic action of rotating
machinery, ASME Paper 74-Pet-16, Petroleum Mechanical Engineering
Conference, Dallas, Texas, Sept 1974.
11.26. Muszynska, A., Partial lateral rotor to stator rubs, I. Mech. E. Conference
Publication 1984-10, Proc. Third International Conference on Vibrations in
Rotating Machinery, Heslington, England, 11-13 Sept 1984, p.327-335.
11.27. Bently, D. E., Basic rotor-to-stator thermal rubs which exhibit rotative
speed (1X) symptom only, Orbit, vol.17, no.3, Sept 1996, p.4-6.
11.28. Gasch, R., Nordmann, R. and Pftzner, H., Rotordynamik, 2nd ed., Springer,
Berlin, 2001.
11.29. Muszynska, A., Misalignment and shaft crack-related phase relationships
for 1X and 2X vibration components of rotor responses, Orbit, vol.10, no.2,
Sept.1989, p.4-8.
11.30. Pfleiderer, C., Petermann, H., Strmungsmaschinen, 6th ed., Springer, 1990.
11.31. Mitchell, J. S., Putting vibration and other operating variables to work in a
monitoring system, Power, May 1977, p.87-89.
11.32. * Centrifugal Compressor for Ultra-High Pressures, Druckschrift MA
25.69 en/9.83, Mannesmann Demag, 1983.
11.33. Eck, B., Ventilatoren, Springer, Berlin, 1957.
11.34. Kostyuk, A. G. and Frolov, V. V., Steam and Gas Turbines (in Russian),
Energoatomizdat, Moskow, 1985.
11.35. Mitchell, J. S., Monitoring the complex vibration characteristics of bladed
machinery, Power, July 1977, p.38-42.
11.36. * Rotating Machinery Information Systems and Services, Bently
Nevada, Publ. L1001-00, April 1993.
11.37. Hayashida, B., Advancement of Turbine Supervisory Instrumentation
continues to help solve machinery problems, Orbit, vol.13, no.1, Feb 1992,
p.6-11.
162 DYNAMICS OF MACHINERY

11.38. Murray, G., Mucci, J., and Brier, S., Analysis of generator rotor unbalance,
Orbit, vol.14, no.1, March 1993, p.25-29.
11.39. Swan, P., Torsional vibration problems with asynchronous motor, Orbit,
vol.18, no.1, March 1997, p.22-24.
11.40. * Monitoring reciprocating compressors, Orbit, vol.11, no.3, Dec 1990,
p.20-23.
11.41. Silcock, D., Reciprocating compressor instrumented for machinery
management, Orbit, vol.17, no.2, June 1996, p.10-12.
11.42. Smith, T., Quantum Chemical uses reciprocating compressor monitor to
improve reliability, Orbit, vol.17, no.2, June 1996, p.14-16.
11.43. Schultheis, S. M., Vibration analysis of reciprocating compressors, Orbit,
vol.17, no.2, June 1996, p.7-9.
10.44. * Vorbeugende Maschineninstandhaltung, Schenck Seminar C 50, Nov
1989, p.70.
11.45. Bognatz, S. R., Alignment of citical and noncritical machines, Orbit, vol.16,
no.1, March 1995, p.23-25.
11.46. Campbell, A. J., Static and dynamic alignment of turbomachinery, Orbit,
vol.14, no.2, June 1993, p.24-29.
11.47. Dodd, V. R., Shaft alignment monitoring cuts costs, Oil and Gas Journal,
Sept 1971.
12.
VIBRATION LIMITS

This chapter presents guidelines and recommendations providing


acceptable vibration limits for different types and sizes of machinery.

12.1 Broadband vibration standards and guidelines

In order to assess the mechanical condition of a machine, vibration levels


measured either on the bearing housings or between the journal and the bearing
case are compared against guideline charts and recommended acceptable limits.
In the development of standards it has been found that machinery can be
subdivided into four categories for the purposes of vibration measurement and
evaluation [12.1]:
1. Reciprocating machinery having both rotating and reciprocating
components, such as diesel engines and certain types of compressors and pumps.
Vibrations are usually measured on the main structure of the machine at low
frequencies.
The Guideline VDI 2063-1985 [12.2], intended for the measurement and
evaluation of mechanical vibrations of reciprocating piston engines and piston
compressors, has proved to be useful in practice, though the same criteria were
applied to all reciprocating machines. It was superseded by the standard ISO
10816-6 [12.3] which gives different vibration limits for various machines.
2. Rotating machinery having rigid rotors, such as certain types of electric
motors, single-stage pumps, and slow-speed pumps. Vibrations are usually
measured on the main structure (such as on the bearing caps or pedestals) where
the vibration levels are indicative of the excitation forces generated by the rotor
because of unbalance, thermal bows, rubs, and other sources of excitation.
164 DYNAMICS OF MACHINERY

Vibration severity is defined as the maximum value of the broadband root-


mean-square velocity in the specified frequency range (typically from 10 to 1,000
Hz), as evaluated on the structure at prescribed points.
The Guideline VDI 2056-1964 [12.4] was the basis for the standards ISO
2372-1974 [12.5] and ISO 2373 [12.6] superseded now by ISO 10816 [12.7]-
[12.12]. ISO standards are used as a basis for the corresponding national standards.
The obsolete standards still contain useful information.
3. Rotating machinery having flexible rotors, such as large steam turbine
generators, multistage pumps and compressors. The machine may be set into
different modes of vibration as it accelerates through one or more critical speeds to
reach its service speed. On such a machine, the vibration amplitude measured on a
structure member may not be indicative of the vibration of the rotor. For example,
a flexible rotor may experience very large amplitude displacements resulting in
failure of the machine, even though the vibration amplitude measured on the
bearing cap is very low. Therefore, it is essential to measure the vibration on the
shaft directly.
The Guideline VDI 2059-1981 (first draft in 1972) comprised 5 parts
[12.13]-[12.17] devoted to general guidelines and four types of turbines. It was the
basis for the standard ISO 7919-1996 which has also 5 parts [12.18]-[12.22].
4. Rotating machinery having quasi-rigid rotors, such as low-pressure
steam turbines, axial-flow compressors, and fans. Such machinery contains a
special class of flexible rotor where vibration amplitudes measured on the bearing
cap are indicative of the shaft vibration.
In addition to the International Standards Organization (ISO), various trade
organizations such as American Petroleum Institute (API), American Gear
Manufacturers Association (AGMA) and the American National Electrical
Manufacturers Association (NEMA) have developed and published vibration
standards, which are widely accepted and applied. In most cases, these standards
have been developed by consensus of consumers and manufacturers, and their use
is considered voluntary.

12.2 Vibration severity charts

The Rathbone Chart (Fig.12.1) was the first vibration guideline chart
produced for the insurance industry whose business depends upon correctly
assessing the mechanical condition of machinery it insures [12.23]. It is limited to
turbines on individual foundations, running at speeds less than 6000 rpm and with
small ratios of shaft vibration to bearing housing or pedestal.
12. VIBRATION LIMITS 165

Six curves limit the zones for different mechanical conditions, ranging
from very smooth to very rough. These categories are for overall (broad-band)
vibration measurements at the machine bearing housing. Above 20 Hz, the
boundaries are defined by lines with slopes of ( 1) on the log-log diagram, which
plots the peak-to-peak displacement amplitude, in mils, as a function of frequency
(1 mil = 25.4 m ). They represent constant peak velocity lines. The line indicating
the sensory perception level [12.24] is also included.

Fig. 12.1 (from [12.23])

Subsequent development of vibration severity charts followed this format.


The chart developed by Blake [12.25] is given in Annex A12.1. Horizontal
displacements measured on bearings have to be multiplied by a service factor.
166 DYNAMICS OF MACHINERY

Figure 12.2 shows the general severity chart for bearing cap measurement
(filtered readings) developed first by H. G. Yates (1949) and reworked in 1964 by
IRD Mechanalysis [12.26]. It was used only as a guide in judging vibrations as a
warning of impending trouble. The diagram plots the peak-to-peak displacement
amplitude versus frequency as constant peak velocity lines, with a step of 2
between severity levels (1in s = 25.4 mm s) . Measurement of the peak-to-peak
vibration level was the current practice in the U.S.A. until 1974.

Fig. 12.2 (from [12.26])

Meanwhile, the VDI Vibration Group developed the Guideline VDI 2056,
first released in 1960, then revised and completed in 1964 [12.27]. The vibration
intensity was defined by the root-mean-square of the vibration velocity. The
guideline was limited to mechanical vibrations above 5 Hz measurable at the
surface, at bearings or at fixing points.
An assessment scale was built up, starting from the average limit of human
perception, 0.112 mm s , and progressing in a ratio of 1.6 (4dB) for the limits of
12. VIBRATION LIMITS 167

vibration intensity levels. The reason was that experience has shown that a 1.6
times increase in the velocity is distinctly perceptible or detectable in its effects and
of importance for the stressing of the machine. A second improvement was the
differentiation of the four quality classes: A good, B allowable, C just
tolerable and D not permissible, for six different groups of machines.
The four groups of machines for which vibration intensity limits have been
suggested are the following: Class I (Group K) - individual parts of engines and
machines, integrally connected to the complete machine in its normal operating
condition; Class II (Group M) - medium sized machines (typically electrical
motors with 15 kW to 75 kW output) without special foundations, rigidly mounted
engines or machines (up to 300 kW) on special foundations; Class III (Group G) -
large prime-movers and other large machines with rotating masses mounted on
rigid and heavy foundations which are relatively stiff in the direction of vibration
measurements; Class IV (Group T) - large prime-movers and other large machines
with rotating masses mounted on foundations which are relatively soft in the
direction of vibration measurements (for example, turbogenerator sets and gas
turbines with outputs greater than 10 MW).
The vibration severity ranges for the four groups of machines are listed in
Table 12.1 [12.4]. The operating zones B and C cover double-step severity ranges.

Table 12.1
168 DYNAMICS OF MACHINERY

Increased recognition of VDI 2056 by both manufacturers and users of


prime movers and driven machinery led to the standard ISO 2372-1974 [12.5],
merely the English version of the VDI 2056 recommendations. The term vibration
intensity was replaced by vibration severity, so that the r.m.s. value of vibration
velocity (over the frequency range 10 to 1000 Hz) was recognized as the best
figure of merit for vibration effects on non-rotating parts of machinery. Table 12.1
can also be found in the above standard. It was maintained in the first release of
ISO 10816-1 [12.7] as a short term expedient only, until the relevant parts of the
standard became available.
Although the absolute values suggested by these criteria are not always
relevant, due to the different mobilities of the machine structures at the
measurement location, they are useful in that they indicate the significance of
various degrees of vibration level increases. For example, a level increase by a
factor of 2.5 (8dB) is a significant change as it is the span of one quality class.
Likewise, an increase by a factor greater than 10 (20dB) is serious as it can take the
classification from good to not permissible.

12.3 Vibration limits for non-rotating parts

This section presents the vibration criteria suggested by the standard ISO
10816 presently in use.

12.3.1 General guidelines

The standard ISO 10816-1 [12.7] provides general guidelines that describe
criteria for the evaluation of vibration based on measurements made on the non-
rotating parts of the machine. These criteria, which are presented in terms of both
vibration magnitude and change of vibration, relate to operational monitoring and
acceptance testing.
This is Part 1 of a series of standards that has been written to: a) cover the
broadband frequency range of both low and high speed machines; b) set the
vibration criteria to include the various operational zones, irrespective of whether
they are increases or decreases; c) incorporate vibration criteria through a
worldwide survey; and d) include unique criteria and measurement procedures for
specific types of machines.
In addition to vibration velocity measurements, which were the primary
criteria in earlier Standards because they related to vibration energy, the ISO 10816
series also includes alternate criteria such as displacement, acceleration, and peak
values instead of r.m.s., as these criteria may be preferred for machines designed
for extra low or high speed operation.
12. VIBRATION LIMITS 169

The measurement of vibration is broadband and the frequency band is


sufficient to ensure that the particular machine is adequately covered, which
depends on the type of machine under consideration. For example, the frequency
range necessary to assess the integrity of a machine with rolling element bearings
should include frequencies higher than those of machines with fluid film bearings.
For long-term operation, the standard shows how to set operational
vibration limits under the form of alarms and trips.
Alarms provide a warning that a defined value of vibration has been
reached or a significant change has occurred, at which remedial action may be
necessary. If an alarm situation occurs, operation can continue for a period whilst
investigations are carried out to identify the reason for the change in vibration and
define any remedial action.
Trips specify the magnitude of vibration beyond which further operation of
the machine may cause damage. If the trip value is exceeded, immediate action
should be taken to reduce the vibration or the machine should be shut down.

12.3.2 Steam turbine sets

The standard ISO 10816-2 [12.8] provides specific guidance for assessing
the severity of vibrations measured on the bearings or pedestals of large turbine
generating sets.
Table 12.2

Shaft rotational speed, rpm


1500 or 1800 3000 or 3600
Zone boundary
Vibration velocity, mm s r.m.s.
A/B 2.8 3.8
B/C 5.3 7.5
C/D 8.5 11.8

The designation of zones is the following:


Zone A - new machines that can be operated without restriction.
Zone B - acceptable for unrestricted long term operation.
Zone C - machines that may be operated for a limited time until a
suitable opportunity arises for remedial action to be taken.
Zone D - vibrations of sufficient severity to cause damage to the
machine.
The vibration measurement system should be capable of measuring
broadband vibrations in mm s r.m.s. over a frequency range 10-500 Hz. If,
170 DYNAMICS OF MACHINERY

however, the instrumentation is also to be used for diagnostic purposes, or for


monitoring during machine run-up or run-down, or overspeed, a wider frequency
may be required.
This standard includes the vibration criteria shown in Table 12.2. It is
based on bearing housing/pedestal vibration velocity amplitude ( mm s r.m.s.) for
turbine generator sets exceeding 50 MW, and with nominal speeds of 1500, 1800,
3000 and 3600 rpm. The values apply to in situ measurement under steady-state
conditions.
It is recommended that the alarm value should not normally exceed 1.25
times the upper limit of zone B. In general, the trip value will be within zone C or
D, but it is recommended that the trip value should not exceed 1.25 times the upper
limit of zone C.

12.3.3 Coupled industrial machines

The standard ISO 10816-3 [12.9] provides specific guidance for assessing
the severity of vibrations on bearings, bearing pedestals, or the housings of coupled
industrial machines when measured in situ. This standard covers the following
machines: steam turbines with power above 50 MW, compressors, industrial gas
turbines with power up to 3 MW, pumps with power up to 1 MW, generators,
electric motors of any type, and blowers with power greater than 300 kW.

Table 12.3
Support Zone Displacement Velocity
class boundary ( m ,r.m.s.) (mm s, r.m.s.)
A/B 37 2.3
Rigid B/C 72 4.5
C/D 113 7.1
A/B 56 3.5
Flexible B/C 113 7.1
C/D 175 11.0

Significant differences in design, types of bearings, and types of support


structures require a division of this standard into two machinery groups, namely: 1)
large machines with rated power above 300 kW, or electrical machines with shaft
heights over 315 mm; and, 2) medium size machines with a rated power above 30
kW up to and including 300 kW, or electrical machines with shaft heights from 180
mm to 315 mm. The larger machines normally have sliding bearings and the range
of operating or nominal speed is relatively broad with ranges from 120 rpm to
15000 rpm. The recommended criteria are shown in Fig. 12.3.
12. VIBRATION LIMITS 171

Classification of the vibration severity zones for large industrial machines


with rated power from 300 kW to 50 MW (group 1) is shown in Table 12.3. The
zone descriptions are the same as in ISO 10816-2.

Table 12.4

Support Zone Displacement Velocity


class boundary ( m , r .m.s.) (mm s, r.m.s.)
A/B 22 1.4
Rigid B/C 45 2.8
C/D 71 4.5
A/B 37 2.3
Flexible B/C 71 4.5
C/D 113 7.1

Classification of the vibration severity zones for medium size industrial


machines with rated power from 15 kW to 300 kW (group 2) is included in Table
12.4.

Fig. 12.3

Based on ISO 10816-3, specific vibration severity standards have been


established for pumps [12.28]. They apply to pumps with multivane impeller
(centrifugal, mixed or axial flow above 15 kW) and separate or integrated driver.
172 DYNAMICS OF MACHINERY

12.3.4 Gas turbine sets

The standard ISO 10816-4 [12.10] provides specific guidance for assessing
the severity of vibrations measured on the bearing housings or pedestals of gas
turbine sets. This standard applies to heavy-duty gas turbines used in electrical and
mechanical drive applications covering the power range above 3 MW, and a speed
range under load between 3000 and 20,000 rpm. Generally, the criteria apply to
both the gas turbine and the driven equipment. However, for generators above 50
MW, the criteria of ISO 10816-2 should be used, and for compressors in the power
range from 30 to 300 kW, the criteria of ISO 10816-3 should be used for assessing
the vibration severity.
The evaluation of zone boundaries based on bearing housing/pedestal
vibration for industrial gas turbines is given in Table 12.5. These criteria assume
that the gas turbines incorporate fluid film bearings, and the vibration
measurements are broadband values taken in situ under normal steady-state
operating conditions.
Table 12.5

Shaft rotational speed, Zone boundary


rpm
A/B B/C C/D
Vibration velocity, mm s r.m.s.
3000-20000
4.5 9.3 14.7

This standard encompasses machines which may have gears or rolling


element bearings, but does not address the evaluation of the condition of those
gears or bearings. The zone descriptors are the same as in ISO 10816-2.

12.3.5 Hydraulic machines

The standard ISO 10816-5 [12.11] provides specific guidance for assessing
the severity of vibrations measured on bearings, bearing pedestals, or housings of
hydraulic machines when measured in situ. It applies to machine sets in hydraulic
power generation, and pump plants where the hydraulic machines have speeds
from 120 to 1800 rpm, shell- or shoe-type sliding bearings, and main engine power
of 1 MW or more. The position of the shaft line may be vertical, horizontal, or at
any arbitrary angle between these two directions.
This Standard includes: turbines and generators, pumps, and electrical
machines operating as motors, pump-turbines, and motor generators, including
auxiliary equipment (e.g., starting turbines or exciters in line with the main shaft).
The standard also includes single turbines or pumps connected to generators or
electric motors over gears and/or radially flexible couplings.
12. VIBRATION LIMITS 173

Fig. 12.4 (from [12.29])


The recommended criteria values (in mm s r.m.s.) vs. shaft rotational
speed (in rpm) for hydraulic machines with nominal power above 1 MW, and
nominal speeds between 120 and 1800 rpm are shown in Fig. 12.4. The zone
descriptions are the same as in ISO 10816-2.

Fig. 12.5 (from [12.30])


174 DYNAMICS OF MACHINERY

Vibration limits set by the Hydraulics Institute for horizontal clear liquid
pumps, measured on bearing housing are given in Fig. 12.5. Vibration tolerances
set by the Hydraulic Institute Application Standards B-74-1: 1967 [12.26] for
centrifugal pumps are shown for comparison in Annex A12.2.

12.3.6 Reciprocating machines

The standard ISO 10816-6 [12.3] establishes procedures and guidelines for
the measurement and classification of mechanical vibrations of reciprocating
machines. In general, this standard refers to vibration measurements made on the
main structure of the machine, and the guide values are defined primarily to secure
a reliable and safe operation of the machine, and to avoid problems with the
auxiliary equipment mounted on the structure.
Table 12.6
Vibration Maximum levels of overall vibration
measured on the machine structure Machine vibration classification number
severity
grade Displacement Velocity Acceleration
m, r.m.s. mm/s,r.m.s. m/s, r.m.s. 1 2 3 4 5 6 7
boundary
1.1 to 1.8 -------17.8----- -----1.12---- -----1.76-----

1.8 to 2.8 -------28.3----- -----1.78---- -----2.79-----


A/B
2.8 to 4.5 -------44.8----- -----2.82---- ------4.42---- A/B
A/B
4.5 to 7.1 -------71.0----- -----4.46---- ------7.01---- A/B
A/B
C
7.1 to 11 -------113----- -----7.07---- ------11.1---- A/B
A/B
C
11 to 18 -------178----- -----11.2---- ------17.6----
C
18 to 28 -------283----- -----17.8---- ------27.9----
C
28 to 45 -------448----- -----28.2---- ------44.2---- D
D C
45 to 71 -------710----- -----44.6---- ------70.1---- D
D C
71 to 112 ------1125---- -----70.7---- ------111----
D
D C
112 to180 ------1784---- -----112---- ------176----
D

Based on experience with similar machines, the damage that can occur
when exceeding the guide values is sustained predominantly by the machine-
mounted components (e.g., turbochargers, heat exchangers, governors, pumps,
filters, etc.), connecting elements of the machine with peripherals (e.g., pipelines),
or monitoring instruments (e.g., pressure gauges, thermometers, etc.). For rigidly
seated reciprocating piston engines, vibration levels are measured at the top edge of
the frame or cylinder cover. This standard generally applies to reciprocating piston
12. VIBRATION LIMITS 175

machines mounted either rigidly or resiliently with power ratings above 100 kW.
The vibration criteria for seven different classes of reciprocating machines are
presented in Table 12.6.
The class definitions are: 1) balanced opposed type rigidly mounted
reciprocating gas compressors; 2) multi-throw type rigidly mounted reciprocating
gas compressors; 3) single-throw type rigidly mounted reciprocating gas
compressors; 4) no example; 5) and 6) industrial and marine diesel engines (<2000
rpm); and, 6) and 7) industrial and marine diesel engines (>200 kW). The zone
descriptions are the same as in ISO 10816-2.
The values in Table 12.6 were derived from constant displacement in the
range 2 Hz to 10 Hz, constant velocity from 10 Hz to 250 Hz and constant
acceleration from 250 Hz to 1000 Hz. Vibration values for reciprocating machines
may tend to be more constant over the life of the machine than for rotating
machines. Therefore zones A and B are combined in this table. In future, when
more experience is accumulated, guide values to differentiate between zones A and
B may be provided.

Fig. 12.6 (after [12.2])

The guideline VDI 2063 [12.2] prescribed as limit values 1 mm peak


displacement in the range 2 Hz to 10 Hz, 45 mm/s r.m.s. velocity (or 68 mm/s peak
velocity) from 10 Hz to 100 Hz, and 4g peak acceleration from 100 Hz to 300 Hz
(Fig. 12.6). The limiting curve is overlaid on the vibration severity grade
nomograph given in the standard ISO 10816-6: 1995 [12.3] in Annex A12.5.
Guidelines for the evaluation of vibrations of reciprocating internal
combustion engine-driven alternating current generating sets are given in the
standard ISO 8528-9: 1995.
176 DYNAMICS OF MACHINERY

12.4 Vibration limits for rotating parts

The first vibration severity chart based on shaft peak-to-peak


displacements measured relative to bearings [12.31] was developed in 1968 for
Dresser Clark centrifugal compressors (see Annex A12.4). Alternative severity
criteria are given in Annex A12.3 [12.30]. This section presents the vibration
criteria suggested by the standard ISO 7919 (based on VDI 2059), presently in use
for measurement on rotating shafts.

12.4.1 General guidelines

The standard ISO 7919-1 [12.18], provides specific guidelines for


vibration measurements on the rotating members of machines. Such machines
generally contain flexible rotor-shaft systems. Change in the vibration condition
may be detected more sensitively by measurements on these rotating elements.
Also, machines having relatively stiff and/or heavy casings, in comparison to the
rotor mass, are typical of those classes of machines for which shaft vibration
measurement are frequently preferred.
Machines such as industrial steam turbines, gas turbines, and
turbocompressors have several modes of vibration in their service speed range,
and, their responses due to unbalance, misalignments, thermal bows, rubs, and the
unloading of bearings may be better observed by measurements on the shafts.
There are three principal factors by which the vibration level of a machine
is judged [12.29], namely: a) bearing kinetic load; b) absolute motion of the rotor;
and, c) rotor clearance relative to the bearing. If the bearing kinetic load is of
concern to ensure against bearing damage, the vibration of the shaft relative to the
bearing structure should be monitored as the overriding criterion. If the absolute
motion of the shaft (a measure of the rotor bending stress) or rotor-bearing
clearance are of concern, the type of measurement used depends on the vibration
level of the structure which supports the relative motion transducer. Hence, if the
vibration level of this support structure is less than 20% of the relative shaft
vibration, the absolute shaft vibration must be measured; and, if this is found to be
larger than the relative shaft vibration, then this will be the more valid
measurement. The rotor clearance to the bearing must be monitored to ensure
against rotor seal and blading rubs which can cause rotor or blading failures.
The shaft vibrations of machines, measured close to the bearings, are
evaluated on the basis of two criteria [12.29]:
1) The reliable and safe running of a machine under normal operating
conditions requires that the shaft vibration displacement remain below certain
limits consistent with, for example, acceptable kinetic loads and adequate margins
on the radial clearance envelope for the machine. Generally, this criterion is taken
12. VIBRATION LIMITS 177

as the basis for the evaluation of a new machine, in the absence of any other
established knowledge of the satisfactory running characteristics for a machine of
that type.
2) Changes in shaft vibration displacement, even though the limits in 1) are
not exceeded, may point to incipient damage or some other irregularity.
Consequently, such changes relative to a reference value should not be allowed to
exceed certain limits. If this reference value changes by a significant amount, and
certainly if it exceeds 25% of the reference level, steps should be taken to ascertain
the reasons for the change and, if necessary, appropriate action should be taken. In
this context, a decision on what action to take, if any, should be made after
consideration of the maximum value of vibration, and whether the machine has
stabilized at a new condition.
The standard ISO 10817-1 [12.32] describes the sensing device
(transducer), signal conditioning, attachment methods, and calibration procedures
for instrumentation to measure shaft vibration.

12.4.2 Steam turbine sets

The standard ISO 7919-2 [12.19], based on VDI 2059-2 [12.14], provides
the special features required for measuring shaft vibrations on the coupled rotor
systems of steam turbine-generating sets for power stations, having rated speeds in
the range 1500-3600 rpm, and power outputs greater than 50 MW. Evaluation
criteria, based on previous experience, are presented which may be used as
guidelines for assessing the vibratory conditions of such machines.

Table 12.7
Shaft rotational speed, rpm
1500 1800 3000 3600
Zone boundary
Peak-to-peak relative displacement of shaft, m
A/B 100 90 80 75
B/C 200 185 165 150
C/D 320 290 260 240

The vibration levels specified here define four quality zones for both
relative and absolute shaft vibration measurement at, or close to, the main load-
carrying bearings, at rated speed and under steady state conditions. Higher levels of
vibration can be permitted at other measuring locations and under transient
conditions, such as start-up and run-down (including acceleration through critical
speed ranges).
178 DYNAMICS OF MACHINERY

The recommended shaft vibration amplitude values for large steam


turbine-generator sets, in micrometers peak-to-peak, measured relative to the
bearings, are shown in Table 12.7 for relative shaft to bearing vibrations, and in
Table 12.8 for absolute vibrations.
Table 12.8
Shaft rotational speed, rpm
1500 1800 3000 3600
Zone boundary
Peak-to-peak absolute displacement of shaft, m
A/B 120 110 100 90
B/C 240 220 200 180
C/D 385 350 320 290

In both tables, zone A represents new machines that can be operated


without restriction; zone B is acceptable for long-term operation; zone C represents
machines that may be operated for a limited time until a suitable opportunity arises
for remedial action to be taken; and zone D is identified as a trip level as these
values are considered to be of sufficient severity to cause damage.

Fig. 12.7 (from [12.14])

Figure 12.7 shows for comparison the limit values for shut-down given in
VDI 2059-2 [12.14]. They represent maximum values of the radius of shaft
precession orbit, hence peak displacements. It can be noticed that the values in
Table 12.7 at the zone boundary C/D are almost the same, though they represent
peak-to-peak values, suggesting to multiply by 2 values in Tables 12.7 and 12.8.

12.4.3 Coupled industrial machines

The standard ISO 7919-3 [12.20], based on VDI 2059-3 [12.15], provides
guidelines for application of evaluation criteria based on shaft vibrations measured
close to the bearings under normal operating conditions. These guidelines are
12. VIBRATION LIMITS 179

presented in terms of both steady running conditions, and any changes that may
occur in these steady values.

Fig. 12.8 (from [12.30])

This standard applies to coupled industrial machines with fluid film


bearings, comprising: turbocompressors, turbines, turbine-generators, and electric
drives, all having maximum rated speeds in the range 1000 to 30,000 rpm, and
powers between 30 kW and 50 MW.
In Fig. 12.8, limits are shown for the evaluation of the peak-to-peak shaft
displacement relative to the bearing, d, as a function of the rotational speed, n.
The three lines at zone boundaries are defined by the following equations
180 DYNAMICS OF MACHINERY

4800
Zone limit A/B d A B (in m ) = ; (12.1)
n (in rpm )

9000
Zone limit B/C d B C (in m ) = ; (12.2)
n (in rpm )

13,200
Zone limit C/D dC (in m ) = . (12.3)
n (in rpm )
D

The numerical values specified in Fig. 12.8 are not intended to serve as the
only basis for acceptance specifications. In general, the vibratory condition of these
machines is usually assessed by consideration of both the shaft vibration and the
associated structural vibration. As a result, this Standard should be used in
conjunction with ISO 10816-3 [12.9]. The zone descriptions of Fig. 12.8 are the
same as in ISO 7919-2.
For comparison, the corresponding chart from the recommendations VDI
2059-3 [12.15] is reproduced in Annex A12.6.

12.4.4 Gas turbine sets

The standard ISO 7919-4 [12.21] applies to heavy-duty gas turbines, used
in electrical and mechanical drive applications (including those with gears), with
fluid film bearings, power outputs greater than 3 MW, and shaft rotational speeds
under load from 3000 to 30,000 rpm. This includes gas turbines directly coupled to
other prime movers such as steam turbines. Aircraft type gas turbines are excluded,
since they differ fundamentally from industrial gas turbines, in the types of
bearings (rolling element), casing flexibility, mounting structure and rotor to stator
weight ratio.
Depending on the construction and mode of operation, there are three types
of industrial gas turbines: 1) single-shaft constant-speed; 2) single-shaft variable-
speed; and, 3) gas turbines having separate shafts for hot-gas generation and power
delivery.
Guidelines are given in Fig. 12.9 for the application of shaft vibration
criteria measured close to the bearings of industrial gas turbines under normal
operating conditions. The zone descriptions are the same as in ISO 7919-2.
Figure 12.9 is basically similar to Fig. 12.8 except for the range of
rotational speeds which starts at 3000 rpm. The three lines defining the zone
boundaries are defined by the same equations (12.1)-(12.3).
The corresponding chart from the recommendations VDI 2059-4 [12.16] is
given for comparison in Annex A12.7.
12. VIBRATION LIMITS 181

Fig. 12.9 (from [12.30])

12.4.5 Hydraulic machine sets

The standard ISO 7919-5 [12.22] lists the special features required for
measuring shaft vibrations on coupled hydraulic sets. This standard applies to all
types of hydraulic machines having nominal speeds between 60 and 3600 rpm,
with fluid film bearings and rated powers of 1 MW or more.
These machines may consist of turbines, pumps, pump-turbines,
generators, motors, and motor-generators, including couplings, gears, or auxiliary
equipment in the shaft line. The position of the shaft may be vertical, horizontal, or
at an arbitrary angle between these two directions.
It is not applicable to pumps in thermal power plants or industrial
installations, hydraulic machines or machine sets having rolling element bearings,
or hydraulic machines with water-lubricated bearings.
182 DYNAMICS OF MACHINERY

Fig. 12.10 (from [12.30])

The guidelines are given for the application of shaft vibration criteria as
measured close to the bearings of coupled hydraulic sets, under normal operating
and steady state conditions, and any changes that may occur in these steady values.
The numerical values specified in Fig. 12.10 present rotor displacements
relative to the bearings vs. shaft rotational speed. It is limited to the range of
nominal rotational speeds from 60 to 2000 rpm. The zone descriptions are the same
as in ISO 7919-2.
For comparison, the corresponding chart from the recommendations VDI
2059-5 [12.17] is reproduced in Annex A12.8.
12. VIBRATION LIMITS 183

12.4.6 Selection of measurements

An umbrella document released as Part 0 of ISO 10816 [12.33] provides


general guidelines for selecting the appropriate vibration standards for a specific
machinery classification. The proposed method includes two basic evaluation
criteria: a) shaft displacement from the journal centerline; and b) stiffness ratio of
pedestal to bearing. The latter determines the ratio of the shaft relative
displacement to the pedestal vibration.

Fig. 12.11 (from [12.29])

Figure 12.11 shows the flow diagram for selection of measurements and
evaluation of vibration severity.
184 DYNAMICS OF MACHINERY

Figure 12.12 gives suggestions when to make absolute vibration


measurements and when to make shaft relative vibration measurements.

Fig. 12.12 (from [12.29])

In general, machines equipped with rolling element bearings will tend to


have high bearing stiffness, a stiffness ratio less than 1, and are better suited to
vibration measurements at the pedestal and/or casing. Conversely, machines using
fluid film bearings and supported on relatively soft pedestals, will have a much
higher stiffness ratio, and are better suited to shaft vibration measurements.
Table 12.9 [12.29] shows example dynamic stiffness ratios, and the
applicability of the reference standard.
12. VIBRATION LIMITS 185

Table 12.9
Dynamic ISO 10816 ISO 7019
Machine stiffness ratio, (pedestal) (shaft)

High pressure turbine 5 Moderate Good
Low pressure turbine 1.5 Moderate Good
Large generator 1.5 Moderate Good
High pressure centrifugal compressor 5 Not Good Good
Large fan 2/3 Good Moderate
Small fan & pump 1/3 Good Moderate
Vertical pump 1/10 Good Not Good
Large steam turbine generator set 1.5 to 3 Moderate Good

12.5 Gear units


ISO 8579-2 [12.34] specifies the methods for determining the mechanical
vibration of individually housed, enclosed, speed-increasing and speed-reducing
gear units. This standard specifies methods for measuring housing and shaft
vibrations, and the types of instrumentation, measurement methods and testing
procedures for determining vibration levels. It applies only to a gear unit under test
and operating within its design speed, load, temperature range and lubrication for
acceptance testing at the manufacturers facility.

Fig. 12.13 (from [12.35])


186 DYNAMICS OF MACHINERY

Figure 12.13 shows guidelines developed by the high speed gear industry
[12.35].
The following peak velocity limit values measured on the bearing caps of
gear boxes were recommended by Jackson [12.36]: smooth - 5 mm/s and less,
acceptable - 5 to 7 mm/s, marginal - 7 to 10 mm/s, planned shutdown repairs 10
to 15 mm/s, and immediate shutdown - 15 mm/s.

12.6 API Standards

The American Petroleum Institute (API) develops consensus standards for


the petroleum and petrochemical industry, covering topics that range from basic
design features of various machine components to conditions regarding critical
speeds, rotor balancing and vibration limits.
Table 12.10 shows vibration limits recommended by various API standards
[12.36].

Table 12.10

Standard Relative shaft displacement, Absolute bearing velocity,


mils peak-peak in/s

API 610 [12.37] 8000


0.12 r.m.s.
nmax

API 611 [12.38] (including runout)


API 612 [12.39]
API 616 [12.41] 12000 -
1.25
API 617 [12.42] nmax
API 672 [12.44]

12000 0.15 pk (10 Hz-2.5 kHz),


API 613 [12.40] 0.5
nmax 4g pk (2.5-10 kHz)

16000
API 619 [12.43] nmax -

API 673 [12.45] - 0.1 peak


12. VIBRATION LIMITS 187

12.7 Industrial buildings

Structures such as buildings, apartment blocks, and plants are subject


to vibrations generated by machinery, road traffic, underground trains, aircraft,
blasting operations, wind, forge hammers, pile drivers, and earthquakes.
Distinction should be made between high-intensity, short-duration
vibrations induced by earthquakes and blasting, and long-duration, usually
smaller-intensity vibrations induced by traffic, compressors, machinery, and other
human activities. While much information is available on the effects of
blasting vibrations from controlled tests on specific types of buildings (reinforced
or prestressed concrete, wood-framed, brick constructions), opinions on the effect
of intermittent or sustained vibrations produced by traffic and factory machinery
are controversial.
It is considered that building damages are not due directly to the effects
of vibrations alone. The risk of damage induced by low-level sustained vibrations
to usual buildings is very small, even when the level of vibrations is considered
intolerable by the occupants.
Buildings are more likely damaged by strong dynamic loading produced
by blasting, earthquakes, or other causes. During vibrations induced later by other
sources, if existing cracks are developed, the structural stiffness can vary in time
and eventually a resonance condition can be reached. This condition can cause the
vibration level to increase beyond safe limits. However, even in these cases,
experience gained in recent years has shown that the resistance to fatigue of steel or
reinforced concrete structures is sufficient to ensure that damage is unlikely to
occur if the level of vibrations can be tolerated by the occupants [12.46].
One cannot establish, with absolute certainty, what constitutes damaging
vibrations for a building. These damaging vibrations depend on building size, type,
and destination. In some norms (e.g.: [12.47]), the concept of damage is used to
define noticeable defects that decrease the buildings capacity to satisfy the
requirements imposed by its proposed use. Thus, for industrial buildings, damages
mean a decrease of either their safety state or the load carrying capacity of structural
elements. Nuclear power plants are not considered herein.
In all cases, damage does not refer to the building collapse or the fracture
of structural elements. From this point of view, the limit values of allowable
vibrations set up a large safety margin against yielding in the ordinary sense of the
word.
Maximum permissible steady-state vibrations have lower levels than
shock-induced short-duration vibrations. In the following, available vibration limits
are classified according to the quantity chosen as criterion in assessing the effect of
vibrations.
188 DYNAMICS OF MACHINERY

12.7.1 Vibration intensity

Based on a review carried out in 1961, information is presented in Fig.


12.14 for estimation of possible vibration-induced damage to buildings [12.46].

Fig. 12.14 (from [12.46])

The limit lines in the chart correspond to constant values of the quantity
3
x 0 f , where x 0 is the displacement amplitude, and f is the frequency of
vibration. This is related to the vibration intensity Z according to the relationship

Z=
a02
f
= 16 4 x02 f 3 [ mm s ]
2 3
(12.4)

where a0 is the amplitude of acceleration.

Taking Z s = 10 mm 2 s 3 as a reference value, the dimensionless


vibration intensity, measured in vibrar, is given by
12. VIBRATION LIMITS 189

S = 10 log
Z
Zs
(
= 22 log x02 f 3 ) [vibrar] (12.5)

Values of both x02 f 3 and S are given for the three zone boundaries in
Fig. 12.14. For comparison, lines of constant peak velocity are indicated on the
chart, as well as the danger limit, according to the 1939 release of DIN 4150
[12.47]. It was considered [12.46] that little risk of damage is probable for values
of x02 f 3 less than 50 mm 2 s 3 , a limit that corresponds to S = 37.37 vibrar .

Fig. 12.15 (from [12.48])

The same criterion was taken into account in [12.48], where a


classification of vibrations according to their effect on buildings is given. The
limit values of ranges are presented in Fig. 12.15, plotted in coordinates peak
displacement vs. frequency with solid lines.
190 DYNAMICS OF MACHINERY

For comparison, r.m.s. velocities of 2.5, 5 and 10 mm/s, as well as peak


velocities of 2.5, 5, 12 and 50 mm/s are also plotted with broken lines. Based on
our experience [12.49], the allowable limit of building vibrations is in the range of
30-40 vibrars and corresponds to 5 mm/s r.m.s. velocity at frequencies between 5
and 50 Hz.

12.7.2 Limits based on vibration velocity

Current opinion is that r.m.s. velocity is a more realistic criterion for


damage than the vibration intensity. From Figs. 12.14 and 12.15 one can see that
constant velocity lines have smaller slopes than constant vibration intensity lines.
Therefore, standards based on constant velocity will give increased weight to
lower frequency vibrations, which more likely can induce structural resonance
than frequencies above 50 Hz.
The German norm DIN 4150 [12.47] indicates that, at conventional
types of structures (industrial buildings and plants, public buildings, offices and
buildings of similar type and destination), damages were not observed for steady-
state horizontal vibrations with a peak velocity less than 5 mm/s. Beyond this
limit, damage occurrence depends on the specific conditions of the case under
investigation.
Practical experience has shown that damages of structural elements do
not occur for peak velocities up to 10 mm/s, even when all the strength capacity is
consumed by the static loading. Calculation of the additional dynamic stresses set
up by vibrations is recommended (if possible) when this limit is exceeded.

Table 12.11

Range Velocity, mm/s r.m.s. Effect


I Below 2.5 damages not possible
II 2.5 5.0 damages very improbable
III 5.0 10.0 damages not probable
IV Over 10.0 damages possible,
stress check necessary

The standard ISO 4866 [12.50] facilitates the rough evaluation of


stationary floor vibrations by measurement of the peak displacement and
frequency. However, vibration limits are expressed in terms of vibration severity
(Table 12.11). This is the maximum r.m.s. velocity, measured as the largest
orthogonal component of vibration, determined at prescribed measuring points on
the structure.
12. VIBRATION LIMITS 191

The safe limit appears to be somewhere below 10 mm/s r.m.s. velocity.


Constant r.m.s. velocity lines of 2.5, 5 and 10 mm/s are plotted in Fig. 12.15 with
broken lines.

Fig. 12.16 (from [12.49])

A simplified vibration severity chart (Fig. 12.16) was published in


[12.49], based on our experience that the upper limit for the range in which
damages from sustained steady-state vibrations are most probable can be chosen
at 7 mm/s r.m.s. velocity. However, as in all vibration standards, this must be
taken only as a guide in judging the vibration level and as a warning of impending
damage.
192 DYNAMICS OF MACHINERY

Annex A12.1

Blake Severity Chart [12.25]


12. VIBRATION LIMITS 193

Annex A12.2
Hydraulic Institute Application Standards B-74-1: 1967

Annex A12.3
In-service vibration severity criteria for centrifugal compressors
as a function of shaft speed Compressed Air and Gas Institute
194 DYNAMICS OF MACHINERY

Annex A12.4
12. VIBRATION LIMITS 195

Annex A12.5

ISO 10816-6, Vibration severity grade nomograph


196 DYNAMICS OF MACHINERY

Annex A12.6

Shaft displacement severity chart for industrial turbosets


VDI 2059 Part 3
12. VIBRATION LIMITS 197

Annex A12.7

Shaft displacement severity chart for gas turbosets


VDI 2059 Part 4
198 DYNAMICS OF MACHINERY

Annex A12.8

Shaft displacement severity chart for hydraulic machine sets


VDI 2059 Part 5
12. VIBRATION LIMITS 199

References

12.1. Maedel, P. H. Jr., Vibration standards and test codes, Shock and Vibration
Handbook, 5th ed., Harris C. ed., McGraw-Hill, 2001, p.19.1-19.11.
12.2. VDI 2063, Measurement and evaluation of mechanical vibrations of
reciprocating piston engines and piston compressors, Sept 1985.
12.3. ISO 10816-6, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 6: Reciprocating machines with power
ratings above 100 kW, 1995.
12.4. VDI 2056, Beurteilungsmastbe fr mechanische Schwingungen von
Maschinen, Okt 1964.
12.5. ISO 2372, Mechanical vibration of machines with operating speeds from 10
to 200 rev s - Basis for specifying evaluation standards, Nov 1974.
12.6. ISO 2373, Mechanical vibration of certain rotating electrical machines with
shaft heights between 80 and 400 mm Measurement and evaluation of the
vibration severity.
12.7. ISO 10816-1, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 1: General guidelines, 1995.
12.8. ISO 10816-2, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 2: Land-based steam turbines and
generators in excess of 50 MW with normal operating speeds of 1500 r min ,
1800 r min , 3000 r min and 3600 r min , 2001.
12.9. ISO 10816-3, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 3: Industrial machines with nominal
power above 15 kW and nominal speeds between 120 r min and 15000 r min
when measured in situ, 1998.
12.10. ISO 10816-4, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 4: Gas turbine driven sets excluding
aircraft derivatives, 1998.
12.11. ISO 10816-5, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 5: Machine sets in hydraulic power
generating and pumping plants, 2000.
12.12. ISO 10816-7, Mechanical vibration Evaluation of machine vibration by
measurements on non-rotating parts, Part 7: Rotordynamic pumps for industrial
application, 2004.
12.13. VDI 2059 - Part 1, Shaft vibrations of turbosets. Principles for measurement
and evaluation, Nov 1981.
200 DYNAMICS OF MACHINERY

12.14. VDI 2059 - Part 2, Shaft vibrations of steam turbosets for power stations,
March 1983.
12.15. VDI 2059 - Part 3, Shaft vibrations of industrial turbosets, Nov 1981.
12.16. VDI 2059 - Part 4, Shaft vibrations of gas turbosets, Nov 1981.
12.17. VDI 2059 - Part 5, Shaft vibrations of hydraulic machine sets, Oct 1982.
12.18. ISO 7919-1, Mechanical vibration of non-reciprocating machines
Measurement on rotating shafts and evaluation criteria, Part 1: General
guidelines, 1996.
12.19. ISO 7919-2, Mechanical vibration of non-reciprocating machines
Measurement on rotating shafts and evaluation criteria, Part 2: Land-based
steam turbines and generators in excess of 50 MW with normal operating
speeds of 1500 r min , 1800 r min , 3000 r min and 3600 r min , 2001.
12.20. ISO 7919-3, Mechanical vibration of non-reciprocating machines
Measurement on rotating shafts and evaluation criteria, Part 3: Coupled
industrial machines, 1996.
12.21. ISO 7919-4, Mechanical vibration of non-reciprocating machines
Measurement on rotating shafts and evaluation criteria, Part 4: Gas turbine
sets, 1996.
12.22. ISO 7919-5, Mechanical vibration of non-reciprocating machines
Measurement on rotating shafts and evaluation criteria, Part 5: Machine sets in
hydraulic power generating and pumping plants, 1997.
12.23. Rathbone, T. C., Vibration tolerance, Power Plant Engineering, Nov 1939,
p.721-724.
12.24. Reiher, H. and Meister, F. J., Die Empfindlichkeit der Menschen gegen
Erschtterungen, Forschung auf dem Gebiete des Ingenieurwesens, vol.2,
no.11, 1931, p.381-386.
12.25. Blake, M. P., New vibration standards for maintenance, Hydrocarbon
Processing and Petroleum Refinery, vol.43, no.1, Jan 1964, p.111-114.
12.26. * A practical guide to in-plane balancing, IRD Mechanalysis, Technical
Paper No. 116, 1981.
12.27. Federn, K., Erfahrungswerte, Richtlinien und Gtemastbe fr die
Beurteilung von Maschinenschwingungen, Konstruktion, vol.10, no.8, 1958,
p.289-298.
12.28. Beebe, R. S., Predictive Maintenance of Pumps Using Condition
Monitoring, Elsevier, Oxford, 2004, p.93.
12.29. Niemkiewicz, J., Standards for vibrations of machines and measurement
procedures, Encyclopedia of Vibration, Braun, S., Ewins, D. and Rao, S.S.,
eds., Academic Press, London, 2002, p.1224-1238.
12. VIBRATION LIMITS 201

12.30. Mechefske, C. K., Vibration standards and acceptance limits, Course


MECH 458, Part 5, Machine condition monitoring and fault diagnostics,
Queens University, Kingston, Canada, 2007.
12.31. * General guide lines for vibration on Clark centrifugal compressors,
Dresser Industries Inc., Clark Turbo Products Division, N.Y., 1971.
12.32. ISO 10817-1, Rotating shaft vibration measuring systems Part 1: Relative
and absolute sensing of radial vibration, 1998.
12.33. ISO 10816-0 (draft by H. Kanki), Guidelines for selecting vibration
evaluation methods, including shaft relative, shaft absolute, and pedestal
vibration measurements, 2003.
12.34. ISO 8579-2, Acceptance code for gears, Part 2: Determination of
mechanical vibration of gear units during acceptance testing, 1993.
12.35. AGMA 426.01, Specification for measurement of lateral vibration on high
speed helical & herringbone gear units, The American Gear Manufacturers
Association.
12.36. Jackson, Ch., Shop testing Is it worth it?, Orbit, vol.19, no.2, June 1998,
p.10..
12.37. ANSI/API Std 610, Centrifugal Pumps for Petroleum, Petrochemical and
Natural Gas Industries, 10th ed., Oct 2004.
12.38. ANSI/API 611, General-Purpose Steam Turbines for Petroleum, Chemical
and Gas Industry Services, Jan 1997.
12.39. ANSI/API 612, Petroleum, Petrochemical and Natural Gas Industries
Steam turbines Special-Purpose Applications, 6th ed., Nov 2005.
12.40. ANSI/API 613, Special Purpose Gear Units for Refinery Service, Chemical
and Gas Industry Services, Jun 1995.
12.41. API Std 616, Gas Turbines for Petroleum, Chemical and Gas Industry
Services, Aug 1998.
12.42. API Std 617, Axial and Centrifugal Compressors and Expander-
Compressors for Petroleum, Chemical and Gas Industry Services, 7th ed., July
2002.
12.43. API Std 619, Rotary-Type Positive-Displacement Compressors for
Petroleum, Petrochemical and Natural Gas Industries, 4th ed., Dec 2004.
12.44. API Std 672, Packaged, Integrally Geared Centrifugal Air Compressors for
Petroleum, Chemical and Gas Industry Services, March 2004.
12.45. API Std 673, Centrifugal Fans for Petroleum, Chemical and Gas Industry
Services, Jan 2002.
202 DYNAMICS OF MACHINERY

12.46. Steffens, R. J., Vibrations in buildings, Part I, Building Research Station


Digest, no.117, May 1970.
12.47. DIN 4150, Erschtterungen im Bauwesen, Teil 1, 2, 3, May 1986.
12.48. Koch, H. W., Ermittlung der Wirkung von Bauwerksschwingungen, VDI-
Zeitschrift, vol.95, 1953, p.733-737.
12.49. Rade, M., Vibration limits for industrial buildings, The Shock and
Vibration Digest, vol.26, no.3, May/June 1994, p.11-14.
12.50. ISO 4866, Mechanical vibration and shock. Vibration of buildings.
Guidelines for the measurement of vibrations and evaluation of their effects on
buildings, 2000.
13.
BALANCING OF ROTORS

Unbalance is the most common malfunction of rotating machines. It


produces specific once per rotation (synchronous) precession of rotors and lateral
vibrations of their supporting structures. Corrective balancing aims to reduce
machine unbalance-related vibrations.
Properly balanced rotors are important for the smooth running and
integrity of rotating machines. However during manufacturing irregularities are
produced by machining errors, cumulative assembly tolerances, distortions due to
heat treatment, flaws or inclusions in castings, and material non-homogeneity.
Irregularities occurring during operation include uneven wearing and erosion,
unsymmetrical buildup of deposits, missing or loose rotor parts, and load-related
and thermal distortion of the rotor. Because of these, the actual axis of rotation
does not coincide with one of the principal axes of inertia of the body, and variable
disturbing forces are produced which result in vibrations. In order to avoid/remove
these vibrations and ensure/establish proper operation, balancing becomes
necessary.
Balancing is the process of adding (or removing) mass on a rotor in order
to shift a central principal inertia axis to coincide with the geometric axis of
rotation. In a thin disc, the mass centre is moved towards the centre of rotation. In a
longer rotor, two or more planes along the axis of the shaft may be selected to
redistribute the rotor mass. An appropriate set of masses is applied (or removed) at
proper angular orientation and as close as possible to the proper plane along the
rotor. This gives a counter effect and balances the unit by making the free
centrifugal forces acting on the rotating body as small as possible.
Balancing is done by drilling, welding, sticking with adhesive, milling,
grinding or attaching screws. Typical items requiring balancing are electrical
armatures, spindles and tool-holders, crankshafts, ventilators, turbo-machinery,
pump impellers, drive assembly components, turbo charger rotors and car wheels.
The forces generated by an unbalance are proportional to the rotating speed
squared. Therefore, the balancing of high-speed equipment is especially important.
204 DYNAMICS OF MACHINERY

13.1 The mass unbalance

The unbalance in a machine may result from its design, from the
manufacturing process, from the assembly of multiple components or during
operation.

13.1.1 Definitions

A mass m rotating at a radius r with an angular speed gives rise to a


centrifugal force F = 2 m r . At a given value of the angular speed, the magnitude
and direction of the centrifugal force is determined by the product m r . In
balancing it is called unbalance [13.1].
The unbalance is defined by the vector quantity
u = mr , (13.1)

whose magnitude u = m r is measured in units of g mm (gram millimeter) or


kg m . Unbalance is independent of the rotational speed, hence implies that the
radius is constant.
In a rotor, the radius is measured with respect to the shaft axis, i.e. the
line connecting the centers of the bearing journals. It is the geometric axis of
rotation attached to the rotor.
In most rotors the condition of unbalance does not change noticeably up
to the operating speed. These rotors are referred to as rigid rotors. Flexible rotors
operating below one third of the first bending critical speed are considered as rigid.
For such rotors the unbalance may be specified as a fixed value, irrespective of the
rotational speed. They may be balanced at any desired speed up to the operating
speed.
Unbalance compensation is carried out by adding (or subtracting)
material to the unbalanced rotor. For a disc-shaped rigid rotor, this is done in such
a way that the sum of the centrifugal forces produced by the initial unbalance and
the correction mass becomes zero. The product of the correction mass with the
correction radius must be equal to the initial rotor unbalance. The added mass must
be placed in the opposite angular position, while the material removing is done in
the same angular position as the initial unbalance. The above considerations apply
only to discs mounted perpendicular to the rotation axis. Several correction masses
can be used (due to the particular form of the rotor) providing that the vector sum
of their individual unbalances counteracts the initial rotor unbalance.
As a rotor spins, the unbalanced mass tends to pull the rotor toward the
bearings on the side the unbalance is located. The point of maximum orbit radius,
13. BALANCING OF ROTORS 205

where the vibration sensed by a fixed non-contacting probe has maximum


amplitude, is called the high side or high spot of motion. The point located at 180 0
phase angle is named the low spot. At very low speeds, the high spot will be in
phase with the unbalanced (heavy) mass. The intersection of its radius with the disc
rim locates the heavy spot. In a flexible rotor, the phase of the measurable high spot
will lag the unknown position of the heavy spot. As speed increases, the high spot
of these rotors begins to lag the heavy spot. At the critical speed, the phase lag is
900 and well beyond the critical speed, the phase lag increases to 1800 (Fig. 2.11).
The high spot can be marked on a spinning disc by holding and gradually
moving a piece of chalk (or marking pen) close enough to cause a mark (streak) on
the disc rim. It can be located with a dial indicator and a stroboscope, or with
proximity transducers and a strip chart recorder or oscilloscope. On small machines
a seismic velocity pickup and a vibration analyzer can also be used. The high spot
is the part of the shaft that would first contact a closely fitting labyrinth seal and
would leave a mark on a shaft at this location. A reference mark called key phasor
is also prescribed on the disc to define the reference with respect to which phase
angles are measured.
For a rigid rotor having significant length, there are four types of rotor
unbalance i.e. static, couple, quasi-static and dynamic [13.2], [13.3].

13.1.2 Static unbalance

Static unbalance is the simplest form of unbalance in which the central


principal inertia axis is displaced parallel to the shaft axis (Fig. 13.1, a).
The problem can be solved with a single correction mass placed in the
same plane as the rotor center of mass. If it is not possible to make a mass
correction in the center portion of the rotor, then equal corrections can be made in-
line at opposite ends of the rotor.
Static unbalance can be usually identified by comparing the magnitude
and phase readings obtained at the support bearings. For rotors supported between
bearings, static unbalance will result in nearly identical magnitude and phase
readings. This does not apply for overhung rotors.

13.1.3 Couple unbalance

In the case of couple unbalance, the shaft axis intersects the central
principal inertia axis at the rotor center of mass.
Couple unbalance is a condition created by a heavy spot at each end of a
rotor but on opposite sides of the centerline, as shown in Fig. 13.1, b. Sometimes it
206 DYNAMICS OF MACHINERY

is referred to as pure dynamic unbalance or swash unbalance. Unlike the static


unbalance, couple unbalance becomes apparent only when the part is rotated and
can be identified by comparing the magnitude and phase readings at the rotor
support bearings. Rotors supported between bearings will typically reveal equal
amplitudes of vibration but phase readings will differ by 1800 . This method of
detecting couple unbalance does not apply to overhung rotors.
In a rotor with overhung discs at both ends, the distance between bearings
is smaller than the distance between the correction planes so that couple unbalance
is of primary importance as compared with the static unbalance.
Unfortunately, only a few balance problems will be pure static or pure
couple. Most balance problems will be a combination of static and couple
unbalance, further classified as quasi-static and dynamic unbalance.

a b

c d
Fig. 13.1

13.1.4 Quasi-static unbalance

In the case of quasi-static unbalance, the central principal inertia axis


intersects the rotating centerline but not at the rotor center of mass.
This type of unbalance can be produced if the rotor were out of balance at
only one end, as shown in Fig. 13.1, c. Installing an unbalanced pulley or coupling,
or reblading only the first stage of a turbine or compressor might cause it. It can be
also thought as a combination of static and couple unbalance where the static
unbalance is directly in line with one of the couple components.
13. BALANCING OF ROTORS 207

Quasi-static unbalance can often be compensated to a satisfactory level


by carrying out a simple single-plane balancing to the end having the higher
vibration level, and by making the mass corrections in a nearby reference plane not
coinciding with the plane through the rotor center of mass.

13.1.5 Dynamic unbalance

In the case of dynamic unbalance, the central principal inertia axis is


inclined with respect to the shaft axis but does not intersect it. It is the most
common type of unbalance for long rotors and can be visualized as a combination
of static and couple unbalance, where the static component is not in line with one
of the couple components. As a result, the central principal axis is both tilted and
displaced from the shaft axis, as shown in Fig. 13.1, d.
Dynamic unbalance problems can only be solved by making mass
corrections in a minimum of two different reference planes. One therefore
frequently refers to dynamic balancing as two plane balancing and to static
balancing as single plane balancing.

13.1.6 Static vs. dynamic balancing

Disc-shaped bodies may be balanced statically as long as the distance


between bearings is reasonably large and the disc has been attached normal to the
shaft axis (no axial run-out). Assembly faults, axial run-out in the bearing seating, a
throw in the ball bearings or internal stresses can produce simultaneous couple
unbalance. Grinding wheels and fan rotors are statically balanced. Turbine wheels
are only statically prebalanced and then the completely assembled rotor is
dynamically balanced. The static part of unbalance may be measured, without
rotation, in the earth gravitational field by rolling on knife edges, suspending as
pendulum or weighing [13.4].
Properly balanced car tires are important for driving comfort and long tire
life. Out-of-balance tires will cause a car to vibrate at speeds between 80 and 120
km/h. Vibrations in the seat or floorboard are produced by static unbalance.
Vibrations in the steering wheel are produced by the wheel wobbling due to
dynamic unbalance.
When mounting a car wheel, the tire light spot, marked by a yellow dot
on the tire sidewall, should be lined up with the rim high spot at the valve stem.
Tires should be balanced when they are mounted on wheels for the first time or
when they are remounted after repair. In the latter case, mud or dirt packed in the
back of the rim or debris embedded in the thread must be first removed. It is also
important to avoid using locking lug nuts heavier than the initial wheel lugs.
208 DYNAMICS OF MACHINERY

13.2 Single plane balancing

Single plane balancing techniques typically used for in-place balancing


are presented in the following. They require the addition of a test trial mass to
produce a change in the original unbalance which can be used to determine the
required balance correction. Common practice is to use a mass which will produce
an unbalance force at the support bearing equal to 10% of the rotor weight
supported by the bearing.

13.2.1 Vector balancing

The diagram in Fig. 13.2 illustrates a vector solution [13.5] for single
plane balancing of a thin disc.

Fig. 13.2

A phase reference is marked on the disc. Then the disc is spun preferably
at the operating speed and the high side is marked on it. The measured response is
expressed by the complex displacement z0 whose magnitude is equal to the
measured amplitude and the phase is equal to the angle between the high side and
the reference timing mark.

Next, a trial mass mt is placed in any position (at 2700 in Fig. 13.2) and,
with the disc running at the same speed as before, the new high side is marked and
the response amplitude is measured. The response is expressed by the complex
13. BALANCING OF ROTORS 209

displacement z1 . It represents the effect of both the original unbalance and the
added trial mass. The vector difference z is then the effect of mt alone.

If mt is rotated counterclockwise by an angle , its response vector


z will rotate by and will be parallel and opposite to z0 . If it is increased in
the ratio z0 z to equal the original unbalance, it will balance the disc. Hence the
correction mass (when mt is removed) is mc = mt z0 z .
This procedure is a variant of the more general influence coefficient
method [13.6] presented in the following.

13.2.2 Influence coefficient method

Rotating unbalanced masses give rise to centrifugal forces proportional to


the radial unbalance vector and the angular velocity squared. In linear systems,
rotor lateral displacements may be expressed as sums of products of a flexibility
coefficient and a centrifugal force, or alternatively in terms of products of speed-
dependent influence coefficients and radial unbalances. Herein, the influence
coefficient represents the response of the rotor to a unit unbalance.
The rotor is spun at constant rotational speed. In a rigid rotor, the heavy
spot lies in the same radial plane as the high side. The response of the unbalanced
rotor is expressed by a complex displacement

z0 = z0 0 = z0 R + i z0 I = z0ei 0 , (13.2)

where z0 is the displacement amplitude measured with a non-contacting probe and


0 is the phase angle between the high spot and a reference timing mark.

This response is produced by a centrifugal force u 2 , where the vector


unbalance u = m r is expressed as the product of the rotor mass m and the
unbalance eccentricity r , of magnitude r and phase lag angle m . This angle
shows the radial plane where the unbalance u is located with respect to the timing
reference mark on the shaft, as measured in a direction opposite to rotation.
The rotor synchronous displacement response may be expressed in terms
of a complex influence coefficient a multiplied by the rotor unbalance u
z0 = a u . (13.3)
It is assumed that the influence coefficient a is a function of rotor speed
only. This implies that, if a small correction mass is placed on the rotor, the
influence coefficient at a particular speed will not change (which is not true in non-
linear systems).
210 DYNAMICS OF MACHINERY

A trial mass mt is placed on the rotor at a given radius rt and at a


(sometimes unknown) phase angle t relative to the reference timing mark. The
trial complex unbalance is
i t
ut = mt rt e . (13.4)
The rotor response, measured at the same speed as the initial response, is

z1 = a ( u + ut ) , (13.5)
or
i 1
z1 = z11 = z1R + i z1I = z1e . (13.6)

By subtracting the initial response, the difference vector is obtained as

z = z1 z0 = z ei . (13.7)

The initial unbalance vector is


z0 z0 z
u= = ut = 0 ei ( 0 ) ut . (13.8)
a z z
If the trial mass is removed, it is desired to find the right location where
to place a compensating mass mc in order to balance the rotor. The placement of
the new mass creates a z vector which is equal to the original z0 vector, but acts
in the opposite direction. This is achieved if the balance correction uc is equal and
opposite to the rotor initial unbalance vector u
z0 i ( 0 ) i t
uc = u = u e i = e u t e = u c e i c , (13.9)
z
where
z0 z
uc = ut = 0 mt rt = mc rt , (13.10)
z z
and
c t = = 0 1800 . (13.11)
The compensation mass
z0
mc = m, (13.12)
z t
may be placed at the same radius rt as the trial mass mt , but should be shifted by
an angle = 0 1800 in the positive direction (if t is not known).
13. BALANCING OF ROTORS 211

If the trial mass is left in the rotor (when it is welded onto it), the
compensating mass produces a trim balance correction. The trim balance
correction vector is given by
z
utrim = ( u + ut ) = 1 . (13.13)
a
If the rotor response z1 is much larger than the original response
amplitude, not only should the trial mass be removed, but also the trial run to get
z1 should be repeated with either a reduced trial mass or a change in its angular
position.

Example 13.1
A rotor of 450 kg has a steady-state response amplitude of 75 m at a
phase angle of 2700 . A trial mass of 5 g is placed at a radius of 225 mm at a
position of 30 0 from the timing mark against rotor rotation. The steady-state
response with the trial mass on the rotor has an amplitude of 50 m at 170 0 .
Determine the correction balance [13.7].
Solution
The initial response is (Fig. 13.3)

z0 = 752700 = 0 i 75 .
The amplitude of the trial unbalance is
ut = mt rt = 5 225 = 1125 g mm .
The trial unbalance vector is

ut = 1125300 = 974.3 + i 562.5 .


The response after the trial mass is applied is

z1 = 501700 = 49.2404 i 8.6824 .


The difference vector

z = z1 z0 = 49.2404 + i 83.6824 = 97.0964120.50 .


The influence coefficient
z
a= = ( 0.7131 + i 86.3034 ) 10 3 = 86.3063 10 3 90.50 .
ut
The original unbalance
212 DYNAMICS OF MACHINERY

z0
u= = ( 0.8690 + i 0.0072 ) 103 = 869179.50 .
a
The correction unbalance

uc = u = 869 i 7.2 = 869359.50


where
uc = mc rt = 3.8622 225 = 869 g mm .

Fig. 13.3

Thus, a compensation mass


z0 75
mc = mt = 5 = 3.8622 g
z 97.0946
should be placed at the radius rt = 225 mm shifted by an angle

= 0 1800 = 270 120.5 180 = 30.50

in the positive direction, i.e. 30.50 anticlockwise.


If the trial mass is left in the rotor, the trim balance correction vector is
13. BALANCING OF ROTORS 213

utrim = ( u + ut ) = 105.3 i 569.7 = 579.3259.50


so that a mass
mtrim = 2.5747 g

should be placed at a radius rt = 225 mm and an angle of 259.50 .

Example 13.2
A rotor has a steady-state response amplitude of 3.4 mm s at a phase
angle of 1160 . A trial mass of 2 g is fixed to the rotor at the same angular position
as the reference mark. With the trial mass, the vibration velocity level is 1.8 mm s
at a phase angle of 420 . Determine the position and the magnitude of the
compensating mass necessary to balance the rotor [13.8].

Fig. 13.4
Solution
The initial response is (Fig. 13.4)

z0 = 3.41160 = 1.4905 + i 3.0559 .


214 DYNAMICS OF MACHINERY

The response after the trial mass is applied is

z1 = 1.8420 = 1.3377 + i 1.2044 .


The difference vector

z = z1 z0 = 2.8281 i 1.8515 = 3.3805327 0 .

A compensation mass
z0 3 .4
mc = mt = 2 = 2.0117 g
z 3.3805
should be placed at an angle

= 0 1800 = 116 327 180 = 310


in the positive direction, i.e. 310 anticlockwise.

Example 13.3
To balance the rotor statically, a machine was run up to its operating
speed and a vibration velocity level of 15 mm s was measured at a phase angle of
550 , after which the machine was stopped. A trial mass of 5 g was fixed to the
rotor at the same angular position as the reference mark. Then the machine was run
up to its operating speed again. The new vibration velocity level was 18 mm s at a
phase angle of 170 0 . Determine the position and the magnitude of the
compensating mass necessary to balance the rotor [13.8].

Solution
The initial response is (Fig. 13.5)

z0 = 15550 = 8.6036 + i 12.2873 .


The response after the trial mass is applied is

z1 = 181700 = 17.7265 + i 3.1257 .


The difference vector

z = z1 z0 = 26.33028281 i 9.1616 = 27.871990 .


A compensation mass
z0 15
mc = mt = 5 = 2.69 g
z 27.87
13. BALANCING OF ROTORS 215

should be placed at an angle

= 0 1800 = 55 199 180 = 3240

or 360 in the positive direction.

Fig. 13.5

13.2.3 Three-trial-mass method

Sometimes it is not possible or practical to obtain phase readings. In such


cases, balancing can be carried out using only amplitude measurements, using a
simple vibration meter connected to an accelerometer mounted on the bearing
[13.2]. The procedure requires one run to obtain the original unbalance amplitude
and three trial runs. On each trial run, a single trial mass is attached at a different
angular position on the rotor. Usually, the same mass is placed at 00 , 1200 , 2400
(or at 00 , 900 , 1800 ) on the rotor, at the same radius.
Geometrically, Sieberts construction (Fig. 13.6) can be used to evaluate
the amount and angular location of the required balance correction. A circle with
the center in O is drawn having the radius proportional to the amplitude of response
to the original unbalance. The relative angular positions A, B, C of the trial mass
mt are marked on the original circle. Using the trial mass positions as the centers,
circles are drawn (at the same scale) having the radius equal to the corresponding
216 DYNAMICS OF MACHINERY

trial run amplitude AT, BT, CT. The three trial run circles intersect at point T. The
line OT is drawn.
The correction mass is

mc = mt OT OA (13.14)

and its location is determined by the angle between the vector OT and the vector
opposite to OA, i.e. OC in this case. It must be placed to have the same phase lag
with respect to the first trial mass position.

Example 13.4
Sieberts construction from Fig. 13.6 is obtained using the following
measurement data. Original response amplitude OA = OB = OC = 20 m .
AT = 30 m - test run with trial mass at 00 . BT = 15 m - test run with trial mass
at 900 . CT = 43 m - test run with trial mass at 1800 .

Fig. 13.6

The result is OT = 31.2 m .


The compensating mass is

OA 20
mc = mt = mt = 0.625 mt
OT 32

and must be placed at 22.30 clockwise with respect to the second trial mass
position.
13. BALANCING OF ROTORS 217

13.3 Two-plane balancing

There are many possible ways to solve two-plane dynamic unbalance


problems in-place [13.9], [13.10], including: a) separate single-plane solutions; b)
simultaneous single-plane solutions; c) the influence coefficient method; and d)
decomposition into static plus couple unbalance. Only the last two methods are
presented in the following [13.6].

13.3.1 Influence coefficient method

First, the original unbalance readings z10 and z 20 are recorded for the two
bearings of the machine. It is assumed that they can be expressed in terms of the
unknown (required) values of unbalance u1 and u2 as
z10 = a11 u 1 + a12 u 2 ,
(13.15)
z 20 = a 21 u 1 + a 22 u 2 ,

where a i j (i , j = 1,2) are complex influence coefficients. To determine these


influence coefficients, a trial unbalance is placed in each plane and the new
resulting amplitudes of motion are measured.
Next, a trial mass producing an unbalance u t1 is added to the first
correction plane and the resultant readings z11 and z 21 at both bearings are
recorded. They can be expressed using the influence coefficients as
( )
z11 = a11 u 1 + u t1 + a12 u 2 ,
z21 = a 21 ( u 1 + u t1 ) + a 22 u 2 .
(13.16)

Subtracting equations (13.15) from (13.16), the influence coefficients a11


and a21 may be determined as

z11 z10 z 21 z 20
a11 = , a 21 = . (13.17)
u t1 u t1

Finally, the trial mass is removed from the first correction plane and a trial
mass is added to the second correction plane. The resultant readings at both
bearings are again recorded.
If the first trial mass is assumed to be left in place, and a second trial mass
producing an unbalance u t 2 is added to the second correction plane, the resultant
readings z12 and z 22 are recorded at both bearings. They can be expressed using
the influence coefficients as
218 DYNAMICS OF MACHINERY

( ) (
z12 = a11 u 1 + u t1 + a12 u 2 + u t 2 , )
z 22 = a 21 ( u 1 + u t1 ) + a 22 (u 2 + u t 2 ) .
(13.18)

Subtracting equations (13.16) from (13.18), the influence coefficients a12


and a22 may be determined as

z12 z11 z 22 z 21
a12 = , a 22 = . (13.19)
ut2 u t2

If the first trial mass is removed, then z11 and z 21 should be replaced by
the original vibration readings z10 and z20 .
To balance the rotor, correction masses should be placed in planes 1 and 2
to generate vibrations equal in magnitude but opposite in direction to z10 and z 20 .

The balance corrections uc1 and uc 2 are given by


1
uc1 u10 + u t1 a11 a12 z12
= = . (13.20)
u c 2 u 20 + u t 2 a21 a22 z 22
Solving for the final balance correction gives
a12 z 22 a22 z 12 a 21 z 12 a11 z 22
u c1 = , u c2 = . (13.21)
a11 a 22 a21 a12 a11 a 22 a21 a12

The values of uc1 and uc 2 represent the trim balance corrections required
if both trial masses are left in place.
If the computations are carried out using the original vibration readings
z10 and z20 , then the computed balances will correspond to the total original
balance correction required in the rotor
1
z11 z10 z12 z10
uc1 u u t 2 z10
= t1 . (13.22)
u
c2 z
21 z 20 z 22 z 20 z 20
u u t 2
t1

Example 13.5
For a machine with a rigid rotor supported in two bearings, vibration
velocity levels and phase angles have been measured as shown in Table 13.1. The
trial mass mt = 2.5 g was mounted on the rotor in turn, in the bearing plane 1 and
13. BALANCING OF ROTORS 219

bearing plane 2, at the same radius and angular position. Calculate the correction
masses and their positions [13.8].
Table 13.1
Vibration readings
Trial mass
Plane 1 Plane 2
None 7.2 mm/s 238 0 z10 13.5 mm/s 2960 z 20
2.5 g in Plane 1 4.9 mm/s 114 0 z11 9.2 mm/s 347 0 z 21
2.5 g in Plane 2 4.0 mm/s 79 0 z12 12.0 mm/s 2920 z22

Solution
The initial responses (Fig. 13.7) are

z10 = 7.22380 = 3.8154 i 6.1059 ,

z20 = 13.52960 = 5.9180 i 12.1337 .

Fig. 13.7
220 DYNAMICS OF MACHINERY

Responses with the trial mass applied in plane 1 are

z11 = 4.91140 = 1.9930 + i 4.4764 ,

z 21 = 9.2347 0 = 8.9642 i 2.0695 .


Responses with the trial mass applied in plane 2 are

z12 = 4.0790 = 0.7632 + i 3.9265 ,

z 22 = 12.02920 = 4.4953 i 11.1262 .


The difference vectors are

z11 z10 = 1.8224 + i 10.5823 = 10.738180.20 ,

z12 z10 = 4.5787 + i 10.0325 = 11.027965.50 ,

z 21 z 20 = 3.0462 + i 10.0642 = 10.515173.10 ,

z22 z20 = 1.4227 + i 1.0075 = 1.7433144.7 0 .


The correction unbalances are

u c1 = 2.5 (0.7559 + i 0.9069 ) = 2.9550.20 ,

u c 2 = 2.5 (0.1606 i 1.1263) = 2.84 81.90 .

The correction masses are

in plane 1: 2.95 g at 50.20 clockwise (opposite the direction of rotation),

in plane 2: 2.84 g at 81.90 counterclockwise (in the direction of rotation).

Example 13.6
The initial vibration readings on a rigid-body rotor are

z10 = 8.6 m630 , z 20 = 6.5 m2060 .

A trial balance mass of u t1 = 10 g is placed at a relative phase angle of


2700 at plane 1. The resulting vibrations at planes 1 and 2 are

z11 = 5.9 m1230 , z21 = 4.5 m2280 .


13. BALANCING OF ROTORS 221

The first trial mass is removed, and a trial balance mass of u t 2 = 12 g is


placed at a relative phase angle of 180 0 at plane 2. The resulting vibration readings
at planes 1 and 2 are

z12 = 6.2 m360 , z22 = 10.4 m1620 .


Determine the balance corrections [13.11].

Solution
The initial responses are
z10 = 3.9043 + i 7.6627 , z20 = 5.8422 i 2.8494 .
The vibration readings at planes 1 and 2, due to the placement at plane 1
of the trial mass (at a unit radius) of unbalance u t1 = 0 i 10 , are

z11 = 3.2134 + i 4.9482 , z 21 = 3.0111 i 3.3442 .


The vibration readings at planes 1 and 2, due to the placement at plane 2
of the trial mass (at a unit radius) of unbalance u t 2 = 12 + i 0 , are

z12 = 5.0159 + i 3.6443 , z22 = 9.8910 + i 3.2138 .


The influence coefficients (13.17) and (13.19) are

a11 = 0.2714 i 0.7118 = 0.7618290.90 ,

a21 = 0.0495 + i 0.2831 = 0.287480.10 ,

a12 = 0.0926 + i 0.3349 = 0.3474105.50 ,

a22 = 0.3374 i 0.5053 = 0.6076303.7 0 .


The correction unbalances (13.22) are

u c1 = 8.9873 i 5.9238 = 10.7640 33.40 ,

u c 2 = 2.5819 + i 5.6392 = 6.202265.40 .

The correction masses are

in plane 1: 10.8 g at 33.40 counterclockwise (in the direction of rotation),

in plane 2: 6.2 g at 65.40 . clockwise (in the opposite direction).


222 DYNAMICS OF MACHINERY

Example 13.7
The run-out readings taken on a rotor during coast down at a very low
speed are 0.5 m2720 and 0.4 m1230 at probe 1 and 2, respectively. They are
not produced by the unbalance, therefore should be subtracted in the balancing
calculation.
The initial vibration readings at design speed are

z10 = 1.8 m1480 , z20 = 3.6 m1150 .


A trial balance mass of u t1 = 4.9 g is placed at a relative phase angle of
1200 at plane 1. The resulting vibrations at planes 1 and 2 are

z11 = 1.1m1780 , z21 = 2.0 m 980 .

The trial mass is removed and placed at a relative phase angle of 2200 at
plane 2. The resulting vibration readings at planes 1 and 2 are

z12 = 2.1m 980 , z 22 = 3.7 m1020 .


Determine the balance corrections [13.11].

Solution
The initial responses are

z10 = 1.81480 0.52720 = 2.1205136.7 0 = 1.5440 + i 1.4535 ,

z 20 = 3.61150 0.41230 = 3.20441140 = 1.3033 + i 2.9273 .


The vibration readings at planes 1 and 2 due to the placement of the trial
mass (at a unit radius) of unbalance u t1 = 2.4500 + i 4.2435 at plane 1 are

z11 = 1.11780 0.52720 = 1.2396154.30 = 1.1167 + i 0.5382 ,

z 21 = 2.0980 0.41230 = 1.646292.10 = 0.0606 + i 1.6451 .


The vibration readings at planes 1 and 2 due to the placement of the trial
mass (at a unit radius) of unbalance u t 2 = 3.7536 i 3.1497 at plane 2 are

z12 = 2.1 980 0.52720 = 2.5978 96.9 0 = 0.3098 + i 2.5792 ,

z22 = 3.71020 0.41230 = 3.3296 99.50 = 0.5513 + i 3.2837 .


The influence coefficients (13.17) and (13.19) are
13. BALANCING OF ROTORS 223

a11 = 0.2054 + i 0.0179 = 0.20611750 ,

a21 = 0.3534 i 0.0888 = 0.3644 165.90 ,

a12 = 0.3406 i 0.0141 = 0.3409 177.60 ,

a22 = 0.1643 + i 0.0429 = 0.1698165.30 .


The correction unbalances (13.22) are

u c1 = 0.0568 + i 7.4599 = 7.4888850 ,

u c 2 = 5.3195 + i 0.0240 = 5.3196179.7 0 .

The correction masses are 7.5 g at 850 in plane 1 and 5.3 g at 179.7 0 in
plane 2 clockwise.
If the second trial mass is left in the rotor, the trim balance is

u trim 2 = u c 2 u t 2 = 3.54116.30 .

The trim balance mass in plane 2 is 3.54 g and should be placed at


116.30 clockwise.

13.3.2 Resolution into static and couple unbalance

For reasonably symmetrical rotors mounted between bearings, the


unbalances at two arbitrary planes, u1 and u 2 , may be vectorially resolved into a
static unbalance u s and a couple unbalance ud [13.11]
u1 = u s ud ,
(13.23)
u2 = u s + ud .

Solving for u s and ud we obtain

u1 + u 2 u u
us = , ud = 2 1 . (13.24)
2 2
The u s components at the two planes are acting in the same radial
direction and generate a centrifugal force at the center of mass. They are equivalent
to a 2 u s unbalance applied at the rotor center of mass. The ud components are
acting 1800 out of phase to each other and create a couple which is a free vector.
224 DYNAMICS OF MACHINERY

The static and couple corrections can be carried out simultaneously, but they are
independent of each other.
The corresponding displacements, z1 and z 2 , can be resolved into the
static (in-phase) component and couple (out-of-phase) components
z1 + z 2 z 2 z1
zs = , zc = (13.25)
2 2
where z s defines the cylindrical mode and zc define the conical mode.

Example 13.8
The original readings on a two-disc rotor are 8 m1300 and 6 m300
at probe 1 and 2, respectively (Fig. 13.8). Determine the response to the original
static and couple unbalance [13.2].
Solution
The vibration readings in the two planes are
z1 = 5.1423 + i 6.1284 , z2 = 5.1622 + i 3.0000 .

Fig. 13.8

The response to static unbalance is


z1 + z 2
zs = = 0.0269 + i 4.5642 = 4.5689.60 .
2
The response to couple unbalance is
z2 z1
zc 2 = = 5.1692 i 1.5642 = 5.4343.20 ,
2

zc1 = 5.4163.20 = 5.1692 + i 1.5642.


13. BALANCING OF ROTORS 225

13.4 Unbalance tolerances

Even after balancing, any rotor will possess a certain residual unbalance.
Permissible residual unbalances are recommended in the standard ISO 1940/1
[13.12]. The standard includes a tentative classification of various types of
representative rotors. For each rotor group, a range of recommended balance
quality grades is given, relating the permissible residual unbalance to the maximum
service speed.

13.4.1 Permissible residual unbalance

In general, the larger the rotor mass m, the greater the permissible
unbalance u per . The specific unbalance is defined as
u per
e per = . (13.26)
m
It corresponds to the mass eccentricity if the residual unbalance is a static
unbalance.

13.4.2 Balance quality grades

Practical experience has shown that for similar rotors, the specific
permissible residual unbalance e per is inversely proportional to the rotor angular
speed = (using the notation from ISO 1940/1)
e per = constant . (13.27)

The product e is the tangential velocity of the centre of mass. In


geometrically similar rotors, with the same peripheral velocity, the same stresses in
the rotor and the same specific bearing loads are produced if the value e per is
kept constant (assuming rigid bearings). The balance quality grades G are based on
this relationship.
The G number is the product of the specific unbalance and the angular
velocity of the rotor at maximum operating speed. It is constant for rotors of the
same type
G = e per = constant . (13.28)
Balance quality grades are separated by a factor of 2.5. However, G
numbers of intermediate value may be used to satisfy special requirements. For
example, a standard pump impeller has a recommended balance quality grade of
226 DYNAMICS OF MACHINERY

6.3. Special conditions may require a better balance quality of G 4.0 to meet the
installation requirement in an area with low structure-borne noise limits [13.13].
The balance quality grades are designated according to the upper limit of
the product e given in millimeters per second (for measured in radians per
second). Plotted against the maximum operating speed, n, the upper limits of e per
are shown in Fig. 13.9.

Example 13.9
How large is the permissible specific residual unbalance e per in a rotor of
the balance quality grade G 6.3 for a service speed n = 3000 rpm ? Determine the
permissible residual unbalances in each of the correction planes if the rotor is
symmetrical and has 40 kg .
Solution
For n = 3000 rpm on the horizontal axis in Fig. 13.9, moving vertically to
the line G 6.3 , then horizontally to the left to the e per axis we obtain e per 20 m
(or 20 g mm kg ). This value can also be calculated. If G 6.3 means that the
permissible tangential velocity of the center of mass is 6.3 mm s , then
v per 6.3 6.3
e per = = = 0.02 mm = 20 m .
3000 314
30
For a rotor of mass m = 40 kg , the total permissible residual unbalance is

u per = e per m = 20 40 = 800 g mm ,

hence 400 g mm in each correction plane.

13.4.3 Classification of rigid rotors

In Table 13.2 [13.12] the most common types of rigid rotor are listed in
groups with the same balance quality grade. The classification is only a
recommendation based on current experience and should be adhered to with care.
For a turbine rotor, a preliminary G-value is selected from Table 13.2 for
the specific application. Then this value is increased up to the next quality grade, as
a result of the unbalance produced by the installation of the coupling, bearing
configuration, salt deposits, corrosion of shaft components, cavitation and thermal
bending. The total permissible unbalance is calculated as shown above.
13. BALANCING OF ROTORS 227

Fig. 13.9 (from [13.12])


228 DYNAMICS OF MACHINERY

Table 13.2
13. BALANCING OF ROTORS 229

The result can be used even for flexible rotors. Using a finite element
model of the rotor, the first modes of bending vibration are calculated, usually all
modes below the machine trip speed and the mode just above the trip speed. Then,
the worst unbalance distribution for each mode is considered, dividing the total
unbalance into suitable individual unbalance components located so that to produce
a maximum response in the respective mode of vibration. The calculated
amplitudes of the unbalance response are then compared to limit values given by
guidelines and standards.

13.5 Multiplane flexible rotor balancing

Flexible rotors running at speeds far below the first critical can be
considered as not being deformed by unbalances and the motion in the two rigid-
body modes of precession can be cancelled by balancing in two planes. For speeds
higher than about half of the first critical, unbalances bend the rotor setting up new
centrifugal forces in addition to the ones balanced by two plane corrections.
The influence coefficient method can be extended to multimass flexible
rotors [13.14]. The aim is to determine those correction masses in a predetermined
set of planes which will minimize measured vibration readings at a series of
sensors and speeds, as predicted by the influence coefficients, relating vibration
readings to mass additions. The influence coefficients are determined
experimentally by applying trial masses to the rotor at one location at a time and
measuring the rotor response at each station where balance masses are to be placed.
The modal method [13.15] aims to balance the rotors, one mode at a time,
by placing proper masses at the antinodes. The set of masses is specifically selected
to leave the already balanced lower modes undisturbed. A unified balancing
approach procedure has been developed [13.16]. It involves the calculation of
modal trial mass sets, employing data derived as in the influence coefficient
method. In general, the number of planes required for the modal trial mass set is
one more than the number of modes which must not be affected.
There are two schools of thought regarding the number of balancing planes
needed at speeds comparable with the critical speeds. One [13.17] is satisfied with
N planes when the Nth critical speed is reached, the other [13.18] stipulates N + 2
planes.

13.5.1 Balancing in N + 2 planes

A rotor consisting of a massless straight shaft supported in B bearings and


carrying N concentrated masses can be balanced perfectly by placing small
correction masses in N + B planes along its length [13.19].
230 DYNAMICS OF MACHINERY

Consider a rotor with a single major mass M (Fig. 13.10) and with an
arbitrary unbalance
uk = mk ek . (13.29)
The displacement of the major mass is
w1 = 11F1 + 1k Fk (13.30)

where 11 is the deflection at station 1 due to a unit force at 1, 1k is the


deflection at station 1 due to a unit force at station k, Fk is the sum of external and
inertia forces acting at station k and F1 = M 2 w1 . For synchronous motion,
neglecting external damping and other external forces at the major mass station

Fk = mk (wk + ek ) 2 = 2 (uk + mk wk ) . (13.31)


The deflection at the major mass is

w1 = 2 [ M w111 + (uk + mk wk )1k ] . (13.32)

It is assumed that the balance correction mass mk is small in comparison


to the major mass M.

Fig. 13.10 (from [13.20])

Solving for the deflection at the major mass station for a series of
unbalances uk

w1 = 1k uk 2 . (13.33)
1 2 M 11
The balancing requirement that ensures zero displacement at the major
mass station is
1k uk = 0 . (13.34)

The bearing force reactions are given by

Fb1 + Fb 2 = 2 [ M w1 + mk (wk + ek )] . (13.35)


13. BALANCING OF ROTORS 231

Neglecting wk compared to ek , the vanishing of the bearing forces


requires that

2 (M w1 + mk ek ) = 0 . (13.36)

If the balance criterion of equation (13.34) is met, then (13.36) becomes

uk = 0 . (13.37)

This is recognized as the first requirement for rigid body balancing.


The third balancing requirement is obtained by summing moments about
the first bearing

Fb1 l = 2 [ M w1l 1 + mk l k (wk + ek )] = 0 . (13.38)

This reduces to
l k uk = 0 . (13.39)

This is recognized as the second requirement for rigid body balancing.


In summary, the requirements for flexible rotor balancing may be stated as
two equations of rigid body balance (13.37) and (13.39), plus a flexible rotor
balance requirement (13.34).
In the N plane method, only one balance correction mass is needed to
reduce the amplitude at the major mass station or shaft antinode to zero. The
balance correction ub1 placed at the major mass station is

1
ub1 =
11
1k uk . (13.40)

Although the amplitude at the major mass station has been reduced to zero,
the transmitted bearing forces are nonvanishing. In order to eliminate the
transmitted bearing forces due to unbalance, as well as reduce the rotor amplitude
of motion while passing through the first critical speed, two additional balance
planes are required. Let ub 2 and ub3 be two additional balance corrections by
masses placed on the rotor. The balance corrections are given by

ub1 + ub 2 + ub3 = u k ,

l 1ub1 + l 2ub 2 + l 3ub3 = l k uk , (13.41)

12 1
ub1 +
11
ub 2 + 13 ub3 =
11 11
1k uk .
232 DYNAMICS OF MACHINERY

13.5.2 Modal balancing

The deflection of a rotor shaft at a critical speed is described as a


precession eigenform or mode shape. Due to damping, precession mode shapes are
three-dimensional curves. In the balancing practice, they are approximated by
planar mode shapes described by curves lying in axial planes, generally non
coincidental [13.20]. For a flexible rotor in two bearings (Fig. 13.11, a) the first
three planar mode shapes are shown in Figs.13.11, b, c, d.

a b c d

e f g h

i j k l

m n o p

q r s t
Fig.13.11 (from [13.18])
13. BALANCING OF ROTORS 233

Modal balancing is based on the assumption that the distributed unbalance


in the rotor (Fig. 13.11, e) can be represented as a sum of modal unbalances (Fig.
13.11, f, g, h), which are proportional to the eigenforms, but lying in different axial
planes. At the same time, the shaft deflection at a particular speed (Fig. 13.11, i)
has a modal expansion by virtue of which it can be resolved into a sum of terms
proportional to the eigenforms (Fig. 13.11, j, k, l).
Running at a critical speed, the deflected shape almost coincides with the
corresponding eigenform, the other terms in the expansion being vanishingly small.
Due to the orthogonality of mode shapes, a particular unbalance distribution
ui (x ) = ai i (x ) , proportional to the ith eigenform i (x ) , can only excite the
lateral deflection wi ( x ) = bi i (x ) , i. e. the deflection which also takes the ith
eigenform. Thus unbalance excitation and bending response shapes are the same
for each term in the series.
This property enables the serial elimination of unbalance terms ui (x ) by
compensating balance masses.
In practice, this means running the rotor up to very near its first critical
speed and adding unbalances (correction masses) u1k (k = 1, 2,...) at particular
locations xk on the rotor. The rotor deflection is thereby reduced enough to allow
running through the first critical speed and almost up to the second.
The process is repeated with another set of balancing masses
u2k (k = 1, 2 ,...) . Note that the individual sets of balancing masses only affect the
bending caused by the corresponding eigenform and no other. This leads to a
systematic balancing procedure with condition equations for the sets of balance
masses [13.18].
From this it is found that balancing out N critical speeds or N eigenforms
requires at least N planes and therefore the addition of at least as many balancing
masses (known as the N plane method).
If there are, for example, two critical speeds in the operating range and one
just over, then at least 2 + 1 = 3 balance planes are necessary.
Let the three correction planes be denoted I, II, III (Fig. 13.11, m). To
compensate the unbalance at the first critical speed, a single mass producing u12
should be added at the antinode of the first eigenmode (Fig. 13.11, n). To
compensate the unbalance at the second critical speed, two masses producing the
unbalances u21 and u23 should be added in antiphase near the antinodes of the
second eigenmode (Fig. 13.11, o). To balance at the third critical speed, three
masses producing u31 , u32 , u33 are located at the antinodes of the third eigenmode
(Fig. 13.11, p). This way the rotor is not balanced as a rigid rotor at low speeds.
234 DYNAMICS OF MACHINERY

The end result of such balancing is subject to small errors, as higher order
unbalance terms, or eigenforms remain uncompensated.
The accuracy can be much improved if the rotor is balanced as a rigid rotor
at low speed. This means that two more balance planes are required (using the
N + 2 plane method), as shown in Fig. 13.11, q, where the five planes are
indicated as I, II, III, IV, V. The individual sets of unbalance masses are then
statically in balance, i.e. the sum and the static moment of the unbalances are zero.
The locations of balance masses are shown in Figs. 13.11, r, s, t. Their magnitudes
and angular positions are measured in the neighbourhood of the critical speeds nk1 ,
nk 2 , nk 3 . One set of masses affects only the deflection of one particular
eigenform. Each set is both statically and dynamically compensated. The method
includes balancing the rotor at low speeds as a rigid rotor. The sets of masses no
longer affect the balance of the rigid rotor.

13.5.3 General remarks

To obtain a complete balance in a given speed range it would be


theoretically necessary to have an infinite number of balancing planes and the same
number of compensating unbalances. In practice, a finite number of planes is used.
Therefore every practical method involves a certain amount of error.
In both the N plane method and the N + 2 plane method, one source of
error is the neglection of the higher-order modes of precession. The main error of
the N plane method is the failure to meet the condition of rigid-body balancing.
Perfect balancing of a flexible rotor with a finite number of masses is
theoretically impossible. For a large rotor with 80 m bearing vibration before
balancing, the above systematic procedures can normally balance the rotor to have
10 m . The difficulties start when one wants to get under 10 m , due to thermal
effects and non-linearities of oil film. Other problems are encountered when the
effect of the second eigenform in the unbalance is so severe that the first critical
speed cannot be reached. Balancing of the first eigenform then requires at least two
or preferably three planes.
With elastic bearings the rotor deformation is not exclusively due to its
flexural rigidity. The stiffness of the bearings and pedestals is also important and
has to be taken into account in balancing machines with soft bearings.
Other details are given in the book [13.4] and the standard ISO 11342-
1998 [13.21].
Two Romanian standards derived from the corresponding ISO versions are
[13.22] and [13.23].
13. BALANCING OF ROTORS 235

References

13.1. ISO 1925, Mechanical vibration, Balancing Vocabulary, 2001.


13.2. * A practical guide to in-place balancing, IRD Mechanalysis, Technical
Paper No.116, 1981.
13.3. Schneider, H., Auswuchttechnik, VDI Taschenbcher T29, Dsseldorf, 1972.
13.4. Kellenberger, W., Elastisches Wuchten, Springer, Berlin, 1987.
13. 5. Thearle, E. L., Dynamic balancing of rotating machinery in the field, Trans.
ASME, vol.56, 1934, p.745-753.
13.6. Chen, W. J. and Gunter, E. J., Introduction to Dynamics of Rotor-Bearing
Systems, Trafford Publ., Victoria, Canada, 2005.
13.7. Somervaile, I. J., Balancing a rotating disc: simple graphical construction,
Engineering, Feb 1954.
13.8. * Static and dynamic balancing, Brel & Kjaer Application Note No. 17-
227.
13.9. Kellenberger, W., Balancing flexible rotors on two generally flexible
bearings, Brown Boveri Review, vol.54, no.9, Sept 1967, p.603-617.
13.10. Dimarogonas, A. D. and Haddad, S., Vibration for Engineers, Prentice Hall,
Englewood Cliffs, NJ, 1992.
13.11. Gunter, E. J. and Jackson, Ch., Balancing of rigid and flexible rotors,
Handbook of Rotordynamics, Ehrich, F. F. ed., McGraw-Hill, New York, 1992.
13.12. ISO 1940-1, Mechanical vibration - Balance quality requirements for rotors
in a constant (rigid) state, Part 1: Specification and verification of balance
tolerances, 2003.
13.13. ISO 1940-2, Mechanical vibration - Balance quality requirements of rigid
rotors, Part 2: Balance errors, 1997.
13.14. Dimentberg, F. M., Theory of balancing flexible rotors, Russian
Engineering Journal, vol.11, 1964.
13.15. Bishop, R. E. D. and Parkinson, A. G., Vibration and balancing of flexible
shafts, Applied Mechanics Reviews, vol.21, no.5, May 1968, p.439-451.
13.16. Darlow, M. S., Smalley, A. J. and Parkinson, A. G., Demonstration of a
unified approach to the balancing of flexible rotors, ASME Paper 80-GT-87,
ASME Gas Turbine Conference, March 1980.
13.17. Bishop, R. E. D. and Gladwell, G. M. L., The vibration and balancing of an
unbalanced flexible rotor, J. Mech. Eng. Sci., vol.1, no.1, 1959, p.66-77.
236 DYNAMICS OF MACHINERY

13.18. Kellenberger, W., Should a flexible rotor be balanced in N or (N + 2)


planes?, ASME Journal of Engineering for Industry, vol.94, 1972, p.548-560.
13.19. Den Hartog, J. P., Mechanical Vibrations, 4th ed., Dover, New York, 1984.
13.20. Gunter, E. J., Barrett, L. E. and Allaire, P. E., Balancing of multimass
flexible rotors, Proc. 5th Turbomachinery Symp., A&M University, Cllege
Station, Texas, Oct 1976.
13.21. ISO 11342, Mechanical vibration Methods and criteria for the mechanical
balancing of flexible rotors, 1998.
13.22. SR ISO 1925: 1995, Echilibrare. Vocabular.
13.23. SR ISO 1940-1: 1994, Vibraii mecanice. Condiii de calitate pentru
echilibrarea rotoarelor rigide. Partea 1: Determinarea dezechilibrului rezidual
admisibil.
14.
RECIPROCATING MACHINES

There are two groups of vibration phenomena of practical importance in


reciprocating machines: a) vibrations transmitted to the foundation by the machine as
a whole, and b) torsional vibrations in the crankshaft and in the shafting of the driven
machinery. They are produced by the periodic accelerations of the crank mechanism
and the periodic variations in cylinder gas pressure. Only the vibratory motions of
the machine frame and the developed reaction forces are studied herein. An example
is also given of fault diagnosis by vibration measurement for an auxiliary diesel
engine. Related topics are the acoustic resonances and pulsation control of the piping
system of reciprocating compressors.

14.1 Single cylinder engines

Internal combustion engines, piston-type compressors and pumps, and other


machinery involving a crank mechanism produce reciprocating forces. The crank
mechanism transfers a reciprocating motion to a rotary motion, or vice versa. After
the mass and center of gravity of each of the moving parts are determined, the forces
resulting from operation of the machine can be evaluated.
Consider a vertical single-cylinder engine. The crank mechanism consists of
three bodies: a) a crank OA which rotates, b) a reciprocating body which slides back
and forth in pure translation; point C is either the piston pin or the crosshead; and c)
a connecting rod AC which joins them (Fig. 14.1, a).

14.1.1 Gas pressure excitation

Consider the effect of fluctuating gas pressure in the engine cylinder. Any
inertia effect is excluded, by assuming that the engine is turning over very slowly at
a constant speed [14.1].
238 DYNAMICS OF MACHINERY

Let P be the pressure force on the piston, which is variable with the time
(or with the crank angle = t ). The gas pressure pushes the piston downward and
presses upward against the cylinder head so that when the entire engine is
considered, the resultant in any direction would be zero.

a b
Fig. 14.1

However, the force P results in a torque about the crankshaft, called the
gas-pressure torque. Indeed, the piston force P (force F1 ) is transmitted to the piston
pin C (or through the piston rod to the crosshead). Neglecting friction, the force F1
is held in equilibrium by the forces F2 = P tan and F3 = P cos . The forces F1 ,
F2 , F3 of Fig. 14.1 are acting on the piston pin (or crosshead). The reaction force to
F2 acts to the right on the guide or frame (Fig. 14.1, b).

The force equal and opposite to F3 is a compression in the connecting rod.


It is transmitted through the connecting rod to the crank pin as force F 4 . By shifting
this force parallel to itself to O we obtain a force F 5 = F 4 and a torque T p which is
the driving torque of the gas pressure
14. RECIPROCATING MACHINES 239

P
Tp = d = P y tan . (14.1)
cos
The force F5 = P cos is taken up by the main bearings at O and can be
resolved into a vertical component F6 = P and a horizontal component
F7 = P tan .

The four forces transmitted to the stationary parts of the engine are: a) P
upward on the cylinder head; b) P tan to the right on the cylinder or the crosshead
guide; c) P downward on the main bearings at O; and d) P tan to the left on the
main bearings at O (Fig. 14.1, b).
The total resultant force on the frame is zero, but there is a resultant torque
P y tan acting clockwise. By the law of action and reaction, this torque must be
equal and opposite to the driving torque d P cos on the crankshaft (acting
anticlockwise in the direction of rotation).
Thus the gas pressure in the cylinder do not causes any resultant force on
the engine frame, but produces only a torque about the longitudinal axis.

14.1.2 Inertia effects

Assume that the piston executes a vertical alternating motion. While the
piston is accelerated downward, there is an upward inertia force F y acting on it
(Fig. 14.2, a), and this force must have a reaction F y pushing downward against
the stationary parts of the engine, which is not balanced internally.
The piston is accelerated downward by a force F 3 = F y cos along the
connecting rod. The force F4 on the crank pin exercises a torque about the
crankshaft axis
Fy
Ti = F4 d = d. (14.2)
cos
Since the piston acceleration is alternating, this inertia torque is also
alternating. The clockwise torque

Ti = F4 d = F2 y

has a counterclockwise reaction torque on the frame, hence of opposite direction


with respect to the gas pressure torque (Fig. 14.2, b).
240 DYNAMICS OF MACHINERY

a b
Fig. 14.2

There are also inertia forces of the rotating parts (the crank and the parts
revolving with it). These can be reduced to zero by counterbalancing the crankshafts.
Figure 14.3 shows the finite element model of a crankshaft in which the
counterweights can be seen.

Fig. 14.3
14. RECIPROCATING MACHINES 241

14.1.3 Kinematics of crank mechanism

Denote l - length of the connecting rod and r crank radius (Fig. 14.4).
Suppose that C and A initially coincide to D (top dead center) and A' , respectively,
in the line of stroke OD [14.2].

Fig. 14.4

The displacement of the piston pin C is


s = DC = DA'+ A' O BO CB ,
s = l + r r cos l cos = r ( 1 cos ) + l ( 1 cos ) . (14.3)
Because
AB = l sin = r sin
one can write
r
sin = sin ,
l
12
r2 1 r2
cos = 1 sin = 1
2
sin 2 1 sin 2 . (14.4)
2 2 l2
l
242 DYNAMICS OF MACHINERY

Using equations (14.4) and (14.3) we obtain

r2
s = r ( 1 cos ) + sin 2 ,
2l
or
r2
s = r ( 1 cos ) + ( 1 cos 2 ) ,
4l
r 2 r
s = r + r cos t + cos 2 t . (14.5)
4l 4l

The velocity of the piston is
ds r
v= = r sin t + sin 2 t .
dt 2 l
The acceleration of the piston is
dv r
a= = r 2 cos t + cos 2 t . (14.6)
dt l
Note that one term varies with the same frequency as the rotation; this is
called the primary term. The term which varies at twice the frequency of rotation is
called the secondary term. The importance of the secondary term is established by
the crank-connecting rod ratio r l . For a connecting rod of finite length, the motion
of the piston is periodic but not harmonic.

14.1.4 Connecting rod and equivalent two-mass system

For purposes of analysis it is desirable to replace the connecting rod by an


equivalent dynamical system composed of two concentrated masses, as shown in
Fig. 14.5.
For the two mass system to be dynamically equivalent to the original
connecting rod, it must satisfy the following requirements: a) same total mass; b)
same center of mass; and c) same mass moment of inertia. These three conditions
can be expressed by the equations
m = m1 + m 2 ,

m2 h = m c , (14.7)

m k G2 = m1c 2 + m 2 ( h c )2 ,
where k G is the radius of gyration of the rod about the piston pin.
14. RECIPROCATING MACHINES 243

In the dynamically equivalent rod, m1 undergoes translation while m 2 is


subjected to translation and rotation.

Fig. 14.5 (after [14.2])

By proportioning the connecting rod with a slight extension beyond the


crank end, it is possible to make h = l . The mass m 2 then coincides with the crank
pin and its motion is that of pure rotation. The two concentrated masses are then
expressed by the equations
c c
m1 = m 1 , m2 = m . (14.8)
l l
We will analyze here only the simpler case where the connecting rod is
replaced by concentrated masses expressed by equations (14.8).
The translating mass mt in C is the sum of the piston mass and the portion
of the connecting rod m 1 . The rotating mass m r is composed of the remaining
portion m 2 and any unbalanced mass of the crankshaft assigned at this position, both
of which will be assumed to be balanced by a counterweight.

14.1.4 Unbalance of a single cylinder engine

The inertia force of the moving parts is in vertical direction and of


magnitude
r
Fy = mt a = mt r 2 cos t + mt r 2 cos 2 t . (14.9)
l
We thus have a primary unbalance at a frequency equal to the rotational
speed, and a secondary unbalance at a frequency equal to twice the rotational speed.
The inertia force mt a has also a torque about the crankshaft equal to
244 DYNAMICS OF MACHINERY

Ti = mt a y tan =
r . (14.10)
= mt r 2 cos t + cos2 t (l cos + r cos t ) tan
l
Replacing
r
tan sin t , cos 1.0 ,
l
equation (14.10) becomes
r r
Ti = mt r 2 2sin t cos t + cos 2 t 1 + cos t . (14.11)
l l
Multiplying out and omitting higher powers of r l yields

r r
Ti = mt r 2 2 sin t cos t + sin t cos 2 t + sin t cos 2 t (14.12)
l l
Using the trigonometric relations
1
sin t cos 2 t = ( sin 3 t sin t ) ,
2
1
sin t cos t = sin 2 t , (14.13)
2
1
cos 2 t = ( 1 + cos 2 t ) ,
2
equation (14.12), giving the inertia torque about the crankshaft, reduces to
1 r 3r
Ti = mt r 2 2 sin t sin 2 t sin 3 t . (14.14)
2 2l 2l
The engine-frame torque differs from the reverse of the crankshaft torque
by the magnitude of the so-called residual couple. This is an inertia couple due to the
connecting rod. It corrects for the error in the angular acceleration of the connecting
rod which is introduced when the common assumption is made that the mass of the
connecting rod is borne at the piston pin and crankpin in inverse proportion to the
distances of these points from the center of gravity of the connecting rod.
The residual couple of the connecting rod usually is negligible in in-line
engine cylinders but is taken into account in radial engine dynamics.
If the pressure variation throughout the cycle of the machine is known, it is
possible to evaluate the gas-pressure torque (14.1) as a function of the crank angle
. This calculation is based on the pressure-volume diagram for a typical cylinder,
obtained from the pressure-volume card obtained experimentally.
14. RECIPROCATING MACHINES 245

From the diagram of the cylinder pressure vs. crank angle (Fig. 14.6, a) the
resulting gas-pressure torque can be calculated as a function of crank angle, hence
time (Fig. 14.6, b).

a b
Fig. 14.6 (after [14.3])

It is usually convenient to make a harmonic analysis of this function in the


following form:


T p = b0 + a1 2 sin + b1 2 cos + a1 sin + b1 cos +
2 2
(14.15)
3 3
+ a3 2 sin + b3 2 cos + a 2 sin 2 + b 2 cos 2 + ....
2 2

In a two-stroke machine, the cycle is complete in a single revolution, and


only integer orders occur. In a four-stoke machine, the cycle requires two
revolutions, and, in general, half-integer as well as integer orders will occur. The
coefficients of all orders up to j = 18 for a number of representative engine cycles
have been tabulated (F. P. Porter 1943).
Neglecting gravity, the total torque delivered to the crankshaft is the sum
of the inertia torque (14.14) and the gas-pressure torque (14.15). The total torque on
the frame is that due to inertia plus the negative of (14.15) due to gas pressure. The
total forces on the frame are due only to inertia.
In a more detailed analysis [14.4], not limited by the approximation (14.4),
the magnitude of the inertia force (14.9) has the form

(
Fy = mt r 2 b1 cos t + b 2 cos 2 t + b4 cos 4 t + b 6 cos 6 t + .... ) (14.16)

where
246 DYNAMICS OF MACHINERY

1 3 15 5
b1 = 1 , b2 = + + + ... ,
4 128
1 3 9 5
b 4 = 3 + 5 + ... , b6 = , (14.17)
4 16 128
and
r
= . (14.18)
l
It contains only even higher order components.
The inertia torque (14.14) has the form

( )
Ti = mt r 2 2 a1 sin t + a 2 sin 2 t + a3 sin 3 t + a 4 sin 4 t + .... (14.19)

where
1 1 15 5 1 1
a1 = + 3 + + ... , a 2 = + 4 + ... ,
4 16 512 2 32
3 9 81 5 1 7
a 3 = + 3 + + ... , a 4 = 2 + 4 + ... , (14.20)
4 32 512 4 32
5 3 75 5 3
a5 = + + ... , a 6 = 4 + ... .
32 512 32

It contains integer harmonic components of all orders. The major unbalance


occurs at twice rotational speed.

14.2 Multi cylinder engines

In the single-cylinder engine there will always be the unbalance due to the
translating mass mt . In the multi-cylinder engine, the unbalance due to mt can be
cancelled by the proper angular spacing of the cranks.

14.2.1 Unbalance force and couples

By combining several cylinders acting on the same drive shaft into a single
rigid frame, it is possible to balance out some of the important harmonics in the
forces and the moments of the individual cylinders. Although many configurations
are possible (see Table 4.1), we will here discuss only the in-line machine, in which
n identical cylinders are equally spaced along a straight line, as shown in Fig. 14.7.
14. RECIPROCATING MACHINES 247

Fig. 14.7 (after [14.4])

Let the crank position (offset angle) be defined by j with respect to the
first crank ( 1 = 0 ) .
Based on equations (14.9) and (14.14), the inertia unbalance of a
counterbalanced multi-cylinder engine consists of a vertical force of magnitude
n

cos ( t + ) + l cos 2 ( t + )
r
Fi = mt r 2 j j (14.21)
j =1
and a yawing moment
n
r
2l sin ( t + ) sin 2 ( t + ) 2l ( )
1 3r
M y = mt r 2 2 j j sin 3 t + j .(14.22)
2 j =1

Even if these forces and moments add up to zero, it is possible to have a


pitching moment about a horizontal axis perpendicular to the center line of the
engine. This moment can be found by summing the moments of the Fy forces about
the first cylinder. If the distance from the jth cylinder to cylinder 1 is c j , the
pitching moment about the first cylinder becomes
248 DYNAMICS OF MACHINERY

n
M z = mt r 2 c j ( )
cos t + j +
r
( )
c j cos 2 t + j , (14.23)
j =1
l
where c1 = 0 .

Table 14.1

Table 14.1 [14.5] illustrates the forces and couples developed by some
multi-cylinder machines for different crank arrangements and numbers of cylinders.
It applies to machines having the same bore and strike for each cylinder. For
compressors at which the bore and strike of the cylinders are not all the same, Table
14.1 should not be used. The unbalanced forces and couples should be computed for
each cylinder and the results superposed.
In multi-cylinder engines and compressors, the net forces and torques are
modified by the cancellation of harmonics among events in the different cylinders.
This cancellation is achieved by the arrangement of cylinders, positioning of cranks
14. RECIPROCATING MACHINES 249

of the crankshaft, ignition sequence, etc. A general treatment of engine balancing


[14.6] is beyond the aim of this presentation. The main results can be summarized as
follows [14.3]:
Multi-cylinder in-line four-stroke-cycle engines, firing at equal intervals
and with parallel lines of stroke are unbalanced with respect to harmonics of order
N 2 , whenever N is an integral multiple of the number of cylinders n. The
reciprocating inertia force contains no odd harmonics after the first. Half-integer
harmonic orders occur only for the gas-pressure torque. The magnitudes of the
unbalanced harmonics are n times the magnitudes of corresponding harmonics for
one cylinder.
The reciprocating inertia forces act in planes normal to the crankshaft and
distributed along the length of the shaft. The unbalanced harmonics are n parallel
forces which alternate in phase in the lines of stroke of each cylinder at frequencies
i n 2 , where i is an integer. The resultant of these forces is at the centroid of the
parallel forces. For a balanced-type crankshaft (rear half is a mirror image of the
front half, implying an even number of cylinders), this point is at the middle of the
crankshaft. The harmonics of the moment of the resultant force with respect to a
reference line are called rocking moments, since the resultant force may produce
pitching vibrations of the engine on its mounts (or on the foundation).
The total engine-frame torque due to both gas-pressure reactions and
dynamic unbalances is obtained by adding the magnitudes of corresponding
harmonics of the gas-pressure torque reaction and inertia couple due to inertia torque
of reciprocating parts. Their effect is to produce rolling vibrations of the engine
frame.
A six cylinder four-cycle engine of 0, 120, 240, 240, 120, 0 deg. crank
shaft has all forces balanced and all moments balanced. Consequently, a V-12 engine
made up of two in-line 6s would also be balanced. An eight-cylinder in-line engine
(0, 180, 90, 270, 270, 90, 180, 0 deg.) is completely balanced.
The balance of the engine presupposes that the engine parts, including the
engine frame, are rigid bodies which do not deform under the influence of a balanced
force system. Actually, the engine is composed of elastic bodies and the balanced
forces may cause deformations which are vibratory in nature. The effects of
elasticity of engine parts is most pronounced in the case of large marine in-line
engines.
Some of the assumptions upon which a theoretical treatment is based often
are not satisfied exactly in the practical cases. A basic assumption is that each
cylinder produces an identical pressure-time history. Irregularities arising from
variations in ignition, fuel distribution to the cylinders, irregular valve operations,
etc., violate this assumption. They usually excite the fundamental harmonic of the
gas-pressure cycle, which is quickly noticeable and usually referred to as engine
roughness [14.3].
250 DYNAMICS OF MACHINERY

14.2.2 Other vibration sources

Apart from being familiar with the potential malfunctions of an engine, an


understanding of the interaction between different malfunction mechanisms is
needed. Interactive fault mechanisms such as misalignment, excessive bearing
clearance, rigid body resonances of the engine on the supporting structure or elastic
resonances of large engine frames, or loss of bolting tightness, occur quite
frequently. A thorough knowledge of all potential malfunctions with their
corresponding symptoms and interactions is mandatory for proper diagnosis of
engine condition.
Other common sources of vibration are the impacts in the fuel injection
system and in the valve train, the piston slap, the lack of tightness of bearing
components, the residual unbalances of rotating parts, various misalignments,
exhaust-gas impulses and looseness of mounting bolts to the subbase.
The fuel injection system produces vibrations due to the impacts of the cam
on the injector train push rod (injection pump) of diesel engines. It influences the
combustion law, namely the ignition delay and the injection rate. Disturbances in the
injected fuel time history and the injection timing have great influence on the
ignition delay, whose variation in different cylinders is indicated by an increase of
fractional order components, especially X and n X 2 , where n is the number of
cylinders. Decrease of ignition delay reduces the level of combustion-generated
vibrations.
The impacts occurring in the valve gear due to clearances produce
vibrations in the gas distribution system. In the four-stroke engine, the camshaft
rotates at half the crankshaft speed, so that any related malfunction will produce an
increase of the one-half order component (X 2 ) and its integer multiples. Increased
clearances in camshaft bearings due to wear as well as improper setting of the
clearance between rockers and valve stems produce an increase of all vibration
frequencies, especially of X 2 , n X 2 and 3n X 2 , where n is the number of
cylinders (three cams for each cylinder). The impact between valves and seats
produce a 2n X 2 component.
Other engine systems (lubricating, cooling, and turbocharging) also
influence the level of overall engine vibrations. They generate frequency harmonics
of orders corresponding to the speed of respective pumps, the gear mesh, etc.
Increased clearances in the running gear produce sharp metallic knocking
at all engine speeds. They can occur between the piston pin and the connecting rod
bush, between the crank pins and the connecting rod or between the crankshaft and
the shell of main bearings. Based on experience, this type of trouble produces a high
first-order component ( 1X ) followed by a string of harmonic components at higher-
order harmonics. This is partly explained by an amplitude modulation of the
vibration signal which is truncated due to the clearance. Spectral analysis of a
14. RECIPROCATING MACHINES 251

truncated waveform yields a number of discrete frequencies, sometimes sum and


difference components. A similar problem occurs with looseness of the engine frame
bolting to the subbase. Strong second-order (2X ) harmonics occur when some bolts
lose their tightness.
Increase of the clearance between the main bearing shell and housing can
determine a periodic variation of the crankshaft support stiffness. Stiffness
nonlinearities produce Mathieu-type vibrations with subsynchronous spectral
components X 2 or X 3 , and their multiples. Usually the X 2 component is
accompanied by integer multiples 1X, 3 X 2 , 2X, 5 X 2 of decreasing amplitude
(with increasing order).
Piston slap is an impact generated by the reversal of the side force acting
on the piston, especially near the top dead center. In large diesel auxiliary engines,
large impact forces occur due to the relatively large piston mass and the piston to
liner clearance, which in some cases overpass combustion generated forces. These
impacts produce an increase of all harmonic components in the frequency spectrum,
especially low order.
One cannot neglect excitations induced by the driven (electrical) machine
or other neighboring engines.

14.2.3 Fault diagnosis of a diesel engine

This section presents an example of dynamic analysis and vibration


measurements on 680 kW, four-stroke, 750 rpm rated speed, non-reversible five-
cylinder diesel engines with direct injection, exhaust turbocharging and charge air
cooling system [14.7].
The dynamic analysis of such an engine yields the following unbalanced
external forcing functions (Fig. 14.8):

Fig. 14.8

a) A pitching moment due to reciprocating inertia forces. It is represented


as a couple acting in the vertical plane containing the lines of stroke of the cylinders,
252 DYNAMICS OF MACHINERY

usually referred to as M 1V (external couple, first order, 1X) and M 2 V (external


couple, second order, 2X).
b) A resultant vertical force, usually small, with the tenth harmonic F10
the first major unbalanced component (10X ) .
c) A rolling moment, due to both the gas-pressure torque reaction and the
inertia torque of reciprocating parts, having the first significant components: the
fifth-order harmonic M 5R ( 5X ) due to gas pressure and the tenth-order harmonic
M 10R ( 10X ) due to both gas-pressure reaction and inertia torque. Higher order
harmonics are of negligible magnitude.
Measurement points were chosen on the mounting feet, on the engine
frame at the crankshaft level, at the top edge of the frame and on the cylinder covers,
at several locations along the engine (between the coupling flange end and the
turbocharger end).
Apart from overall level measurements, frequency spectra of the peak-to-
peak vibration velocity were plotted, in the frequency range 4-100 Hz. A
characteristic spectrum, plotted at 750 rpm crankshaft speed and partial load, is
shown in Fig. 14.9

Fig. 14.9 (from [14.7])


14. RECIPROCATING MACHINES 253

a b

c d
Fig. 14.10 (from [14.7])
254 DYNAMICS OF MACHINERY

Worthy of note are the large amplitude components 2X and 2X. The first
is produced by the unbalanced pitching moment M 2 V , the second corresponds to the
ignition rate in five cylinders. Diesel engine-electric generator alignment was
checked during measurements to minimize the misalignment-induced 2X spectral
component.
The harmonic 5X can be produced by the rockers or by the unbalanced
rolling moment M 5 R . The 7X component corresponds to the impact rate of the
camshaft (two cams/cylinder for valves plus one cam/cylinder for injection). The
X 2 component, which corresponds to the ignition rate in a cylinder and to the
camshaft speed, is associated with irregularities in the injection timing.
In order to explain the relatively large amplitude of components 2X and
2X, measurements were performed at various engine speeds, at a point near the
cylinder covers, with the engine in idling condition.
At 500 rpm (Fig. 14.10, a) the component 2X has the highest level, while
2X is slightly lower. At 600 rpm (Fig. 14.10, b), the component 2X occurs at 25
Hz, with a level three times higher than at 500 rpm. At 680 rpm (Fig. 14.10, c) 2X
occurs at 28.3 Hz, having a level 2 times lower than at 600 rpm, while the
component 2X increases to 5 mm s . At 750 rpm (Fig. 14.10, d), 2X decreases to
8.5 mm s , while 2X, which now occurs at 25 Hz, increases to 12 mm s , having the
highest level.
It was concluded that, at 25 Hz, a structural resonance occurs, excited at
600 rpm by the component 2X and at 750 rpm (the rated speed), by the component
2X.
Measurements have shown that the relative level of the 2X and 2X
components decreases at points located in the middle of the engine, both at the
engine mounting feet level and at the top edge of the frame (Fig. 14.11). The
relatively low frequency of the recorded vibrations and the amplitude maps plotted
along the height and the length of the engine excluded consideration of lateral
resonances of the flexible engine frame, in the so-called H-shape and X-shape modes
of vibration.
It was concluded that the resonance is a rigid body resonance of the whole
engine-generator set on the flexible engine room floor structure.
The main cause of engine vibration does not lie, in this case, in the engine
itself, but in the insufficient stiffness of its supporting structure. Location of the
engine with the middle on a local transverse stiffener was not sufficient to prevent
the pitching vibrations induced by the unbalanced moment M 2 V . Supplementary
stiffeners have been added, to transmit the load from the engine base to adequate
bulkhead plating and deck stiffeners, which eliminated the resonance condition.
14. RECIPROCATING MACHINES 255

In some engines, loosening of the engine bolting to the subbase yielded


higher 1X and 2X components, with a series of higher multiples, which made
identification of vibrations stemming from the engine itself more difficult (Fig.
14.12).

Fig. 14.11 Fig. 14.12

The effect of other malfunctions was studied, mainly by simulating various


faults. Irregularities in combustion produce large X 2 , 2X and 7X harmonic
components (Fig. 14.9). In fact, all measured engines were maladjusted, compression
checks showing differences up to 15% between cylinders for the same engine. When
the injection was shut off in the cylinder next the measurement point (Fig. 14.13) or
in the adjacent cylinder (Fig. 14.14), the X 2 component decreased at a level
corresponding to measurements at idleness.
A broken valve spring, producing a recognizable clattering noise, gave rise
to larger 2X and 5X components in the frequency spectrum measured at a point
near the respective cylinder. Excessive valve clearance produced deliberately during
measurements had the same effect. Other faults such as worn main bearing half-
shells, a damaged roller in the fuel injection pump, misalignment preloads caused by
offset camshaft bearing housings, and problems with the nozzle needle of the fuel
injection valve gave minor or unnoticeable changes in the frequency spectrum.
In some cases the diagnosis was made difficult because the faults caused
the engine to run unevenly and spectrum averaging attenuated some changes in the
frequency spectra.
256 DYNAMICS OF MACHINERY

Fig. 14.13 Fig. 14.14

14.3 Reciprocating compressors and piping systems

Piston-type reciprocating compressors are currently used in some refining


and gas-processing facilities. The major concern is to avoid or eliminate excessive
vibrations and dynamic stresses caused by mechanical and pulsation-induced
shaking forces. This is achieved by the reduction and control of pulsation levels as
well as the restraint of piping and use of elbows only where required. This section
presents basic information about the features of reciprocating compressors, as
described by the API Standard 618 [14.8], acoustic resonance phenomena and
pulsation control in piping systems, as described in reference [14.9].

14.3.1 Compressor cylinder manifold system

Figure 14.15 shows a typical compressor cross section and the customary
nomenclature. The cylinder is connected to the compressor frame via the distance
piece and the crosshead guide, both of which have flexibility in axial, transverse, and
torsional direction, and each of which is of rather complicated geometry. Their
14. RECIPROCATING MACHINES 257

correct modeling is a condition for the accurate analytical prediction of the


mechanical natural frequencies, associated mode shapes, and cyclic stresses in the
compressor cylinder-manifold system, in order to avoid coincidences between the
mechanical and acoustical resonant frequencies.

Fig. 14.15 (from [14.9])

Figure 14.16 shows various components of an ideally simple two cylinder


system. Besides the cylinders, there are the suction and the discharge bottle which
often have internals in the form of choke tubes and baffles (Fig. 14.17).

Fig. 14.16 (from [14.10])


258 DYNAMICS OF MACHINERY

In addition, the discharge bottle is usually restrained by wedge clamps of


finite stiffness. The bottles are flexibly connected to the cylinders via the nozzles.
One end of the nozzle is connected to the cylinder by a bolted flange joint, and the
other end of the nozzle penetrates the bottle with a welded joint (branch connection).
Additional suction and discharge piping, additional stages, connected units, etc.,
distort the ideal features of Fig. 14.16 and make the mechanical analysis of this
system even more complex.

Fig. 14.17 (after [14.10])

14.3.2 Excitation forces

Reduction in excitation force levels is often achieved by attenuating pressure


pulsation amplitudes through the insertion of passive filters into the gas flow path.
Attenuation of the pressure pulsations in reciprocating compressor installations
results in improved compressor performance and reduction of dynamic pressure
drop, i.e. additional pressure drop due to pulsations, with the expense of unwanted
losses in the static pressure drop, produced by the filtering devices.

Bottle unbalance forces


The predominant pulsation induced excitation arises from the unbalanced
forces occurring in the suction and discharge bottles. These forces are the result of
differential pressures acting axially on surfaces within the bottle (end caps and
baffles). They can be calculated based on data on amplitude and relative phase of
pressures within the bottle. The dynamic components of these forces are represented
as a function of frequency under the form of the net shaking force spectrum.

Cylinder internal pressure forces


The gas force, or load, is equal and opposite to the cylinder internal pressure
force, i.e. the cylinder internal pressure force acts on the heads of the cylinders, the
opposing gas force acts on rod. The net cylinder internal pressure force is calculated
from the head end internal cylinder pressure times the area of the head end minus the
14. RECIPROCATING MACHINES 259

crank end internal cylinder pressure times the area of the crank end head. The
internal cylinder pressure force is plotted versus crank angle in Fig. 14.18.

Compressor mechanical unbalance forces


These forces are caused by accelerations of the piston assembly and
crankshaft throws and are transmitted to the compressor frame via the bearing
reactions. This type of force is determined from the manufacturers balance data and
pressure-volume cards.

Fig. 14.18 (from [14.8])

The compressor inertial force versus crank angle (time) is nearly sinusoidal
(broken line in Fig. 14.18). As a result, depending on the connecting rod length to
stoke ratio, the amplitude of the second harmonic is about 20% of the fundamental.
There are no significant harmonics above two times crank shaft speed. The gas force
plus the inertial force represents the total load (solid line).

Cylinder stretch forces


Figure 14.19 illustrates, schematically, an additional important excitation
force - the cylinder stretch. Under the action of compression loads within the
cylinder, each cylinder will tend to stretch and shrink once per revolution of the
crankshaft. Thus, there is a strong first order component of cylinder stretch motion,
with the various cylinders out-of-phase with each other due to the crank angle
phasing. The magnitude of cylinder stretch can often be 0.25 mm in peak-to-peak
amplitude and occasionally reaches 0.4 mm. This movement can, in some cases,
cause excessive stresses in the attached nozzles and bottles. First order cylinder
stretch has sometimes been a significant contributor in breaking nozzles and baffles.
260 DYNAMICS OF MACHINERY

Fig. 14.19 (from [14.10])

Shaking forces in piping


Acoustic shaking forces act in the axial direction of pipe runs causing
vibration in the axial direction of that pipe run. But maximum vibration actually
occurs in the transverse direction of adjoining piping that runs perpendicular to the
piping where the shaking force is acting. Figure 14.20 shows the action of shaking
forces on common adjoining piping configurations, including spans with pinned to
nearly fixed ends, L-bends and U-bends.

Fig. 14.20 (from [14.8])


14. RECIPROCATING MACHINES 261

14.3.3 Pulsation analysis

Because of the nature of reciprocating compressors, pressure pulses are


generated and transmitted into the piping system. Pulsations are a major cause of
reduced reliability and lost efficiency in compressor systems and their piping.
Unbalanced forces caused by pulsations at piping elbows, surge volumes, etc. can
result in high vibration levels and cause fatigue failures of piping, supports and
nozzles. Reflection of pulsations back to suction or discharge valves can cause
changes in the valve opening time, distortion of the pressure-volume card and
reduction in cylinder capacity and efficiency, as well as increasing valve
maintenance. Plane wave theory is satisfactory for the analysis of pulsations in
reciprocating compressors in typical piping systems in the petrochemical industry.

14.3.3.1 Pulsation excitation mechanisms


Reciprocating compressors generate flow modulations which in turn
generate pressure pulsations. The flow modulations are a result of intermittent flow
through the suction and discharge valves. They are superimposed upon the steady
(average) flow.
Figure 14.21 shows a schematic of a compressor cylinder. The suction flow
qS enters the cylinder, and the discharge flow q D exits the cylinder.

Fig. 14.21 (after [14.8])

The magnitude and shape of the flow pulses through the compressor valves
are determined by the physical, geometrical and mechanical characteristics of the
compressor (rotational speed, bore, stroke, loading, compression ratio, etc.).
The velocity of the piston is approximately sinusoidal in shape, due to the
finite ratio of connecting rod length to crank radius. Since the flow is based on the
product of the piston velocity and the piston swept area, the shape of the discharge
262 DYNAMICS OF MACHINERY

flow curve at the piston face is of the same shape as the piston velocity curve. Two
simplified examples are shown in Figs. 14.22 and 14.23.

Fig. 14.22 (from [14.8])

Figure 14.22, a shows the discharge valve flow versus time for a single-
acting cylinder. During compression, the suction and discharge valves are closed.
When the pressure in the cylinder reaches the discharge back pressure, the discharge
valve opens, and the flow versus time (or crank angle) wave through the valve has
the shape of the corresponding portion of the piston velocity curve (a quarter of a
sinusoid for l r = ). As the cylinder reaches the top dead center, the discharge
valves close, and the flow returns to zero.

Fig. 14.23 (from [14.8])


14. RECIPROCATING MACHINES 263

A frequency analysis of the flow wave is shown in Fig. 14.22, b. Due to the
repetitive action of the compressor cylinder, excitation is generated only at discrete
frequencies, which are integer multiples of the running speed. The highest harmonic
amplitude occurs at 1 running speed (for a single cylinder end), with the levels
decreasing at higher harmonics.
For a double acting cylinder ( l r = 5 and no valve loses), the flow versus
time diagram contains two flow slugs slightly different and not 180 0 apart in time
(Fig. 14.23, a). The cylinder produces flow excitation at all integer harmonics of
running speed as shown in Fig. 14.23, b.
Assuming no interaction between the piping (i.e., no reflected acoustic
waves), the pressure wave out of the cylinder takes on a shape, as a function of the
crank angle, as shown in Fig. 14.24.

Fig. 14.24 (from [14.11])

Figure 14.25, a shows the p-V diagram for a cylinder which is not affected
by pulsations, while Fig. 14.25, b shows a diagram distorted due to pulsation. For
this type of p-V card, the discharge pressure is higher than desired, and the suction
pressure is lower. The valve opening and closing times are also distorted. The
capacity is lower than calculated for the ideal case, resulting in decreased efficiency.

a b
Fig. 14.25 (from [14.11])
264 DYNAMICS OF MACHINERY

Actual dynamic pressure data taken from a natural gas compressor is shown
in Fig. 14.26 as pressure-volume (p-V diagram) and pressure-time data.

Fig. 14.26 (from [14.11])

Notice that the ideal p-t wave (Fig. 14.24) and the actual p-t wave (Fig.
14.26) are definitely non-sinusoidal, which results in pressure pulsations at the
higher harmonic frequencies, as seen in the overlaid frequency spectrum. Comparing
the p-V card with an ideal p-V diagram in Fig. 14.25 one can notice the dynamic
character during the discharge (top of the curve). This distortion of the p-V diagram
14. RECIPROCATING MACHINES 265

comes from acoustic resonances of the discharge piping as the pressure pulses are
reflected back into the cylinder. Strong acoustic responses of the piping can distort
the p-V card resulting in compressor overloading.
The complex interaction which occurs between the piping and compressor
can cause a variety of actual p-V cards among different cylinders and compressors.
Considering the phasing of pressure waves resulting from multiple compressor
cylinders operating with head-ends and crank-end pockets, the shape and frequency
content of pressure-time waves can become very complex. The interaction of the
piping with the cylinders further distorts the picture since these pressure pulses can
excite acoustic natural frequencies (resonance).

14.3.3.2 Acoustic resonance phenomena


The flow pulses caused by the reciprocating action of the compressor create
pressure pulses or waves that move through the piping system. As the disturbance
propagates through the medium, portions of the gas are alternately compressed or
expanded from the equilibrium state.
The wavelength of the pressure wave is
a
= , (14.24)
f
where a is the acoustic velocity and f = 2 is the frequency. This equation
describes the spatial distribution of pressure maxima and minima of the acoustic
wave.
Standing waves
In order for acoustic or pulsation waves to reinforce and result in resonance,
reflections of acoustic waves are necessary. Full reflections occur at closed or open
ends. An acoustic compression pulse is reflected by a closed end as a compression
pulse; an open end will reflect it as a rarefaction pulse. Partial reflections occur at
pipe section discontinuities. Pulsations can cause pressure forces at a restriction such
as reducers, elbows, pipe caps, orifices or partially closed valves.
The superposition of an incident wave and a reflected wave, being the sum
of two waves travelling in opposite directions, will give rise to a standing wave.
Acoustic standing waves are like the natural modes of mechanical vibrating systems.
They are defined by a natural frequency and two distinct mode shapes, one for
pressure and one for velocity.
The acoustical response in the piping is a function of both the mechanical
properties of the compressor, the thermo physical properties of the gas, and the
acoustical circuit defined by the attached piping. When a particular harmonic of
running speed is near or coincident with an acoustical natural frequency, the acoustic
266 DYNAMICS OF MACHINERY

response (dynamic pressure amplitude) is amplified. These resonances can be simple


organ pipe type resonances or complex modes involving all of the piping.
Figure. 14.27 provides the pressure and velocity mode shapes for the second
mode of a closed-closed pipe.

Fig. 14.27 (from [14.11])

At the pressure nodes, pulsation pressures are minimal. If a pressure


transducer were placed at these nodes, no pulsation pressure would be detectable.
The velocity mode shape is shown by the bottom trace in Fig. 14.27. The
velocity amplitude is a maximum at the pressure nodes (maximum kinetic energy)
and zero at the pressure maxima (maximum potential energy), except at the piston
surface, which is not a true velocity node due to piston motion. Because of the
resonant condition, the gas velocity at the velocity maxima may be greater than the
piston velocity.
When pipe lengths are equal to multiples of the wavelength, resonance can
occur. In addition, pipe resonances can occur when the lengths coincide with one
half or one quarter of the wavelength, with the right combination of end conditions
(open or closed).
Most practical piping is open at least at one end. The compressor discharge
line ends in a bottle or a manifold. Likewise, the suction pipe almost always begins
with an open end. According to API 618 [14.8], if the diameter reduction is two-to-
14. RECIPROCATING MACHINES 267

one, a contraction can be treated as a closed end. If a pipe is connected to another


pipe having a diameter that is at least twice as large, it can be considered to be open-
ended. Closed-closed modes are occasionally encountered in pulsation control
bottles and acoustic filters.

Half wave resonance


For half wave resonances, both end conditions must be the same, i.e. open-
open or closed-closed. The first three pressure mode shapes are shown in Fig. 14.28.
Resonances occur at multiples of the half-wave frequency. The acoustic resonance
frequency is
na
f = , (14.25)
2L
where L is the effective length of pipe and n = 1, 2 , 3,... The length should be
corrected for entrance and exit effects [14.11].

Fig. 14.28 (from [14.8])

Quarter wave resonance


For quarter wave resonances, end conditions must be opposite, i.e. one open
end and one closed. The first three pressure mode shapes for an open-closed pipe are
depicted in Fig. 14.29. The acoustic resonance frequency is
na
f = , (14.26)
4L
where L is the effective length of pipe and n = 1, 3, 5,... (odd integers). The length
should be corrected for end effects.
268 DYNAMICS OF MACHINERY

Fig. 14.29 (from [14.8])

A quarter-wave resonance can cause erroneous dynamic pressure


measurements when a pressure transducer is connected to a main line with a short
nipple and a valve forming a quarter-wave stub. When its length is tuned to the
pulsations in the main line, the needle of the pressure gage will wobble or indicate
severe pressure variations which do not actually exist in the main line [14.11].

Acoustic resonances
The existence of quarter and half-wave modes alone do not constitute
resonances. Resonance occurs when a compression wave is generated at a frequency
equal to an acoustical natural frequency. The build up in amplitude occurs because a
reflected wave arrives at the proper time to reinforce the wave at the compressor.
The arrival of the reflected wave is dependent upon the path length of the piping
elements. Therefore, the standing wave pattern amplitude is reinforced so that the
actual maximum pulsating wave amplitude is substantially greater than the induced
level. Since the large pressure amplitudes are the element of most concern, the
pressure antinodes are the areas of concern.
Figure 14.30 illustrates how the response of an acoustic system varies with
the excitation frequency. The figure shows a piston operating in a pipe with a closed
end opposite the piston. Since this system behaves as a closed-closed pipe, the
resonant frequencies are n a 2 L , which for the given dimensions occur at 20, 40, 60,
80, and 100 Hz. The plot in the figure represents the pressure amplitudes at point A,
located at the piston, as the piston excitation frequency is varied from 0 to 100 Hz.
As the frequency increases, the resonant amplitudes decrease. This is
because the lowest modes, starting with the fundamental, have the most energy and,
thus, are more dangerous. The responses at the nonresonant frequencies are small but
nonzero.
At points A and C, the peak pressures at resonance are approximately equal,
while at point B, the only resonances observed are at 40 and 80 Hz, which are the
even-numbered modes, for which point B is a pressure node.
14. RECIPROCATING MACHINES 269

Fig. 14.30 (from [14.12])

If no damping is present, the pressure fluctuations at the antinodes would,


theoretically, be infinite. Actual piping systems have acoustic damping as a result of
the following mechanisms: a) viscous fluid action (intermolecular shearing), b)
transmission (lack of total reflection) at line terminations, junctions, diameter
changes, and c) piping resistance (pipe roughness, restrictions, orifices).
Therefore, damping of acoustic modes may be accomplished by placement
of resistance elements, such as an orifice, which will work most effectively at
velocity maxima.
Note that acoustic resonances of piping systems for constant-speed
compressors can usually be adjusted to detune them from the compressor harmonic
frequencies by locating the resonances between harmonics and avoid pulsation
amplification. However, for a variable speed compressor, detuning resonances
becomes impossible and requires an acoustical filter.

Lumped acoustic elements


Certain components of the piping system may be viewed as lumped
elements. Lumped acoustic elements are: a) the acoustic compliance, represented by
a volume which acts as a stiffness or storage element and opposes a change in
applied pressure; b) the acoustic inertance, characterizing a mass of gas contained in
a relatively small diameter pipe which, when forced into motion, has the property of
opposing a change in volume velocity; and c) the acoustic resistance, an orifice
which dissipates energy when the gas is forced through the smaller diameter
opening. These acoustic elements are directly analogous to the mechanical stiffness,
270 DYNAMICS OF MACHINERY

mass and damping elements, or to the electrical capacitance, inductance and


resistance. Moreover, the acoustic volumetric flowrate is analogous to the
mechanical displacement and the electrical current, while the acoustic pressure is
analogous to the mechanical force and the electrical voltage.
Systems consisting of only the two reactive elements (a gas inertia and
compliance, a mass-spring system, an L-C circuit) are one degree of freedom
oscillators. The equations of their natural frequencies are essentially equivalent.

14.3.3.3 Pulsation control in piping systems


The control of pulsations can be accomplished by judicious use of filters and
tuned absorbers. Both of these types of acoustic elements amplify pulsations at their
resonant frequencies; however, by their careful design and application, they can be
used to attenuate energy.

Fig. 14.31 (from [14.11])


14. RECIPROCATING MACHINES 271

A filter is designed such that its resonance is located at a frequency where


the pulsation energy does not exist, and the maximum attenuation frequencies are
located where the pulsation energy does exist. A tuned absorber uses a resonant
component to take energy from the main system and relocate the resonance where it
is controllable. The attenuation characteristics of four acoustic components are given
in Fig. 14.31.
The fundamental acoustic properties of piping components defined
previously can be used to describe methods of controlling pulsations in piping
systems. These methods include: a) use of side-branch resonators (Helmholz
resonators), b) use of a surge volume for compressor cylinders, c) use of baffles and
choke tubes within surge volumes, and d) use of dissipative components such as
orifice plates, perforations, etc.

Helmholtz resonators
The side-branch resonator (Fig. 14.31, a) can be an effective dynamic
absorber in an acoustic system. It is a choke-volume system which, attached to a
pipe, creates an antiresonance. For long neck resonators
a A
fr = , = . (14.27)
2 V L +1 2 A

Its use should be limited to constant speed systems where the resonator is
tuned to a major frequency of pulsation. The resonator will pull pulsation energy out
of the main line. However, the pulsations will be amplified in the resonator. It must
be mechanically restrained to prevent vibrations in the cantilever mode.
Acoustic resonances of the nozzles to a filter bottle in a reciprocating gas
compressor generally have a strong response, because the pulsating flow from the
cylinder flows directly into the nozzle. The nozzle and cylinder are similar to the
elements of a Helmholtz side branch resonator, where the cylinder internal passages
and clearance pockets form the volume and the gas in the nozzle is the oscillating
mass. The nozzle resonant frequency can be estimated from the equation for the
Helmholtz resonator (14.27). Because the cylinder generates strong pressure pulses
over a range of harmonic frequencies, the probability is high that one of the
harmonic frequencies will match the nozzle resonant frequency.
In high flow, multi-cylinder compressor stages where several cylinders
discharge into a surge volume or filter bottle, nozzle resonances may be near the
bottle passbands simply because of typical dimensions.

Surge volumes
A surge volume (Fig. 14.31, c) can be quite effective in attenuating
pulsations of a compressor, particularly if it can be located near the discharge flange.
272 DYNAMICS OF MACHINERY

Although it has qualities similar to a filter, it is not a true filter. The maximum
attenuation of the outlet pulsations occur away from the resonant frequency.
The attenuation factor is approximately

2
pin 1 1
= 1+ m , (14.28)
pout 4 m

D2
where the expansion ratio m = and d inlet diameter, D bottle diameter.
d2
The resonant frequency is

a 1 + 2
fr = , (14.29)
2 V
where
Aj
j = , (14.30)
L j + 1 2 Aj

and a acoustic velocity, V chamber volume, L j - choke tube lengths, A j - choke


tube area.
Economic considerations limit the size of surge bottles, and therefore impose
practical limits on the degree of the overall acoustic attenuation which can be
achieved. While the surge volume reduces the outlet pulsations, the pulsations within
the volume can be at resonance, so that it must be adequately supported.

Reactive acoustic filters


An acoustic filter is designed to reflect as much of the incident energy as
possible and, thereby, transmit as little as possible. One of the simplest low-pass
filters is a pipe with either a constriction or an added enlarged section. They are
commonly used in the design of automobile mufflers, gun silencers, and sound-
absorbing plenums used in the ventilating systems. Another filter is a bottle with an
internal baffle and a choke tube (Fig. 14.31, d).
One of the most common types of acoustic filters used in reciprocating
equipment involves the use of two volumes joined by a relatively small diameter
pipe, to create a volume-choke-volume filter. Figure 14.32 shows various forms of
the volume-choke-volume filter. If properly designed, the filter components behave
as basic lumped acoustic elements, i.e. the bottles behave as acoustic compliances
and the choke acts like an acoustic inertia. These lumped characteristics are valid as
long as the excitation frequencies are below the open-open resonant frequency of the
choke tube and the closed-closed resonant frequencies of the bottles.
14. RECIPROCATING MACHINES 273

Fig. 14.32 (from [14.13])

These devices have pulsation response characteristics at low frequencies like


that shown in Fig. 14.33. At frequencies above its resonance frequency, known as
the Helmholtz frequency f H , transmitted pulsation levels drop off rapidly.

Fig. 14.33 (from [14.14])

The Helmholtz resonant frequency of an ideal filter with no piping attached


is given by

Ac 1
fH =
a + 1 , (14.31)
2 Lc V1 V 2

where Lc = L c + 0.6 d c , d c - the choke diameter, L c - the length of choke tube,
A c - the area of choke tube, V 1 - the volume of primary bottle and V 2 - the volume
of secondary bottle.
For equal volumes, the Helmholtz frequency is approximately [14.11]
274 DYNAMICS OF MACHINERY

a c
fH = . (14.32)
2 V1

where the acoustical conductivity c is given by

Ac
c = . (14.33)
Lc + 1 2 Ac

In addition to the Helmholtz resonance of a two-chambered filter, other


internal resonances of elements such as compressor internals, nozzles, etc., may exist
and will have the effect of passing particular frequencies. They produce peaks
(pass bands) in the high frequency portion of the filter frequency response (Fig.
14.31, d). The number of pass bands occurring can be minimized by making the
effective length of the choke equal to the lengths of the bottles.
In low mole weight gas systems, reactive filters are impractical due to the
high speed of sound values. Pulsation control can be accomplished through the use
of ample surge volumes and resistive or pressure drop elements.

Dissipative devices
The most frequently employed dissipative device is the orifice plate. Forcing
the flow through small openings, substantial pressure drops can be obtained.
Orifices are the cheapest pulsation control device and the one most amenable
to quick fix solutions. If a pulsation problem is uncovered in the field, it is
relatively easy to add an orifice to the system (usually at a flange), compared to
adding an accumulator or acoustic filter. Of course, the likelihood that simply
placing an orifice in the most available location will solve the problem is not very
high. Their sizing and placement has to be guided through an accurate acoustic
simulation of the entire system.
The effectiveness of dissipative devices is frequency-dependent. Their
performance is better at high frequencies. Orifices are most effective when placed at
or near the location of a velocity antinode in the mode shape of the mode to be
attenuated. In order to provide some appreciable damping to the system, the orifice
diameter should be not larger than one-half of the pipe diameter [14.15].

14.3.4 Piping vibration

To avoid potential vibration problems in compressor and piping systems, the


single most important concept is to avoid resonance. This is best achieved by
focusing on two factors of design: a) try to minimize the magnitude of the harmonic
forcing functions as described in Section 14.3.3, and b) make revisions to the piping
14. RECIPROCATING MACHINES 275

support structure to change its natural frequency or to revise the piping layout to
change the location where the harmonic force is applied.

14.3.4.1 Design separation margins for natural frequencies


Field experience has confirmed that a 10% shift of the natural frequency
away from resonance results in acceptable vibration levels (a reduction by a factor of
five to ten, depending on damping) when the primary cause of the high vibration was
resonance. Allowing for 10% uncertainty in natural frequency predictions, a 20%
design separation margin is recommended.
The operation of double acting compressors results in significant inherent
forces at one and two times rotational speed. Resonance could be avoided at these
frequencies if the predicted natural frequencies should be at least 20% above two
times the compressor rotational speed.
The two design separation margin guidelines adopted by the API 618
standard are: a) minimum predicted natural frequencies should be greater than 2.4
times maximum compressor rotational speed, and b) predicted natural frequencies
should be separated at least 20% from frequencies with significant excitation forces.

14.3.4.2 Piping span natural frequencies


Actual piping span natural frequencies deviate from the theoretical beam
natural frequencies, since the configurations that exist in typical plant piping have
boundary conditions that differ from ideal values. Nevertheless, ideal beam theory
gives a valuable starting point for understanding piping vibration behavior.
Simplified natural frequency formulas can be used to evaluate the piping system
with a minimum of detailed computer analyses.

Straight piping spans


For a straight uniform piping span, the natural frequency can be calculated
using the following relationship
EI
f = , (14.34)
2 Al 4
where: f is the span natural frequency, E Youngs modulus, density, A pipe
cross section area, I cross section moment of inertia, l span length,
frequency factor.

By substituting the material properties for steel, E = 30 106 lb in 2 and


= 0.283 lb in 3 , equation (14.34) can be simplified [14.16] to
k
f = 223 , (14.35)
L2
276 DYNAMICS OF MACHINERY

where k = I A is the radius of gyration, inches, and L is the length of span, ft (!).

Note that this equation does not include the weight of the fluid and the
insulation. The frequency factors, , for calculating the first two natural frequencies
for ideal straight piping spans are given in terms of the overall span length in Fig.
14.34.

Fig. 14.34 (after [14.17])


14. RECIPROCATING MACHINES 277

Piping bends
The natural frequencies of selected pipe configurations with piping elbows
(L-bends, U-bends, Z-bends, and three dimensional bends) were calculated using the
ANSYS finite element program to generate frequency factors for the first two modes
of vibration. Values are given in Fig. 14.34 for bends with equal span length and a
total length L, calculated using a curved beam (elbow) element at corners. Frequency
factors for a range of bend aspect ratios are published in reference [14.17].

Effect of concentrated masses


Applying Rayleighs method, the first natural frequency of a beam with a
concentrated mass can be calculated by
f
fw = , (14.36)
P
1+
W
where f w is the pipe span natural frequency with concentrated weight, f pipe span
natural frequency without concentrated weight, P concentrated weight, W weight
of beam span, - weight correction factor.
Weight correction factors to be used in calculating the natural frequencies of
ideal piping spans for weights at the maximum deflection locations are given in Fig.
14.35. If two weights are located in one span, Dunkerleys formula can be used to
calculate the effect of the second weight.
The frequency for one weight P1 is

f
f1 = . (14.37)
P1
1+
W
If the second weight in the span is considered by itself, the equation is
f
f2 = . (14.38)
P2
1+
W
The frequency for the span with both weights is given by the following
equation
1
f 12+ 2 = . (14.39)
1 1 1
+
f12 f 22 f 2

Correction factors for non-ideal end conditions are suggested in [14.17].


278 DYNAMICS OF MACHINERY

Fig. 14.35 (from [14.18])


14. RECIPROCATING MACHINES 279

With the above approach, clamp spacings can be selected which ensure that
the piping spans will be resonant above some selected frequencies. Table 14.2
[14.13] gives the recommended maximum clamp spacing for minimum natural
frequencies from 10 to 50 Hz.
For complicated piping configurations a finite element analysis is required.
These include flange flexibilities, flexibility of structures on which pipe supports are
mounted, branch connection flexibilities, dynamic pipe-soil interaction, compressor
frame flexibility, etc.
Table 14.2

14.3.4.3 Allowable vibration amplitudes


In cases where high vibrations are noticed, the engineer must have some
simple criteria to judge the severity of vibrations. Screening criteria have been
developed to eliminate the necessity of a comprehensive analysis of every span in
the piping system.

API 618 Design Vibration Guideline


The piping system design vibration criteria adopted by API Standard 618 is
shown in Fig. 14.36.
The figure is based on: a) a constant allowable vibration amplitude of 0.5
mm peak-to-peak for frequencies below 10 Hz, and b) a constant allowable vibration
velocity of approximately 32 mm s peak-to-peak for frequencies between 10 and
200 Hz. In fact, no single vibration guideline can completely account for the wide
variation in geometry and supporting of actual compressor and piping systems. The
adopted design vibration limit balances between typically acceptable vibration levels
for large slow speed versus smaller high speed compressor piping systems.
280 DYNAMICS OF MACHINERY

Fig. 14.36 (from [14.8])

Vibration displacement amplitude as a function of stress


The severity of piping span lateral vibration displacement amplitudes can be
assessed by comparing the maximum resonant vibration-induced dynamic stresses to
an allowable fatigue limit stress.
The low cycle fatigue curves for carbon steel given in the ASME USAS
B31.7-1969 (Fig. 14.37) can be used to obtain an acceptable fatigue limit stress
[14.20].

Fig. 14.37 (from [14.19])


14. RECIPROCATING MACHINES 281

The ANSI/ASME Code OM3-1987 [14.21] uses this stress versus cycles-to-
failure curve as a basis for specifying criteria for evaluating the vibration-induced
stresses in nuclear power plant piping for preoperational and startup testing. The
code defines the allowable fatigue stresses as 0.8 times the allowable alternating
stress intensity at 10 6 cycles which is 13,000 psi zero-to-peak (89.5 MPa).
The vibration-induced dynamic stresses in a piping span vibrating at
resonance has been shown to be related to the maximum vibration amplitude in the
span [14.18]. The relationship is given by the equation below
D
= Kd y (SCF ) , (14.40)
L2
where dynamic stress, psi, K d deflection stress factor, y maximum vibration
amplitude measured between nodes (normally at supports), mils, D outside pipe
diameter, inches, L span length, ft, SCF stress concentration factor
( 1 psi = 6.895 kPa , 1 ft = 0.3048 m , 1in = 25.4 mm ).

The deflection stress factor K d is a function of the boundary conditions and


the vibration mode shape at resonance. The deflection stress factors for the first two
modes of the ideal Bernoulli-Euler beams and the piping configurations with elbows
are given in Fig. 14.34 for equal leg lengths and in reference [14.17] for various
values of the leg length ratio.
The allowable vibration amplitude can be calculated based on the fatigue
limit using the relationship [14.18]
a L2
ya = , (14.41)
(SCF )(SF ) K d D
where a allowable stress, psi, K d deflection stress factor, SCF stress
concentration factor and SF safety factor.
If the API 618 allowable of 13,000 psi zero-to-peak is used as the endurance
limit combined with a stress concentration factor of 4.33, a safety factor of 2, and a
stress deflection factor of 3000 (applicable for a fixed-fixed pipe), the allowable
vibration in peak-to-peak mils can be calculated. Equation (14.41) becomes
L2
ya = . Rule of Thumb (14.42)
D
This can be used conservatively as a screening criterion for straight runs of
piping or for piping with bends. Note that the pipe diameter is measured in inches,
while the span length is measured in feets. This criterion is overly conservative for
cantilever beams. If the measured vibrations exceed the screening criterion, the
vibration induced stresses are not necessarily excessive, and more detailed
calculations using computer programs are required.
282 DYNAMICS OF MACHINERY

Vibration velocity amplitude as a function of stress


In a piping span vibrating at resonance, it is also possible to relate the
maximum stress to the measured velocity [14.16]. In order to develop a closed-form
solution of the dynamic stress as a function of the velocity, the radius of gyration has
to be expressed as a function of the outside diameter of the pipe. A comparison of
the radius of gyration for different sizes of pipe versus the diameter shows that, for a
significant range of pipe sizes, this is approximately 0.34 D0 , where D0 is the
outside pipe diameter. By making this substitution for the radius of gyration, the
stress in an ideal beam can be expressed as a constant K v , referred to as the velocity
stress factor, multiplied by the maximum velocity measured in the piping span,
times the stress concentration factor
= K v v SCF . (14.43)

where dynamic stress, psi, v is the maximum velocity in the pipe span, in sec .
Some velocity stress factors are given in Fig. 14.34 for ideal straight spans and
piping bends with equal legs. Values for different leg length ratios are presented in
reference [14.17].
The actual velocity is a function of the fatigue limit and is given in equation
(14.44) where a safety factor (usually 2) is included to account for system unknowns
a
v= . (14.44)
K v SF SCF
Based on an allowable fatigue limit of 13,000 psi zero-to-peak, a maximum
velocity stress factor of 318, a stress concentration factor of 5, and a safety factor of
2, the allowable zero-to-peak velocity is equal to [14.17]
13000
va = = 4 in sec . (14.45)
318 2 5
For spans with weights, the allowable velocity is [14.22]
va = 2 in sec 50 mm s . (14.46)

Dynamic strain criteria


For typical piping with an ultimate tensile strength of less than 80,000 psi,
the fatigue limit from ASME B31.7 is 26,000 psi peak-to-peak. Since the stress is
equal to the dynamic strain times the modulus of elasticity, the allowable strain
would be 866 10 6 in in . If a stress concentration factor of 4.33 and a safety factor
of 2 is used, a safe allowable strain reading for a gage mounted near the area of high
stress concentration would be 100 10 6 in in or 100 microstrain.
The guidelines for the interpretation of the strains are the following [14.19]:
14. RECIPROCATING MACHINES 283

Strain: < 100 p-p acceptable.


Strain: 100 < < 200 p-p marginal. (14.47)
Strain: 200 < p-p failure possible.

14.3.4.4 Solutions to piping vibration problems


Since the span natural frequency is an inverse function of the square of the
span length, the most effective way to solve a mechanical resonance is to add pipe
restraints, such as piers, supports or clamps to shorten the vibrating span. Many
times, temporary bracing with hydraulic jacks, wooden beams and wedges can be
used to confirm that a support at a particular location will reduce the vibrations.
Some of the general guidelines which can be used in selecting modifications
to detune the mechanical resonances are outlined below [14.12]:
1. Pipe supports and clamps should be installed on one side of each bend, at
all heavy weights, and at all piping discontinuities.
2. The support and clamp stiffness should be adequate to restrain the shaking
forces in the piping to the desired amplitudes and should be greater than twice the
basic span stiffness in order to effectively enforce a node at the support location.
3. Vents, drains, bypass, and instrument piping (appurtenances) should be
braced to the main pipe to eliminate relative vibration between the small-bore piping
and the main piping.
4. Restraints, supports, or gussets should not be directly welded to pressure
vessels or the piping unless they are subjected to the appropriate heat treatment. It is
more desirable to add a saddle-type clamp around the pipe and weld the braces to the
clamp.
5. To resist vibration, the piping clamps should have contact with the pipe
over 180 degrees of the circumference. Rubber or gasket-type material can be used
between the clamp and the pipe to improve the contact.
6. The piping span natural frequency should not be coincident with the
excitation frequencies.

14.3.4.5 Design procedure using reactive pulsation control


The use of reactive filtering in conjunction with control of mechanical
natural frequencies results in a safe margin between significant pulsation-induced
forces and mechanical natural frequencies. The procedure for designing reactive
filters consists of the following steps:
1. Determine choke diameter and approximate length based on allowable
pressure drop.
284 DYNAMICS OF MACHINERY

2. Determine volume-choke-volume filter design to filter all harmonics of


running speed. Generally, the filter frequency is set at 50-80 percent of 1 running
speed for heavy gases, or between 1 and 2 running speed for lighter gases.
3. Perform pulsation simulation to determine pulsation levels and
( )
acceptability of filter design. Determine maximum frequency f p of significant
pulsation and force in piping (Fig. 14.45, d).
4. Determine minimum allowable mechanical natural frequency ( f m ) based
( )
on f p . Set f m 1.5 f p .

5. Locate vibration restraints near all concentrated masses (e.g., valves).


6. Use pipe support span tables (Table 14.2) to determine additional support
locations based on f m .
7. Determine minimum stiffness (k) of each support: k 2 lateral span
2 48 E I
stiffness = ( l = support span).
l3
Use of this acoustic filtering concept in conjunction with control of
minimum piping mechanical natural frequencies provides a high level of confidence
that resonance will be avoided.

References

14.1. Den Hartog, J. P., Mechanical Vibrations, 4th ed., Dover, New York, 1985.
14.2. Thompson, W. T., Vibration. Theory and Applications, George Allen &
Unwin, London, 1966.
14.3. Magrath, H. A., Rogers, O. R., and Grimes, C. K., Shock and vibration in
aircraft and missiles, Ch. 47 in Shock and Vibration Handbook, C. M. Harris and
Ch. E. Crede, eds., McGraw-Hill, New York, 1961.
14.4. Crandall, S. H., Rotating and reciprocating machines, Ch. 58 in Handbook of
Engineering Mechanics, W. Flgge, ed., McGraw-Hill, New York, 1962.
14.5. Richart, F. E. Jr., Hall, J. R. Jr. and Woods, R. D., Vibrations of Soils and
Foundations, Prentice Hall, Englewood Cliffs, N.J., 1970.
14.6. Maas, H. and Klier, H., Krfte, Momente und deren Ausgleich in der
Verbrennungskraftmaschine, Die Verbrennungkraftmaschine, Neue Folge, Band
2, Springer, Wien, 1981.
14. RECIPROCATING MACHINES 285

14.7. Rade, M., Diagnosis of an auxiliary diesel engine vibration problem with
signature analysis, Machine Vibration, vol.1, 1992, p.58-63.
14.8. * Reciprocating Compressors for Petroleum, Chemical, and Gas Industry
Services, ANSI/API Standard 618, 5th ed., 2007.
14.9. Bloch, H. P., Compressors and Modern Process Applications, Wiley, New
York, 2006.
14.10. Lifson A. and Dube, J. C., Specifying reciprocating machinery pulsation and
vibration requirements per API-618, American Gas Association
Distribution/Transmission Conference, Las Vegas, Nevada, May 4-6, 1987.
14.11. Wachel, J. C. et al, Vibrations in Reciprocating Machinery and Piping
Systems, Engineering Dynamics Incorporated, Technical Report EDI 85-305, 2nd
ed., 2nd Printing, 1988.
14.12. Wachel, J. C. and Tison, J. D., Vibrations in reciprocating machinery and
piping systems, Proc. 23rd Turbomachinery Symposium, Texas A&M University,
College Station, Texas, 1994, p.243-272.
14.13. Atkins, K. E., Pyle, A. S. and Tison, J. D., Understanding the pulsation and
vibration control concepts in the new API 618 Fifth Edition, 2004 Gas Machinery
Conference, Albuquerque, New Mexico, Oct. 4-7, 2004
14.14. Corbo, M. A. and Stearns, Ch. F., Practical design against pump pulsations,
Proc. 22nd International Pump Users Symposium, Turbomachinery Laboratory,
Texas A&M University, Feb.28-March 3, 2005, p.137-177.
14.15. Price, S. M., and Smith, D. R., Sources and remedies of high-frequency piping
vibration and noise, Proc. 28th Turbomachinery Symposium, Texas A&M
University, College Station, Texas, 1999, p.189-212.
14.16. Wachel, J. C., Piping vibration and stress, Proc. Machinery Vibration
Monitoring and Analysis Seminar , Vibration Institute, April 1981.
14.17. Wachel, J. C., Morton, S. J. and Atkins, K. E., Piping vibration analysis,
Proc. 19th Turbomachinery Symposium, Texas A&M University, College
Station, Texas, 1990, p.119-134.
14.18. Wachel, J. C, Displacement method for determining acceptable piping
vibration amplitudes, International Pressure Vessels and Piping Codes and
Standards, PVP-vol.313-2, ASME 1995, p.197-208.
14.19. Wachel, J. C., Field investigations of piping systems for vibration-induced
stresses and failures, Pressure Vessels and Piping Conference, Orlando, Florida,
June 27 July 2, 1982, ASME Bound Volume No.H00219, 1982.
14.20. * Nuclear Power Piping, USAS B31.7-1969 ASME Code, New York,
1969.
286 DYNAMICS OF MACHINERY

14.21. * Preoperational and Initial Startup Vibration Testing of Nuclear Power


Plant Piping Systems, ANSI/ASME Operations & Maintenance Standards/Guides
Part 3, ASME, New York, 1990.
14.22. Wachel, J. C. and Smith, D. R., Vibration troubleshooting of existing piping
systems, Engineering Dynamics Incorporated Report 91903, July 1991.
14.23. * Nivele admisibile de vibraii pentru conducte din instalaii chimice i
rafinrii, Ministerul Industriei Chimice, NTR 11230-85, ICITPR Ploieti, iulie
1985.
Index

Accelerometers 70 Diesel engine 251


Acceptance region plots 111 Discharge bottle 258
Acoustic resonance 265, 268 Discoloration 56
Amplitude demodulation 28 Dresser Clark Chart 194
probability density 18 Dynamic unbalance 207
Angular contact bearings 4
Eddy current transducer 88
API 618 279
Electrical machines 151
API standards 186
Envelope analysis 28
Axial compressors 145
Enveloping 62
Balance quality grades 225, 228 Energy operator 62
Balancing of rotors 203 Errors 54
in N+2 planes 229
Bearing frequencies 6
Fault diagnostics 121
wear 14
Feature extraction 59
Blake Severity Chart 192
Flaking 14
Bode plots 108
Fluid induced instabilities 127
Brinelling 16
FM0 63
Burning 56
FM4 64
Cage damage 15 Fretting 16
Cascade plots 109 Frosting 56
Case crushing 57 Full spectrum plots 104
Centrifugal compressors 142
Galling 55
fans 144
Gas pressure excitation 237
pumps 141
Cepstrum analysis 35, 69
torque 238
Gas turbines 150, 172, 180
Condition indicators 59
Gear errors 54
monitoring 115
Gears 39, 185
Connecting rod 242
Ghost frequencies 48
Contact ratio 42
Glazing 15
Couple unbalance 205, 223
Grooving 16
Coupled industrial machines 170, 178
Cracked shafts 138 Half spectrum plots 104
Crank mechanism 241 Half-wave resonance 267
Creeping 16 Heavy spot 205
Crest factor 17, 62 Helmholtz resonator 271
Cylinder-manifold system 256 High spot 205
Cylinder stretch forces 259 Hydraulic machines 172, 181
Demodulation 63 Indentations 14
Denting 15 Industrial buildings 187
Diagnosis process 120 Inertia torque 239
288 DYNAMICS OF MACHINERY

Influence coefficient method 209, 217 Reciprocating compressors 152, 256


ISO 1940 225 machines 174, 237
ISO 7919 176 Residual unbalance 227
ISO 10816 168 Reverse dial indicator method 156
Rigid rotors 226
Kurtosis 22, 61
Rolling element bearings 1
Laser alignment 159 Root mean square 17, 61
Looseness 135, 137 Rotor-bearing clearance 136
Lumped acoustic elements 269 Rubbing 130
Run-to-failure maintenance 117
M6A 66
M8A 66 Scoring 55
Machine deterioration 115 Scuffing 55
Maintenance strategies 117 Shaft alignment 155
Mass unbalance 204 Shock pulse method 30
Misalignment 123 Shaking forces 260
Modal balancing 232 Sieberts construction 215
Mode shape plots 106 Single cylinder engines 237
Modulation plane balancing 208
amplitude 49 Skewness 21
effects 48 Skidding 2
frequency 51 Smearing 14
Multi cylinder engines 246 Spalling 14, 57
Multiplane balancing 229 Spike energy 25
Spinning 16
NA4 65 Standard deviation 21
NB4 67
Standing waves 265
Oil debris analysis 67 Static unbalance 205, 223
Oil whip 128 Statistical moments 21
whirl 128 Steam turbines 146, 169, 177
Orbits 101 Surge volume 271
Orifice plate 274 Tapered roller bearings 7
Piping vibration 274 Three-trial-mass method 215
Pitting 56 Time base plots 101
Plots 101 Time-frequency analysis 72
polar 108 Time synchronous averaging 59
Predictive maintenance 118 Tooth deflection 46
Preventive maintenance 117 engagement 45
Proximity probes 88, 97 fracture 58
P-V card 263 wear 47
Pulsation analysis 261 Transducers 85
control 270 Trend plots 107
excitation mechanism 261 Two-plane balancing 217
Quarter-wave resonance 267 Unbalance 121
Quasi-static unbalance 206 couples 244, 246
force 243, 246
Radial preload 123 tolerances 225
Rathbone Chart 164
Reactive acoustic filters 272 VDI 2056 166
INDEX 289

VDI 2059 196, 197, 198


Vector balancing 208
Velocity pickups 91, 97
Vibration intensity 188
limits 163
magnitude 80
measurement 75
peak-to-peak 83
severity 81, 195
charts 164
Volume-choke-volume filter 272
Wear 56

You might also like