Sunteți pe pagina 1din 152

1

Vibration knowledge Bank


DLI Engineering Corp.

Introduction
Overview of Maintenance Practice
Introduction to Vibration
Vibration Transducers
The FFT Analyzer
Machine Vibration Monitoring
Machine Vibration Analysis
Machine Diagnosis
Estimating Vibration Severity
Index
Glossary of Terms
Introduction

More:

Foreword
Foreword

The purpose of this book is to serve as a reference text for the maintenance engineer
and technician who are working with state of the art machinery maintenance
technology. Broadly speaking, the subject is the principles of vibration theory and
analysis as they apply to the determination of machine operating characteristics and
deficiencies. The first chapter underscores the importance of vibration analysis in the
field of predictive maintenance and root cause failure analysis. The chapters on
vibration theory and frequency analysis lay the groundwork for the chapter on
machine fault diagnostics based on vibration measurement and analysis. A
systematic approach is used here to guide the reader through a logical sequence of
steps to determine a machine's condition by detailed examination of vibration
signatures.
Some of the terminology used here may not be familiar to all readers, and for this
reason a fairly complete glossary is included as the final chapter. The author wishes
to sincerely thank Mr. Peter Bunker, of VIPAC, in Tasmania, for supplying some
information in the section on amplitude demodulation.
The author of this book welcomes comments and suggestions from readers. Please
address inquiries to:
Ron Bodre
DLI Engineering Corp.
253 Winslow Way West
Bainbridge Island, WA 98110
206-842-7656
FAX 206-842-7667
2




Overview of Maintenance Practice

More:

Survey of Machinery Maintenance Practices
Components of a Maintenance Program
Case Studies
Survey of Machinery Maintenance Practices

Presented here is an overview of maintenance programs and techniques as practiced
in the early 1990s in a wide variety of industrial areas. Most of the information
presented here was collected from shore side industrial plants, but it is equally
applicable to maintenance of shipboard mechanical systems. While the emphasis is
on predictive maintenance, other disciplines are described and evaluated.
Machinery maintenance practices have greatly changed and evolved over the last 15
years, and it is instructive to study this development. We will first look at the basic
goals of any maintenance system:
More:

Maintenance Program Goals
Historical review
Where are we today?
Maintenance Program Goals

The most important goal of any maintenance program is the elimination of machine
breakdowns. Very often a catastrophic breakdown will cause significant peripheral
damage to the machine, greatly increasing the cost of the repair. Complete
elimination of breakdowns is not at present possible in practice, but it can be
approached by a systematic approach to maintenance.
The second goal of maintenance is to be able to anticipate and accurately plan for
maintenance needs. This means spare parts inventories can be minimized and
overtime work largely eliminated. Repairs of mechanical systems are ideally planned
for scheduled plant down times.
Goal number three is to increase plant production readiness by significantly reducing
the chance of a breakdown during operations, and to maintain system operational
capacity through reduced down time of critical machines. Ideally, the operating
condition of all the machines would be known and documented.
The last goal of maintenance is to provide predictable and reasonable work hours for
maintenance personnel.
Historical review

3
In order to gain some perspective on modern maintenance programs, we will look at
the history of maintenance practices a little more closely.
The earliest type of maintenance was run-to-failure, where the machine was run until
a fault caused it to fail in service. This is obviously an expensive approach, with the
major part of the cost being the unpredictability of the machine condition. It is
surprising to learn how much of present day maintenance activity is of this type.
Eventually, maintenance people hit on the idea of periodic preventive maintenance,
where machines are disassembled and overhauled on regular schedules. The theory
is that if machines are overhauled before their expected service life is exceeded, they
will not break down in service. Preventive maintenance has been around for a long
time, but became much more prominent in the early 1980s, as we will see.
In the last ten years, predictive maintenance has become popular, where the
machine is repaired only when it is known to have a fault. Smoothly running
machines are not interfered with, on the theory that you shouldn't "fix it if it ain't
broke".
The most recent innovation in maintenance is called pro-active, and it includes a
technique called "root cause failure analysis", in which the primary cause of the
machine failure is sought and corrected.
We will evaluate these various maintenance philosophies shortly.
Where are we today?

In 1991, an international survey of the majority of types of industrial plants was
made, and it was found that all four of the above mentioned maintenance techniques
were being used, and in about the following amounts:
Over one-half of maintenance hours are spent in a reactive mode, performing
emergency repairs on an unscheduled basis.
Less then ten percent of hours are spent on preventive maintenance.
Less then forty percent of maintenance activity is predictive, and
Very little time is spent on pro-active techniques.
From these numbers, it can be seen that, like Thomas Edison when he invented the
phonograph, we have "barely scratched the surface" in bringing maintenance
practices into the 20th century!
It makes sense for a modern machinery maintenance program to include elements of
all these techniques, and to find out why; we will now take a closer look at them:
Components of a Maintenance Program

More:

Run-to-Failure Maintenance
Periodic Preventive Maintenance
Predictive Maintenance
Benefits of Predictive Maintenance
Pro-active Maintenance
Benefits of Pro-active Maintenance
Run-to-Failure Maintenance
4
Run-to-failure maintenance is sometimes called "crisis maintenance" or "hysterical
maintenance" for good reason. This has been the dominant form of maintenance for
a long time, and its costs are relatively high because of unplanned downtime,
damaged machinery, and overtime expenditure.
In this mode, management and the maintenance department are controlled by the
vagaries of their machines, and the actual status of the overall plant machinery is
only vaguely known. This makes it nearly impossible to plan for maintenance needs,
and what is worse, impossible to predict the state of overall system readiness.
Run-to-failure should be a very small part in a modern program, but there are some
instances where it does make sense. An example is a plant, which employs a great
number of similar machines that are not expensive to replace or to repair. When one
breaks down, others are scheduled to take up the slack and production is not
affected very much.
Periodic Preventive Maintenance

From run-to-failure, we progress to periodic preventive maintenance, which is
sometimes called "historical" maintenance. This is where the histories of each
machine type are analyzed and periodic overhauls are scheduled to occur before the
statistically expected problems occur. It has long been known that most groups of
similar machines will exhibit failure rates that are somewhat predictable if averaged
over a long time. This gives rise to the so-called "Bathtub Curve" which relates
failure rate to operating time, as follows:



If this curve applied to all machines of the group, and if the shape of the curve is
known, preventive maintenance could be used advantageously, but unfortunately,
this is not the case in practice.
Preventive maintenance also includes such activities as changing lube oil and filters,
periodic cleaning and inspecting, etc. Maintenance activity may be scheduled on the
basis of calendar time, machine operating hours, number of parts produced, and so
on. Preventive maintenance became very popular in the early 1980s when small
computers began to be used for planning and tracking maintenance work.
In a famous study of preventive maintenance by United and American Airlines, it was
found that for a large class of rotating machines, the failure rate greatly increased
just after the periodic overhauls -- in other words, the overhaul reduced the
reliability of the machines. It is as if the machine reverts to the beginning of the
bathtub curve after each overhaul.
5
From this study and subsequent observations, it was found that periodic overhauls
result in 20 % to 25 % of startup failures. About ten percent of these can be
attributed to defective new bearings.
It is obvious that preventive maintenance is an inefficient use of resources for most
machines; however, there are cases where it can be used to good effect. Examples
are machines which exhibit wear related to use such as rock and ore crushers, and
machines that are subject to corrosion such as equipment handling caustic
substances.
Predictive Maintenance

The next improvement in maintenance technology was the advent of predictive
maintenance, which is based on the determination of a machine's condition while in
operation. The technique is dependent on the fact that most machine components
will give some type of warning before they fail. To sense the symptoms by which the
machine is warning us requires several types of non-destructive testing, such as oil
analysis, wear particle analysis, vibration analysis, and temperature measurements.
Use of these techniques to determine the machine condition results in a much more
efficient use of maintenance effort compared to any earlier types of maintenance.
Predictive maintenance allows plant management to control the machinery and
maintenance programs rather than vice versa. In a plant using predictive
maintenance, the overall machinery condition at any time is known, and much more
accurate planning is possible.
Predictive maintenance utilizes many different disciplines, by far the most important
of which is periodic vibration analysis. It has been shown many times over that of all
the non-destructive testing that can be done on a machine, the vibration signature
provides the most information about its inner workings.
Certain machines, which would affect plant operations adversely if they were to fail,
can be subjected to continuous vibration monitoring, in which an alarm is sounded if
the vibration level exceeds a predetermined value. In this way, rapidly progressing
faults are prevented from causing catastrophic failures. Most modern turbine-driven
equipment is monitored in this way.
Oil analysis and wear particle analysis are important parts of modern predictive
programs, especially in critical or very expensive equipment. Thermography is the
measurement of surface temperature by infrared detection, and is very useful in
detecting problems in electrical switchgear and other areas where access is difficult.
Motor current signature analysis is another technique that is very useful in detecting
cracked or broken rotor bars while the motor is in operation, and electrical surge
testing of motor stators is used for detecting incipient electrical insulation failure.
Benefits of Predictive Maintenance

The major benefit of predictive maintenance of industrial mechanical equipment is
increased plant readiness due to greater reliability of the equipment. The trending
over time of developing faults in machines can be carefully done so as to plan
maintenance operations to coincide with scheduled shutdowns. Many industries
report from two to ten percent productivity increases due to predictive maintenance
practices. Similar percentages of increased mission readiness are expected in
shipboard systems.
6
Another benefit of predictive maintenance is reduced expenditures for spare parts
and labor. Machines that fail while in service often cost ten times as much to repair
than if the repair were anticipated and scheduled.
A great many new machines fail soon after startup due to built-in defects or
improper installation. Predictive techniques can be used to assure proper alignment
and overall integrity of the installed machine when first brought into service. Many
plants base the acceptance of new machine installations on a clean bill of health as
determined by vibration measurements.
Predictive maintenance reduces the likelihood of a machine experiencing a
catastrophic failure, and this results in an improvement in worker safety. There have
been many cases of bodily injury and even death due to sudden machine failures.

Pro-active Maintenance

The latest innovation in the field of predictive maintenance is so-called pro-active
maintenance, which uses a variety of technologies to extend the operating lives of
machines and to virtually eliminate reactive maintenance. The major part of a pro-
active program is root cause failure analysis, which is the determination of the
mechanisms and causes of machine faults. The fundamental causes of machine
failures can thus be corrected, and the failure mechanisms can be gradually
engineered out of each machinery installation.
It has been known for a long time that imbalance and misalignment are the root
causes of the majority of machine faults. Both of these conditions place undue forces
on bearings, shortening their service life. Rather than continually replacing worn
bearings in an offending machine, a far better policy is to perform precision balance
and alignment on the machine, and then to verify the results by careful vibration
signature analysis.
New Installations

It is also well known that many machines contain defects when newly installed.
These defects can range from improper installation, caused by poor footings and
poor alignment, to defective parts within the machine, such as bad bearings, bent
shaft, and so forth. A pro-active maintenance program will include testing on new
installations for the purpose of certification and verification that the performance is
held to a rigid standard. The same standards are applied to rebuilt and overhauled
equipment.
This type of testing can also lead to the establishment of specific performance
specifications that in many cases are more stringent than the equipment
manufacturer's specifications and tolerances.
An essential part of a pro-active approach is the training of maintenance personnel in
the application of the underlying principles.
Precision Alignment

It has been reported in the TAPPI journal that precision alignment resulted in
extending bearing life by a factor of eight in a large class of rotating machines. Other
reported benefits were a seven percent savings in overall maintenance costs and a
twelve percent increase in machine availability. Machine breakdowns attributed to
misalignment were cut in half.
7
Another benefit of precision alignment is a power saving. A recent study documented
an average of eleven percent power saving by precision alignment in a group of
simple pump-motor assemblies. This is because less power is expended in flexing the
coupling, vibrating the machine, and heating the bearings. The dollar saving in this
case due to reduced power consumption was more than twice the maintenance costs
on these machines!
Benefits of Pro-active Maintenance

A successful pro-active maintenance program will gradually design the problems out
of the machines over a period of time, resulting in greatly extended machine life,
reduced down time, and expanded production capacity. One of the best features of a
pro-active approach is that the techniques are natural extensions of those used in a
predictive program, and they are easily added to existing programs.
It is apparent today that we need a balanced approach to maintenance, including the
appropriate use of preventive, predictive, and pro-active methods, and these
elements are not independent, but should be integral parts of a unified maintenance
program.

Case Studies
Figure of Merit

The DLI Engineering automated diagnostic software system examines all vibration
data from these ships and produces repair recommendations for several hundred
machines on each ship. The repair records and follow-up reports are also kept. This
system diagnoses the machine problems and makes specific recommendations for
repairs. It also assigns to each machine a "figure of merit" (FOM) that is inversely
proportional to the overall condition of the machine. (Some have said it should be
called a "figure of demerit"). The figure of merit scale is normalized such that an
FOM of 100 indicates the machine should be scheduled for repair. Higher values
indicate worse conditions, and lower values indicate acceptable conditions.
The accompanying table summarizes the entire pacific aircraft carrier fleet in terms
of average figure of merit of all the monitored machines from 1986 to 1992.
It is seen that at the start of the program, the average FOM is 111, and in 1992, the
average FOM is less than 90. This means the average machine on any ship in the
PACFLT was in need of repair in 1986, but in 1992, the average machine was in
acceptable condition.
1986 1987 1988 1989 1990 1991 1992
Average
FOM

111

108

103

101

98

92

89
Specific Machine Faults

It is also instructive to look at the types of repair that were requested by the
automated diagnostic system and then performed.

Ship No. of Balance Alignment Bearings Other
8
Machines
CV41,43

1755

22

10

25

43

CV59-67

4877

21

13

25

41

CVN68,69

527

28

15

22

35

AVT16

383

20

19

13

48


Here it can be seen that Balance, Alignment, and Bearing problems account for more
than half of all machine repairs. Of course, these fault categories are interrelated, for
most bearing problems are caused by imbalance and misalignment. It is said that
only a few percent of rolling element bearings achieve their design lifetime.
Perhaps more interesting is the repair recommendation record for the USS America
(CV66), shown below. This is the relation between vibration survey dates and
percentage of machines recommended for overhaul. The Feb. 88, Aug. 90, and Aug.
93 surveys were made just after overhauls.

Feb.
88
Post
OH
Sep.
89
Aug.
90
Post
OH
Mar 91 Aug.
92
Aug.
93
Post
OH
Percent of all
machines
needing repair

12

8

10

7.5

7

13


The fact that the percentages of problems post overhaul were higher than at other
times indicates that the overhauls created more problems than they fixed. This is in
keeping with the earlier airline study that showed high numbers of startup failures
after overhaul.
This pattern is not as apparent on most of the ships surveyed.

Introduction to Vibration


More:

What is Vibration?
Energy and Power Considerations
Linear and Non-Linear Systems
Frequency Analysis
Octave Band and One-Third Octave Band Analysis
Linear and Logarithmic Amplitude Scales

What is Vibration?

In its simplest form, vibration can be considered to be the oscillation or repetitive
motion of an object around an equilibrium position. The equilibrium position is the
position the object will attain when the force acting on it is zero. This type of
9
vibration is called "whole body motion", meaning that all parts of the body are
moving together in the same direction at any point in time.
The vibratory motion of a whole body can be completely described as a combination
of individual motions of six different types. These are translation in the three
orthogonal directions x, y, and z, and rotation around the x, y, and z-axes. Any
complex motion the body may have can be broken down into a combination of these
six motions. Such a body is therefore said to possess six degrees of freedom. For
instance, a ship can move in the fore and aft direction (surge), up and down
direction (heave), and port and starboard direction (sway), and it can rotate
lengthwise (roll), rotate around the vertical axis (yaw), and rotate about the port-
starboard axis (pitch).
Suppose an object were restrained from motion in any direction except one. For
instance, a clock pendulum is restricted from motion except in one plane. It is
therefore called a single degree of freedom system. Another example of a single
degree of freedom system is an elevator moving up and down in an elevator shaft.
The vibration of an object is always caused by an excitation force. This force may be
externally applied to the object, or it may originate inside the object. It will be seen
later that the rate (frequency) and magnitude of the vibration of a given object is
completely determined by the excitation force, direction, and frequency. This is the
reason that vibration analysis can determine the excitation forces at work in a
machine. These forces are dependent upon the machine condition, and knowledge of
their characteristics and interactions allows one to diagnose a machine problem.
More:

Simple Harmonic Motion
Equations of Motion
Dynamics of Mechanical Systems
Vibration Amplitude Measurement
The Concept of Phase
Vibration Units
Summary of Amplitude Units:
Displacement, Velocity and Acceleration
Complex Vibration
Simple Harmonic Motion

The simplest possible vibratory motion that can exist is the movement in one
direction of a mass controlled by a single spring. Such a mechanical system is called
a single degree of freedom spring-mass system. If the mass is displaced a certain
distance from the equilibrium point and then released, the spring will return it to
equilibrium, but by then the mass will have some kinetic energy and will overshoot
the rest position and deflect the spring in the opposite direction. It will then
decelerate to a stop at the other extreme of its displacement where the spring will
again begin to return it toward equilibrium. The same process repeats over and over
with the energy sloshing back and forth between the spring and the mass -- from
kinetic energy in the mass to potential energy in the spring and back.
The following illustration shows a graph of the displacement of the mass plotted
versus time.

10


If there were no friction in the system, the oscillation would continue at the same
rate and same amplitude forever. This idealized simple harmonic motion is almost
never found in real mechanical systems. Any real system does have friction, and this
causes the amplitude of vibration to gradually decrease as the energy is converted to
heat. The following definitions apply to simple harmonic motion:
T = The period of the wave.
The period is the time required for one cycle, or one "round trip" from one zero
crossing to the next zero crossing in the same direction. The period is measured in
seconds, or milliseconds, depending on how fast the wave is changing.
The unit for frequency is
the Hz, named after
Heinrich Hertz, the
German scientist who
first investigated radio.

F = The Frequency of the wave, = 1/T
The frequency is the number of cycles that occur in one
second, and is simply the reciprocal of the period.
Equations of Motion

If the position, or displacement, of an object undergoing simple harmonic motion is
plotted versus time on a graph as shown above, the resulting curve is a sine wave,or
sinusoid, and is described by the following equation:

where d = instantaneous displacement,
D = maximum, or peak, displacement
w= angular frequency, = 2f
t = time
This is the same curve that the sine function from trigonometry generates, and it can
be considered the simplest and most basic of all possible repetitive wave forms. The
mathematical sine function is derived from the relative lengths of the sides of a right
triangle, and the sine wave is actually a plot of the value of the sine function versus
angle. In the case of vibration, the sine wave is plotted as a function of time, but one
cycle of the waveform is sometimes considered to equal 360 degrees of angle. More
will be said about this when we consider the subject of phase.
11
The velocity of the motion described above is equal to the rate of change of the
displacement, or in other words how fast its position is changing. The rate of change
of one quantity with respect to another can be described by the mathematical
derivative, as follows:

where v = instantaneous velocity.
Here we see that the form of the velocity function is also sinusoidal, but because it is
described by the cosine, it is displaced by 90 degrees. We will see the significance of
this in a moment.
The acceleration of the motion described here is defined as the rate of change of the
velocity, or how fast the velocity is changing at any instant:

where a = instantaneous acceleration.
Note here also that the acceleration function is displaced by an additional 90 degrees,
as indicated by the negative sign.
If we examine these equations, it is seen that the velocity is proportional to the
displacement times the frequency, and that the acceleration is proportional to the
frequency squared times the displacement. This means that at a large displacement
and a high frequency, very high velocities result, and extremely high levels of
acceleration would be required. For instance, suppose that a vibrating object is
undergoing 0.1 inch of displacement at 100 Hz. The velocity equals displacement
times frequency, or
,
Acceleration equals displacement times frequency squared, or
a = 0.1 x 10000 = 1000 inches per second per second.
One G of acceleration equals 386 inches per second per second, so this
acceleration is

Now, see what happens if we raise the frequency to 1000 Hz:
, and

Thus, we see that in practice, high frequencies can not be associated with
high displacement levels.
Dynamics of Mechanical Systems

A small compact physical structure, such as a marble, can be thought of as simply a
mass. It will move in response to an external force applied to it, and Newtons laws
of motion will govern its movement. Simply put, Newton's laws dictate that if the
marble is at rest, it will remain at rest unless acted on by an external force, and if in
12
motion it will continue in motion unless acted on by an external force. If it is
subjected to an external force, its acceleration will be proportional to that force.
Most mechanical systems are more complex than a simple mass, and they do not
necessarily move as a whole when subjected to a force. Mechanical systems, such as
rotating machines, are not infinitely rigid, and have varying degrees of flexibility at
different frequencies. As we will see, their motion in response to an external force is
dependent on the nature of that force and the dynamic characteristics of their
mechanical structure, and is often difficult to predict. The disciplines of Finite
Element Modeling (FEM) and Modal Analysis are dedicated to predicting how a
structure will respond to a known force. We will not discuss these fields further, for
they are very complex, but it is instructive to look into how forces and structures
interact if we are to understand the usefulness of vibration analysis of machines.


Vibration Amplitude Measurement

The following definitions apply to the measurement of mechanical vibration
amplitude.



Peak Amplitude (Pk) is the maximum excursion of the wave from the zero or
equilibrium point.
Peak-to-Peak Amplitude (Pk-Pk) is the distance from a negative peak to a positive
peak. In the case of the sine wave, the peak-to-peak value is exactly twice the peak
value because the waveform is symmetrical, but this is not necessarily the case with
all vibration waveforms, as we will see shortly.
Root Mean Square Amplitude (RMS) is the square root of the averageof the squared
values of the waveform. In the case of the sine wave, the RMS value is 0.707 times
the peak value, but this is only true in the case of the sine wave. The RMS value is
proportional to the area under the curve -- if the negative peaks are rectified, i.e.,
made positive, and the area under the resulting curve averaged to a constant level,
that level would be proportional to the RMS value.
13


The RMS value of a vibration signal is an important measure of its amplitude. As
mentioned before, it is numerically equal to the square root of the average of the
squared value of amplitude. To calculate this value, the instantaneous amplitude
values of the waveform must be squared and these squared values averaged over a
certain length of time. This time interval must be at least one period of the wave in
order to arrive at the correct value. The squared values are all positive, and thus so
is their average. Then the square root of this average value is extracted to get the
RMS value.
Average Amplitude, which is
simply the arithmetic average of
the signal level over time, is not
used in vibration measurements,
and we will not consider it further.

The RMS value must be used in all calculations
regarding power or energy in a waveform. An
example of this is the 117 volt AC line. The 117
volts is the RMS value of the voltage, and it is
used in calculations of the wattage (power) drawn
by devices connected to it. Remember that the
RMS value of a sine wave is 0.707 times the peak
value, and this is the only wave form where this is
true. We will see shortly that this is important.

The Concept of Phase

Phase is a measure of relative time difference between two sine waves. Even though
phase is truly a time difference, it is almost always measured in terms of angle,
either degrees or radians. This represents normalization to the time taken by one
cycle of the wave in question, without regard to its true time period.
The phase difference between two waveforms is often called a phase shift. A phase
shift of 360 degrees is a time delay of one cycle, or one period of the wave, which
actually amounts to no phase shift at all. A phase shift of 90 degrees is a shift of 1/4
of the period of the wave, etc. Phase shift may be considered positive or negative,
i.e., one waveform may be delayed relative to another one, or one waveform may be
advanced relative to another one. These conditions are called phase lag and phase
lead respectively.
14


In this example, the lower curve is shifted 90 degrees with respect to the upper
curve. This is a time lag of one-fourth of the period of the wave. You could also say
the upper waveform has a 90 degree phase lead.
Phase can also be measured with reference to a particular time. An example of this is
the phase of an imbalance component in a rotor with reference to a fixed point on
the rotor, such as a key way. To measure this phase, a triggerpulse must be
generated from a certain reference point on the shaft. This trigger can be generated
by a tachometer or some type of optical or magnetic probe that senses a
discontinuity on the rotor, and is sometimes called a "tach" pulse.


Phase of a Rotor

The phase angle can be measured from the reference position either in the direction
of rotation or opposite to the direction of rotation, i.e., phase lag or lead, and
different equipment manufacturers use different conventions. In the DLI Balance
program software for the DC-7, either direction may be selected at the operator's
preference.

Vibration Units

So far, we have been looking at the displacement of a vibrating object as a measure
of its vibration amplitude. The displacement is simply the distance from a reference
position, or equilibrium point. In addition to varying displacement, a vibrating object
will experience a varying velocity and a varying acceleration. Velocity is defined as
the rate of change of displacement, and in the English system is usually measured in
units of inches per second. Acceleration is defined as the rate of change of velocity,
and in the English system, is usually measured in units of G, or the average
acceleration due to gravity at the earth's surface.
15
The displacement of a body undergoing simple harmonic motion is a sine wave as we
have seen. It also turns out (and is easily proved mathematically), that the velocity
of the motion is sinusoidal. When the displacement is at a maximum, the velocity will
be zero because that is the position at which its direction of motion reverses. When
the displacement is zero (the equilibrium point), the velocity will be at a maximum.
This means that the phase of velocity waveform will be displaced to the left by 90
degrees compared to the displacement waveform. In other words, the velocity is said
to lead the displacement by a 90-degree phase angle.
Remembering that acceleration is the rate of change of velocity, it can be shown that
the acceleration waveform of an object undergoing simple harmonic motion is also
sinusoidal, and also that when the velocity is at a maximum, the acceleration is zero.
In other words, the velocity is not changing at this instant. Then, when the velocity
is zero, the acceleration is at a maximum -- the velocity is changing the fastest at
this instant. The sine curve of acceleration versus time is thus seen to be 90 degrees
phase shifted to the left of the velocity curve, and therefore acceleration leads
velocity by 90 degrees.
These relationships are shown here:
'

Note here that the acceleration is 180 degrees out of phase with the displacement.
This means the acceleration of a vibrating object is always in the opposite direction
to the displacement!
It is possible to define another parameter that is the rate of change of acceleration,
and it is called "jerk". Jerk is what you feel when your car comes to a stop if you
maintain a constant brake pedal pressure. It is really the sudden cessation of the
deceleration. Elevator manufacturers are interested in measuring jerk, for it is the
variation in acceleration that elevator passengers are especially sensitive to.
Summary of Amplitude Units:

In the English system of measurements, displacement is usually measured in mils
(thousandths of an inch), and the peak-to-peak value is used by convention.
Velocity is usually measured in inches per second, and the convention is to use the
peak value or the RMS value. The peak value is the most commonly used, not
because it is better, but because of long tradition.
Acceleration is usually measured in Gs, where 1 G is the acceleration due to gravity
at the earth's surface. The G is not actually an acceleration unit -- it is simply an
16
amount of acceleration we experience as inhabitants of the earth. Acceleration is
sometimes measured in inches per second per second (in/sec2), or m/sec2, that are
true units. One G is equal to 386 inches/sec2 or 9.81 meters/sec2.
The process of converting a signal from displacement to velocity or velocity to
acceleration is equivalent to the mathematical operation of differentiation.
Conversely, the conversion from acceleration to velocity or velocity to displacement
is mathematical integration. It is possible to perform these operations in vibration
measuring instruments and thus to convert from any system of units to any other
one. From a practical standpoint however, differentiation is an inherently noisy
process, and is seldom done. Integration, on the other hand, can be done very
accurately with inexpensive electrical circuitry. This is one reason that the
accelerometer is the de facto standard transducer for vibration measurement, for its
output is easily integrated once or twice in order to display velocity or displacement.
Integration is not suitable, however, for signals of very low frequencies (Below 1 Hz),
for in this region the noise level increases and the accuracy of the integration
process itself suffers. Most commercially available integrators operate correctly
above one Hz, which is sufficiently low for almost all vibration applications.

Displacement, Velocity and Acceleration


This means that a plot of
vibration velocity will slope
upwards as frequency rises
compared to the same signal
plotted as displacement.

A vibration signal plotted as displacement vs.
frequency can be converted into a plot of velocity vs.
frequency by a process of differentiation, as we have
defined earlier. Differentiation involves a multiplication
by frequency, and this means the vibration velocity at
any frequency is proportional to the displacement
times the frequency. For a given displacement, if the
frequency is doubled, the velocity will also double, and
if the frequency is increased tenfold, the velocity is
also increased by a factor of ten.
In order to obtain acceleration from velocity, another differentiation is required, and
this results in another multiplication by frequency. The result is that for a given
displacement, the acceleration is proportional to the frequency squared. This means
that the acceleration curve slopes upward twice as steeply as the velocity curve.
To illustrate these relationships, consider how easy it is to move your hand back and
forth over a distance of one foot at one cycle per second, or 1 Hz.. It might be
possible to attain the same hand displacement at 5 or 6 Hz. But consider how fast
your hand would be moving if it had the same 1 foot displacement at 100 Hz, or
1000 Hz!

Newton's second law of
motion states that force
equals mass times
acceleration.

Now consider the great force that would be required to move
your hand a foot at these higher frequencies. Force equals
mass times acceleration according to Newton, so the force
required goes up as the square of the frequency. This is the
reason we never see high acceleration levels combined with
high displacement values. The very large forces that would
be required are simply not found in practice.
From these considerations, it can be seen that the same vibration data plotted in
displacement, velocity, and acceleration will have very different appearances. The
displacement curve will greatly emphasize the lowest frequencies, and the
17
acceleration curve will greatly emphasize the highest frequencies at the expense of
the lowest ones.
The relationship between levels of displacement, velocity, and acceleration versus
frequency in standard English units of mils peak-to-peak, inches per second peak,
and G RMS are expressed by the following equations:


18
The three curves shown above display the same information, but the emphasis is
changed. Note that the displacement curve is difficult to read at higher frequencies,
and acceleration has enhanced higher frequency levels. The velocity curve is the
most uniform in level over frequency. This is typical of most rotating machinery, but
in some cases the displacement or acceleration curves will be the most uniform. It is
a good idea to select the units so the flattest curve is attained -- this provides the
most visual information to the observer. Velocity is the most commonly used
vibration parameter for machine diagnostic work.




Complex Vibration

In a linear mechanical system, all the
vibration components will exist together,
and none will interfere with any other. In
the case of a non-linear system, the
vibration components will interact and
generate new components which are not
in the forcing function. See also the
section on linear systems in the Machine
Monitoring chapter.

Vibration is the motion resulting from an
oscillating force, and for a linear
mechanical system, the vibration
frequency will be the same as the forcing
frequency. If there are several forcing
frequencies occurring at the same time,
then the resulting vibration will be a
summation of the vibration at each
frequency. Under these conditions the
resulting waveform of the vibration will
not be a sinusoid, and may be very
complex.



19

Certain machines, especially very
slow speed ones, produce vibration
wave forms that are relatively easy
to interpret directly. See also the
section on Time Domain Analysis in
the Machinery Monitoring chapter.

In the diagram, the high frequency and the low
frequency vibration add together to make the
complex waveform. In simple cases like this, it is
relatively easy to find the frequencies and
amplitudes of the two components by
examination of the wave form, but most vibration
signals are far more complex than this, and can
be extremely difficult to interpret. In a typical
rotating machine, it is often hard to get very
much information about the inner workings of the
machine by looking at the vibration wave form,
although in certain cases wave form analysis is a
powerful tool, as will be discussed in the chapter
on machine vibration monitoring.

Energy and Power Considerations

Energy is required to produce vibration and in the case of machine vibration, this
energy comes from the source of power to the machine. This energy source can be
the AC power line, an internal combustion engine, or steam driving a turbine, etc.
Energy is defined as force multiplied by the distance over which the force acts, and
the SI unit of energy is the Joule. One Joule of energy is equivalent to a force of one
Newton acting over a distance of one meter. The physical concept of work is similar
to that of energy, and the units used to measure work are the same as those for
measuring energy.
The actual amount of energy present in the machine vibration itself is usually not
very great compared to the energy required to operate the machine for its intended
task.
Power is defined as the rate of doing work, or the rate of energy transfer, and
according to the SI, it is measured in Joules per second, or Watts. One horsepower is
equivalent to 746 watts. Power is proportional to the square of the vibration
amplitude, just as electrical power is proportional to the voltage squared or the
current squared.
According to the law of conservation of energy, energy cannot be created or
destroyed, but it can be transformed into different forms. The vibratory energy in a
mechanical system is ultimately dissipated in the form of heat.
More:

Mechanical Structures
Natural Frequencies
Resonance
Mechanical Structures

In analyzing the vibration of a machine, which is a more or less complex mechanical
system, it is useful to consider the sources of vibration energy and the paths in the
machine that this energy takes. Energy always moves, or flows, from the source of
the vibration to the energy absorber where it is converted into heat. In some cases,
20
this may be a very short path, but in other situations, the energy may travel
relatively long distances before being absorbed.
The most important absorber of energy in a machine is friction, which can be sliding
friction or viscous friction. Sliding friction is represented by relative motion between
parts of the machine, and an example of viscous friction is the oil film in a journal
bearing. If a machine has very little friction, its vibration level tends to be fairly high,
for the vibration energy builds up due to the lack of absorption. On the other hand, a
machine with greater inherent friction will have lower vibration levels because the
energy is absorbed quickly. For example, a machine with rolling element bearings
(often called anti-friction bearings), generally vibrates more than a machine with
sleeve bearings, where the oil film acts as a significant absorber of energy. The
reason that airplane structures are riveted together rather than being welded into a
solid unit is that the riveted joints move slightly, absorbing energy by sliding friction.
This keeps vibrations from building up to destructive levels. Such a structure is said
to be highly damped, and the damping is actually a measure of its energy absorption
capability.

Natural Frequencies

Any physical structure can be modeled as a number of springs, masses, and dampers.
Dampers absorb energy, but springs and masses do not. As we saw in the previous
section, a spring and a mass interact with one another to form a system that
resonates at their characteristic natural frequency. If energy is applied to a spring-
mass system, it will vibrate at its natural frequency, and the level of the vibration
depends on the strength of the energy source as well as the absorption or damping
inherent in the system. The natural frequency of an undamped spring-mass system
is given by the following equation:

where Fn = The natural frequency
k = the spring constant, or stiffness
m = the mass
From this, it is seen that if the stiffness increases, the natural frequency also
increases, and if the mass increases, the natural frequency decreases. If the system
has damping, which all physical systems do, its natural frequency is a little lower,
and depends on the amount of damping.
The multitude of spring-mass-damper systems that make up a mechanical system
are called "degrees of freedom", and the vibration energy put into a machine will
distribute itself among the degrees of freedom in amounts depending on their natural
frequencies and damping, and on the frequency of the energy source. For this reason,
the vibration will not be uniformly distributed in the machine. For instance, in a
machine driven by an electric motor, a major source of vibration energy is residual
imbalance in the motor rotor. This will result in a measurable vibration at the motor
bearings. But if the machine has a degree of freedom with a natural frequency close
to the RPM of the rotor, its vibration level can be very high, even though it may be a
long distance from the motor. It is important to be aware of this fact when
evaluating the vibration of a machine -- the location of the maximum vibration level
21
may not be close to the source of the vibration energy. Vibration energy frequently
travels great distances along pipes, and can wreak havoc when it encounters a
remote structure with a natural frequency near that of its source.
Resonance


Examples of highly
resonant mechanical
systems are bells and
tuning forks.

Resonance is an operating condition where an excitation
frequency is near a natural frequency of the machine
structure. A natural frequency is a frequency at which a
structure will vibrate if deflected and then let go. A typical
structure will have many natural frequencies. When
resonance occurs, the resulting vibration levels can be
very high and can cause rapid damage.

Under no circumstances
should a machine be
operated at a speed
corresponding to a
resonance!

In a machine that produces a broad spectrum of vibration
energy, a resonance shows up in the vibration spectrum as
a peak whose frequency is constant even as the machine
speed is varied. The peak may be quite sharp, or may be
broad; depending on the amount of effective damping the
structure has at the frequency in question.
In order to determine if a machine has prominent
resonances, one of several tests can be performed to find
them:
The "Bump Test" -- The machine is impacted with a heavy mass such as a wooden
four by four or the booted heel of the foot of a football player while recording
vibration data. If a resonance is there, the machine vibration will be at the natural
frequency as it dies away.
The "Run Up" or "Coast Down" -- The machine is turned on, or turned off, while
taking vibration data and tachometer data. The time wave form will show maxima
when the RPM matches natural frequencies.
"Variable Speed Test" -- With a machine whose speed can be varied over a wide
range, the speed can be varied while taking vibration and tachometer data. The data
are interpreted as in the run up test.
The figure below shows an idealized response curve of a mechanical resonance. The
behavior of a resonant system when subjected to an external force is interesting and
somewhat counter intuitive. It depends strongly on the frequency of the excitation
force. If the forcing frequency is lower than the natural frequency -- in other words
to the left of the peak -- then the system behaves like a spring, and the
displacement is proportional to the force. The spring of the spring-mass combination
making up the resonant system is dominant in determining the response of the
system. In this spring-controlled region, the system behaves in agreement with our
intuition, responding with greater motion as greater force is applied to it, and the
motion is in phase with the force.
In the region above the natural frequency, the situation is different. Here, the mass
is the controlling element, and the system looks like a mass to an input force. This
means its acceleration is proportional to the applied force, and the displacement is
relatively constant with changing frequency. The displacement is out of phase with
the force in this region -- when you push against the system, it moves toward you
and vice versa!
22
At the resonance itself, the system looks completely different to an applied force.
Here, the mass and spring elements effectively cancel each other out, and the force
sees only the damping, or friction, in the system. If the system is lightly damped, it
is like pushing on air. When you push on it, it recedes from you on its own.
Consequently, you cannot apply much force to the system at resonance, and if you
continue to try, the vibration amplitude builds up to very high values. It is the
damping that controls the motion of a resonant system at its natural frequency.


Examples of resonances in machines are the so-called critical frequencies of rotating
shafts.
The phase angle between the excitation source vibration and the response of the
structure is always 90 degrees at the natural frequency
In the case of long rotors such as turbines, the natural frequencies are called "critical
frequencies" or "critical speeds," and care must be taken that these machines are
not operated at speeds where 1X or 2X correspond to these critical frequencies.
Linear and Non-Linear Systems

To assist in understanding the transmission of vibration through a machine, it is
instructive to investigate the concept of linearity and what is meant by linear and
non-linear systems. Thus far, we have discussed linear and logarithmic amplitude
and frequency scales, but the term "linear" also refers to the characteristics of a
system which can have input and output signals. A "system" is any device or
structure that can accept an input or stimulus in some form and produce a
corresponding output or response. Examples of systems are tape recorders and
amplifiers, which operate on electrical signals, and mechanical structures, whose
inputs are vibration forces, and whose outputs are vibration displacements, velocities,
or accelerations.
More:

Definition of Linearity
Non-Linearities in Systems
Non-Linearities in Rotating Machines
23
Definition of Linearity

A system is said to be linear if it meets the following two criteria:
1. If input x to the system results in output X, then an input of 2x will produce
output of 2X. In other words, the magnitude of the system output is proportional to
the magnitude of the system input.
2. If input x produces output X, and input y produces output Y, then an input of x +
y will produce an output of X + Y. In other words, the system handles two
simultaneous inputs independently, and they do not interact within the system.
Implicit in these criteria is the fact that a linear system will not produce any
frequencies in the output that are not present in the input.
Note that there is nothing in these criteria that says the system output is the same
as the system input, or even that it resembles the system input. For instance, the
input could be an electric current, and the output could be a temperature. In the
case of mechanical structures such as machines, we will consider the input to be a
vibratory force and the output to be the measured vibration itself.
Non-Linearities in Systems

Absolutely perfect linearity does not exist in any real system. There are many
different types of non-linearity, and they exist in varying degrees in all mechanical
systems, although many actual systems approach linear behavior, especially with
small input levels. If a system is not perfectly linear, it will produce frequencies in its
output that do not exist in its input. An example of this is a stereo amplifier or tape
recorder that produces harmonics of its input signal. This is called "harmonic
distortion", and it degrades the quality of the music being reproduced. Harmonic
distortion almost always gets much worse at high signal levels. An example of this is
a small radio that sounds relatively "clean" at low volume levels, but sounds harsh
and distorted at high volume levels.
Many systems are very nearly linear in response to small inputs, but become non-
linear at higher levels of excitation. Sometimes a definite threshold exists in which
input levels only a little above the threshold result in gross non-linearity. An example
of this is the "clipping" of an amplifier when its input signal level exceeds the voltage
or current swing capacity of its power supply. This is analogous to a mechanical
system where a part is free to move until it hits a stop, such as a loose bearing
housing that can move a little before being stopped by the mounting bolts.
Non-Linearities in Rotating Machines

As has been discussed, the vibration of a machine is actually its response to forces
caused by moving parts in the machine. We measure the vibration at various
locations on the machine, and deduce from these vibrations the magnitude of the
forces. In measuring the frequency of the vibration, we assume the forces occur at
the same frequency as the response, and that the measured levels are proportional
to the magnitudes of the forces. This rationale assumes that the machine is linear in
its response to forcing functions, and this is a reasonable assumption for most
machines.
However, as a machine wears and clearances increase, or if it develops cracks or
loose parts, its response will no longer be linear, and the result is that the measured
vibration can be quite different in character from the forcing functions. For instance,
24
an unbalanced rotor imparts a sinusoidal force at a frequency of 1X to the bearing,
and this force does not contain any other frequency. If the mechanical structure of
the machine is non-linear, this sinusoidal force will be distorted, and the resulting
vibration will occur at harmonics of 1X as well as 1X. The extent and magnitude of
the harmonic content of the vibration is a measure of the degree of non-linearity of
the machine. For instance, the vibration of a journal bearing contains greater and
greater numbers and magnitudes of harmonics as the bearing clearance increases.
Flexible couplings are non-linear when misaligned, and this is the reason their
vibration signature contains a strong second harmonic of 1X. Worn couplings that are
misaligned often produce a strong third harmonic of 1X. When forces acting at
different frequencies interact in a non-linear way in a machine, the result is the
generation of sum and difference frequencies -- new frequencies that are not present
in the forcing functions themselves. These sum and difference frequencies are the
sidebands found in spectra of defective gearboxes, rolling element bearings, etc. In
the case of a gearbox, one forcing frequency is the gear mesh and another is the
rpm of the gear. If the gear is eccentric or otherwise misshapen, the rpm will
modulate the gear mesh resulting in sidebands. Modulation is always a non-linear
process, creating new frequencies that do not exist in the forcing functions.
Frequency Analysis

To get around the limitations in the analysis of the wave form itself, the common
practice is to perform frequency analysis, also called spectrum analysis, on the
vibration signal. The time domain graph is called the waveform, and the frequency
domain graph is called the spectrum. Spectrum analysis is equivalent to transforming
the information in the signal from the time domain into the frequency domain. The
following relationships hold between time and frequency:



A train schedule shows the equivalence of information in the time and frequency
domains:
25


The frequency representation in this case is much shorter than the time
representation. This is a "data reduction".
Note that the information is the same in both domains, but that it is much more
compact in the frequency domain. A very long schedule in time has been compressed
to two lines in the frequency domain. It is a general rule of the transformation
characteristic that events that take place over a long time interval are compressed to
specific locations in the frequency domain.
More:

Why perform Frequency Analysis?
How to perform Frequency Analysis
Examples of some wave forms and their spectra
Modulation Effects
Beats
Why perform Frequency Analysis?

In the figure below, note that the individual frequency components are separate and
distinct in the spectrum, and that their levels are easily identified. It would be
difficult to extract this information from the time domain waveform.
26


It has been argued that
the primary reason for
the widespread use of
frequency analysis is the
wide availability of the
inexpensive FFT analyzer!

In the next figure, we see that events that are overlapped
and confused in the time domain are separated into
individual components in the frequency domain. The
vibration waveform contains a great deal of information
that is not apparent to the eye. Some of the information is
in very low-level components whose magnitude may be
less than the width of the line of the waveform plot.
Nevertheless, such very low-level components may be
important if they indicate a developing problem such as a
bearing fault. The essence of predictive maintenance is the
early detection of incipient faults, so we must be sensitive
to very small values of vibration signals, as we will see
shortly.


In the next figure, a very low-level component represents a small developing fault in
a bearing, and it would have been unnoticed in the time domain or in the overall
27
vibration level. Remember that the overall level is simply the RMS level of the
vibration waveform over a broad frequency range, and that a small disturbance such
as the bearing tone shown here could double or quadruple in level before the overall
RMS would be affected.



On the other hand, there are circumstances where the waveform provides more
information to the analyst than does the spectrum.








How to perform Frequency Analysis

Before we investigate the procedure of performing spectrum analysis, we will look at
the various types of signals we will be working with.
From a theoretical and practical standpoint, it is possible to divide all time domain
signals into several groups. These different signal types produce different types of
spectra, and to avoid errors in performing frequency analysis, it is instructive to
know their characteristics.
28


More:

Stationary Signals
Deterministic Signals
Non-Stationary Signals

Stationary Signals

The first natural division of all signals is into either stationary or non-stationary
categories. Stationary signals are constant in their statistical parameters over time.
If you look at a stationary signal for a few moments and then wait an hour and look
at it again, it would look essentially the same, i.e. its overall level would be about the
same and its amplitude distribution and standard deviation would be about the same.
Rotating machinery generally produces stationary vibration signals.
Stationary signals are further divided into deterministic and random signals. Random
signals are unpredictable in their frequency content and their amplitude level, but
they still have relatively uniform statistical characteristics over time. Examples of
random signals are rain falling on a roof, jet engine noise, turbulence in pump flow
patterns and cavitation.

Deterministic Signals

Deterministic signals are a special class of stationary signals, and they have a
relatively constant frequency and level content over a long time period. Deterministic
signals are generated by rotating machines, musical instruments, and electronic
function generators. They are further divisible into periodic and quasi-periodic signals.
Periodic signals have waveforms whose pattern repeats at equal increments of time,
whereas quasi-periodic signals have waveforms whose repetition rate varies over
time, but still appears to the eye to be periodic. Sometimes, rotating machines will
produce quasi-periodic signals, especially belt-driven equipment.
Deterministic signals are probably the most important in vibration analysis and their
spectra resemble the following:
29



Most quasi-periodic signals
are actually a combination
of several harmonic series.

Periodic signals always produce spectra with discrete
frequency components that are a harmonic series. The
term "harmonic" comes from music, where harmonics
are multiples of the fundamental frequency.
Non-Stationary Signals

Non-stationary signals are divided into continuous and transient types. Examples of
non-stationary continuous signals are the vibration produced by a jackhammer and
the sound of a fireworks display. Transient signals are defined as signals which start
and end at zero level and last a finite amount of time. They may be very short, or
quite long. Examples of transient signals are a hammer blow, an airplane flyover
noise, or a vibration signature of a machine run up or run down.

Examples of some wave forms and their
spectra

Following are some waveforms and spectra that illustrate some important
characteristics of frequency analysis. While these are idealized in the sense that they
were made from an electronic function generator and analyzed with an FFT analyzer,
they do show certain attributes that are commonly seen in machine vibration spectra.
30


A sine wave consists of a single frequency only, and its spectrum is a single point.
Theoretically, a sine wave exists over infinite time and never changes. The
mathematical transform that converts the time domain waveform into the frequency
domain is called the Fourier transform, and it compresses all the information in the
sine wave over infinite time into one point. The fact that the peak in the spectrum
shown above has a finite width is an artifact of the FFT analysis, which will be
discussed later.
A machine with imbalance has an excitation force that is a sine wave at 1X, or once
per revolution. If the machine were perfectly linear in response, the resulting
vibration would be a pure sine wave like the one shown above. In many poorly
balanced machines, the waveform does resemble a sine wave, and there is a large
vibration peak in the spectrum at 1X, or one order.

31

Here we see that a harmonic spectrum results from a periodic waveform, in this case
a "clipped" sine wave. The spectrum contains equally spaced components, and their
spacing is equal to 1 divided by the period of the waveform. The lowest of the
components above zero frequency is called the fundamental, and the others are
called harmonics. This waveform came from a signal generator, and it can be seen
that it is not symmetrical about the zero line. This means it has a "DC." component,
and this is seen as the first line at the left in the spectrum. This is to illustrate that a
spectrum analysis can go all the way to zero frequency, or in common terminology,
to DC.
In vibration analysis of machinery, it is not usually desirable to include such low
frequencies in the spectrum analysis for several reasons. Most vibration transducers
do not have response to DC, although there are accelerometers that are used in
inertial navigation systems that do have DC response. For machine vibration, the
lowest frequency that is generally considered of interest is about 0.3 orders. In some
machines this will be below 1Hz. Special techniques are required to measure and
interpret signals below this frequency.

Note that because this
spectrum consists of discrete
points, the signal is by
definition deterministic!

It is not uncommon in machine vibration signatures to
see a waveform which is clipped something like the
one shown above. What this usually means is there is
looseness in the machine, and something is restricting
its motion in one direction.



The signal shown above is similar to the previous one, but it is clipped on both
positive and negative sides, resulting in a symmetrical waveform. This type of signal
can occur in machine vibration if there is looseness in the machine and motion is
restricted in both directions. The spectrum seems to have harmonics, but they are
actually only the odd-numbered harmonics. All the even-numbered harmonics are
missing. Any periodic waveform that is symmetrical will have a spectrum with only
odd harmonics! The spectrum of a square wave would also look like this.
32
Sometimes the vibration spectrum of a machine will resemble this if there is extreme
looseness and the motion of the vibrating part is restricted at each extreme of
displacement. An unbalanced machine with a loose hold-down bolt is an example of
this.



Shown above is a short impulse produced by a signal generator. Note that its
spectrum is continuous rather than discrete. In other words, the energy in the
spectrum is spread out continuously over a range of frequencies rather than being
concentrated only at specific frequencies. This is characteristic of non-deterministic
signals such as random noise and transients. Note that the level of the spectrum
goes to zero at a particular frequency. This frequency is the reciprocal of the length
of the impulse, therefore the shorter the impulse, the greater its high frequency
content. If the impulse were infinitely short (the so-called delta function, in
mathematics), then its spectrum would extend from 0 to infinity in frequency.
By examining a continuous spectrum, it is usually impossible to tell whether it is the
result of a random signal or a transient. This is an inherent limitation of Fourier-type
frequency analysis, and for this reason it is a good idea to look at the wave form
when a continuous spectrum is encountered. As far as machine vibration is
concerned, it is of interest to the analyst whether impacting is occurring (causing
impulses in the wave form) or random noise (for example, from cavitation) exists in
the signal.
A rotating machine seldom produces a single impulse like this, but in the "bump test",
this type of excitation is applied to the machine. Its vibration response will not be a
classic smooth curve like this one, but it will be continuous with peaks corresponding
to the natural frequencies of the machine structure. This spectrum shows that the
impulse is a good input force to use in this type of test, for it contains energy over a
continuous frequency range.
33


If the same impulse that produced the previous spectrum is repeated at a constant
rate, the resulting spectrum will have an overall envelope with the same shape as
the spectrum of the single impulse, but it will consist of harmonics of the pulse
repetition frequency rather than a continuous spectrum.
A bearing produces this type of signal with a definite defect in one of the races. The
impulses can be very narrow, and they will always produce an extensive series of
harmonics.
Modulation Effects

Modulation is a non-linear effect in which several signals interact with one another to
produce new signals with frequencies not present in the original signals. Modulation
effects are the bane of the audio engineer, for they produce "intermodulation
distortion", which is annoying to the music listener. There are many forms of
modulation, including frequency and amplitude modulation, and the subject is quite
complex. We will now look at the two primary types of modulation individually.

34



It is rare to see frequency
modulation by itself; most
machines will produce
amplitude modulation at the
same time as frequency
modulation!

Frequency modulation (FM) is the varying in frequency
of one signal by the influence of another signal, usually
of lower frequency. The frequency being modulated is
called the "carrier". In the spectrum shown above, the
largest component is the carrier, and the other
components which look like harmonics, are called
"sidebands". These sidebands are symmetrically
located on either side of the carrier, and their spacing
is equal to the modulating frequency.
Frequency modulation occurs in machine vibration spectra, especially in gearboxes
where the gear mesh frequency is modulated by the rpm of the gear. It also occurs
in some sound system loudspeakers, where it is called FM distortion, although it is
generally at a very low level.
35


This example shows amplitude modulation at about 50% of full modulation
Notice that the frequency of the waveform seems to be constant and that it is
fluctuating up and down in level at a constant rate. This test signal was produced by
rapidly varying the gain control on a function generator while recording the signal.

This type of signal is often
produced by defective
bearings and gears, and can
be easily identified by the
sidebands in the spectrum.

The spectrum has a peak at the frequency of the carrier,
and two more components on each side. These extra
components are the sidebands. Note that there are only
two sidebands here compared to the great number
produced by frequency modulation. The sidebands are
spaced away from the carrier at the frequency of the
modulating signal, in this case at the frequency at which
the control knob was wiggled. In this example, the
modulating frequency is much lower than the modulated
or carrier frequency, but the two frequencies are often
close together in practical situations. Also these
frequencies are sine waves, but in practice, both the
modulated and modulating signals are often complex.
For instance, the transmitted signal from an AM radio
station contains a high-frequency carrier, and many
sidebands resulting from the carrier modulation by the
voice or music signal being broadcast.

A vibration and acoustic signature similar to this is
frequently produced by electric motors with rotor bar
problems.



36


Beats





It is almost impossible to tell beating
from amplitude modulation by
looking at the waveform, but they
are fundamentally different
processes, caused by different
phenomena in machines. The
spectrum tells the story.

This waveform looks like amplitude modulation,
but is actually just two sine wave signals added
together to form beats. Because the signals are
slightly different in frequency, their relative
phase varies from zero to 360 degrees, and
this means the combined amplitude varies due
to reinforcement and partial cancellation. The
spectrum shows the frequency and amplitude
of each component, and there are no sidebands
present. In this example, the amplitudes of the
two beating signals are different, causing
incomplete cancellation at the null points
between the maxima. Beating is a linear
process -- no additional frequency components
are created.
Electric motors often produce sound and vibration signatures that resemble beating,
where the beat rate is at twice the slip frequency. This is not actually beating, but is
in fact amplitude modulation of the vibration signature at twice the slip frequency.
Probably it has been called beating because it sounds somewhat like the beats
present in the sound of an out of tune musical instrument.
The following example of beats shows the combined waveform when the two beating
signals are the same amplitude. At first glance, this looks like 100% amplitude
modulation, but close inspection of the minimum amplitude area shows that the
phase is reversed at that point.

37
'

This looks like 100% amplitude modulation!
This example of beats is like the previous one, but the levels of the two signals are
the same, and they cancel completely at the nulls. This complete cancellation is quite
rare in actual signals encountered in rotating equipment.



Earlier we learned that beats and amplitude modulation produce similar waveforms.
This is true, but there is a subtle difference. These waveforms are enlarged for clarity.
Note that in the case of beats, there is a phase change at the point where
cancellation is complete.




38

Octave Band and One-Third Octave Band
Analysis
Logarithmic Frequency Scaling

So far, the only type of frequency analysis discussed has been on a linear frequency
scale, i.e., the frequency axis is set out in a linear fashion. This is suitable for
frequency analysis with a frequency resolution that is constant throughout the
frequency range, commonly called "narrow band" analysis. The FFT analyzer
performs this type of analysis.
There are several situations where frequency analysis is desired, but narrow band
analysis does not present the data in its most useful form. An example of this is
acoustic noise analysis where the annoyance value of the noise to a human observer
is being studied. The human hearing mechanism is responsive to frequency ratios
rather than actual frequencies. The frequency of a sound determines its pitch as
perceived by a listener, and a frequency ratio of two is a perceived pitch change of
one octave, no matter what the actual frequencies are. For instance if a sound of 100
Hz frequency is raised to 200 Hz, its pitch will rise one octave, and a sound of 1000
Hz, when raised to 2000 Hz, will also rise one octave in pitch. This fact is so precisely
true over a wide frequency range that it is convenient to define the octave as a
frequency ratio of two, even though the octave itself is really a subjective measure of
a sound pitch change.
This phenomenon can be summarized by saying that the pitch perception of the ear
is proportional to the logarithm of frequency rather than to frequency itself.
Therefore, it makes sense to express the frequency axis of acoustic spectra on a log
frequency axis, and this is almost universally done. For instance, the frequency
response curves that sound equipment manufacturers publish are always plotted in
log frequency. Likewise, when frequency analysis of sound is performed, it is very
common to use log frequency plots.

The vertical axis of
an octave band
spectrum is usually
scaled in dB.

The octave is such an important frequency interval to the ear
that so-called octave band analysis has been defined as a
standard for acoustic analysis. The figure below shows a typical
octave band spectrum where the ISO standard center
frequencies of the octave bands are used. Each octave band has
a bandwidth equal to about 70% of it center frequency. This type
of spectrum is called constant percentage band because each
frequency band has a width that is a constant percentage of its
center frequency. In other words, the analysis bands become
wider in proportion to their center frequencies.

39

It can be argued that the frequency resolution in octave band analysis is too poor to
be of much use, especially in analyzing machine vibration signatures, but it is
possible to define constant percentage band analysis with frequency bands of
narrower width. A common example of this is the one-third-octave spectrum, whose
filter bandwidths are about 27 % of their center frequencies. Three one-third octave
bands span one octave, so the resolution of such a spectrum is three times better
than the octave band spectrum. One-third octave spectra are frequently used in
acoustical measurements.
A major advantage of constant percentage band analysis is that a very wide
frequency range can be displayed on a single graph and the frequency resolution at
the lower frequencies can still be fairly narrow. Of course, the frequency resolution at
the highest frequencies suffers, but this is not a problem for some applications such
as fault detection in machines.
In the chapter on machine fault diagnosis, it will be seen the narrow band spectra
are very useful in resolving higher-frequency harmonics and sidebands, but for the
detection of a machine fault, no such high resolution is required. The vibration
velocity spectra of most machines will be found to slope downwards at the highest
frequencies, and a constant percentage band (CPB) spectrum of the same data will
usually be more uniform in level over a broad frequency range. This means that a
CPB spectrum takes better advantage of the dynamic range of the instrumentation.
One-third octave spectra are sufficiently narrow at low frequencies to show the first
few harmonics of run speed, and can be used effectively for the detection of faults if
trended over time.
The use of constant CPB spectra for machine monitoring is not very well recognized
in industry with a few notable exceptions such as the US Navy submarine fleet.
Linear and Logarithmic Amplitude Scales

It may seem to be best to look at vibration spectra with a linear amplitude scale
because that is a true representation of the actual measured vibration amplitude.
Linear amplitude scaling makes the largest components in a spectrum very easy to
see and to evaluate, but very small components may be overlooked completely, or
are at best difficult to assign a magnitude to. The eye is able to see small
components about 1/50th as large as the largest ones in the same spectrum, but
anything smaller than this is essentially lost. In other words, the dynamic range of
the eye is about 50 to 1
Linear scaling may be adequate in cases where the components are all about the
same size, but in the case of machine vibration, beginning faults in such parts as
bearings produce very small signal amplitudes. If we are to do a good job of trending
the levels of these spectral components, it is best to plot the logarithm of the
40
amplitude rather than the amplitude itself. In this way, we can easily display and
visually interpret a dynamic range of at least 5000 to 1, or more than 100 times
better than the linear scaling allows.
To illustrate different types of amplitude presentations, the same vibration signature
will be shown in linear and two different types of logarithmic amplitude scales.
It might be said that the dynamic range of the eye, when looking at linear spectra, is
about 34 dB.
More:

Linear Amplitude Scaling
Logarithmic Amplitude Scaling
The Decibel
dB Values vs. Amplitude Level Ratios
Unit Conversions
VdB Levels vs. Vibration Levels in ips
Linear Amplitude Scaling


Note that this linear spectrum shows the larger peaks very well, but lower level
information is missing. In the case of machine vibration analysis, we are often
interested in the smaller components of the spectrum, i.e., in the case of rolling
element bearing diagnosis. This subject will be covered in detail in the chapter on
Machine Vibration Monitoring.
Logarithmic Amplitude Scaling

41
The
spectrum above plots the logarithm of the vibration level rather than the level itself.
Since this spectrum is on a log amplitude scale, multiplication by any constant value
simply translates the spectrum up on the screen without changing its shape or the
relationship between the components.
Multiplication of the signal level translates into addition on a log scale. This means
that if the amount of amplification of a vibration signal is changed, the shape of the
spectrum is not affected. This fact greatly simplifies visual interpretation of log
spectra taken at different amplification factors -- the curves are simply translated up
or down on the graph. With a linear scaling, the shape of the spectrum changes
drastically with different degrees of amplification.
The next spectrum is presented in decibels, a special type of log scaling that is very
important in vibration analysis.







42

The Decibel


The decibel (dB) is defined by the following expression:


where: LdB = The signal level in dB
L1 = Vibration level in Acceleration, Velocity, or Displacement
Lref = Reference level, equivalent to 0 dB
The Bell Telephone Labs introduced the concept of the decibel before 1930. It was
first used to measure relative power loss and signal to noise ratio in telephone lines.
It was soon pressed into service as a measure of acoustic sound pressure level.
The vibration velocity level in dB is abbreviated VdB, and is defined as:



or




43

The Systeme
Internationale, or SI, is
the modern replacement
for the metric system.

The reference, or "0 dB" level of 10-9 meter per sec is
sufficiently small that all our measurements on machines
will result in positive dB numbers. this standardized
reference level uses the SI, or "metric," system units, but
it is not recognized as a standard in the US and other
English-speaking countries. (The US. Navy and many
American industries use a zero dB reference of 10-8
m/sec, making their readings higher than SI readings by
20 dB.)
The VdB is a logarithmic scaling of vibration magnitude, and it allows relative
measurements to be easily made. Any increase in level of 6 dB represents a doubling
of amplitude, regardless of the initial level. In like manner, any change of 20 dB
represents a change in level by a factor of ten. Thus any constant ratio of levels is
seen as a certain distance on the scale, regardless of the absolute levels of the
measurements. This makes it very easy to evaluate trended vibration spectral data;
6 dB increases always indicate doubling of the magnitudes.
dB Values vs. Amplitude Level Ratios

The following table relates dB values to amplitude ratios:

dB Change Linear Level
Ratio
dB Change Linear Level
Ratio
0

1

30

31

3

1.4

36

60

6

2

40

100

10

3.1

50

310

12

4

60

1000

18

8

70

3100

20

10

80

10,000

24

16

100

100,000


It is strongly recommended that VdB be used as the vibration amplitude scaling
because so much more information is available to the viewer compared to linear
amplitude units. Also, compared to a conventional log scale, the dB scale is much
easier to read.
Unit Conversions

Acceleration and Displacement can also be expressed on dB scales. The AdB scale is
the most used one, and its zero reference is set 1 micro G, commonly abbreviated G.
It turns out that AdB = VdB at 159.2 Hz. VdB levels, AdB levels, and DdB levels are
related by the following formulas:
44
Any vibration parameter --
displacement, velocity, or
acceleration can be
displayed on a dB scale.
The reference quantities
for 0 dB on these scales
were chosen such that the
dB levels of all three
quantities are the same at
a frequency of 159.2 Hz,
which is equal to 1000
radians per second.


Acceleration and Velocity in linear units are calculated from dB levels as follows:


It is convenient to
remember the following
rule of thumb:
At 100 Hz, 1G = 120
AdB = 124 VdB = 2.8
mils p-p.



Note that the time domain wave form is always represented in linear amplitude units
- it is not possible to use a log scale in the wave form plot because some of the
values are negative, and the logarithm of a negative number is not defined.
VdB Levels vs. Vibration Levels in ips

Peak level is the de facto
standard unit for vibration
velocity measurements,
even though RMS level
would make more sense in
most cases.



Following is a convenient conversion table for relating VdB
levels to inches per second peak:


VdB ips peak VdB ips peak VdB ips peak
60

.0006

90

.018

120

.56

62

.0007

92

.022

122

.70

64

.0009

94

.028

124

.88

66

.0011

96

.035

126

1.1

68

.0014

98

.044

128

1.4

70

.0018

100

.056

130

1.8

72

.0022

102

.070

132

2.2

74

.0028

104

.088

134

2.8

76

.0035

106

.11

136

3.5

45
78

.0044

108

.14

138

4.4

80

.0056

110

.18

140

5.6

82

.0070

112

.22

142

7.0

84

.0088

114

.28

144

8.8

86

.011

116

.35

146

11.1

88

.014

118

.44

148

14.0
Vibration transducers
Overview


An early vibration
transducer is the human
finger! An earlier, and
much more sensitive one
is the lateral line organ of
the fishes.

The vibration transducer is a device that produces an
electric signal that is a replica, or analog, of the vibratory
motion it is subjected to. A good transducer should not add
any spurious components to the signal, and should produce
signals uniformly over the frequency range of interest.
Different types of transducers respond to different parameters of the vibration source,
as shown in the following table:
Name: Sensitive To:
Proximity Probe

Displacement

Velocity Probe

Velocity

Accelerometer

Acceleration


On the following pages, we will examine the characteristics of these transducers.
The Proximity Probe
The Velocity Probe
The Accelerometer
The Proximity Probe


One very common type of
proximity probe is known
commercially as a
"Proximiter", which is a
trademark of the Bentley
Nevada Company.

The Proximity Probe, also called an "Eddy Current Probe"
or "Displacement Transducer", is a permanently mounted
unit, and requires a signal-conditioning amplifier to
generate an output voltage proportional to the distance
between the transducer end and the shaft. It operates on a
magnetic principle, and is thus sensitive to magnetic
anomalies in the shaft -- care should be taken that the
shaft is not magnetized to assure the output signal is not
contaminated. It is important to realize that the transducer
measures relative displacement between the bearing and
the journal, and does not measure total vibration level of
the shaft or the housing. The displacement transducer is
very commonly installed in large machines with journal
bearings where it is used to detect bearing failure and to
shut the machine down before catastrophic failure occurs.
46


These transducers are frequently used in pairs oriented 90 apart, and can be
connected to the vertical and horizontal plates of an oscilloscope to display the
"orbit", or path of the journal as it migrates around in the bearing.
The frequency response of the displacement transducer extends from DC (0 Hz) to
about 1000 Hz.
The Velocity Probe



Velocity Transducer

Some velocity transducers are made with a moving coil outside a stationary magnet.
The principle of operation is the same. 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
The velocity probe was one of the first vibration transducers to be built. It consists of
a coil of wire and a magnet so arranged that if the housing is moved, the magnet
tends to remain stationary due to its inertia. The relative motion between the
magnetic field and the coil induces a current that is proportional to the velocity of
motion. The unit thus produces a signal directly proportional to vibration velocity. It
is self-generating and needs no conditioning electronics in order to operate, and it
has a relatively low electrical output impedance making it fairly insensitive to noise
induction.
In spite of these advantages, the velocity transducer 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
47
and the magnet make up a low-frequency resonant system with a natural frequency
of about 10 Hz. 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.
The Accelerometer



Piezo-Electric Accelerometer
The compression-type accelerometer, diagrammed here, was the first type to be
developed. The shear type, which is arranged so the active element is subjected to
shear forces, is generally preferred.
There are also other designs for accelerometers
The piezo-electric accelerometer can be considered the standard vibration transducer
for machine vibration measurement. It is made in several different configurations,
but the illustration of the compression type serves to describe the principle of
operation.
The seismic mass is clamped to the base by an axial bolt bearing down on a circular
spring. The piezo-electric element is squeezed between the mass and the base.
When a piezo-electric material experiences a force, it generates an electric charge
between its surfaces. There are many such materials, with quartz being one of the
most commonly used. There are also synthetic ceramic piezo materials that work
well, and in some cases, work at higher temperatures than quartz is able to do. If
the temperature of a piezo material is increased, finally the so called "curie point", or
"curie temperature" is reached, and the piezo-electric property is lost. Once this
happens, the transducer is defective and not repairable.
When the accelerometer is moved in the up and down direction, the force required to
move the seismic mass is born by the active element. According to Newton's second
law, this force is proportional to the acceleration of the mass. The force on the
crystal produces the output signal, which is therefore proportional to the acceleration
of the transducer. Accelerometers are inherently extremely linear in an amplitude
sense, meaning they 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.
This range of acceleration levels can span an amplitude range of about 108, which is
160 dB! No other transducer can match this performance.
48
The piezo-electric accelerometer is very stable over long periods of time, and will
maintain its calibration if it is not abused. The two ways that accelerometers can be
damaged are subjecting them to excessive heat and dropping onto a hard surface. If
dropped more than a few feet onto a concrete floor or steel deck, the accelerometer
should be re-calibrated to be sure the crystal is not cracked. A small crack will cause
the sensitivity to be reduced and also will greatly affect the resonance, and thus the
frequency response. It is a good idea to calibrate accelerometers about once a year if
they are in service with portable data collectors.
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 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.

When using an ICP accelerometer,
care must be
taken not to subject it to
acceleration levels where the
output voltage will exceed several
volts. Otherwise, the internal
preamplifier will be overloaded
and data distortion will result!

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
DC 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 to 0.1
Hz if very low frequency data is required.
When an ICP accelerometer is connected to the power source, it takes a few seconds
for the amplifier to stabilize, and during this time, any data the unit is collecting will
be contaminated by a slowly varying voltage ramp. For this reason, there must be a
time delay built into data collectors to assure the unit is stable. If the delay is too
short, the time waveform will have an exponentially shaped voltage ramp
superimposed on the data, and the spectrum will show a rising very low-frequency
characteristic sometimes called a "ski slope". This should be avoided because the
dynamic range of the measurement is compromised.
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.
49

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.

FFT Analyzer
Background

This section will cover the operation and theory of the FFT analyzer, which is the
most commonly used piece of signal analysis equipment in the vibration field. Many
workers think of the FFT analyzer as a "magic box," into which you put a signal and
out of which comes a spectrum. The assumption usually is that the spectrum tells
the truth -- the box cannot lie. We will see that this assumption is valid in many
cases, but we will also see that we can be misled, for there are several pitfalls in the
process of digital signal analysis. One of the purposes of this section is to help you
avoid falling into any of the pitfalls, and if you do, how to crawl out smelling like a
rose.
FFT analysis is but one type of digital spectrum analysis, but we will not concentrate
on the other types because they do not apply directly to the VMS program.
Spectrum Analysis

Spectrum analysis, which is defined as the transformation of a signal from a time-
domain representation into a frequency-domain representation, has its roots in the
early 19th century, when several mathematicians were working on it from a
theoretical basis. But it took a practical man, an engineer with a good mathematical
background, to develop the rationale upon which almost all our modern spectrum
analysis techniques are based. That engineer was Jean Baptiste Fourier, and he was
working for Napoleon during his invasion of Egypt on a problem of overheating
cannons when he derived the famous Fourier Series for the solution of heat
conduction. It may seem a far cry from overheating cannons to frequency analysis,
but it turns out that the same equations apply to both cases. Fourier later
generalized the Fourier series into the Fourier Integral Transform. The advent of
digital signal analysis naturally led to the so-called Discrete Fourier Transform and
the Fast Fourier Transform or FFT
More:

Forms of the Fourier Transform
The Fourier Series
The Fourier Integral Transform
50
The Discrete Fourier Transform
The Fast Fourier Transform
Analog to Digital Conversion
Aliasing
Leakage
Windows
The Hanning Window
Overlap Processing
The Picket Fence Effect
Averaging
Time Synchronous Averaging
Pitfalls in the FFT
Forms of the Fourier Transform

There are four forms of the Fourier Transform, as follows:
Fourier Series -- Transforms an infinite periodic time signal into an infinite discrete
frequency spectrum.
Fourier Integral Transform -- Transforms an infinite continuous time signal into an
infinite continuous frequency spectrum
Discrete Fourier Transform (DFT) -- Transforms a discrete periodic time signal into a
discrete periodic frequency spectrum
Fast Fourier Transform -- A computer algorithm for calculating the DFT
They will be discussed in more detail in the next section.
The Fourier Series

The Fourier Series operates on a time signal that is periodic, i.e., a time signal whose
waveform repeats over and over again out to infinite time. Fourier showed that such
a signal is equivalent to a collection of sine and cosine functions whose frequencies
are multiples of the reciprocal of the period of the time signal. The rather unexpected
result is that any wave shape whatsoever, as long as it is not infinite in length, can
be represented, as the sum of a collection of harmonic components, and the
fundamental frequency of the harmonic series is 1 divided by the length of the wave
shape. The amplitudes of the various harmonics are called the Fourier coefficients,
and their values can be calculated easily if the equation for the wave shape is known.
They can also be calculated graphically from the wave shape itself. A certain physics
class is known to have done this with the silhouette of Marilyn Monroe. They posted
the MM coefficients on the bulletin board as an "in" joke.
Fourier Coefficients

The calculation of the Fourier coefficients is defined as a mathematical
transformation from the time domain to the frequency domain. One important fact
emerges from the Fourier Series, and that is that the original waveform can be
reconstructed from the frequency coefficients; in other words it is possible to
transform from the frequency domain back to the time domain without loss of
information. The Fourier series is perfectly adequate for performing frequency
analysis on periodic waveforms; that is to say on deterministic signals.

51
The Fourier Integral Transform

The natural extension of the Fourier series to encompass time signals of infinite
length, i.e., non-repetitive continuous signals, is the Fourier Integral Transform, or
more simply the Fourier Transform. This integral will transform any continuous time
signal of arbitrary shape into a continuous spectrum extending to infinite frequency.
An interesting characteristic of the Fourier Transform is that an event encompassing
a short time interval will be spread out over a wide frequency range and vice versa.
This was seen in the Introduction to Vibration chapter where a spectrum of a short
impulse is shown.
The Discrete Fourier Transform

Neither the Fourier Series nor the Fourier Transform lends itself easily to calculation
by digital computers. To overcome this hurdle, the so-called Discrete Fourier
Transform, or DFT was developed. Probably the first person to conceive the DFT was
Wilhelm Friederich Gauss, the famous 19th century German mathematician, although
he certainly did not have a digital computer on which to implement it. The DFT
operates on a sampled, or discrete, signal in the time domain, and generates from
this a sampled, or discrete, spectrum in the frequency domain. The resulting
spectrum is an approximation of the Fourier Series, an approximation in the sense
that information between the samples of the waveform is lost. The key to the DFT is
the existence of the sampled waveform, i.e., the possibility of representing the
waveform by a series of numbers. To generate this series of numbers from an analog
signal, a process of sampling and analog to digital conversion is required. The
sampled signal is a mathematical representation of the instantaneous signal level at
precisely defined time intervals. It contains no information about the signal between
the actual sample times.
If the sampling rate is high enough to ensure a reasonable representation of the
shape of the signal, the DFT does produce a spectrum very close to a theoretically
true spectrum. This spectrum is also discrete, and there is no information between
the samples, or "lines" of the spectrum. In theory, there is no limit to the number of
samples that can be used, or the speed of the sampling, but there are practical
limitations we must live with. Most of these limitations are the result of using a
digital computer as the calculating agent.
The Fast Fourier Transform

In order to adapt the DFT for use with digital computers, the so-called Fast Fourier
Transform (FFT) was developed. The FFT is simply an algorithm for calculating the
DFT in a fast and efficient manner.
Cooley and Tukey are credited with the discovery of the FFT in 1967, but it existed
much earlier, although without the digital computers needed to exploit it. The FFT
algorithm places certain limitations on the signal and the resulting spectrum. For
instance, the sampled signal to be transformed must consist of a number of samples
equal to a power of two. Most FFT analyzers allow 512, 1024, 2048, or 4096 samples
to be transformed. The frequency range covered by FFT analysis depends on the
number of samples collected and on the sampling rate, as will be explained shortly.


52

Analog to Digital Conversion

The first step in performing an FFT analysis is the actual sampling process, which is
illustrated here:
Analog to
Digital Conversion

The sampling is an analog, not digital, process and is accomplished with a "sample
and hold" circuit. The output of this circuit is a sequence of voltage levels that are
fed into an analog to digital converter (ADC). Here the voltage levels are converted
into digital words representing each sampled level. The accuracy of the sampled
levels depends in part on the number of bits in the digital words. The greater the
number of bits, the lower the noise level and the greater the dynamic range will be.
Most FFT analyzers use 12-bit words and this produces a dynamic range of about 70
dB (3,100:1). Fourteen bit words can achieve 80 dB (10,000:1) dynamic range.
It can be seen here that the sampling rate determines the highest frequency in the
signal that can be encoded. The sampled waveform cannot know anything about
what happens in the signal between the sampled times. Claude Shannon, the
developer of the branch of mathematics called information theory, determined that
to encode all the information in a signal being sampled, the sampling frequency must
be at least double the highest frequency present in the signal. This fact is sometimes
called the Nyquist criterion.
Aliasing

It is important that there is no information in the sampled waveform near the
sampling frequency to avoid a problem called aliasing.
53


Aliasing
Here the actual signal is represented in black and the sampled representation of it is
in gray. The vertical lines represent the sampling frequency. Note that if the
sampling frequency is the same as the sampled frequency, each sample is the same
size, and the output of the sampling circuit will be a constant direct voltage --
obviously having no relation to frequency of the input signal.
Now note what happens if the actual signal is higher in frequency than the sampling
frequency. The sampler output looks like a very low frequency, and again it is not a
correct representation of the actual signal. This phenomenon is called aliasing, and it
can lead to gross errors unless it is avoided. The best way to avoid aliasing is to pass
the input signal through an analog low-pass filter whose cut-off frequency is less
than one-half the sampling frequency. In most modern FFT analyzers, the sampling
frequency is set to 2.56 times the filter cut-off frequency. The filter must have a very
sharp cut off characteristic, or roll off, and this means it will also have Phase Shift
that can affect the data if one needs phase information near the upper end of the
frequency span of the analyzer. To avoid this, select a frequency span so the
frequency in question is in the lower half of the frequency range. This is important in
performing balancing with an FFT analyzer, where phase of the 1X vibration signal is
needed.
Aliasing also occurs in other media, such as motion pictures. For instance, sometimes
in western movies the wagon wheel spokes may appear stopped, or rotating
backward. This is optical aliasing, for a movie is a sampled representation of the
original motion. Another example of optical aliasing is the stroboscope, which is set
to flash at a rate equal to or near the rotation rate of the object being observed,
making it appear stationary or slowly turning.
Sampling Rules for Digital Signal Analysis
The data path must contain an analog Anti-Aliasing low-pass filter
You must sample at least twice as fast as the highest frequency to be analyzed
The Frequency Response of the analysis depends on the sampling frequency
These rules apply to all FFT analysis, and the analyzer automatically takes care of
them. The anti-aliasing filter is internally set to the appropriate value for each
frequency range of the analyzer. The total sampling time is called the time record
length and the nature of the FFT dictates that the spacing between the frequency
54
components in the spectrum (also called the frequency resolution) is 1 divided by the
record length. For instance, if the frequency resolution is one Hz, then the record
length is one second, and if the resolution is 0.1 Hz, then the record length is 10
seconds, etc. From this it can be seen that in order to perform high resolution
spectrum analysis relatively long times are required to collect the data. This has
nothing to do with the speed of the calculations in the analyzer; it is simply a natural
law of frequency analysis.
Leakage

The FFT analyzer is a batch processing device; that is it samples the input signal for
a specific time interval collecting the samples in a buffer, after which it performs the
FFT calculation on that "batch" and displays the resulting spectrum
If a sinusoidal signal waveform is passing through zero level at the beginning and
end of the time record, i.e., if the time record encompasses exactly an integral
number of cycles of the waveform, the resulting FFT spectrum will consist of a single
line with the correct amplitude and at the correct frequency. If, on the other hand,
the signal level is not at zero at one or both ends of the time record, truncation of
the waveform will occur, resulting in a discontinuity in the sampled signal. This
discontinuity is not handled well by the FFT process, and the result is a smearing of
the spectrum from a single line into adjacent lines. This is called "leakage"; it is as if
the energy in the signal "leaks" from its proper location into the adjacent lines.
The shape of the "leaky" spectrum depends on the amount of signal truncation, and
is generally unpredictable for real signals.







55

Windows

In order to reduce the effect of leakage, it is necessary to see to it that the signal
level is zero at the beginning and end of the time record. Multiplying the data
samples by a so-called "windowing or "weighting function, which can have several
different shapes, does this. The most common forms of windows and their uses are
considered next.


If there is no windowing function used, this is called "Rectangular", "Flat", or
"Uniform" windowing. In the figure above, the effect of the data truncation can be
seen as discontinuities in the windowed waveform. The FFT analyzer only knows
what is in the time window, or time record. It assumes the actual signal contains the
discontinuities, and they are the cause of the leakage seen in the previous figure.
Leakage could be avoided if the input waveform zero crossings were synchronized
with the sampling times, but this is impossible to achieve in practice.
More:

Windowing for Transient Signals
Windowing for Transient Signals



In the case where the input signal is a transient, it will by definition begin and end at
zero level, and as long as it is entirely within the time record, no truncation will occur,
and the analysis will be correct because the FFT sees the entire signal. It is very
important that the entire transient fit into the record, and the record length is
dependent upon the frequency range of the analysis. Most FFT analyzers allow the
user to see the time record on the screen, so it can be assured that this condition is
met.

56

The Hanning Window




The Hanning window, after its inventor whose name was Von Hann, has the shape of
one cycle of a cosine wave with 1 added to it so it is always positive. The sampled
signal values are multiplied by the Hanning function, and the result is shown in the
figure. Note that the ends of the time record are forced to zero regardless of what
the input signal is doing.
While the Hanning window does a good job of forcing the ends to zero, it also adds
distortion to the wave form being analyzed in the form of amplitude modulation; i.e.,
the variation in amplitude of the signal over the time record. Amplitude Modulation in
a wave form results in sidebands in its spectrum, and in the case of the Hanning
window, these sidebands, or side lobes as they are called, effectively reduce the
frequency resolution of the analyzer by 50%. It is as if the analyzer frequency "lines"
are made wider. In the illustration here, the curve is the actual filter shape that the
FFT analyzer with Hanning weighting produces. Each line of the FFT analyzer has the
shape of this curve -- only one is shown in the figure.
If a signal component is at the exact frequency of an FFT line, it will be read at its
correct amplitude, but if it is at a frequency that is one half of delta F (One half the
distance between lines), it will be read at an amplitude that is too low by 1.4 dB.
The illustration shows this effect, and also shows the side lobes created by the
Hanning window. The highest-level side lobes are about 32 dB down from the main
lobe.

57
The measured amplitude of the Hanning weighted signal is also incorrect because the
weighting process removes essentially half of the signal level. This can be easily
corrected, however, simply by multiplying the spectral levels by two, and the FFT
analyzer does this job. This process assumes the amplitude of the signal is constant
over the sampling interval. If it is not, as is the case with transient signal, the
amplitude calculation will be in error, as shown in the figure below.



The Hanning window should always be used with continuous signals, but must never
be used with transients. The reason is that the window shape will distort the shape of
the transient, and the frequency and phase content of a transient is intimately
connected with its shape.
The measured level will also be greatly distorted. Even if the transient were in the
center of the Hanning window, the measured level would be twice as great as the
actual level because of the amplitude correction the analyzer applies when using the
Hanning weighting.
A Hanning weighted signal actually is only half there, the other half of it having been
removed by the windowing. This is not a problem with a perfectly smooth and
continuous signal like a sinusoid, but most signals we want to analyze, such as
machine vibration signatures are not perfectly smooth. If a small change occurs in
the signal near the beginning or end of the time record, it will either be analyzed at a
much lower level than its true level, or it may be missed altogether. For this reason,
it is a good idea to employ overlap processing. To do this, two time buffers are
required in the analyzer. For 50% overlap, the sequence of events is as follows:
When the first buffer is half full, i.e., it contains half the samples of a time record,
the second buffer is connected to the data stream and also begins to collect samples.
As soon as the first buffer is full, the FFT is calculated, and the buffer begins to take
data again. When the second buffer is filled, the FFT is again calculated on its
contents, and the result sent to the spectrum-averaging buffer. This process
continues on until the desired number of averages is collected.


58
Overlap Processing

Overlap processing can only be achieved if the time required to calculate the FFT is
shorter than the time record length. If this is not the case, the spectral calculations
will lag behind the data acquisition leaving gaps of unanalyzed signal. See also the
paragraph on real time speed later in this section.

If the overlap is 2/3, i.e., 66.7%, then the overall time weighting of the data will be
flat, and there is no advantage to using a greater overlap. Most data collection for
machinery analysis uses 50% data overlap, which provides adequate amplitude
accuracy for most vibration work.
Here is a summary of the relationship between sampling rate, number of samples,
time record length, and frequency resolution that affect FFT analysis. The sampling
rate in samples per second, times the time record length T in seconds, equals the
number of samples N. In the FFT analyzer, the number of samples N is constrained
to a power of two.


FFT Fundamentals
59

The FFT algorithm, operating on N samples of time data produces N/2 frequency
lines. Thus a time record of 512 samples will generate a spectrum of 256 lines. FFT
analyzers generally do not display the upper spectral lines because of the possibility
of their being contaminated by aliased components. This is because the anti-aliasing
filter is not perfect, and has a finite slope in its cut-off range. Therefore, a 256 line
spectrum will be displayed as a 200 line spectrum, and a 512-line spectrum will be
displayed as a 400 line spectrum, etc.
The frequency resolution, DF, is equal to the frequency span divided by the number
of lines, and this is equal to 1/T. Conversely, the time record length T equals 1/DF.
From this it can be seen that as the frequency resolution increases (smaller DF), the
time record length also increases in proportion. For this reason, to create a high-
resolution spectrum requires a relatively long time to acquire the data.
The Picket Fence Effect

As has been mentioned before, the FFT spectrum is a discrete spectrum, consisting
of estimates of what the spectral level is at specific frequencies. These frequencies
are determined by the analysis parameters that are set up in the analyzer, and have
nothing to do with the signal being analyzed. This means there may be, and probably
are, peaks in the true spectrum of the signal that are between the lines of the FFT
analysis. This also means that in general, the peaks in an FFT spectrum will be
measured too low in level, and the valleys will be measured too high. Moreover, the
true frequencies where the peaks and valleys lie will not be those indicated in the
FFT spectrum.




This phenomenon is called resolution bias error, or more commonly, the picket fence
effect. In other words, looking at an FFT spectrum is a little like looking at mountain
range through a picket fence.
Averaging

One of the important functions of the FFT analyzer is that it is easily able to do
averaging of spectra over time. In general, the vibration signal from a rotating
machine is not completely deterministic, but has some random noise superimposed
on it. Because the noise is unpredictable, it alters the spectrum shape, and in many
cases can seriously distort the spectrum. If a series of spectra are averaged together,
the noise will gradually assume a smooth shape, and the spectral peaks due to the
deterministic part of the signal will stand out and their levels will be more accurately
60
represented. It is not true that simply averaging FFT spectra will reduce the amount
of the noise -- the noise will be smoothed but its level will not be reduced.
There are two types of averaging in general use in FFT analyzers, called linear
averaging and exponential averaging. Linear averaging is the adding together of a
number of spectra and then dividing the total by the number that was added. This is
done for each line of the spectra and the result is a true arithmetic average on a line-
by-line basis. Exponential averaging generates a continuous running average where
the most recently collected spectra have more influence on the average than older
ones. This provides a convenient form to examine changing data but still have the
benefit of some averaging to smooth the spectra and reduce the apparent noisiness
of them.
Time Synchronous Averaging

Time synchronous averaging, also called time domain averaging, is a completely
different type of averaging, where the waveform itself is averaged in a buffer before
the FFT is calculated. In order to do time domain averaging, a reference trigger pulse
must be input to the analyzer to tell it when to start sampling the signal. This trigger
is typically synchronized with an element of the machine that is of interest.
The average gradually accumulates those portions of the signal that are
synchronized with the trigger, and other parts of the signal, such as noise, are
effectively averaged out. This is the only type of averaging which actually does
reduce noise.
More information on applications of time synchronous can be found in the next
chapter on Machine Vibration Monitoring.
Pitfalls in the FFT

This is a summary of the pitfalls that plague the FFT analysis technique. This is not to
say that FFT analysis is no good -- on the contrary, it has revolutionized the analysis
of vibration data. The important fact is that the problems with FFT analysis can be
overcome by proper technique, and the residual effects that remain can be reduced
to insignificant levels.
Sampling causes aliasing
Time limitation causes leakage
Discrete frequencies in the calculated spectrum causes the picket fence effect.
Machine vibration monitoring
Introduction

It has been shown many times over that the vibration signature of an operating
machine provides far more information about the inner workings of the machine than
any other type of non-destructive test. A bearing that has a small developing defect
will cause a telltale change in the machine vibration, as will an imbalance condition,
a misalignment, or any of a myriad of other faults. Vibration analysis, properly
applied, allows the technician to detect small developing mechanical defects long
before they become a threat to the integrity of the machine, and thus provides the
necessary lead-time to schedule maintenance to suit the needs of the plant
management. In this way, plant management has control over the machines, rather
than the other way around.
61
Vibration measurement and analysis is the cornerstone of Predictive Maintenance,
which stands in sharp contrast to the historical "run-to-failure" type of maintenance
practice. Numerous studies, such as those conducted by the Electric Power Research
Institute (EPRI), have shown that on average, the cost to industry for maintenance
will be reduced by more than 50% if a predictive maintenance program is used
instead of run-to-failure.
History of Vibration Analysis used for
Machinery Maintenance


The first vibration meters were introduced in the 1950s, and they measured the
overall, or "broad band" level of machine vibration, either in peak-to-peak mils
(thousandths of an inch) of vibratory displacement, or in inches per second (IPS) of
vibration velocity. A little later, tunable analog filters were added to the meters in
order to discriminate between different frequency components, and thus to produce
a sort of vibration spectrum.
The 1970s brought forth the personal computer and the advent of digital signal
processing that led to the FFT analyzer, and it made quick work of calculating a
frequency spectrum from a recorded vibration signal. The first such analyzers were
quite bulky, weighing as much as 75 pounds, and this made them more suited as
laboratory instruments than portable units for field use.
The 1980s saw the exploitation of the microprocessor on a single silicon chip, and
the battery-powered truly portable digital signal analyzer quickly followed this. It is
this device, coupled with a computer program that stores the data and takes care of
the logistics of vibration data collection that has revolutionized the application of
vibration analysis to machinery diagnostics.
Practical Aspects of Vibration Measurement

More:

Test Point Location
Vibration Sensor Orientation
Triaxial Measurements
Orientation Examples
Sensor Mounting Pads -- "Blocking"
Vibration Surveys
62
Test Point Location

In general, it is desirable to locate the test transducer as close as possible to the
bearing with solid metal between the bearing and the sensor. Avoid bearing caps,
which are of thin metal and are thus poor conductors of vibration energy. If possible,
pick test point locations so that there is no metal-to-metal joint between the bearing
and the sensor. The joint between the end bell and stator housing of a motor is an
example of this. Fan housings on the ends of motors are also to be avoided.


In general, it has been found that for motors of less than about 50 HP, one test point
is adequate, but for motors over 50 HP, each bearing should have its own test point.
In any machines that are especially sensitive to bearing damage, and bearing
problems should be detected as early as possible, each bearing should have its own
test point.
Another consideration in the integrity of the path between the bearing itself and the
transducer: If the motor and bell is a solid casting, it will effectively transmit
vibration with little loss of high frequencies, but if it contains one or more metal to
metal connections, the high frequencies will be significantly distorted.
Vibration Sensor Orientation

In any machinery-monitoring program, it is extremely important that the data is
collected in exactly the same manner each time a measurement is taken. This is to
assure that the data is repeatable and can be trended over time. For this reason, it is
not recommended that hand-held transducers be used. By far the most reliable data
is collected when the transducer is stud mounted to the machine surface.
Triaxial Measurements

To assist in the determination of machine problems, it is very helpful to have
vibration data from each measurement point in three directions. These directions are
called Axial, Radial, and Tangential. Axial is the direction parallel to the shaft in
question, radial is the direction from the transducer to the center of the shaft, and
tangential is 90 degrees from radial, tangent to the shaft.
63



Alignment of Vibration Axes
Orientation Examples

The following diagram shows the six possible orientations of the sensor for a
horizontal machine.


For vertical machines, 'R' is Radial, 'T' is Tangential, and 'A' is vertical:
Sensor Mounting Pads -- "Blocking"

When using a triaxial accelerometer, it is extremely important that it be installed in
exactly the same location each time the data is collected, and also that it be oriented
in the same direction. One way of assuring this is to use permanently affixed
mounting blocks on the machine.
The cylindrical mounting block, or "pad", is a bronze disc with a central tapped hole
and a key way at the edge that receives an indexing pin on the transducer itself. The
transducer that is sensitive along the axis of the mounting screw is channel No. 1,
the axis in the direction of the key way is channel No. 2, and the axis perpendicular
to this is channel No. 3. The pad is normally attached to the machine with a hard,
strong adhesive such as Versilok type 204 structural adhesive.
As was mentioned above, it is very important that the orientation of the block is
known by the software, and if a block is replaced, the new one must be oriented in
the same direction. The VTAG states the proper orientation of each block. The
installation of the mounting blocks is sometimes referred to as "blocking" a machine.
64


Sensor Mounting Pad
Vibration Surveys

When performing a vibration survey of a group of machines, the following points
should be considered in order to assure consistency of the data from one
measurement time to the next.
More:

Test Conditions
Operating Conditions
Warm-up
Visual Inspection
Test Conditions

The vibration signature of a machine is strongly dependent on the operating
parameters as well as its physical condition. These operating parameters include
such things as running speed, load, pump discharge pressure, and compressor
delivery pressure.
The machine must be in its normal operating condition when vibration data is
collected. If this is not the case, the vibration signature will not match the vibration
signatures previously recorded, and trending vibration levels over time becomes
impossible. Running speeds of induction motors depend on the load, and should not
vary from one collection time to the next by more than a few percent. This means
that load conditions must be as nearly as possible the same.

The vibration level contributed by extraneous sources, such as nearby machines,
must also be the same for each data collection time. Do not collect data with
adjacent machines turned off if the previous spectra were recorded with them
running. This is especially true with strong background vibration levels, as in the
engine room of a ship. Propulsion diesels must be operating at the same speed for
each data collection session!
Operating Conditions

It is imperative that when collecting data, the test RPM is very near the RPM that
was used for the previous tests. In turbine-driven equipment, the speed should be
65
verified by the use of a portable stroboscopic or other tachometer, and it must be
running at a constant, not varying, speed!
Gauge pressures should reflect normal operating conditions. Pump testing with
discharge valves closed is discouraged, but if a pump must be tested in a
recirculating condition, the recirculation valve may be partially closed to attain a
normal discharge pressure.
Warm-up

All machines should be tested in a fully warmed-up condition. Machine temperature
will affect alignment and operating clearances due to thermal expansion. A cold
machine will have a different vibration signature than a warmed-up machine,
sometimes extremely different.
Visual Inspection

Visual inspection of an operating machine while vibration testing is important, for
valuable clues to machinery condition can often be uncovered. RPM and discharge
pressure, etc., should be noted. The following items should be checked:
Are there any unusual noises present?
Do any bearings feel hotter than normal?
Can you feel any excessive vibration level?
Is there anything unusual about the operation of the
machine?
Are there any fluid or steam leaks obvious?
Do the gauge readings look normal?
Does the machine operator have any comments on
machine condition?
The Concept of Spectrum Comparison

More:

Vibration Measurement Parameters
Machinery Testing Schedule
Trending of Vibration Data
The Reference Spectrum
Forcing Frequencies
Order Normalization
Evaluating Machine Vibration Spectra
Vibration Measurement Parameters

As we saw in the Introduction to Vibration chapter, it is possible to examine the
same vibration signal in terms of Acceleration, Velocity, or Displacement. It is seen
that velocity at any frequency is proportional to the displacement times the
frequency, and the acceleration at any frequency is proportional to velocity times
frequency, which means it is also equal to displacement times frequency squared.
66


Machinery Testing Schedule

It is important to begin a vibration-monitoring program of manageable size and then
gradually expand it as you gain experience. The most important machines to monitor
should be those that are critical to the plant's productivity and/or have a poor
maintenance record. Variable speed machines, extremely complex machines, and
reciprocating machines should not be included at first.
For a successful monitoring program, machinery measurements must be carried out
on a scheduled periodic basis. Most equipment should be tested monthly, with
certain less important machines on a 3-month schedule. Weekly testing is common
for critical machines. In any case, it is important to tailor your measurement
schedule to suit the machines and their condition. As experience is gained, it will be
easy to revise the testing schedule accordingly.
Trending of Vibration Data

Trending is the storage of vibration signatures recorded at specific time intervals and
plotting the changes in vibration levels at the forcing frequencies vs. time. An
upward trend in level indicates a developing problem.
The simplest way to utilize the concept of vibration trending is to establish a
representative vibration spectrum of a normally operating machine as a reference,
and compare this reference to spectra measured at later times on the same machine.
The comparison of the spectra is made possible by order normalization, which will be
discussed shortly. When performing the spectral comparison there are several
important points that need to be addressed:
67
-
The operating conditions of the machine when measuring the new vibration data must
match as closely as possible the conditions under which the reference spectrum was
recorded. Otherwise, the spectra will not be comparable and gross errors can be made.
-
The vibration data must be recorded in exactly the same way that the reference data
were measured. The transducer must be mounted in exactly the same location, and its
calibration must be accurate. If possible, the same transducer should be used for all
successive measurements on the machine.
-
When taking vibration data with an FFT analyzer, or data collector, it is important to
average several instantaneous spectra together to reduce random variations and the
effects of extraneous noise in the measured signal. The number of spectral averages
recorded to produce the spectra must be sufficient to produce a uniform and steady
signature. Usually from six to ten averages will do this, but on some machines with a
relatively high random noise content in their vibration signature, longer averaging times
may be needed. A rule of thumb is to record a spectrum with several averages and then
immediately record another one with twice as many averages. If the spectra are
significantly different, the number of averages should be doubled again and another
spectrum recorded. If the latter two spectra are similar, then the previous number of
averages is adequate for this machine.
The Reference Spectrum

When performing trending, it is extremely important to be sure that the reference
spectrum to which the subsequent test spectra will be compared is truly
representative of the machine.
More:

Averaged Vibration Signatures
The Spectrum Mask
Averaged Vibration Signatures

Long experience has shown that an excellent way to generate a meaningful reference
is to average several spectra together from machines of the same type. If there are
a number of similar machines in a plant the statistical average of their reference
spectra is a good indication of the overall characteristics of that particular machine. A
series of similar machines in good working order will produce vibration spectra that
are similar to one another, but will have random variations in level. The spectra of
the machines are averaged together and the standard deviations in level at each
significant frequency are calculated.
Some types of machines are so individualistic that when averaged together, the
standard deviation between the vibration magnitudes is so great that the average is
essentially meaningless. In this case, each machine must be used by itself to
produce a meaningful reference by averaging a series of measurements over a fairly
long time period, and generating a mask from this average reference spectrum.
There are many situations where a large selection of similar machines is not
available, and in this case, the averaged reference spectra are taken on the same
machine at different times. When averaging spectra from a group of machines to
make a reference spectrum, care must be taken to see that the spectra to be
averaged are valid and that the machines they come from are not defective. One of
the most important jobs of the vibration analyst is to be sure that the average
reference spectra are valid and representative of the machines in question. Do not
68
confuse reference spectrum averaging to produce a reference signature with
spectrum averaging done at the time of vibration data collection, as described above.
The Spectrum Mask

As we have seen, healthy machines will show minor deviations in their vibration
spectra because of small load variations, temperature variations, line voltage
variations, and background noise level fluctuations. These variations in vibration
signatures can cause false alarms to be generated if the raw spectrum is directly
compared to a valid reference spectrum. For this reason, it is desirable to generate a
so-called mask spectrum from the reference spectrum. The mask is a new spectrum
made by increasing the levels in the reference spectrum by various amounts at
different frequencies. For instance the mask might be 6 dB above the reference at 1X,
but only 4 dB above the reference at 2X.
A good staring point for establishing the mask is to add one standard deviation in
level at each spectral peak to the averaged reference spectrum. A large class of
machines will be found to produce averaged spectra with fairly small standard
deviations, and with these machines in particular it is a good idea to perform the
spectrum averaging and then generate the mask by adding one standard deviation
to the average spectrum at each frequency. A group of machines which exhibits
large standard deviations in level when making the reference will be more difficult to
deal with in generating the mask, and the mask levels will have to be higher than
one standard deviation above the reference.
The determination of the shape of the mask spectrum can be fairly complicated, and
it depends on the machine in question and normal variation in its vibration spectral
levels at different frequencies. This can only be determined by looking at a series of
historical spectra and applying good judgment and a good knowledge of the machine
itself.
Forcing Frequencies

The value of vibration analysis of machinery is based on the fact that specific
elements in the rotating parts of any machine will produce forces in the machine that
will cause vibration at specific frequencies. One of the most important of the forcing
frequencies is the RPM of the shaft, and it arises from the fact that any rotor will
always have a certain amount of residual imbalance. This imparts a radial centripetal
force on the bearings, causing the structure to vibrate at the 1X, or fundamental,
frequency. The so-called bearing tones, which are characteristic of each bearing
geometry, are forces generated by defects in the races and rolling elements of the
bearing itself. Gear tooth-mesh frequencies come from the individual impacts of gear
teeth against each other, and the tooth-mesh frequency is equal to the number of
teeth on the gear times the gear RPM. Vane pass or blade pass frequencies are
similar to tooth mesh and are equal to the number of vanes in an impeller or number
of blades in a fan times the RPM. Each forcing frequency will create a peak in the
vibration spectrum, the amplitude of the peak being dependent on the severity of the
condition that causes it. Thus the frequency indicates the type of problem and the
amplitude indicates its severity.
As an example of a simple forcing frequency, the ceiling fan illustrated below would
produce vibration component each time a blade struck the fly swatter, giving rise to
a peak in the spectrum at 5 times the turning speed.
69

The figure below, showing a centrifugal air compressor, illustrates some of the
forcing frequencies in the spectrum.

Following is an example of forcing frequency calculation for a gear-driven machine:


Let us assume that the motor/gear/fan components have the following element
counts:
Machine
Component
Elements of
Component
Number of
Elements
Motor Cooling Fan

Fan Blades

11

Motor Rotor

Rotor Bars

42

Drive Pinion

Gear Teeth

36

Driven Gear

Gear Teeth

100

Fan

Fan Blades

9

In this case of a multiple shaft machine, we must consider that the fundamental
frequencies of the motor and fan shafts are different. Let us assume that the motor
is again running at 1780 RPM. To calculate the fan shaft RPM, we must first find the
reduction ratio of the gearbox. To find this we would look at the number of gear
teeth on each of the gears. Divide the drive pinion tooth count by the driven gear
tooth count:

or

70


Next, multiply this ratio by the motor shaft RPM to find the fan shaft RPM;





We would now say that the fundamental frequency of the motor is 1780 CPM and the
fundamental frequency of the fan is 640.8 CPM.
We multiply the number of elements on each component by the fundamental
frequency of the shaft from which it rotates. The components that are on the motor
shaft will be multiplied by 1780 CPM and the components on the fan shaft will be
multiplied by 640.8 CPM. To make this easier, let us separate the components with
their corresponding shafts:
Motor Shaft Elements Forcing Frequency,
CPM
Rotation

1

1,780

Motor Cooling Fan

11

19,580

Motor Rotor

42

74,760

Drive Pinion

36

64,080


Fan Shaft Elements Forcing Frequency
Rotation

1

640.8

Driven Gear

100

64,080

Fan

9

5,767.2


The Frequency Axis
When plotting vibration spectra from rotating machines, you have several choices of
units for the frequency axis. Probably the most natural unit is the cycle per second,
or hertz (Hz). Another unit in common use is Revolutions Per Minute (RPM), or
Cycles per Minute (CPM). Hz is converted to CPM by multiplying by 60. Many people
feel that CPM is a convenient scale to use because the machines are described in
terms of RPM. This practice results in quite large numbers for the frequency axis,
however, and many other people prefer to use Hz because the smaller numbers are
more convenient.




71
Order Normalization

Order normalization is performed by most condition monitoring software, and under
certain conditions it is possible for the software to select the wrong peak as the 1X
component. For this reason it is important for the analyst to verify that the
normalization was correctly done if a spectrum looks vastly different than other
spectra taken from the same machine. In such a case, the analyst must re-normalize
the spectrum.
Following is a non-order normalized spectrum, scaled from zero to 30,000 RPM.


Conventional Vibration Spectrum

Note that many peaks appear to be equally spaced, but it may be difficult to tell
which one near 20,000 CPM is a shaft harmonic.
The next figure is a normalized spectrum scaled from 0 to 10 orders. Note that the
harmonics of turn speed are integers on the frequency scale, and that the peak
below 7X is immediately seen as a non-synchronous component.



Order Normalized Spectrum

72
Order normalization of spectra has the following
advantages:

The fundamental turn speed is instantly recognizable
at 1.0 order.
Harmonics of the turn speed will be integers
A second shaft in a gear-driven machine will have an
order equal to the gear ratio
Excitation frequencies, such as gear mesh and pump
vane pass, are readily recognized because their
order is equal to the number of elements
Bearing tones will be non-integer, often the only
major non-integer components
Sidebands around bearing tones will be easily
recognized because they will be at the tone order
1, 2, etc.
Most important: Because machine speed is almost
never exactly the same from test to test, the peaks
in the spectrum will not be at the same frequencies,
and the spectra cannot be averaged. Normalized
spectra have the peaks at the same orders from test
to test, and they can be averaged
Evaluating Machine Vibration Spectra

Most machines have a relatively simple set of vibration forcing frequencies,
determined by the geometry of the machine and its speed. The existence of other
frequencies than the forcing frequencies, such as harmonics of 1X, in the vibration
signature of the machine indicates non-linearities, and the combined magnitude of
these new frequencies is a good indicator of the overall health of the machine. As a
machine wears, its clearances typically become greater and its vibration signature
becomes more complex due to generation of harmonics and sidebands.
In trending the vibration level of a machine over time, a rise in the level of a forcing
frequency indicates a change in the mechanism causing that particular forcing
frequency, but does not necessarily indicate any damage to the machine. For
instance, an increase in 1X at a motor bearing indicates an increasing imbalance
condition, but if harmonics of 1X begin to appear, this indicates damage, such as
bearing clearance increases, looseness, or cracking of the structure. Therefore, a
strong 1X vibration means the rotor should be balanced, but the appearance of
harmonics of 1X means the bearing and surrounding structure should also be
inspected for damage.
Machine vibration analysis
Machine Vibration Analysis

More:

Introduction
Time Domain Analysis
Cepstrum Analysis
Statistical Properties of Vibration Signals
Amplitude Demodulation
73
Root Cause Failure Analysis
Introduction

The steps in manual (non-automatic) machine vibration analysis are:
-
Identifying vibration peaks in the spectrum and relating them to forcing frequencies
-
Determining the severity of machine problems from the amplitudes and relationships
between the vibration peaks.
-
Making the appropriate repair recommendations based on the severity of the machine
problem.
In order to do a proper job of vibration analysis, several tools are needed: If the
vibration spectra are being analyzed on a computer, a calculator and Vibration Test
and Analysis Guide (VTAG) for the machine in question are required. If the vibration
spectra have been printed on paper, then a straight edge and ten-point divider are
desirable. Previous vibration data and average vibration data are also helpful if
available.
More:

The Vibration Test and Analysis Guide (VTAG)
Checking for Data Validity
Step-by-Step Analysis of Spectra
Identifying the First Order (1X) Peak
The Vibration Test and Analysis Guide (VTAG)

The VTAG contains important information about the design of the machine, the test
points and their locations, the frequency ranges to be tested, and the forcing
frequencies to be expected. The VTAG should be consulted before any vibration
analysis is attempted.
Following is an example of a VTAG:
74

Checking for Data Validity

After determining the shaft rotation rate and locating it on the spectrum (it will be
the first order in a normalized spectrum), the vibration analyst must check the
validity of the spectrum. Data validity can be corrupted by such things as incorrect
labeling of accelerometer orientation or position, improper accelerometer attachment,
rapid accelerometer temperature changes, and incorrect machine operating
conditions.
75
When data are to be compared to previously collected data from the same point,
similar test conditions must be maintained, especially machine speed, load, and
operating temperature.

The integrity of the accelerometer cable is crucial to the collection of valid data. If
the central conductor in the cable is intermittent or open, the measured signal will
consist mostly of random noise, and if the cable shield is intermittent or broken, the
data will be contaminated with 60 Hz noise and harmonics. (50 Hz in countries with
50 Hz power lines.) In electrically driven machines, the 60 Hz line frequency will
produce a series of 120 Hz harmonics in the vibration spectrum, as explained in the
section on electrically induced vibration.
If an accelerometer is exposed continuously to a higher temperature than that for
which it is rated, it will become desensitized, and the data it senses from then on will
be worthless. Some accelerometers will operate up to 400 degrees F, but most give
up the ghost at about 200 degrees F.
Care must be taken that the accelerometer is not dropped onto a hard surface lest
the piezo-electric element be damaged. If the element is cracked, the stiffness of the
internal assembly will decrease, reducing the resonant frequency of the
accelerometer, and this can greatly change its sensitivity at high frequencies.
Step-by-Step Analysis of Spectra

In preparation for the diagnostic techniques described in the next chapter, the first
steps of analysis should be performed as follows:
This procedure assumes the vibration spectra are printed on paper. When viewing
spectra on the computer screen, similar procedures are used, as explained in the
software instructions.
Note that all the following steps are greatly simplified if the spectra are order
normalized.
Identifying the First Order (1X) Peak

The first step in machine vibration analysis is to identify the spectral peak
corresponding to shaft rotation rate, or the so-called 1X peak. This will be the 1X in a
normalized spectrum. It is important to check to be sure the normalization was done
correctly. It is also called the first order peak. In multiple-shaft machines, each shaft
will have a characteristic 1X peak, and these are then located by the analyst.
More:

Single-shaft machine
Multiple-shaft machine
Single-shaft machine

Mark the harmonics of 1X on the spectra. This is simplified if you use a ten-point
divider.
Identify the fan blade pass frequency and mark it on the spectra. This is the number
of blades multiplied by the RPM. Note the harmonics of blade pass frequency if they
are prominent.
Look for bearing tones, which are between the harmonics of the 1X run speed and
not synchronous with it. Mark them on the spectra. There are other machine
components besides bearings that generate non-synchronous tones. Probably the
most common one is belt drives.
Multiple-shaft machine

76
Identify and mark the 1X and harmonics of the pump on the spectra. The pump RPM
can be found from the VTAG, or can be calculated from the motor speed and gear
ratio as follows: If the motor is turning 1780 and the gear ratio is 2.3 to 1, then the
pump speed is:



Identify and mark the pump vane pass frequency and harmonics, if any, on the
spectra. The vane pass is the number of vanes times the pump RPM.
Search the spectra for non-synchronous components that could be bearing tones, or
consult the VTAG for bearing tone frequencies, and mark them on the spectra.
After performing all these tasks, you are ready for the work described in next
chapter on Machine Diagnostics.

Time Domain Analysis

More:

The Waveform vs. the Spectrum
FFT Analyzer Setup for Waveform Collection
Acceleration vs. Velocity
Phase in the Time Domain
The Wave Form as an analytical tool
Synchronous Averaging
Analyzer Set-Up for Synchronous Averaging
Case Histories using Synchronous Averaging
The Waveform vs. the Spectrum

Time Domain Analysis is simply the use of the waveform instead of the spectrum to
help diagnose machine problems. As we learned in the frequency analysis section of
the Vibration Fundamentals course, the spectra of an impulse or transient and of a
random signal may look almost exactly alike. This is true even though the parent
time signals are very different in character.
The waveform immediately shows the difference, however, and therefore it is a good
idea for the analyst to examine the waveform when the spectrum may not provide all
the information needed to make a complete diagnosis.
FFT Analyzer Setup for Waveform Collection

When setting up an analyzer to store waveforms, an important point should be born
in mind, and that is that the frequency range normally convenient for looking at a
spectrum is usually not suited to looking at the waveform. Most FFT analyzers, with a
few notable exceptions, do not allow you to set up specific sampling rates or time
domain record lengths - you must set them up in terms of frequency span and
frequency resolution. Remember from the FFT Analysis chapter that the time record
length used by the analyzer to calculate the spectrum is the reciprocal of the line
spacing, or resolution, of the spectrum.
Spectra are generally scaled so relatively wide frequency ranges can be examined,
and the FFT analyzer of necessity acquires a short time record. For instance, a 400-
77
line spectrum extending from DC to 1000 Hz will have a line spacing of 1000/400, or
2.5 Hz. The time record length used to calculate this spectrum is 1/2.5, or 0.4
seconds. This time record, which is the actual waveform, will show details that
happen in that 0.4-second time span, but in practice, when looking at a machine
vibration waveform, we are often looking for events that occur over a much longer
time than that. If we are looking for beats in the vibration signature of an electric
motor, or of the combined vibration of two machines running at slightly different
speeds, we need to see a waveform that is at least several seconds long.
To acquire a waveform lasting five seconds, we need to set up a line spacing of 1/5
Hz, and this can be done by adjusting the number of lines of resolution and the
frequency span to suitable values.
To find out the sampling rate of the waveform, and thus it time resolution, again we
need to get the information from the spectrum characteristics. The sampling
frequency of the time record for most analyzers is 2.56 times the highest frequency
in the spectrum. Thus a frequency span of 100 Hz implies a sampling frequency of
256 samples per second, and a span of 1000 Hz requires a sampling rate of 2560
samples per second.
Remember that a meaningful time record contains many more data points that the
usual spectrum, and therefore you need to take care that you have enough memory
available in your data collector to store the waveform data. For this reason, it is best
to use the lowest sampling rate and the shortest time record length that will provide
the needed data. For example if you just want to resolve beats in waveform that only
occur once in several seconds, the sampling rate need not be very high - 50 samples
per second is probably fast enough. This corresponds to a frequency span of 50/2.56,
or 19.53 Hz. So you can select 20 Hz in the frequency span set up.
On the other hand, if you want to examine a waveform that might have interesting
glitches at 50 times per second, then you need to sample fast enough to resolve
each glitch. You might sample at 1000 samples per second, and this requires a
frequency span of 1000/2.56, or about 390 Hz.
A good rule of thumb to memorize is that the time record length depends only on the
line spacing of the FFT spectrum and the sampling rate depends only on the
frequency range of the FFT spectrum, and they are independently adjustable. We will
return to this subject of time resolution versus frequency resolution soon, when we
look into Synchronous Averaging.
Acceleration vs. Velocity

When collecting a waveform, it is important to note that we generally want to see
high-frequency information. When integrating a signal from acceleration to velocity,
the high-frequency response is greatly reduced in level, and many small irregularities
in the waveform are not visible. Therefore, it is a generally valid rule of thumb that
we should use vibration acceleration rather than velocity as the measurement
parameter.
However there are cases, especially in low-speed machines, when low-frequency
information is interesting to us, and then integration to velocity, or even to
displacement, can be used. Of course, the integration must be done to the input
signal from the accelerometer. It is not possible to digitally integrate the acceleration
waveform.


78

Phase in the Time Domain

The concept of phase is interesting in that it can be considered a property of the time
domain waveform or a property of the frequency domain spectrum. Phase is nothing
more than the time difference between a reference time and a measured time. It
usually is represented in units of milliseconds in the waveform, and as degrees of
angle in the spectrum. One complete rotation corresponds to 160 degrees of phase
angle.
As we will see later in this book when we look at diagnostics of machine problems,
we will see that there are many different areas where phase gives us important
diagnostic information, balancing being one of the most important.
Even though phase really is defined in the time domain, it is usually much easier to
measure and display in the frequency domain. For this reason, we will not address it
further here.
The Wave Form as an analytical tool

More:

Orbit Plots
Impacts vs. Random Noise
Random of Non-Periodic Impacts
Truncation or flattening of the signal
Low-frequency events
Crest Factor
Orbit Plots

Orbit plots are plots that appear on an oscilloscope screen when two proximity
probes are connected to the horizontal and vertical inputs, respectively. The
proximity probes are installed in a sleeve bearing, one oriented vertically and one
oriented horizontally. Under these conditions, vertical motion of the shaft center line
will move the oscilloscope dot vertically, and horizontal motion of the shaft will move
the dot horizontally. Thus, when the shaft is turning, the dot will be stationary if the
center of the shaft is stationary in the bearing. If the shaft is migrating around in the
bearing, the dot will follow, and it will trace out the motion of the shaft center in real
time. The instrumentation setup is shown schematically below:

Under normal operating conditions, the scope will show a circle, indicating the oil film
has equal stiffness and thickness in all directions, and there is some imbalance
causing the centerline to move in a circle.
79

The orbit at the left above shows an ideal condition of the shaft in the bearing, while
the one on the right shows that the shaft is moving more in the vertical direction
than it is in the horizontal direction. This may mean the bearing is worn in an oval
pattern, with more vertical clearance than horizontal clearance.

An orbit pattern like the one shown above indicates shaft motion that moves in a
figure 8 pattern. It is vibrating twice as fast in one direction than in the other one.
This can be caused by excessive clearance in one direction, or a bearing worn into an
oval shape. A pattern like this is a danger sign, for the journal is likely to develop
metal-to-metal contact with the bearing, causing extensive damage to both.
Impacts vs. Random Noise

Impacts may be caused by rolling element bearings where the rollers encounter a
crack or small spall in a race. If there is a lot of external noise present, the spectrum
may not show a well-defined peak at the bearing tone frequency. However, the
acceleration waveform will usually exhibit repetitive peaks with their repetition rate
equal to the period of the bearing ball pass frequency.

A loose machine component that hits something at a rate that is unrelated to
machine speed will generate impacts that are usually not precisely periodic, and can
be essentially random in their spacing. In the vibration spectrum of such a condition,
the impacts will produce a continuous noise spectrum that can extend to very high
frequencies. The spectrum could be confusing in that other sources or wide band
noise will look almost the same. In the waveform, the impacts are clearly seen.
Truncation or flattening of the signal

In many cases of looseness, such as a bearing pillow block that rises up a little
during part of the rotation and then contacts the base for the rest of the cycle, the
waveform will be flattened on one side. This will result in harmonics in the spectrum,
but other types of waveform distortion will also produce harmonics. The waveform
gives a quick identification of this kind of looseness where motion is restricted in one
direction.

80
Low-frequency events

In some cases the vibration signal might have a discontinuity once in a while that
when transformed into the frequency domain will be so low in frequency it will not be
clearly seen in the spectrum. An example of this is a low-speed gearbox which has
one broken or cracked tooth on the large gear. A waveform recorded over a long
time will show the discontinuities.
Crest Factor

The Crest factor, also sometimes called the "peak-to-RMS-ratio", is defined as the
ratio of the peak value of a waveform to its RMS value. It is a pure number, without
units. The crest factor of a sine wave is 2, or1.414; i.e. the peak value is 1.414
times the RMS value. A typical vibration signal from a machine with a large
imbalance and no other problems will have a crest factor of about 1.5, but as the
bearings begin to wear, and impacting begins to happen, the crest factor will become
much greater than this. The reason that the crest factor is so sensitive to the
existence of sharp peaks in the waveform is that the peaks do not last very long in
time, and therefore do not contain very much energy. The RMS value is proportional
to the amount of energy in the vibration signal.
Examples of actual crest factors measured on a cam rider on a large low-speed
machine are shown below. The cam rider contains a roller bearing. The vibration
signature, which is scaled in velocity units (ips), shows a little irregularity that is no
doubt due to some roughness in the bearing. There is essentially no low-frequency
motion here because the accelerometer is oriented in the radial direction of the
bearing shaft, which is at 90 degrees from the motion of the cam rider.
Note the RMS value is 0.017 ips, and the crest factor is 8.9. In other words, the peak
value is 8.9 times the RMS value.

Radial Velocity Waveform Without Fault
The next figure, below, shows the same measurement point, at a later time. Here,
the waveform shows that the bearing has an obvious fault in one of the races. The
RMS level remains low at 0.086 ips, but the crest factor has risen to 19. This shows
that a simple vibration meter that is only sensitive to RMS level is not able to detect
a defective bearing, at least in this case.

81


The next figure, below, was measured at the same point on the machine as the
previous waveforms, but in the direction of motion of the cam rider rather than
perpendicular to it. The measurement was made before the bearing developed the
fault. Here, we see the low-frequency content due to the movement of the cam. The
small noise bursts are caused by the minor bearing damage as was shown in the first
figure, above.
Note the RMS value is 0.45 ips and the crest factor is 1.7.

Tangential Waveform Without Fault
The figure below is from the same measurement point and direction as the one
above, except it was recorded after the bearing fault developed, as in the second
figure, above.
82

Tangential Velocity Waveform With Fault
Note that the bearing fault is clearly visible in the sharp spikes similar to the ones in
the Radial with fault figure above. Note also that the RMS value is 0.45 ips, the same
as in the previous figure, indicating that the bearing fault did not add significantly
more energy to the vibration signature. The interesting fact here is that the crest
factor of 1.8 is only slightly higher than before, even though the sharp vibration
spikes are present. In this case, the large low-frequency signal masks the spikes and
they do not show up as an elevated crest factor. This is a good illustration that the
crest factor alone can sometimes be misinterpreted unless the vibration waveform is
actually observed. This condition does no occur very often in practice, however.
Synchronous Averaging

Synchronous averaging, also sometimes redundantly called "Time Synchronous
averaging, was discussed earlier as a method of background noise reduction in
spectra of complex signals. Now, we will look at it as a means of greatly increasing
the information we can extract from the time-domain vibration waveform.
Synchronous averaging is a fundamentally different process than the usual spectrum
averaging that is generally done in FFT analysis. It is used to greatly reduce the
effects of unwanted noise in the measurement. The waveform itself is averaged in a
time buffer before the FFT is calculated, and the sampling of the signal is initiated by
a trigger pulse input to the analyzer. If the trigger pulse is synchronized with the
repetition rate of the signal in question, the averaging process will gradually
eliminate the random noise because it is not synchronized with the trigger. However,
the signal that is synchronous with the trigger will be emphasized, as shown below:

83


When you do time domain averaging on the vibration signal from a real machine, the
averaged time record gradually accumulates those portions of the signal that are
synchronized with the trigger, and other parts of the signal, such as noise and any
other components such as other rotating parts of the machine, etc., are effectively
averaged out. This is the only type of averaging that actually does reduce noise.

An important application of time synchronous averaging is in the waveform analysis
of machine vibration, especially in the case of gear drives. In this case, the trigger is
derived from a tachometer that provides one pulse per revolution of a gear in a
machine. This way, the time samples are synchronized in that they all begin at the
same exact point in the angular position of the gear.
Consider a gearbox containing a pinion with 13 teeth and a driven gear with 31 teeth.
If a tachometer is connected to the pinion shaft, and its output is used to trigger an
analyzer capable of time synchronous averaging, the averaged waveform will
gradually exclude vibration components from everything except the events related to
the pinion revolution. Any vibration caused by the driven gear will be averaged out,
and the resulting waveform will show the vibration caused by each individual tooth
on the pinion.

Note that in the figure above, the lower averaged waveform indicates one damaged
tooth on the pinion.
Analyzer Set-Up for Synchronous Averaging

Here, the FFT analyzer is used not as a frequency analyzer, but as a time-domain
averager. Many, but not all, FFT analyzers have this capability. We will see that the
setup parameters for synchronous averaging are generally quite different from those
needed for spectrum analysis.
84
There are two important considerations in setting up a synchronous averaging test,
the Time Record Length, and the Time Domain Resolution. The length of the time
record is usually set up to be at least a little longer than one revolution of the of the
shaft you are looking at, or one cycle of the event you are studying. There is usually
no reason to use a longer time record, since you simply lose resolution in the time
domain. The fixed number of samples is spread over a longer time; so short events
during the motion you are analyzing could be lost.
More:

Time Record Length
Time Domain Resolution
Time Record Length

Time record length in an FFT analyzer, as was discussed in the frequency analysis
chapter, is the reciprocal of the frequency resolution of the spectrum. In other words,
a spectrum with frequency resolution (or line spacing) of 10 Hz is generated from a
time sample lasting 1/10
th
second. So, suppose in the example in the figure above
the pinion is rotating at 1800 RPM and we want to look at one revolution. 1800 RPM
is 30 Hz (30 revolutions per second), so one revolution will take 1/30
th
of a second.
To set up an FFT analyzer to collect a 1/30
th
second time record requires that its
frequency resolution, or line spacing, must be 30 Hz. In order to acquire a little more
than one revolution, you might use 1/25
th
second as the time record length. There
are many theoretical combinations of frequency range and resolution that could be
used to do this - i.e., a span of 5000 Hz with 200 lines resolution, a span of 2500 Hz
and 100 lines of resolution, or a span of 250 Hz at 10 lines resolution.
Time Domain Resolution

The resolution in the time domain depends on the sampling rate of the D/A converter
in the analyzer - the higher the sampling rate, the greater the resolution. The FFT
requires that the sampling rate be 2.56 times the maximum frequency being
analyzed. Therefore, a span of 5000 Hz will set the A/D converter to 5000 x 2.56 =
12,800 samples per second. If our time record length is 1/25 second, then the time
resolution is 1/25
th
of 12,800, or 512 samples in the time record. This means we will
get 512 "snapshots of the gear in one revolution. Since there are 13 teeth on the
pinion in question, our 512 samples per revolution will produce for us 512/13 or
about 39 snapshots of each tooth. This is more than adequate to examine each tooth
in the averaged waveform.
On the other hand, had we used 250 Hz at 10 lines of resolution, we would have only
250 x 2.5 = 640 samples per second, and 1/25 second will only contain about 25
samples, or about 2 samples per tooth of the gear, which would probably not be
adequate to examine the gear in enough detail.
This same basic setup can be used to examine the driven gear rather than the pinion.
If it is not possible to put a tachometer on the shaft of the large gear, a frequency
divider can be placed in the tachometer output, and if its division ratio is 13:31, its
output pulses will coincide with each revolution of the large gear. Then, the time-
averaged waveform will show the teeth of the driven gear rather than the pinion. The
setup is shown in the figure below:

85


Since the driven gear is running slower than the pinion, the analyzer must be set up
to acquire a time record length corresponding to a little more than 1 revolution of the
large gear.
If the averaged waveform were also subjected to a frequency analysis with the
analyzer, the predominant frequency in the spectrum would be 13X, which is the
gear mesh frequency, as in the previous case with the pinion.
The technique can be applied to gearboxes with multiple gears as long as the
numbers of teeth on each gear are known, and the frequency divider has the
capability of performing the proper ratio multiplication.
Case Histories using Synchronous Averaging

The following data are from actual industrial machines whose vibration waveforms
were analyzed by synchronous averaging as well as conventional vibration spectrum
analysis. In these examples, the averaged waveform presented a much clearer and
unequivocal picture of the nature of the machine faults than the vibration spectrum
did. However, it must be emphasized that synchronous averaging is not suitable for
use in finding modulation effects such as sidebands and other conditions where high
resolution in the frequency domain are required.
More:

Low-speed Radar Gearbox
Tabletting Press Example
Low-Speed Corrugation Press Gearbox
Low-speed Radar Gearbox

This example is the acceleration spectrum of a large low speed (4 RPM) gear driving
a long-range radar antenna. The frequency range is from 0Hz to 6Hz. Note that the
spectrum is noisy, and it is difficult to tell it contains anything else but noise.
86

The next graph shows the synchronous-averaged waveform from which this
spectrum came :

Note that the time record length is 90 seconds.
There is a sharp peak every 15 seconds. Since the gear speed is 4 RPM, 15 seconds
is the time it takes to make one revolution. The peaks are the result of one defective
tooth in the gear. Each time the tooth contacts the pinion gear an impulse is
generated. The information is immediately assimilated form the waveform, while the
spectrum is nowhere near as informative.
Tabletting Press Example

Following is a simplified diagram of the drive mechanism for a large tabletting press.
The speed of the large ring gear is 56 RPM. This machine was moved from one
building to another, and encountered an accident while en route. It was decided to
perform a vibration analysis on it to find any hidden damage.

The following vibration spectra were measured eight days apart. The lower spectrum
was the first to be collected. Note the gear mesh frequency at 3640 RPM, with the
two prominent sidebands spaced at the pinion RPM from the GMF. This indicated
87
modulation of the ring gear speed at the pinion rate, and is probably caused by a
bent pinion shaft or uneven pinion tooth wear.

It was then decided to collect a synchronous averaged waveform with a time span of
about 3 ring gear revolutions. The waveform from the earlier measurement is shown
below:

This is not remarkable in that it does not show any significant repetitive pattern.
Now, note the upper spectrum in the previous figure, made 8 days later. Here we
see sidebands around the GMF at the ring gear turning speed. This indicates the ring
gear vibration level is fluctuating once per revolution. This could indicate a localized
fault in the ring gear that causes momentary level changes. Again, a synchronous
averaged waveform was collected, and is shown below. The time duration is about 3
seconds, so the plot encompasses 3 revolutions of the ring gear. Note the very
definite repeating pattern of a high vibration level. The conclusion was that the ring
gear was damaged by the accident, probably by impacting the pinion, and the
damage increased rapidly as the machine ran.



88

Low-Speed Corrugation Press Gearbox

This next example is from a heavy-duty corrugation press in a paper plant.
The gearbox in question had an input shaft with a 23-tooth pinion driving a large bull
gear with 132 teeth, which in turn drives another bull gear of the same size. The two
gears are connected to large steel rollers about 24 inches in diameter. The two
output gears turn at 52 RPM and the pinion turns at 302 RPM.

The lubricating oil in the gearbox was regularly subjected to analysis, and the last
report stated that there were iron particles in the oil. The maintenance supervisor
asked if we could determine the source of the oil contamination, and the first thing
we did was to examine the vibration spectra measured near the bearings. The
spectra looked normal, without evidence of bearing tones, so we suspected that the
metal was coming from one or more of the gears. See the vibration spectrum in the
figure below:

The problem then was to identify the faulty gear (if any), so as to allow the
maintenance effort to proceed without delay.
We decided to perform synchronous averaging of the gearbox vibration signature.
When we did synchronous averaging on each of the two bull gears, there was no
indication of any defect in the waveform. But, when we performed the same test on
the pinion, the waveform told another story.
The waveform shows a little over two revolutions of the pinion. There is an obvious
area on the gear where the meshing with the bull gear was very noisy and non-
uniform. We called for an inspection of the gear, and an access plate was removed
so we could look at the gears. We found that the keyway in the pinion shaft was
badly worn such that the gear could be rotated back and forth on the shaft by about
1/2 tooth at the edge of the gear. There was also visible clearance between the shaft
and the bore of the pinion. The bull gears showed no sign of damage. The averaged
waveform is shown below:
89

The figure below is a photograph of the pinion showing the spalling in the pinion bore
caused by the gear turning back and forth on the shaft:


We called one of the engineers at the gearbox factory and described the situation.
He said the problem occurred during installation when the interference fit between
the pinion and the shaft was too loose. He said the shaft and pinion would have to be
replaced, and very soon, to avoid a catastrophic failure.


90

Cepstrum Analysis

More:

Cepstrum Terminology
Cepstrum Terminology

The term "cepstrum" is simply "spectrum" with the first four letters in reverse order.
The various parameters of the cepstrum have been given somewhat whimsical
names, summarized below:
Spectrum Cepstrum
Frequency

Quefrency

Harmonic

Rahmonic

Magnitude

Gamnitude

Phase

Saphe

Filter

Lifter

High-Pass

Short-Pass

Low-Pass

Long-Pass

Fundamental

Mundafental

Quefrency is the horizontal axis of the cepstrum, and has the units of periodic time.
Rahmonics are cepstral components that are spaced at equal increments of time.
Following are the spectrum and cepstrum plots of the vibration signature of a belt-
driven machine. In the spectrum, the harmonic cursor is set to 8.35 Hz, which is the
fundamental belt frequency. Note that many harmonics are highlighted. The second
figure is the cepstrum of this spectrum. The cursor is set at 0.119 seconds, which is
the period of the 8.35 Hz component in the spectrum. Note how much simpler the
cepstrum is compared to the spectrum. The peaks at 0.119 sec and rahmonics of
0.119 sec indicate a strong periodicity in the spectrum; i.e. a distinct harmonic series.
If this machine were to be monitored in a predictive maintenance program, its
spectra would be collected and trended over time. To determine the belt condition,
the levels of the various harmonics of belt frequency would need to be noted, for the
level of the fundamental is not a good indicator of belt condition.
If the cepstrum were used for the same type of trending, only the component at
0.119 sec would need to be considered, since its level is dependent on the levels of
all the harmonics of 8.35 Hz in the spectrum.
91
Spectrum from a
belt-driven machine

Cepstrum from a
belt-driven machine
Statistical Properties of Vibration Signals

More:

Amplitude Probability Distribution
Kurtosis
Amplitude Probability Distribution

The vibration signature of a machine always has some random variation, i.e., its
instantaneous value is not predictable. Nevertheless, the probability of a given value
falling within a certain amplitude range is predictable in a statistical sense. For
example, consider a short sample of the vibration velocity signal from an operating
machine. The vibration velocity V at any instant can vary in a random manner about
some mean value. Suppose the velocity scale is divided into a series of small
divisions DV. Then, the statistical probability that the signal will be in any given
division can be measured by noting the time the signal spends in each division
divided by the total time the signal is monitored. The probability density is a
measure of the distance away from the mean value the amplitude will be, plotted
against the amplitude.
The most familiar probability density curve is the famous "normal", or Gaussian,
distribution, also popularly known as the "bell-shaped curve".
92
The RMS value of a signal with a Gaussian distribution is equal to the Standard
Deviation of the signal, and is abbreviated with the Greek letter s (sigma). A random
vibration signal will produce a Gaussian distribution, and experience shows that
healthy machines also produce Gaussian distributions. As faults develop in machines,
the amplitude distribution curve changes shape; for instance, a small bearing fault
will introduce "spikes" in the vibration wave form, and this will increase the level of
the "tails" of the distribution curve, as shown below. The U.S. Navy has studied the
use of the amplitude distribution in machine monitoring for some time, but it has not
been generally adopted by industry.




Kurtosis

One mathematical representation of the deviation of an amplitude distribution from
Gaussian is the so-called "fourth moment", or kurtosis. The Gaussian distribution has
a kurtosis of 3, and higher values of kurtosis indicate increased crest factor of the
vibration signal. Kurtosis is a valid measure of the degradation of a machine, but it
does not give any indication of the diagnosis of the problem. It has been reported
that kurtosis is especially well suited to monitoring of reciprocating machines for fault
detection.
One possible advantage of using kurtosis as a fault detection parameter is that it
does not need to be trended over time to be effective. A kurtosis of 3 generally is
taken as indicating a healthy machine, with higher values indicating progressive
states of fault progression.
Amplitude Demodulation

More:

What is Amplitude Modulation?
Beats
Amplitude Modulation in Machine Vibration Signatures
Amplitude Demodulation Applied to Bearing Analysis
Evaluating Demodulated Bearing Spectra
Appearance of Demodulated Spectra
Demodulation Case Histories
What is Amplitude Modulation?

Amplitude modulation is defined as the multiplication of one time-domain signal by
another time-domain signal. The signals may or may not be complex in nature, i.e.,
either or both signals may contain harmonics components. It is impossible to have
93
amplitude modulation unless at least two different signals are involved. The signals
may be electrical in nature, or they can be vibration signals. Modulation is inherently
a non-linear process, and always gives rise to frequency components that did not
exist in either of the two original signals



Amplitude Modulated Wave Form
If the amplitude-modulated signal shown here is passed through a frequency
analyzer, the following spectrum is the result. The highest peak is the carrier
frequency. The right-hand peak is the "upper sideband, and has a frequency of the
carrier frequency plus the modulating frequency. The left-hand peak or "lower
sideband has a frequency of the carrier minus the modulating frequency. The
sidebands are sometimes called "sum and difference frequencies because of their
symmetrical spacing around the carrier.
Amplitude modulation also occurs in sound reproducing equipment, where it is called
Intermodulation Distortion. The sum and difference frequencies are not in musical
harmony with the tones that cause them, making intermodulation a particularly
noticeable form of sound distortion.


Spectrum of Modulated Wave Form

Rectified Wave Form

Recovered Modulating Signal

94
This process of demodulation is exactly what happens in an AM radio -- the carrier is
a very high frequency signal generated by the radio station, and the modulating
signal is the voice or music that constitutes the program. The radio receives the
modulated carrier, amplifies it, and rectifies ("detects) it to recover the program.
Beats

If two sounds, vibrations, or electrical signals have nearly the same frequency and
they are linearly added together, their combined amplitude will fluctuate up and
down at a rate equal to the difference frequency between them. This phenomenon is
called "beating, and is very commonly seen in practice. For instance, a musician
tunes his instrument by listening for beats between two tones that are nearly the
same pitch.
A beating waveform looks very much like amplitude modulation, but it is actually
completely different. A spectrum analysis of beats produces only the two frequency
components that are combined -- there are no new frequencies such as sidebands
present. It is easy to confuse beats with amplitude modulation, but a spectrum
analysis will show the difference. In general, beats are benign, and do not imply
faults in machines. For example, the sound of two similar machines running side by
side at slightly different speeds will often produce audible beats. This is simply the
sounds made by the machines combining in air to produce the amplitude fluctuations.
Amplitude Modulation in Machine Vibration
Signatures

Many machines produce vibration signatures which contain amplitude modulation,
and as we have seen in the introduction to vibration section, amplitude modulation
causes sidebands to occur in the vibration spectrum. Several types of machine
problems can be diagnosed by detailed examination of these sidebands. Examples of
machines that produce amplitude modulation are gearboxes, where the tooth mesh
frequency is modulated by the turn speed of each gear, and rolling element bearings,
where bearing tones can be modulated by turn speed or the fundamental train
frequency of the bearing.
In the case of gearboxes, an eccentric gear or bent shaft will cause the tooth mesh
tone to be stronger during the portion of the revolution of the gear where the radius
is increasing -- the driven gear is actually being accelerated in its rotation during this
time. The part of the rotation where the radius is decreasing places less force on the
gear teeth, and the tooth mesh tone is less strong. (The tooth mesh tone is also
frequency modulated at the same time, and this also creates sidebands in the
spectrum, but for this discussion, we will only consider amplitude modulation). Any
other defect in the gear, such as a cracked or spalled tooth, will also cause the tooth
mesh tone to be irregular, and will result in modulation of the tone and consequent
sidebands in the spectrum.
Since the gears in the gearbox usually rotate at different speeds, the amplitude
modulation due to each gear will be at a different rate, and the resulting sidebands
will be of different spacing. This allows the diagnosis of gearbox faults narrowed
down to specific gears and/or shafts by analyzing the sideband patterns in the
vibration spectra.
In bearings, modulation of the bearing tones occurs in several ways. If the inner race
has a small defect such as a crack, this defect will move in and out of the bearing
load zone at the rate of the shaft rpm. This assumes the inner race is rotating and
95
that it is in a horizontal machine where gravity imparts a radial, rather than axial,
force on the bearing. The bearing tone will be strongest when the defect is in the
load zone, and weakest when it is out of the load zone. This means the inner race
ball pass frequency will be amplitude modulated, and its spectrum will have
sidebands spaced apart by the rpm of the race. In contrast to this, a fault in the
outer race, which is stationary, will always be in the load zone of the bearing, and no
modulation will occur, and no sidebands around the outer race frequency will be
produced.
If a rolling element has a defect, this roller will move in and out of the load zone also,
but will do so at the fundamental train frequency (FTF) rather than the rpm. This is
because the rollers are migrating around in the bearing at the cage rpm. This
condition will produce amplitude modulation of the ball spin frequency, and spectral
sidebands will be spaced apart at the FTF.
Demodulation Applied to Bearing Analysis

If a rolling element bearing has a defect such as a crack in one of the races, there
will be an impact created every time a rolling element passes over it. These impulses
cause the bearing race to "ring at its natural frequency, just as a bell will ring when
tapped. The race is not free to vibrate very much because it is held in the bearing
housing, so its ringing is very highly damped. This gives rise to a series of very,
short "pings which occur at the rate of the ball passing, as illustrated below:
Evaluating Demodulated Bearing Spectra

When looking at demodulated spectra, it is important to keep in mind that they are
not the same as normal vibration spectra. Spectral components in the demodulated
spectrum of a bearing at the bearing forcing frequencies do not represent actual
vibration at these frequencies. This is because the high-pass filter has filtered all the
energy at these forcing frequencies out of the signal before the demodulation was
performed. The spectrum of the demodulated vibration signal indicates the influence
of the bearing faults on a high-frequency band of vibration that is not related to the
forcing frequencies. Even though the vibration sensor is an accelerometer, the
demodulated spectrum should not be scaled in acceleration units. This has led to
confusion in the industry as to what is the proper amplitude unit to use in displaying
demodulated spectra. It is the opinion of the author that a simple scaling in voltage
decibels without reference to any physical vibration parameters is best. DLI
Engineering Corp. has selected decibels related to 1 millivolt as the default scaling
for demodulated spectra in the DC-7B data collector. This is abbreviated dBmV. The
one-millivolt reference is not particularly significant, but it assures that all dB values
likely to be encountered in practice will be positive, and the numbers will be in
ranges that are familiar to persons used to working in dB Velocity.
The use of the decibel, which is a logarithmic ratio rather than a unit, is appropriate
because the demodulated spectra are not evaluated in terms of absolute levels, but
rather as signal to noise ratios, as will be described in the next section.








96
ThThe resulting waveform is actually an example of amplitude modulation, with the
ringing frequency corresponding to the carrier and the envelope of the pings is the
modulating signal. If the signal is passed through a spectrum analyzer, there will be
almost no energy at the ball pass frequency in the spectrum, but there will be a
component at the natural "ringing frequency and there will be sidebands around it
spaced apart by the ball pass frequency. In practice, it is usually very difficult to see
these sidebands in a machine spectrum, most due to extraneous noise components
that mask them. Also, the ringing frequency is usually quite high, sometimes over 10
kHz, and these frequencies are difficult to pick up reliably.


If the signal is demodulated by being rectified, the result is a series of impulses
spaced apart by the ball pass frequency, as shown below. The rectification turns the
negative-going parts if the waveform to become positive:


Rectified Ping Signature
The signal is then passed through a low-pass filter to remove the fluctuations due to
the ringing frequency, and smoothing the pulses. These pulses, when subjected to
frequency analysis will produce a strong component at the ball pass frequency, along
with harmonics of it. This is because they have more area under them, contributing
more energy at their fundamental repetition rate that is the outer race ball pass
frequency of the bearing.

Smoothed Rectified Bearing Pings
In order to separate the high frequency ringing of the bearing race from the rest of
the vibration signature of the machine, the accelerometer signal is passed through a
high-pass filter tuned to about 2.5 kHz. This filter removes all the lower frequency
components due to rotation rates and their harmonics, and effectively isolates the
modulated natural frequencies. This results in a very great increase in signal to noise
ratio, and is one of the main reasons for the sensitivity of demodulation in detecting
small bearing defects. This is the most important benefit of amplitude demodulation
as a machine diagnostic tool. The block diagram of an effective amplitude
demodulation scheme is shown below.

Appearance of Demodulated Spectra

The typical demodulated spectrum from an accelerometer connected to a rolling
element bearing will usually have a fairly uniform and level "noise floor with discrete
peaks rising above it, as shown in the following figure. If the loading on the machine
increases, the entire noise floor and peaks will rise in level, but, and this is the
crucial part, the relative height of the peaks above the noise floor will remain almost
97
exactly the same. This means the loading of the machine is not nearly as important
as it is in measuring vibration directly, and the demodulated spectra are more
consistent in their appearance.
The noise floor in demodulated spectra is generally quite smooth and uniform in level,
in contrast to the random nature of the noise in conventional spectra. This is because
the initial high-pass filter filters out almost all the random noise in the vibration
signal.







Typical Appearance of a Demodulated Acceleration Spectrum
The following hypothetical spectra represent the progression of damage in a rolling
element bearing. It must be borne in mind that this is meant to be a guide only, and
must not be taken as an absolute standard that applies to all machines. The data
presented here are a distillation of the analysis and verification of many hundreds of
demodulated machine spectra collected over a period of about 10 years on a variety
of industrial machines. However, there is no substitute for knowing the
characteristics of your particular machine and especially trending the rate of increase
in bearing tones in demodulated spectra.


Stage 1
The figure shown above is a conventional vibration spectrum in velocity dB alongside
a demodulated spectrum of the same measurement scaled in dB volts. The VdB
spectrum shows a few run speed harmonics and a normal noise floor at a low level.
The demodulated spectrum shows a smooth noise floor at an arbitrary level we can
use as a reference.

98
Stage 2
The next figure above shows the first stage of bearing degradation due to a tiny flaw
in the inner race. The conventional spectrum shows very little if any bearing tones
and the same residual looseness indicated by the harmonics of run speed. The
demodulated spectrum, however, shows the bearing tone at 2 - 3 dB above the
reference noise floor, and also some run speed harmonics. Run speed harmonics in
the demodulated spectrum indicate a small increase in looseness due to bearing
clearance increasing. They may or may not be apparent at this stage.
At this stage, the bearing need not be replaced, but its condition should be
monitored closely.

Stage 3
The next stage of degradation is shown above. The conventional spectrum still does
not show any bearing tones at significant levels. The demodulated spectrum has
bearing tones at 5 to 10 dB above the reference noise floor. The bearing is in poor
condition, but still may have significant service life left.


Stage 4
Here, the bearing has degraded to unacceptable condition. Bearing tones appear in
the velocity spectrum, and also appear with run speed sidebands in the demodulated
spectrum. Note that the entire demodulated spectrum has risen in level about 10 dB,
and the bearing tones are 10 dB or more above the floor.


Stage 5
99
In this stage, the bearing needs to be replaced immediately. Bearing tones with 1X
sidebands appear in both spectra, along with run speed harmonics in the
demodulated spectrum. Note that the noise floor in the demodulated spectrum has
risen by nearly another 10 dB.


Stage 6
In the figure above, the spectra indicate total failure of the bearing is imminent.
Bearing tones are missing in both spectra because the fault has become distributed
over the race, rather than being localized. The increased harmonic content in the
conventional spectrum is due to increased clearance between the balls and the races.
Demodulation Case Histories

To illustrate the effectiveness of amplitude demodulation, several actual case
histories are presented below. In some cases, conventional vibration velocity spectra
are compared to the demodulated spectra. The relatively noise free demodulated
spectra are typical of those found in industrial machines.
More:

Motor-driven Centrifugal pump
Outdoor Conveyer Belt system
Crane Gearbox
Motor-driven Centrifugal pump

The first case history is interesting in that it shows that demodulation is very
powerful in improving diagnostic accuracy compared to conventional vibration
analysis. The machine is a direct-coupled motor pump combination. Conventional
vibration spectral analysis over a frequency range of 20 orders of run speed detected
a bad bearing, but the spectrum alone could lead one to the wrong bearing in the
machine.
Here is the vibration spectrum measured at the coupled end of the motor. Note the
non-synchronous bearing tone at just over 100 Hz and 92 VdB (0.022 ips) level.
There is also a second harmonic of this component visible. This strongly indicates a
defective rolling element bearing.









100










The next spectrum was measured at the free end of the pump, and is shown below:
Note that the same bearing tone and second harmonic are present here, but the
level of the tone is 12 dB (4 times) higher in level than it was on the motor coupled
end. This is an indication that the bad bearing is in the pump and not the motor. The
fact that the bearing tone appears in both locations is because this is a small
machine, and the bearing vibration travels across the machine.

















The next pair of spectra tells a different story. Following is an amplitude-
demodulated spectrum taken at the pump free end. The frequency span of the
demodulated spectrum is 20 orders. Note that there are no bearing tones visible.

Next is the demodulated spectrum from the motor-coupled end, shown below: Here
we have a series of bearing tone harmonics that rise above the noise floor more than
101
15 dB! This is a sure-fire indication that the problem bearing is in the motor and not
the pump. This is a good indication of the power of demodulation to localize vibration
sources, especially in the case of rolling element bearings.


the higher-level bearing tone in the conventional vibration spectrum of the pump
free end? If we look at the noise floor near the bearing-tone frequency in the
conventional spectrum from the pump free end, we see a hump, or "haystack that is
102
no doubt caused by a mechanical resonance. This resonance amplifies the noise floor
by about 15 dB. The bearing tone is also amplified by the resonance, accounting for
its elevated amplitude.
An important aspect of the demodulation process is that the signal actually being
demodulated is very high in frequency since it has been passed through a high-pass
filter, usually above 2.5 kHz or so. This high-frequency energy does not travel very
well through mechanical structures, so the information in the demodulated spectrum
comes from very close to the accelerometer. This is the reason the demodulated
spectrum from the motor-coupled end is not affected by the mechanical resonance
that amplifies the bearing tone in the base band spectrum.
Outdoor Conveyer Belt system

The following example is from an outdoor heavy-duty conveyor belt used to
transport earth. The belt runs uphill and is about mile long and is four feet wide.
The drive is through a large gearbox at the upper turn-around roller turning at about
70 RPM. The data shown in the example was measured at the far end of the belt at
the lower turn-around roller bearing. The bearings in these rollers are spherical roller
bearings with two sets of rollers each. These bearings will tolerate a significant
amount of angular misalignment due to the spherical shape of the outer race. They
will not, however, tolerate very much thrust loading.
The conventional vibration spectrum, shown below, was measured in the radial
direction on the bearing housing:

The cursor is set at the turn speed of 71 RPM, with harmonics cursors activated. The
frequency scaling is in orders of run speed. The large peaks at about 12X with a
harmonic and at14X with 2 harmonics could be due to bearings in the small support
rollers under the belt, and not from the bearing being monitored. The spectrum is
complex, with a high noise floor due to the several hundred other rollers that support
the belt. It is difficult to make a definite diagnosis from spectrum since the machine
is so complex.
103
The following figure is the demodulated spectrum measured from the same location.
Its frequency span is 35 orders of the roller turn speed.

Note here strong peak at 14.9 orders with 1X sidebands around it. The noise floor is
quite flat and uniform compared to the conventional spectrum. The small flag
indicates the turn speed component at 1 order.
14.9X is recognized as a bearing tone since it is not synchronous with the roller turn
speed and has 1X sidebands around it. The sidebands indicate that it represents an
inner race fault in the bearing. This is because the fault goes in and out of the
bearing load zone at once per revolution, modulating the amplitude of the bearing
tones at the turn speed.
The bearing was replaced, and the next figure is a photograph of the inner race of
the old bearing:

It is seen that the fault is localized and only occurs on one side of the race. This
indicates that the bearing experienced thrust loading, relieving the force on the other
half of the race.
The following figure shows the outer race of the same bearing:
104

The outer race damage is localized in the load zone of the bearing.
This is a classic case of the demodulated spectrum showing an inner race fault and
also eliminating almost all of the contaminating noise that was so evident in the
vibration base band spectrum.
The defective bearing was replaced, and the following figure shows the conventional
vibration spectrum form the same measurement location:

Note that this spectrum resembles the first conventional spectrum, shown earlier.
There are still plenty of run speed harmonics and noise components. This illustrates
that the bearing fault did not present a very effective signature in the original
spectrum. This, of course, is due to the excessive complexity and noisiness of the
spectrum.
Now, look at the demodulated spectrum, taken at the same time and location as the
previous conventional spectrum, shown next:
105

In this spectrum, we find no bearing tones, and the smooth, uniform noise floor that
is characteristic of a classic demodulated spectrum. This illustrates the effectiveness
of the demodulation process in zeroing in on localized faults and eliminating noise
components that come from more distant parts of the machine.
Crane Gearbox

The following example is from a hoist gearbox input shaft cylindrical roller bearing.
The first spectrum is a conventional vibration spectrum measured in the radial
direction on the bearing housing. The frequency scaling is in orders of shaft speed of
the bearing, and the markers on the spectral peaks indicate the harmonics of this
speed :
106

This is a very complex, noisy spectrum, and the numerous strong harmonics indicate
extreme looseness in the machine. The dotted cursors are set to 6.31 orders of run
speed, and they correspond to an outer race fault frequency of the bearing. But, with
this much looseness shown in the spectrum, it is difficult to say how badly the
bearing is damaged, since looseness elsewhere in the machine could produce such a
noisy spectrum.
The figure below shows the demodulated spectrum taken from the same
measurement point:
107

This spectrum shows harmonics of the outer race bearing fault frequency rising
about 25 dB above the noise floor. It is an excellent example of a severely damaged
bearing spectrum. Note that the first harmonic is quite well defined in frequency, and
the higher harmonics are progressively "smeared" in frequency. This indicates there
is some jitter in the frequency of the fault tone, probably caused by excessive
distributed damage to the outer race.
The next figure, below, contains photographs of the outer and inner races of the
bearing: (The inner race picture is taken with more magnification than the outer race
picture.)

108

The excessive damage to the outer race is impressive, and it is surprising that the
inner race is damaged so little. The inner race is indented rather than spalled,
whereas the outer race is heavily spalled.
Machine Diagnosis

After the vibration signatures are verified as to validity and the spectral peaks,
especially the 1X components positively identified, can the diagnosis of machine
problems begin. The following section discusses a variety of machine problems and
illustrates them with their typical vibration signatures.
In analyzing vibration spectra from rotating machines, it is important to note that
individual faults are seldom seen by themselves. Care must be taken in the
interpretation of vibration signatures since different faults can cause spectral
components at the same frequencies.
More:

Imbalance
Misalignment
Journal Bearings
Rolling Element Bearings
Mechanical Looseness
Electrically Induced Vibration
Turbines
Pumps
Fans
Couplings
Drive Belts
Gearboxes
Centrifugal Compressors
Reciprocating Machines
Diagnostic Summary Charts
Imbalance

More:

Calculating the Imbalance Force
Couple Imbalance
109
Severity of Imbalance
Calculating the Imbalance Force

,
where F = the imbalance force, Im = the mass, r = its distance from the pivot, and
w (omega) is the angular frequency, equal to 2p times the frequency in Hz..

From this, it is seen that the force on the pivot is proportional to its distance from
the center of rotation and to the speed squared.
A rotor containing a heavy spot is not exactly equivalent to the stone on a string. In
the case of the stone, the center of gravity of the system is the center of the stone
itself, whereas the CG of a rotor with imbalance is outside the imbalance mass and is
near the axis of rotation of the rotor.

If the structure holding the bearings in such a system is infinitely rigid, the center of
rotation is constrained from moving, and the centripetal force resulting from the
imbalance mass can be found from the above formula. This force is borne by the
bearings. Now, consider a hypothetical machine where the bearings are not rigidly
supported, but are suspended on springs.

Under these conditions the shaft centerline is not constrained, and the rotor will
rotate around its center of gravity. The 1 x RPM force on the bearings will be very
small because it is only required to accelerate the bearings to the above mentioned
amplitude. The double amplitude of vibration of the bearings will be equal to twice
the distance between the CG and the centerline of the rotor. Moreover, the
amplitude of bearing vibration is constant regardless of the rotor speed, provided the
speed is higher than the natural frequency of the spring-rotor system. It is seen here
that the vibration amplitude has nothing to do with the above centripetal force
formula.
At speeds well below the natural frequency, the system is said to be "spring
controlled", and the centripetal force formula holds. Speeds above the natural
frequency are in the "mass-controlled" region where the amplitude is constant, and
the bearing forces are not so easily predictable, be dependent on the equivalent
mass of the bearings and springs.

110


Static Imbalance
Couple Imbalance



Couple Imbalance

With pure imbalance, either static or dynamic, the axial 1X and 2X vibration levels
will be low
Severity of Imbalance

The severity of imbalance depends on both the type and size of the machine as well
as the vibration level. To assess imbalance severity, average 1X levels for healthy
machines of the same type should be used as a comparison. If the second order
peak is as large as the first order, you should suspect misalignment.
The following levels are guidelines for general use in diagnosing imbalance for
machines running at 1800 or 3600 RPM. Very high-speed machines have lower
tolerance levels.
1X Vibration Level,
VdB
Diagnosis Repair Priority
Less than 108 VdB

(0.141 ips)

Slight Imbalance

No recommendation

108 VdB -- 114 VdB

(0.141 - 0.282 ips)

Moderate Imbalance

Desirable

115 VdB -- 124 VdB

(0.316 - 0.891 ips)

Severe Imbalance

Important

More then 125 VdB

(>1.00 ips)

Extreme Imbalance

Mandatory

The measured vibration level at 1X depends on the stiffness of the machine
mounting as well as the amount of imbalance, with spring-mounted machines
showing more 1X than solidly mounted machines for the same degree of imbalance.
The overall size of the machine also affects the allowable 1X level as follows:
1X Vibration Level,
VdB
Machine Type Repair Priority
109 VdB (0.158 ips)

Small Single-stage
Pump

Desirable

118 VdB (0.447 ips)

Large Hydraulic Pump

Desirable

111
116 VdB (0.355 ips)

Medium Sized Fan

Desirable

The tangential and radial 1X levels should be compared. The more nearly equal they
are, the more likely that imbalance is the cause. In any case, the direction in which
the machine has the least stiffness will be the direction of the highest 1X level.
More:

Imbalance in Vertically Mounted Machines
Imbalance in Overhung Machines
Sources of Imbalance
Imbalance in Vertically Mounted Machines

Vertical machines, such as pumps, are usually cantilevered from their foundation,
and they usually show maximum 1X levels at the free end of the motor regardless of
where the vibration source is. To isolate motor imbalance from pump imbalance, it
may be necessary to break the coupling and run the motor solo while measuring 1X.
If the 1X level is still high the problem is the motor; otherwise it is the pump.
Imbalance in Overhung Machines

In a machine with an out of balance overhung, or cantilevered, rotor such as a fan
will produce 1X vibration in the axial direction as well as some radial and tangential
at the nearest bearing to the rotor. This is because the imbalance creates a bending
moment on the shaft, causing the bearing housing to move axially. Examples of
overhung rotors are close-coupled pumps, axial flow fans, and small turbines.


Overhung Rotor Imbalance
The bearing closest to the overhung rotor will usually show the highest radial 1X-
vibration levels.
Sources of Imbalance

The following machine problems are among the conditions that will create imbalance:
Uneven dirt accumulation on fan rotors
Lack of homogeneity in cast parts, such as bubbles,
blow-holes, porous sections
Rotor eccentricity
Roller deflection, especially in paper machines
Machining errors
Uneven mass distribution in electric motor rotor bars
or windings
Uneven erosion and corrosion of pump impellers
Missing balance weights
Bowed Shaft
Misalignment

Misalignment is a condition where the centerlines of coupled shafts do not coincide.
If the misaligned shaft centerlines are parallel but not coincident, then the
misalignment is said to be parallel misalignment. If the misaligned shafts meet at a
point but are not parallel, then the misalignment is called angular misalignment.
112
Almost all misalignment conditions of machines seen in practice are a combination of
these two basic types.
More:

Parallel Misalignment
Angular Misalignment
General Misalignment
Temperature Effects on Alignment
Causes of Misalignment
Bent Shaft
Parallel Misalignment



Parallel Misalignment
If the machine speed can be varied, the vibration due to imbalance will vary as the
square of the speed. If the speed is doubled, the imbalance component will rise by a
factor of four, while misalignment-induced vibration will not change in level
Following is a typical vibration spectrum from a misaligned machine.

Angular Misalignment

Angular misalignment produces a bending moment on each shaft, and this generates
a strong vibration at 1X and some vibration at 2X in the axial direction at both
bearings, and of the opposite phase. There will also be fairly strong radial and/or
transverse 1X and 2X levels, but in phase.


Angular Misalignment
Misaligned couplings will usually produce fairly high axial 1X levels at the bearings on
the other ends of the shafts as well!

113
General Misalignment

Most cases of misalignment are a combination of the two above described types, and
diagnosis is based on stronger 2X peaks than 1X peaks and the existence of 1X and
2X axial peaks. Take care that high axial 1X levels are not caused by imbalance in
overhung rotors.
Misalignment produces a variety of symptoms on different types of machines, and
the average vibration signatures for healthy machines should be consulted to
determine allowable 1X and 2X levels.

Temperature Effects on Alignment

The best alignment of any machine will always occur at only one operating
temperature, and hopefully this will be its normal operating temperature. It is
imperative that the vibration measurements for misalignment diagnosis be made
with the machine at normal operating temperature.
Causes of Misalignment

Misalignment is typically caused by the following conditions:
Inaccurate assembly of components, such as motors, pumps, etc.
Relative position of components shifting after assembly
Distortion due to forces exerted by piping
Distortion of flexible supports due to torque
Temperature induced growth of machine structure
Coupling face not perpendicular to the shaft axis
Soft foot, where the machine shifts when hold down bolts are torqued.
Bent Shaft




Bent Shaft
Journal Bearings

Most journal-bearing problems will generate spectral peaks at lower frequencies than
1X, and these are called sub-synchronous peaks. Sometimes harmonics of these
sub-synchronous peaks are also created, indicating severe degradation of the
bearing. Here are some things to look for in diagnosing journal bearings:
More:

Oil Whirl
_AIntroduction to Machine Vibration-2
Oil Whip
Journal Looseness
Journal Thrust Bearings
114
Oil Whirl

Oil Whirl is a condition in which a strong vibration occurs at between 0.38X and
0.48X. It never shows up at precisely 0.5X, but is always a little lower in frequency.
It is caused by excessive clearance and light radial loading, which results in the oil
film building up and forcing the journal centerline to migrate around in the bearing
opposite the direction of rotation at less than one-half RPM. Oil whirl is a serious
condition and needs to be corrected when found, for it can deteriorate fairly quickly
to the point where metal-to-metal contact occurs in the bearing.
Oil Whip


The solutions for oil whip and oil whirl are suitably small bearing clearances and
adequate radial loading. When bringing a large turbine up to speed, it is important to
pass through the critical frequencies very quickly to prevent the buildup of oil whip.
Journal Looseness



Journal or bearing housing looseness
One half, one third, and one fourth-order harmonics are sometimes called sub
harmonics.
Journal Thrust Bearings

Worn thrust bearings usually present strong axial components at the first few
harmonics of 1X. Worn Kingsbury bearings with 6 shoes will generate a peak at 6X.
This vibration peak is predominantly in the axial direction.
Rolling Element Bearings

Many years of experience have shown that in practice, less than 10 % of all bearings
will run for their design lifetime. About 40 % of bearing failures are attributed to
improper lubrication, and about 30% of failures are from improper mounting, i.e.
115
misalignment or "cocking". About 20 % fail for other reasons, such as overloading
and manufacturing defects, etc.





These are the formulas for calculating the frequencies of the bearing tones from the
bearing geometry, but they are a little imprecise because the axial loading and
slippage affects them in an unpredictable manner.

The number of rollers in most bearings is usually between 8 and 12, but in very large
diameter bearings, such as the ones found in paper machines, the number of rollers
can be much higher.
More:

Rolling Element Bearing Wear
Sidebands
Misaligned ("cocked") Rolling Element Bearings
Rolling Element Bearing Looseness
Rolling Element Bearing Wear

The first stages of bearing defects will produce telltale non-synchronous vibration
frequencies called "bearing tones" and their harmonics. Bearing tones at 0.006
inches per second peak (81 VdB) or higher are considered significant. Sometimes a
new bearing will produce bearing tones, possibly because of damage during
installation, shipping, or defective manufacture.
116




If the bearing defect is very small in size, such as a crack in one of the races, the
vibration signature will show harmonics of the bearing tone with little or no
fundamental frequency present. If the defect begins as a spall over a larger area of
the race, the bearing tone fundamental will usually be higher in level than the
harmonics. As the defect becomes worse, the overall level of the bearing tones will
increase, as will the overall broadband noise level.
Sidebands

If the defect is on the inner race of the bearing, the turning speed will amplitude
modulate the bearing tones, and this will cause sidebands around the bearing tones,
spaced apart at 1X, to appear. The amplitude modulation comes from the fact that
the defect on the inner race moves in and out of the bearing load zone once per
revolution. While in the load zone, the defect produces vibration at the ball pass
frequency, but when it is out of the load zone, very little vibration is produced at this
frequency. This accounts for the amplitude modulation of the bearing tone and the
consequent sidebands. Sidebands spaced at 1X around bearing tones are a sure sign
of advanced bearing wear. Sometimes, if a rotor is strongly out of balance, an inner-
race bearing defect will not produce amplitude modulation or sidebands. This is
because the centrifugal force due to imbalance keeps the inner race loaded at the
same location on its periphery all the time.
Another example of sidebands in bearing spectra involves the Fundamental Train
Frequency (FTF). This is the rate at which the cage holding the rollers rotates in the
bearing. If one roller is spalled, cracked, or worse yet, in several pieces, it will make
a lot of noise when it is in the load zone of the bearing, but will be quiet when not in
the load zone. It will move in and out of the load zone at the FTF rate because it
migrates around the bearing with the cage. This causes amplitude modulation of the
bearing tones at the FTF rate, and the result is sidebands around the bearing tones
spaced apart by the FTF.
117



The final stage of bearing wear is sometimes called the "thermal" stage, where the
bearing becomes hot, breaking down the lubricant, leading to catastrophic failure
which can include melting of the rolling elements and/or the races.
The key to effective predictive maintenance of bearings is the trending of bearing
tone levels over time from their onset. Sometimes a bearing condition will progress
from a very small defect to complete failure in a relatively short time, so early
detection requires sensitivity to very small vibration signature components. The
analyst should be aware that some types of machines will show bearing tones in the
average spectra. Diagnosis is made on the basis of significant increases from these
average values. Any significant bearing tone should be carefully watched for signs of
worsening.
Misaligned ("cocked") Rolling Element Bearings



Cocked Bearing
Rolling Element Bearing Looseness

Excessive clearance in a rolling element bearing will produce harmonics of 1X,
usually in the range from 2X to 8X. Extreme looseness will commonly produce one-
half order components, i.e., components at multiples of 0.5X. Looseness in other
parts of the machine will also produce 1X harmonics and sometimes 0.5X harmonics,
so this is not a conclusive sign of bearing clearance problems.
Mechanical Looseness


118


Mechanical Looseness
More:

Rotating Looseness
Non-Rotating Looseness
Rotating Looseness
Non-Rotating Looseness

Looseness between a machine and its foundation will increase the 1X vibration
component in the direction if the least stiffness. This is usually the horizontal
direction, but it depends on the physical layout of the machine. Low-order 1X
harmonics are also commonly produced if the looseness is severe. It is often hard to
tell imbalance from foundation looseness or flexibility, especially in vertical machines.
If 1X tangential is much greater than 1X radial, looseness is suspected. If 1X
tangential is lower than or equal to 1X radial, then imbalance is suspected.
Foundation flexibility or looseness can be caused by loose bolts, corrosion, or
cracking of mounting hardware.

Electrically Induced Vibration


More:

AC Electric Motors
Sources of Vibration
Mechanical Sources of Vibration in Motors
Rotor Bar Problems
D.C. Motors
AC Electric Motors

There are two types of AC electric motors; the synchronous motor and the induction
motor, and single phase or 3-phase current may power each of the types. In
industrial applications, 3-phase motors are by far the most common, owing to their
higher efficiency than single-phase units. The synchronous motor is much less
119
prevalent than the induction motor, but is used in some special applications requiring
absolutely constant speed, or for power factor correction. Induction and synchronous
motors are similar in many respects, but differ in some details.





More:

Synchronous Motors
Induction Motors
Synchronous Motors




An interesting characteristic of the synchronous motor is that is the rotor is "over
excited, i.e., if its magnetic field is greater than a critical strength, the motor
behaves like an electrical capacitor connected to the power line. This has been used
for power factor correction in industrial plants that use large numbers of induction
motors.

Induction Motors


120
Induction Motor Rotor
Because the induction motor works by magnetic repulsion rather than attraction like
the synchronous motor, it has been called a "repulsion induction" motor.
If there were no friction in the system, the rotor would turn at synchronous speed,
but the motor would produce no useful torque. Under this condition, there would be
no relative motion between the rotor bars and the rotating stator field, and no
current would be induced in them. As soon as any load is applied to the motor the
speed is reduced, causing the rotor bars to cut the magnetic lines of force of the
stator field, and creating the repulsion force in the rotor. The induced magnetic field
in the rotor migrates around in the direction of the rotation, and the speed of this
migration is dependent on the applied load. This means the RPM will always be less
than synchronous speed. The difference between the actual speed and synchronous
speed is called the "slip". The greater the slip, the greater the induced current in the
rotor bars, and the greater the output torque. The current in the stator windings also
increases in order to create the larger currents in the bars.
For these reasons, the actual speed of an induction motor is always dependent on
the load.
Sources of Vibration

Twice the line frequency (120 Hz in the US) is always a measurable vibration
component in an electric motor. The attraction between the stator and rotor varies at
this rate, and the iron itself changes dimension a little in the presence of the varying
magnetic field due to magnetostriction.
More:

Slip-related vibration
Slot Pass Frequency
Shorted Laminations
Slip-related vibration

Irregularities in the rotor bars will cause vibration at the slip frequency times the
number of poles in the motor. For instance, in a two-pole motor, any particular rotor
bar will be aligned with the rotating magnetic pole created by the stator at two times
for every "slip cycle". The slip cycle is the synchronous speed divided by the slip
speed. For instance, in a 3450 RPM motor, the synchronous speed is 3600 RPM, and
the slip frequency is 3600 - 3450 = 150 RPM. Then, 3600 150 = 24, which is the
slip cycle. This means for every 24 revolutions of the rotor, the same rotor bar will
be exactly aligned with the same polarity of the rotating magnetic pole, and will be
aligned with the opposite rotating pole once every 12 revolutions.
If one rotor bar has more resistance than the others due to a crack or break, it will
have less current induced in it when it is aligned with the poles, and this will produce
a little less torque at this point in its slip cycle. Thus, the torque will be modulated at
the slip frequency times twice the number of poles. This frequency is also called the
pole pass frequency. Pole pass is seen in a vibration component in the signature, and
also results in sidebands around the 1X vibration component and around the 120 Hz
component.
Another vibration component in electric motors is the so-called slot pass frequency.
This frequency is the number of stator slots times the RPM. The stator slots contain
the conducting windings, and their finite number creates a non-uniformity, or
"cogging" in the rotating magnetic field that in turn causes a vibration component.
The rotor bars are also in slots, and the rotor bar pass frequency is also sometimes
called a slot pass frequency, and is different than the stator slot pass.
121
Slot Pass Frequency

Another vibration component in electric motors is the so-called slot pass frequency.
This frequency is the number of stator slots times the RPM. The stator slots contain
the conducting windings, and their finite number creates a non-uniformity, or
"cogging" in the rotating magnetic field that in turn causes a vibration component.
The rotor bars are also in slots, and the rotor bar pass frequency is also sometimes
called a slot pass frequency, and is different than the stator slot pass.
Shorted Laminations

The rotor and stator of AC motors are made of thin laminations that are isolated
from each other. This prevents magnetically induced currents from circulating in the
iron and cause heating. If the laminations are shorted together in some locations,
local heating and resultant thermal warping will occur. Shorted laminations also
cause higher 120 Hz vibration levels.
Electric motors suffer from all the mechanical ailments common to other rotating
machines, with a few additions, as discussed below.
More:

Rotor Thermal Bow
Air Gap Eccentricity
Loose Rotor
Eccentric Rotor
Loose Windings
Rotor Thermal Bow

Uneven heating of the rotor due to unbalanced rotor bar current distribution causes
the rotor to warp, or "bow", and rotor bow results in an imbalance condition with all
its usual symptoms. It can be detected by the fact that it goes away when the motor
is cold.
If the air gap is not uniform, the forces on the rotor are not balanced, resulting in
high magnetically induced vibration at 120 Hz. The magnetic attraction is inversely
proportional to the square of the distance between the rotor and stator, so a small
eccentricity causes a relatively large vibration.
Air Gap Eccentricity

If the air gap is not uniform, the forces on the rotor are not balanced, resulting in
high magnetically induced vibration at 120 Hz. The magnetic attraction is inversely
proportional to the square of the distance between the rotor and stator, so a small
eccentricity causes a relatively large vibration.
Loose Rotor

Sometimes the rotor can slip on the shaft, usually intermittently depending on
temperature, and this causes severe vibration at 1X and harmonics. Abrupt changes
in load or line voltage can instigate this.
Eccentric Rotor

If the rotor is not round, it will cause 1X excitation and unbalanced magnetic forces
that cause vibration at slip frequency times the number of poles. This component will
disappear immediately when the power is cut, and this is a confirming test.

Loose Windings

If the electrical windings of the motor stator are even a little loose, the vibration
level at 120 Hz will be increased. This condition is very destructive because it
abrades the insulation on the wire, leading to shorted turns and eventual short
122
circuits to ground and stator failure. In some large machines such as AC generators,
loose windings will generate one-half order harmonics of the 120 Hz excitation
frequency.
Rotor Bar Problems

An important failure mode of large electric motors is the cracking and subsequent
heating and breaking of the rotor bars, especially in motors that experience frequent
starts under load. The starting condition places the heaviest stress on the rotor bars
because they are carrying the highest current since the rotor is running at much
lower than synchronous speed. The high currents cause heating and expansion of the
bars relative to the rotor itself, and differences in the electrical resistance of the
individual bars result in uneven heating and uneven expansion. This leads to
cracking of the joints where the bars are welded to the shorting ring. As soon as a
crack develops, the resistance of that bar increases, increasing its heating, and
consequently worsening the crack. At the same time, the adjacent rotor bars
experience increased currents because of the reduced current in the broken bar.
This scenario results in localized heating of the rotor, causing it to warp. See the
paragraph on Rotor Thermal Bow, above.
More:

Rotor Bar Monitoring via Motor Current Analysis
The condition of the rotor bars in an induction motor can be measured by performing
a high-resolution frequency analysis on the input current to the motor.
The presence of a defective rotor bar will cause the motor torque to be reduced
slightly every time a pole of the rotating magnetic field passes by it. This happens at
twice the slip frequency, for both the north and south poles of the field cause a
momentary reduction. This reduction in torque also results in a reduction in the input
current to the motor at the same rate -- this is a result of conservation of energy.
This periodic reduction in the motor current is actually an amplitude modulation of
the motor current. The amount of modulation is related to the severity of the rotor
bar problem.
A good way to detect motor current modulation is to perform a frequency analysis on the
current, and look at the sidebands around 60 Hz spaced at twice the slip frequency. This can
be done with a current clamp placed around one phase of the input line and connecting it to
a spectrum analyzer.


123


For this test, the motor must be operating under load, for with no load, the slip will
be very slow, and no appreciable torque is being developed.
Only one phase of a three-phase motor need be measured.
The spectrum analyzer must be capable of generating a high-resolution spectrum
from 0 Hz to about 70 Hz, or a zoom spectrum from 50 Hz to 70 Hz. A frequency
resolution of 1600 lines is desirable in order to separate the 2X slip sidebands from
other sidebands caused by load variations, etc.
The high resolution and the zoom spectrum are desirable because the slip frequency
sidebands will be very close in frequency to the 60 Hz line frequency. For instance,
for a motor turning 1760 RPM, the slip frequency will be 1800 - 1760 = 40 RPM,
which is equivalent to 0.667 Hz. The sidebands will be spaced at twice this frequency,
or 1.334 Hz.
If the sidebands are 55 to 60 dB (1,000:1) down from the 60 Hz peak, the rotor bars
are considered good, but if they rise to 40 dB (100:1) below the 60 Hz peak,
damaged rotor bars are indicated. It is possible to calibrate a system like this to
relate the actual number of open bars to the sideband level if the number of bars in
the rotor is known.



The spectrum above is from a 1760-RPM motor with rotor bar problems.
D.C. Motors


124

Misshapen or pitted commutator segments or improper brush contact can cause
excessive vibration in D.C. motors with the commutator. The frequency will be at the
segment pass frequency, which is the number of commutator segments times the
RPM.


If the 360 Hz peak in the vibration spectrum rises significantly, the likely cause is
probably open circuited field windings, loose electrical connections, or malfunctioning
SCRs.
Turbines

Gas and steam turbines are essentially similar mechanically, with gas turbines
having the added complication of a combustion chamber. Gas turbine vibration
signatures commonly contain a broadband vibration component caused by the
combustion noise.
More:

Turbine Diagnostics
Turbine Diagnostics

Turbines often exhibit a strong component at the so-called blade rate, which is the
number of turbine blades times the RPM of the rotor. The magnitude of this
component is dependent on the internal geometry of the unit. If this changes, as for
instance by a cracked, warped, or pitted blade, the blade pass component in the
vibration signature will change, usually for the worse.
If the turbine blades wear uniformly, the blade pass frequency is quite uniform, but if
a portion of the rotor is damaged, such as a broken blade, the blade pass component
will be modulated by the RPM of the rotor or by the number of nozzles in the turbine
times the RPM, causing sidebands in the spectrum.
Pumps

There are many types of pumps in common use, and their vibration signatures vary
over a wide range. When monitoring pump vibration, it is important that the
operating conditions are uniform from one measurement to the next to assure
consistent signatures. Suction pressure, discharge pressure, and especially air
induction and cavitation will affect the vibration signature.
More:

Centrifugal Pumps
Gear Pumps
Screw Pumps
Centrifugal Pumps

The following spectrum, containing broadband high-frequency noise, indicates
cavitation in a centrifugal pump due to low inlet pressure.
125

Cavitation in Centrifugal Pumps

Cavitation produces this type of spectrum at all measurement points of the pump
and the housing


Gear Pumps

Gear pumps are commonly used for pumping lube oil, and they almost always have a
strong vibration component at the tooth mesh frequency, which is the number of
teeth on the gear times the RPM. This component will be highly dependent on the
output pressure of the pump. If the tooth mesh frequency changes significantly, such
as the sudden appearance of harmonics or sidebands in the vibration spectrum, it
could indicate a cracked or otherwise damaged tooth.


126
Typical Gear Pump Spectrum
Screw Pumps

The screw type pump can generate a multitude of frequency components in the
vibration spectrum. Thread wear or damage will usually produce strong harmonics of
the thread rate, which is the number of threads times the RPM.
Fans

Most fans are either axial flow propeller-type fans, or are centrifugal. Fans, especially
when they are handling particle-laden air or gas, are prone to uneven buildup of
detritus on the blades. This causes imbalance, and should be corrected as soon as it
is diagnosed. If any of the blades become deformed, cracked, or broken, the blade
pass frequency vibration peak will increase in level, and if there are many blades,
sometimes 1X sidebands will appear around the blade pass frequency.
More:

Axial Flow Fans
Centrifugal Fans
Axial Flow Fans






Centrifugal Fans

127


A common problem in centrifugal fans is uneven supply air velocity distribution
across the inlet, and this causes increased vibration levels at the blade pass rate. If
the fan is out of balance and is overhung, high 1X vibration will occur in axial as well
as both radial directions.
Defective blades can also cause 1X sidebands around the blade pass frequency.
Couplings

Couplings exist in many types and configurations, and defects in them usually cause
symptoms similar to misalignment. Frequently coupling problems will produce
stronger 1X vibration components than simple misalignment does. If the coupling is
not true, i.e., has non-parallel flange faces, a vibration similar to angular
misalignment is produced.
Coupling imbalance is also a common problem, and results in high 1X and 2X radial
and tangential components.
Coupling wear can produce all the symptoms of misalignment and looseness. Three-
jaw motor couplings that contain spacers of improper length will cause strong axial
and radial components at 3 times shaft RPM.

Drive Belts

Belt drives are relatively inexpensive types of power transmissions, but they are
prone to many problems. There are many types of drive belts, and all are subject to
wear and damage. Belts should be frequently inspected for damage and should be
kept at the proper tension and kept clean.
More:

Mismatched, Worn, or Stretched Belts
Eccentric Sheaves, Sheave Runout
Sheave Misalignment
Belt Resonance, or Belt Slap
Mismatched, Worn, or Stretched Belts

Mismatched, worn, or stretched belts, especially Vee belts, will generate vibration at
the fundamental belt pass frequency and harmonics of it. Usually the second
harmonic is dominant if there are two sheaves in the system. The fundamental belt
frequency FBF is given by the following formula. It is always sub-synchronous,
meaning it is lower in frequency than 1X.

128


Where D = Sheave Diameter
L = Belt Length
RPM = Turn speed of sheave D
Eccentric Sheaves, Sheave Runout

Eccentric sheaves will generate strong 1X radial components, especially in the
direction parallel to the belts. This condition is very common, and mimics imbalance.
This can be checked by removing the belts and measuring again. 1X vibration of an
eccentric sheave or a sheave with runout will usually also show up at the other
sheave.


Eccentric Sheave
Sheave Misalignment

Sheave misalignment will generate strong axial 1X components and axial harmonics
of the fundamental belt frequency.


Belt Resonance, or Belt Slap



129

Belt Resonance

Centrifugal Compressors

Centrifugal compressors generate spectra similar to centrifugal fans in that the vane
pass frequency will be dominant. Damaged or eroded vanes will cause increases in
the level at the vane pass, and also will usually produce 1X sidebands around the
vane pass. Compressor surge is a fluid dynamic problem at the compressor output
port that usually causes vibration at less than 1X frequency. It is often caused by
improper output pressure.
Below is a typical vibration spectrum from a 6-vaned centrifugal compressor.


Vane Pass Harmonics
Reciprocating Machines

The most common types of reciprocating machines are piston pumps and
compressors and internal combustion engines. In all these machines, the piston rate
(usually 1X) is dominant, along with the firing rate for 4-cycle engines. Vibration
levels as high as 125 VdB (1.0 inches per second peak) are not uncommon for
healthy machines such as these. The analyst must judge the machine condition by
comparison to previous levels rather than applying absolute reference levels.
Many reciprocating engines have turbo chargers, and they are diagnosed like other
rotating turbines and compressors. Camshaft gear problems are also common, and
can be seen by looking for the tooth mesh frequency. If the engine has a torsional
vibration damper on the shaft, it can fail, greatly increasing vibration at the
frequency of the first crankshaft torsional vibration mode. This frequency must be
obtained from the engine maker.
Variable displacement piston pumps are much smoother than compressors, and lend
themselves well to vibration analysis. If harmonics of the piston rate are present in
130
significant levels, it usually indicates a piston drive linkage problem. A very
prominent tone at piston frequency fundamental may indicate a worn spot on the
wobble plate.

Diagnostic Summary Charts

The following charts summarize most of the information on machine diagnostics, but
are not meant to be exhaustive.
More:

Imbalance
Misalignment
Bent Shaft
Journal Bearing Problems
Rolling Element Bearing Problems
Rolling Element Bearings -- Continued
Mechanical Looseness
Electric Motor Problems
Pump Problems
Turbine Problems
Fan Problems
Compressor Problems
Drive Belt Problems
Gear Problems
Diesel Engines
Imbalance

Vibration
Source

Exciting
Frequency
Dominant
Plane
Amplitude Spectral
Envelope
Characteristics
Comments
Mass
Imbalance:
Static

1X

radial*

steady

narrow band

Rotor bow
due to
thermal
stresses may
cause a rise
in amplitude
with
temperature


Dynamic


1X radial generally, some
1X harmonics
Most common
form of
imbalance.
Couple


1X radial, axial
Overhung
Rotor
1X axial, radial
131


*The Radial plane includes the tangential direction in all the tables.
Misalignment


Vibration
Source

Exciting
Frequency

Dominant
Plane

Amplitude

Envelope
Characteristic
s

Comments
Angular
Misalignment
1X, 2X axial steady narrow band Most
misalignment is
a combination of
parallel and
angular.

Parallel
Misalignment
1X, 2X radial steady narrow band On long coupled
spans 1X will be
higher.

Combined
Angular &
Parallel
1X, 2X radial, axial steady narrow band Misalignment is
also indicated
by multiples of
2X

Cocked
Bearing
2X, increased
1X, and
bearing tones
radial,
tangential,
axial
high,
steady
narrow band Usually
accompanied by
axial
components.

Misaligned
Impeller
2X, increase
in vane rate
harmonics
radial steady narrow band Usually
accompanied by
low axial
amplitudes.

Gear
Misalignment
Strong gear
mesh rate

radial,
tangential,
axial
steady Usually 1X
sidebands
around gear
mesh rate
Harmonics of
gear mesh are
common
Bent Shaft


Vibration Source

Exciting Frequency

Dominant Plane

Amplitude

132
Envelope
Characteristics

Comments

Slightly Bowed
Shaft
1X , 2X axial,
radial,
tangential
steady narrow band Run-out @ rotor
mass appears as
imbalance.

Shaft Bent at
Coupling

1X, 2X axial,
radial,
tangential
steady narrow band,
maybe with
2X, 3X
harmonics
Run-out @
coupling appears
as misalignment
Journal Bearing Problems


Vibration Source

Exciting Frequency

Dominant Plane

Amplitude

Envelope
Characteristics

Comments

Oil Whirl

0.38X to
0.48X
radial Sharp peak
Oil Whip 0.38X to
0.48X
radial Sharp peak Run-out at rotor
mass appears as
imbalance.
Run-out at coupling
appears as
misalignment.

Excessive Bearing
Clearance
1X
harmonics
radial hump in 1X
harmonic series

4X to 8X and/or 7X
to 15X
Journal
Bearing
looseness,
Rattling
0.5X, 1X radial 0.5X harmonics
Journal
Thrust
Bearings,
1X; shoe
rate of
Kingsbury
axial 1X harmonics; shoe
rate harmonics for
Kingsbury
Usually six shoes
133
Kingsbury

Rolling Element Bearings -- Continued


Vibration
Source
Exciting
Frequency

Dominant Plane

Amplitude

Envelope
Characteristics

Comments

Defective Cage
Broken at one
point
Broadband
noise

radial


broad band

Low level
noise

Cage broken in
pieces
Broadband
noise
radial broad band Noise due to
debris from cage
in the bearing
Inadequate
Preload
or Lubrication
High-
frequency
noise
"haystack"
radial

broad
band
Natural
frequency of
the bearing.

Bearing
Looseness
turning on the
shaft
1X harmonics

radial


narrow band

No. and
amplitude of the
harmonics is a
function of
looseness.

Loose in
housing
1X, 2X, 4X

radial narrow band
Extreme
Looseness
or failure
0.5X, and
harmonics
Rising low
frequency noise
floor.
Excessive
Bearing

4X to 8X and/or

134
Clearance
1X harmonics

radial

7X to 15X

Cocked Bearing 1X, 2X,
bearing tones
axial, radial high,
steady
narrow band 180 phase
difference in axial
measurements on
each side of
bearing housing
Mechanical Looseness


Vibration Source

Exciting Frequency

Dominant Plane

Amplitude

Envelope
Characteristics

Comments

Foundation Looseness 1X
harmonics
Usually
Tangential
steady narrow
band
Indicated by
foundation
flexibility.

Journal Bearing
Looseness
1X
harmonics
radial steady narrow
band
Harmonics may
extend to 10X.
Extreme journal bearing
looseness
0.5X half-
order
harmonics
radial steady Sometimes
0.25 X
harmonics also
present
Electric Motor Problems


Vibration Source

Exciting Frequency

Dominant Plane

Amplitude
135

Envelope
Characteristics

Comments

Improper Brush -
Commutator Contact
(D.C. Motors)


Radial,
Transverse
narrow
band
n = any
positive
integer
C = Number
of
commutator
segments.
X = machine
speed (RPM)

Broken Rotor Bars 2 x Slip x
No. of
Poles
Radial,
Transverse
Sometimes
Beating
narrow band Sometimes
causes 2 x
Slip
sidebands
around 120
Hz
Induction Motor Slot
Pass


Radial,
Transverse
narrow
band
S = rotor
slot pass
frequency,
Hz
B = number
of rotor bars
X = rotor
speed (RPM)
120 = twice
line
frequency ,
Hz

Pump Problems


Note: V = number of pump vanes
T = number of gear teeth
S = number of screw threads

Vibration
Source

Exciting
Frequency

Dominant Plane
136

Amplitude

Envelope
Characteristics

Comments

Non-Rotating
Looseness

1X, 2X,
3X

radial steady narrow
band
Harmonics can extend to
10 X
Rotating
Looseness
(rotors, impellers,
etc.)
1X radial varies
startup to
startup
narrow
band
Sometimes 0.5X harmonics
also
Centrifugal Pumps
with (V) Vanes

Vane
pass =
VX
radial fluctuating
vane rate
harmonics
of pump
In large pumps strongest
amplitude occurs at vane
rate.
In smaller pumps
strongest amplitudes occur
at harmonic of vane rate.
Gear Pumps with
(T)
Teeth

tooth
mesh =
TX
radial More than one discharge
volute (as in multiple rotor
pumps) will create
harmonics of tooth mesh
frequency.

Rotor Rub

0.5X, 1X radial steady narrow
band
May excite rotor critical.
Screw Pumps

SX radial S = number of screw
threads
Cavitation or
Starvation

random radial fluctuates broadband Random noise sometimes
up to 20 kHz
Turbine Problems


Vibration Source

Exciting Frequency

Dominant Plane

Amplitude

Envelope
Characteristics
137

Comments

Rotor Clearance
Problem
blade pass
rate
radial steady narrow band
Turbine Blade
Damage
1X and
harmonics,
blade pass
radial steady narrow band harmonics
usually higher in
level than 1X
1X sidebands
around blade
pass

Fan Problems


Vibration Source

Exciting Frequency

Dominant Plane

Amplitude

Envelope
Characteristics

Comments

Fan Housing, Blade
Clearance Problem
blade rate
= X times
No. of
blades
radial,
axial,
tangential
steady narrow
band
Sometimes
blade rate
harmonics
present
Fan Imbalance

1X Radial,
tangential
steady narrow
band

Fan Pitch Problem

1X axial steady narrow
band

Uneven Air Velocity

blade rate radial,
tangential
steady narrow
band

Compressor Problems


Vibration Source

Exciting Frequency

138
Dominant Plane

Amplitude

Envelope
Characteristics

Comments

Diffuser Type

blade
rate
radial,
transverse
steady narrow
band
Sometimes
blade rate
harmonics are
present
Piston Type

2X radial, axial steady narrow
band
1X harmonics
common
Drive Belt Problems


Vibration Source

Exciting
Frequency

Dominant Plane

Amplitude

Envelope
Characteristics

Comments

Mismatched, Worn
or Stretched Belts
multiples
of belt
frequency
(B). 2B
usually
strongest
radial
in line
with
belts
may be
unsteady
beating if
the 2B is
near speed
of either
shaft
B is always less than 1X
Eccentric and/or
Unbalanced Sheaves

1X shaft radial steady Easily confused with imbalance
Drive Belt or Sheave
Face Misalignment
1X driver axial steady Confirm with strobe light
Drive Belt
Resonance
varies radial may be
unsteady
Belt resonance with no
relationship to rotational
speed.
139
Improper Belt
Tension

Can produce belt resonance
(see above) Can increase
bearing wear
Gear Problems


Vibration
Source

Exciting
Frequency

Dominant
Plane

Amplitude

Envelope
Characteristics

Comments

Improper Tooth
Contact (Mesh)
X times Gear
Tooth Count
radial,
axial
steady,
sometimes
with beats
narrow
band
Often with 1X of
either gear sidebands
Gear Eccentricity 1X, tooth mesh radial narrow
band
Gear may be
balanced but
mounted on an
unbalanced rotor.
1X sidebands around
tooth mesh are
present
Gear
Misalignment
2X, tooth
mesh
axial steady narrow
band
1X sidebands around
mesh rate

Pitch Line
Runout,
Mass Unbalance
or
Faulty Tooth
1X plus tooth
mesh frequency
radial for
spur
gears
plus axial
for single
or double
helical

steady narrow
band
1X sidebands around
tooth mesh
Machining Errors tooth mesh,
"Ghost"
frequency,
usually not
synchronous with
1X
radial,
axial
steady narrow
band
Machining errors due
to hobbing and
cutting table drive
gear mesh problem
can cause "ghost"
component.
140
Planetary Gear
Problems
Strongly
dependent on
gearbox
geometry
radial steady narrow
band
Sidebands around
tooth mesh spaced at
planet orbiting rate
are common.
Diesel Engines



Vibration Source

Exciting
Frequency

Dominant Plane

Amplitude

Envelope
Characteristics

Comments

Cylinder Misfiring

0.5 X radial,
tangential
5 mils narrow
band
No diagnostics can
be done until all
misfiring is corrected
Imbalance,
Misalignment Torque
Reaction,
Weak Foundation
1X radial,
tangential
5 mils narrow
band
1X and No. of firing
strokes per
revolution are
caused by torque
reaction, which in
turn can cause
misalignment
Harmonic Balancer 2X radial,
tangential
4 mils narrow
band
Correct by rotating
the balancer
Cylinder Combustion
Forces
1.5X,
2.5X, etc
radial,
tangential
4.3 mils narrow
band
Normal operating
levels






141

Estimating Vibration Severity

More:

Absolute Vibration Levels
Absolute Vibration Levels

Once a specific machine problem is identified by its vibration signature, the next
question should be "Is the problem bad enough to require maintenance?". There is
no general agreement on how to do this but we will look at several approaches that
have proven to be reasonably successful in practice:


More:

Rathbone Chart
ISO Standard 2372
MIL-STD-167-1 and MIL-STD-167-2
NAVSEA Technical Specification S9073-AX-SPN-010/MVA
Commercial Standards (DLI Machinery Vibration Severity Chart)
Rathbone Chart

The Rathbone Chart, devised by T. C. Rathbone in 1939, compares overall vibration
velocity to varying degrees of machine smoothness. Rathbone made no inferences as
to frequency content of the vibration, or to machine size. The Rathbone Chart is
considered obsolete today, and is presented here only for historical interest.
Machine Running Condition Overall Vibration Velocity
Very Rough

0.628 ips peak

Rough

0.314 ips peak

Slightly Rough

0.157 ips peak

Fair

0.0785 ips peak

Good

0.0392 ips peak

Very Good

0.0196 ips peak

Smooth

0.0098 ips peak

Very smooth

0.0049 ips peak

ISO Standard 2372

The ISO standard number 2372 provides vibration amplitude acceptance guidelines
for rotating machinery operating from 600 to 12000 RPM. It specifies overall
vibration velocity levels rather than spectral levels, and can therefore be quite
misleading.
ISO 2372 specifies the RMS vibration velocity limits on a basis of machine
horsepower, and covers a frequency range from 10 Hz to 1000 Hz. Because of the
142
limited high frequency range, rolling element-bearing problems can be easily missed.
This standard is considered obsolete, and is about to be rewritten.
Level,
VdB
Level,
IPS
Less Than
20 HP
20 to 100
HP
More
Than100 HP
125

1.00

Not Permissible

Not Permissible

Not Permissible

121

0.63

Not Permissible

Not Permissible

Just Tolerable

117

0.40

Not Permissible

Just Tolerable

Just Tolerable

113

0.25

Just Tolerable

Just Tolerable

Allowable

109

0.16

Just Tolerable

Allowable

Allowable

105

0.10

Allowable

Allowable

Good

101

0.06

Allowable

Good

Good

97

0.04

Good

Good

Good
MIL-STD-167-1 and MIL-STD-167-2

These standards that date from 1974 are an attempt to
provide a threshold vibration level as a function of
frequency for acceptance testing of rotating machinery.
MIL-STD-167-1 covers internally excited vibration of all
rotating equipment except reciprocating machinery, and
MIL-STD-167-2 covers reciprocating machinery,
propulsion systems and shafting. They have been used
for many years, and are generally considered to be quite
outdated. They are based on a displacement (mils peak)
spectrum, which is actually equivalent to a constant
velocity of 0.13 inches per second (107 VdB) above
1200 RPM. These standards continue to be used as a
rough reference for acceptable vibration levels for
medium sized simple machines such as electric motor
pumps, but they should not be used as any absolute
standard.


NAVSEA Technical Specification S9073-AX-SPN-
010/MVA

This is a more recent standard, dated 1978, and was issued by the Naval Sea
Systems Command (NAVSEA). It is based on average vibration signatures, and
states that the acceptance criterion after machine overhaul is no higher than one
standard deviation ("one sigma") above average spectral levels. The acceptance
criterion during normal operation is the average spectral level plus two sigma.
143
Commercial Standards (DLI Machinery
Vibration Severity Chart)

The chart shown here can be applied to a large number of rotating machines with
reasonable confidence. It is a distillation of data from a wide range of industrial
machinery, and is considered more up to date and useful than the above mentioned
standards.
Vibration
Level
Below 30 Hz 30 Hz - 1000
Hz
Above 1000
Hz
Extreme

10 mils p-p

125 VdB rms

0.282 G rms

Excessive

4.2 mils p-p

117 VdB rms

0.020 G rms

Tolerable

1.5 mils p-p

108 VdB rms

0.007 G rms

Acceptable

0.6 mils p-p

100 VdB rms

0.003 G rms


The same information is shown graphically below:










144
Index

A

acceleration 12, 15, 16, 47, 48
Acceleration 68
amplitude 10, 12, 13, 14, 16, 28, 34, 36, 37, 39, 42, 47
Amplitude demodulation 83
Amplitude modulation 83, 84, 85
Angular Misalignment 109
Average 13, 15
B

Bearing Mobility 102
bearing tones 71, 64, 114, 115
Beats 36, 37
Belt drives 64, 128
blade rate 125
bump test 32
C

carrier 34, 35
Cavitation 125
Centrifugal Compressors 133
Centrifugal Fans 133
Commercial Standards 148
Compressor surge 133
continuous spectrum 32
Couplings 110
crest factor 69, 83
Crest factor 69
critical 22, 69
D

D.C. Motors 124
dB 42, 47, 71
degrees of freedom 9
Deterministic 28, 30, 32
differentiation 16
displacement 10, 14, 15, 16, 31, 45
Displacement 62, 68, 134
Dynamic Imbalance 106
E

Eccentric Sheaves 129
excitation force 9, 29
F
145
fans 108, 133
fault 1-3, 8, 26, 39
FFT 29, 53
FFT analyzer 62
First Order 64
Forcing Frequencies 71, 61
frequency 9, 10, 16, 19, 24, 26, 28, 29, 30, 32, 34, 35, 36, 45, 46, 47, 48
Frequency Analysis 27, 29, 32
frequency domain 24, 25, 29
FTF 115
Fundamental Train Frequency 115
G

G 15, 16
Gear eccentricity 132
Gear pumps 126
ghost 63
H

harmonic 15, 28, 30
harmonic series 28
harmonics 64, 68, 111, 112, 114, 115, 116, 122, 127, 128, 129, 134
I

ICP 48
imbalance 61, 71, 104, 106, 107, 108, 117, 122, 127, 128, 129
Induction Motors 67
Integration 16
ISO Standards 147
J

jerk 15
journal bearing 111
K

Kurtosis 83
L

linear 19, 29, 34, 36, 39, 40, 41, 43, 47
Log 39, 43
logarithm 43
Loose Windings 122
looseness 68, 116, 128
M

Magnetostriction 120
mils 62, 147
146
Mil-Std 167-2 146
misalignment 109, 110, 113, 128, 129
Modal Analysis 12
Modulation 34, 35, 36
N

NAVSEA Technical Specification 148
non-synchronous components 64
O

Oil Whirl 111
Orbit plots 67
oscillation 9, 10
overhung 107, 108, 110, 128
P

Parallel Misalignment 109
Peak Amplitude 13
Peak-to-Peak 13, 16
pendulum 9
period 10, 13, 28, 30, 50
phase 10, 13, 15, 36, 37, 47
Pumps 107, 110, 125, 134, 147
Q

quasi-periodic 28
R

Random Noise 70, 68
Reciprocating Machines 69, 134
resonance 21
RMS 13, 26
Rolling Element Bearings 68, 113
Root Mean Square 13
rotor 71, 107, 108, 120, 121, 122, 125
rotor bar 121, 122
rotor bow 122
S

screw type pump 127
Sensor Mounting Pads 65
Sheave Misalignment 129
sidebands 34, 35, 36, 115
signals 16, 19, 27, 28, 30, 32, 34, 35, 36, 37, 45
Simple Harmonic Motion 9
sine wave 10, 13, 15, 29, 36
sinusoid 19
slip 120, 122
147
slot pass frequency 121
spectrum 24, 25, 27, 29, 31, 32, 34, 35, 36, 40, 48, 62, 69, 71, 61, 63, 68, 125,
126, 132
Spectrum Comparison 68
Spectrum Mask 70
stationary 27, 28, 46
stator 62, 120, 121, 122
Synchronous Averaging 71, 73, 75, 78
Synchronous Motors 118
T

tachometer 14
ten-point divider 61, 64
time domain 24, 26, 29
tooth-mesh 71
transducer 62, 69
trending 39, 67, 69, 115
trigger 14
Truncation 68
Turbines 22, 108, 125, 134
V

vane pass 71, 64, 133, 134
VdB 42, 44, 107, 114, 134, 147
velocity 15, 16
Velocity 44, 62, 147
vibration 9, 10, 13, 16, 17, 26, 28, 29, 31, 34, 35, 40, 45, 46, 47
Vibration Severity Chart 148
vibration spectrum 62, 69, 71, 127
VTAG 61
W

wave form 11, 13, 15, 19, 26, 29, 31, 32, 35, 36, 43, 48, 68
waveform 65
whole body motion 9
Glossary of Terms

Acceleration
Accelerometer
Algorithm
Aliasing
Alignment
AM
Amplitude
Amplitude Modulation
Analog
Analog to Digital Conversion
Analysis Parameters
Angular Frequency
Anti-Aliasing Filter
148
Apodize, Apodization
Asynchronous
Attachment Pad/Block
Attenuation
Auto correlation
Averaging
Axial
Background Noise
Balancing
Ball Pass Frequency
Ball Spin Frequency
Band Pass Filter
Bandwidth
Barcode
Baseline Spectrum
Bearing Tones
Beat Frequency
Bin
Bit
Blade Pass Frequency
Block
Bode Plot
Bow
BPI, BPFI
BPO, BPFO
BS, BSF
Brinnelling
Broad band
Buffer
Bump Test, Impact Test
Calibration
Carrier Frequency
Cavitation
Center of Gravity
Centrifugal Force
Centripetal Force
Cepstrum
Charge Amplifier
Cluster
Coherence
Correction Weight
Coulomb Damping
Couple Imbalance
Crest Factor
Critical Damping
Critical Speed
Cross Correlation
Cycle
Damped Natural Frequency
Damping
dB
Decibel
Degree of Freedom
149
Demodulate, Demodulation
Detector
Deterministic
Differentiation
Digital
Discrete
Discrete Fourier Transform
Displacement
Displacement Transducer
Distortion
Domain
Dynamic Imbalance
Dynamic Range
Eccentricity
Eddy Current
Eddy Current Probe
Engineering Units, EU
EU
Excitation
Expert System
Fast Fourier Transform (FFT)
Fatigue
Fault Frequency
FEM
FFT
FFT Analyzer
Filter
Flattop Window
Fluid-Film Bearing
Forced Vibration
Forcing Frequencies
Foundation
Fourier, Jean Baptiste
Fourier Transform
Fourier Analysis
Free Running
Free Vibration
Frequency
Frequency Domain
Frequency Response
FT, FTF
Fundamental Frequency
Fundamental Train Frequency
G
Gear-Mesh Frequency
Ghost Frequency
GPIB
Ground Loop
Hanning Window
Hamming Window
Harmonics
Hertz
High-Pass Filter
150
HTF
Hunting Tooth
Hysteresis
Hz
ICP Accelerometer
Imbalance
Impact Test
Impedance, mechanical
Inertia
Integration
Integrator
Isolation
Jerk
Keyphasor
Kurtosis
Leakage
Level
Line Spectrum
Linear, Linearity
Low Pass Filter
Magnetostriction
Mask
Mechanical Impedance
Micrometer
MIL-STD-167-1
Mils
Mobility
Modal Analysis
Mode of Vibration
Mode Shape
Modulation
Narrow band Analysis
Natural Frequency
Node
Noise
Noise Floor
Non-Linear
Non-Linear Damping
Normal Mode of Vibration
Normalization
Nyquist frequency
Nyquist Plot
Octave
Oil Whip
Oil Whirl
Orbit
Orders
Order Analysis
Orthogonal
Oscillation
Overall Level
Overlap Processing
Peak
151
Peak-to-Peak (Pk-Pk) Value
Pendulum
Period
Periodic
Phase
Phase Angle
Phase Shift
Phasor
Picket Fence Effect
Pickup
Piezo-electric
Piezo-electric Transducer
Pink Noise
Power Factor
Power Spectral Density
Preload
Pressure Waves
Prime Mover
Principal Inertia Axis
Proximity Probe
PSD
Q
Quasi-Periodic
Radial
Radian
Random
Rectangular Window
Resolution Bias Error
Resonance
Resonant Frequency
Response Spectrum
Rigid Rotor
RMS
Roll Off, Rolloff
Running Speed
Runout
Scalar
Seismic
Selectivity
Sensitivity
Shock
Shock Pulse Meter
Shorting Ring
SI
Sidebands
Signal
Signal Conditioning
Signature
Single Degree of Freedom
Simple Harmonic Motion
Sine Wave
Sinusoid
Ski Slope
152
Acceleration
The time rate of change of velocity, usually measured in Gs in the English system of
measurements, and in meters per second per second (m/s
2
) in the SI system. It is
interesting to note that the G is not actually a unit of acceleration, but is the
magnitude of the acceleration due to gravity at the earth's surface. This causes some
undue complexity in converting parameters between acceleration, velocity, and
displacement. The value of G amounts to 32.2 feet per second per second.

S-ar putea să vă placă și