Sie sind auf Seite 1von 487

Machine Condition Monitoring

and
Fault Diagnostics

Chris K. Mechefske

Queens University

The focus of this book is on the definition and description of machine condition monitoring and
fault diagnosis. Included are the reasons and justification behind the adoption of any of the
techniques presented. The motivation behind the decision making in regard to various
applications is both financial and technical. Both of these aspects are discussed, with the
emphasis being on the technical side. The book defines machinery failure (causes, types, and
frequency), and describes basic maintenance strategies and the factors that should be considered
when deciding which to apply in a given situation. Topics considered in detail include transducer
selection and mounting location, recording and analysis instrumentation, display formats and
analysis tools (time domain, frequency domain, modal domain, and quefrency domain based
strategies). The discussion of fault detection is based primarily on standards and acceptance
limits in the time and frequency domains. The topics of machine testing and fault trouble
shooting are also addressed. The discussion of fault diagnostics is divided into sections that focus
on different forcing functions, specific machine components, specific machine types, and
advanced diagnostic techniques. While the focus of the book is on vibration based techniques,
some information related to alternative techniques is included.

Copyright Chris K. Mechefske, 2012


Contents
1. Introduction

2. Machinery Failure

2.1 Causes of Failure


2.2 Types of Failure
2.3 Frequency of Failure
2.4 The History of Maintenance Expectations and Methods

3. Basic Maintenance Strategies

3.1 Reactive (Run-to-Failure, Breakdown) Maintenance


3.2 Scheduled (Preventive) Maintenance
3.3 Condition-Based (Predictive, On-Condition) Maintenance
3.4 Factors which Influence Maintenance Strategy

4. Machine Condition Monitoring

4.1 Periodic Monitoring


4.1.1 Listing and Categorization of Machinery
4.1.2 Machinery Knowledge
4.1.3 Route Selection and Definition
4.1.4 Measurement Parameters and Points
4.1.5 Baseline Data
4.1.6 Frequency of Data Collection
4.1.7 Selection of Test Equipment
4.1.8 Screening
4.1.9 Trending
4.1.10 Alarms
4.1.11 Reports
4.1.12 Fault Diagnosis
4.2 Continuous Monitoring

5. Basic Vibrations

5.1 Classification of Vibration Signals by Type of Motion


5.1.1 Simple Harmonic Motion
5.1.2 Periodic Motion
5.1.3 Random Motion
5.1.4 Transient Motion
5.2 Classification of Vibration Signals by Type of Excitation
5.2.1 Free Vibration
5.2.2 Forced Vibration
5.2.3 Self-induced Vibration

ii
5.3 Wave Fundamentals (Amplitude, Frequency, Phase)
5.4 Basic Spring-Mass Model of Vibration
5.5 Basic Spring-Mass-Damper Model of Vibration
5.5.1 Example: Car Model
5.6 Vibration Signal Fundamentals (Displacement, Velocity and Acceleration)
5.7 Vibration Descriptors
5.7.1 Average Vibration Signal Strength
5.7.2 Peak Amplitude
6.1.3 Peak-to-Peak Amplitude
6.1.4 Root Mean Square (RMS)
6.1.5 Crest Factor
5.8 Frequency Response Functions and Natural Frequencies
5.9 Time Waveform versus Frequency Spectra Analysis
5.10 Overall Levels versus Frequency Spectra Analysis

6 Basic Vibration Signal Analysis

6.1 Time Domain Signal Analysis


6.1.1 Time Waveform Analysis
6.1.2 Time Domain Indices
6.1.3 Statistical Methods and Parameters
Probability Density Functions
Probability Density Moments
Crest Factor versus Kurtosis
6.2 Frequency Domain Signal Analysis
6.2.1 Band Pass Analysis
6.2.2 Frequency Domain Indices
6.2.3 Signature Spectral Analysis
Decibel (dB) Units
Windowing to Prevent Leakage
Overlapping and Zero Padding
6.2.4 Cascades (Waterfalls Plots)
Natural Frequencies
6.3 Modal Domain Signal Analysis
6.3.1 Mode Shapes
6.3.2 Deflection Shape Analysis
6.3.3 Operational Deflection Shape Analysis compared to Mode Shape Analysis
6.4 Quefrency Domain Signal Analysis

7. Transducer Selection

7.1 Selecting a Measurement Parameter


7.2 Noncontact Displacement Transducers (Proximity (Eddy Current) Probes)
7.2.1 Design and Operation
7.2.2 Orbits
7.2.3 Calibration

iii
7.2.4 Mounting and Location
7.2.5 Non-contact Displacement Sensors as triggers and Speed Sensors
7.2.6 Advantages and Disadvantages
7.3 Laser Based Non-Contact Displacement Sensors
7.4 Velocity Transducers (Electro-Mechanical)
7.4.1 Design and Use
7.4.2 Advantages and Disadvantages
7.5 Acceleration Transducers
7.5.1 Principles of Operation
7.5.2 Sensitivity versus Mass
7.5.3 Design and Operation
7.5.4 Calibration
7.5.5 Mounting and Location
7.5.6 Cables and Ground Loops
7.5.7 Advantages and Disadvantages

8. Signal Collection and Analysis Instrumentation and Analysis Methods

8.1 Basic Vibration Meters


8.1.1 Oscilloscopes
8.1.2 Overall Vibration Level Meters
8.1.3 Shock Pulse Meter
8.2 Data Collectors
8.3 Frequency Domain FFT Analyzers
8.3.1 Order Tracking Analyzers
8.3.2 Selecting the Correct Analyzer
8.4 Data Recorders
8.5 Vibration Signal Sampling Alternatives
8.5.1 Uniform Time Sampling
8.5.2 Uniform Position Sampling
8.6 Data Sampling Rates and Resolution Issues
8.7 Aliasing
8.7.1 Anti-Aliasing Filters
8.8 Time Synchronous Averaging
8.8.1 Negative Averaging
8.9 Orbits
8.10 Enveloped (Demodulated) Spectra
8.11 Modal Analysis (revisited)

9. Fault Detection

9.1 General Standards and Guidelines


9.1.1 Standards Based on Vibration Severity
9.1.2 Standards Based on Machinery Type
9.1.3 Standards based on Statistical Limits
9.2 Acceptance Limits

iv
9.3 ISO Standard on Training and Certification in Machine Condition Monitoring
9.4 Frequency Domain Limits
9.4.1 Limited Band Monitoring
9.4.2 Constant Bandwidth Limits
9.4.3 Constant Percentage Bandwidth Limits

10 Fault Diagnostics

10.1 Overview of Machine Fault Diagnostics


10.1.1 Frequency Spectra Analysis
10.1.2 Synchronous, Sub-Synchronous and Non-Synchronous Responses
10.2 Fault Diagnosis based on Forcing Functions
10.2.1 Resonance
10.2.2 Unbalance
Definition of Unbalance
Causes of Unbalance
Unbalance Correction Methods
Types of Unbalance
Static Unbalance
Couple (Moment) Unbalance
Quasi-Static Unbalance
Dynamic Unbalance
Rotor Motions
Runout versus Unbalance
The Effect of Rotational Speed (on Unbalance Forces)
Single Plane Balancing (general)
Two-Plane (Dynamic) Balancing
Field (In-Situ) Balancing
Sensors for Measuring Unbalance
The Effect of Rotational Speed (on the Phase of Measurements)
Single Plane Balancing (details)
The Vector Method of Single Plane Balancing
Summary of Unbalance
10.2.3 Bent Shafts and Bowed Rotors
10.2.4 Misalignment
Types of Misalignment
Parallel Misalignment
Angular Misalignment
Bearing Misalignment
Alignment Methods
Reverse Dial Alignment Method
Face and Rim Alignment Method
Alignment Procedures
Soft Feet
Sag in Alignment Rods
Thermal Growth

v
Vibration Caused by Misalignment
10.2.5 Distinguishing between Unbalance and Misalignment
10.2.6 Mechanical Looseness
10.2.7 Rubs
10.2.8 Oil Whirl and Oil Whip
10.2.9 Beating and Amplitude Modulation
10.2.10 Structural Vibrations
10.2.11 Foundation Problems
10.2.12 Summary of Identification and Correction of Forcing Function Faults
10.3 Faults Diagnosis Based on Specific Machine Components
10.3.1 Rolling Element Bearings
10.3.2 Gears and Gearboxes
10.3.3 Belts
10.4 Faults Diagnosis Based on Specific Machine Types
10.4.1 Electric Motors
10.4.2 Fans
10.4.3 Pumps
10.4.4 Compressors
10.4.5 Steam and Gas Turbines
10.4.6 Reciprocating Machines

11 Machine Testing

11.1 Test Plans


11.2 Selection of Test Equipment
11.3 Site Inspection
11.4 Acceptance Tests
11.5 Baseline Tests
11.6 Resonance and Critical Speed Testing
11.7 Specifications
11.8 Environment and Mounting
11.9 Presentation of Data
11.10 Reports

12 Trouble Shooting

12.1 General Steps


12.2 Identify the Problem
12.3 Gather Information
12.4 Determine Possible Forcing Functions
12.5 Determine Where to Take Data and What Data Collection Equipment to Use
12.6 Take Vibration Data
12.7 Analyze Vibration Data (and any other data available)
12.8 Make Recommendations

13 Advanced Methods of Machine Condition Monitoring

vi
13.1 Automated Machine Condition Monitoring
13.1.1 Expert Systems
13.1.2 Fuzzy Logic based Signal Analysis
13.1.3 Artificial Neural Networks
13.1.4 Support Vector Machines
13.1.5 Novelty Detection
13.1.6 Automated Machine Condition Monitoring A Case Study
13.2 Model based Spectral Estimation
13.3 Minimum Variance based Spectral Estimation
13.4 Time-Frequency Analysis
13.5 Short Time Fourier Transforms (STFT)
13.6 Wigner-Ville Distributions
13.7 Wavelets
13.8 Independent Component Analysis

14 Non-Vibration based Techniques

14.1 Visual Monitoring


14.2 Performance Monitoring
14.3 Temperature Monitoring
14.3.1 Thermography (thermal mapping)
14.4 Acoustic Emission (AE)
14.5 Oil Quality Analysis
14.5.1 Lubrication Fundamentals
14.5.2 Base Stock Considerations
14.5.3 Mineral versus Synthetic Oils
14.5.4 Hydrostatic, Elasto-Hydrodynamic and Boundary Lubrication
14.5.5 Additives and Their Functions
14.5.6 Oil Quality Parameters
14.5.7 Water in Oil
14.5.8 Fuel in Oil
14.5.9 Soot in Oil
14.5.10 Glycol in Oil
14.6 Oil Sampling
14.6.1 General Guidelines for Drawing Samples
14.6.2 General Safety Considerations
14.6.3 Sources of Error in Data
14.6.4 Recommended Oil Sampling Frequencies
14.6.5 Information to be Supplied with Oil Sample
14.6.6 Drain Port Sampling
14.6.7 Drain Port Vacuum Sampling
14.6.8 Drain Line Tap Sampling
14.6.9 Drain Line Vacuum Sampling
14.6.10 Portable Off-Line Sampling
14.6.11 Dedicated Off-Line Sampling

vii
14.6.12 Probe-On Vacuum Sampling
14.6.13 Drop-Tube Vacuum Sampling
14.6.14 Pressurized Line Sampling Low Pressure Tap Sampling
14.6.15 Pressurized Line Sampling Portable Tap Sampling
14.6.16 Pressurized Line Sampling Low Pressure Ball Valve Sampling
14.6.17 Pressurized Line Sampling High Pressure Sampling
14.7 Wear Particle Analysis
14.7.1 Wear Particle Generation (Wear Particle Types)
14.7.2Controlling Particle Contamination Filters
14.7.3 Origins of Particles
14.7.4 Trending using Wear Particles
14.7.5 Collection Techniques
14.7.6 Particle Counting
14.7.7 Visual Analysis Methods - Analytical Ferrography
14.7.8 Spectrametric Oil Analysis

viii
Chapter 1

Introduction
Approximately half of all operating costs in most processing and manufacturing operations can
be attributed to maintenance. This is ample motivation for studying any activity that can
potentially lower these costs. Machine condition monitoring and fault diagnostics is one of these
activities. Machine condition monitoring and fault diagnostics can be defined as the field of
technical activity in which selected physical parameters, associated with machinery operation,
are observed for the purpose of determining machinery integrity. Once the integrity of a machine
has been estimated, this information can be used for many different purposes. Machine load
capacity and maintenance activity timing are the two main tasks that link directly to the
information provided. The ultimate goal in regard to maintenance activities is to schedule only
what is needed at any given time, which results in optimum use of resources. It should also be
noted that condition monitoring and fault diagnostic practices may also be applied to improve
end product quality control and as such can also be considered as process monitoring tools.

Ideally, machine condition monitoring and fault diagnostics allows for the accurate estimation of
current machine condition as well as remaining machine useful life. However, failures do occur.
In many instances these failures are not only costly (due to loss of the physical asset and lost
production while the machine is repaired or a replacement is brought on line), but may be
dangerous to personnel and/or the environment. Figures 1 and 2 show some examples of
relatively large scale failures. Figure 3 shows a smaller scale failure that could be just as costly
depending on the process that the bearing was supporting.

Figure 1.1 Examples of large scale turbine shaft failures.

1
Figure 1.2 Examples of large scale gearbox failures.

Figure 1.3 Examples of small scale rolling element bearing failures.

2
The definition stated above implies that, while machine condition monitoring and fault
diagnostics is being treated as the focus here, it must also be considered in the broader context of
plant operations. With this in mind, it is appropriate to begin with a description of what is meant
by machinery failure and a brief overview of different maintenance strategies and the various
tasks associated with each. A short description of different vibration sensors, their modes of
operation, selection criteria, and placement for the purposes of measuring accurate vibration
signals will then follow. Data collection and display formats will be discussed with the specific
focus being on standards common in condition monitoring and fault diagnostics. Machine fault
detection and diagnostic practices will make up the remainder of this book. The progression of
information provided will be from general to specific. The hope is that this will allow a broad
range of individuals to make effective use of the information provided.

Before going further it should be noted that the focus of this book is on vibration based
measurements and dynamic signal analysis. The reasons for this focus will be highlighted in
Chapter 5. However, Chapter 14 contains descriptions of some common non-vibration based
machine condition and fault diagnostic techniques. These techniques include (but are not limited
to) measurement and analysis of the following parameters:

oil quality (including contamination)


wear particles
force
sound pressure (intensity)
temperature
output (machine performance)
product quality
odour
visual inspection

3
Chapter 2

Machinery Failure
Most machinery is required to operate within a relatively close set of limits. These limits, or
operating conditions, are designed to allow for safe operation of the equipment and to ensure
equipment or system design specifications are not exceeded. They are usually set to optimize
product quality and throughput (load) without overstressing the equipment. Generally speaking,
this means that the equipment will operate within a particular range of operating speeds. This
definition includes both steady-state operation (constant speed) and variable speed machines,
which may move within a broader range of operation but still have fixed limits based on design
constraints. Occasionally, machinery is required to operate outside these limits for short times
(during start-up, shutdown, and planned overloads).

The main reason for employing machine condition monitoring and fault diagnostics is to
generate accurate, quantitative information on the present condition of the machinery. This
enables more confident and realistic expectations regarding machine performance. Having at
hand this type of reliable information allows for the following questions to be answered with
confidence:

Will a machine stand a required overload?


Should equipment be removed from service for maintenance now or later?
What maintenance activities (if any) are required?
What is the expected time to failure?
What is the expected failure mode?

Machinery failure can be defined as the inability of a machine to perform its required function.
Failure is always machinery specific. For example, the bearings in a conveyor belt support pulley
may be severely damaged or worn, but as long as the bearings are not seized, it may be
considered as having not failed. One the other hand, depending on the critical nature of the
machinery involved, a severely worn bearing may not be tolerable due to the short lead time
before complete failure (bearing secure) or the increased energy required to roll the bearing or to
the noise that is generated or reduced product quality. Different machinery and different
applications involving the same machinery may not tolerate the same operating conditions or the
same degree of component wear before being deemed to have failed. A computer disk drive with
only a very slight amount of wear or misalignment or looseness resulting in noisy operation
(while all other performance criteria are still being met) may still be considered as having failed
if the required operation of that device mandates quiet operation.

It should also be noted that the failure of an entire system may be the result of the failure of a
sub-system or individual component. As an example consider the tires on a car. A flat tire is the
failure of a component (the tire) which is part of a sub-system (the suspension) of the entire car.
A flat tire may render a car no longer drivable even though all the other systems are functioning
normally. This example considers a car with standard tires (not drive-flat tires). This

4
example can also be used to point out that in an extreme case a car can still be driven (slowly and
for a relatively short distance) on a flat tire. There are serious consequences to this action (the
potential destruction of an otherwise repairable tire), but it may be considered a reasonable
option (after weighing all the alternatives and costs involved).

There are also other physical considerations that may suggest that a machine is no longer
performing adequately. Many of the most common situations and machine/component types will
be discussed fully in this book. Also, purely economic considerations may result in a machine
being classified as obsolete and it may then be scheduled for replacement before it has worn
out. Safety considerations may also require the replacement of parts in order to ensure the risk
of unexpected failure is minimized. This is particularly true in situations involving the transport
of humans (planes, trains and automobiles).

2.1 Causes of Failure

When we disregard the gradual wear of machinery as a cause of failure, there are still many other
specific causes. These are perhaps as numerous as the different types of machines. There are,
however, some generic categories that are listed and described here.

Deficiencies in the original design may result in a machine that is inherently under-designed for
a particular purpose or expected load. Defective materials or poor material processing may result
in a machine that is weaker than expected due to flaws in the original material, such as poor
quality steel that is not up to the standards specified (lower strength, low resistance to corrosion,
etc.). Improper assembly either of the machine or improper commissioning of large installations
may result in situations where a machine is not capable of meeting the expected design
requirements. In general the above causes of failure typically result in failure or noticeably poor
performance early in the useful life of machines (during the testing or commissioning stages).

Inappropriate maintenance includes too little, too much and of course the wrong type of
maintenance activity. An example of a common inappropriate maintenance activity is performing
the wrong type of lubrication. This includes using the wrong type of lubricant, using too much or
too little lubricant as well as lubricating too often or too seldom. This also includes improper re-
assembly following maintenance (see above). Excessive or inappropriate operational demands
may also cause premature failure. The actual failure mechanisms involved in cases of excessive
or inappropriate operation are the same as those that would result after a longer period of normal
use. The failures just occur sooner in the useful life of the machine.

2.2 Types of Failure

As with the causes of failure, there are many different types of actual failure. At this stage, these
types will be subdivided into only two broad categories - catastrophic and incipient. Catastrophic
failures are sudden and complete. Incipient failures are partial and usually gradual. In all but a
few instances, there is some advanced warning as to the onset of failure. That is, the vast
majority of failures pass through a distinct and relatively lengthy incipient phase. The goal of
machine condition monitoring and fault diagnostics is to detect this early onset of failure,
diagnose the condition, and trend its progression over time. The time until ultimate failure can

5
then hopefully be better estimated, and this will allow plans to be made to avoid catastrophic
repercussions. These plans typically include the scheduling of some form of maintenance (repair
or replacement) of the machine, sub-system or component(s). This, of course, excludes failures
caused by unforeseen and uncontrollable outside forces such as earthquakes.

(a) (b)

Figure 2.1 a) A small crack in a car windshield - Incipient Failure, b) The crack has passed
through the incipient phase and the result is total loss of function - Catastrophic Failure.

Figure 2.2 Incipient failures detected early and then corrected lead to avoidance (or at least
delay) of catastrophic failures (and the associated costs).

2.3 Frequency of Failure

Anecdotal and statistical data describing the frequency of failures can be summarized in what is
called a bathtub curve. Figure 2.3 shows a typical bathtub curve. These curves are applicable
to an individual machine or population of machines of the same type being used in the same way.
The beginning of a machines useful life is usually characterized by a relatively high rate of
failure. These failures are referred to as wear in failures. They are typically due to such things

6
as design errors, manufacturing defects, assembly mistakes, installation problems and
commissioning errors. As the causes of these failures are found and corrected, the frequency of
failure decreases.

The machine then passes into a relatively long period of operation, during which the frequency
of failures occurring is relatively low. The failures that do occur mainly happen on a random
basis. This period of a machines life is called the normal wear period and usually makes up
most of the life of a machine. There should be a relatively low failure rate during the normal
wear period when operating within design specifications.

Wear In Normal Wear Wear Out


(much longer than
Failure wear in and wear
Rate or out phases)
Likelihood
of Failure

Time in Service

Figure 2.3 A typical bathtub curve for an individual machine or population of machines.

As a machine gradually reaches the end of its designed useful life, the frequency of failures again
increases. These failures are called wear out failures. This gradually increasing failure rate at
the expected end of a machines useful life is primarily due to metal fatigue, wear mechanisms
between moving parts, corrosion, and/or obsolescence. The slope of the wear out part of the
bathtub curve is machine and load dependent. The rate at which the frequency of failures
increases is largely dependent on the design of the machine and its operational history. If the
bathtub curve increases in slope sharply within the wear out section of the graph, the machine is
likely under-designed to meet the load expected or the machine has endured a severe operational
life (experienced numerous overloads). If the machinery is over-designed or experiences a
relatively light loading history, the slope of this part of the bathtub curve will increase only
gradually with time (see Figure 2.4).

While the effects of load are dominant in the wear out stage of the bathtub curve, it is more
accurate to expect these effects to be somewhat obvious throughout the entire useful life of a
machine (as shown in Figure 2.5).

7
Wear In
Normal Wear Wear Out
Failure
Rate or
Likelihood (a)
of Failure
(b)

Time In Service

Figure 2.4 A typical bathtub curve showing the typically observed results (predominantly in the
wear out stage) of different operational conditions over the useful life of the machine a) under
designed for purpose and/or frequently overloaded b) over designed for purpose and/or
frequently lightly loaded.

Wear In Normal Wear Wear Out

Increasing Duty
Failure
Rate

Time in Service

Figure 2.5 A typical bathtub curve showing the true overall effects of different operational
conditions over the useful life of the machine.

2.4 The History of Maintenance Expectations and Methods

Over the course of the last 70 years there has been an evolving set of expectations from
maintenance activities in terms of what could be done and what the expected outcomes could
achieve. Figure 2.6 shows a summary of how these expectations have changed over that time.
These changes can be group into three broad categories. Before the end of World War II
maintenance (apart from regular lubrication) was largely conducted only when a machine or
component actually broke. Failure detection was typically limited to when the machine could no
longer function. That is, only after catastrophic failure. Only in rare cases would a failure be
detected before ultimate failure, usually by chance or by direct visual inspection. This
maintenance method served the machines of that time relatively well as they were typically over-
designed to meet the required function and were generally relatively simple in design.

8
Figure 2.6 Changing maintenance expectations as a function of time.

During World War II it was observed that more complex machines (aircraft engines) often had
an increased failure rate just after a maintenance function that required some amount of
disassembly and then reassembly. Tighter control needed to be placed on the maintenance
activities in order to achieve the required quality. These practices were quickly adopted by
industry in the following decades in order to improve plant availability, increase equipment
useful life and reduce costs.

In the last three or four decades there has been a heightened desire to further reduce maintenance
costs because of global industrial competition. This, coupled with the concurrent development of
improved sensor technology and faster computer chips, has lead to the development of a third
generation of maintenance expectations. Maintenance decisions and activities are now driven by
the same motivations as previously, but also include a need for increased equipment reliability,
improved safety, and reduced risk to the environment.

As new monitoring and diagnostic methods and tools are developed, maintenance expectations
will surely continue to evolve. Figure 2.7 shows an historical perspective of how the classical
emphasis on overhauls and administrative systems to control maintenance activities has changed
and now includes many new developments in a number of different fields.

At the same time as the changes in maintenance expectations and methods has taken place, there
has also been a gradual change in our understanding of equipment failure rates. Figure 2.8 shows
this change as a function of time. Early machines were considered to be relatively reliable until
they reached the end of their useful life when the failure rate increased (pre 1950). Following the
observations made on aircraft engines during World War II the traditional bathtub curve was
recognized as a better predictor of failure rates during the useful life of machinery. As mentioned
above, this was partly due to the rapid increase in the mechanical complexity of machines at that
time.

9
Figure 2.7 Changing maintenance methods as a function of time.

Currently there are a host of curves that can be used to describe equipment failure rates.
Mechanical components will still typically follow the bathtub curve distribution. However, other
types of equipment need to be described differently. In Figure 2.8 the upper three curves on the
right hand side can be used to show the failure rates of mechanical components and systems. The
degree to which the component is well designed and/or pre-use tested dictates the shape of the
wear in part of the failure rate curve. The wear out section is still dictated by the factors
described above. The bottom three curves on the right hand side of Figure 2.8 are better used to
describe the failure rates of electronic equipment. These components are typically not age or
usage related. Once they pass the initial testing phase they will not wear out. Failures are
random or related to other factors.

It is worthwhile pointing out again at this stage that these curves may be considered descriptive
of the expected failure rate of a population of machines or the expected likelihood of failure of
an individual machine.

Figure 2.8 Changing views of expected equipment failure rates as a function of time.

10
Chapter Summary

Generally (outside of start-up and shutdown) machinery is required to operate at a design


specific constant speed and load.
Machinery failure is defined based on performance, operating conditions, and system
specifications.
Machinery failure can be defined as the inability of a machine to perform its required
function.
Causes of machinery failure can be generally defined as being due to deficiencies in the
original design, material or processing, improper assembly, inappropriate maintenance, or
excessive operational demands.
The frequency of failure for an individual machine or a population of similar machines can
be summarized using a bathtub curve.
Our understanding of equipment failure rates has evolved over time as the design and use of
engineering equipment has changed (and will continue to change).

11
Chapter 3

Basic Maintenance Strategies


The three most commonly applied maintenance strategies are shown in Figure 3.1. Reactive
maintenance (also know as run-to-failure or breakdown maintenance) is the simplest, requiring
no initial investment in special training or equipment, but is also the most limited, being largely
unable to prevent costly catastrophic failures. Scheduled maintenance (also known as preventive
maintenance) is when standard maintenance activities are conducted at regular intervals. In this
way catastrophic failures are usually prevented, but there is typically also the added cost of doing
maintenance work prematurely (with some machinery useful life still remaining). Condition
based maintenance (also know as on-condition or predictive maintenance) is the practice of
regularly or continuously monitoring some aspect of the machine under consideration in order to
detect the earliest stages of incipient failure. The correct maintenance activities can then be
planned at the appropriate time. This strategy results in maintenance being conducted only when
needed, but requires the effective collection and analysis of appropriate measures related to
machine condition.

Philosophy: Fix it when it breaks

Benefit: Cost:
Reactive
Zero initial investment Costly catastrophic
breakdowns

Philosophy: Change it out every ____ hours

Benefit: Cost:
Scheduled
Reduced catastrophic Premature work
breakdowns

Philosophy: Does it need to be fixed?

Condition Benefit: Cost:


Based Maintenance done when Requires effective use of
needed information

Figure 3.1 Three common basic maintenance strategies and some associated advantages and
costs.

3.1 Reactive (Run-to-Failure, Breakdown) Maintenance

Reactive (run-to-failure, breakdown) maintenance is a strategy where maintenance, in the form


of repair work or replacement, is only performed when machinery has failed. Reactive
maintenance, while seemingly out of date, is still appropriate when the following situations exist:

12
The equipment under consideration is redundant. That is, the equipment is running in parallel
with at least one other machine of the same type or there is a backup machine ready to take
the load at short notice.
Low cost spares are available. This may be spare components for repair of individual
machines and/or spare machines for replacement of failed machines.
The process is interruptible without additional cost or there is stockpiled product.
All known failure modes are safe.
There is a known long mean time to failure (MTTF) or a long mean time between failures
(MTBF). That is, under normal circumstances the machine (on average) is expected to
operate without failure for a relatively long period of time.
There is a low cost associated with secondary damage. That is, any resulting damage that
may occur due to the failure of the machine under consideration is minimal.
Quick repair or replacement is possible. That is, in addition to there being spare parts
available for repair work and/or a spare machine available for replacement of the failed
machine, there are also enough appropriately trained maintenance staff available at short
notice and the cost of these individuals is reasonable.

An example of the application of reactive maintenance can be found when one considers the
standard household light bulb. There are often more than one light bulbs lighting a given area
(redundancy). Low cost spares can be kept on hand or can be acquired relatively quickly and
easily. The failure mode is safe. Light bulbs typically last a relatively long time, so failures are
infrequent. There is typically no secondary damage when a light bulb burns out. Quick
replacement is possible by almost anyone. This device satisfies all the requirements above and
therefore the most cost-effective maintenance strategy is to replace burnt out light bulbs as
needed.

Figure 3.2 shows a schematic diagram demonstrating the relationship between a machines time
in service, the load (or duty) placed on the machine, and the estimated remaining capacity of the
machine. Whenever the estimated capacity curve intersects with (or drops below) the load curve,
a failure will occur. At these times, repair work must be carried out. If the situation that exists
fits within the rules outlined above, all related costs (repair work and downtime) will be
minimized when using reactive maintenance. Note that the load on the machine affects the slope
of the estimated machine capacity line, indicating that the higher the load the faster a machines
useful life can be considered as being used up. In this figure (and the two following) it is
assumed that the maintenance activity renews the machine to its previous full capability. This
may not always be the case. It is also assumed that there can be no loading of the machine (and
therefore no production) during the time spent doing maintenance work. Again, this may not
always be the case in reality.

13
Machine Capacity (Est.) Failures

Estimated
Capacity
(blue) and
Load (red)

Machine Duty (Load)

Time in Service Maintenance

Figure 3.2 Time versus estimated capacity and actual load (reactive maintenance).

3.2 Scheduled (Preventive) Maintenance

When specific maintenance tasks are performed at set time intervals (or duty cycles) in order to
maintain a significant margin between machine capacity and actual duty, the type of maintenance
is called scheduled or preventive maintenance. Scheduled maintenance is most effective under
the following circumstances:

Data describing the statistical failure rate for the machinery is available. That is, there is a
long history (with the company or available from the equipment manufacturer) that shows
the statistics of failures for the machine under known loads. From these statistics an accurate
estimate of the failure rate can be determined.
The failure distribution is narrow. This means that the MTBF is accurately predictable.
Maintenance restores full (or close to full) integrity of the machine so that the maintained
machine can be expected to achieve the same performance as a new machine (including
lasting the same period of time before failure is expected).
A single, known failure mode dominates. That is, the machine or component can be expected
to fail in the same manner each time it fails.
There is low cost associated with regular overhaul/replacement of the equipment. This refers
to the cost of the labour required to conduct the maintenance work as well as consumables
that may be required for each repair/replacement.
Unexpected interruptions to production are expensive, but scheduled interruptions are not so
bad. An example of a process that may be less costly to interrupt on a scheduled basis would
be a plastic (say nylon) processing plant. An unexpected interruption of the process could
result in piping becoming filled with solid product. Recovery from such an interruption

14
would be much more costly then recovery from a scheduled interruption when all the pipes
could be drained prior to shutting down the pumps.
Low cost spares are available.
Costly secondary damage from failure is likely to occur. This refers to the situation where a
failure itself may cause subsequent damage to machinery or product upstream, downstream
or nearby the failed equipment.

An example of scheduled maintenance practices can be found under the hood of your car. Oil
and oil filter changes on a regular basis are part of the scheduled maintenance program that most
car owners practice. A relatively small investment in time and money on a regular basis acts to
reduce (but not eliminate) the likelihood of a major failure taking place. Again, this example
shows how when all, or most, of the criteria listed above are satisfied, overall maintenance costs
are minimized over the useful life of the system.

Figure 3.3 shows a schematic diagram demonstrating the relationship between a machines time
in-service, the load (or duty) placed on the machine and the estimated remaining capacity of the
machine when scheduled maintenance is being practiced. In this case, maintenance activities are
scheduled at regular intervals in order to restore machine capacity before a failure occurs. In this
way, there is always a margin between the estimated capacity and the actual load on the machine.
If this margin is always present, there should theoretically never be an unexpected failure, which
is the ultimate goal of scheduled maintenance.

Machine Capacity
(Est.)

Estimated
Capacity
and Load Margin
Margin {

Machine Duty (Load)


Time In Service Maintenance

Figure 3.3 Time versus estimated capacity and actual load (scheduled maintenance).

3.3 Condition Based (Predictive, On-Condition) Maintenance

Condition based maintenance requires that some means of assessing the actual condition of the
machinery is used in order to optimally schedule maintenance, in order to achieve maximum
production, and still avoid unexpected catastrophic failures. Condition based maintenance should
be employed when the following conditions apply:

15
Expensive or critical machinery is under consideration. Obviously, machinery that is critical
to production and machinery that is particularly expensive are good candidates for condition
based maintenance because of the high cost of lost production from a critical machine and
the high cost of replacement.
There is a long lead-time for replacement parts (no spares are readily available). This is often
the case with extremely large or complex machines. However, it may also be the case in
situations where local or global demand temporarily outstrips supply. Such was the case
during the recent construction and mining boom when large tires (those used on very large
haul trucks and loaders) became difficult to acquire. This was due to the sharp increase in
demand that came about relative quickly. The principle manufacturer of these tires
(Firestone) could not increase production capacity quickly enough to meet that demand.
The process is uninterruptible (both scheduled and unexpected interruptions are excessively
costly). This may be due to the process itself or simply due to lost sales or penalties payable
to clients because of missed shipments.
Equipment overhaul is expensive and requires highly trained (and expensive) people.
Highly skilled maintenance people are unavailable.
The costs of the monitoring program (acquisition and running) are acceptable.
Failures may be dangerous.
The equipment is remote or mobile (which can significantly increase the cost of repair work).
Failures are not indicated by degeneration of normal operating response. That is, regular or
standard monitoring of the machine or process performance is not enough to indicate a
failure or imminent failure.
Secondary damage may be costly.

An example of condition based maintenance practices can again be found when considering your
car, but this time we consider the tires. Reactive maintenance is not appropriate for tires because
of the potentially dangerous circumstances that may surround a flat tire. Schedule maintenance
(replacement) is also not appropriate because not all tires on all cars wear at the same rate. Tire
wear rates depend largely on driver habits and road conditions. Regular inspections of the tires
(air pressure checks, looking for cracks and scratches, measuring the remaining tread, listening
for slippage during cornering) can all be used to make an assessment of the remaining life of the
tires and also the risk of catastrophic failure. In order to minimize costs and risk, the tires can be
replaced before they are worn out completely, but not before they have given up the majority of
their useful life. A measure of the actual condition of equipment is required to allow utilization
of maintenance resources optimally.

Figure 3.4 shows a schematic drawing that demonstrates the relationship between a machines
time in service, the load (or duty) placed on the machine, and the estimated remaining capacity
of the machine when condition based maintenance is being practiced. Note that the margin
between duty and capacity is allowed to become quite small (smaller than for scheduled
maintenance), but the two lines never touch (as in reactive maintenance). This results in a longer
time between maintenance activities than for scheduled maintenance. Maintenance tasks are

16
scheduled just before a failure is expected to occur, thereby optimizing the use of resources. This
requires that there exists a set of accurate measures that can be used to assess the machine
integrity. Note again that a change in load will change the slope of the machine capacity curve
(reducing load typically extends machine useful life). This technique can be used to fully
optimize maintenance activities by providing flexibility in the timing of these activities.

Machine Capacity
(Est.)
Reduced Load
Estimated
Capacity
and Load Minimum Margin

Machine Duty (Load)


Time in Service Maintenance

Figure 3.3 Time versus estimated capacity and actual load (condition based maintenance).

Each of these maintenance strategies has its advantages and disadvantages and situations exist
where one or the other is appropriate. It is the maintenance engineers role to decide on and
justify the use of any one of these procedures for a given machine. There are also instances
where a given machine will require more than one maintenance strategy during its operational
life, or perhaps even at one time, and situations where more than one strategy is appropriate
within a particular plant. Examples of these situations include; i) the need for an increased
frequency of monitoring as the age of a machine increases and the likelihood of failure increases,
and ii) the scheduling of overhauls at the maximum time interval during the early stages of a
machines useful life, with monitoring in between looking for unexpected failures (until the
typical failure rates and modes are discovered.

Finally, there are still other maintenance strategies that are practiced, but these typically fall near
the ones described above, but have been developed for application in particular industries. For
example, Reliability Centered Maintenance (RCM) was originally developed in the airline
industry in order to carefully control the scheduled maintenance activities that are applied to
aircraft. RCM is now applied across a broad range of industries and involves a careful review of
risks and practices in order to set the optimum schedule for maintenance activities. Total
Productive Maintenance (TPM) was developed in the manufacturing industry and links the
maintenance activities to production in order to minimize costs, again when considering the
schedule of maintenance tasks. Proactive maintenance is the term sometimes used to describe
efforts that are made to investigate the root cause of failures and then redesign machines and/or
processes to eliminate these failures.

3.4 Factors which Influence Maintenance Strategy

17
While there are some general guidelines for choosing the most appropriate maintenance strategy,
each case must be evaluated individually. Principal considerations will always be defined in
economic terms. Sometimes, a specific company policy (such as safety or the desire to be an
industry technology leader) will outweigh all other considerations. Below is a list of factors (in
no particular order) that should be taken into account when deciding which maintenance strategy
is most appropriate for a given situation or machine:

Classification (size, type) of the machine.


Critical nature of the machine relative to production.
Cost of replacement of the entire machine.
Lead-time for replacement of the entire machine.
Manufacturers recommendations.
Failure data (history), MTTF, MTBF, failure modes.
Redundancy.
Safety (plant personnel, community, environment).
Cost and availability of spare parts.
Personnel costs, administrative costs, monitoring equipment costs.
Running costs for a monitoring program (if used).

Chapter Summary

Maintenance strategies can be divided into three main types: 1) reactive, 2) scheduled, and 3)
condition based maintenance.
No one strategy should be considered as always superior or inferior to another.
Reactive, or breakdown maintenance, is a strategy where maintenance, in the form of repair
work or replacement, is only performed when machinery has failed.
When specific maintenance tasks are performed at set time intervals (or duty cycles) in order
to maintain a significant margin between machine capacity and actual duty, the type of
maintenance is called scheduled or preventive maintenance.
Condition based maintenance requires that some means of assessing the actual condition of
the machinery is used in order to optimally schedule maintenance, in order to achieve
maximum production and still avoid unexpected catastrophic failures.

18
Chapter 4

Machine Condition Monitoring


The goal of machine condition monitoring and fault diagnostics is to:

Detect the onset of equipment deterioration.


Diagnose the existing condition.
Trend equipment condition or the progression of a fault, over time.
Prognose (predict) when ultimate failure will occur.

This excludes failures caused by unforeseen and uncontrollable outside forces such as
earthquakes and fire. These individual machine condition monitoring and fault diagnostics tasks
will be discussed in more detail later in this chapter. All of these activities combined should
allow more time for maintenance planning.

With the understanding that condition based maintenance may not be appropriate in all situations
(as discussed in the previous chapter), if the decision has been made to apply machine condition
monitoring and fault diagnostics in a selected part of a plant or on a specific machine, the
following is a list of potential advantages that should be realized.

Increased machine availability and reliability. Availability refers to whether or not a machine
is available for duty at a particular time when it is called upon. Availability is a measure of
the percentage of time a machine is ready to be operated or is in operation. A machine may
be unavailable due to being in a failed state or being under repair. Reliability refers to the
percentage of time that a machine is able to perform its designed duty. One can easily
imagine situations where a machine may be highly reliable (when in operation), but only
available a small percentage of the time (if it is too often experiencing scheduled
maintenance). An example of this would be a car that is frequently in the shop for repair and
therefore rarely driven (reliable when driven, but not regularly available). Alternatively, one
can also imagine a situation where a machine is almost always available, but not particularly
reliable (such as a spare machine that is never maintained). An example of this would be a
back-up generator that is never tested, lubricated, refueled, or otherwise maintained (always
available for use, but highly unreliable when called upon for duty).
Improved operating efficiency. This refers to efficiency of operation (low unit cost of
production) rather than efficiency related to other costs, such as maintenance costs.
Improved risk management (less downtime). In many industries (such as electrical utilities)
the costs of lost production are essentially too high to allow any unplanned or planned
interruptions. Loss of production then becomes a risk management balance.
Reduced maintenance costs (better planning). On balance, when using condition based
maintenance the overall costs (including the costs of monitoring) should be lower than other
strategies applied in the same set of circumstances.

19
Reduced spare parts inventories. Knowing when maintenance needs to be preformed should
allow for spare parts or replacement machines to be ordered from manufacturers on a just-in-
time basis.
Improved safety. Knowing the condition of machines should make the workplace safer by
allowing a clearer estimate of the likelihood of failure.
Safe short-term overloading of a machine is possible through improved knowledge of the
machine condition.
Extended operational life of the machine. Having an accurate estimate of a machines actual
condition not only allows for relatively safe overloads, but also provides an opportunity for
operators to scale back on load and/or throughput in order to preserve a machine in operation
(albeit reduced) until required maintenance can be scheduled efficiently. This may allow
additional time for acquiring the correct spare parts and the proper trained staff as well as
preventing the untimely stoppage of a process.
Improved customer relations (less planned/unplanned downtime). This may relate to external
or internal (other departments within a large organization) customers.
Elimination of chronic failures (root cause failure analysis and redesign). Having a record of
the condition of a machine may provide valuable information when attempting to fully
understand repeated failures.
Reduction of post-overhaul failures due to improperly performed maintenance or reassembly.

There are, of course, also some disadvantages that must be weighed in the decision to use
machine condition monitoring and fault diagnostics. These disadvantages are listed below.

Monitoring equipment costs are usually significant, but depend on the scale of the monitoring
to be done.
Operational costs (running the program) are also largely dependent on the scale of the
monitoring to be done.
Skilled personnel needed (training costs, higher salaries for trained technical staff).
Strong management commitment needed. Because it is always more difficult to win support
from management for reducing costs than it is for increasing revenue, a strong commitment
from management in terms of recognizing the long term benefits of machine condition
monitoring and fault diagnostics, is essential.
A significant run-in time to collect machine histories and trends is usually needed. Many
programs are cancelled or reduced during this time of building up a track record of
trustworthy results.

The ultimate goal of machine condition monitoring and fault diagnostics is to get useful
information on the condition of equipment to the people who need it in a timely manner. The
people who need this information include operators, maintenance engineers and technicians,
managers, vendors, and suppliers. These groups will need different information at different
times. The task of the person or group in charge of condition monitoring and diagnostics is to

20
ensure that useful data is collected, that data is changed into information in a form required by,
and useful to, others and that the information is provided to the people who need it when they
need it. For example, an operator will often need second to second feedback on the condition of a
machine or process, particularly if the lead time between incipient failure and catastrophic failure
is short (as in most high speed machinery). This information will need to be in the form of
overall vibration levels and alarms only. More detail is not required. Maintenance teams
however, may need weekly, monthly and even yearly data (trends, alarm reports) as well as a
complete picture of the current state of a machine (raw vibration data, frequency spectra) in order
to determine existing condition and project future estimates of potential failure. A manager may
only need quarterly or annual data showing the most general picture of the state of the machinery
(maintenance history, failure rates, cost/benefit of monitoring) in order to make longer term
decisions. Further general reading can be found in these references: Mitchell (1981), Lyon
(1987), Mobley (1990), Rao (1996), and Moubray (1997).

As mentioned previously the focus of this book will be on vibration based data. Still, other types
of data can be useful for assessing machine condition and these should not be ignored. These
include physical parameters related to lubrication analysis (oil/grease quality, contamination),
wear particle monitoring and analysis, force, sound, temperature, output (machine performance),
product quality, odor, and visual inspections. All of these factors may contribute to a complete
picture of machine integrity. The types of information that can be gleaned from the data include
existing condition, trends, expected time to failure at a given load, the type of fault existing or
developing, and the type of fault that caused failure.

The specific tasks which must be carried out to complete a successful machine condition
monitoring and fault diagnostics program include detection, diagnosis, prognosis, postmortem,
and prescription.

Detection requires data gathering, comparison to standards, comparison to limits set in-plant
for specific equipment, and trending over time.
Diagnosis involves recognizing the types of fault developing (different fault types may be
more or less serious and require different action) and determining the severity of given faults
once detected and diagnosed.
Prognosis, which is a very challenging task, involves estimating (forecasting) the expected
time to failure, trending the condition of the equipment being monitored, and planning the
appropriate maintenance timing.
Postmortem is the investigation of root-cause failure analysis, and usually involves some
research-type investigation in the laboratory and/or in the field, as well as modeling of the
system.
Prescription is an activity that is dictated by the information collected and may be applied at
any stage of the condition monitoring and diagnostic work. It may involve recommendations
for altering the operating conditions, altering the monitoring strategy (frequency, type), or
redesigning the process or equipment.

21
The tasks listed above have relatively crisp definitions, but there is still considerable room for
adjustment within any machine condition monitoring and fault diagnostic program. There are
always questions concerning such things as how much data to collect and how much time to
spend on data analysis, that need to be considered before the final program is put in place. As
mentioned previously, things such as equipment class, size, importance within the process,
replacement cost, availability, and safety need to be carefully considered. Different pieces of
equipment or processes may require different monitoring strategies.

Table 4.1 below shows several common fault types and the measurement parameter that could be
used to reveal the existence of such faults or changing overall machine condition. As can be seen
in this table, vibration based monitoring has the potential to provide moderate to excellent
condition assessment in most applications.

Table 4.1 Machine Condition Evaluation

4.1 Periodic Monitoring

Periodic monitoring involves intermittent data gathering and analysis with portable, removable
monitoring equipment. On occasion, permanent monitoring hardware may be used for this type

22
of monitoring strategy, but data is only collected at specific times. This type of monitoring is
usually applied to non-critical equipment where failure modes are well known (historically
dependable equipment) and/or spares are readily available. The main focus of periodic
monitoring is the trending of condition and severity level checks, with problems triggering more
rigorous investigations.

Periodic monitoring has become the principal component of machine condition monitoring and
fault diagnostics programs in many industries. Modern data collectors have made the routine
collection, trending and analysis of many machines cost effective. Regular monitoring can
provide a level of protection for critical equipment and the capability to evaluate critical
equipment should a machine defect or deterioration be detected. Machines are selected for
monitoring and their monitoring priorities are established before detailed plans are made.
Baseline data are used to define the normal operating conditions for a machine and to establish
the data needed for effective monitoring. The goal of any monitoring program is to select
measurements that provide the greatest sensitivity to any change in machine condition. A
procedure should be chosen for each machine when a program is initiated and modified as new
information is obtained.

4.1.1 Listing and Categorization of Machinery

The first step in a monitoring program is to list the machines in the plant. Machines should be
categorized according to a hierarchy based on the criticality of the machine to plant operation.
Machines may be ranked in four grades: A, critical; B, critical or failure-prone; C, critical, but
spared; D, non-critical. The periodic monitoring program should focus on categories C and D.
However, as mentioned above equipment in categories A and B may also benefit from periodic
monitoring, but should first be considered as candidates for continuous monitoring. Table 4.2
provides definitions for these categories.

Table 4.2 Machinery classification for monitoring.

Machine Classification Result of Failure


A: critical Unexpected shutdown or failure will cause significant
production losses
B: critical or failure prone Unexpected shutdown or failure reduces but does not interrupt
production
C: critical, but spared Light-duty service causes inconvenience in operation but no
interruption of production; repair costs may justify
D: non-critical Production will not be affected by failure; repair cost does not
justify monitoring

4.1.2 Machinery Knowledge

Knowledge of the characteristics of machinery is essential to conducting efficient vibration


analyses. The more information available about the machine (design, construction, supports,
operational responses and defect responses), the easier will be the diagnosis of defects and
malfunctions. Machinery knowledge includes the following things.

23
Broad characteristics such as rotational frequencies, gear mesh, vane pass and bearing defect
frequencies.
Vibration, temperature gradients, or pressure initiated by an operating component or system.
Vibration responses to process changes.
Characteristics identified with specific machine type.
Known natural frequencies and mode shapes.
Sensitivity to instability from wear or changes in operating conditions.
Sensitivity to vibration from mass unbalance, misalignment, distortion, and other
malfunction/defect excitations.

Certain responses (vibration, temperature, pressure) can be related to components of the systems
(such as bearings, pumps, fans, compressors, and gear teeth). Vibration responses displayed in
the frequency domain (more on this in later chapters) include centre frequencies (those
frequencies close to or at the fault characteristic defect frequency) and sidebands (response
frequencies that are removed from the characteristic defect frequency by one or more multiples
of the main shaft rotating speed). Table 4.3 lists some characteristic defect frequencies for
common machine components. Table 4.4 lists some common sources of machine excitation and
responses based on machine type. Many of the terms used in Tables 4.3 and 4.4 have not yet
been defined. Definitions will be provided as these particular fault types are discussed in later
chapters.

Table 4.3 Component sources of machine excitation and response.

Component Frequency
Antifriction (rolling ball pass frequency inner and outer race
element) bearings fundamental train frequency
rotating unit frequency
ball spin frequency
Hydrodynamic journal whip and whirl frequencies
bearings
Gears rotating unit frequency
gear-mesh frequencies (and harmonics)
system natural frequencies (gear-tooth defects)
Blade wheels and impellers rotating unit frequency
(fans and pumps) vane pass and blade pass frequencies (and
harmonics of vane and blade pass frequencies)
Rotors trapped fluid rotational frequency
natural frequencies (resonances)
higher harmonics
Couplings and universal orders of rotating frequency
joints
Reciprocating mechanisms rotating frequency and multiples (orders)

24
Table 4.4 Sources of machine excitation and response based on machine type.

Machine Type Characteristics


Centrifugal Machines - Impeller
Types
centrifugal pumps rolling element bearings (stiff)
vane-passing frequencies and their multiples
centrifugal compressors sleeve or tilt pad bearings (soft)
large casing-to-rotor weight ratio
vane-passing frequencies from impeller
fans vane-passing frequencies
system aerodynamics
pedestal characteristics often important
Bladed Machines
axial-flow compressors blade-passing frequencies and their multiples
steam turbines - mechanical drives blade pass frequencies
critical speeds (resonances)
rubs and mass unbalance
steam turbines for power generation blade pass frequencies
mass unbalance
gas turbines blade pass and gearing frequencies
instability and rubs
Power Transmission Equipment
gearboxes gear-mesh frequency and higher harmonics
casing and gear tooth resonances
pitch-line runout
torsional responses
fluid drives slip-frequency excitation
bearing whirl
Motors/Generators slip frequency modulation
pole-induced structural vibration
thermal-induced excitation
high synchronous-motor excitation at start-up
stator shorts
Reciprocating Machines
engines casing distortion
bearing-induced foundation vibration
high torsional excitation by inertia and pressure
pumps and compressors high torsional excitation by inertia and pressure
Small Equipment antifriction bearing failures
looseness
belt and gear-drive problems

25
Details of each machine that should be kept on file should include the following.

The location and name of the equipment.


The author and date on which the information was acquired.
A generic description of the machine.
Plant asset numbers.
A sketch andor photos of the installation for reference and background information.

Figure 4.1 shows a sample survey request form that could be used to specify machine details
prior to data collection. Not listed on this particular form is whether or not repeat monitoring is
required and if so at what frequency. The mechanical component list is critical in structuring the
database and for follow-up analysis. Minimum requirements for this type of form are machine
type, manufacturer, horsepower, operating speed, bearings, model number, serial number, and
internal configuration. Sketches of machine layout and/or photographs may also be useful.

SURVEY REQUEST
PLANT BY DATE

AREA EQUIPMENT:

GENERAL DESCRIPTION: ASSET NUMBER:

MECHANICAL COMPONENT
INFORMATION:
ROLL DIAMETER: AVG. SPEED:

BEARING (MOTOR, ROLL) MOTOR NAMEPLATE DATA:


INBOARD H.P. S.F.
OUTBOARD RPM FRAME

REDUCER: TEETH BEARINGS


ST.
TYPE: 1 RED:
RATIO: ND
2 RED:
MFG.: RD
3 RED:
MODEL #:
SERIAL #:

Figure 4.1 Example of a survey request form.

26
4.1.3 Route Selection and Definition

The data collection route selected can be based on plant layout, machine train (process stream),
machine type or type of data required. Plant layout and machine train routes are the most
commonly used. Routes based on plant layout follow the floor plan and progress from one
machine to another. Routes based on machine train require data be taken on all machines in the
production or processing of a product line. Machine type based routes include only the same type
machines. Routes based on the type of data being collected require that all points have similar
processing spectral, overall, band, HFD (high frequency detection), or similar components.
Routes should be set to follow from machine to machine naturally and easily. Routes should be
tailored to meet the needs of the plant, the equipment and the operator. An initial route should be
a small number of machines, which can be expanded as desired.

There are a variety of ways to keep track of vibration data collection points. Manual systems
based on a map or plan of the plant are useful for small plants or local routes. However, for
larger monitoring programs a more formalized data point organizational program may be
required. Figure 4.2 shows a bar code based monitoring point location tracking system. Each
data collection point is identified with a bar code. These types of systems reduce the likelihood
of errors in data collection (the incorrect labelling of data collected in the field) particularly as
some data collection may require vibration data to be collected in more than one direction at one
single point. In order to keep a monitoring program on track it is important to; i) not change
personnel too often, ii) keep personnel well trained, and iii) keep things simple.

Figure 4.2 Example of a survey barcode data point location tracking system.

27
4.1.4 Measurement Parameters and Points

Measurement points are typically identified after the route has been selected. The measurement
report should include: unit ID, location and position, measurement units, measurement data,
previous measured amplitude, percentage change from last measurement, and alarm status. In a
new program, data may be taken in all directions (radial to the rotating shaft in the horizontal and
vertical directions and axially (parallel to the rotating shaft)). The readings should be in the plane
with the greatest support flexibility. Six measurements over three points may be necessary. Later,
redundant measurement points may be dropped. Measures are intended to be sensitive to
machine condition and frequency spans can be adjusted according to baseline data results.

The optimum point configuration should provide measurements that are in response to vibration
and provide alarm levels that announce threshold crossing and initiate analytical data collection.
Time waveform and spectral (frequency domain) data with detailed resolution (if possible)
should be collected upon reaching alarm levels. Frequency ranges that allow for analysis of
operating speed and multiples of operating speed (orders) and bearing defects are required.

A measurement point is assigned for each end of a motor, each rotational component of a
reducer/increaser, and each end of a roll. The orientation of a transducer may be radial, axial, or
both depending on machine type, bearing, service and application. The points should be
organized under the respective machine, area, and plant into a hierarchy within of the data base.

The hierarchical outline should contain the following.

A description of the plant (the highest category).


A subsystem of machines (the second category).
A single machine from the group at the next level.
The individual measurement points that are used to evaluate the machine.

4.1.5 Baseline Data

Baseline data are used to evaluate the condition of machines of the same design operating at
different normal vibration levels. Baseline data provide the initial data for selecting a trending
type, the trending database and information for setting alarms. The typical representation of
baseline data is the trending of overall measurement parameters (overall vibration levels).
Transient, start-up and coast-down, data are used to evaluate class A equipment (500 HP and
above).

4.1.6 Frequency of Data Collection

The frequency of monitoring depends on the following factors.

The mean time to failure (MTTF). The more imminent the failure the more frequent the
data should be collected.

28
Criticality of the machine. The more critical the machine the more frequent the data
should be collected.
The number of spares. The fewer spares available the more frequent the data should be
collected.
Production and failure repair costs. The higher the repair and lost production costs the
more frequent the data should be collected.
Available personnel.
Monitoring costs.

General guidelines for determining the frequency of data collection include the following.

Initially any arbitrary span such as a week, month or several months may be suitable.
The records from a machine should be reviewed to assess past failure frequency.
Important factors include loss of production, machine replacement and personnel costs.
If the machine is performing well monitor less frequently.
If the schedule does not allow for consistent work, decrease the number of machines
monitored.
Periodic monitoring is an expensive way to compensate for unreliable machines.
With the exception of critical machines, quarterly monitoring should be adequate.
High-speed machines typically require more frequent monitoring.

4.1.7 Selection of Test Equipment

The selection of instrumentation used depends on the operation of the program, number of data
points and the depth of analysis that is likely to be carried out. The minimum requirements for
data collector are as follows.

Capable of trending overall vibration levels and specified frequency band readings.
Performing analogue and digital integration (to allow post-collection conversion of
vibration data between displacement, velocity and acceleration). More on this in later
chapters.
Providing up to 6,400 lines of resolution with a dynamic range of 72 dB. More on this in
later chapters.
Selection of at least Hanning or uniform windows for frequency domain data processing.
More on this in later chapters.

4.1.8 Screening

Screening is used to assess, at relatively low cost, when a problem is developing in a specific
machine. It allows time for analysis and to prepare for repairs. Techniques vary in sophistication

29
and effectiveness. The effectiveness depends on the device and the type of machine being
monitored. Allowances must be made for changes in operating conditions.

Simple methods include the following.

Devices including screw drivers, wires, and stethoscopes for listening to a machine.
RMS meters used to monitor overall vibration levels. The doubling of vibration level
(regardless of the absolute level) usually indicates some corrective action is necessary.
High frequency defect meters are used to detect the activity around the natural frequency
of the sensor. More on these in a later chapter.
Simple screening devices are most useful if non-destructive pulses and noise are not
present.

Limitations of simple screening devices include the following.

May not be capable of distinguishing a new fault if the level is low.


A low-vibration level related fault may be masked by the normal vibration of another
component.
May be ineffective in the presence of random noise and vibration.

More elaborate methods include band filtering, spectrum analysis, time waveform analysis and
enveloping (amplitude demodulation). Band filtering displays changes in vibration in distinct
frequency bands. Faults may be separated into broad categories such as mass unbalance,
misalignment, bearing defects, and gear-mesh defects. Distinct frequency ranges can then be
screened with electronic data collectors looking for changes in specific frequency ranges that
indicate changes in the related specific machine components. Spectrum and time waveform
analysis are also used to study the frequencies and energy in rolling element bearing defects in
low-speed machines. The peak vibration obtained from a time waveform or peak-detection
circuit may be more sensitive to bearing condition than spectral measurements.

Because of its usefulness in detecting the earliest stages of rolling element failures, the basic
principles of the enveloping (amplitude demodulation) method are presented here. They will also
be further discussed in a later chapter.

In some high-speed machines (>3,600 RPM), rolling element bearing defects may appear
in the high-frequency range (5 kHz 40 kHz).
In order to highlight the high frequencies, the lower frequencies (containing operating
speed and gear meshing induced vibration) are first filtered out (high pass filtered).
The filtered signal is then demodulated (rectified and then low pass filtered) to produce a
signal free of natural frequencies and other excitations related to operating speed.
A spectrum of the demodulated signal typically shows the bearing frequencies and the
nature of the defects more clearly than a frequency spectrum based on the original
vibration signal.

30
4.1.9 Trending

Trending is carried out to reveal the vibration related or process characteristics based on the
recordings that are collected on a regular basis. Monthly trending is the most popular method.
Filtering or band trending may be useful in some cases. Trending is flexible in terms of
frequency ranges, measures (RMS or peak), and process characteristics. Normalization may be
needed on vibration data before trending in order to compare machines operating under different
loads.

Figures 4.3 to 4.5 show examples of trending carried out with overall vibration levels, band
filtered, and broad band frequency information respectively.

Figure 4.3 Trend of overall (all frequencies averaged together) vibration levels (no alarm).

Figure 4.4 Trend of sub-synchronous (below shaft rotational speed) vibration levels (no alarm).

31
Figure 4.5 Trend of rotational speed (1X) and other frequencies up to 1200 Hz.

4.1.10 Alarms

Two or three alarm levels are typically used in the trending process. An alert alarm means a
detailed vibration analysis should be performed. It is established on the basis of condition
changes as indicated by 2.0 to 2.5 times changes in vibration level. That is, the overall vibration
level increases by 2.0 to 2.5 times its typical level. It may imply maintenance action needs to be
taken or that more frequent monitoring is required. A warning alarm indicates more serious
problems. This alarm level should lead to full-scale analysis or maintenance. The time for action
is limited. A fault alarm means failure is close if no maintenance action is taken. Maintenance
action may include balancing, repair, redesign or more careful installation.

Establishing realistic alarm levels requires knowledge of machine condition and vibration
signals. This topic will be discussed more fully in a later chapter. In a newly established
program, the alarms may be set up based on information about other equipment, experience of
others and standards (i.e. vibration charts). Alarm levels should be reviewed from time to time
and changed to reflect the experience gained during the monitoring program. Table 4.5 shows a
general machine condition evaluation guideline. These numbers can be used to set alarm levels
for general purpose machines such as medium sized motor-pump sets.

As an example, consider a 300 HP pump with rolling element bearings that is being operated at
1,200 RPM. The alarm levels, according to Table 4.5 should be as listed below. More discussion
of vibration signal measurement parameters such as Peak and RMS will follow in subsequent
chapters.

Alert: 0.12 inches per second (RMS)


Warning: 0.28 inches per second (Peak)
Shut down: 0.6 inches per second (Peak)

32
Table 4.5 Vibration guidelines for general machine condition evaluation.

Condition Limits
RMS velocity Peak velocity
in/sec in/sec
Acceptance of new or repaired equipment <0.08 <0.16

Unrestricted operation normal <0.12 <0.24


Surveillance 0.12 0.28 0.24 - 0.7
Unsuitable for operation >0.28 >0.7

Table 4.6 shows another general machine condition evaluation guideline. These numbers can be
used to set alarm levels for larger machines supported on journal bearings such as turbine
generators. As an example, consider an 18,000 HP turbine with journal bearing clearances of 8
mils (0.2 mm) and an operating speed of 10,000 RPM. The alarm levels, according to Table 4.6
should be as listed below. Note that a mil is equal to one one-thousandth of an inch. The term
mils and thou are often used interchangeably.

Alert: 0.2 x 8 = 1.6 mils (peak-to-peak)


Warning: 0.4 x 8 = 3.2 mils (peak-to-peak)
Shut down: 0.6 x 8 = 4.8 mils (peak-to-peak)

Table 4.6 Evaluation of rotor/bearing vibration.

Maintenance Allowable Rotor Clearance


3,600 RPM 10,000 RPM
Normal 0.3 0.2
Surveillance 0.3 0.5 0.2 - 0.4
Shut down at next convenient time 0.5 0.4
Shut down immediately 0.7 0.6

Figure 4.7 shows a generic trend of overall vibration levels showing regular samples and alarm
levels. This figure shows an idealized version of what a vibration trending plot should reveal.
Namely, the gradual change in the overall vibration level (representing the gradual change in
machine condition) in relation to set warning levels that are used to trigger some appropriate
action.

The spectra that follow (Figures 4.8 to 4.11) represent the same data displayed in different
formats. The comparison shows that, although the overall level and the sub-synchronous levels
remain low (Figures 4.8 and 4.9), other vibration characteristics have changed significantly.
Specific higher frequencies increase while the 1X (1 times shaft rotational speed) peak decreases.

33
Figure 4.7 Generic trend of overall vibration levels showing regular samples and alarm levels.

Figure 4.8 Trend of overall (all frequencies) vibration levels (no alarm).

Figure 4.9 Trend of sub-synchronous (below shaft rotational speed) vibration levels (no alarm).

34
Figures 4.10 and 4.11 show examples of trending carried out with two different selected
frequency band vibration levels. In each case there are different alarm levels set and there are
different trends. These types of analysis tools are used to focus in on specific frequency bands
where particular frequency responses would be expected for specific faults in particular
components. This evidence shows that the overall reading does not always allow an accurate
assessment of machinery condition.

Figure 4.10 Trend of selected frequency (9-35 times shaft rotational speed) vibration levels (alert
level alarm).

Figure 4.11 Trend of selected frequency (36-65 times shaft rotational speed) vibration levels
(alert level alarm).

35
4.1.11 Reports

Machine condition monitoring and fault diagnostics reports provide the following general
information.

Record of case histories, alarms and actions.


Technical data that result in a more efficient program.

Data compression and retrieval should be planned for accessibility and utility. Data compression
techniques are preferred in order to minimize the storage space for long term programs, but of
course need to preserve all important information contained in the collected data. A report should
contain the following detailed information.

A description of each measurement point (including all pertinent information about the
machine and operating conditions at the time the measurement was taken).
All measurement parameters.
Date of measurement.
Previous and latest values.
Percentage change.
Trend plots, spectra and alarm status.
Recommendations and executive summary.

4.1.12 Fault Diagnosis

Fault diagnostics requires inspection and analysis of vibration data in many different forms. In
later chapters the details of vibration signal analysis will be fully discussed. For now it is enough
to point out that different faults generate different related characteristics in the vibration signal
and these characteristics can be used to diagnose the different faults present. It is important to
remember that a full analysis of all the data available is the best way to diagnose the fault present
and at the same time eliminate other potential faults from consideration. With some diagnostic
experience, one can recognize faults relatively quickly.

Bearing defects (for example) may be recognized by their high frequency peaks and patterns of
high frequency and non-synchronous frequency energy spaced by 1 order (1X, 1 times shaft
rotational speed). Specific machine component technical details are not always needed when
these general diagnostic patterns are evident (see Figure 4.12).

The time waveform provides yet another helpful tool in vibration analysis (Figure 4.13). High
levels of impacting and ringing suggest that the balls or rollers of a bearing are passing over a
race defect. Again, with some experience, this pattern suggests concern regardless of the overall
trend or lack of bearing ID.

36
Figure 4.12 Evidence of relatively high frequency, non-synchronous vibration (rolling element
bearing defect likely).

Figure 4.13 Evidence of ringing vibration caused by rolling element bearing defect.

Figure 4.12 Bearing geometry defined bearing defect frequencies (dotted lines).

37
Accurate diagnosis of a rolling element bearing inner race defect is also possible without
knowing the bearing geometry associated with the bearing ID. Figure 4.14 shows how fault
frequency overlays (dotted lines) can be useful in such as case. A much more detailed discussion
of fault diagnostics will be presented in Chapter 10.

4.2 Continuous Monitoring

Constant or very frequent data collection and analysis is referred to as continuous monitoring.
Permanently installed monitoring systems are typically used, with sampling and analysis of data
done automatically. This type of monitoring is carried out on critical equipment (expensive to
replace, with downtime and lost production also being expensive). Changes in condition trigger
more detailed investigation or possibly an automatic shutdown of the equipment. The same
detection, trending and fault diagnosis techniques are again employed during continuous
monitoring, but are often fully automated.

Chapter Summary

Potential advantages of machine condition monitoring include increased machine availability


and reliability, improved efficiency, reduced costs, extended operational life, and improved
safety.
Some of the disadvantages of condition monitoring include monitoring equipment costs,
operational costs, and training costs.
The ultimate goal of machine condition monitoring and fault diagnostics is to get useful
information on the condition of equipment to the people who need it, in a timely manner.
The specific tasks which must be carried out to complete a successful machine condition
monitoring and fault diagnostics program include detection, diagnosis, prognosis,
postmortem, and prescription.
Periodic monitoring is used to assess the condition and changes in the condition of machines.
Measurements are selected that provide the greatest sensitivity to a change in machine
condition with the least complexity and data processing.
For a new periodic monitoring program, machines should be listed and placed in a
hierarchical order of importance to production.
Equipment knowledge paves the way for accurate machine fault and condition analysis and
should be consolidated in one table.
Data collection routes are based on plant layout, machine train, machine type, or data type.
Measures and measurement points are selected for efficient collection of data pertaining to
condition; redundant measurements should be eliminated as experience is gained with the
program.
Frequency spans used in measurement are based on the defect frequencies of the machine.

38
Baseline data provide a reference for evaluating changes in condition.
The frequency of data collection is based on mean time to failure (MTTF) of machine
components, cost of failure, available personnel, number of spares, and monitoring costs.
Chronic problems dilute the resources for monitoring and should be solved.
Screening can be a low-cost method for detecting changes in machine condition.
Frequency band screening may be necessary to obtain the sensitivity required to assess
changes in machine condition in complex machines with rolling element bearings.
Trending provides the opportunity to compare screening measures and the alarm settings that
either initiate analysis, more frequent monitoring or repair.
Alarms are used in periodic monitoring to alert the data collector that a significant change in
condition has occurred.
Two or three alarm levels are typically established on the basis of a 2 or 2.5 times increase in
the trended measure.
The effect of process changes on a trended measurement parameter must be taken into
account during trending.
Reporting is used to document case histories, record alarms, and request maintenance action.
Report formats should report the facts in a simple manner to the proper authority.
Reports should include case histories of unusual problems, information about pre- and post-
overhaul conditions, alarms and warnings, and information about time and materials and cost
accounting.

39
Chapter 5

Basic Vibrations
Mechanical vibration is the oscillation of a mechanical system about a reference position.
Vibration may occur in any direction and in multiple directions at the same time. Vibration is an
everyday phenomenon that can be seen in our everyday lives at home (the way a diving board
moves after the diver has launched into the air), during transport (the way a car suspension
moves when traveling over a pothole, the way an airplane wingtip moves in response to
turbulence) and at work (the way the elevator moves up and down just as it comes to a stop at
each floor). When mechanical systems are properly designed vibration responses are usually
minor and in most cases undetectable by human senses. However, the vibration response is still
there and may become excessive if the operation of the system moves outside the designed
operational limits.

Vibration (even when small in amplitude) is often a destructive and annoying side effect of a
useful process. These vibrations are the result of forces that may cause accelerated wear and
fatigue in mechanical components. However, sometimes vibration is generated intentionally to
perform a useful task (such as material transport or material separation). Vibration in machines is
the result of dynamic forces in machines which have moving parts and in structures which are
connected to those machines. Different parts of a machine will vibrate with different frequencies
and amplitudes. Accelerated wear and fatigue processes caused by mechanical vibrations are
often responsible for the ultimate breakdown of a machine.

The study and description of mechanical vibrations covers a wide range of material. Here the
focus will be constrained to a definition of a simple one degree-of-freedom, lumped parameter
vibration model. This model is suitable for describing the vibration response in most mechanical
systems and allows subsequent development of mechanical system fault detection and diagnosis
procedures, which is the focus here. Topics such as modal analysis, operational deflection shape
analysis and multi degree-of-freedom system analysis will be briefly described, but largely left
for more advanced treatments of the subject.

Mechanical vibration analysis provides a useful tool for gaining insight into mechanical system
response and condition for a variety of reasons. It is a benign, non-intrusive data collection
method which can be performed without having to take the machine off-line. It is applicable to a
broad range of machinery (electric motors, pumps, fans, engines, turbines) and machine
components (bearings, gears, couplings, belts). Mechanical vibration analysis is applicable to a
broad range of fault types (such as general wear, fracture, misalignment, and unbalance). The
associated hardware used to measure and record mechanical vibrations is relatively economical
and reliable. This hardware is also relatively commonplace in industry, and is relatively straight-
forward to use. Finally, vibration signals are well suited for continuous on-line monitoring which
is extremely useful when dealing with expensive and/or critical machinery.

The focus here is on the use of mechanical vibrations for machine condition monitoring and fault
diagnostics. Mechanical vibration signal analysis also find uses in many other situations and for
many other purposes. A comparison between the vibrations inside a package with external

40
applied vibrations can reveal the effectiveness of packaging. Package testing can also be used to
monitor vibration and shock that a product may experience during transport. Vibration
measurements may be used to determine maximum impact acceleration levels experienced
during collisions (inside crash dummies). Vibration (shock) measurements may also form part of
a vehicle safety system (airbag deployment). The measurement of shock exposure experienced
by sensitive cargo during transport is another application.

Vibration signal analysis is currently the dominant machine condition monitoring and fault
diagnosis technology used in industry. It is particularly useful when applied to machines with
rotating components. As mentioned in a previous chapter, there are various levels of analysis that
may be performed once a vibration signal has been recorded. Broad band trending is typically
used to determine the overall functioning of a machine. Narrow band trending is used to focus on
particular fault conditions (unbalance, misalignment) that may be dominant in a particular type
of machine. Spectrum (signature) analysis is used to focus on specific frequency components
in the vibration signal in order to highlight specific frequency responses that relate to specific
machine component deterioration or failure (bearings, gears).

As mentioned earlier, vibration based machine condition monitoring and fault diagnostics is not
the only strategy that may be employed to detect and diagnose changing machine conditions.
Even though all methods should be considered as potentially useful sources of data, Figure 5.1
shows that vibration is generally considered more successful at detecting a wider variety of fault
conditions than other methods.

Figure 5.1 Typical fault types versus different monitoring methods that may be used to detect
machinery deterioration.

Vibration by itself is usually identified as a maintenance problem if; i) the machine reliability
and/or maintainability are affected, or ii) the machine operation is considered unacceptable due
to the vibration levels.

Machine vibration can be categorized according to the source of vibration. It is typically either
rotary mechanical, other mechanical, hydraulic, and/or electrical. Rotary mechanical vibrations
include mass unbalance (non-uniform mass distribution on a rotating part), coupling

41
misalignment (centerlines of coupled machines not aligned), bent shaft or rotor (heavy rotors
bend if stationary for long periods, shafts are prone to damage during operation), and faulty
bearings, gears, belts, and pulleys.

Other mechanical vibrations include looseness (loose parts hit against limits of movement
creating off-harmonic frequency components (between regular harmonic frequencies)) and soft
foot (non-level hold-down bolts can result in distortion of machine casing - hence change in
stiffness and a change in the structural resonance of the machine). The stiffness change and
resulting shift in the resonant frequency can be remedied by loosening the hold-down bolts and
shimming the base of the machine (leveling), then tightening the hold-down bolts. Excitation of
any other structural resonance (of the machine or supporting structure) also fall into this
category.

Hydraulic vibrations are due to vane or blade passing in pumps, fans and compressors (frequency
= (number of vanes) x (rotational speed)), cavitation (causes erosion damage, resulting in
broadband, high frequency vibration), and piping vibration (often related to piping system
resonances). Electrical vibrations are typically found in A.C. motors (unbalanced electrical load
on windings of the motor or unbalanced rotor bars) and transformers (60 Hz harmonics due to
magnetostriction of transformer core iron). Figure 5.2 shows a variety of common vibration
problems and the approximate percentage of occurrence. It should be noted that the percentages
listed do not sum to 100 percent. This is due to the compounding of problems that often occurs.
Several vibration problems may exist on the same machine at the same time, but only one is
usually dominant.

Figure 5.2 Typical vibration problems and the approximate percentage of occurrence.
(Note: Imbalance means exactly the same thing as Unbalance)

42
5.1 Classification of Vibration by Type of Motion

Before getting into a full mathematical description of vibration signals, it suits the needs of the
subject matter to first describe vibrations in several general terms based on the type of motion
that may be seen and the type of excitation that causes the vibration. Both of these descriptive
assessments will become more useful later in the book.

5.1.1 Simple Harmonic Motion

The simplest form of vibration is referred to as simple harmonic motion. In simple harmonic
motion the exact position of a vibrating body or particle may be predicted from a sinusoidal
equation of motion. The mathematical description is shown in equation 5.1.

x(t ) A sin(t ) 5.1

The terms in equation 5.1 are defined below.

x(t) - instantaneous displacement (m)

A - maximum amplitude (m)

- angular velocity (Radians/Second)

- phase angle (Radians)



f = frequency, 2f f
2
T = cycle/period, T 1/ f

Figure 5.3 shows a graphical representation of simple harmonic motion. The vertical position of
a point on the circle at the left can be plotted as that point moves around the circumference of the
circle. The vertical position of that point as a function of time defines simple harmonic motion.

60

-0.6
1 -
0 1

Figure 5.3 Graphical description of simple harmonic motion.

5.1.2 Periodic Motion

43
Periodic motion is any motion that repeats itself in equal time periods. This classification
includes harmonic motion (simple harmonic motion), as well as pulses, and any other repeated
vibratory motion. Figures 5.4 to 5.6 show examples of periodic motion. While somewhat
artificial, Figure 5.4 shows a clearly periodic dynamic signal. This representation shows how the
repeated (periodic) nature of the signal may not involve the exact repetition of shape during each
cycle, but does involve the general repetition of shape.

0.8

0.6

0.4
Amplitude
0.2

-0.2

-0.4

-0.6

-0.8

20 30 40
Time (ms)

Figure 5.4 Clearly periodic vibratory motion.

Figure 5.5 shows a periodic dynamic signal with slightly more variability in the exact shape of
each cycle. The peak amplitudes of each cycle are changing gradually (diminishing in this case,
but they could also be gradually increasing in amplitude) as time progresses. This represents a
dynamic signal generated by a defective rolling element bearing. The spikes in the signal show
the interaction of rolling elements with defects in a raceway. The changing peak amplitude
suggests that the signal represents the vibrations that are being generated as the rolling elements
move through the load zone. As the rolling elements exit the load zone the energy in each rolling
element-defect interaction will be reduced. If the rolling elements were entering the load zone
they would gradually increase in amplitude.

Figure 5.5 Periodic vibratory motion with changing peak amplitude.

44
Figure 5.6 shows a periodic dynamic signal with significant variability in the shape of each
cycle. The peak amplitudes of each cycle are similar and not changing as time progresses. This
may be representative of a dynamic signal generated by a rolling element bearing without any
defects or in an extremely advanced stage of deterioration (more on bearing faults in a later
chapter). The spikes in the signal are at approximately the same amplitude as the background
noise. The only way to distinguish the two potential conditions is to look carefully at the
magnitude of the vibration signal. While signal amplitude units are not listed in Figures 5.5 and
5.6, if they are considered to be from the same machine operating under the same conditions and
taken with the same instrumentation, it is clear that the amplitude of the signal in Figure 5.6 is
significantly larger than in Figure 5.5, leading to the conclusion that this signal represents a
bearing in an advanced stage of deterioration.

As will be discussed at length in later chapters, the analysis results described above, which are
based on a relatively quick and rudimentary consideration of the vibration signal, need to be
confirmed with other analysis methods and potential other measurement data (vibration and/or
other) in order to assure trustworthy conclusions.

Figure 5.6 Periodic vibratory motion with peak amplitudes only slightly above background noise
levels.

5.1.3 Random Motion

Random vibratory motion is that motion which is not deterministic. That is, the motion is not
repeatable in any way that allows the exact value of any future amplitude to be determined.
Statistics based on the motion history may be well defined, but the exact amplitude as a function
of time is not obtainable. The random vibration signal will typically contain all frequencies in
given bands. Random vibration is often generated by machine looseness as the impacts from
loose components may occur randomly (unrelated to machine speed). Other forms of looseness
may generate vibration signals that are related to machine speed (loose foot see later chapter).
Random vibration typically results in broad band responses in the frequency domain.

45
1

0.8

0.6

Amplitude
0.4

0.2

-0.2

-0.4

-0.6

-0.8

-1
0 10 20 30 40 50 60 70 80 90
Time (ms)

Figure 5.6 Random vibratory motion.

5.1.4 Transient Motion

Transient motion includes any motion other than the above. Typically transient vibration, as
suggested by the name, is impulsive in nature, but not regularly repeated. Figure 5.7 shows
samples of the different types of vibratory motion.

Figure 5.7 Different types of vibratory motion.

5.2 Classification of Vibration by Type of Excitation

The cause of all vibratory motion is the existence of one or more pulsating forces (excitation).
The cause of vibration is governed by the process under consideration, the tolerances that exist
within the system and any defects that may be present. Examples of sources of vibration
excitation include mass unbalance and misalignment.

5.2.1 Free vibration

Free vibration excitation is when the vibratory oscillations occur at the system natural frequency.
The motion is initiated by a transient external force. The motion continues for at least a short
time after the initial force input has disappeared.

46
5.2.2 Forced vibration

System oscillation occurs at the frequency of a driving force input, which may or may not be at
the system natural frequency.

5.2.3 Self-induced vibration

Self-induced vibration of a system is the vibration resulting from conversion of energy within a
system. The excitation is caused by non-oscillatory energy within the system being converted to
oscillatory excitation.

5.3 Wave Fundamentals (Amplitude, Frequency, Phase)

Basic vibration signal attributes are the amplitude (magnitude or severity) of the vibration, the
frequency (rate) of oscillation, the frequency content (all the different frequencies that may be
contained in the signal), the phase relationship of the vibration signal to some reference or of one
vibration signal to another, and the offset of the vibration signal from some baseline.

Vibration wave amplitude is the maximum value at a given location. The units are typically
given in Volts (if the output has not been converted to an appropriate scale which represents the
units being measured) or in units that reflect what is being measured (mm, inches, m/sec,
inches/sec, Gs, m/sec2). Several conventions exist that result in the amplitude typically being
recorded as an RMS value, a Peak value or a Peak-to-Peak value. Frequency is measured in the
number of complete cycles per unit time. The units are typically cycles per second (CPS or
Hertz), cycles per minute (CPM), or orders of operating speed (RPM). The wave period is the
amount of time for one complete cycle to take place. The period is the inverse of frequency and
has units of seconds.

Figure 5.8 shows a sample vibration signal (which happens to be simple harmonic motion). This
figure shows a signal with an amplitude of 1.0 units, a frequency of one cycle per 6 seconds, and
zero offset.

Figure 5.8 A sample vibration signal (simple harmonic motion).

47
In machine condition monitoring applications phase (degrees) is typically referred to the time
relationship between vibrations at the same frequency, but at different locations on a particular
machine. Figure 5.9 shows three example vibration signals that may have been recorded from
different locations on a particular machine. The phase differences shown here are 90 degrees
(Relative Phase 1) and 180 degrees (Relative Phase 2). In these cases the second and third
signals lag behind the first signal by 90 and 180 degrees respectively. That is the phase
difference between these signals.

Phase may be measured in the time waveform using an oscilloscope, dual-channel analyzer,
phase meter or strobe light. Phase indicates the time relationship between two sensors (signals).
Applications where phase should be measured include situations involving machine
misalignment, frequency of critical speed measurements as well as determination of the location
of the heavy spot on a rotor (balancing applications).

Figure 5.9 Phase relationships between three vibration signals.

Figure 5.10 shows the relationship between two measured signals. One is the measured vibration
signal (upper trace on left hand side) and the other is the measured phase reference signal (lower
trace on left hand side). In this case the phase reference signal is represented as a marker on the
same scale as the vibration signal.

Figure 5.10 Phase relationships may be used to determine the location of the heavy spot on a
rotor.

The phase reference signal (also called a phase trigger signal) may take other forms as well
(square waves, spikes), but basically always marks a reference point on the vibration signal that
is at the same location every rotation of the shaft. This allows the location of the maximum
vibration amplitude (associated with the heavy spot in the case of mass unbalance) to be linked

48
to a known position on the rotating shaft. This is critical information for determining the best
corrective action (placement of a counterweight) to reduce the excessive vibration levels. More
detail relating phase measurements to balancing activities will be presented in a later chapter.
The phase trigger sensor is often called a keyphasor as it typically uses a key or keyway on the
shaft as a reference point. In this case both sensors are the same type, but they may be different
sensor types in different applications.

Phase measurement is basically a measure of the time between two events. This may be a
vibration signal taken from one location in a rotating shaft and a reference signal that allows the
vibration signal to be linked to a known location on the shaft. Alternatively, phase may be used
to relate two vibration signals that represent the same vibration signal, measured from the same
rotating shaft, but measured at two different locations (as represented in Figure 5.9 above). This
typically requires that the signals of interest have one dominant frequency component. If they do
not, signals are typically filtered until only a single frequency remains before the phase
measurements are made.

Figure 5.11 shows a situation where the absolute phase measurement is being taken. The phase
represents the time between the most recent zero crossing (the trigger point that starts the timer)
and the next maximum amplitude value (the point where the timer is stopped).

Figure 5.11 Absolute phase measurement.

A keyphasor is a transducer that detects the location of a reference point on a rotating shaft.
These reference points are typically keys or empty keyways. When the transducer detects the
once per turn marker, an output signal (keyphasor timing event) is generated. The event is a
sudden negative (or positive) change in signal voltage when the leading edge of the keyway
passes next to the probe. Figure 5.12 shows a keyphasor sensor and output relative to shaft
rotational position. Figure 5.13 shows a keyphasor signal that is 180 degrees away from the
known high spot on a rotating shaft. The output plot (amplitude versus time) shows the
measured vibration amplitude relative to the measured phase (triggered tachometer pulse).

49
Figure 5.12 Keyphasor transducer and output signal for use in phase determination.

Figure 5.13 Keyphasor transducer and output signal together with measured vibration signal.

Relative phase measurement compares the timing of equivalent events on two vibration signals.
This is often used when comparing two vibration signals taken from different locations on the
same machine. This type of measurement requires that the two signals have the same frequency
(as is the case if the signals are taken from the same machine). The same measurement units
must also be used. The transducers should be in the same plane or at least have the same
direction. Figure 5.14 shows two vibration signal that display a relative phase lag.

50
A B

Figure 5.14 Two vibration signals that display a relative phase lag. A lags B because it reaches
the same amplitudes at delayed times.

In general, two vibration signals will have a phase difference if they reach positive peaks at
different times. The phase difference can be measured and expressed in units of degrees. Phase
measurements are important for the diagnosis of several different machine faults. It is
particularly important when balancing rotating machinery. The amplitude and phase information
can be combined to produce a picture of the deflection or mode shape of a running machine
(more on this in later chapters). Figure 5.15 shows a set of sensors that may be used to reveal
modal shapes or operational deflection shapes. The vibration signal from the sensors will contain
the same frequency information, but they will be in phase or out of phase depending on the shape
of the object being monitored. In this case the upper schematic in Figure 5.15 shows a shaft
vibrating with a shape that is typical of a first natural frequency (first mode shape). The lower
schematic in the same figure shows a shaft vibrating with a shape that is typical of a second
natural frequency (second mode shape). The relative phase and amplitude between each sensor
can be used to define the responses. The vibration at the source of a machine problem always
happens first in time. As the vibration propagates away from the source location, it experiences a
time lag.

Figure 5.15 Vibration measurement locations that can be used to show a machine operating
deflection shape (if the measurements include phase).

51
5.4 Basic Spring-Mass Model of Vibration

Once a (theoretical) system of a mass and a spring (no damping) is set in motion it will continue
this motion with constant frequency and amplitude. The system is said to oscillate with a
sinusoidal waveform (simple harmonic motion). Figure 5.16 shows the simplest form of such a
vibrating system (a mass supported on a spring that has been set in motion). The figure shows the
resulting simple harmonic motion (on left) with the equation that describes the motion. The
period is T seconds and the frequency is 1/T Hz. Because there is no damping this frequency in
Hertz (fn) is also exactly equal to the system natural frequency in Radians per second (n). In this
case the system natural frequency can be found to be,

k
n 2 f n 5.2
m

where k is the system spring stiffness and m is the system mass.

The sinusoidal waveform which emerges when a mass and a spring oscillate can be described by
its amplitude (D) and period (T). Frequency is defined as the number of cycles per second and is
equal to the reciprocal of the period. By multiplying the frequency by 2 the angular frequency is
obtained, which is again proportional to the square root of the spring constant k divided by mass
m. The frequency of oscillation is called the natural frequency fn. The whole sine wave can be
described by the formula d = D sin nt, where d = instantaneous displacement and D = peak
displacement.

Figure 5.16 Simplest form of a vibrating system.

Figure 5.17 shows the same spring-mass model of a vibrating system in a time-lapse sequence
(maximum displace vertically on far left and far right). If damping is ignored, there is a perfect
conversion of kinetic energy (at neutral displacement there is maximum velocity and maximum
kinetic energy) to potential energy (at maximum displacement there is zero velocity and
maximum potential energy) and back twice per vibration cycle.

52
Figure 5.17 Spring-mass model showing energy transfer within vibrating system.

Figure 5.18 highlights the relationship between the natural frequency and system mass. Because
the mass term is in the denominator of the natural frequency equation, as the mass increases the
natural frequency decreases (and the period of the oscillations increases). Conversely, as the
mass decreases the natural frequency will increase (and the period of the oscillations will
decrease).

Figure 5.18 Spring-mass model showing change in natural frequency with change in system
mass.

5.5 Basic Spring-Mass-Damper Model of Vibration

A slightly more realistic model of one degree-of-freedom vibrating systems includes a damping
term as well as the spring and mass terms. The damper extracts energy from the vibrating
system. This reduces the vibration amplitude and slightly reduces the frequency of vibration.
Figure 5.19 shows a schematic diagram of a spring-mass-damper model of a one degree-of-
freedom vibration system.

53
F(t)
x(t)

Mass M

C K

Figure 5.19 Spring-mass-damper model of a vibrating one degree-of-freedom system.

The equation of motion (x(t) = the displacement of the system as a function of time) comes from
the force balance equation (equation 5.3),

Mx(t ) Cx (t ) Kx (t ) F (t ) 5.3

where C is the damping term which is proportional to the velocity of the vibrating mass.

The total solution to the equation of motion has two parts. The transient solution (x1) and the
steady state part (x2). We are usually more interested in the steady state solution, but will
consider both here for completeness.

The transient state solution applies for the case where the applied force equals zero ( F (t ) 0 )
and the system responds freely. This is referred to as the general form solution.

x1 (t ) Ae s1t Be s2t 5.4

S1,2 2 1 0 5.5

In equation 5.4 the terms A and B are initial conditions.

In equation 5.5 the term 0 K M is the natural frequency, and C 2 M is the damping
0
ratio.

There are three special cases of transient vibration.

1. Underdamped 1
2. Critically Damped 1
3. Overdamped 1
For the underdamped case ( 1 ) the response is defined as,

54
x1 (t ) Ae 0t sin( 1 2 0t ) 5.6

Oscillations occur with frequency 0 and the amplitude decays exponentially as a function of
time (See Figure 5.20).

For the critically damped case ( 1 ) the response is defined as,

x1 (t ) ( A Bt )e 0t 5.7

There is a quick restoration to equilibrium state with no oscillations (See Figure 5.20).

For the overdamped case ( 1 ) the response is defined as,

x1 (t ) Ae s1t Be s2t 5.8

In equation 5.8 the terms s1 and s2 are real numbers. The equation defines an exponential
amplitude decay without oscillations (See Figure 5.20).

2
Underdamped, =0.1
Critically damped, =1.
1.5 Overdamped, =3.

x(t) 0.5

-0.5

-1
0 0.5 1 1.5 2 2.5 3
Time (s)

Figure 5.20 Comparison of underdamped, critically damped and overdamped spring-mass-


damper model responses.

In all underdamped cases it should be noted that the decay of the vibration oscillations changes
with changing damping. That is, the decay rate increases with increased damping and decreases
with decreased damping. The natural frequency (also referred to as the damped natural
frequency) also changes, albeit only slightly, with changing damping (decreasing and increasing

55
damped natural frequency with increasing and decreasing damping respectively). Figure 5.21
shows these relationships.

Figure 5.21 Spring-mass-damper model showing change in decay rate (and slight change in
damped natural frequency) with a change in system damping.

The steady state solution to the equation of motion ( when F (t ) F0 sin(t ) ) can be written in
the form,

F0
x2 (t ) sin(t ) 5.9

C K 2 M
The total solution to the equation of motion is then the sum of these two partial solutions.

x(t ) x1 (t ) x2 (t ) 5.10

While the transient response is useful for describing and understanding the dynamic behaviour of
a system due to the physical characteristics of the system (spring stiffness, mass and damping),
the primary interest here is in the steady state response of the system due to some continuous
forcing function input. This is the response that is typically generated by a mechanical system
during operation (due to rotating or reciprocating masses, and/or the interaction of mechanical
components).

Recall that the equation that describes simple harmonic motion is:

x(t ) A sin(t ) 5.11

This is of course the same equation as was introduced earlier (equation 5.1) and describes the
displacement of a reference point of a system in response to some forcing function.

56
5.5.1 Example: Car Model

To understand how mechanical systems respond to forcing function inputs, consider an everyday
example the motor vehicle. A wide range of different inputs can cause vibrations in motor
vehicles. These inputs include external factors such as the wind the road surface as well as
internal factors such as fuel combustion, mechanical unbalance, engine fan noise, and
misalignment of various components. All vibrations experienced by the driver and other
occupants are the result of mechanical dissipation of energy in response to some forcing function
input.

Consider for now only one source of potential forcing function input the road surface. Also
consider the vehicle suspension system as a linear single degree-of-freedom system. Using the
single degree-of-freedom system model for the suspension system the suspension can be
modeled as shown in Figure 5.22. For the sake of simplicity, and because all the individual
wheels and associated suspension components are the same for each wheel can be considered as
identical for most common cars, Figure 5.22 represents only one wheel and associated
suspension system also known as a car model.

Assuming that the unsprung mass (the wheel and tire) is small (but not negligible) compared to
that of the vehicle, K represents the spring stiffness (linear). The spring stores energy when
stretched or compressed and acts to oppose motion proportional to position. An unstretched or
uncompressed spring suggests that no force is acting on the system. The parameter C is the
damping coefficient of the shock absorber, which is modeled as a viscous damper. The shock
absorber dissipates energy rather than storing it and opposes motion proportional to velocity.
When the velocity is zero, there is no force acting through the damper.

y(t) output
(vehicle vibration)
Mass of
vehicle, M

C shock damping K spring stiffness


x(t) input
(road surface)
F(t)

Figure 5.22 Single degree-of-freedom system model for a car suspension ( car model)

If we consider the force input from the road surface to be a sinusoidal signal of a fixed
frequency, as shown in Figure 5.23(a), then the output will also be a single frequency sinusoidal
signal, as shown in Figure 5.23(b). The frequency of the two signals will be the same, but they
will potentially be shifted in phase and of different amplitude. The amount of phase shift and
amplitude difference between the input and output signals is a function of the system structural
characteristics and the frequency of the input excitation.

57
Road Input Vehicle Output

Ampl. Ampl.

Time Time
(a) (b)

Figure 5.23 Single degree-of-freedom system model for a car suspension ( car model)
(a) is a single frequency input excitation, (b) is the single frequency output.

Evaluation of Figure 5.23 reveals two important quantities that define the relationship between
the input excitation and the output signal gain and phase shift. Gain is the change in amplitude
(usually a decrease, but occasionally an increase) from input excitation to output signal (often
expressed in decibels). Gain is defined as the ratio of output amplitude to input amplitude. The
phase shift is the change in the position of the output vibration signal relative to the input
vibration signal. The frequency of the output does not change relative to the input. Figure 5.24
shows the gain and phase shift relationship between input excitation and output. Notice in this
case the gain is less than one (1) and the amplitude of the output is smaller than the input. If the
gain were larger than one (1) the amplitude of the output would be larger than the input.

Road Input Vehicle Output

Gain
Ampl.
Ampl.

Time Time
Phase Shift

Figure 5.24 Gain and phase shift relationship between input excitation and output.

Consider now the gain and phase shift relationship between input excitation and output for a
system over a range of frequencies. In order to do this we need to introduce what is known as the
Transfer Function (TF). Figure 5.25 a graphical representation of a transfer function
relationship over a range of frequencies for a particular mechanical system. The gain and phase
shift at any particular frequency may be found from these plots.

58
0
Gain Phase
(dB) 1 (degrees) 90

180
Freq. Freq.

Figure 5.25 Gain and phase relationship between input excitation and output over a range of
frequencies.

While the amplitude scale is not shown in Figure 5.25 or Figure 5.26 the gain at low frequencies
is one or close to one and the phase shift is zero (see Figure 5.26(a)). There is little change in the
gain as the frequency increases, until the system natural frequency is approached when the gain
quickly increases with increasing frequency. The phase shifts approaches 90 as the frequency
gets close to the natural frequency (see Figure 5.26(b)). Above the natural frequency, the gain
decreases at a constant rate (usually rapidly). The phase shift approaches 180 as the frequency
continues to increase above the first natural frequency (see Figure 5.26(c)).

It needs to be noted now (and it will be discussed in more detail later) that there may be more
than one natural frequency in a given mechanical system. In all cases, low frequency inputs are
passed through the system (gain equals one) while high frequency inputs are attenuated. Such a
system is acts as a low pass filter. All mechanical systems act as low pass filters for the
following reason. High frequencies require higher speeds to reach the same amplitudes as lower
frequencies. All machines have a maximum velocity (due to inertia). Once the maximum
velocity is reached, higher frequencies can only be reached by reducing the amplitude.

Gain Gain Gain


(dB) (dB) (dB)

Freq Freq Freq

Phase Phase Phase


(degrees) (degrees) (degrees)

Freq Freq Freq

(a) (b) (c)

Figure 5.26 Gain and phase relationship between input excitation ((a) below the first natural
frequency, (b) at the first natural frequency, (c) above the first natural frequency).

The increase in gain and dramatic phase shift that occur at the natural frequency of the system
are referred to as a mechanical resonance. A mechanical resonance exists at (or close to) every

59
system natural frequency. As stated earlier systems may have several natural frequencies. Many
system responses or forcing function frequencies exist at or close to a resonance. It is essential to
consider the existence of these resonances when designing new machines and when maintaining
existing machines.

As noted earlier system damping affects the response of the system. The peak gain occurs
slightly below the system resonance frequency due to damping. Increased damping results in
lower peak gain. Increased damping also results in reduced phase shift slope. Changes in
damping result in only minimal changes in gain and phase shift at low and high frequencies.

Figure 5.27 shows the relationship between gain and of the ratio of input excitation to system
natural frequency. In general as the damping decreases the gain increases. In particular, at (and
close to) the natural frequency (where the ratio of input excitation to system natural frequency
equals 1) the gain grows to larger than one. This means that the output signal amplitude is
amplified by the system and becomes larger than the input excitation signal amplitude. This
situation can be dangerous in mechanical systems. Extreme amplitude output vibration due to
relatively small amplitude input excitation can quickly lead to severe loads exerted on
components. These high loads may result in inacceptable machine operating conditions and/or
abrupt catastrophic failure.

In order to see this type of system behavior first hand a simple example can be relatively easily
performed. Using a small weight (say a stapler) and a support spring (say a few thick elastic
bands tied together) it is possible to form a simple single degree-of-freedom system the stapler
supported by the elastic bands. By holding the free end of the elastic bands between ones fingers
the system can be excited at a variety of frequencies. First try an input of low frequency (your
hand moving up and down very slowly). The output (the movement of the stapler) should mimic
your hand (same amplitude of displacement and the same phase (moving up as your hand moves
up and moving down as your hand moves down). Now try an input of very high frequency (move
your hand up and down as quickly as possible, but with relatively low amplitude). You should
find that it is difficult to move your hand at high frequency and high amplitude (because of the
inertia of your hand and arm). The stapler should be moving only slightly (small amplitude
displacement) and any motion that does take place should be approximately 180 out of phase
from the input (your hand moves up as the stapler moves down and vice versa).

Now the tricky part, starting with a low frequency input, gradually increase the frequency. At a
certain frequency (the system natural frequency) the output signal amplitude should be far
greater than the input signal amplitude. With a relatively small amplitude input the system
responds with a significantly amplified output. This is the system resonance. If you watch
carefully you should see that the output (stapler motion) should be 90 out of phase with the
input excitation (your hand motion). Be careful, the large amplitude motion of the stapler will
put significant stress on the elastic bands.

60
Figure 5.27 Gain as a function of the ratio of input excitation to system natural frequency (note
that as the damping decreases the gain increases at (and close to) the natural frequency where
the ratio of input excitation to system natural frequency equals 1).

Typically only outputs can be measured, not inputs. To complicate matters, a different transfer
function exists from each vibration forcing function that relates an input point to an output
(measurement) point. Not all (if any) transfer functions are known due to their complex nature.
As a result analyzing transfer functions and inputs separately is extremely challenging.

Damping is usually modeled as linear. Using this model results in force approaching zero as
velocity slows. This is, of course, not true in real systems. In reality, the damping force levels off
as velocity approaches zero. This relationship is shown in Figure 5.28 below.

Damping
Force

Velocity

Figure 5.28 Damping force as a function of velocity. Damping force levels off as velocity
approaches zero.

In summary, vibration is the mechanical dissipation of energy in response to a mechanical input.


All mechanical systems act as low pass filters of vibration inputs. In a simple linear system, the
response to a sinusoidal input is a sinusoidal output with the same frequency, but different phase
and amplitude. A system response to vibration input depends on the frequency of the input. The
change in amplitude and phase shift of the output relative to the input is slight at low frequencies,
but is dramatic close to the system natural frequency (resonance) and above. In vibration analysis

61
it is essential to consider both the specifics of the input and the system characteristics (transfer
function) such as resonances and non-linearities.

5.6 Vibration Signal Fundamentals (Displacement, Velocity and Acceleration)

Equation 5.10 describes the total displacement of the system output (at a particular reference
point) as a function of input excitation and system characteristic parameters. Equation 5.11 (and
equation 5.1) describe the steady state displacement response, which is of more interest here. The
velocity (the rate of change of displacement with time, m/s) of the same reference point can be
described as the derivative of the displacement as a function of time (see equation 5.12).

v(t ) x (t ) A sin(t ) 5.12
2

The acceleration (the rate of change of velocity with time, m/s2) of the same reference point is
the derivative of the equation that describes the velocity (which is also the second derivative of
the equation that describes the displacement (equation 5.13).

a (t ) x(t ) 2 A sin(t ) 5.13

Figure 5.29 shows the relationships that exist between displacement, velocity and acceleration of
a single reference point. Note the different phase relationships and the relative amplitudes. The
phase of the velocity lags the displacement by 90 degrees and the acceleration lags the velocity
by 90 degrees (lags the displacement by 180 degrees). The amplitudes also vary due to the
appearance of frequency as a term that multiplies the maximum amplitude in the velocity
equation and frequency squared as a term that multiplies the maximum amplitude in the
acceleration equation.

The same relationships are shown in Figure 5.30 with the amplitudes of the displacement,
velocity and acceleration normalized. The absolute amplitudes still have the same relationship as
in Figure 5.29. Figure 5.31 shows the same relationships using integrals rather than derivatives.
These relationships are important as they allow measurements made using one set of instruments
to be easily converted to another (perhaps more appropriate) format.

62
0.25

0.2

0.15

0.1

0.05
Amplitude
0

-0.05

-0.1

-0.15
displacement
-0.2 velocity
acceleration
-0.25
0 5 10 15 20 25 30 35 40
Time (ms)

Figure 5.29 Graphical relationships between displacement, velocity and acceleration of a single
reference point.

Figure 5.30 Graphical and numerical (via differentiation) relationships between displacement,
velocity and acceleration (normalized amplitude).

63
Figure 5.31 Graphical and numerical (via intigration) relationships between displacement,
velocity and acceleration (normalized amplitude).

5.7 Vibration Descriptors

Vibration descriptors are single number quantities that are derived from a vibration signal and
can be used as a fast and easy way to compare signals. These signals can be either two signals
from different machines or two different locations on one machine or the same location sampled
at two different times.

5.7.1 Average Vibration Signal Strength

The simplest of these descriptors is the average of the signal, which, as the name implies, is the
average vibration level of the signal over a given period of time. The average is defined
mathematically in equation 5.14.
1
xav
T x(t ) dt 5.14

5.7.2 Peak Amplitude

The peak value (zero-to-peak) term indicates the peak vibration level of the signal measured
from zero to the maximum positive of negative value of the vibration signal. The peak value is
defined in equation 5.15.

x p absmax x(t ) x 5.15

5.7.3 Peak-to-Peak Amplitude

The peak-to-peak value indicates the total maximum fluctuation in the vibration signal.
This value is defined in equation 5.16. This value is particularly useful when determining the
total overall movement of an object that is vibrating. It is usually used to describe the maximum

64
displacement of a rotating shaft inside a journal bearing. In this case the maximum total
displacement is important as it indicates how close the rotating shaft is to exceeding the
clearance within the bearing (and potentially causing a rub).

x p p maxx(t ) minx(t ) 5.16

5.7.4 Root Mean Square (RMS)

The RMS (root mean square) value is defined in equation 5.17 and is useful because it is
proportional to the energy in the vibration signal. The relationships between these parameters
(for a sinusoidal waveform) are shown in Figure 5.32.

1
x(t ) dt
2
xRMS 5.17
T

xp
Amplitude x p p
xav xRMS

time

Figure 5.32 The relationships between various vibration signal descriptors (for simple harmonic
motion).

With respect to the amplitude of a simple harmonic motion signal, A, in the original equation of
simple harmonic motion:

x(t ) A sin(t ) 5.11

The RMS value can be defined as that shown in equation 5.18.

xRMS 0.707 A 5.18

In this case the peak value is x p A , the peak to peak value is x p p 2 A and the average
amplitude is xav 0.637 A .

It must be noted that equation 5.18 and the equalities defined immediately above are true for
simple harmonic motion only. If the vibration signal has a different character (periodic, but not

65
simple harmonic motion) the simplification above does not hold. The RMS value in such a case
would not equal 0.707 A. Rather, the RMS value must be calculated directly from equation 5.17.
Figure 5.33 shows the relationships between peak value, peak to peak value and the RMS for a
signal that is not simple harmonic motion. Such a signal more accurately represents what is
likely to be measured from a rotating machine.

Figure 5.33 The relationships between various vibration signal descriptors (random or periodic
(non-simple harmonic) motion).

It should be noted that some dynamic signal analyzers will use the simplification noted above in
the estimation of Peak, Peak-to-Peak and RMS values. These calculations will result in under
estimates of the Peak and Peak-to-Peak values based on RMS or the over estimate of RMS based
on Peak or Peak-to-Peak values. As an example of how these approximations may creep into the
vibration signal analysis Figure 5.36 to Figure 5.42 show the standard signal analysis that may be
carried out on data collected from vibrating machines.

5.7.5 Crest Factor

Another numerical descriptor based on the time waveform that is useful for indicating the
character of the vibration signal is the Crest Factor. The Crest Factor is the ratio of the Peak
value to the RMS value. This ratio is sensitive to the impulsive nature of a vibration signal (the
amount of short duration, high amplitude spikes in the signal as related to the overall signal
strength). As spikes appear in the signal the Peak value increases relative to the RMS value,
which will not change significantly until the number of spikes per time interval increases
significantly. These few spikes indicate the earliest stages of failure, (typically in rolling element
bearings). As the impulsive nature of the signal begins to dominate the signal, the Crest Factor
value decreases (the RMS level increases relative to the high, but relatively constant Peak level).
As such, the Crest Factor is a good indicator of the early stages of rolling element bearing
failure, but not a good long term trending parameter. Figure 5.34 shows the change in the value
of the Crest Factor parameter as the Peak value and the RMS value of a vibration signal change
with time.

66
Figure 5.34 Changing values of the Crest Factor relative to changing values of Peak and RMS.

Figure 5.35 shows a schematic of a vibrating machine that is producing a signal that closely
approximates simple harmonic motion. It should be noted that this signal is not uncommon in
simple machines (rotors in fluid filled bearings without other sources of vibration noise close
by). Figure 5.36 shows a sample signal analysis screen from a standard data analysis instrument
(more detail will be provided on instrumentation and dynamic signal analysis in a later chapter).

X = Time
Y = Amplitude

Figure 5.35 Schematic of a rotor in a bearing housing and the vibration response (simple
harmonic motion).

The frequency of the signal is the inverse of its period (T). The period of time from the vertical
set mark line to the cursor is (recorded at the lower right of the plot above): DTIM = 22.46
milliseconds, which equals 0.02246 seconds. Therefore, the equation below can be used to
calculate the frequency. This value is the same as that generated by the analysis instrumentation,
as seen in the bottom right hand corner of Figure 5.36 (FREQ: 44.52).

1 1
F 44.5Hz 2671 CPM 5.19
T 0.02246

67
Figure 5.36 Screen capture from a standard vibration signal analysis instrument showing period
of dominant frequency which is used to calculate frequency.

Figure 5.37 shows the same vibration signal with details of the displacement calculation. The
Peak-to-Peak displacement is shown in the upper right hand corner of the plot and also shown as
the calculated difference between the maximum negative displacement (-1.902 mils) and the
maximum positive displacement (+1.917 mils) for a total of 3.819 mils (or 0.003819 inches)
Peak-to-Peak displacement. Recall that one mil is equal to one one-thousandth of one inch (1 mil
= 0.001 inches) or 1/40 mm (1 mil = 0.025 mm). The time required for this movement is 11.72
msec (or 0.0117 seconds). The average velocity during this interval is then calculated as:

displacement 0.003819
V 0.326 in / sec 5.20
Time 0.0117

Figure 5.37 Screen capture from a standard vibration signal analysis instrument showing the
maximum positive and negative peak values.

The value calculated for the velocity is actually the average velocity determined from the total
time taken for the displacement to move from the maximum negative to the maximum positive

68
position. In reality, the velocity would be zero at the extremes (maximum negative and
maximum positive displacement) and a maximum velocity (somewhat greater than the average)
would exist at the neutral (At Rest) position (as shown on Figure 5.38).

When a Fast Fourier Transform is performed (more on the details of FFT analysis in a later
chapter) on the waveform shown above the spectrum shown in Figure 5.39 results. The
amplitude of the primary frequency matches the Peak-to-Peak amplitude of the waveform. This
result is due to the fact that the time waveform is a simple harmonic motion waveform. This
exact relationship does not hold true if the time waveform is periodic, but not simple harmonic
motion (as will be shown in the following figures).

Figure 5.38 Screen capture from a standard vibration signal analysis instrument showing the
maximum positive and negative displacements and the locations of maximum velocity.

Figure 5.39 Screen capture from a standard vibration signal analysis instrument showing the
frequency spectrum of the displacement time waveform shown in Figure 5.36.

69
Figure 5.40 shows a time waveform of a periodic vibration signal (which is not simple harmonic
motion). The value underlined in the upper right hand corner of the plot above is the RMS value.
It resembles the value that would be measured with an AC voltmeter and represents an accurate
estimate of the true RMS value for this dynamic signal.

Figure 5.41 shows the same time waveform as in Figure 5.40, but the Peak value (underlined in
the upper right hand corner of Figure 5.41) is the RMS value multiplied by 1.4141. This value is
referred to as an RMS peak value and as can be seen from the figure is not equal to the true Peak
value. It is an approximation based on the (in this case incorrect) assumption that the time
waveform can be approximated as simple harmonic motion. It can be clearly seen in both Figure
5.40 and Figure 5.41 that the true Peak value is closer to 0.4 mils.

Figure 5.40 Screen capture from a standard vibration signal analysis instrument showing the time
waveform of a periodic vibration signal (not simple harmonic motion) and calculated true
RMS value.

Figure 5.41 Screen capture from a standard vibration signal analysis instrument showing the time
waveform of a periodic vibration signal (not simple harmonic motion) and calculated RMS Peak
value.

70
The Peak-to-Peak value shown in the upper right hand corner of Figure 5.42 is the RMS value
multiplied by 2.828. This value is referred to as an RMS Peak-to-Peak value. The true Peak-to-
Peak value exceeds 0.8 mils (as shown on the plot). Therefore, remember that RMS based Peak
and Peak-to-Peak values do not represent the true Peak and Peak-to-Peak values. Despite the fact
that many analyzers display the Peak and Peak-to-Peak values calculated from all periodic
waveforms in this format, pure sine waves (simple harmonic motion) are the only exception to
this rule. The same holds true for Peak and Peak-to-Peak values calculated from frequency
spectra that have been calculated from periodic (but not simple harmonic motion) waveforms.

Figure 5.42 Screen capture from a standard vibration signal analysis instrument showing the time
waveform of a periodic vibration signal (not simple harmonic motion) and calculated RMS Peak-
to-Peak value.

5.8 Frequency Response Functions and Natural Frequencies

From the discussion above it should be clear that all vibration signals are the result of two
factors. These are the physical nature of the system (mass, stiffness, damping) and the type of
excitation that is driving the excitation. Two simple models of vibrating systems were presented
(the mass-spring model and the mass-spring-damper model) and two basic excitation responses
were discussed (free and forced vibration). Before going on to discuss in more detail the types of
responses to all the different types of forcing functions that may be generated by machines, a
more complete description of the response of systems based on the inherent physical parameters
will be presented.

Figure 5.43 shows a simplified version of a frequency response function (FRF) (middle right),
the phase shift (lower right) and several sample time waveform responses (upper right) at
different frequencies for a single degree-of-freedom mass-spring-damper model of a mechanical
system (left). This figure shows that the response of a system in the lower frequencies (below the
system natural frequency) is in-phase with the forcing function and of equal amplitude to the
forcing function (location (a) on the Frequency axis). The response of a system in the higher
frequencies (above the system natural frequency) is 180 out-of-phase with the forcing function
and of significantly reduced amplitude to the forcing function (location (c) on the Frequency

71
axis). The response of a system at or close to the system natural frequency is out-of-phase by 90
with the forcing function and of significantly amplified amplitude to the forcing function
(location (b) on the Frequency axis).

This type of system response is typical of all mechanical systems. This type of response of also
referred to as a low-pass filter. All the lower frequencies are passed through the system
unaffected in terms of frequency and amplitude, while high frequencies are filtered out
(significantly reduced in amplitude). All systems respond at the natural frequency with amplified
response vibrations. However, systems with relatively low levels of damping may have
dangerously high response levels that could quickly damage machinery due to the high forces
developed and/or the high displacements (see Figure 5.44). For this reason, exciting natural
frequencies (resonance frequencies) by operating machines at or close to a shaft rotational speed
equal to the natural frequency should be avoided. These speeds are referred to as critical
speeds and will be discussed in more detail in a later chapter.

(a) (b) (c)

Figure 5.43 Frequency response function, phase shift and sample response waveforms of a single
degree-of-freedom mass-spring-damper model.

Figure 5.44 Relative response amplitudes for varying levels of system damping.

Because few machines actually respond as single degree-of-freedom systems, frequency


response functions for multiple degree-of-freedom systems are also useful in defining the

72
response of a system to input excitations. Figure 5.45 shows the frequency response function,
phase shift and sample response waveforms of a two degree-of-freedom mass-spring-damper
model. The overall response is a combination of the two single degree-of-freedom systems and
results in two natural frequencies and a two-stage phase shift. Figure 5.46 shows a multi degree-
of freedom system frequency response function (with multiple natural frequencies).

Figure 5.47 shows a multi degree-of freedom system frequency response function representing a
model of a rotor mounted on two bearings which are in turn mounted on a foundation that is
separate from the fixed earth. It should be noted that this type of model representation and
estimation of the frequency response function has significant potential to explain the measured
vibration response of a system to internally or externally generated forcing functions. The
measured vibration response will always be a combination of the input excitation and the system
response (as shown in Figure 5.48).

Figure 5.45 Frequency response function and phase shifts of a two degree-of-freedom mass-
spring-damper model.

Figure 5.46 Frequency response function of a multiple degree-of-freedom mass-spring-damper


model.

73
Figure 5.47 Frequency response function of a multiple degree-of-freedom mass-spring-damper
model representing a rotor on two bearings on a foundation.

Figure 5.48 Total vibration response is a sum of the input excitation forces and the characteristic
system frequency response.

The total vibration response from an operating machine is usually dominated by machine
operating characteristic frequencies generated internally. These frequencies are typically closely
related to specific components inside the machine and the operating speed of the machine. These
relationships allow for the use of frequency spectra to detect and diagnose machine condition
using a vibration signal. Only when one of these frequencies overlap with a system natural
frequency will a resonance be generated. If the operating speed happens to match a natural
frequency, this speed is referred to as a critical speed (as mentioned earlier). Well designed
machines should not excite resonances, but changes in the design on site, changes to the machine
and changes to the machine mounting structure can all change the natural frequencies of a
system. A previously safe operating speed may then become one that excites a resonance.

Figure 5.49 Total vibration response is usually dominated by machine operating characteristic
frequencies.

74
5.9 Time Waveform versus Frequency Spectra Analysis

As is evident from the figures above, the time waveform and the frequency spectra can be used
to determine vibration signal characteristics that help define the excitation that is causing the
vibration as well as the system characteristics that are also part of the response. There are
situations where one or the other of this two vibration data presentation formats should be
favoured. This will be discussed in more detail in a later chapter, but for now an example of a
vibration signal that contains two simple harmonic motion signals at different frequencies will
illustrate the usefulness of looking a vibration data in both formats rather than relying on only
one in all cases.

The waveform in Figure 5.50 shows a simple harmonic motion signal with regularly varying
amplitude. This type of signal is referred to as a signal with amplitude modulation (or an
amplitude modulated signal). This signal is only slightly more complicated than those shown on
the preceding pages. The primary carrier signal is at a higher frequency and the modulating
signal is at a lower frequency. The time interval marked on Figure 5.50 shows 31.25 msec
(0.03125 sec) between carrier signal peaks. The frequency of this signal is determined by
dividing 1 by this interval time (the period of the simple harmonic motion), as shown in equation
5.21 below.

Figure 5.50 Simple harmonic motion signal with amplitude modulation (note carrier signal
period).

1 1
F 32.0 Hz 5.21
T 0.03125

The modulation of the carrier waveform has a period equal to the time between maximum peaks.
This time interval can be measured (as shown in Figure 5.51) at 252 msec (0.252 sec). The
frequency of the modulation signal is determined by dividing 1 by this interval time (the period
of the modulation signal), as shown in equation 5.22 below.

75
1 1
F 3.969 Hz 5.22
T 0.252

Figure 5.51 Simple harmonic motion signal with amplitude modulation (note modulation signal
period).

From the relatively simple analysis of the time waveform above the carrier frequency and the
modulation frequency were both determined. When considering the frequency spectrum of this
signal however (see Figure 5.52), the two frequencies of interest are not as clearly evident. The
dominant carrier frequency is obvious at 32 Hz, but the modulation signal frequency is not
obvious at or near 4 Hz. Instead the modulation frequency shows up most clearly as what are
referred to as sidebands on both sides of the carrier frequency. These sidebands are separated
from the main carrier frequency response by a span of 3.969 Hz (the approximate frequency of
the modulation signal).

Figure 5.52 Frequency spectrum of simple harmonic motion signal with amplitude modulation
(note frequency span separating sidebands).

76
Depending on the magnitude of the modulation signal relative to the carrier signal and the
resolution of the frequency spectrum, the sidebands may be difficult to observe. This suggests
that frequency spectra may not be the best tool for analyzing vibration signal that may contain
amplitude modulation. This topic will be reviewed and expanded upon again in a later chapter as
will the use of logarithmic amplitude scales to help highlight small amplitude frequency spectra
responses in the presence of high amplitude responses.

5.10 Overall Levels versus Frequency Spectra Analysis

Another example worth considering shows the potential pitfalls associated with only monitoring
overall vibration levels of the time waveform or the frequency spectra as compared to monitoring
the amplitudes of different frequencies within the frequency spectra. As mentioned above (and as
will be discussed in detail in a later chapter) the overall level (the Peak value for example) of a
vibration signal maybe a good indicator of overall machine condition. However, reliance on any
one overall measure of machine condition may not suit all applications.

Figure 5.53 shows a schematic of the signal processing paths that would be followed when
generating overall levels and frequency spectra from a vibration signal. Using modern
instrumentation the calculation of the frequency spectra takes no more time and only a small
amount of additional effort (during instrument setup). The results, however, may require more
time to interpret, hence the reliance on single number indices for overall monitoring.

Figure 5.53 Schematic of the signal processing paths that would be followed when generating
overall levels and frequency spectra from a vibration signal.

Figure 5.54 shows two different applications where both overall levels and frequency spectral
analysis may be applied. In the case of the fan it is clear that the increasing amplitude of the
frequency component that is linked to the dominant fault frequency in the fan is the dominant
amplitude in the entire frequency spectrum. Therefore, monitoring the overall level is
automatically linked to the level of the fault characteristic frequency and will give a good
indication of the graduate deterioration of the machine. In the case of the gearbox, initially the
dominant frequency in the spectrum is not linked to the frequency of most interest. The mid-
frequency vibrations (perhaps from a bearing) start out low and then only gradually increase to a
level above the originally dominant lower frequency components (shaft rotation frequency) and
higher frequency components (gear mesh vibration frequencies). The frequency component that

77
is linked to the bearing fault only affects the overall level magnitude after the vibration level at
that frequency has increased significantly above the original level at that frequency. In this case
monitoring only the overall level would result in a relatively late warning of the gradually
deteriorating condition of the bearings. This could result in insufficient warning of an incipient
fault developing, or no warning at all.

Figure 5.54 Frequency spectral analysis versus overall level monitoring in different applications.

Chapter Summary

Vibration (even when very small in amplitude) is often a destructive and annoying side effect
of a useful process.
The study and description of mechanical vibrations covers a wide range of topics.
Mechanical vibration analysis provides a useful insight into mechanical system response and
condition for a variety of reasons. It is non-intrusive, can be performed on-line, is applicable
to a broad range of machinery and machine components, is applicable to a broad range of
fault types, and the associated hardware used to measure and record mechanical vibrations is
relatively economical and reliable.
Mechanical vibration signals can be classified based on the type of motion that is present
(simple harmonic, periodic, random, transient).
Mechanical vibration signals can be classified based on the type of excitation that caused the
vibration (free vibration, forced vibration, self-induced vibration).
Vibration waves can be defined in terms of amplitude, frequency and phase.
Spring-mass and spring-mass-damper models are a useful tools for describing the dynamic
response of mechanical systems.
The basic elements of vibratory motion are displacement, velocity and acceleration. These
parameters are related in a relatively simple way that allows the calculation of any parameter
from any other such that the measurement of one can be used to determination the others.

78
Here are several basic vibration signal time waveform descriptors (Peak value, Peak-to-Peak
value, RMS value, Crest Factor) that allow the definition of most simple waveforms.
Mechanical vibrations provide a means of obtaining useful insight into the condition of
machines because they are related to both the system characteristics (mass, stiffness,
damping) and the excitation forces being applied.
The frequency spectra, calculated from the time waveform, may provide useful insight into
the mechanisms that are responsible for exciting the responses seem in the vibration signal.
Care should be exercised when analyzing vibration signals as useful information is usually
contained in both the time waveform and the frequency spectra.

79
Chapter 6

Basic Vibration Signal Analysis


How much information is there in a vibration signal? It depends on how you look at it.

Vibration signals can be displayed in a variety of different formats. Each format has advantages
and disadvantages, but generally the more processing that is done on the dynamic signal, the
more specific information is highlighted and the more extraneous information is discarded. The
display formats that will be discussed here are the time domain, the frequency domain, the modal
domain, and the quefrency domain. Within each of these display formats, several different
analysis tools (some specific to that display format) will be described.

6.1 Time Domain Signal Analysis

The time domain refers to a display or analysis of vibration data as a function of time. The
principal advantage of this format is that little or no data are lost prior to inspection. This allows
for detailed analysis. However, the disadvantage is that there is often too much data for easy and
clear fault diagnosis. The time-domain analysis of vibration signals that will be discussed here is
subdivided into the following sections; Time Waveform Analysis, Time Waveform Indices,
Probability Density Functions, and Probability Density Moments.

6.1.1 Time-Waveform Analysis

Time-waveform analysis involves the visual inspection of the time-history of the vibration
signal. The general nature of the vibration signal can be clearly seen and distinctions made
between sinusoidal, random, repetitive, and transient events. Non-steady-state conditions, such
as run-up and coast-down, are most easily captured and analyzed using time waveforms. High-
speed sampling can reveal such defects as broken gear teeth and cracked bearing races, but can
also result in extremely large amounts of data being collected, much of which is likely to be
redundant and of little use.

Figure 6.1 shows the time waveform signal from a pump bearing. The signal shows a highly
impulsive signal with relatively high overall signal energy (velocity, inches per sec.). The signal
is somewhat repetitive, but shows no signs of a dominant underlying shaft rotational component.
This signal suggests that there could be some significant bearing deterioration already in
developed. The amplitude levels could be used to compare the peak vibration amplitude (or the
RMS level) to standards or acceptance limits to determine the state of the bearing. (More on
standards and acceptance limits in a later chapter.) The signal also shows that a detailed analysis
of the condition of the equipment being monitored may not be possible because of the large
amount of data in the signal.

Figure 6.2 shows the time waveform signal from a rolling element bearing. The signal again
shows a highly impulsive signal. This plot shows the vibration signal measurement in Gs
(acceleration). The signal is more obviously repetitive, but again shows no signs of a dominant
underlying shaft rotational component. This signal suggests that there could be some early onset

80
bearing deterioration developing. The peak amplitude levels are relatively low, but should be
compared to standards or acceptance limits to determine the state of the bearing. The obvious
intermittent spikes followed by ringing suggests that the bearing rolling elements are interacting
with race faults (cracks or pits). The signal shows that a detailed analysis of the condition of the
equipment is possible, but needs to be confirmed with other data or comparison to other analysis
results.

Figure 6.1 Time waveform of pump bearing signal.

Figure 6.2 Time waveform of rolling element bearing signal

Figure 6.3 shows three samples of time waveforms that represent vibrations of a shaft in a
journal bearing. In these cases there are no (or at least minimal) background vibrations. Here is
some background noise, but the dominant vibration frequency is clearly at one times the
rotational frequency of the shaft. Figure 6.3(a) shows one dominant frequency which may be due
to shaft unbalance. Figure 6.3(b) and Figure 6.3(c) show a dominant one times rotational speed
frequency and also a once per revolution secondary frequency. In Figure 6.2(b) the secondary
frequency is 180 out of phase with the primary frequency and it Figure 6.2(c) the secondary
frequency is approximately 90 out of phase with the primary signal. These responses are more
suggestive of a misalignment problem rather than unbalance. These figures show how the time
waveform, in some situations, can be a useful fault detection and diagnostic analysis tool.

81
(a) (b) (c)

Figure 6.3 Time waveforms showing vibration signals taken from a shaft in a journal bearing.

Figure 6.4 shows two views of the same vibration time waveform. The upper view shows several
seconds and the lower view focuses in on just a few hundred milliseconds. Both views offer
opportunities to extract useful information about the signal under consideration. The longer time
view shows a peak modulation (also known as beating, to be discussed in detail in a later
chapter) with a repetition rate of about every of a second. There are two possible causes for
this response. It may be due to the interaction of two vibration sources that are close in
frequency, but not exactly at the same frequency. This response may also be due to a single
excitation that is being modulated by a significantly lower frequency signal.

The shorter time waveform shows a spike that occurs once per revolution of the shaft. Close
inspection shows that this spike does not occur every revolution, but seems to come and go. This
may suggest a rolling element bearing fault on the inner race. Each time the fault passes through
the load zone the spikes show up in the vibration signal. When the fault is not in the load zone
the spikes are not present.

Figure 6.4 Time waveforms showing vibration signals of different time lengths showing different
levels of detail.

82
6.1.2 Time-Waveform Indices

As discussed in the previous chapter under the heading of vibration descriptors, time-waveform
indices are single number values calculated in some way based on the raw vibration signal and
used for trending and comparisons. These indices significantly reduce the amount of data that is
presented for inspection, but highlight differences between samples. Examples of time-
waveform-based indices include the average signal amplitude (also referred to as the mean signal
amplitude, see Chapter 5.7.1), peak level (maximum vibration amplitude within a given time
signal, see Chapter 5.7.2), peak-to-peak level (maximum positive to maximum negative vibration
amplitude within a given time signal, see Chapter 5.7.3), root-mean-square (RMS) level (reduces
the effect of spurious peaks caused by noise or transient events, see Chapter 5.7.4), and the Crest
Factor which is the ratio of the peak level to the RMS level (indicates the early stages of rolling-
element-bearing failure, see Chapter 5.7.5). All of these measures are affected adversely when
more than one machinery component contributes to the measured signal.

6.1.3 Statistical Methods and Parameters

Probability Density Functions

The probability of finding the instantaneous amplitude value from a vibration signal within a
certain amplitude range can be represented as a probability density function. Typically, the shape
of the probability density function in these cases will be similar to a Gaussian (or normal)
probability distribution. Fault conditions will have different characteristic shapes. Figure 6.5
shows two probability density functions. One is characteristic of normal machine operating
conditions, and the other represents a fault condition. A high probability at the mean value with a
wide spread of low probabilities is characteristic of the impulsive time-domain waveforms that
are typical for rolling-element-bearing faults. This type of display format can be used for
condition trending and fault diagnostics.

Faulty Bearing
Probability Density (dB)

Normal Bearing

Normalized Vibration Amplitude

Figure 6.5 Normalized vibration amplitude vs. probability density (normal and faulty bearings).

Probability Density Moments

83
Probability density moments are single-number indices (descriptors), similar to the time-
waveform indices except they are based on the probability density function. Odd moments (first
and third, mean and skewness) reflect the probability density function peak position relative to
the mean. Even moments (second and fourth, standard deviation and kurtosis) are proportional to
the spread of the distribution. Perhaps the most useful of these indices is the kurtosis, which is
sensitive to the impulsiveness in the vibration signal and therefore sensitive to the type of
vibration signal generated in the early stages of a rolling-element-bearing fault. Because of this
characteristic sensitivity, the kurtosis index is a useful fault detection tool. However, it is not
good for trending. As a rolling-element-bearing fault worsens, the vibration signal becomes more
random, the impulsiveness disappears, and the noise floor increases in amplitude. The kurtosis
increases in value during the early stages of a fault, and decreases in value as the fault worsens.

Crest Factor versus Kurtosis

The Crest Factor and kurtosis parameters are both good for detection of faults that typically
generate large amplitude spikes in the vibrations signal at regular intervals, but not regularly
enough to change the overall average (or RMS) level. They also share the same poor trending
ability. That is, as the fault worsens the values of the Crest Factor and the kurtosis typically
decrease in value after an initial increase. This makes trending using these parameters
untrustworthy. The only slight advantage to using the Crest Factor as a measure over the kurtosis
is that Crest Factor meters are available as stand alone survey tools, while the kurtosis parameter
needs to be calculated as part of a standard set of statistical paramenters and is not typically
offered as a stand alone meter.

6.2 Frequency Domain Signal Analysis

The frequency domain refers to a display or analysis of the vibration data as a function of
frequency. The time-domain vibration signal is typically processed into the frequency domain by
applying a Fourier transform, usually in the form of a fast Fourier transform (FFT) algorithm.
The principal advantage of this format is that the repetitive nature of the vibration signal is
clearly displayed as peaks in a frequency spectrum at the frequencies where the repetition takes
place. This allows for faults, which usually generate specific characteristic frequency responses,
to be detected early, diagnosed accurately, and trended over time as the condition changes.
However, the disadvantage of frequency-domain analysis is that a significant amount of
information (transients, non-repetitive signal components) may be lost during the transformation
process. This information is non-retrievable unless a permanent record of the raw vibration
signal has been made. There are also some serious errors that may take place during the FFT
process that need to be considered. More on this later in this chapter.

Figures 6.6 shows graphically the basic concept behind frequency spectral analysis. It is well
understood that several sinusoidal waves can be summed together and will result in a more
complex waveform. In 1827 Joseph Fourier (a French engineer) discovered that in theory any
waveform could be constructed by a potentially infinite series of pure sinusoid signals summed
together. Conversely, any waveform, regardless of how complex can be separated into a
(potentially infinite) series of pure sinusoids. This series of sinusoids that result from the
decomposition of any waveform is known as a Fourier series. The Fourier transform is then the

84
transformation of a series of sinusoids into a complex waveform or vise versa. In 1965 two
geophysicists (of J. W. Cooley and J. W. Tukey) derived a formulation that allows for the
Fourier transform to be calculated with a drastically reduced number of computations. Their
algorithm is known as the Fast Fourier Transform (FFT). When implemented digitally the FFT
can convert a dynamic signal to individual sinusoids almost instantly. When the power
(amplitude) of these individual sinusoids are plotted together with frequency on the horizontal
axis, it is known as a frequency spectrum.

Figure 6.6 Schematic showing how sinusoids can be combined to form more complex signals.

Figure 6.7 shows how the frequency content of a relatively simple dynamic signal can be
represented in a 3D plot with time, frequency and amplitude taking up the three axes. When this
information is viewed orthogonally facing the amplitude-time plane, the time waveform is seen
(the sum of all the individual frequencies). When this same information is viewed orthogonally
facing the amplitude-frequency plane, the frequency spectrum is seen (the power (amplitude) at
each individual frequency), as shown in Figure 6.8.

Amplitude Frequency

F2

F1

Time

Figure 6.7 Dynamic signals considered in 3D can be seen as having characteristics in the time,
amplitude and frequency domains.

85
Amplitude

F1 Frequency F2

Figure 6.8 In the frequency domain only the frequency and amplitude are highlighted.

Figure 6.9 shows graphically the process of moving from the time domain waveform to the
frequency domain (spectrum) via the FFT algorithm. As noted previously, any changes in the
amplitude of any individual frequency (or set of frequencies) over a period of time can be used to
assess machine condition. This is particularly true as many of the frequency components in a
spectrum are generated by particular machine components.

Figure 6.9 Schematic of the FFT process extracting the frequency content from a dynamic signal
and displaying it as a function of frequency and amplitude.

Figure 6.10 shows another example of several pure sinusoidal waveforms at selected frequencies
and the frequency spectra associated with each sinusoid. The frequency responses in Figures
6.10(a), (b) and (c) all show clear individual peaks at the associated frequency only with
amplitudes that reflect the amplitude of the pure sinusoid. When these individual sinusoids are
combined in Figure 6.10(a) the time waveform shows the individual waveform shapes overlaid
with one another and the frequency spectrum shows peak responses at the associated frequencies.

This example is a somewhat simplified case. All the sinusoids have an integer number of cycles
and therefore result in clear sharp peaks in the frequency domain. In most cases when dealing
with recorded dynamic signals this is not the case. This problem (known as leakage) will be
discussed in a later chapter.

86
Figure 6.10 Pure sinusoids and their associated individual frequency spectra (b, c, d) and the
same sinusoids combined with the associated frequency spectra (a).

Figure 6.11(a) shows a set of sinusoidal waveforms at different frequencies and amplitudes and
how they can be combined to generate an approximation of a square wave. Figure 6.11(b) and
Figure 6.12 show the same process in reverse. A square wave (and several other shapes) are
shown as the sum of many sinusoids. These constituent sinusoids can be separated out and
displayed as a series of frequency responses in a frequency spectrum. Note that a square wave is
made up of only odd harmonics of the primary frequency (fo in this case) that gradually diminish
in amplitude. A sawtooth wave is made up of odd and even harmonics that again gradually
diminish in amplitude.

(a) (b)

Figure 6.11 Component frequencies of a square waveform (a) and a sawtooth waveform (b).

87
Figure 6.12 Frequency components of several different waveforms.

Figures 6.13 and 6.14 show a time waveform and associated frequency spectra respectively that
represent the shape that might be expected when measuring vibration signal from machines in
the field. The time waveform shows a degree of relatively low frequency repetitiveness as well
as a large amount of high frequency content (potentially noise). It is difficult to see from an
inspection of the time waveform what particular machine components may be involved or what
condition may exist. The frequency spectra clearly shows several dominant frequency responses
at frequencies below 200 Hz. These are likely associated with particular machine components
and/or operations. There are also many lower amplitude frequency responses between 200 Hz
and 600 Hz. These may be associated with secondary operations in or near the machine or
components of the machine that are generating lower amplitude excitation signals.

It should also be noted that while many of the strongest peaks in the frequency spectra show
clear frequency distinction they also show some leakage. That is, they have slight shoulders or
slopes close to zero amplitude. This is particularly noticeable in the responses at the higher
frequencies where it seems there are many lower amplitude responses clustered together (see the
frequency responses from 200 Hz to 600 Hz. There are strategies that can be employed to reduce
the effects of leakage in the frequency domain and these will be discussed in detail in a later
chapter.

Figure 6.13 Vibration time waveform from a real machine.

88
Figure 6.14 Frequency spectra based on data presented in Figure 6.13.

As mentioned previously different machine components tend to generate specific frequency


responses when the machine is operating at a specific primary shaft rotational speed. Figure 6.15
shows these relationships graphically. The vibration signal recorded at any particular location on
the machine will be a resultant of all the internal (and potentially some external) excitations. The
sometimes quite complex waveform that results may be difficult to interpret in the time domain.
Using the FFT to extract the individual frequency components that make up the time waveform
allows for the different faults and other internal conditions that are linked to those frequencies to
be examined separately. Such is the strength of the frequency domain as a fault diagnostic tool.

Figure 6.15 The relationships between frequency responses and different machine components.

Figure 6.16 shows another graphical example of how the vibration signal taken from the bearing
supporting a rotating shaft will initially result in a relatively complex time waveform that can
subsequently be transformed into the frequency domain to reveal specific frequency responses

89
that are linked to different faults that exist or are developing in the machine. The excitation
forces are often the result of defects, wear of components, and/or design and installation errors.
Frequencies are used to relate machine faults to the forces that cause the vibration.

Figure 6.16 Time waveform and frequency spectral representations of vibration data measured
from a bearing supporting a rotating shaft.

Table 6.1 Forcing frequencies associated with machines.

Source Frequency (multiple of RPM)


Fault Induced
mass unbalance 1X (frequency in once per revolution)
Misalignment 1X, 2X
bent shaft 1X
mechanical looseness odd orders of X
casing and foundation 1X
distortion
antifriction bearing bearing frequencies, not integer ones
impact mechanisms multi-frequency depending on waveform
Design Induced
universal joints 2X
asymmetric shaft 2X
gear mesh (n teeth) nX
coupling (m jaws) mX
fluid-film bearings (oil whirl) 0.43X to 0.47X
blades and vanes (m) mX
reciprocating machines half & full multiples of speed, depending
on design

90
6.2.1 Band-Pass Analysis

Band-pass analysis is perhaps the most basic of all frequency-domain analysis techniques, and
involves filtering the vibration signal above and/or below specific frequencies in order to reduce
the amount of information presented in the spectrum to a set band of frequencies. These
frequencies are typically where fault characteristic responses are anticipated. Changes in the
vibration signal outside the frequency band of interest are not displayed.

6.2.2 Frequency-Domain Indices

It has been noted that frequency spectra are more sensitive to changes related to machine
condition (Mathew, 1987). Because of this sensitivity, several single number indices based on
the frequency spectra have been proposed. Like the time-waveform indices, frequency-domain
indices reduce the amount of information in frequency spectra to a single number. Because they
are based on the frequency spectra, they are generally more sensitive to changes in machine
condition than time domain indices. They are used as a means of comparing original spectra or
previous spectra to the current spectra. Several frequency domain indices are listed below.

1 N
Arithmetic Mean = 20 log Ai / 10 5
N i 1

Ai - amplitude of ith frequency spectrum component


N - total of frequency spectrum components

1 N Ai 5

Geometric Mean = 20 log
/ 10

N i 1 2

1 N Ai 2
Matched Filter Root Mean Square, Mfrms = 10 log
i 1 A ( ref )
N i

Ai (ref) - amplitude of ith component in the reference spectrum

6.2.3 Signature Spectral Analysis

The signature spectrum (Braun, 1986) is a baseline frequency spectrum taken from new or
recently overhauled machinery. It is then later compared with spectra taken from the same
machinery that represent current conditions. The unique nature of each machine and installation
is automatically taken into account. Characteristic component and fault frequencies can be
clearly seen and comparisons made manually (by eye), using indices, or using automated pattern
recognition techniques.

91
Figure 6.17 shows several samples of frequency spectra being used in signature analysis. The
figure shows a spectrum taken just after installation as a reference to the as new condition
(upper left) and a broadened version (upper right) of the same spectrum that is to be used for
comparison against future spectra calculated from vibration signals recorded at later times and
which can be expected to represent different machine conditions. The spectra in the lower half of
the figure show a series of measurement data collected at one month intervals. It is clear from
these plots that the responses in the subsequent spectra are changing and once particular
responses at particular frequencies exceed the reference spectrum further inspection and/or
analysis of the collected data may be needed to determine the type and severity of the fault that is
developing.

Figure 6.17 Signature spectral analysis.

Decibel (dB) units

Decibel (dB) units are simply a logarithmic scale representation of a dynamic signal. They are
calculated using equation 6.1 and require a reference value that sets the scale. This type of
display scale is an effective way to display very small values together with very large values.
Because vibration signal (particularly in the frequency domain) often result in strong and weak
signal response (representing different machine components) being presented in the same plot,
the dB scale is quiet useful.

A
dB 20 log10 rms 6.1
Aref

Arms = RMS value of a parameter

Aref = Reference value of the parameter

92
Figure 6.18 shows a linear scale (top) and a logarithmic scale (bottom) for comparison. It is clear
from this figure that relatively small amplitude values will be easier to see on the logarithmic
scale than the linear scale. As such, the logarithmic scale is better suited for displaying signals
that have a wide range of amplitude responses.

Linear Scale

11 10 100

Logarithmic Scale

1 10 100 1000

Figure 6.18 Signature spectral analysis.

Figure 6.19 shows the same frequency spectrum plotted using a linear amplitude scale (left) and
a logarithmic amplitude scale (right). It is clear from this figure that the logarithmic scale is ideal
for displaying data that is spread over a large dynamic range. In this case the dynamic range is 0
to 1000. Incremental increases in the amplitude of the spectrum are more clearly seen on the
logarithmic scaled plot. Another advantage is that changes in the amplitude at different levels are
displayed equally in the logarithmic scale. This makes it easier to judge changes in machinery
condition based on changes in the frequency spectrum amplitude at specific frequencies.

Figure 6.19 Linear and logarithmic amplitude scales used to display the same frequency
spectrum.

93
Figure 6.20 shows the same frequency spectrum plotted using a linear frequency scale (top) and
a logarithmic frequency scale (bottom). It is clear from this figure that the logarithmic scale
effectively compresses the frequency scale at the higher frequencies. This could be useful if the
lower frequencies (1X say) are of more interest. However, if it is necessary to see a clear
separation of responses in the higher frequency range, a linear frequency scale would be
preferred.

Figure 6.20 Linear and logarithmic frequency scales used to display the same frequency
spectrum.

Table 6.2 shows the standard reference values for calculating dB scales for displacement,
velocity and acceleration. Table 6.3 shows the dB increase compared to the linear multiplication
of a given value for a range of values. These values show that a very large range of values (in the
linear scale) can be represented in a much smaller dB (logarithmic) scale. This makes the large
dynamic range often found in vibration signals easier to display and interpret.

Table 6.2 Standardised reference values (ISO standard).

Parameter Displacement Velocity Acceleration

1 10 6 g
1 10 12 m
Reference or
110 9 m
s 1 10 6 m
s2

94
Table 6.3 dB increase compared to the linear multiplication factor.

dB increase Linear Multiplication


6 x2
10 x3
20 x 10
30 x 30
40 x 100
50 x 300
60 x 1000
70 x 3000

The dB scale is often associated with acoustic measurements. Figure 6.21 shows the linear and
logarithmic (dB) amplitude scales for a variety of acoustic noises. Again it is clear from this
figure that a very large linear dynamic range can be more easily displayed and interpreted using a
dB scale.

Figure 6.21 Linear and logarithmic (dB) amplitude scales for acoustic noises.

95
Windowing

As mentioned previously for a perfect conversion of time domain information to frequency


domain information an infinitely long set of sampled data is required. This of course is not
practical. The FFT algorithm needs a finite set of sample data to process. In fact this set needs to
be 2N in size (N is an integer) for the algorithm to be most efficient. To make the most accurate
conversion of time domain information to frequency domain information the FFT requires the
sampled data to start and end at zero amplitude. If the sampled signal does not begin and end at
zero amplitude, leakage will occur resulting in energy loss and false peaks to appear in the
frequency spectrum. Windows are used to force the end points of the data to zero.

Figure 6.22 shows the representation of a simple periodic input signal in the time domain (top).
Also shown is a section of the original signal representing a sampled signal taken from the
original (middle). Finally there is also shown an assumed input signal based on the sampled
signal (bottom). In this case, because the sample happens to begin and end with zero amplitude it
represents an integer number of sinusoid cycles. The FFT algorithm is applied to a finite length
signal sample and assumes that this sample is representative of the longer original and that the
signal is a pure sinusoid (which it is in this case).

actual input

Sampled time record


(integer # of cycles no
leakage)

assumed input

Figure 6.22 Actual, sampled and assumed representations of a dynamic signal when the
beginning and the end of the sampled signal are zero amplitude.

Figure 6.23 Representations of a dynamic signal when the beginning and the end of the sampled
signal are not zero amplitude.

However, if the sampled signal is not periodic in time domain, as shown in Figure 6.23, leakage
will occur. The sample signals no longer contain integer number of sinusoid cycles. A
discontinuity occurs in the sample signal and the FFT then assumes that this discontinuity occurs
regularly in the original signal and generates frequency content that describes it. Much like when

96
generating frequency spectra from square wave or triangular wave input signals (with obvious
discontinuities) the FFT inserts frequency content that is not really there in order to describe the
discontinuity.

Because most vibration signals that are recorded from rotating machines are not pure sinusoids
there will always be a non-integer number of cycles in the sampled signal and spectral leakage
due to boundary discontinuities will occur. One strategy to reduce the effect of leakage is to
employ windows that force the value of the sample signal to zero at the ends. Figure 6.24 shows
a window function (top), a sample signal to be analysed (middle) and the windowed signal
(bottom). The window function is simply a weighting function that the sampled signal is
multiplied by to reduce the end values to zero. Figure 6.25 shows a window function applied to a
continuous signal.

Figure 6.24 Window function applied to a sampled signal to reduce the end values to zero.

Figure 6.25 Window function applied to a continuous signal.

Figure 6.26 shows a sample signal (a), a typical window function (b), the windowed signal (c),
and the resulting frequency spectra generated from the original non-windowed signal (d - dashed
line) and from the windowed signal (d - solid line). It should be noted in this case that there is
not a significant difference between the two frequency spectra because the original sample signal
is a pure sinusoid and end effects are not that significant in this case. However, the spectrum
from the windowed signal does show that the shoulder frequency responses (sidebands) are
reduced. This modest improvement in resolution comes at the cost of a somewhat reduced
amplitude of the primary frequency response. This is due to the fact that a significant amount of
the signal has been significantly reduced in amplitude. This deficiency associated with

97
windowing can be compensated by a technique known as zero padding and over lapping. The
topics will be discussed later in this chapter.

Figure 6.26 Sample signal (a), window function (b), windowed signal (c), frequency spectra of
the original non-windowed signal (d - dashed line) and the frequency spectra of the windowed
signal (d - solid line).

There are many different types of windows. Figure 6.27 shows several of these and the shape of
frequency spectra that can be expected when using these windows. Hanning, Hamming,
Blackman and Bartlett-Hanning windows and commonly applied in machine condition
monitoring and fault diagnosis applications. A rectangular window (also known as a boxcar
window), which is effectively no windowing, should be used when performing modal analysis.
The input signal for modal analysis is often an impact excitation that generates a free vibration
response in the structure under consideration. The early stages of this type of response are of
critical importance when calculating the modal response. A window that attenuated the early part
of the sampled signal would effectively reduce that important part of the signal to zero.

Rectangular window Hanning window Hamming window

Bartlett window Blackman window Kaiser window

Tukey window Cosine window Bartlett-Hanning

Nuttall window Blackman-Harris Blackman-Nuttall

Figure 6.27 Several different types of window functions and the shape of frequency spectra that
can be expected when using these windows.
Overlaping and Zero Padding

98
Each time a frequency spectrum is generated with a FFT calculation only a fraction of the new
data are required. Figure 6.28 shows a long sample signal being cut into smaller sections for
processing through the FFT algorithm. Note that each small sample will in turn be windowed (as
shown in Figure 6.29) before the FFT is calculated. Averaging in the time domain or in the
frequency domain (as discussed previously) will result in a final frequency spectrum that is
relatively free of random noise. However, the averaging process requires a relatively long signal
to start with.

Figure 6.28 Relatively long signal being cut into smaller sections for FFT calculations and then
averaged to a final frequency spectrum (no overlapping).

Figure 6.29 Windowed signal section prior to FFT.

Figure 6.30 shows overlapped shorter sections of the time waveform prior to FFT processing. A
significantly shorter overall signal length is required to extract the same degree of accuracy in
the frequency spectrum. It may appear as if some sections of the vibration signal are used twice
making the averaging process less accurate. However, when considering that the window takes a
significant amount of information out of the beginning and end of each time sample (see Figure
6.29), the overlapping does not effectively cause the loss on any raw data. Testing using various
lengths of overlapped signal has shown this to be true.

Because the FFT process requires the sampled signal to be an exact length (a factor of 2N, where
N is an integer) it is sometime necessary to add zeros to the data sample in order to achieve this
exact sample signal length. This does not typically add any significant error to the resulting
frequency spectra because the end points have been forced to zero amplitude by the windowing

99
process already. In fact, the added signal length can have a positive affect on the degree of
frequency resolution that is possible.

Figure 6.30 Overlapped smaller sections of the time waveform prior to FFT.

6.2.4 Cascades (Waterfall Plots)

Cascade plots (also known as waterfall plots) are successive spectra plotted with respect to time
and displayed in a three-dimensional manner. Changing trends can be seen easily, which makes
this type of display a useful fault detection and trending tool. This type of display is also used
when a transient event, such as a coast-down, is known to be about to occur. Cascade plots can
also be linked to the speed of a machine. In this case, the horizontal axis is labeled in multiples
of the rotational speed of the machine. Each multiple of the rotational speed is referred to as an
order.

Order tracking is the name commonly used to refer to cascade plots that are synchronously
linked to the machine rotational speed via a tachometer. As the speed of the machine changes,
the responses at specific frequencies change relative to the speed, but are still tracked in each
time-stamped spectra by the changing horizontal axis scale.

Figure 6.31 shows a simplified cascade plot with a response frequency of one times the rotating
speed (1X) and another response at approximately one-half operating speed (1/2X). The 1X
response shows an increase in amplitude as time progresses. This represents the machine passing
through the first critical speed (where the machine rotational speed matches the first natural
frequency). The 1/2X response represents the response due to oil whirl. Oil whirl is the motion of
the shaft in a journal bearing as the oil wedge that supports the shaft moves around the journal.
This frequency is typical just less than shaft rotational speed. In a case where the change in
shaft speed is too slow (the machine passes through the critical speed too slowly) or when the
bearing is too lightly damped, the response at the first natural frequency may excite the oil whirl
into a condition that is referred to as oil whip. Oil whip is when a natural frequency (resonance)
is excited and continues to be excited even after the machine speed has passed through the
critical speed. This is a dangerous situation as mechanical resonances can lead to extremely large

100
vibration forces and displacements. Figure 6.32 shows a more realistic example of the same thing
described in Figure 6.31.

Rotor Speed Oil Oil Whip


Mass Unbalance
Whirl

Critical Speed (system Frequency


natural frequency is excited)

Figure 6.31 Simplified cascade plot showing oil whirl and oil whip.

Figure 6.32 Realistic cascade plot showing oil whirl and oil whip.

Figure 6.33 shows another example of a cascade plot. In this case the plot shows the responses at
1X and higher multiples of the shaft rotational speed. The natural frequencies of the system are
revealed as the responses at different rotational speeds display increased amplitudes at different
frequencies. These responses are typically lower in amplitude than the first natural frequency
response.

101
Figure 6.33 Cascade plot showing natural frequency excitations at 1X and higher multiples of
shaft rotational speed.

Natural Frequencies

Natural frequencies are determined by the design of the machine or a component. They are the
properties of the system and are dependent on the distribution of mass within the machine or on
the component, the stiffness of the structural components, and the damping of the system. Every
system has a number of natural frequencies. If a forcing frequency is close to a natural
frequency, a resonance occurs and results in high vibration levels. If forcing frequency is a
multiple of the operating speed, this speed is called a critical speed.

6.3 Modal Domain

Modal analysis is not traditionally considered a machine condition monitoring and fault
diagnostics tool, but is included here because of the ever-increasing complexity of modern
machinery. Often, unless the natural (free and forced response) frequencies of machinery, their
support structure, and the surrounding buildings are fully understood, a complete and accurate
assessment of existing machinery condition is not possible. A complete overview of modal
analysis will not be provided here, but a specific approach to modal analysis (operational
deflection shape [ODS] analysis) will be described.

ODS analysis is like other types of modal analysis in that a force input is provided to a structure
or machine and then the response is measured. The response at different frequencies defines the
natural frequencies of the structure or machine. Typically, an impact or constant frequency force
is used to excite the structure. In the case of ODSs, the regular operation of the machinery
provides the excitation input. With vibration sensors placed at critical locations and a reference
signal linking together all the recorded signals, a simple animation showing how the machine or
structure deflects under normal operation can be generated. These animations, along with the
frequency information contained in each individual signal, can provide significant insights into
how a machine or structure deforms under a dynamic load. This information, in turn, can be a
useful in addition to other data when attempting to diagnose problems.

102
6.3.1 Mode Shapes

The mode shape is the shape assumed by a system as it vibrates at a natural frequency. Mode
shapes are associated with all natural frequencies. A mode shape does not provide information
on system absolute motions, only on the relative motion between machine and structural
components. Absolute motions are determined by damping and vibration forces.

Figure 6.34 shows the mode shapes of a vibrating beam. In this simple case the beam is
supported in such a way as to allow for free vibration of any point. That is, no part of the beam is
fixed to a support. This is a relatively common approach when testing structures or components
for their natural frequency responses during modal testing. However, when modal testing in the
field one or more locations on the machine or structure are typically considered as fixed
because they are the points where the machine or structure is held in place. The mode shapes
shown in Figure 6.34 display the shape and relative amplitude of the beam as it is vibrating at
each of the first three natural frequencies. In most cases the complexity of the mode shapes
increase with frequency and the relative vibration amplitudes decrease with frequency.

Figure 6.34 Mode shapes of a vibrating beam.

Figure 6.35 Typical rotor with disk.

103
Figure 6.35 shows a typical rotor with disk mounted at mid-span. The mode shapes and natural
frequencies that will be present in such a case will be dependant on the physical dimensions and
material properties of the rotor and disk as well as on the support bearing physical properties.
Bearings supporting shafts may be considered as soft, intermediate or stiff. Typically journal
(fluid filled) bearings are soft or intermediate while rolling element bearings are considered as
stiff. Bearings are also the primary way that system damping is affected. Oil viscosity and
bearing running temperature affect bearing damping characteristics as does the internal design.
This topic will be discussed again in a later chapter.

Figure 6.36 shows the potential responses of a rotor with mounted disk at three different natural
frequencies and with three different bearing stiffnesses. The mode shapes reflect the fact that
shaft movement at the bearings is larger in amplitude with decreasing bearing stiffness. Soft
bearings typically have lower friction, but also allow for more movement under vibratory
excitation.

Figure 6.36 Mode shapes of a flexible rotor with disk.

6.3.2 Deflection Shape Analysis

Traditionally, an operational deflection shape (ODS) has been defined as the deflection of a
structure at a particular frequency. However, an ODS can be defined more generally as any
forced motion of two or more points on a structure. Specifying the motion of two or more points
defines a shape. Stated differently, a shape is the motion of one point relative to all others.
Motion is a vector quantity, which means that it has location and direction associated with it.

Measuring ODSs can help answer vibration related questions such as; How much is a machine
moving?, Where is it moving the most, and in what direction?, What is the motion of one point
relative to another (Operating Deflection Shape)?, Is a resonance being excited?, What does its
mode shape look like?, Is there structure-born noise?, and Do corrective actions reduce noise or
vibration levels?.

As discussed previously all vibration is a combination of both forced and resonant vibration.
Forced vibration can be due to, internally generated forces such as unbalances or external loads
such as ambient excitation. An operating deflection shape contains the overall vibration for two

104
or more degrees-of-freedom (DOF) on a machine or structure. That is, the ODS contains both
forced and resonant vibration components. On the other hand, a mode shape characterizes only
the resonant vibration at two or more DOFs. Resonant vibration typically amplifies the vibration
response of a machine or structure far beyond the design levels for static and/or dynamic loading.
Resonant vibration is the cause of, or at least a contributing factor to many of the vibration
related problems that occur in structures and operating machinery. To understand any structural
vibration problem, the resonances of a structure need to be identified. A common way of doing
this is to define its modes of vibration. Each mode is defined by a natural (modal) frequency,
modal damping, and a mode shape.

An ODS can be defined from any forced motion, either at a moment in time, or at a specific
frequency. Having acquired either a set of sampled time domain responses, or computed (via the
FFT) a set of frequency domain responses, an operating deflection shape is defined as: The
values of a set of time domain responses at a specific time, or the values of a set of frequency
domain responses at a specific frequency. Figure 6.37 shows the display of an ODS from a set of
impulse response measurements. This type of response is more accurately referred to as the
modal response as the excitation is predominantly external.

Figure 6.37. Time domain ODS from impulse responses.

Real continuous structures have an infinite number of degrees-of-freedom and an infinite number
of modes. From a testing point of view, a real structure can be sampled spatially at as many
DOFs as needed. There is no limit to the number of unique DOFs at which measurements can be
made. Because of time and cost constraints, it is typical to only measure a small subset of the
measurements that could be made on a structure. From this subset of measurements an accurate
definition of the resonances within the frequency range of interest is possible. Of course, the
broader the spatial sampling on the surface of the structure, the more definition possible in the
ODS and mode shapes.

In general, an ODS is defined with a magnitude and phase value at each point on a machine or
structure. To define an ODS vector properly, at least the relative magnitude and relative phase
are needed at all response points. In a time domain ODS, magnitude and phase are implicitly

105
assumed. This means that either all of the responses have to be measured simultaneously, or they
have to be measured under conditions which guarantee their correct magnitudes and phases
relative to one another. Simultaneous measurement of all responses means that a multi-channel
acquisition system, that can simultaneously sample all of the response signals, must be used.
This requires lots of transducers and signal conditioning equipment, which is expensive.

If the structure or machine is undergoing, or can be made to undergo, repeatable operation, then
response data can be acquired one channel at a time. Repeatable operation typically means
steady state operation when applying ODS to rotating machines. To be repeatable, data
acquisition must occur so that exactly the same time waveform is obtained in the sampling
window, every time one is acquired. For repeatable operation, the magnitude and phase of each
response signal is unique and repeatable, so ODS data can be acquired using a single channel
analyzer. An external trigger is usually required to capture the repeatable event in the sampling
window.

A single channel analyzer can be used to acquire ODSs using a tachometer pulse as the trigger.
In this case, spectral magnitude and phase data at any fixed multiple of the rotational speed
(order) of the machine can be acquired with a single channel analyzer, but only if the operation is
repeatable. Steady state operation is achieved when the auto power spectrum of a response signal
does not change over time, or from measurement to measurement.

For steady state operation, ODS data can also be measured with a 2 channel FFT analyzer or
acquisition system. The cross spectrum measurement contains the relative phase between two
responses, and the auto power spectrum of each response contains the correct magnitude of the
response. Since the two response signals are simultaneously acquired, the relative phase between
them is always maintained. No special triggering is required during steady state operation.

An ODS FRF (frequency response function) is a different 2-channel measurement that can also
be used when excitation forces cannot be measured. The advantage of the ODS FRF is that the
ODS FRF has peaks at resonances, thus making it easy for locating resonances. An ODS FRF
also requires a reference (fixed) response measurement along with each response measurement.
Each ODS FRF is formed by replacing the magnitude of each cross spectrum between a response
and the reference response with the auto spectrum of the response. The phase of the cross
spectrum is retained as the phase of the ODS FRF. This new measurement contains the correct
magnitude of the response at each point, and the correct phase relative to the reference response.
Evaluating a set of ODS FRF measurements at any frequency yields the frequency domain ODS
for that frequency. Figure 6.38 shows the display of an ODS from a set of ODS FRF
measurements.

In addition, modal parameters (natural frequency, damping, & mode shape) can be obtained from
a set of FRF measurements. In general, the following statement can be made, All experimental
modal parameters are obtained from measured ODSs. Stated differently, modal parameters are
obtained by post-processing (curve fitting) a set of ODS data. In other words, a set of FRFs can
be thought of as a set of ODSs over a frequency range. At or near a resonance peak, the ODS is
dominated by the mode at that frequency. Therefore, the ODS is approximately equal to the
mode shape. This concept becomes clearer when sine wave excitation is used.

106
Figure 6.38 ODS displayed directly from ODS FRF data.

6.3.3 Operational Deflection Shape Analysis compared to Mode Shape Analysis

Even though all experimental mode shapes may be obtained from measured ODSs, mode shapes
are different from ODSs in the following ways.

Each mode shape is defined for a specific natural frequency. An ODS can be defined at any
frequency.
Mode shapes are only defined for linear, stationary structures. ODSs can be defined for non-
linear and non-stationary structures.
Mode shapes are used to characterize resonant vibration. ODSs can characterize resonant as
well as non-resonant vibration.
Mode shapes do not depend on forces or loads. They are inherent properties of the structure.
ODSs depend on forces and/or loads. They will change if the loads change.
Mode shapes only change if the material properties or boundary conditions change. ODSs
will change if either the modes or the loads change.
Mode shapes do not have unique values or units. ODSs do have unique values and units.
Mode shapes can answer the question, What is the relative motion of one DOF versus
another? ODSs can answer the question, What is the actual motion of one DOF versus
another?

107
6.4 Quefrency Domain Signal Analysis

A quefrency domain (Randall, 1981, 1987) plot results when a Fourier transform of a frequency
spectra (log scale) is generated. As the frequency spectra highlight periodicities in the time
waveform, so the quefrency cepstra highlights periodicities in a frequency spectra. This
analysis procedure is particularly useful when analyzing gearbox vibration signals where
modulation components in spectra (sidebands) may overly complicate the spectrum, but are
easily detected and diagnosed in the cepstrum.

Figure 6.39 shows several different representations of vibration data. Figure 6.39(a) shows a time
domain signal with clearly repetitive components. Figure 6.39(b) shows the frequency spectrum
for the same signal and highlights the repetitive nature of the time waveform with responses at
specific frequencies. Figure 6.39(c) shows a frequency spectrum with large number of frequency
responses which may be difficult to analyze because of the density of the information. Figure
6.39(d) shows the quefrency cepstrum calculated by taking the FFT of the spectrum in Figure
6.39(c). The cepstrum shows a clearer picture of the repetitive components from Figure 6.39(c).
As mentioned previously this signal processing technique is useful at extracting information
from frequency spectra that contain a large amount of closely spaced frequency responses, as are
often generated by gear trains.

Figure 6.39 Time domain, frequency domain and quefrency domain representations of vibration
signals.

108
Chapter Summary

Generally, the more processing that is done on the dynamic signal, the more specific useful
information is highlighted and the more extraneous information is discarded.
The primary display formats used in machine condition monitoring are the time domain, the
frequency domain, the modal domain, and the quefrency domain.
The time domain refers to a display or analysis of the vibration data as a function of time,
allowing for little or no data to be lost prior to inspection.
Time domain analysis includes: waveform analysis, time waveform indices, time
synchronous averaging, negative averaging, orbit analysis, probability density functions, and
probability density moments.
The frequency domain refers to a display or analysis of the vibration data as a function of
frequency, where the time domain vibration signal is typically processed into the frequency
domain by applying a Fourier transform, usually in the form of a FFT algorithm.
The principal advantage of frequency domain analysis is that the repetitive nature of the
vibration signal is clearly displayed as peaks in the frequency spectrum at the frequencies
where the repetition takes place. This allows for faults, which usually generate specific
characteristic frequency responses, to be detected early, diagnosed accurately, and trended
over time as the condition deteriorates.
Frequency-domain analysis includes the use of band pass analysis, shock pulse (spike
energy), envelope spectrum, signature spectrum, cascades (waterfall plots), masks, and
frequency domain indices.
Quefrency-domain analysis involves a Fourier transform of a frequency spectra (log scale).
As the frequency spectra highlight periodicities in the time waveform, so the quefrency
cepstra highlights periodicities in a frequency spectra.

109
Chapter 7

Transducer Selection
A transducer is a device that senses a physical quantity (vibration in this case, but it can also be
temperature, pressure, etc.) and converts it into an electrical output signal, which is linearly
proportional to the measured variable. As such, the transducer is a vital link in the measurement
chain. Accurate analysis results depend on an accurate electrical reproduction of the measured
parameters. If information is missed or distorted during measurement, it cannot be recovered
later. Hence, the selection, placement, and proper use of the correct transducer are important
steps in the implementation of a condition monitoring and fault diagnostics program.

Considerable research and development work has gone into the design, testing, and calibration of
sensors (transducers) for a wide range of applications. The most basic requirements of
transducers are that they be; i) correct for the task, ii) properly mounted, iii) in good working
order (properly calibrated), and iv) fully understood in terms of operational characteristics. The
first considerations when selecting a sensor is the frequency range within which that sensor will
be expected to accurately measure a vibration signal.

7.1 Selecting a Measurement Parameter

A general guideline for transducer selection is presented graphically in Figure 7.1. This figure
suggests that displacement measurements should be preferred in the frequency range up to 1000
Hz (1500 Hz maximum). Velocity measurements can be made in the frequency range of 50 Hz to
1000 Hz (for electromechanical devices) and from 10 Hz to 2000 Hz (for piezoelectric devices).
Acceleration measurements can be accurately made at frequencies above 1.0 Hz. For constant
velocity vibration amplitude across all frequencies, a displacement transducer is more sensitive
in the lower frequency range, while an accelerometer is more sensitive at higher frequencies.
While it may appear as if the velocity transducer is the best compromise, transducers are selected
to optimize sensitivity over the frequency range that is expected to be recorded.

Figure 7.1 is meant to show the relative sensitivity of different sensor types as a function of
frequency. This figure is a generalization of the relative sensitivities and should not be
interpreted as prescriptive guide. There are many reasons why in a particular application one
sensor type may be selected over another. This will be discussed in detail later in this chapter.

Transducers usually require amplification and conversion electronics to produce a useful output
signal. These circuits may be located within the sealed sensor unit or in a separate box. There are
advantages and disadvantages to both of these configurations but they will not be detailed here.

Important considerations that will be considered throughout this chapter include frequency
response range (which must be compatible with the frequencies generated by mechanical
components of the machine), signal-to-noise ratio, sensitivity, natural frequency range of the
sensor, output signal strength, weight and size of the sensor, temperature limitations, and
measurement units translated to desired analysis units. Sensor frequency spans should cover the
maximum frequency components of interest. Clipping occurs when the amplitude range of the

110
measured signal greater than the maximum amplitude range (dynamic range) of the transducer.
To increase spectral resolution, it is necessary to acquire multiple data samples or split data
points into more spans.

Relative Amplitude Response


1 Acceleration

Velocity
1.

Displacement

0.

0. 1. 1 10 1,00 10,00
Frequency (Hz)

Figure 7.1 Frequency versus response amplitude for various sensor types.

Traditional vibration sensors fall into three main classes. These classes are; non-contact
displacement transducers (also known as proximity probes or eddy current probes), velocity
transducers (electro-mechanical or piezoelectric), and accelerometers (piezoelectric). Force and
frequency considerations dictate the type of measurements and applications that are best suited
for each transducer. Recently, laser-based non-contact velocity/displacement transducers have
become more commonplace. These are still relatively expensive because of their extreme
sensitivity, and hence are still predominantly used in the laboratory setting.

The type of motion sensed by displacement transducers is the relative motion between the point
of attachment and the observed surface. Velocity transducers and accelerometers measure the
absolute motion of the structure to which they are attached.

Figure 7.2 Frequency versus relative amplitude for various sensor types.

111
Figure 7.2 shows the ranges in terms of frequency response and relative amplitude response over
which displacement, velocity and acceleration sensor function. Again this is a general guide and
not meant to be prescriptive. There are many situations where sensors could find use outside
these guidelines and of course new sensors are constantly being designed to extend the useful
ranges of different types of sensors.

7.2 Non-contact Displacement Transducers (Proximity (Eddy Current) Probes)

These types of sensors find application primarily in fluid film (journal) radial or thrust bearings.
With the rotor resting on a fluid film there is no way to easily attach a sensor directly to the shaft
and measure its vibration. Nor is it effective to attach a sensor to the bearing housing because the
a journal bearing is designed to allow the shaft to move somewhat inside the bearing seal
clearances. A non-contact approach is then the best alternative. Non-contact measurements
indicate shaft motion and position relative to the bearing. Radial shaft displacements and seal
clearances can be conveniently measured. Another advantage of using non-contact displacement
probes is that when they are used in pairs, set 90 apart, the signals can be used to show shaft
dynamic motion (orbit) within the bearing.

Non-contacting eddy current displacement transducers are typically permanently mounted on


machines. They have a measurement range of from 0 mils to 80 mils (0 2 mm). The sensitivity
is usually around 200 mV/mil. The probe should be shielded and grounded.

Figure 7.3 Non-contact Eddy current type probes showing different probe tips, a connection box,
and a labeled schematic.

112
Figure 7.4 Non-contact Eddy current type probes showing internal circuitry and different
probe/wiring connection types.

The linearity and sensitivity of the proximity probe depends on the target conductivity and
porosity (porosity affects resistance). Calibration of the probe on the specific material in use is
recommended. This type of sensor is capable of both static and dynamic measurements, but
temperature and pressure extremes will affect the transducer output. The probe will detect small
defects in the shaft (cracks or pits), and these may seem like vibrations in the output signal.

Figure 7.5 shows a schematic diagram of how non-contact displacement sensors could be
mounted to measure radial and axial displacement. Figure 7.6 shows two non-contact
displacement sensors positioned 90 apart in order that their combined outputs can be used to
generate orbit plots. Note the clearance close to the sensor tip. Magnetic material needs to be
kept away from the tip in order to maintain maximum sensitivity to the measured surface. Figure
7.7 shows several different orientations that can be used to mount non-contact displacement
sensors. Because these sensors are not orientation sensitive, they can be used in any orientation
and maintain their standard functional properties.

Figure 7.5 Schematic showing radial and axial non-contact displacement sensor mounting
positions.

113
Figure 7.6 Non-contact displacement sensor positions 90 apart for generating orbits from the
combined output.

Figure 7.7 Schematic showing various non-contact displacement sensor mounting positions.

7.2.1 Design and Operation

Non-contact displacement probes work in the following manner. The probe is excited from an
external source at a frequency of 1.5 kHz. The excitation produces a magnetic field radiating
from probe tip. Eddy currents are induced into any conductive material close to the tip. Eddy
currents are electrical currents that flow on the surface of a conductive material when moving
through a magnetic field. Energy is extracted from the probe as the distance from tip to the
conductive material is varied. The energy extracted is proportional to the distance to the
conductive material. The linear range is limited and proper calibration is essential, as well be
discussed below.

The probe excitation frequency acts as a relatively high frequency carrier wave for the signal.
Riding on top of the carrier signal is the low frequency displacement signal. When the carrier
signal is stripped off using low pass filtering, the remaining signal is the displacement as a
function of time. See Figure 7.8(a).

7.2.2 Orbits

Figure 7.8 shows a single channel and dual channel measurement result. When the two channels
are plotted against one another, they clearly show what are known as shaft orbits. These orbits

114
define the dynamic motion of the shaft in the bearing, and are valuable fault detection and
diagnostic tools.

t
tpu
Ou

Y Direction Displacement
Vibration Amplitude
Modulated Carrier

Vibration Direction
Output
Vibration

Time

me
Ti
Time

Y Direction Displacement
Y Output
Vibration

Vibration
Demodulator
Direction X Output

X Direction Displacement
Time

(a) (b)
Figure 7.8 (a) Shaft displacement with one sensor; (b) shaft orbit with two sensors (x direction
versus y direction displacement assumes shaft is circular).

Of course the shaft motion is rarely exactly circular. Figure 7.9 shows an example of shaft
motion that is measured with two non-contact displacement probes 90 apart. This type of
measurement and plot of rotational displacement signal provides an excellent window into the
actual motion of the rotating shaft within its bearings.

Figure 7.9 Example shaft motion measured with two non-contact displacement probes 90 apart.

7.2.3 Calibration

The calibration curve for non-contact displacement sensors has three regions. These regions are
shown in Figure 7.10. The first region is when the probe is in contact with, or very near the shaft.
The output in this region is 0 volts DC (or close to zero and highly non-linear). The next region

115
is the linear region where the change in gap distance produces a proportional change in DC
output. The typical linear gap is between 0.25 mm and 2.25 mm. The industry standard output in
this range is 8 volts/mm. The third region is farther away than the linear region where the supply
voltage is reached and the system looses constant proportionality.

Supply

Output (Volts)
Linear

Gap Distance
(0.25 mm to 2.25 mm)

Figure 7.10 Typical calibration curve for a non-contact displacement sensor.

The system sensitivity is equal to the slope of the linear part of the calibration curve, or basically
the change in output divided by the change in gap within the linear region. Probe linearity
depends on target conductivity, porosity, surface conditions (cracks, rust, pits, etc.), temperature
and pressure influences. Finally non-contact displacement sensors are capable of both static and
dynamic measurements.

Figure 7.11 Typical non-contact displacement sensor sensitivities for different materials.

7.2.4 Mounting and Location

Installation of these sensors requires a rigid mounting. Adaptors for quick removal and
replacement without machine disassembly can be useful. The minimum tip clearance from all
adjacent surfaces should be two times the tip diameter. Probe extensions must be checked to
ensure that the resonant frequency of the extension is not excited during data gathering. As with
all sensors, care must be taken when handling the cables and the connections must be kept clean.

116
Figure 7.12 shows a non-contact displacement sensor inserted into a journal bearing. Note the
metal free zone around the sensor tip to ensure maximum sensitivity. Figure 7.13 shows a typical
non-contact displacement probe tip extension. Care should be taken when using tip extensions in
order that the extension natural frequency is not excited. Figure 7.14 shows a photograph of a
mounting extension. Again, care should be taken when designing and using extensions in order
that the extension natural frequency is not excited. The extension shown in Figure 7.14 is not a
particularly good design as it is poorly supported on the right hand side and may have a low (and
easily excited) natural frequency.

Figure 7.12 Typical non-contact displacement sensor installation.

Figure 7.13 Typical non-contact displacement sensor probe tip extension.

Figure 7.14 Poorly designed non-contact displacement sensor extension.

117
7.2.5 Non-contact Displacement Sensors as Triggers and Speed Sensors

As mentioned previously non-contact displacement sensors are often used as triggers to show a
once per shaft revolution marker or determining phase relationships and for measuring speed (as
a tachometer). Figure 7.15 shows a non-contact displacement sensor being used to detect a
keyway as it travels past the sensor once per shaft revolution. This type of use of a non-contact
displacement sensor is referred to as a keyphasor. When the keyphasor transducer observes a
once per revolution marker (the keyway) it generates a keyphasor timing event. The event is the
sudden negative change in signal voltage when the leading edge of the keyway passes next to the
probe. The voltage drops because there is no material in the empty keyway.

Figure 7.15 Non-contact displacement sensor being used as a keyphasor.

Figure 5.16 and Figure 5.17 show several different keyphasor trigger signals. These signals and
the trigger points used depend on whether or not there is a key in the keyway as well as on
whether the leading or trailing edge of the key (or empty keyway) is used as the trigger point.

Figure 7.16 Several different keyphasor signal trigger points (key in keyway).

118
Figure 7.17 Several different keyphasor signal trigger points (no key in keyway).

Proximity probes can also be used as speed sensors. Figure 7.18 shows a probe being used with a
gear ring. With a known number of teeth on the gear it is relatively easy to use the output from
the sensor (a trigger at each leading and trailing edge of each gear tooth) to calculate the speed of
the shaft at any point in time. The proximity probe signal would need to be recorded along with
the vibration signal if detailed analysis was planned for later.

Figure 7.18 Proximity probe being used as a shaft rotational speed sensor.

7.2.6 Advantages and Disadvantages

In summary the following advantages and disadvantages can be listed in association with the use
of proximity probes.

119
Table 7.1 Advantages and disadvantages of non-contact displacement sensors.

Advantages Disadvantages
Measures shaft dynamic motion Sensitive to surface imperfections and
magnetism
Measures shaft static position Sensitive to material properties
Excellent signal response between DC and Shaft surface must be conductive
90,000 CPM (1.5 kHz)
No natural frequency throughout transducer Low dynamics signal response above 90,000
operating range CPM (1.5 kHz)
Simple calibration External power supply required
Solid state electronics Correct probe to cable impedance must be
maintained
Rugged and reliable construction Minor temperature sensitivity
Available in many physical configurations Sensitive to interference from adjacent
proximity probes
Suitable for installation in harsh environments Sensitive to probe mounting bracket
resonances
Multiple machinery applications for the same Potentially difficult to install
transducer

7.3 Laser Based Non-Contact Displacement Sensors

Laser based non-contact displacement transducers (vibrometers) use the Doppler effect as it
applies to laser light to measure distances very accurately. The moving object reflects light from
the incoming laser beam and the Doppler frequency shift is used to measure the component of
velocity which lies along the axis of the laser beam. A low power Helium-Neon laser (eye safe)
is used. These sensors can operate up to 200 meters away from the moving object that is under
measurement. The device uses the phase difference between the measured signal and an internal
reference light source to determine the phase change.

Figure 7.19 Scanning laser vibrometer system.

120
7.4 Velocity Transducers (Electro-Mechanical

There are two general types of velocity transducers. They can be distinguished by considering
the mode of operation. The two types are electro-mechanical and piezoelectric crystal based.
Piezoelectric crystal-based transducers will be discussed in the next section, so the focus here
will be on electro-mechanical.

7.4.1 Design and Operation

Electro-mechanical velocity transducers function with a permanent magnet (supported by


springs) moving within a coil of wire. As the sensor experiences changes in velocity, as when
attached to a vibrating surface, the movement of the magnet within the coil is proportional to
force acting on the sensor. The current in the coil, induced by the moving magnet, is proportional
to velocity, which in turn is proportional to the force. This type of device is known as self-
generating and produces a low impedance signal; therefore, no additional signal conditioning is
generally needed.

Figure 7.20 shows two schematics and a photograph of electro-mechanical velocity sensors.
Figure 7.20(a) shows a configuration where the magnet moves inside the coil of wires. Figure
7.20(b) shows a configuration where the coil of wires moves outside a fixed magnet. Figure
7.20(c) shows a photograph of an electro-mechanical velocity sensor.

(a) (b) (c)

Figure 7.20 Electro-mechanical velocity transducers.

Electro-mechanical velocity sensors require the spring suspension system to be designed with a
relatively low natural frequency. These devices have good sensitivity, typically above 10 Hz, but
their high-frequency response is limited (usually to around 1500 Hz) by the inertia of the system.
Some devices may obtain a portion, or all, of their damping electrically. This type must be
loaded with resistance of a specific value to meet design constraints. These are usually designed
for use with a specific data collection instrument and must be checked and modified if they are to
be used with other instruments. Figure 7.21 shows a plot of the sensitivity vs. frequency for an

121
electro-mechanical velocity transducer. Note that the sensor natural frequency is below the linear
operating range of the device.

Amplitude Response
Natural 10 Hz Frequency
Frequency

Figure 7.21: Output sensitivity versus frequency for an electro-mechanical velocity transducer.

While electro-mechanical velocity transducers can be designed to have good dynamic range
within a specific frequency range, there are several functional limitations. Because a damping
fluid is typically used to provide most of the damping, this type of transducer is limited to a
relatively narrow temperature band, below the boiling point of the damping liquid. The
mechanical reliability of these sensors is also limited by the moving parts within the transducer,
which may become worn or fail over time. This has resulted in this type of transducer being
replaced by piezoelectric sensors in machine condition monitoring applications. The orientation
of the sensor is also limited to only the vertical or horizontal direction, depending on the type of
mounting used. Finally, as a damped system, such as an electro-mechanical velocity transducer,
approaches its natural frequency, a shift in phase relationships may occur (below 50 Hz). This
phase shift at low frequencies will affect analysis work.

7.4.2 Advantages and Disadvantages

In summary the following advantages and disadvantages can be listed in association with the use
of electro-mechanical velocity sensors.

Table 7.2 Advantages and disadvantages of electro-mechanical velocity sensors.

Advantages Disadvantages
Measures casing absolute vibration Sensitive to mounting fixture and transducer
orientation
Easily attached to machinery externals, Operates above transducer natural frequency of
piping, baseplates or structures about 600 CPM (10 Hz)
Good signal response between 900 and Poor signal response below 900 CPM (15 Hz) and
90,000 CPM (15 to 1,500 Hz) above 90,000 CPM (1,500 Hz)
Self-generation signal electronics Difficult calibration check
No special wiring required Unable to measure shaft vibrations or position
Available in several configurations Potential for failure due to fatigue of moving
internal parts
Temperature sensitive with a typical upper limit of
250 F (120 C) and lower limit of 30 F (-1 C)

122
7.5 Acceleration Transducers

7.5.1 Principles of Operation

By far the most commonly used transducers for measuring vibration are accelerometers. These
devices contain one or more piezoelectric crystal element (natural quartz or man-made ceramic),
which produce voltage when stressed in tension, compression or shear. This is the piezoelectric
effect. The voltage generated across the crystal pole faces is proportional to the applied force.

Accelerometers and force transducers work in the same way. Accelerometers can be used to
measure absolute vibration levels on casings or bearing housings. They contain a small mass and
piezoelectric crystal. The piezoelectric crystal produces an electric charge output (that can easily
be converted to voltage) proportional to the applied acceleration or force. Small size
accelerometers have lower sensitivity yet higher frequency range. Large accelerometers have
higher sensitivity but lower frequency range.

Accelerometers generally have a linear response over a wide frequency range (10 Hz to 20 kHz),
with specialty sensors linear down to 0.5 Hz and up to 50 kHz. This wide linear frequency range
and the broad dynamic amplitude range make accelerometers extremely versatile sensors. In
addition, the signal can be electronically integrated to give velocity and displacement
measurements. This type of transducer is relatively resistant to temperature changes, reliable
(having no moving parts), produces a self-generating output signal meaning no external power
supply is needed unless there are onboard electronics, is available in a variety of sizes, is usually
relatively insensitive to non-axial vibration (3% of main axis sensitivity), and can function well
in any orientation. Signals from this type of transducer contain significantly more vibration
components than other types. This means that there is a large amount of information available in
the raw vibration signal.

Figure 7.22 shows the sensitivity versus frequency relationship for a typical accelerometer. This
figure shows the linear range and the extent to which responses will be linear at the extremes of
the frequency range (upper and lower frequencies) as well as the sensitivity deviation at those
extremes. These limits are commonly reported on the sensor specification sheet.

Figure 7.22: Sensitivity versus frequency relationship for a typical accelerometer.

123
Figure 7.23 shows a typical piezoelectric frequency response curve with a typical frequency
range. Note that accelerometers have a much higher natural frequency than velocity sensors. The
ratio of output to input is also constant over a wider range of frequencies than for velocity
sensors. Measurement takes place at frequencies below the natural frequency of the seismic mass
meaning that there is no worry about phase effects entering the measured signal.

Figure 7.23: Typical piezoelectric frequency response curve.

Figure 7.24 shows another typical accelerometer frequency response curve with a typical
frequency range (in Hz and in CPM cycles per minute) and with a typical sensitivity scale (in
milliVolts per G). Note that accelerometers typically have their sensitivity expressed in mV/G.
Attenuation begins to set in below around 15 Hz, but specialty sensors are available with lower
limits to the linear frequency response. Sensor structural resonance is typically around 30 kHz,
but again specialty sensors are available with higher limits to the linear frequency response.

Figure 7.24: Typical piezoelectric frequency response curve (with scale).

The linear range of the frequency response curve for accelerometers is broad, but still limited at
the upper and lower ends. These limits must not be ignored. Typically the lower limit is not of as
much concern as the upper limit. The lower limit causes the response signal to attenuate at low

124
frequencies, which is a concern because vibration signal responses at those frequencies are
already likely to be low energy. Some important information could be lost. However, these lower
frequency limits are so low that it is unlikely that responses in this range will be of interest in
typical machine condition monitoring and fault diagnosis applications. As mentioned previously,
there are specialty accelerometers that can be used in situations where extremely low frequency
responses are important.

Responses at the upper limit of linear response are another story. The upper limit separates the
linear frequency response range from the sensor natural frequency (resonance). If vibration
signal components exist in the range of frequencies that could excite a resonance response in the
sensor the output from the sensor would be amplified far above the true measure of the input.
This exaggeration of the vibration signal at those frequencies could lead to incorrect machine
condition assessments as well as damage the sensor.

Figure 7.25 shows two typical accelerometer frequency response curves. Figure 7.25(b) shows
two potential frequency responses from unknown machine related forcing functions. Both of
these responses occur within the linear range of the accelerometer. The resulting measured
response can be expected to be close to the amplitude of the actual vibration generating the
signal.

(a) (b)

Figure 7.25 Typical accelerometer frequency response curves with (a) linear range highlighted
and (b) potential vibration responses within the linear range.

Figure 7.26(a) shows a typical accelerometer frequency response curve with two vibration signal
input excitations. One (at the lower frequency) is in the linear response range. The other (at the
higher frequency) is close to the sensor natural frequency. Figure 7.26(b) shows the input to the
accelerometer (bottom plot) and the output from the accelerometer (top plot). The measured
acceleration at the lower frequency is an accurate representation of the input vibration. However,
the measured acceleration signal at the higher frequency is misrepresented by the non-unity gain
that exists close to the sensor natural frequency and is amplified well beyond the true response
amplitude. As mentioned earlier this exaggeration of the vibration signal at those frequencies
could lead to incorrect machine condition assessments as well as damage the sensor.

There are two relatively simple ways to avoid the problem described above, where the natural
frequency can incorrectly amplify a vibration signal resulting in misinterpretation of the

125
vibration signal measurement. Figure 7.27 shows the frequency response curve for a higher
natural frequency accelerometer. Using an accelerometer with a higher natural frequency moves
the higher frequencies of non-linear response away from potential excitation frequencies. This is
a suitable solution as long as the excitation frequencies are known to stay within the new broader
linear range.

However, if the frequencies that are known to exist in the vibration signal are at or close to the
natural frequency of the accelerometer in use, the only way to avoid these signals showing up in
the analysis results as exaggerated responses is to remove them from the measured signal. This
removal is done by low-pass filtering the vibration signal before it is recorded (see Figure 7.28).
Most analyzers automatically set this filtering to occur at the same time as the sampling
frequency is set. This type of filtering is referred to as anti-aliasing filtering.

(a) (b)

Figure 7.26 Typical accelerometer frequency response curves with (a) vibration responses in the
linear range and close to the sensor natural frequency and (b) true (lower) and exaggerated
(upper) frequency responses close to the sensor natural frequency.

Figure 7.27 Frequency response curve for a higher natural frequency accelerometer.

126
Figure 7.28 Accelerometer frequency response curve and low pass filter response curve.

7.5.2 Sensitivity versus Mass

The mass of an accelerometer more than anything else defines the sensitivity. In general, large
mass accelerometers will be more sensitive than small mass accelerometers. The trade-off is that
the larger mass accelerometers will have a lower natural frequency that leads to a narrower linear
range of the frequency response curve. Figure 7.29 shows the relationship between a relatively
high mass accelerometer with high sensitivity, but narrow linear frequency response range
(lower natural frequency) and a relatively low mass accelerometer with lower sensitivity, but
broader linear frequency response range (higher natural frequency)

The mass of an accelerometer also becomes important when measuring vibrations of light weight
test objects. The additional mass of the sensor can significantly alter the vibration levels and
frequencies at the measuring point. The rule of thumb in this regard is that the mass of
accelerometer should be less than 1/10 of the mass of the vibrating component. This is not
typically a concern in machine condition monitoring and fault diagnosis applications as most
machines are relatively robust.

Figure 7.29 Accelerometer frequency response curves showing the relationship between sensor
sensitivity and natural frequency.

The sensitivity of an accelerometer is typically given in units of mV/G and is defined as the ratio
of the change in Voltage output over the change in acceleration (measured in Gs). Table 7.3
shows a list of common experiences and the associated accelerations (in Gs). Typical machine
induced vibration levels can be as high as 50Gs.

127
Table 7.3 Common experiences and the associated accelerations.

Most accelerometers are designed to have maximum sensitivity in one primary direction. The
transverse sensitivity is typically quiet low, but still enough to affect the primary vibration
measurement direction output if the transverse vibration levels are severe. Figure 7.30 shows an
accelerometer and its main axis of sensitivity, the transverse sensitivity and the axis of maximum
sensitivity.

Figure 7.30 Transverse sensitivity in accelerometers.

7.5.3 Design and Operation

The word "Piezo" is a Greek term which means to squeeze. When strained, piezoelectric
elements create a charge. Crystalline quartz is a stable and sensitive piezoelectric material which
has been used in accelerometers for many years. However, man-made ceramic material has
largely replaced quartz because it can be designed for a wider range of applications and has a
longer stable useful life.

Figure 7.31(a) shows a schematic piezoelectric crystal reacting to an applied load. The strain that
results from an applied load (applied force) causes the ions in the crystal to react by grouping
together with other ions of like charge. This behavior takes place regardless of the type of strain.
Figure 7.31(b) shows this reaction for compressive, flexural, and shear strain.

128
(a) (b)

Figure 7.31 Schematic showing piezoelectric crystals reacting to applied loads.

The term ICP is a registered trademark of the accelerometer manufacturer PCB. It stands for
Integrated Circuit Piezoelectric. The built-in electronics in an ICP accelerometer convert the
high-impedance charge signal that is generated by the piezoelectric sensing element into a usable
low-impedance voltage signal that can be readily transmitted over ordinary two-wire or coaxial
cables to any voltage readout or recording device. The electronics within ICP accelerometers
require excitation power from a constant-current regulated, DC voltage source. Most data
collectors and FFT analyzers contain internal analog to digital conversion cards that also provide
the excitation needed by accelerometers (Figure 7.32(a)). Alternatively, an external power supply
can also be used (Figure 7.32(b)).

(a) (b)

Figure 7.32 Schematic showing two ways of providing excitation power to an accelerometer with
internal circuitry (a) and a photograph of an external excitation power supply.

Compression mode accelerometers are a common design because of their simplicity and
robustness. Figure 7.33 shows a pair of diagrams displaying the internal construction of
compression mode accelerometers. The piezoelectric material is held in place between the
accelerometer mass above and a substantial base. When acted on by a force the peizo-electric
crystal is compressed (or stretched) between the mass and the base. This strain causes the
piezoelectric effect. Also shown are the built in electronics and the connector. Figure 7.34 shows
a pair of diagrams displaying compression mode accelerometers each with a different design of
isolating baseplate. These plates are sometimes needed to isolate the crystal from heat or bending
strain not related to the vibration signal. Figure 7.35 shows another design configuration for a
compression mode accelerometer and some of the many different shapes and sizes available.

129
Figure 7.33 Compression mode accelerometers.

Figure 7.34 Compression mode accelerometers with crystal isolation.

Figure 7.35 Compression mode accelerometers come in many shapes and sizes.

Shear mode accelerometers are another design style that is popular. This design type allows bias
compensation to be easily implemented when uneven crystal loading occurs due to baseplate
distortion during mounting. Figure 7.36 shows the internal configuration of two different shear
mode accelerometers. Figure 7.37 shows the popular delta shear design.

130
Figure 7.36 Shear mode accelerometers.

Figure 7.37 Delta shear type accelerometer.

Flexure mode accelerometer s are used when high sensitivity is needed. Figure 7.38 shows a
schematic diagram of a flexure mode accelerometer.

Figure 7.38 Flexure mode accelerometer.

Most accelerometers are uni-axial. That is, they are sensitive to acceleration forces only in one
direction. The technical specifications for a general purpose accelerometer are listed below to
give an example of the performance details that can be expected.

131
High sensitivity, quartz shear ICP accel., 100 mV/g, 2.5 to 10k Hz, 10-32 top
connection, ground isolation
Sensitivity: 100 mV/g (10.2mV/(m/s))
Measurement Range: 50 g pk (490m/s pk)
Broadband Resolution: 0.002 g rms (0.02m/s rms)
Frequency Range: 2.5 to 10000 Hz
Weight: 1.2 oz (35gm)

Tri-axial accelerometers are sensors that have the built in capability of measuring acceleration in
three dimensions simultaneously. The technical specifications for a tri-axial accelerometer are
listed below to give an example of the performance details that can be expected.

Tri-axial, thru-hole mounting, ceramic shear ICP accelerometer, 10 mV/g, 0.5 to 2k Hz,
ground isolated, 4-pin conn.
Sensitivity: 10 mV/g (1.02mV/(m/s))
Measurement Range: 500 g pk (4905m/s pk)
Broadband Resolution: 0.0005 g rms (0.005m/s rms)
Frequency Range: 0.5 to 2000 Hz
Electrical Connector: 1/4-28 4-Pin
Weight: 0.55 oz (15.5gm)

Other types of accelerometers include miniature, high temperature resistance, high sensitivity,
low frequency sensitive, shock resistant, and many more.

Figure 7.39 Some of the many different types of accelerometers.

7.5.4 Calibration

Accelerometers are designed to retain a constant sensitivity for long periods of time. However,
they may need calibrating if damaged by dropping or high temperatures. A known amplitude and
frequency source (or another accelerometer that has a known calibration) should be used to
check the calibration of accelerometers from time to time. Figure 7.40 shows a calibration check

132
setup for accelerometers. If this type of calibration equipment is not available a quick calibration
check can be performed by holding two accelerometers together and shaking them. The response
from both should be the same. Of course, most manufacturers and some testing labs also provide
calibration services.

Figure 7.40 Calibration check setup for an accelerometer.

7.5.5 Mounting and Location

Installation of accelerometers requires as rigid a mount as possible. Permanent installation with


studs or bolts is usually best for high speed machinery where high-frequency vibration
measurements are required. The close coupling between the machine and the sensor allows for
direct transmission of the vibration to the sensor. Stud mounting requires a flat surface to give
the best amplitude linearity and frequency response. This type of mounting is expensive and may
not be practical if large numbers of measurements are being recorded with a portable instrument.

Magnetic mounts have the advantage of being easily movable and provide good repeatability in
the lower frequency range, but have limited high-frequency sensitivity (4 to 5 kHz). Hand-held
measurements are useful when conducting general vibration surveys, but usually result in
significant variation between measurements. The hand-held mount is least expensive but only
offers frequency response below 1 kHz. For machine condition monitoring and fault diagnostics
applications, there will typically be a combination of these three mounting methods used,
depending on the equipment being monitored and the monitoring strategy employed.

Some other common mounting methods are to use epoxy or some other permanent adhesive for
mounting or wax as a temporary mounting adhesive. The type of mounting affects the frequency
response curve of the accelerometer. The common mounting methods and the frequency limits
for a 100 mV/g accelerometer are listed below.

Hand-held (500 Hz)


Magnetic (2,000 Hz)
Adhesive (2,500 4,000 Hz)
Wax (5,000 Hz)
Threaded (6,000 10,000 Hz)

Figure 7.41 shows a plot of the response curves for the same accelerometer with different
mounts. The more permanent and rigid the mount the higher the natural frequency and the wider
the linear frequency response.

133
Figure 7.41 Accelerometer response versus frequency for various types of mounts.

A threaded mount (stud or screw) is usually used for permanent sensor placement. It offers the
best frequency response, particularly in the high frequency range. Only threaded mounts allow
sensors to respond approaching the designed transducer specifications. Figure 7.42(a) shows a
cross section of a threaded accelerometer mount. Adding a coupling fluid (grease or oil)
improves vibration transmissibility by filling small voids in the mounting surface. This
consequently increases the mounting stiffness. For permanent mounting substitute the fluid with
an epoxy or other type of adhesive. Isolating the sensor through the mount is useful when
electrical isolation is required. A mica washer and isolated stud is typically used.

(a) (b)

Figure 7.42 Accelerometer threaded mount.

A magnetic mount will only work with compatible surface materials. They should not be used
for acceleration amplitudes exceeding 200Gs. If a magnetic surface is not available a metal
washer or plate can be epoxied into position and used as a magnetic pad for the accelerometer.
Magnetic mounts allow for good repeatability with the accelerometer being removable and re-
installable. A magnetic mount is a good option for non-permanent installations. If the surface is
not level a mounting pad can be used (epoxied in place or a dual rail mount can also be used on
curved surfaces (common on machines). Magnetically mounted accelerometers should not be
placed onto a machine vertically. Contact should be made with the machine surface with the
tilted edge of the accelerometer and then it should be gently tilted towards the vertical in order to

134
avoid damaging the sensor with a severe vertical landing. Figure 7.43 shows several different
options for magnetically mounting accelerometers.

(a) (b)

Figure 7.43 Accelerometer magnetic mounts.

Wax should be used only as a temporary mounting method. It is not suitable for high temperature
situations. Hand-held accelerometer mounting can be performed with a round or sharp pointed
probe tip. This method is convenient, but offers poor repeatability in measurements and is not
suitable for frequencies over 1 kHz.

Figure 7.44 Accelerometer hand-held mounting.

Transducer Location

In general, transducers should be placed as close to bearings as possible (and within the load
zone). Radially placed transducers are to measure radial forces such as due to mass unbalance.
Axially placed transducer are to measure axially-directed forces such as due to misalignment.
Figure 7.45 shows the locations and directions that should be used for vibration measurements.
Note that these locations may result in redundant data being collected. Only after analysis can the
decision be made to reduce the number of measurement locations in order to stream-line the
monitoring program.

Measurements should also be made as close to the vibration source as possible. Figure 7.46
shows how material transitions should be avoided if possible and the sensors should be placed as

135
close to bearings as possible. The primary reason for this is that there is a significant amount of
vibration signal transmission loss in homogeneous and non-homogeneous materials. The farther
the measurement point from the source of the vibration, the weaker the vibration forces at the
measurement point. Figure 7.47 shows this relationship for homogeneous and non-homogeneous
materials.

Figure 7.45 Accelerometer mounting directions.

Figure 7.46 Accelerometer mounting locations (close to bearings).

Figure 7.47 Vibration signal transmission loss in homogeneous and non-homogeneous materials.

136
Vibration measurements should always be made within or as close to the load zone as possible.
Figure 7.48 and Figure 7.49 show the load zones for pulleys and gears respectively.

Figure 7.48 Pulley load zones.

Figure 7.49 Gear load zones.

As with the other types of vibration sensors, accelerometers have certain limitations. Because of
their sensitivity and wide dynamic range, accelerometers are also sensitive to environmental
input not related to the vibration signal of interest. Temperature (ambient and fluctuations) may
cause distortion in the recorded signal. General purpose accelerometers are relatively insensitive
to temperatures up to 250C. At higher temperatures, the piezoelectric material may depolarize
and the sensitivity may be permanently altered. Temperature transients also affect accelerometer
output. Shear-type accelerometers have the lowest temperature transient sensitivity. A heat sink
or mica washer between the accelerometer and a hot surface may help reduce the effects of
temperature.

137
Figure 7.50 Accelerometer sensitivity versus ambient temperature fluctuation.

Piezoelectric crystals are sensitive to changes in humidity. Most accelerometers are epoxy
bonded or welded together to provide a humidity barrier. Moisture migration through cables and
into connections must be guarded against. Large electro-magnetic fields can also induce noise
into cables that are not double shielded.

If an accelerometer is mounted on a surface that is being strained (bent), the output will be
altered. This is known as base strain, and thick accelerometer bases will minimize this effect.
Shear-type accelerometers are less sensitive because the piezoelectric crystal is mounted to a
center post not the base.

7.5.6 Cables and Ground Loops

Repeated bending of a cable can cause tribo-electric effects resulting in a noisy measurement.
The easiest way to guard against this type of noise entering the measured signal is to secure all
cables to prevent bending and other cable movement.

Figure 7.51 Cables should be secured.

Ground loops develop between a sensor and a data collector if the two devices are grounded at
different potentials. Noise that is transmitted between the sensor and the data collector as a result
of the different potential (see Figure 7.52) can be difficult to detect. It is prudent to take steps to
prevent this situation. Solutions include establishing a connection between the two ground
potentials (closing the loop, see Figure 7.53(a)) and/or isolating the sensor with an electrically

138
isolating washer (see Figure 7.53(b)). A mica washer can act to electrically isolate the sensor and
thermally isolate the sensor.

Figure 7.52 Accelerometer ground loops may cause noise in the measured signal.

Figure 7.53 Accelerometer ground loops may be removed using (a) a common ground and (b) an
isolated sensor.

7.5.7 Advantages and Disadvantages

In summary the following advantages and disadvantages can be listed in association with the use
of accelerometers.

139
Table 7.4 Advantages and disadvantages of accelerometers.

Advantanges Disadvantages
Measures casing or structural absolute motion Sensitive to mounting technique and surface
condition
Easily attached to machinery, piping, Unable to measure shaft vibrations or position
baseplates or structures directly
Good signal response between 900 and External power source required
600,000 CPM (15 to greater than 10,000Hz)
Flat phase response throughout transducer Low dynamic signal response below 600 CPM
operating range (10 Hz)
Solid state electronics with rugged and reliable Transducer cable sensitive to noise, motion and
construction electrical interference
Operates below mounted natural resonant Temperature limitation of 250 F (120 C) for
frequency ICP Transducers
Special units are available for high temperature Extended frequency range often requires signal
applications filtration
Double integration often suffers from low
frequency noise

Chapter Summary

A transducer is a device that senses a physical quantity and converts it into an electrical
output signal, which is proportional to the measured variable.
The selection, placement, and proper use of the correct transducer are important steps in the
implementation of a condition monitoring and fault diagnostics program.
The transducer must be correct for the task, properly mounted, in good working order
(properly calibrated), and fully understood in terms of operational characteristics.
Traditional vibration sensors fall into three main classes: noncontact displacement
transducers, velocity transducers, and accelerometers.
Transducers are selected to optimize sensitivity over the frequency range that is expected to
be recorded.

140
Chapter 8

Signal Collection and Analysis Instrumentation and Analysis Methods


Vibration data are acquired from a machine by a transducer that converts the mechanical
vibration to an electrical signal (typically in volts). The quality of the signal depends on
transducer selection, mounting and location (among other things). Data acquisition is key to
condition monitoring and fault diagnosis. Quality data acquisition requires sound and cost-
effective planning.

A review of measurement parameters shows the following. Displacement can be measured in


absolute or relative terms. Absolute displacement is typically used for the measurement of low
frequency structural vibration (0 - 20 Hz). It relates to stress and motion and can be measured
with a double integrated accelerometer or contacting transducer. Relative displacement
transducers typically have a wider frequency range (0 1000 Hz). It is measured with a
proximity probe (non-contacting transducer) and relates to stress and motion.

Velocity is a more general measure for machinery monitoring and analysis. It covers the medium
frequency range (10 1,000 Hz) and relates to component energy and fatigue. Velocity is
measured with an accelerometer with integration or an electro-mechanical device. Velocity
measurements are often used in rough assessments of condition in the time domain without
considering frequency.

Acceleration is the most common general measure for machinery monitoring and analysis.
Acceleration sensors are particularly sensitive at higher frequencies (>1,000 Hz), but also has a
wide frequency response range (1 10,000 Hz). Acceleration relates to force and the sensors
used are called accelerometers. This measure is particularly suitable for gear mesh and rolling
element bearing defects.

Table 8.1 Summary of vibration measurement parameters.

Measurement Frequency Physical Application


Parameter Span Parameter
Relative 0 1,000 Hz Stress/Motion Relative motion in bearings and
Displacement casings
Absolute 0 20 Hz Stress/Motion Structural motion
Displacement
Velocity 10 1,000 Hz Energy General machine condition,
medium frequency vibrations
Acceleration 1 10,000 Hz Force General machine condition,
medium to high frequency
vibrations

Knowing the frequency span that is suitable for the different types of measurement parameter
(and sensor) is important when deciding the type of hardware to employ for collecting vibration
data. It is also important to know the appropriate frequency span for different faults conditions

141
that may be present in the vibration signal. Table 8.2 shows some default frequency spans that
should be used to collect data when measuring vibrations from different machines. Table 8.3
shows more examples of measurement parameters and frequency spans for selected machines.

Table 8.2 Default frequency spans when measuring vibrations from different machines.

Machine Component Frequency Span of Response


Shaft vibration 10 x RPM
Gearbox 3 x Gear mesh
Rolling element bearings 10 x Ball pass inner race
Pumps 3 x Vane pass
Motors and generators 3 x 2 x Line frequency
Fans 3 x Blade pass
Journal bearings 10 x RPM

Table 8.3 Measurement parameters and frequency spans for selected machines.

Machine HP/MW Speed (RPM) Bearing Type Measure(s) Transducer


Frequencies (Hz)
Gearbox- 9 MW 7,500 RPM input fluid film displacement-shaft proximity
single 1,200 RPM output acceleration-casing probe
reduction GM = 3 000 Hz accelerometer
Gearbox- 400 HP 1,800 RPM input 15 rolling acceleration accelerometer
double 200 RPM output elements
reduction GM = 375 725 Hz
Steam 18,000 HP 5,000 RPM fluid film displacement-shaft proximity
Turbine probe
Steam 500 MW 3,600 RPM fluid film displacement-shaft proximity
Turbine probe
Gas Turbine 20 MW 9,000 RPM fluid film displacement-shaft proximity
rolling element acceleration-casing probe
accelerometer
Gearbox- 9 Mw 7,500 RPM input fluid film displacement-shaft proximity
single 1,200 RPM output acceleration-casing probe
reduction GM = 3,000 Hz accelerometer
Gearbox- 400 HP 1,800 RPM input 15 rolling acceleration accelerometer
double 200 RPM output elements
reduction GM = 375 & 725
Hz
Steam 18 000 HP 5,000 RPM fluid film displacement-shaft proximity
Turbine probe
Steam 500 MW 3,600 RPM Fluid film displacement-shaft proximity
Turbine probe

142
Table 8.3 Measurement parameters and frequency spans for selected machines (Contd).

Machine HP/MW Speed (RPM) Bearing Type Measure(s) Transducer


Frequencies (Hz)
Gas Turbine 20 MW 9,000 RPM fluid film displacement-shaft proximity
rolling element acceleration-casing probe
accelerometer
Large 4,000 HP 3,600 RPM fluid film displacement-shaft proximity
Induction probe
Motor
Induction 200 HP 1,800 RPM 8 rolling velocity-casing integrated
Motor element accelerometer
or velocity
Diesel 400 HP 1,800 RPM fluid film velocity-casing integrated
Engine accelerometer
or velocity
High- 18,000 HP 5,000 RPM fluid film displacement-shaft proximity
Performance velocity-casing probe
Centrifugal integrated
Pump accelerometer
or velocity
Centrifugal 200 HP 1,800 RPM 12 rolling velocity-casing integrated
Pump element accelerometer
Reciprocatin 200 HP 300 RPM 15 rolling velocity-casing integrated
g Pump element accelerometer
or velocity
Centrifugal 1,000 HP 5,000 RPM fluid film displacement-shaft proximity
Compressor probe
Reciprocatin 500 HP 480 RPM fluid film velocity-casing integrated
g accelerometer
Compressor or velocity
Dryer Roll 300 RPM 26 rolling velocity-casing integrated
element accelerometer
or velocity

The tables above show the type of data that can be collected from a wide variety of machines
using vibration sensors. However, knowing the type of sensor to use and the vibration frequency
span that should be recorded is not enough to allow complete analysis of machine condition
based on vibration signals. There are other details related to where exactly to measure the
vibration and how to make these measurements and carry out the vibration signal analysis. Some
basic vibration signal analysis was discussed in Chapter 6, but the following will highlight more
of the detail needed for accurate fault detection and diagnosis.

143
Figure 8.1 shows a few of the wide variety of machines that could benefit from vibration based
machine condition monitoring and diagnostics. Knowing where and how to collect useful
vibration data from these (and other) machines is not a simple task. There are many factors to
consider. The first part of this chapter steps through some of the basic considerations that need to
be factored into these types of decisions. The two most basic and simple strategies for success
are to i) keep things as simple as possible and ii) remember to consider the context.

Figure 8.1 Industrial machines.

Keeping things simple is self explanatory. Using simple tools and straight forward methods can
often be the key to solving the majority of problems. Considering the context within which a
machine condition is defined may or may not involved more complex data collection and
analysis. It may simply mean that one considers all the different data sources available (vibration
and non-vibration) before proceeding to more complex data collection and analysis methods.
Information such as the age of a component or machine, the failure history, the mode of
operation (speed, acceleration, duty, etc.) are all important when outlining a vibration data
collection and analysis strategy.

Most industrial machines have four main states of operation during which data acquisition can
take place. These are; off, run up, steady state, and coast down.

Steady state data is taken while the machine is held at steady operating conditions. Frequencies
linked to the machine kinematics remain constant and this allows the use of basic signal
processing methods to easily establish a baseline and monitor for changes.

Run up (when the machine is first turned on) data is taken as the machine starts and accelerates
to full steady state running speed. Run up data collection is often used to pass new or re-built
machinery at commissioning. Collecting run up data often takes a machine through the resonant
(natural) frequencies. This must be done quickly so as not to excite resonances. This type of data
collection often requires extra signal processing techniques to extract characteristic frequencies.

144
With the speed of the machine changing throughout the data collection period the characteristic
frequencies will change requiring careful analysis.

Coast down (run down) data is taken while the machine decelerates from full speed down to a
full stop. This type of data collection is the same similar to run up in that the speed of the
machine is changing constantly during the data collection stage. If the coast down is done
without breaking it may reveal more of the natural frequencies below the running speed because
there will be no acceleration (or deceleration) forces acting to artificially stiffen the structure.

Figure 8.2 Vibration signal spectra from run up tests.

Vibration measurements can also be collected while the machine is at rest (switched off). This
type of data collection and analysis is called modal analysis. When the machine is at rest it may
be excited with an impulsive input. Such a test is called an impact or bump test and involves
striking the machine with an instrumented hammer to cause a transient input excitation. The
input and output are both measured and used to determine the systems transfer function and
resonances. A steady state sinusoidal input or a variety of sinusoidal inputs (referred to as a
swept sinusoidal excitation) may also be used. Modal analysis will be discussed more fully later
in this chapter.

8.1 Basic Vibration Meters

8.1.1 Oscilloscopes

Time domain instruments are generally only able to provide a time domain display of the
vibration waveform. Some devices have limited frequency domain capabilities. While this
restriction may seem limiting, the low cost of these devices and the fact that some vibration
characteristics and trends (such as shaft displacements (orbits) in fluid film bearings, and
transients) show up well in the time domain make them valuable tools. Synchronous time

145
averaging (and negative averaging) also falls into time domain signal analysis best performed
using these tools.

Oscilloscopes are the most common form of this device. Shaft displacements (orbits), transients
and synchronous time averaging (and negative averaging) are some of the analysis strategies that
can be employed with this type of device. A transducer converts the mechanical vibration into a
proportional electrical signal (typically in Volts). The electrical signal can be converted into
engineering units (EU) with a scale factor (i.e. mV/G). The oscilloscope measures and displays
voltages versus time. Time waveforms, orbits and markers can be displayed by oscilloscopes.

Figure 8.3 Typical digital oscilloscope.

Oscilloscopes are all digital now, but analogue devices are still relatively common in some
laboratories and industries. Oscilloscopes are commonly used for triggering. This form of
triggering is used to initiate data acquisition at a specified time or amplitude and control data
acquisition. Auto trigger is used for continuous sampling. Slope or magnitude based triggering is
used for sampling at specified conditions. Oscilloscopes can also make use of an external
intensity input to measure phase and to relate a shaft marker position to another signal
representing the vibration amplitude.

8.1.2 Overall Vibration Level Meters

Vibration meters are generally small, hand-held (portable), inexpensive, simple to use, self-
contained devices that give an overall vibration level reading. They are used for walk-around
surveys and measure velocity and/or acceleration. Generally, these devices have no built-in
diagnostics capability, but the natural frequency of an accelerometer can be exploited to look for
specific machinery faults.

Figure 8.4 Typical overall vibration level meter.

146
As an example, rolling element bearings generally emit spike energy during the early stages of
deterioration. These are sharp impacts as rollers strike defects (pits, cracks) in the races. A spike
energy meter is an accelerometer that has been tuned to have its resonant frequency excited by
these impacts, thus giving a very early warning of deteriorating bearings. See section 8.1.3
below.

8.1.3 Shock Pulse (Spike Energy) Meter

A shock pulse (or spike energy) meter in an accelerometer that has been electrically and
mechanically tuned (during device design) to 32 kHz. This frequency is an ideal carrier
frequency for transients caused by shocks. These devices are a special subset of the more general
purpose vibration level meter. They are simple, inexpensive, good for fault detection only
particularly in high speed rolling element bearings. It should be noted that this devices are load
dependant so changes in load will affect the output and could be mistaken for changing machine
condition.

Figure 8.5 Typical shock pulse (spike energy) meter.

The shock pulse method of vibration signal analysis is specifically for use in analyzing signals
from rotating rolling bearings and can form the basis of an efficient machine condition
monitoring program. From the innovation of the method in 1969 it has now been further
developed and broadened and is now a worldwide accepted methodology for condition
monitoring of rolling bearings and machine maintenance.

The difference between shock pulse and regular vibration signals is that the shock pulse signal is
the short duration impact excitation caused by an impulsive event. Regular vibration caused by
other machine related forcing functions is lower in energy and low in frequency. Figure 8.6
shows a shock pulse vibration response in relation of a regular vibration signal.

It may seem dangerous to excite an accelerometer at its natural frequency, but because the early
stages of bearing deterioration typically exhibit relatively low energy very short duration spikes,
there is little chance that the sensor will be damaged. The high sensitivity at the resonance

147
frequency is also typically accompanied by damping or filtering to remove excitation at other
frequencies.

Figure 8.6 Shock pulse vibration compared to a regular vibration signal.

Figure 8.7 shows a shock pulse meter and several bearing fault signals that show a gradually
deteriorating condition. Figure 8.7(a) shows s series of vibration signals with the typical spikes
beginning to appear and than becoming more frequent within the signal as the fault worsens.
Figure 8.7(b) shows the initial shock pulse pressure wave (excited by the roller interaction with a
crack or pit on the raceway), the damped oscillation of the accelerometer (at or close to 32 kHz)
and the output shock pulse level signal.

(a) (b)

Figure 8.7 Shock pulse meter and associated bearing fault signals.

148
8.2 Data Collectors

Most vibration data collectors available today for use in machine condition monitoring and fault
diagnostics are microcomputer based. They are used together with vibration sensors to measure
vibration, to store and transfer data, and for frequency domain analysis. Considerably more data
can be recorded in a digital form, but the cost of these devices can also be considerable. Another
advantage of most data collectors is the ability to use these devices to conduct on-the-spot
diagnostics or balancing. They are usually used with a desktop personal computer to provide
permanent data storage and a platform for more detailed analysis software. Data collectors are
usually used on general-purpose equipment. Simple devices are good for monitoring overall
levels and trending.

The strength of modern data collectors are the analysis capabilities that are built in. These
depend on the supporting software of course. Electronic data collectors may be programmed to
acquire and store selected vibration parameters such as overall vibration, spectra, and various
plots. They are capable of performing trending and comparing with previously-collected
information. Most collectors include FFT algorithm as a standard feature and have high
resolution (from 100 to 64,000 lines representing the frequency spectra).

Data collectors come in many different shapes and sizes. Figure 8.8 show examples of large (old)
and small (new) data collectors. Figure 8.9 shows examples of many more data collectors.
Specifics and brand names are changing regularly, but there seems to be a relatively broad
selection regularly available. Cost of course depends largely on capability. There is a continuing
trend toward smaller devices that can be used via USB connection with laptops. I this way the
data sampling (analogue to digital conversion) remains onboard the data collection device, while
the data analysis software resides on the laptop and benefits from ever increasing computer speed
for data processing tasks.

Figure 8.8 Examples of large (old) and small (new) data collectors.

149
Figure 8.9 More examples of data collectors.

Figure 8.10 Portable data collectors in use.

8.3 Frequency Domain FFT Analyzers

The frequency domain analyzer is perhaps the key instrument for diagnostic work. Different
machine conditions (unbalance, misalignment, looseness, bearing flaws) all generate
characteristic patterns that are usually visible in the frequency domain. While data collectors do
provide some frequency domain analysis capability, their main purpose is data collection.
Frequency domain analyzers are specialized instruments that are specialized for the analysis of
vibration signals. As such, they are often treated as a laboratory instrument. Generally, analyzers
will have superior frequency resolution, filtering ability (including anti-aliasing), weighting
functions for the elimination of leakage, averaging capabilities (both in the time and frequency
domains), envelope detection (demodulation), transient capture, large memory, order tracking
capability, cascade/waterfall display, and zoom features. Dual-channel analysis is also common.

Figure 8.11 Frequency domain FFT analyzer.

150
8.3.1 Order Tracking Analyzers

Tracking analyzers are typically used to record and analyze data from machines that are
changing speed. This usually occurs during run-up and coast-down of large machinery or turbo-
machinery. These measurements are typically used to locate machine resonances and unbalance
conditions. The tracking rate is dependent on filter bandwidth, and there is a need for a reference
signal to track speed (tachometer input). These devices usually have dual channel input, variable
input sensitivity and a large dynamic range.

8.3.2 Selecting the Correct Analyzer

All FFT Analyzers come with a complete set of specifications. These specification sheets can be
used to select the correct analyzer for a particular application. An example is set out below where
a variety of factors are highlighted for the purpose of analyzer selection.

Type of signal or
transducer

Speed input is important


for rotating machinery
Record keeping
capability

Enveloping for bearing


frequencies

Basic Filtering Capability

151
Transducer specific parameters

The range of signal we can


measure

Based on PC processor

Memory (Is 6MB enough?)

Sample rate of acquisition.

Useful for gear


measurements.

Useful for inspecting


spectra and tracking live
data
Triggers for: Basic bump
test, event triggered
acquisition.

152
Dictated by maximum sample rate

An important parameter

Windows

Automatically adjusting filter (wrt


speed)

Observing individual bands

Proprietary database of bearing


characteristic frequencies

153
8.4 Data Recorders

Tape recorders (for historical information) are direct recording devices for the long term storage
of vibration signals. While they may still be used in some situations most modern data collectors
use hard drive digital storage devices. Standard analogue direct tape recorders are very rare now.
They have a relatively poor response at low frequencies. A high frequency response requires a
high tape speed. Frequency Modulated (FM) recorders have good low frequency response (0 - 5
kHz) and a more linear reproduction across all frequencies. Digital Audio Tape (DAT) recorders
are smaller in size, have a large dynamic range (70 dB), are accurate, have good frequency
response and typically have built in filters for anti-aliasing filtering before the analogue to digital
conversion.

Regardless of the type of data storage system being used there are a few general guidelines that
should be followed when recording data. One should record carefully what the recordings
represent, make regular calibration checks (internal and external) as well as monitor the input
while recording to watch for anomalies in the data.

Calibration instruments such as multi-meters, signal generators, shakers, filters, envelope


detectors are all very important. They need to be handled with care as this is the reference for all
recordings.

8.5 Vibration Signal Sampling Alternatives

There are two main types of data sampling commonly conducted. These are i) uniform time
sampling and ii) uniform position sampling.

8.5.1 Uniform Time Sampling

Uniform time sampling is the most common type of vibration data sampling in machine
condition monitoring and fault diagnosis applications. Samples are taken at uniform time
intervals. This type of sampling is used on steady state tests and impulse tests. Uniform time
sampling may also be used for run up and run down tests if the speed rate of change is slow.

If uniform time sampling is used when the shaft is speeding up (is slowing down) the sampling
will not represent constant locations on the shaft (see Figure 8.12). The frequency of the sampled
sinusoid changes during data acquisition and smearing takes place in the frequency spectrum.

8.5.2 Uniform Position Sampling

This type of data sampling is also known as order tracking. The sampling rate changes as a
function of the machine operating speed. Samples are taken at uniform position intervals. This
type of sampling is used for run up and coast down tests when the rate of change of speed is
high. This type of sampling is shown in Figure 8.13. Note that the frequency remains constant
and there is no smearing in the frequency spectrum.

154
Figure 8.12 Uniform time sampling when the shaft is changing speed.

Figure 8.13 Uniform position sampling when the shaft is changing speed.

155
Variable sample rate can be controlled by hardware (adjustable filters, the old way) or software
(oversampling and then resampling at variable rate). The hardware method is seldom used now
as it cannot keep up with rapid changes in speed.

8.6 Data Sampling Rates and Resolution Issues

Data sampling is the process of converting from an analogue input to digitized data prior to
storage and/or FFT processing. The number of samples stored in the analyzer buffer depends on
the number of lines selected. The computer records the values as equally-spaced Y (amplitude)
and X (time/position) components. Resolution is determined by the sampling time, (1/Ts). No
frequency lower than 1/Ts can be resolved. The number of bins (lines) dictates the number of
samples (Nsamples = 2 Nlines). The sampling rate must be at least 2 times the highest frequency of
interest (Fs = 2 Fmax). The data acquisition time equals the number of bins (or lines) divided by
the maximum frequency (Fmax).

Original analog Digitally sampled


amplitude
amplitude

time time

Figure 8.14 An analogue signal and the digitally (uniform time) sampled equivalent.

A general data acquisition signal flow chart is shown in Figure 8.15. The only slight difference
between this flow chart and typical digital data collection processes is the presence of the
multiplexer. Modern data collection systems typically have one analogue to digit conversion chip
per channel. Older units may rely on a multiplexer to store and than distribute sampled values to
a single A/D converter. The reduced cost and size of these electronics means that most systems
have dedicated A/D converters for each data channel.

Figure 8.15 Data acquisition signal flow chart.

156
The Resolution of a recorded signal is the smallest voltage increment that will cause a bit
change. Higher resolution leads to a smaller detectable voltage changes and better representation
of the amplitude being measured as a function of time. However, higher resolution results in
more data and a balance must be maintained. Equation 8.1 defines the relationship between
resolution (Q), the full scale votage range (EFSR) and the number of bits required to represent the
level of resolution (M).
EFSR
Q 8.1
2M
Figure 8.16 shows the signal resolution possible with 3-bit resolution. As is clear from Figure
8.16 the resolution is somewhat course resulting in a poor ability to distinguish between closely
spaced amplitude levels. Figure 8.17 shows the resolution possible with a variety of different
numbers of bits available for amplitude representation.

Figure 8.16 Signal resolution possible with 3-bit resolution.

Figure 8.17 Signal resolution possible with a variety of different bit resolutions.

157
As an example the following in is a calculation of how many bits would be required to achieve a
resolution of 0.001% of full scale. Required resolution: 0.001% = 0.001/100 = 0.00001

EFSR 1
Q 0.00001
2M 2M
Solving for M gives M = 16.6 and rounding up gives 17 bits. Therefore, 17 or more bit will
allow a resolution of greater than 0.001%.

The number of channels that are recorded and/or analyzed depend on how many signals may
require analysis at one time. Here is of course a cost associated with collecting multiple channels
of data. Here are two choices for sampling methods. The first option is to sample and hold using
a multiplexer and then send the signal (one at a time) to a single A/D converter. The second
choice is to have dedicated circuitry for each channel. As mentioned above, most modern
analyzers have a dedicated A/D converter for each channel on board.

However, given that there are still multiplexed analyzers in use it may be useful to discuss how
to handle data analysis on these devices. Basically a multiplexer will sample and hold a value for
each channel simultaneously (using sample and hold circuits). The multiplexer circuit then scans
each captured voltage and converts the analogue sample to a digital representation one at a time.
Phase relations are preserved. In most cases the user is allowed to control the number of scanned
channels from software. This way, the multiplexer is not scanning unused channels. The fewer
channels in use, the faster the acquisition rate.

8.7 Aliasing

When digitally sampling an analogue signal, one of the main concerns is aliasing of the sampled
signal. For uniform time sampling, the sampling occurs at a specific rate with respect to time.
This is called the Sampling Frequency (fs). As a general rule, the sampling frequency used (fs)
must be at least twice the desired maximum measured frequency (fm).

( fs >= 2 x fm ) 8.2

Conversely, the Nyquist Frequency defines the highest frequency that can be accurately
analysed from data sampled at a given rate. This frequency is the sampling rate. The practice
of sampling twice as fast as the frequency of interest occurs does not ensure that the sampled
signal will be an accurate representation of the analogue waveform. Experience suggests that at
least 8 to 10 times the highest expected frequency should be the sampling frequency.

If a signal is measured at a sampling rate that is too slow, aliasing may occur. Aliasing is when a
sampled signal represents an analogue signal that is not actually present in the original signal.
Figures 8.18-8.20 show sinusoidal signals that are sampled too slowly and the resulting (aliased)
sampled signal.

158
Figure 8.18 Original analogue signal sampled above the Nyquist frequency (left) and the same
signal sampled below the Nyquist frequency resulting in an aliased sampled signal (right).

Figure 8.19 A 30 Hz signal (original) sampled at 1000 samples per second (top left) and the same
30 Hz signal sampled at 100 Hz (top right), 25 Hz sample frequency (bottom left), and 11 Hz
sample frequency (bottom right). The top two sample frequencies are above the Nyquist
frequency (no aliasing), while the bottom two sample frequencies are below the Nyquist
frequency (aliasing occurs).

159
(a) (b)

(c) (d)

Figure 8.20 Frequency spectra calculated from a 30 Hz signal sampled at (a) 1000 samples/sec,
(b) 100 samples/sec, (c) 25 samples/sec and (d) 11 samples/sec. (c) and (d) are aliased.

8.7.1 Anti Aliasing Filters

Aliasing is obviously not a problem if the sampling can always be done fast enough to be well
above the Nyquist frequency. However, there are cases it may not be possible or practical to
sample fast enough to avoid all potential aliasing. In these cases the high frequency components
most be removed (filtered) from the measured signal before the signal is digitized. These filters
are known as anti-aliasing filters.

An ideal anti aliasing filter would block 100% of the signal above a specified cut off frequency.
In the case shown in Figure 8.21 the ideal filter would pass all frequencies in the PASSBAND
frequencies and completely block all frequencies in the STOPBAND. Of course ideal anti
aliasing filters do not exist in reality. Generally, anti-aliasing filters are set to about 80% of the
sampling frequency to account for the transition zone (see Figure 8.22). Because frequencies
near the transition zone (both below and above) can be aliased, this frequency range is
considered to be unreliable for analysis purposes. In most cases the filter cut off frequency is set

160
well below the Nyquist frequency so that the frequency range used for any analysis will be free
of all potentially aliased components. Once a signal passes through an anti-aliasing filter it is said
to be band limited.

Figure 8.21 Ideal low pass (anti aliasing) filter.

Figure 8.22 Realistic low pass (anti aliasing) filter.

There are four basic types of signal low pass filters. These are shown in Figure 8.23.

Figure 8.23 Realistic low pass (anti aliasing) filter.

161
8.8 Time Synchronous Averaging

Time synchronous averaging is the averaging together of a set of vibration signal samples that
have the same trigger point and the same length. That is, the sample segments represent the exact
same number of machine cycles and they are averaged over time. The result is that the
synchronous components of the signal are reinforced and the non-synchronous components are
averaged out. Time synchronous averaging removes background noise and non-synchronous
events (random transients). It is useful where multiple shafts and speeds are being monitored in
close proximity (such as in a paper mill). A reference signal (tachometer input) in needed.

It is important that the speed of the machine being monitored be kept steady during the sampling
procedure. A speed fluctuation as small as 0.1% can cause jitter. That is the variability in the
time samples become too great and the averaging begins to remove some of the repetitive signal
as well as the random parts. A remedy for this is to have the sampling frequency regulated by the
rotating speed. Averaging of frequency spectra is also common. That is all the frequency spectra
are calculated from non-averaged raw vibration signals and then the spectra are averaged
together. This process is more convenient as it does not require collecting vibration data that is
linked to a trigger. However, the results are not as satisfactory. Figure 8.24 shows samples of
both types of averaging. Note that averaging in the time domain is more effective at removing
the background noise.

(a) (b)

Figure 8.24 Frequency spectra (a) averaged in the time domain and (b) averaged in the frequency
domain.

162
Figures 8.25 to 8.29 shows samples of various signals together with random noise. The effects of
the different types of averaging are seen in these figures.

Figure 8.25 Sample sinusoid signal (top), random noise (middle) and the same sinusoid signal
mixed with the random noise (bottom).

8 Averages

64 Averages

256 Averages

Figure 8.26 Sample sinusoid signal with noise, 8 averages (top), 64 averages (middle) and 256
averages (bottom).

163
Raw Signal (a)

16 Averages (b)

Spectra of Averaged (c)


Signal

16 Spectral Averages (d)


of Raw Signals

Figure 8.27 Strong sinusoid signal with noise, raw signal (a), 16 averages (b), spectra of
averaged signal (c) and 16 spectral averages of raw signal.

Raw Signal (a)

16 Averages (b)

Spectra of Averaged (c)


Signal

16 Spectral Averages (d)


of Raw Signals

Figure 8.28 Weak sinusoid signal with noise, raw signal (a), 16 averages (b), spectra of averaged
signal (c) and 16 spectral averages of raw signal.

164
Raw Signal (a)

128 Averages (b)

Spectra of Averaged (c)


Signal

Figure 8.29 Square wave with noise, raw signal (a), 128 averages (b), spectra of averaged signal
(c).

It will be noted here that the FFT and the inverse FFT can be used as effective filters. After the
FFT has been calculated the unwanted frequency components in the spectrum can be set to zero
and then this new truncated spectrum can be use to regenerate the original signal less the noise
that was in the original. Figures 8.30 to 8.32 show this process.

Raw Signal (a)

FFT (b)

Unwanted Components (c)


Set to Zero

Filtered Signal (d)

Figure 8.30 Sinusoids with noise, raw signal (a), FFT results (b), unwanted frequency
components set to zero (c) and reconstituted filtered signal (d).

165
Raw Signal (a)

FFT (b)

Unwanted Components (c)


Set to Zero

Filtered Signal (d)

Figure 8.31 Sinusoid with noise, raw signal (a), FFT results (b), unwanted frequency components
set to zero (c) and reconstituted filtered signal (d).

Raw Signal (a)

FFT (b)

Unwanted Components (c)


Set to Zero

Filtered Signal (d)

Figure 8.32 Square wave with noise, raw signal (a), FFT results (b), unwanted frequency
components set to zero (c) and reconstituted filtered signal (d).

8.8.1 Negative Averaging

Negative averaging is when a baseline signal is recorded and then subsequently removed from
all the signals before calculation of the FFT. This process reveals changes and transients only

166
while removing the repetitive parts of the signal. This process is useful when analyzing signal
collected from isolated equipment or components.

Figure 8.33 Sinusoid with noise, averaged (left), negative averaged (right).

8.9 Orbits

Vertical motion of the shaft centre will cause vertical movement of the oscilloscope marker.
Horizontal motion of the shaft centre will cause the marker to move horizontally. Combining
these outputs creates a plot known as an orbit. Figure 8.34 shows an elliptical orbit shape which
is the result of more clearance in the one direction than in the orthogonal direction. While an
ideal orbit shape would be circular, a non-circular orbit may not always indicate a problem.
Machines with a difference in stiffness in the vertical and horizontal direction will present a
naturally elliptical orbit shape.

Y
Y Direction Displacement

Vibration
Direction X

X Direction Displacement

Figure 8.34 An elliptical orbit suggesting more clearance in the one direction than in the
orthogonal direction.

Figure 8.35 shows a figure 8 orbit plot suggesting a shaft that is vibrating twice as fast in one
direction than the other. Figure 8.36 shows a set of orbits that indicate the influence of radial
shaft preloads on the motion of the shaft. Table 8.4 shows a variety of common orbit patterns.

167
Y Direction Displacement
X Direction Displacement

Figure 8.35 A figure 8 orbit plot suggesting a shaft that is vibrating twice as fast in one
direction than the other.

Figure 8.36 Orbits showing the influence of radial shaft preloads.

Table 8.4 Common orbit patterns.

168
Table 8.4 Common orbit patterns (Contd).

Tracing the centre of rotation of a shaft in a journal bearing can also be used to show the
progression of the vibration amplitude and direction during start up and/or changing operational
conditions (see Figure 8.37). Note that if the shaft centre-line moves above the split line of the
bearing the condition is considered as unstable because of the forces that will be exerted on the
top half of the bearing housing.

Figure 8.37 Centre of rotation of a shaft in a journal bearing during start up.

169
8.10 Enveloped (Demodulated) Spectra

Enveloped spectra (also referred to as demodulated spectra) are generated when the time
waveform is demodulated (high pass filtered, rectified, and then low pass filtered). This process
removes noise and transient signal components. The effect of impulsive components (in most
cases due to early stage bearing deterioration) is highlighted. This is a preferred method for the
detection of bearing damage in complex machinery. It needs to be noted that while this is a
useful detection method, because the impulsive nature in faulty bearing vibration signal
disappears as the fault worsens, it is not a good long term trending tool. Figure 8.38 shows the
steps involved in vibration signal enveloping.

Figure 8.38 Typical steps in vibration signal enveloping.

Bearing defect impacts cause structural natural frequencies to be excited (structural ringing).
Figure 8.39 shows a typical accelerometer vibration waveform with imbedded defect signal.

Figure 8.39 A typical accelerometer vibration waveform with imbedded defect signal.

The band-pass filtered vibration waveform then shows the defect modulation of the machine
structural resonance (low frequency modulation removed). Figure 8.40 shows a full-wave
vibration signal. Figure 8.41 shows a rectified and full wave detected (enveloped) vibration
signal produced by a peak detector. Figure 8.42 shows the FFT analysis result of an enveloped

170
vibration signal. The amplitude modulation is evident by the sidebands around the ball pass
frequency and its harmonics.

Figure 8.40 A full-wave rectified vibration signal.

Figure 8.41 Enveloped vibration signal produced by a peak detector.

Figure 8.42 FFT analysis result of an enveloped vibration signal. The amplitude modulation is
evident by the sidebands around the ball pass frequency and its harmonics.

Figure 8.43 shows the FFT analysis result of several different enveloped vibration signals
representing different fault conditions.

171
Figure 8.43 FFT analysis results of several different enveloped vibration signals representing
different fault conditions.

While the methodology for generating an enveloping spectrum may seem straightforward, valid
results depend on careful application. The following issues should be carefully considered before
applying enveloping techniques to machinery monitoring.

Early Detection: Enveloping provides early detection of faults that would otherwise be obscured
by larger components of the machine vibration signature. When a fault is identified in an
enveloped spectrum, failure is not necessarily imminent. Monitoring should continue or be
increased to trend progression of the fault. Defects will not be detectable until they have

172
progressed to the point that their interaction with other components is repetitive, not random.
Check against other data using other measurement techniques, as available.

Eligible Machines: Enveloping techniques can be used to detect faults in machine components
with repetitive metal-to-metal interaction. However, because enveloping is not a direct
measurement, many extraneous factors can add to or diminish the enveloped signal. Several
machine components or characteristics can prevent successful enveloping implementation.
Joints, interfaces, gaskets, and fluid-film or squeeze-film dampers prevent high-frequency signal
transmission critical for enveloping. High-frequency operational noise may overshadow signals
of interest in reciprocating machines, variable frequency drive motors, and others.
Electromagnetic interference may also introduce itself into the cabling between the transducer
and the signal processing device and compromise signal integrity.

Transducer Selection: The frequency response of the transducer must include the expected range
of machine resonance frequencies (ranging from 1 kHz to more than 40 kHz). The resonance
frequency of the mounted transducer must be sufficiently far away from the machine frequencies
of interest to avoid interactions. The transducer should be extremely reliable to ensure trending
consistency.

Transducer Mounting: Enveloping measurements are highly dependent on transducer mounting


method and location. Even a slight change in mounting location can yield quite different results.
A solid, repeatable mounting is essential to help ensure that changes observed in data are due to
changes in machine condition not variations introduced due to changes in collecting the data. A
flat, clean (bare metal) surface for mounting the transducer is critical. Handheld transducer
applications can be especially susceptible to variation based on changes in applied pressure,
mounting angle, and other variables introduced by the person taking the reading. When practical,
consider affixing the transducer in a way that reduces variability (stud mounting).

Since the high frequency signal on which enveloping depends does not usually travel very far
within a machine, the transducer mounting should have a short transmission path from the
machinery component of interest, with as little damping of the high frequency energy as
possible. Any metal interface or discontinuity in the machine causes significant signal
attenuation, and fluid films at any interface can completely stop transmission of a signal.
Consequently, faults which are detected with enveloping can be expected to be located near the
measurement transducer.

Fault Identification: Because of the correlation between fundamental spectrum frequencies and
fault sources, flawed components can often be identified before the bearing is removed and
physically examined, allowing spare parts to be ordered in advance and work procedures to be
written with the knowledge of precisely what needs to be changed. Frequencies associated with
specific machine components and natural resonances must both be considered when configuring
the enveloping technique to ensure valid data.

Since improper lubrication, caused by inadequate, excessive, or contaminated lubricants, can


cause frequency components to appear in the enveloping spectrum, lubrication should be
checked first when faults appear. Progression of faults is often indicated by the presence of more

173
bearing frequency components and an overall increase in the noise floor of the enveloping
spectrum. Overall, the fundamental fault frequencies in the spectrum are the most important for
correlation with physical defects.

Severity Prediction: Enveloping provides valuable information for machinery management.


However, enveloping by itself does not give all the information necessary for reliable and
accurate prediction of the condition of a machine component. In the enveloping spectrum,
frequency can be correlated to a specific machine component, but increasing magnitude is not
necessarily correlated to the progression of the fault. In fact, a well-known phenomenon is that
the acceleration enveloping amplitude may actually decrease as bearing failure becomes more
imminent. As a bearing continues to wear, its small, vibration inducing flaws begin to smooth
out, and the characteristic ringing caused by the flaws (and detected by enveloping) decreases.

Trending Consistency: Data must be collected periodically and consistently to ensure the
integrity of the trend. As noted above, this includes using the same transducer, in the same
location mounted, the same way, to reduce gross and systematic errors. Then, the trend can be
evaluated to see the progression of any defects. Permanently mounted transducers are
recommended.

Frequency Variation: The absolute frequency of enveloping signals is directly dependent on shaft
rotation speed. In order for frequencies to be correlated with potential faults, machine speed must
be known and relatively constant. Otherwise, the magnitude of frequency components may be
affected by frequency dependent machinery and instrumentation responses rather than changes in
defect severity.

8.11 Modal Analysis (revisited)

Modal domain analysis is the investigation of natural frequencies and the deflection shapes at
these frequencies. Natural frequencies can be excited by external forces (wind, impacts) or self
excited (called Operational Deflection Shapes). An instrumented impulse (also called impact)
hammer is a standard way to excite a free vibration response in a machine or structure. The
response will tend to vibrate at one or more of the machine or structure natural frequencies.
Depending on the tip hardness and the size (mass) of the impact hammer, different frequencies
will be preferentially excited. Listed below are a set of typical operational specifications for an
impulse hammer.

Modally Tuned Impulse Hammer with force sensor and tips, 0 to 100 lbs, 50 mV/lb
Sensitivity: 50 mV/lbf (11.2mV/N)
Measurement Range: 100 lbf pk (440N pk)
Frequency Range: 8000 Hz
Frequency Range: 2500 Hz
Frequency Range: 750 Hz
Hammer Mass: 0.34 lb (0.16kg)

In theory an impact at any location will excite all the natural frequencies. However, in real
situations the location of the impact site and the location of the sensors measuring the response

174
are important. Figure 8.45 shows a bearing casing and impact hammer and measurement sensor.
Typical applications where modal analysis is done involve repeated impacts to excite the
structure and one or more sensors place at various locations covering the extent of the structure
under consideration. Figure 8.46 shows an impact hammer with the components labeled and
typical input and output signals.

Figure 8.44 Impact hammer and various tips of different hardness. Impact hammer tips of
different hardness are used to excite different frequency ranges.

Figure 8.45 Bearing casing and impact hammer and measurement sensor.

Figure 8.46 Impact hammer components and typical input and output signals.

Not every impact is ideal for taking a vibration signal measurement. Figure 8.47 shows a variety
of input signals (one good and the rest faulty and not useable as modal analysis inputs). Just as
not every input is usable, even with an ideal impact excitation, sometimes the output is
defective in some way and not suitable for use in modal analysis. Figure 8.48 shows a variety of
output signals (one good and the rest faulty and not useable for modal analysis).

175
Figure 8.47 A variety of input signals (one good and the rest faulty and not useable as modal
analysis inputs).

Figure 8.48 A variety of output signals (one good and the rest faulty and not useable for modal
analysis).

176
Mode shapes are the shapes assumed by a system as it vibrates at a natural frequency. A mode
shape does not provide information on system absolute motions, only relative motions. Absolute
motions are determined by damping and vibration forces.

Figure 8.49 Tacoma narrows bridge disaster. Bridge natural frequency (resonance) was excited
by wind gusts and eventually destroyed the bridge.

Figure 8.50 Car body modal analysis results.

Figure 8.51 Cylinder vibration modes FEA results (free and fixed end supports).

177
Figure 8.52 Cylinder vibration modes test results (free and fixed end supports).

Figure 8.53 Rotor vibration modes (soft and hard supports).

Chapter Summary

Vibration meters are generally small, hand-held (portable), inexpensive, simple to use,
selfcontained devices that give an overall vibration level reading.
Most vibration data collectors available today for use in machine condition monitoring and
fault diagnostics are microcomputer based. They are used together with vibration sensors to
measures vibration, to store and transfer data, and for frequency domain analysis.

178
Frequency domain analyzers are specialized instruments that emphasize the analysis of
vibration signals, and as such they are perhaps the key instrument for diagnostic work.
Time domain instruments are generally only able to provide a time domain display of the
vibration waveform.
Tracking analyzers are typically used to record and analyze data (locate machine resonances
and unbalance conditions) from machines that are changing speed. This usually occurs
during run-up and coast-down of large machinery or turbo-machinery.
Most digital data recording is currently done on solid state media.
Vibration signal sampling can take place with uniform time spacing between samples or with
uniform position (degrees of shaft rotation) spacing between samples.
The faster the sampling rate the higher the resolution possible, but the more raw data there is
to deal with.
Aliasing becomes a problem if the sampling rate is too low for the given situation or if there
is no anti-aliasing filtering.
Time synchronous averaging helps remove noise and random components from the vibration
signal, but requires a tachometer input signal as well.
Orbit display show the movement of a shaft inside a journal bearing and are useful fault
detection and diagnostic tools
Enveloped (demodulated) frequency spectra is a useful method of detecting the earliest
stages of rolling element bearing failure.
Modal analysis can reveal machine and structural natural frequencies and the relative motion
of the machine or structure during those vibration responses.

179
Chapter 9

Fault Detection
In many discussions of machine condition monitoring and fault diagnostics, the distinction
between fault detection and fault diagnosis is not made. Here, they have been divided into
separate sections in order to highlight the differences and clarify why they should be treated as
separate tasks. Fault detection can be defined as the departure of a measurement parameter from
a range that is known to represent normal operation. Such a departure then signals the existence
of a faulty condition. Given that measurement parameters are being recorded, what is needed for
fault detection is a definition of an acceptable range for the measurement parameters to fall
within. There are two methods for setting suitable ranges: (1) comparison of recorded signals to
known standards and (2) comparison of the recorded signals to acceptance limits.

9.1 General Standards and Guidelines

Standards are documented agreements containing technical specifications or other precise criteria
to be used consistently as rules, guidelines, or definitions of characteristics, to ensure that
materials, products, processes and services are fit for their purpose. A good standard represents
consensus of opinion, is easy to understand and use, and contains no ambiguities or loopholes.
Standards are intended to set criteria for rating or classifying the performance of equipment or
material and to provide a basis for comparison of the maintenance qualities of pieces of
equipment of the same type. In addition standards are intended to test equipment whose
continuous operation is necessary for industrial or public safety, to provide a basis for the
selection of equipment or material, and to setup a procedure for the calibration of equipment.

One of the best known sources of standards is the International Organization for Standardization
(ISO). These standards are technology oriented and are set by teams of international experts. ISO
Technical Committee 108 (Mechanical Vibration and Shock), Sub-Committee 5 is responsible
for standards for condition monitoring and diagnostics of machines. This group is further divided
into a number of working groups who review data and draft preliminary standards. Each working
group has a particular focus such as terminology, data interpretation, performance monitoring, or
tribology-based machine condition monitoring.

The detailed scope of TC 108, SC 5 includes terminology, excitation, vibration control, human
exposure, measurement and calibration, test methods, and condition monitoring and diagnostics.
Working groups within SC 5 include those that are focused on terminology, data interpretation
and diagnostic techniques, performance monitoring and diagnostics, tribology, prognostics,
formats and methods for presenting data, training and certification, monitoring and diagnostics of
machines, monitoring and diagnostics of electrical equipment, and thermal imaging.

ISO Standards are developed according to the principles of consensus amongst participants and
industry-wide voluntary participation. While ISO is perhaps the most widely known
standardization organization, there are several others that are focused on specific industries.
Examples of these include the International Electrical Commission, which is primarily product
oriented, and the American National Standards Institute (ANSI), which is a non-government

180
agency. There are also different domestic government agencies that vary from country to
country. National defense departments also tend to set their own standards. Figure 9.1 shows an
overview of ISO standards that are in existence and those that are in progress within TC 108
SC5.

Figure 9.1 An overview of ISO standards that are in existence and those that are in progress
within TC 108 SC5

Table 9.1 ISO standards related to machine condition monitoring and fault diagnositics.

ISO Reference Title


18431-1 Mechanical Vibration and Shock Signal Processing Part 1:
General Introduction
18434-1 Condition Monitoring and Diagnostics of Machines Thermal
Imaging
18436-1 Condition Monitoring and Diagnostics of Machines Requirements
for Training and Certification of Personnel Part 1: Requirements
for Certifying Bodies and the Certification Process
18436-2 Condition Monitoring and Diagnostics of Machines Part 2:
Vibration Condition Monitoring and Diagnosis
18436-3 Condition Monitoring and Diagnostics of Machines Accreditation
of Organisation and Training Specialists - Part 3: Accreditation of
Certification Bodies
18436-4 Condition Monitoring and Diagnostics of Machines Part 4:
Lubrication Management and Analysis

181
Table 9.1 ISO standards related to machine condition monitoring and fault diagnositics (Cont`d).

ISO Reference Title


18436-5 Condition Monitoring and Diagnostics of Machines Part 5:
Thermography
18436-6 Condition Monitoring and Diagnostics of Machines Part 6:
Diagnostics and Prognostics
18436-7 Condition Monitoring and Diagnostics of Machines Part 7:
Condition Monitoring Specialists
14830-1 Condition Monitoring and Diagnostics of Machines Tribology
Based Monitoring of Machines Part 1: General Guidelines
14830-2 Condition Monitoring and Diagnostics of Machines Tribology
Based Monitoring of Machines Part 2: Lubricant Sampling
13372 Condition Monitoring and Diagnostics of Machines Vocabulary
13372-1 Condition Monitoring and Diagnostics of Machines Vibration
Condition Monitoring: General Procedures
13374-1 Condition Monitoring and Diagnostics of Machines Data
Processing, Communication and Presentation Part 1: General
Guidelines
13374-2 Condition Monitoring and Diagnostics of Machines Data
Processing, Communication and Presentation Part 2: General
Data Processing and Analysis Procedures
17359 Condition Monitoring and Diagnostics of Machines General
Guidelines
22349 Condition Monitoring and Diagnostics of Machines Condition
Based Maintenance Optimization Part 1: General Guidelines
13379 Condition Monitoring and Diagnostics of Machines Data
Interpretation and Diagnostic Techniques General Guidlines

9.1.1 Standards Based on Vibration Severity

It is an oversimplification to say that vibration levels must always be kept low. Standards depend
on many things, including the speed of the machinery, the type and size of the machine, the
service (load) expected, the mounting system, and the effect of machinery vibration on the
surrounding environment. Standards that are based on vibration severity can be divided into two
basic categories:

1. Small-to-medium sized machines: These machines usually operate with shaft speeds of
between 600 and 12,000 rpm. The highest broadband RMS value usually occurs in the
frequency range of 10 to 1000 Hz.

182
2. Large machines: These machines usually operate with shaft speeds of 600 to 1200 rpm. If the
machine is rigidly supported, the machines fundamental resonant frequency will be above
the main excitation frequency. If the machine is mounted on a flexible support, the machines
fundamental resonant frequency will be below the main excitation frequency.

While general standards do exist, there are also a large number of standards that have been
developed for specific machines. Figure 9.2 shows a table with generic acceptance limits based
on vibration severity and class of machine based on size.

Figure 9.2 Acceptance limits based on vibration severity levels (zone A - new machine; zone B -
acceptable; zone C - monitor closely; zone D - damage occurring).

The classes in Figure 9.2 are defined as follows.


Class I - dividual components, integrally connected with complete machine (electric motors up
to 15 kiloWatts)
Class II - Medium sized machines (15 - 75 kiloWatt electric motors)
Class III - Large prime movers on heavy, rigid foundations
Class IV - Large prime movers on relatively soft, light-weight foundations

Figure 9.3 shows a table with generic acceptance limits based on vibration severity and the type
of rotor supports in use. Figure 9.4 shows a vibration severity grade monograph that can be used
to relate displacement, velocity and acceleration.

183
Figure 9.3 Acceptance limits based on vibration severity levels and the type of rotor supports in
use.

Figure 9.4 Vibration severity grade monograph that can be used to relate displacement, velocity
and acceleration.

184
Figure 9.5 Rathbone chart showing unbalance severity.

185
Figure 9.6 Blake Chart chart showing general vibration severity levels.

186
.

Figure 9.7 Dresser-Clark-Jackson chart for proximity shaft vibration measurements on certain
centrifugal compressors.

9.1.2 Standards Based on Machinery Type

Because different machines that are designed to perform approximately the same task tend to
behave in a similar manner, it is not surprising that many standards are set based on machinery
type. Figure 9.8 and Figure 9.9 show two sets of ISO standards that deal with different types of
machines. Figure 9.8 shows ISO standard 7919 series which deals with measurements made on
rotating shafts. Figure 9.9 shows standard 10816 series which deals with measurements made on
non-rotating parts (typically bearing casings). These two sets of standards represent two different
types of machine. Standard 7919 refers to machines with shafts that are supported on relatively
soft journal bearings (where the vibration is not transmitted through the bearing and needs to be
measured directly from the shaft). Standard 10816 refers to machines that are supported on
relatively stiff journal bearings or rolling element bearings (where the vibration is transmitted
through the bearing and can be measured from the bearing casing).

187
Mechanical vibration of non-reciprocating machines
ISO 7919 Series
- Measurement on rotating shafts and evaluation criteria

7919-1:1996 Part 1: General Guidelines

Part 2: Land-based steam turbines and generators in excess of 50 MW with normal operating
7919-2: 2001
speeds of 1500 r/min, 1800 r/min, 3000 r/min and 3600 r/min

7919-3: 1996 Part 3: Coupled industrial machines

7919-4: 1996 Part 4: Gas turbine sets

7919-5: 1997 Part 5: Machines set in hydraulic power generating and pumping plants

Figure 9.8 ISO standard 7919 series which deals with measurements made on rotating shafts.

Mechanical vibration
ISO 10816 Series
- Evaluation of machine vibration by measurements on non-rotating parts

10816-1: 1995 Part 1: General Guidelines

Part 2: Land-based steam turbines and generators in excess of 50 MW with normal operating
10816-2: 2001
speeds of 1500 r/min, 1800 r/min, 3000 r/min and 3600 r/min
Part 3: Industrial machines with normal power above 15kW and nominal speeds between 120
10816-3: 1998
r/min and 15000 r/min when measured in situ

10816-4: 1998 Part 4: Gas turbine sets excluding aircraft derivatives

10816-5: 2000 Part 5: Machines set in hydraulic power generating and pumping plants

10816-6: 1995 Part 6: Reciprocating machines with power ratings above 100 kW

10816-7 Part 7: Rotodynamic pumps for industrial application

Figure 9.9 ISO standard 10816 series which deals with measurements made on non-rotating parts
(typically bearing casings).

Figure 9.10 shows a decision chart for selecting between ISO standard 10816 and 7919. Figure
9.11 shows a monograph for selecting between ISO standard 10816 and 7919. Both of these
figures use a combination of the shaft relative displacement from the centre-line (xf = a function
of the bearing length (l) and the shaft length (L)) and a ratio () of the bearing and pedestal
stiffnesses (Z1 and Z2) to select the appropriate standard to apply.

Figure 9.12 shows a chart that uses dynamic stiffness ratio as a means of selecting between ISO
standard 10816 and 7919 for different machine types. Figure 9.13 shows a graphic figure which
complements the chart in Figure 9.12, showing the dynamic stiffness ratio as a means of
selecting between ISO standard 10816 and 7919 for different machine types.

In general machines can be divided into four primary categories. There are listed and described
below.

1. Reciprocating machinery: These machines may contain both rotating and reciprocating
components (e.g., engines, compressors, pumps).

188
2. Rotating machinery (rigid rotors): These machines have rotors that are supported on rolling
element bearings (usually). The vibration signal can be measured from the bearing housing
because the vibration signal is transmitted well through the bearings to the housing (e.g.,
electric motors, single-stage pumps, slow-speed pumps).

3. Rotating machinery (flexible rotors): These machines have rotors that are supported on
journal (fluid film) bearings. The movement of the rotor must be measured using proximity
probes (e.g., large steam turbines, multistage pumps, compressors). These machines are
subject to critical speeds (high vibration levels when the speed of rotation excites a natural
frequency). Different modes of vibration may occur at different speeds.

4. Rotating machinery (quasi-rigid rotors): These are usually specialty machines in which some
vibration gets through the bearings, but it is not always trustworthy data (e.g., low-pressure
steam turbines, axial flow compressors, fans).

Figure 9.10 Decision chart for selecting between ISO standard 10816 and 7919.

189
Figure 9.11 Monograph for selecting between ISO standard 10816 and 7919.

ISO 10816 ISO 7919


Machine Dynamic Stiffness Ratio,
(pedestal) (shaft)
High Pressure Turbine 5 Moderate Good
Low Pressure Turbine 1.5 Moderate Good
Large Generator 1.5 Moderate Good
High Pressure Centrifugal Compressor 5 Not Good Good
Large Fan 2/3 Good Moderate
Small Fan & Pump 1/3 Good Moderate
Vertical Pump 1/10 Good Not Good
Large Steam Turbine Generator Set 1.5 to 3 Moderate Good

Figure 9.12 Chart using dynamic stiffness ratio as a means of selecting between ISO standard
10816 and 7919 for different machine types.

190
Figure 9.13 Decision chart using dynamic stiffness ratio as a means of selecting between ISO
standard 10816 and 7919 for different machine types.

Figure 9.14 Monograph for selecting between ISO standard 10816 and 7919 also showing some
common machine types.

191
Electric motors fall under ISO standard 2373. Acceptable vibration levels a motor size
dependent. All measurements should be made at no load. This standard applies to three phase
AC and DC motors with shaft height (the vertical distance from the base of the motor to the
centerline of the shaft) between 80 and 400 mm. The criterion for vibration severity is given in
terms of the RMS value of velocity amplitude in the frequency range from 10 to 1000 Hz when
measured with instrumentation which meets the requirements of ISO 2954. These measurements
are to be made on the machine installed on a free suspension. The motor should be operated at
rated voltage and nominal frequency (for AC motors) and at its nominal speed (for machines
with several speeds). Measurements of vibration should be carried out under no load operation at
the temperature reached by the motor after a sufficient period of no load operation

Figure 9.15 Recommended limits of vibration severity for electric motors (ISO standard 2373).

When measuring vibration severity levels on pumps they should be operating in a non-cavitating
mode. The suction piping must be arranged so as provide a straight uniform flow to the pump.
Piping must be connected in such a way so as to avoid undue strain on the pump. Shaft coupling
must be aligned to within the manufacturers recommendations.

Figure 9.16 Monograph relating peak-to-peak displacement, peak velocity and peak acceleration
(Metric units) at different frequencies and showing the limits of acceptable levels.

192
Figure 9.17 Monograph relating peak-to-peak displacement, peak velocity and peak acceleration
(Imperial units) at different frequencies.

Steam turbine generator sets have vibration severity limits set based on absolute shaft
displacement (flowing ISO standard 7919 (part 2)).

Figure 9.18 Vibration classification guide values for large steam turbines (ISO standard 7919)
(part 2)). A - good, B Acceptable, C - Monitor closely, D Unacceptable.

Reciprocating machinery vibration measurements are type and size dependant, load and
mounting dependant, should be done at greater than 3,000 RPM (frequency 2 - 300 Hz).

193
Figure 9.19 Vibration classification numbers and guide values for reciprocating machines (ISO
standard 10816-6).

Industrial turbo machinery (high speed) vibration severity limits require shaft displacement
measurement relative to bearings (ISO standard 7919 (part 3)).

Figure 9.20 Vibration classification guide values for industrial turbo machinery (ISO standard
7919 (part 3)). A - good, B Acceptable, C - Monitor closely, D Unacceptable.

194
Centrifugal compressors require in-service vibration severity criteria to be measured as a
function of shaft speed (Compressed Air and Gas Institute standard).

Figure 9.21 Vibration classification guide values for centrifugal compressors (Compressed Air
and Gas Institute standard).

Gear shaft vibration (displacement amplitude) versus frequency is still an ISO draft standard.

Figure 9.22 Vibration classification guide values for gear shafts (ISO draft standard).

It is important to note that the early discovery of faulty conditions is a key to optimizing the
maintenance effort by allowing the longest possible lead-time for decision making. As well as
the overall vibration levels being monitored, the rates of change are also important. The rate of
change of a vibration level will often provide a strong indication of the expected time until
absolute limits are exceeded. In general, relatively high but stable vibration levels are of less
concern than relatively low but rapidly increasing levels.

An example of how acceptance limits may be used to detect faults and trend condition is
provided when the gradual deterioration of rolling-element bearings is considered. Rolling-

195
element bearings generate distinctive defect characteristic frequencies in the frequency spectrum
during a slow, progressive failure. Vibration levels can be monitored to achieve maximum useful
life and failure avoidance. Typically, the vibration levels increase as a fault is initiated in the
early stages of deterioration, but then decrease in the later stages as the deterioration becomes
more advanced. Appropriately, set acceptance levels will detect the early onset of the fault and
allow subsequent monitoring to take place even after the overall vibration level has dropped.
However, rapid bearing deterioration may still occur due to a sudden loss of lubrication,
lubrication contamination, or a sudden overload. The possibility of these situations emphasizes
the need for carefully selected acceptance limits.

It should also be noted that changes in operating conditions, such as speed or load changes, could
invalidate time trends. Comparisons must take this into consideration.

9.1.3 Standards based on Statistical Limits

Statistical acceptance limits are set using statistical information calculated from the vibration
signals measured from the equipment that the limits will ultimately be used with. As many
vibration signals as possible are recorded, and the average of the overall vibration level is
calculated. An alert or warning level can then be set at 2.5 standard deviations above or below
the average reading. This level has been found to provide optimum sensitivity to small changes
in machine condition and maximum immunity to false alarms. A distinct advantage to using this
method to set alarm levels is the fact that the settings are based on actual conditions being
experienced by the machine that is being monitored. This process accommodates normal
variations that exist between machines and takes into account the initial condition of the
machine.

9.2 Acceptance Limits

Standards developed by dedicated organizations are a useful starting point for judging machine
condition. They give a good indication of the current condition of a machine and whether or not
a fault exists. However, judging the overall condition of machinery is often more involved.
Recognizing the changing machinery condition requires the trending of condition indicators over
time. The development and use of acceptance limits that are close to the normal operating values
for specific machinery will detect even slight changes in condition. While these acceptance limits
must be tight enough to allow even small changes in condition to be detected, they must also
tolerate normal operating variations without generating false alarms. There are two types of
limits:

1. Absolute limits represent conditions that could result in catastrophic failure. These limits are
usually physical constraints such as the allowable movement of a rotating part before contact
is made with stationary parts.

2. Change limits are essentially warning levels that provide warning well in advance of the
absolute limit. These vibration limits are set based on standards and experience with a
particular class of machinery or a particular machine. Change limits are usually based on
overall vibration levels.

196
It is important to note that the key to failure prevention is early discovery of deterioration. The
rates of change are also important. The expected time until limits are exceeded is also useful
information. In general high but stable vibration levels are of less concern than low but rapidly
increasing levels. A small percentage change at a high vibration level is more significant than a
large percentage change at a low level. It should also be noted that changes in operating
conditions can invalidate time trends, speed or load changes may alter trends, and comparisons
must take this into consideration.

When judging overall condition it is important to recognize changing machinery condition (trend
over time), develop and use acceptance limits, measure vibration data while the machine under
consideration is operating close to normal operating conditions to detect changes in condition. In
this way the alarm limits set will tolerate normal operating variations without false alarms

9.3 ISO Standard on Training and Certification in Machine Condition Monitoring

Training and certification within the ISO set of standards is covered under ISO standard 18436.
This standard encompasses the following parts, under the general title of Condition monitoring
and diagnostics of machines Requirements for training and certification of personnel. Parts 3
7 are still under preparation. Tables 9.2 to 9.5 show some of the training requirements for the
different categories (levels) of certification.

Part 1: Requirements for certifying bodies and the certification process.


Part 2: Vibration condition monitoring and diagnostics.
Part 3: Requirements for training bodies.
Part 4: Lubrication management and analysis.
Part 5: Thermography.
Part 6: Diagnostics and prognostics.
Part 7: Condition monitoring specialists.

Table 9.3 Recommended minimum duration of cumulative training (hours).

Table 9.3 Recommended minimum duration of cumulative experience (months).

197
Table 9.4 Examples of certification examination details.

Table 9.5 Overview of training hours needed per subject per category.

Other related ISO standards include:

ISO 6954 (1984) Mechanical vibration and shock Guidelines for the overall evaluation of
vibration in merchant ships.

ISO 8528/9 (1995) Reciprocating internal combustion engine driven alternating current
generating sets Part 9: Measurement and evaluation of mechanical vibrations.

ISO 1940/1 (2002) Mechanical vibration Balance quality requirements of rigid rotors
Part 1: Specification and verification of balance tolerances.

198
9.4 Frequency-Domain Limits

Judging vibration characteristics within the frequency spectra is sometimes a more accurate
method of detecting and trending fault conditions. It can also provide earlier detection of specific
faults because, as mentioned previously, the frequency domain is generally more sensitive to
changes in the vibration signal that result from changes in machine condition. Spectral
components are directly linked to forcing functions often providing for more accurate for
trending and diagnostics and early detection of specific faults. The different specific methods are
listed and described below.

9.4.1 Limited Band Monitoring

In limited band monitoring, the frequency spectrum is divided into frequency bands. The total
energy or highest amplitude frequency is then trended within each band. Each band has its own
limits based on experience. Generally, ten or fewer bands are used. Small changes in component-
specific frequency ranges are more clearly shown using this strategy. Bandwidths and limits
must be specific to the machine, sensor type, and location.

Narrowband monitoring is the same as limited band monitoring, except it has finer definition of
the bands. Figures 9.23, 9.24, and 9.25 show limited band frequency monitoring schemes
(amplitude limits in band limited frequency ranges) for monitoring rolling element bearings,
Fluid film bearings, and gearboxes respectively.

Figure 9.23 Limited band frequency monitoring schemes (amplitude limits in band limited
frequency ranges) for monitoring rolling element bearings.

199
Figure 9.24 Limited band frequency monitoring schemes (amplitude limits in band limited
frequency ranges) for monitoring fluid film bearings.

Figure 9.25 Limited band frequency monitoring schemes (amplitude limits in band limited
frequency ranges) for monitoring gearboxes.

9.4.2 Constant Bandwidth Limits

When limited band monitoring is practiced and the bands have same width at high and low
frequencies, the procedure is called constant bandwidth monitoring. This technique is useful for
constant speed machines where the frequency peaks in the spectra remain relatively fixed.

Figure 9.26 shows a set of constant bandwidth acceptance limits with calculated alert and danger
alarm levels for each of the different frequency spans.

200
Figure 9.26 Constant bandwidth acceptance limits showing the calculated alert and danger alarm
levels.

9.4.3 Constant Percentage Bandwidth Limits

Constant percentage bandwidth monitoring involves using bandwidths that remain a constant
percentage of the frequency being monitored. This results in the higher frequency bands being
proportionally wider than the lower frequency bands. This allows for small variations in speed
without the frequency peaks moving between bands, which may have different acceptance limits.

Figure 9.27 Constant percentage bandwidth acceptance limits showing the calculated alert and
danger alarm levels.

Regardless of the type of spectral monitoring that is practiced, it is important to establish an


appropriate reference spectrum that represents the machine operation in an as new condition.

201
The spectra from one good machine may be selected to represent the best condition for a
population of like machines. A composite reference put together using vibration signals from all
machines averaged together is another way to establish a good reference. Each individual
machine may also have its own reference spectrum (statistically derived if enough data is
available). All measurement samples must represent the machine in good condition and the
samples must be taken under normal operating conditions.

When setting the minimum threshold values for the trends that are expected there are some
important considerations to keep in mind. Trends and levels are important. Low levels will
typically have a wide percentage variation than high levels. A minimum level that is well below
the average will allow normal variation without triggering a false alarm. This requires knowledge
of the machine and operating conditions.

Chapter Summary

Fault detection can be defined as the departure of a measurement parameter from a range that
represents normal operation. Such a departure signals the existence of a faulty condition.
ISO Technical Committee 108, Sub-Committee 5 is responsible for standards for Condition
Monitoring and Diagnostics of Machines. Standards are based on machinery type or
vibration severity.
The development and use of acceptance limits that are close to normal operating values for
specific machinery will detect even slight changes in condition.
Statistical acceptance limits are set using statistical information calculated from the vibration
signals measured from the equipment that the limits will ultimately be used with.
Judging vibration characteristics within the frequency spectra is sometimes a more accurate
method of the early detecting and trending of fault conditions because the frequency domain
is generally more sensitive to changes in the vibration signal that result from changes in
machine condition.
Frequency domain limits include limited band monitoring, constant bandwidth limits, and
constant percentage bandwidth limits.
In general, the severity of machine condition is assessed by using the amplitude of vibration
As a result of variation in design and low signal strength, rolling element bearings and gears
require the evaluation of amplitudes and frequencies
Principal measures for bearing cap (casing) vibration are peak or RMS velocity and
acceleration
All bearing cap measures should be stated in either RMS or peak and should not be mixed
Shaft vibration severity is evaluated using relative displacement peak to peak, bearing
clearance, and rotor speed
Shaft vibration is the preferred basis for evaluating machines with a large casing to rotor
weight ratio. For casing measurements, a significant service factor (3-5) must be used

202
Chapter 10

Fault Diagnostics (and Correction)


Depending on the type of equipment being monitored and the maintenance strategy being
followed, once a faulty condition has been detected and the severity of the fault assessed, repair
work or replacement will be scheduled. However, in many situations, the maintenance strategy
involves further analysis of the vibration signal to determine the actual type of fault present. This
information then allows for a more accurate estimation of the remaining life, the replacement
parts that are needed, and the maintenance tools, personnel, and time required to repair the
machinery. For these reasons, and many more, it is often advantageous to have some idea of the
fault type that exists before decisions regarding maintenance actions are made.

There are obviously a large number of potential different fault types. The description of these
faults can be systemized somewhat by considering the type of characteristic defect frequencies
generated (synchronous to rotating speed, sub-synchronous, harmonics related to rotating speed,
non-synchronous responses). Such a systemization requires a focus on frequency-domain
analysis tools (primarily frequency spectra). While this organization strategy is effective, it
inherently leaves out potentially valuable information from other display formats. For this
reason, the various faults that usually develop in machinery are listed here in terms of the forcing
functions that cause them, specific machine elements and specific machine types. In this way, a
diagnostic template can be developed for the different types of faults that are common in a given
facility or plant.

10.1 Overview of Machine Fault Diagnostics

Frequencies are the key to analysis and fault diagnosis. Vibrations are produced by vibratory
forces due to machine wear, design and installation defects. Impulse forces may excite system
natural frequencies. The machine operating frequency and its multiple or fractional orders give
reference and clues to faults.

Fault diagnosis techniques rely on data in the time waveform and orbits. These presentation
formats provide insight into the physical characteristics of the motions of the shaft and casing.
Phase shows the time relationship between vibration and various references points on a rotating
shaft. Frequency spectra show the vibration activity at specific frequencies and specific locations
on a machine.

Table 10.1 lists the different common diagnostic techniques used in the analysis of rotating
machinery.

203
Table 10.1 Diagnostic techniques for rotating machinery.

Technique Use Description Instrument


time waveform modulation, pulses, amplitude vs. time analog and digital
analysis phase, truncation, glitch oscilloscope, FFT
analyzer
orbital analysis shaft motion, sub- relative displacement of digital vector filter,
synchronous whirl rotor bearing in XY oscilloscope
direction
phase analysis force/motion relative time between strobe light, digital
relationships, force and vibration vector filter, analog and
vibration/space signals or between two or digital oscilloscope, FFT
relationship more vibration signals analyzer
spectrum analysis direct frequencies, amplitude vs. frequency FFT analyzer, electronic
natural frequencies, data collector
sidebands, beats, sub-
harmonics, sum and
difference frequencies

10.1.1 Frequency Spectrum Analysis

Frequency spectrum analysis is used to identify operating speed and its multiples, dominant
frequencies, non-synchronous multiples of operating speed (from rolling element bearings), beat
frequencies (amplitude modulation), natural frequencies and sidebands due to gearbox wear.
Diagnosing faults from the vibration spectrum requires the identification of the following
characteristics in a spectrum reveal information about a fault condition.

High first order (1X) peak


High second order (2X) peak
First order harmonics (multiples of 1X)
Presence of sidebands
Non-synchronous peaks
Synchronous Peaks
Sub-synchnonous peaks
Directional peaks (present in the vibration measurements in one direction but not another)

10.1.2 Synchronous, sub-synchronous and non-synchronous responses

Synchronous responses are features in the frequency spectrum that are exact multiples of the
shaft rotational speed (1X). Faults that are represented by this type of response include
unbalance, misalignment, looseness, bent shaft, blade and vane wear, and gear faults.

204
Sub-synchronous responses in the frequency spectrum occur in the area of the spectrum below
the first order (below 1X). Faults that are represented by responses in this range include oil whirl,
cage frequency of a rolling element bearing, belt frequencies, turbulence in pumps, rubs, and
looseness.

Non-synchronous responses are features that are not an exact multiple of the first order response.
Faults that generate non-synchronous responses include cavitation, resonances, oil whip and
whirl, rolling element bearing faults, and unrelated signals from other machines.

Figure 10.1 Frequency spectrum showing synchronous, sub-synchronous, and non-synchronous


vibration responses.

10.2 Fault Diagnosis based on Forcing Functions

Listed and described below are a variety of forcing functions that can result in accelerated
deterioration of machinery or are the result of damaged or worn mechanical components. The list
is not meant to be exhaustive and is in no particular order.

Table 10.2 Forcing frequencies associated with machines.

Source Frequency (multiple of RPM)


Fault Induced
mass unbalance 1X (frequency in once per revolution)
misalignment 1X, 2X
bent shaft 1X
mechanical looseness odd orders of X
casing and foundation distortion 1X
antifriction bearing bearing frequencies, not integer ones
impact mechanisms multi-frequency depending on waveform

205
Table 10.2 Forcing frequencies associated with machines (Contd).

Source Frequency (multiple of RPM)


Design Induced
universal joints 2X
asymmetric shaft 2X
gear mesh (n teeth) nX
coupling (m jaws) mX
fluid-film bearings (oil whirl) 0.43X to 0.47X
blades and vanes (m) mX
reciprocating machines half &full multiples of speed, depending
on design

10.2.1 Resonances

The analysis of resonance problems is beyond the scope of this chapter. However, some basic
description is provided here because of the high likelihood that at some time a resonance will be
excited by repetitive or cyclic forces acting on or nearby a machine. A resonance is the so-called
natural frequency at which all things tend to vibrate. A machines natural resonant frequency is
dictated by the relationship shown in equation 10.1 below. The term n is the natural frequency,
k is the spring stiffness, and m is the mass. Most systems will have more than one resonance
frequency. These resonances (also called modes) can be excited by any forcing function that is at
or close to that frequency. They may also be excited by impacts due to the fact that an impact is
likely to contain most frequencies (see chapter 5). The response amplitude can be 10 to 100
times that of the forcing function. The term critical speed is also used to refer to resonances
when the machine rotating speed equals, or is close to, the natural frequency.

n k m 10.1

The amount of response amplification depends on the damping in the system. A highly damped
system will not show signs of resonance excitation, while a lightly damped system will be prone
to resonance excitations. Resonances can be diagnosed by monitoring the vibration level while
the speed of rotation of the machine is changed. A resonance will cause a dramatic increase in
the 1X vibration levels as the speed is slowly changed. Most machines are designed to operate
well away from known resonance frequencies, but changes to the machine (support structure,
piping connections, etc.) and proximity to other machines may excite a resonance.

Frequency spectra that contain resonance responses will likely have a peak that is constant in
amplitude and frequency regardless of the operating speed. This is because the resonant
frequencies are tied to the structural stiffness and mass, not the speed of the machine. The peak
width may vary depending on the damping of the structure in question.

206
Energy at, or close to, the natural frequency (or an order (multiple) of it) must be present to
excite the resonance. This can take the form of shaft speed (first order, second order, etc.), vane
pass frequency and/or blade pass frequency.

A critical speed is a rotor speed at which the operating speed (or a multiple) is equal to a natural
frequency. Resonance is excited under low damping conditions (less than 15% of critical
damping) when a natural frequency is equal to a forcing frequency. Correction includes tuning
the natural frequency away from the excitation frequency by adding mass or stiffness to the
structure or increasing the damping. Because all systems will have some resonant response at the
critical speed, passing critical speeds quickly while bringing a machine up to normal operating
speed is an important operational strategy. The waterfall plot is a good diagnostic tool for
identification of resonant frequencies.

Figure 10.2 Waterfall plot showing critical speeds during machine startup.

Figure 10.3 Waterfall plot showing critical speeds during machine coastdown.

An interference diagram (also known as a Campbell diagram) is another good way to show the
relationship between operating speed and critical speeds (resonances). The following figures

207
show the 1X, 2X, 3X, and 4X shaft speed harmonic frequency responses for a machine that is
changing speed (RPM on horizontal axis). As the speed of the machine increases the primary
excitation frequencies (1X, 2X, 3X, and 4X shaft speed) excite the different natural frequencies.
If any of these excitations fall within the main operating range (or ranges) there is a danger of
resonances developing. Ideally, all the natural frequencies should all fall outside the main
operating range(s).

In general the higher natural frequencies will not display the same degree of serious vibration
response compared to the first and second natural frequencies (for reasons discussed in chapter
5). Figure 10.4 and 10.5 below show sample interference diagrams.

Figure 10.4 Interference diagram showing 1X operating speed exciting the first natural
frequency.

Figure 10.5 Interference diagram showing 1X operating speed and 3X and 4X operating speed
exciting the first natural frequency and the 2nd, 4th, and 5th natural freqencies respectively.

The previous figures illustrate system responses that assumed constant natural frequencies for all
operating conditions (shaft speeds). In reality, rotor natural frequencies often depend on the shaft

208
rotation rates due to the induced gyroscopic effects and/or variable hydrodynamic conditions in
the fluid bearings. Figure 10.6a shows a comparison of an interference diagram, with the shaft
critical speeds highlighted, and a modal plot showing natural frequencies as a function of
operating speed. In this plot, 1F, 2F and 3F are the natural frequencies, while 1B, 2B and 3B are
estimates of the damping ratios for the different natural frequencies as a function of changing
shaft speed. There is little change in 1B and 3B as the shaft speed changes, resulting in little
change in 1F and 3F (both in terms of response frequency and response amplitude). However, 2B
does change significantly as a function of changing shaft speed, resulting in changes to 2F (both
in terms of response frequency and response amplitude).

Analytically computed values of the natural frequencies as a function of the shaft's rotation speed
can be used to generate what is known as a "whirl speed map". Such a chart can be used in
turbine design as shown in the numerically calculated Campbell diagram example illustrated in
Figure 10.6b. Analysis shows that there are well-damped critical speeds in the lower speed range
(1 and 3) which should not excite natural frequency responses. There are also several lightly
damped critical speeds in the higher speed range (5 and 6). Another critical speed at mode 4 is
observed at 7810 rpm (130 Hz) which is in the dangerous vicinity of the nominal shaft speed
(7413 rpm), but it has 30% damping - enough to safely ignore it. The goal of good rotor design is
to define a rotor design where the operating speed is located well away from any natural
frequencies that are not significantly damped.

Experimentally measured vibration response spectrum plotted as a function of the shaft's rotation
speed (waterfall plot), usually show peak locations for each slice corresponding to the natural
frequencies.

Figure 10.6a A comparison of an interference diagram (lower) with critical speeds (highlighted)
and a modal plot (upper) showing natural frequencies as a function of operating speed.

209
Figure 10.6b A comparison of an interference diagram (upper) with critical speeds (highlighted)
and a damping ratio (lower), both showing natural frequencies as a function of operating speed.

10.2.2 Unbalance

Unbalance (also referred to as imbalance) exists when the center of mass of a rotating component
is not coincident with the center of rotation. Note that this definition applies specifically to rotors
that have only on disk or flywheel in place and therefore the unbalance exists only in one plane.
It is practically impossible to fabricate a component that is perfectly balanced; hence, unbalance
is a relatively common condition in a rotor or other rotating component (flywheel, fan, gear,
etc.). The degree to which an unbalance affects the operation of machinery dictates whether or
not it is a problem.

The causes of unbalance include excess mass on one side of the rotor, low tolerances during
fabrication (casting, machining, assembly), variation within materials (voids, porosity,
inclusions), non-symmetry of design, aerodynamic and fluid forces, and temperature changes.
The vector sum of all the different sources of unbalance can be combined into a single vector
which represents an imaginary heavy spot on the rotor. If this heavy spot can be located and the
unbalance force quantified, placing an appropriate weight 180 from the heavy spot will

210
counteract the original unbalance. Unbalance can result in excessive bearing wear, fatigue in
support structures, decreased product quality, power losses, and disturbed adjacent machinery.

Unbalance results in a periodic vibration signal with the same amplitude each shaft rotation
(360). A strong radial vibration at the fundamental frequency, 1X, (1 x rotational speed) is the
characteristic diagnostic symptom. If the rotor is overhung, there will also be a strong axial
vibration at 1X. The amplitude of the response is related to the square of the rotational speed,
making unbalance a dangerous condition in machinery that runs at high rotational speeds. In
variable speed machines (or machines that must be run-up to speed gradually), the effects of
unbalance will vary with the shaft rotational speed. At low speeds, the high spot (location of
maximum displacement of the shaft) will be at the same location as the unbalance. At increased
speeds, the high spot will lag behind the unbalance location. At the shaft first critical speed (the
first resonance), the lag reaches 90, and at the second critical and above, the lag reaches 180.
This phase relationship that relates the location of the heavy spot to the location of maximum
displacement vibration amplitude is an important one when correcting mass unbalance. Figure
10.7 shows the phase relationship of the centre of the orbit (centre of rotation) versus the centre
of mass (centre of gravity) as a function of shaft speed (through critical speed).

Figure 10.7 Phase relationship of the centre of the orbit (centre of rotation) versus the centre of
mass as a function of shaft speed (through critical speed).

Definition of Unbalance

Unbalance is the unequal distribution of weight on a rotor about its centerline. Unbalance is a
condition in which a rotor imparts vibration force to its bearings as a result of centrifugal forces.
Equation 10.2 defines the force developed by an unbalance. Figure 10.8 shows a schematic of a
shaft cross-section with an added mass causing unbalance and an equivalent unbalance caused by
voids (and the correction counterbalancing mass).

Fc = mr2 10.2

Fc = centrifugal force, m = mass, r = radius from centre of rotation, = rotational speed.

211
Counterbalancing
mass
Voids

Vector
sum of
voids

Figure 10.8 A schematic of a shaft cross-section with an added mass causing unbalance (left) and
an equivalent unbalance caused by voids (and the correction counterbalancing mass) (right).

Figure 10.9 Journal bearing schematic with nomenclature (left) and showing oil wedge
supporting shaft (right).

Figure 10.10 Normal journal bearing balanced forces.

212
Figure 10.11 Journal bearing instability unbalanced forces.

Figure 10.12 Increased velocity may generate an unbalance force.

Journal bearings are designed to support a rotating shaft on a thin wedge of oil. A thin fluid film
is developed between the two surfaces moving relative to one another (the rotating shaft and the
stationary journal). The oil film is entrained (dragged along with) by the the spinning shaft.
Hydrodynamic pressure is created in the film, effectively floating the shaft in the journal and
carrying any applied loads. The high velocity profile within the thin film creates the pressure
gradient that acts to support the rotating shaft.

213
Figure 10.13 Typical radial forces in a fluid film bearing at the support point (left) and typical
bearing oil velocity profile at the point of minimum oil film thickness.

Figure 10.14 shows a plot of the coefficient of friction in the thin film as a function of oil
viscosity, load and shaft speed. The coefficient of friction reaches a minimum for a specific set
of oil viscosity, load and shaft rotational speed values and is unstable above this range.

Figure 10.14 Coefficient of friction in the thin film as a function of oil viscosity, load and shaft
speed.

Figure 10.15 shows and plot of the stability threshold speed versus a combination of the oil
viscosity, shaft speed, bearing length, bearing diameter, bearing load, shaft radius and clearance,
known as the Sommerfeld number.

214
Figure 10.15 Stability threshold speed versus viscosity, shaft speed, bearing length, bearing
diameter, bearing load, shaft radius and clearance.

Centrifrugal force due to unbalance can be defined using the units of Once-Inches (see Figure
10.16) or Gram-Inches (see Figure 10.17). It is a result of the actual unbalance mass and the
distance the mass is from the centre of rotation.

Figure 10.16 Centrifrugal force due to unbalance in units of Once-Inches.

215
Figure 10.17 Centrifrugal force due to unbalance in units of Gram-Inches.

Causes of Unbalance

The most basic description of the cause of unbalance is an excess of mass on one side of a rotor.
The resulting centrifugal force pulls the rotor toward the heavy side. Excess mass on one side of
a rotor may be caused by low tolerances during fabrication (casting, machining, assembly),
variation within materials (voids, porosity, inclusions, variable density, finishes, etc.), non-
symmetry of design (motor windings, part shapes, locations), and non-symmetry in use
(distortion, size changes, shifting parts due to stress, aerodynamic forces, and temperature
changes).

Manufacturing processes are a major cause of unbalance. Cost plays a role in achieving good
balance. That is, a extremely well balanced rotor can be achieved, but it may cost a lot.
Unbalance can be corrected by adding or removing mass from the rotor at the appropriate
location. There are advantages and disadvantages to both these practices. Typically unbalance
problems should be considered last because any changes after a rotor has been balanced are
likely to require that the rotor be balanced again.

Even small amounts of extra mass, such as dust on the fan blades, attached to a rotor that has a
large diameter will generate large unbalance forces (see Figure 10.18).

Note that some machinery is designed to operate out-of-balance, such as shakers, sieves, and
some types of material transport systems.

216
Figure 10.18 Dust on fan blades - a common source of unbalance forces.

Figure 10.19 Typical time waveform and frequency spectrum representing unbalance.

Unbalance Correction Methods

As mentioned previously unbalance may be corrected by adding or removing mass from a rotor.
The addition of mass typically results in up to a 20:1 vibration amplitude reduction on the first
try (if done carefully). If space limitations exist more than one addition of mass may be required
with the vector sum of all added masses balancing the original unbalance.

Addition of mass may be done by adding solder or epoxy. This method takes more time than
some other methods and the centre of mass (centre of gravity) of the added mass is difficult to
control. However, it is low cost. The addition of standard washers that are bolted or riveted into
place is also quick. However, the incremental sizes mean that an exact balance mass may not be
possible without cutting a washer or some other adjustment. The addition of mass through the
adding of pre-manufactured weights is similar to the adding of washers as incremental sizes are
also used. It is a quick method and handy as most rotors have flywheels or other suitable
locations for adding these masses. Finally cut to size masses may be added by welding them in
place. This method takes the most time, but can be the most accurate.

217
Figure 10.20 Typical hardware for balancing (balance wheel left, pre-manufactured masses
centre, assorted hardware right).

Removal of mass may result in a 10:1 vibration amplitude reduction on the first try (again, if
done correctly). Material removal methods include drilling, milling and grinding. Drilling is
quick and very accurate. Milling is also accurate, but best used for large corrections because of
the large cost. Grinding is basically a trial and error method where accurate removal of mass is
difficult.

Types of Unbalance

There are four basic types of unbalance. These are; static (or force) unbalance, quasi-static
unbalance, couple (or moment) unbalance, and dynamic unbalance.

Static Unbalance (force unbalance)

Static unbalance is also known as force unbalance. This type of unbalance is when the principal
axis of inertia is displaced parallel to the shaft axis. This type of unbalance is found mostly in
narrow, disk-shaped parts (fly wheels, grinding wheels, fans, turbine wheels). To correct this
type of unbalance a single correction mass is placed opposite the centre-of-gravity in a plane
perpendicular to shaft axis and intersecting the centre-of-gravity. A technique known as knife
edge balancing is also possible. This is where the rotor is placed in a cradle or on a support that
allows it to spin freely. The heavy spot will rotate to the bottom due to gravity.

Figure 10.21 Rotor with static unbalance showing unbalance force and other parameters.

218
Figure 10.22 Concentric disk with static unbalance.

Figure 10.23 Eccentric disk causing static unbalance.

Figure 10.24 Two disks of equal mass and identical static unbalance.

Static unbalance vibrations measured at different locations (opposite ends of the rotor for
example) will be in-phase and steady in amplitude as long as the speed is constant. The
amplitude will increase as the square of the change in the speed of rotation. That is, if the speed
increases by a factor of 3 the vibration amplitude will increase by a factor of 9. Vibration at 1X
RPM is always present and dominates the frequency spectrum.

Figure 10.25 Rotor with static unbalance has vibrations at all locations that are in-phase.

219
Figure 10.26 Rotor with disk and static unbalance on soft bearings. Note that the shaft
vibration at each bearing is significant, but still in-phase.

Figure 10.27 Rotor with disk and static unbalance on hard bearings. Note that the shaft
vibration at each bearing is very small and the vibration beyond the bearing is out-of-phase with
the shaft between the bearings.

Figure 10.28 Rotor with an overhung disk and static unbalance on hard bearings. Note that the
shaft vibration at each bearing is very small and the vibration beyond the bearing is out-of-phase
with the shaft between the bearings. In this case the vibration amplitude at the disk in is large
compared to other locations due to the lack of support at that end of the shaft.

220
Couple Unbalance (moment unbalance)

Couple unbalance is also known as moment unbalance because there are two unbalance forces
that cause a couple (or moment) to act about the long axis of the shaft. In couple unbalance the
principal axis of inertia intersects the shaft axis at the centre of gravity (centre of mass). There
are two equal unbalance forces developed at opposite ends of shaft and 180 apart. Dynamic
balancing methods needed.

Figure 10.29 Rotor with couple unbalance showing unbalance forces.

Figure 10.30 Couple unbalance in a solid rotor

Figure 10.31 Couple unbalance in two disks of equal mass.

Couple unbalance vibrations will be 180 out-of-phase when measured at locations at opposite
ends of the shaft. Vibration at 1X RPM is always present and dominates the frequency spectrum.

221
The amplitude will increase as the square of any change in the speed of shaft rotation. There may
be high amplitude axial vibrations as well as radial vibrations.

Figure 10.32 Rotor with couple unbalance has vibrations at opposite ends of the shaft that are
out-of-phase.

Figure 10.33 Couple unbalance on an outboard rotor component.

Figure 10.34 The disk-shaped rotor mounted with minimum axial swash motion (A and B)
requires only single-plane balancing because only static unbalance is likely. The rotors with
mounting errors similar to C and D may exhibit unacceptable moment unbalance and must be
balanced in two planes.

Overhung rotor unbalance vibrations are at typically found at 1X RPM and in the axial and radial
directions. The axial vibrations tend to be in-phase while the radial vibrations may have unsteady
phase readings. Overhung rotors usually have a combination of static and couple unbalance.

222
Quasi-Static Unbalance

Quasi-static unbalance is, as the name implies, a static unbalance where the principal axis of
inertia intersects the shaft axis at a point other than the centre of mass. It is in fact a combination
of static and couple unbalance conditions.

Figure 10.35 Quasi-static unbalance in a solid rotor.

Figure 10.36 Couple plus static unbalance resulting in quasi-static unbalance in a solid rotor.

Figure 10.36 Rotor assembly with unbalance in the coupling resulting in quasi-static unbalance.

Dynamic Unbalance

Dynamic unbalance is when the principal axis of inertia is neither parallel to, nor intersects the
shaft axis. Dynamic unbalance must be corrected in at least two planes perpendicular to the shaft
axis.

223
Figure 10.37 Dynamic unbalance showing unbalance forces.

Figure 10.38 Dynamic unbalance in a solid rotor.

Figure 10.39 Couple plus static unbalance resulting in dynamic unbalance in a solid rotor.

Figure 10.40 Dynamic unbalance may be considered as a multitude of unbalanced disks together
on a rotor.

224
Figure 10.41 Turbines and turbo-fans are where a multitude of unbalanced disks may exist
together on a rotor resulting in dynamic unbalance.

Rotor Motions

Static unbalance results in a rotor motion where all points vibrate in the same direction at the
same time. The measured vibration signals are in-phase. Couple unbalance results in rotor
motion where points at opposite ends of the rotor vibrate in opposite directions at the same time.
The vibration signals measured at opposite ends of the rotor are out-of-phase. In the case of
quasi-static unbalance (static and couple unbalance) the apex of vibration is moved away from
centre of mass. In the case of dynamic unbalance the only thing that can be said is that the
vibrations along the shaft (or rotor) are complex.

Figure 10.42 The effect of unbalance on free rotor motion.

Runout versus Unbalance

Often times the terms runout and unbalance are used interchangeably. However, runout is also
often used to describe the linear radial displacement of the outside surface of the rotor. Linear
radial displacement of the outside surface of a shaft or rotor may be due to things other than
unbalance, such as voids, density changes, rust pits or other damage. Zero runout may not
indicate a balanced rotor.

225
The Effects of Rotational Speed (on Unbalance Forces)

As mentioned previously, at low speeds the high spot (the location of maximum displacement
radial of a shaft) is at the same location as the unbalance mass. As the shaft speed of rotation
increases the high spot will lag behind the unbalance mass location. At the first shaft critical
speed (first resonance) the high spot lag behind the unbalance mass location reaches 90. At the
second critical speed (the second resonance) and above the lag reaches and remains at 180.

Figure 10.43 Angle of lag and migration of axis of rotation. At low speed well below the first
critical speed the high spot and the heavy spot are at the same location (A). As the shaft speed
increases the high spot tends to lag behind the heavy spot (B). At the first critical speed (first
resonance) the high spot lags behind the heavy spot by 90 (C). At the second critical speed
(second resonance) and above the high spot lags the heavy spot by 180 (D).

The correlation between the displacement of the center of gravity with an unbalance mass (force)
is an important relationship. A clear understanding of this relationship is required when
correcting for unbalance, setting balancing procedures, and selecting or setting bearing
tolerances. In the case of disk shaped rotors this relationship is relatively straight forward and
simple. For long rotors several simplifying assumptions need to be made. This chapter will deal
only with simple disk shaped rotors.

As an example, consider a rotating disk spinning (see Figure 10.44) with an angular velocity ,
with an unbalance mass u at radius r. The disk has a centre of mass at S which is displaced by a
value e (eccentricity), creating a centrifugal force F. Also consider that the weight of disk is 999
oz. A mass of 1 oz. is added to the disk to create an unbalance. The total weight, W, is then 1000
oz. If the unbalance mass is placed 10 inches from center of rotation, the unbalance force created
is U = 10 in 1 oz = 10 oz in.

226
Figure 10.44 Rotating disk with unbalance mass example.

Notwithstanding that the vibration levels generated will be dependent on the shaft speed of
rotation, the information above can be used to determine the displacement of the centre of mass
from the geometric centre of the disk. The centre of mass will rotate at a distance e from the shaft
rotational axis. This eccentricity is a function of the mass of the shaft (disk) and the unbalance
mass (as indicated below)

U=We

10 oz. in = 1000 oz. e


U oz.in
e=
W oz.
10oz.in
e= = 0.01 in
1000oz

Figure 10.45 Rotating disk with unbalance mass and calculation of the centre of mass relative to
the centre of rotation.

The example above used a value for the unbalance force of 10 oz. inches. This came from the 1
oz. mass place 10 inches from the centre of rotation of the disk. The same unbalance force could
come from a 2 oz. mass placed 5 inches from the centre of rotation. This flexibility in creating an

227
unbalance force by placing different sized masses at different distances from the shaft centre of
rotation is useful when adding mass to a rotating shaft in order to correct for an existing
unbalance.

Figure 10.46 Different unbalance mass sizes and locations that result in the same unbalance
force.

There are two important points to be noted. The displacement of the centre of mass from the
centre of rotation, e, is always only 1/2 of the measured relative vibration amplitude. That is,
displacement vibration sensors always measure twice the shaft eccentricity. As the centre of
mass moves around the shaft centre of rotation the eccentricity moves around as well, effectively
doubling the resulting vibration magnitude. Also, the relationship between the rotor mass and the
unbalance and the eccentricity is true only for disk shaped rotors experiencing static unbalance.

Figure 10.47 is a monograph relating rotor mass, static unbalance and eccentricity. It can be used
to quickly determine any one of the three related parameters knowing any of the other two. The
following example will show how this graph can be used.

A pulley with mass of 20kg has a static unbalance mass u = 5g added at a radius r = 100mm.
Calculate the magnitude of the center of gravity eccentricity e.

u r 5 g 100mm .005kg 0.1m


Solution: e 0.000025m 25 m
m 20kg 20kg

Alternatively, Figure 10.47 can be used to find e from the known values for rotor mass (20kg)
and static unbalance (500g.mm). See Figure 10.48 for the detailed graphical solution.

228
Figure 10.47 Monograph relating rotor mass, static unbalance and eccentricity.

229
Figure 10.48 Graphical solution to the above example (finding eccentricity).

By including the speed of rotation of the shaft (or disk) the unbalance force can be determined.
Continuing with the previous example the shaft speed of rotation in included in the example
below.

The disc in the previous example with the static unbalance u = 5g at a radius r = 100mm rotates
at a speed of n = 6,000rpm (f =100cps = 1=Hz). Calculate the centrifugal force in Newtons
induced by the unbalance.

Solution: Fu u r 2

Fu 0.005kg 0.10m 2 100


2

Fu 197 N

As in the previous example this solution can also be obtained using an appropriate monograph.
Figure 10.49 shows a monograph relating static unbalance, shaft speed of rotation and unbalance
force.

Figure 10.50 shows the detailed graphical solution to the above example (finding the unbalance
force).

230
Figure 10.49 Monograph relating static unbalance, shaft speed of rotation and unbalance force.

231
Figure 10.50 Graphical solution to the above example (finding the unbalance force).

232
The total unbalance force is shared between the support bearings. Figures 10.51 and 10.52 show
the relationships between the total unbalance force acting on the rotor and the unbalance force
acting on each supporting bearing for the cases of static unbalance and couple unbalance.

Figure 10.51 Centrifugal forces and bearing forces static unbalance.

Figure 10.52 Centrifugal forces and bearing forces couple unbalance.

For the case where the unbalance weight is near one end of the rotor, equation 10.3 can be used
to calculate the displacement of the principal axis of inertia away from the shaft axis of rotation
at each bearing.
mr mrjh
d= 10.3
W m Ix Iz

d = displacement of principal axis of inertia from shaft axis at the bearing


W = rotor weight
m = unbalance mass
r = radius of unbalance
h = distance from center of gravity to plane of unbalance
j = distance from center of gravity to bearing
Ix = moment of inertia around transverse axis
Iz = polar moment of inertia around shaft axis

Since Ix and Iz may not be known, it is acceptable in most cases to assume that the unbalance
causes parallel displacement of the principal axis of inertia. In reality most shafts are circular and

233
therefore have a relatively straight forward equation for determining Ix and Iz. The result is (as
just stated) that the unbalance causes parallel displacement.

Single-plane balancing (general)

Typical rotors which often require only single-plane balancing, especially in the assembled
condition, include, fans, grinding discs, die chucks, pulleys, flywheels, clutches, gears, and pump
impellers. Figure 10.53 shows two typical strategies for balancing disk shaped rotors. The rotor
may be removed from the machine and balanced on a fixture that allows the rotor to move freely
under to the effects of gravity. Alternatively, the rotor may also be balanced in place. There are
advantages and disadvantages to both these procedures which will be listed and discussed later in
this chapter.

Figure 10.53 Horizontal balancing machine (left) and instrumentation for an in-situ balancing
(right).

Two-plane (dynamic) balancing

Typical rotors which should be balanced in two planes include paper machine rolls, steam
turbine/generator sets, electric motor and generator armatures, crushing and cutting rotors,
machine tool spindles, grinding rolls, fans and blowers with long distances between the end
plates and compressor rotors.

Figure 10.54 Multi plane balancing machines for use in balancing long rotors or rotors with
multiple disks.

234
Field Balancing (in-situ)

Field balancing is basically a balancing of a rotor in place. There are many advantages to this
technique (as listed below), but also some disadvantages. Generally the most significant
disadvantage is that field balancing often is not able to correct for unbalance as effectively as
balancing done on a balancing machine. The full list of advantages to field balancing include the
following.

the rotor is balanced on its own bearings


balancing takes place at normal operating speed
balancing takes place at normal load
the rotor is driven in the same manner as during normal operation
there is no tear down, re-assembly and re-alignment necessary
an in-place trim balance (following re-installation) is not required
down time is greatly reduced
generally simple procedures are employed which require only the starting and stopping of
the machine (which may be time consuming) and adding or removing of correction
weights

Figure 10.55 Field balancing can be done with a two channel data collector (one channel collects
vibration data and one channel collects phase (speed) data).

Before starting the balancing task it is important to determine if mass unbalance is the actual
problem by performing a complete vibration analysis to eliminate and/or correct all other
vibration problems. If mass unbalance is not the problem, look into and correct other problems
(which may include excessive bearing clearance, looseness, resonance, and/or misalignment). If
mass unbalance is the cause of the excessive vibration, continue with pre-balancing checks.

The nature of the unbalance problem is also important. Different types of unbalance need to be
corrected in different ways. A complete vibration analysis should be carried out to determine the
type of unbalance problem that exists. An inspection of the rotor should also be done to
determine whether or not the rotor is clean. Rotor stability (structural, thermal) should also be

235
assessed as part of the overall vibration analysis. Start-up and/or coast-down tests may be
required to determine shaft rotational critical speeds (if these are not already known). The
balance weights already in place should be located and noted. The details of any balance planes
or rings in place should also be determined.

Sensors for Measuring Unbalance

Any of the three standard sensors for measuring vibration signal in machine condition
monitoring applications will work for measuring the vibration signals for balancing. However, it
should be noted that proximity probes are the most direct measure as they show output in terms
of displacement which are directly linked to unbalance forces and the shaft response. Velocity
transducers and accelerometers provide indirect measures which contain what is referred to as
electronic lag. This is the difference in the output quantity as it relates to the displacement (see
description below). Recall from chapter 5 that velocity lags displacement by 90 and acceleration
lags displacement by 180.

Sensors for measuring phase are also required. A strobe light can be used. These devices are
simple to use, but may result in a less accurate phase reading as multiples of rotational speed are
sometimes not clearly distinguishable from the primary rotational speed. Photoelectric sensors
and proximity probes are accurate phase reading devices and relatively inexpensive.

Proximity Probe Measurement.

The instantaneous vibration amplitude measured by a displacement transducer is,

d = D/2 sin (t)

d = instantaneous displacement
=2f
f = frequency (Hz)
t = time (sec)
D = Peak-to-Peak displacement

Using displacement for balancing simplifies phase measurements and calculation of correction
weight placement. No electronic phase lag exists between the vibration signals and the shaft
rotational speed (position) signal.

Figure 10.56 Proximity probe vibration signal and key phasor shaft position signal (no electronic
lag).

236
Velocity Probe Measurement.

The instantaneous vibration amplitude measured from a velocity transducer is,


d D
v= d cos(t )
dt 2
v = instantaneous velocity
=2f
f = frequency (Hz)
t = time (sec)
D = Peak-to-Peak displacement

Both electronic and mechanical phase lag may exist in these signals and need to be compensated.
The electronic lag comes from the fact that velocity measurements lag displacement by 90.
With the speed measured using a strobe or key phasor there may be an additional 90 of
mechanical phase lag depending on the speed of the shaft (above or below the first natural
frequency (critical speed).

Acceleration Measurement.

The instantaneous vibration amplitude measured with an accelerometer transducer is,


d 2D
a= v sin(t )
dt 2
a = instantaneous acceleration
=2f
f = frequency (Hz)
t = time (sec)
D = Peak-to-Peak displacement

Again electronic lag exists in these signals. The electronic lag comes from the fact that
acceleration measurements lag displacement by 180. Mechanical phase lag may exist in these
signals as well depending on the speed of the shaft (above or below the first natural frequency
(critical speed) and may need to be compensated.

The effect of Rotational Speed on the Phase of Measurements

As mentioned previously, the force caused by the mass unbalance (heavy spot rotating) will lead
the vibration peak (high spot) by 0 to 180 depending the shaft rotational speed. When the
operating speed is less than the first critical speed the phase lag in between 0 and 90. When the
operating speed is close to the first critical speed the phase lag approaches exactly 90. When the
operating speed is beyond the first critical speed the phase lag is between 90 and 180. At
operating speeds above second critical speed the phase lag typically does not exceed 180.

237
Figure 10.57 and Figure 10.58 show diagrams that describe graphically the relationship between
shaft rotational speed and the phase lag between the timing of the maximum vibration amplitude
(at 1X rotational speed) and the location of the unbalance mass that is causing the vibration.

Figure 10.57 The relationship between rotor speed, relative maximum displacement and phase
lag between the radial location of the maximum vibration amplitude and the radial location of the
force due to the unbalance mass.

Figure 10.58 Schematics of a rotor operating well below the shaft first critical speed (left) and
above the shaft first critical speed.

238
Figure 10.59 Establishing the phase angle and speed reference can be done with a photo-electric
sensor and a reference mark (reflective or black tape).

Single Plane Balancing (details)

There are four basic steps involved in single plane balancing. These are listed below.

1 The as is run, which establishes an accurate state of affairs.


2 The trial weight run, which allows the definition of how the rotor will react to the
addition of a known trial weight.
3 The calculation of the required permanent correction and the placement of that
permanent correction weight.
4 - The final run to check on the new balance conditions.

It should be noted that the first step in the balancing process assumes that all other potential
vibration causing problems or conditions have been eliminated as causes of any excess vibration
and there is no other work to be done on the machine. It also assumes that the machine operating
characteristics (critical speeds) are already well known.

Figure 10.60 Photo (left) and schematic (right) showing the external view of a machine about to
undergo balancing. The number 1 shows where the measurement plane is located and A
shows where the correction plane (where weights could be added) is located.

239
In single plane balancing the as is run provides the baseline data that shows the condition of
the machine (in terms of vibration response amplitude and phase). It also ensures repeatability
in data collection, which helps confirm that it is unbalance that is the problem as well as
confirms the reliability of the data collection instrumentation. An example of such a
measurement simply states the vibration amplitude (peak) and position of the maximum
amplitude in relation to some reference on the rotating shaft. (10 at 30 suggests 10 mils peak-to-
peak at 30 phase relative to some reference).

The next step will provide information related to the location and weight of unbalance causing
the measured vibration. This is done by first adding a trial weight to the rotor. In single plane
balancing the trial weight run shows the response due to the addition of a known unbalance
weight in a known location. This information can then be used to locate the location of the
unbalance mass that is causing the vibration problem. Note again that it is important to know
how close the machine operating conditions are to any critical speeds. Ideally the operating
conditions will be far away from the critical speeds, but if they are close (just below or just
above) a resonance related response may be generated. If it is suspected that the operating speed
is close to a critical speed there will be a phase change immediately after shut down and before
coast down. Dramatic changes in phase mean operational speed is close to a resonance.

When calculating the unbalance weight an extra 10% should be added onto the weight of the
rotor to account for vibrations absorbed into the bearings and supports. The trial weight should
be approximately equal to the expected unbalance weight divided by an amplification factor
(determined from the location of the operating frequency on the system response curve more
on this later). The trial weight should be added opposite the existing unbalance location in order
to ensure that the unbalance condition is not worsened. Following this, the trial weight run
vibration data is collected.

The trial weight (WT) calculation formula (equation 10.4) is shown below.

WT= 56,375.5 (W/N2e) (ounces) 10.4

W - rotor weight (lb)


N rotor speed (RPM)
e eccentricity (in)

If no vibration response is obtained, either the trial weight is too small or the problem is not mass
unbalance.

The final calculation of the trial weight resultant vector in single plane balancing involves
combining the trial weight vector and the initial unbalance vector. Typically this correction is
determined graphically, but may also be done using vector calculations.

The final run is a restart of the machine under consideration to check vibration levels. If these
new measured levels are unacceptable this measurement should be treated as a new as is run. It
is not wise to tinker with weights already added.

240
The Vector Method of Single Plane Balancing

The following is a simple example of single plane balancing. Please refer to Figure 10.61 as you
read the following section.

Note and mark the high spot (a) and the amplitude (oa).
Place a trial weight WT at the selected location (to generate a response opposite the original
unbalance vibration response). Note and mark the location of the new high spot (b) and its
amplitude (ob).
The vector difference ab = ob oa is the effect of WT alone.
Move the WT in the same direction and angle to make (ab) parallel to and opposite (oa).
The trial weight is increased or decreased in the ratio (oa/ab) equal to the original unbalance.

Figure 10.61 Simple example of single plane balancing #1.

Another simple example of single plane balancing, but involving a graphical solution (and a few
more details) is described below. Please refer to Figure 10.62 as you read the following section.
Detailed procedures for constructing a vector diagram for single-plane balancing are included.

1. Mark the direction of rotor rotation on the graph.


2. Mark the direction of positive phase angle.
3. Establish a scale of numbers (mils per division) so the vectors are large but do not exceed the
graph overall dimensions.
4. The original vibration O (5 mils at 190) is plotted on the graph.
5. The location of the trial weight (WT) is plotted (30) and its size (75 grams) are noted on the
graph.
6. Plot the vibration (original + trail, O+T) obtained after the trial weight has been added to the
rotor. The rotor must be operated at the same speed as when the original data (O) were
acquired.
7. The difference between O and (O+T) is the effect of the trial weight.

241
8. The effect of the trial weight is obtained by drawing a line between O and (O+T).
9. O + T must be equal to (O+T). That is, the arrow T must point to the end of (O+T). Vectors
add heads to tails and subtract heads to heads.
10. T is now repositioned with its tail at the origin by moving it parallel and maintaining the
same length.
11. Draw a line opposite O from the origin.
12. The goal in balancing is to add a trial weight that will create a T vector directly opposite and
equal to O.
13. The angle between T and the line opposite O is 36 and it determines how far and in what
direction the trial weight must be moved.
14. The trial weight is multiplied by the ratio of the original vibration to the effect of the trial
weight (5/3.4 in this case, 75 g (5/3.4) = 110 g) to determine the final correction weight.
15. The angle between T and O is used to determine the location of the final balance weight.

242
Figure 10.62 Simple example of single plane balancing #2.

Weight splitting may be used to place multiple balance masses in locations that allow the total
balance force to sum to a suitable amplitude and direction. This is often done where the
geometry of the machine will not allow for a suitably sized correction weight to be placed in one
particular location. A parallel rule can be used to determine graphically the magnitudes of the
weights at desired locations (see Figure 10.63) by the lengths of the vectors. Weight combination
is the inverse process used to determine the location and magnitude of the combined weights.

243
Figure 10.63 Weight splitting (left) and weight combination (right).

Acceptable vibration levels standards exist particularly for machines that are being balanced for
the first time or after some major work. Two such standards are the Blake chart and the Dresser-
Clark chart.

The modified Blake chart (see Figure 10.64) for field balancing shows acceptable vibration
levels in displacement, velocity and acceleration parameters.

The Dresser-Clark chart (see Figure 10.65) shows proximity probe measurements and
recommendations for appropriate action.

Multiple plane balancing (which will not be covered in detail in this book) requires multiple
correction planes to be investigated. Typically the number of bearings plus one equals the
number of planes that need to be considered. This rule of thumb depends on the flexibility of
rotor. Several trial runs may be required to begin with. Cross plane effects need to be considered
and the solution usually involves a matrix solution.

244
Figure 10.64 Modified Blake chart for field balancing showing acceptable vibration levels.

Figure 10.65 Dresser-Clark chart showing proximity probe measurements and recommendations
for appropriate action.

245
Summary of Unbalance

Mass unbalance of a rotor results when the mass centre is not at the same location as the
geometric centre.
Mass unbalance causes a rotating force at the frequency of shaft speed.
The amount of mass unbalance force depends on the location of the mass centre from the
geometric centre, the weight of the object, and the square of the speed.
Balancing is a procedure in which a balance weight that creates a force equal to the mass
unbalance is placed opposite the effective location of the mass unbalance.
The heavy spot is the angular location of the mass unbalance on the rotor.
The high spot is the angular location of the peak of vibration (displacement).
The high spot is measured during the balancing process; however, the balance weight must
be positioned opposite the heavy spot.
Either displacement, velocity, or acceleration can be measured; however, displacement is
preferred.
The high spot may lag the heavy spot as a result of electronic (instrument) and mechanical
lag. The amount of mechanical lag depends on the shaft speed relative to the first critical
speed.
Balancing should not be performed until it is evident that misalignment, excessive bearing
clearance, looseness, and distortion are not the cause of the vibration at operating speed.
The rotor should be clean and structurally sound prior to balancing.
Trial or calibration weights are used to obtain the mechanical lag.
The rule of thumb for selecting a trial weight is that it should create a force of not more than
10% of the rotor weight.
The vector method is used to determine the size and location of the correction weight.
Vibration is measured on the machine with and without the trial weight.
The vector difference is determined to assess the effect of the trial weight.
The trial weight is moved relative to the effect vector so that it is opposite the original
unbalance vector.
The size of the trial weight is adjusted so that the effect vector is the same length as the
original unbalance vector.
Allowable field unbalance values are obtained from vibration severity levels in ISO 2372
(rms) and the modified Blake chart.

Often persistent unbalance problems are dealt with be replacing or redesigning the support
bearings. Before this is (usually expensive and time consuming) strategy is pursued, the
following checklist should be considered.

Check that the oil reservoir contains the correct lubricant.

246
Check the oil quality for proper density, viscosity, water content, and other contaminants.
Check for proper oil supply pressure, temperature and system control.
Check the oil flow rate to each bearing and verify that orifices are properly installed and that
orifice diameters are both correct and reasonable.
Check the oil drain temperatures and relative flow rates.
Check that the bearing is properly installed with respect to shaft rotation.
Check that anti-rotation pins are properly installed with respect to shaft rotation.
Check that the shaft to bearing clearance is correct.
Check that the bearing to housing clearance is correct.
Check that the bearing liner is not distorted or warped.
Check that the bearing splitline is not sealed with silicone or some other incompressible
sealant.
Check for other mechanical changes in the train that would influence bearing load (changing
a gear coupling to a large diaphragm coupling for example).
Check rotor balance records and the last set of transient start-up data.
Check couple alignment for proper cold offset and hot running position.
Check for proper temperatures from imbedded thermocouples and RTDs.
Check bearing temperature trends (day to night, week to week, etc.).
Check to be sure that shaft is level when hot and running.
Check bearings, seals and couplings for evidence of electrical discharge.
Check pads and backing for evidence of wear, cracking or fretting.
Check bearings for evidence of edge wear.
Check for proper position of the rotor/shaft within the bearing with proximity probes.
Check shaft vibration for normal 1X running speed vibration amplitude and phase.
Check shaft vibration for any abnormal frequency components.
Check the attachment of the bearing housing to the casing and/or baseplate.
Check grout condition and the attachment of baseplate to the foundation.

10.2.3 Bent Shafts and Bowed Rotors

A special form of unbalance is caused by a bent shaft or bowed rotor. These two conditions are
essentially the same; only the location distinguishes them. A bent shaft is the term used when the
rotating shaft is located outside the machine housing. A bowed rotor is inside the machine
housing. This condition (refered to as a bent shaft or bowed rotor) is typically seem on large
machines (with heavy rotors) that have been allowed to sit idle for a long time. Gravity and time
cause the natural sag in the rotor to become permanent.

247
The vibration spectrum from a machine with a bent shaft or bowed rotor is identical to
unbalance, largely because it is an unbalanced condition. Bent shafts and bowed rotors are
difficult to correct (straighten), so they need to be balanced by adding counterweights as
described above. The best way to avoid this condition is to keep heavy shafts/rotors rotating
slowly when the machine is not in use. This condition may also be caused by local (uneven)
heating of the shaft due to a rub (discussed later in this chapter)

Figure 10.66 Bent shaft (or bowed rotor) causing unbalance vibration.

10.2.4 Misalignment

While misalignment can occur in several different places (between shafts and bearings, between
gears, etc.), the most common form is when two machines are coupled together. In this case,
there are two main categories of misalignment: (1) parallel misalignment (also known as offset)
and (2) angular misalignment. Parallel misalignment occurs when shaft centerlines are parallel
but offset from one another in the horizontal or vertical direction, or a combination of both.
Angular misalignment occurs when the shaft centerlines meet at an angle. The intersection may
be at the driver end or driven end, between the coupled units or behind one of the coupled units.
Most misalignment is a combination of these two types.

Misalignment is another major cause of excessive machinery vibration. It is usually caused by


improper machine installation. Flexible couplings can tolerate some shaft misalignment, but
misalignment should always be minimized.

The vibration caused by misalignment results in excessive radial loads on bearings, which in turn
causes premature bearing failure. Elevated 1X vibration with harmonics (usually up to the third,
but sometimes up to the sixth) in the frequency spectrum are the usual diagnostic signatures. The
harmonics allow misalignment to be distinguished from unbalance. High horizontal relative to
vertical vibration amplitude ratios (greater than 3:1) may also indicate misalignment.

The heat of operation causes metal to expand resulting in thermal growth. Vibration readings
should be taken when the equipment is cold and again after normal operating temperature has
been reached. The changes in alignment due to thermal growth may be minimal, but should
always be measured since they can lead to significant vibration levels.

Types of Misalignment

Parallel Misalignment (offset)

248
Parallel misalignment (offset) is defined as misalignment that results when shaft centre lines are
parallel but offset from one another. This type of misalignment may involve displacement in the
horizontal, vertical or a combination of both directions.

Figure 10.67 Parallel misalignment.

Angular Misalignment

Angular misalignment is defined as the misalignment that occurs when shaft center lines meet at
an angle.

Most misalignment is a combination of parallel and angular misalignment.

Figure 10.68 Angular misalignment.

Bearing Misalignment (cocked bearing)

Bearing misalignment is defined as when the shaft center lines are properly aligned, but the
bearings on one side of the coupling are misaligned with the shaft and/or with another bearing).
That is they are not mounted in the same plane (not normal to shaft). When bearings are
misaligned the machine tends to distort in use (soft foot, uneven base, thermal growth).

A cocked bearing will cause considerable axial vibration. This type of misalignment may also
cause a twisting motion with approximately 180 of phase shift from top to bottom and/or from
side to side as measured in the axial direction of the same bearing housing. Attempts to align the
coupling or balance the rotor will not alleviate the problem. The bearing must be removed and
correctly re-installed.

249
Figure 10.69 Cocked bearing.

Figure 10.70 Axial vibration frequency spectrum from a misalignment condition.

Alignment Methods

There are two primary alignment methods; the reverse dial method, and the face and rim method.
Both of these methods my involve dial indicators or laser sensors to measure the actual
misalignment distance. However, before going on to describe these two methods there are a few
simpler methods that can be used in situations that do not require extreme accuracy. These
alternatives include using feeler gages (as shown in Figure 10.71), straight edges (as shown in
Figure 10.72), callipers (as shown in Figure 10.73), and micrometers (as shown in Figure 10.74).

Figure 10.71 Feeler gage used to measure alignment.

250
Figure 10.72 Straight edge used to measure alignment.

Figure 10.73 Calliper used to measure alignment.

Figure 10.74 Micrometer used to measure alignment.

251
Reverse Dial Alignment Method

The reverse dial alignment method is simple and accurate. Brackets are used on both shafts on
opposite sides of the coupling. Each bracket holds a rod which spans the coupling. Both rods rest
on dial indicators attached to the opposite bracket. Long rods between the brackets improve
accuracy. For long spool pieces the position dial indicator stems may be placed against the spool
piece surface.

Figure 10.75 Reverse dial alignment method.

Figure 10.76 Reverse dial alignment method with the dial mounted on the bracket to reduce rod
sag.

Figure 10.77 Dial indicator.

252
Figure 10.78 Reverse dial alignment method with the dial placed against the spool piece surface.

Face and Rim Alignment Method

The face and rim method uses one bracket to hold both dials. The brackets are placed as far apart
as possible to increase accuracy. The rod is placed as high above shaft center line as possible to
amplify even small angles of misalignment. Long rods sag due to their own weight and the dial
indicator weight.

Figure 10.79 Face and rim method.

253
Figure 10.80 Face and rim method with the dial indicator mounted on the bracket to reduce sag.

Figure 10.81 Various alignment measurement configurations.

Alignment Procedures

Alignment procedures are relatively simple. First, the dial indicators are set to zero after
mounting. Four equally spaced (rotationally) readings are then taken (rotate brackets 90 each
time). If a full rotation is not possible, three readings spaced at 90 is usually sufficient. The sum
of the opposite readings (for each individual gage 180 apart) should always be equal. It is
important that the shafts rotate together during the measurement procedure. It is also important
that the shafts do not uncouple during the alignment procedure (axial shaft movement causes
inaccuracies). Both shafts must always rotate in the same direction (coupling backlash causes
inaccuracies). If the coupling is designed to allow some axial movement, this must be restricted.

254
Figure 10.82 Sample alignment results.

Laser based alignment methods are common. The same basic procedure applies, only a laser
beam is used in place of the rod. A laser detector measures beam deflection as the shaft is
rotated. The emitter and detector are placed on opposite shafts. Long coupling spans are possible
primarily because there is no sag (resulting in high accuracy). Laser alignment systems are
typically quick and relatively easy to set up. However, they may also be relatively costly to
purchase.

Figure 10.83 Laser alignment.

Soft feet and frame twists

Another condition that is in fact a type of mechanical looseness, but often masquerades as
misalignment, unbalance, or a bent shaft, is soft foot. Soft foot occurs when one of a machines
hold-down bolts is not tight enough to resist the dynamic forces exerted by the machine. That
part of the machine will lift off and set back down as a function of the cyclical forces acting on
it. All the diagnostic signs associated with mechanical looseness will be present in the vibration
signal.

255
If the foundation (hold-down points) of a machine does not form a plane, then tightening the
hold-down bolts will cause the casing and/or rotor to be distorted. This distortion is what leads to
the misalignment, unbalance, and bent shaft vibration signatures. In order to check for a soft
foot, the vibration level must be monitored while each hold-down bolt is loosened and then
retightened. The appearance and/or disappearance of the diagnostic indicators mentioned above
will determine if soft foot is the problem. When a machines vibration levels cannot be reduced
by realignment or balancing, soft foot could well be the cause. Foot movement of more than 1
mil. (1/40 mm) should be shimmed. Each foot should be tightened and loosened in a set
sequence .

Figure 10.84 Detecting soft feet poor contact between floor and machine feet.

Figure 10.85 Detecting soft feet - loose hold-down bolts.

256
Figure 10.86 Detecting soft feet phase analysis may reveal approximately 180 phase
difference between vertical measurements on the machine foot, base plate and the foundation.

When lifting and moving a large machine it is important to use jack screws wherever possible for
vertical and horizontal movements. A machine should never be moved using a sledge hammer.
Piping connections should be disconnected during alignment Dial indicators can be kept in place
during reconnection as a check for machine movement. Piping should have its own supports and
not rely on the machine it is attached to for support. Dial indicators can be used to estimate rod
sag. Rod sag, even though small in most cases, should not be ignored.

Sag in Alignment Rods

Sag in round rods can be calculated using the following equation.


K
Sg = 2.829 10-5 L3 4 r 4
D d
Sg - dial indicator sag (mils)
L - length of rod (inches)
D - rod outside diameter (inches)
d - rod inside diameter (inches)
Kr - spring constant of circular rod (mild steel)
Kr = 1.334 (D2 - d2) (L + Wd)
Wd - weight of dial indicator (ounces)

Sag in square rods can be calculated using the following equation.

K
Sg = 2.833 10-5 L3 4 s 4
D d
Sg - dial indicator sag (mils)
L - length of rod (inches)

257
D - rob outside width or height (inches)
d - rod inside width or height (inches)
Ks - spring constant of square rod (mild steel)
Ks = 1.699 (D2 - d2) (L + Wd)
Wd - weight of dial indicator (ounces)

Figure 10.87 Alignment rod sag using dial indicators.

Figure 10.88 Strategies for reducing alignment rod sag. Angled supports (upper left),
counterweights (lower left), rods of variable dimension (right).

258
Figure 10.89 Alignment in long shafts requires special attention to detail over the entire length of
the shaft.

Thermal growth

The heat of most operations causes the metal in the machine to expand. Vertical growth is
typically the most important as it is measured against a stationary foundation. Thermal growth
may be minimal but should always be measured. Laser based equipment, micrometers, and
alignment bars are all subject to thermal growth. Typically it is best to take readings when the
machine is cold and again after normal operating temperature has been reached.

Calculating machine moves is handled automatically by most computer based alignment


systems. Manual methods typically involve first plotting the shaft center lines from the movable
and fixed machine. The difference between the center lines on the graph paper is the distance
that the movable machine must be moved. It is advisable to always check in the vertical and
horizontal directions.

259
Vibration Caused by Misalignment

The vibration caused by misalignment may result in excessive radial loads on bearings and
premature bearing failure. High 1X vibration with high harmonics up to the 6th harmonic are
common in the frequency spectra. One of these frequency spectra characteristics may be seen
without the other. Misalignment caused vibration is often mistaken as unbalance, looseness or
excessive clearance. High horizontal to vertical vibration amplitude ratios (greater than 3:1) may
also indicate misalignment.

Figure 10.90 Typical time waveform of misalignment. 180 of phase shift between 1X and 2X
components (left), 90 of phase shift between 1X and 2X components (right).

Figure 10.91 Misalignment time waveform and frequency spectra. Radial (left) and axial (right).

Some other operating speed related machinery faults include eccentricity of rotating parts. This
fault results in vibration at 1X (due to the unbalance that is created by the non-symmetry).
Distortion of a machine casing (due to soft foot or some other cause) results in internal preload
on the bearings. The vibrations signal typically contains 1X and higher-order vibrations.

10.1.5 Distinguishing between Unbalance and Misalignment

Unbalance and misalignment are perhaps the two most common causes of excessive machinery
vibrations. They also have similar characteristic vibration indicators. Table 10.3 summarizes the
similarities and difference in order to help distinguish between them.

260
Table 10.3 Characteristics that Can Help Distinguish between Unbalance and Misalignment.

Unbalance Misalignment
High 1X response in frequency spectra. High harmonics of 1X relative to 1X.
Low axial vibration levels. High axial vibration levels.
Measurements at different locations are in Measurements at different locations are 180
phase. out of phase.
Vibration levels are independent of Vibration levels are dependent on temperature
temperature. (change during warm-up).
Vibration level at 1X increases with rotational Vibration level does not change with rotational
speed. speed.
Centrifugal force increases as the square of the Forces due to misalignment remain relatively
shaft rotational speed. constant with changes in shaft rotational speed.

10.2.6 Mechanical Looseness

While there are many ways in which mechanical looseness may appear, there are two main
types: (1) a bearing loose on a shaft and (2) a bearing loose in a housing. A bearing that is loose
on a shaft will display a modulated time signal with many harmonics. The time period of
modulation will vary and the time signal will also be truncated (clipped). A bearing that is loose
in its housing will display a strong fourth harmonic, which can sometimes be mistaken for the
blade-pass frequency on a four-blade fan. These faults may also look like rolling-element-
bearing characteristic defect frequencies, but always contain a significant amount of wideband
noise.

Another way to diagnose mechanical looseness is by tracking the changes in the vibrations signal
as the condition worsens. In the early stages, mechanical looseness generates a strong 1X
response in the frequency spectrum along with some harmonics. At this stage, the condition
could be mistaken for unbalance. As the looseness worsens, the amplitude of the harmonics will
increase relative to the 1X response (which may actually decrease). The overall RMS value of
the time waveform may also decrease. Further deterioration of the condition results in fractional
harmonics increasing in amplitude. These harmonics are most visible in signals taken when the
machine is only lightly loaded. These harmonics show up because of the clipping described
above.

Figure 10.92 Mechanical looseness impacts (twice per revolution left) and frequency spectrum
(right) showing fractional harmonics.

261
Mechanical looseness signals often have phase that is unstable and may vary from one
measurement to the next, particularly if the rotor shifts position on the shaft from one start-up to
the next. This will result in a frequency spectrum that contains many harmonics. This type of
response can be caused by a loose bearing liner, excessive bearing clearance or a loose impeller
on a shaft.

Figure 10.93 Sample frequency spectrum representing mechanical looseness.

Figure 10.94 Sample time waveform and frequency spectrum representing mechanical looseness
(fan vibration due to excessive bearing clearance).

10.2.7 Rubs

Rubs are caused by excessive mechanical looseness or oil whirl. The result is that moving parts
come into contact with stationary parts. The vibration signal generated may be similar to that of
looseness, but is usually clouded with high levels of wideband noise. This noise is due to the
impacts. If the impacts are repetitive, such as occurring each time a fan blade passes, there may
be strong spectral responses at the striking frequency. The impacts may excite natural

262
frequencies and sub-harmonic frequencies may be present in the frequency spectrum depending
on the location of the rotor natural frequencies.

In many cases, rubs are the result of a rotor pressing too hard against a seal. In these cases, the
rotor may heat up asymmetrically and develop a bowed shape. Subsequently, a vibration signal
will be generated that shows unbalance. To diagnose this condition, it will be noted that the
unbalance is absent until the machine comes up to normal operating temperature.

Figure 10.95 Rub time waveform (often truncated) and frequency spectrum (with sub-, and
fractional-harmonics, and an excited resonance).

10.2.8 Oil Whirl and Oil Whip

Oil whirl occurs when the fluid in a lightly loaded journal bearing does not exert a constant force
on the shaft that is being supported and a stable operating position is not maintained. In most
journal bearing designs, this situation is prevented by using pressure dams or tilt pads to insure
that the shaft rides on an oil pressure gradient that is sufficient to support it. During oil whirl, the
shaft pushes a wedge of oil in front of itself and the shaft then migrates in a circular fashion
within the bearing clearance at just less than one half the shaft rotational speed. The rotor centre
is actually moving inside the bearing in the opposite direction from shaft rotation.

Because of the inherent instability of oil whirl, in many situations where oil whirl occurs, the
time waveform will show intermittent whirl events. The shaft makes a few revolutions while
whirl is present and then a few revolutions where the whirl is not present. This beating effect is
often evident in the time waveform and can be used as a diagnostic indicator.

Persistent oil whirl usually requires a replacement of the bearing. However, temporary measures
to mitigate the detrimental effects include changing the oil viscosity (changing the operating
temperature or the oil), running the machine in a more heavily loaded manner, or introducing a
misalignment that will load the bearing asymmetrically. This last course of action is of course
not recommended for more than relatively short-term relief.

263
Figure 10.96 Hydrodynamic bearing pressure profile.

Figure 10.97 Frequency spectrum representing oil whirl.

Figure 10.98 Some typical journal bearing pressure dam configurations that help prevent oil
whirl.

Oil whip occurs when a sub-synchronous instability (oil whirl) excites a critical speed
(resonance), which then remains at a constant frequency regardless of speed changes. Oil whip
often occurs at two times the critical speed because, at that speed, oil whirl matches the critical

264
speed. Figure 10.99 shows a waterfall (cascade) plot of a mass unbalance that excites oil whirl
and oil whip. Note how the oil whip locks on to the critical speed resonance.
Oil Whirl Oil Whip
Mass Unbalance

Critical Speed (system natural Frequency


frequency is excited)

Figure 10.99 Waterfall (cascade) plot of a mass unbalance that excites oil whirl and oil whip.

Figure 10.100 Waterfall (cascade) plot of a natural frequency that excites oil whip.

10.2.9 Beating and Amplitude Modulation

A beat frequency (or beating) is the result of two closely spaced frequencies going into and out
of phase as a function of time. The wideband spectrum will show one peak pulsating up and
down as a function of time.

Zooming in to this peak will show that there are actually two closely spaced individual peaks.
The difference between the peaks is the beat frequency.

265
Figure 10.101 Beating from two adjacent pumps operating at different speeds

Figure 10.102 Another example of beating. Frequency spectrum (top) and time waveform
(bottom).

Amplitude modulation is similar to beating. The frequency of the waveform seems to be


constant, but the amplitude is fluctuating up and down at a constant rate. 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. The sidebands
are spaced away from the carrier at the frequency of the modulating signal.

266
Figure 10.103 Amplitude modulation. 5000 Hz carrier and 200 Hz modulator. Time waveform
(left) and frequency spectrum (right).

In Figure 10.103 the modulating frequency is much lower than the modulated or carrier
frequency, but the two frequencies are often close together in practical situations (see next
example). Also these frequencies are sine waves, but in practice, both the modulated and
modulating signals are often complex. A vibration and acoustic signature similar to this is
frequently produced by electric motors with rotor bar problems.

Figure 10.104 Amplitude modulation. 200 Hz carrier and 190 Hz modulator. Time waveform
(left) and frequency spectrum (right).

Beating versus Amplitude Modulation

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. The spectrum in the case of beating shows the frequency and amplitude
of each component, and there should be no sidebands present.

A signal that looks like amplitude modulation, but is actually just two sine wave signals (of
slightly different frequency) added together to form beats will have relative phase variation
between the two signals from zero to 360 degrees. This means the combined amplitude varies
due to reinforcement and partial cancellation. In this case, the amplitudes of the two beating
signals are different, causing incomplete cancellation at the null points between the maxima (see
Figure 10.105). Beating is a linear process and no additional frequency components are created.

267
Figure 10.105 Amplitude modulation (left) and beating (right) time waveforms. Note the phase
shift (null points) between the maxima in the beating signal.

10.2.10 Structural Vibrations

Structural vibrations can range dramatically in amplitude and frequency. Large amplitude, low
frequency vibrations can be excited in multistory buildings during an earthquake or by the wind.
These vibrations are sometimes the result of a building resonance being excited. While these
sources of structural vibration are important, the source that we are concerned with here is that of
machinery operating as part of a buildings utility system, as part of the production plant, or
construction equipment close-by. Fans, blowers, compressors, piping systems, elevators, and
other building service machines all produce vibrations in a building and, if they are not properly
isolated they can cause disruption and/or damage to other machines or processes operating close-
by. The same is true of heavy machinery operating within a plant (stamping machines, presses,
forges, etc.) and construction equipment. High-impact and repetitive vibrations can excite
resonances large distances from the source of the excitation.

10.2.11 Foundation Problems

Machine foundations provide rigidity and inertia so that the machine stays in alignment. The
energy generated by a machine in the form of vibrations is transmitted, reflected, or absorbed by
the foundation. Especially on larger machines, the foundation is paramount to successful
dynamic behavior. Maximum energy is transmitted through the foundation to the earth when the
mechanical impedance of the foundation is well matched to that of the source of vibration. That
is, the source of vibration and the foundation should have the same natural frequency. If this is
the case, all frequencies of vibration below the natural frequency will be transmitted by the
foundation to earth. A poor match will mean that more energy is reflected or absorbed by the
foundation, which could effect the operation of the machine attached. Changing foundations can
grossly affect amplitude and phase measurements, which means that vibration measurements can
be used to detect a changing foundation or hold-down system.

268
10.2.12 Summary of Identification and Correction of Forcing Function Faults

Fault Frequency Spectrum, Time Waveform, Orbit Correction


Shape
critical speed 1X, 2X, 3X, amplified vibration due to proximity tune natural frequency
etc. of operating speed to natural
frequency
mass 1X distinct 1X with much lower values field or shop balancing
unbalanced of 2X, 3X, etc.; elliptical and
circular orbits; constant phase
misalignment 1X, 2X, distinct 1X with equal or higher perform hot and/or cold
occasionally values of 2X, 3X; 1X axial alignment
3X
shaft bow 1X dropout of vibration around critical heat or peening to straighten
speed in Bod plot rotor (allow rotor to float
axially)
fluid film 1X, high 1X, high 1/2X, sometimes 1-1/2 replace bearing
bearing wear subharmonics, or orders; cannot be balanced
and orders
excessive
clearance
resonance 1X, 2X, 3X, high balance sensitivity high change structural natural
etc. amplitude vibration at order of frequency
operating speed
looseness 1X plus large high 1X with lower-level orders, shim and tighten bolts to
number of large 1/2 order, low axial vibration obtain rigidity
orders, 1/2X
may show up
eccentricity 1X high 1X machine journal for
concentricity
thermal 1X 1X has varying phases angles and compromise balance or
variability amplitude with load remove problem
distortion 1X and orders 1X from preload of bearings, 2X line relieve soft foot
frequency, air gap on motor

269
10.3 Fault Diagnosis based on Specific Machine Components

10.3.1 Rolling Element Bearings

Rolling element bearings produce very little vibration (low level random signal) when they are
fault free, and have very distinctive characteristic defect frequency responses when faults
develop. This, and the fact that most damage in rolling element bearings occurs and worsens
gradually, makes fault detection and diagnosis on this component relatively straightforward.

Faults occur in rolling element bearings for many reasons. Table 10.4 lists some of the more
common causes of rolling element bearing failure. Faults due to normal use usually begin as a
single defect caused by metal fatigue in one of the raceways or on a rolling element. The
vibration signature of a damaged bearing is dominated by impulsive events at the ball or roller
passing frequency. As the damage worsens, there is a gradual increase in the vibration response
at the characteristic defect frequencies followed by a drop in these amplitudes and an increase in
the broadband noise. In machines where there is little other vibration that would contaminate or
mask the bearing vibration signal, the gradual deterioration of rolling-element bearings can be
monitored by using the crest factor or the kurtosis measure (see previous chapters for
definitions).

Table 10.4 Common causes of rolling element bearing failure.

Causes of Rolling Element Bearing Failure


Wrong Lubrication (includes lack of lubrication) 43%
Improper Mounting 27%
Change in Internal Clearance
External Vibration 21%
Parasitic Loads (due to misalignment, unbalance, etc.)
Contamination and Hostile Environment
Run to L10 Life 9%

A key factor in being able to accurately detect and diagnose rolling element bearing defects is the
placement of the vibration sensor. Because of the relatively high frequencies involved,
accelerometers should be used and placed on the bearing housing as close as possible to, or
within, the load zone of the stationary outer race.

Specific applications can also pose significant challenges to fault diagnosis. Low speed machines
have bearings that generate low energy signals and require special processing to extract useful
bearing condition indications. Machines that operate at varying speeds also pose a problem
because the characteristic defect frequencies are continuously changing. Bearings located close
to, or within, gearboxes are also difficult to monitor because the high energy at the gear meshing
frequencies masks the bearing defect frequencies.

Factors affecting rolling element bearing life include lubrication (type, quality, quantity and level
of contamination), temperature, alignment (misalignment), load (static and dynamic), speed,

270
installation and adjustment of the machine, improper fit (critical press fit, size), vibration from
external sources, and the passage of electrical current through the bearing.

The most common bearing failure mode is spalling. Microscopic fissures develop on a raceway
and/or the rolling elements. These small fissures gradually worsen (become larger) and fatigue
eventually causes pits to form on the surface. Another cause of failure is Brinelling. This results
from excessive static stress. The raceway and/or rollers become deformed at the roller/raceway
contact (which is typically very small in area) False Brinelling is caused by vibration damage
while the machine is not rotating. Again the small contact area between the rollers and the
raceway focus the vibration forces and cause small marks (and/or local areas of work hardening)
in the load carrying surface. Localized faults in the raceway, deformed raceways, damaged cage
and cracked raceway are also causes of rolling element bearing failure.

Figure 10.106 Rolling element bearing dimensional definitions.

Outer Race

Inner Race

Rollers

Cage

Figure 10.107 Rolling element bearing components.

271
Figure 10.108 Rolling element bearing configurations. (a) Ball, (b) Cylindrical, (c) Thrust, (d)
Tapered, (e) Double row.

Figure 10.109 Nomenclature of rolling element bearings.

Figure 10.110 Types of rolling element (ball) bearings.

272
Figure 10.111 Types of rolling element (cylindrical) bearings.

Figure 10.112 Small pits (spalls) on rollers (left) and raceways (right) are the earliest signs of
rolling element bearing failure.

Figure 10.113 Large pits on raceways (left) and rollers (right) are advanced signs of rolling
element bearing failure.

273
Figure 10.114 The progression of bearing failures typically follows the sequence shown here.
Upper left strains on the outer race eventually generate subsurface fatigue cracks. Lower left
These cracks grow and eventually reach the surface of the raceway. Right When the cracks
become large enough to surround small amounts of raceway metal, that piece of metal breaks
away (spalls off) and a pit is formed.

Figure 10.115 Small pits result in distinct impacts at the ball pass frequency as the rolling
elements move past these depressions.

Symptoms of extreme bearing wear include extreme vibration, excessive heat, noise, seizure
between inner and outer race, and finally catastrophic failure. Early detection of a rolling element
fault is possible through vibration monitoring and this can avert a catastrophic failure. While
vibration is the most common monitoring method, there are others. These include ultrasonic
monitoring, acoustic emissions (AE) monitoring, temperature monitoring (using infrared and/or
thermocouple based sensors).

Rolling element defects generate pulse-like forces at one or a combination of bearing


frequencies. The magnitude of vibration depends on machine design, load and defect severity.
The bearing frequencies are unique to bearing geometry and operating speed. Bearing

274
frequencies are usually below 1,000 Hz. Velocity is typically the better choice of measurement
parameter, but acceleration is also commonly used.

As bearing damage worsens there is typically a gradual increase in the magnitude of the
frequency response at the characteristic defect frequencies followed by a drop in these
amplitudes and an increase in the broadband noise. This is due to the fact that as the rollers
continually pass over the small original defects, they tend to flatten these defects and the impacts
are not as distinct. At the same time the other fatigue cracks in the raceway are gradually
generating more pits. Eventually these pits may join together and the distinct impacts will
disappear from the vibration signal to be replaced by a broadband noise as the rollers move along
a rough pit-filled raceway.

Vibration measurement can be made with velocity transducers or accelerometers. All


measurements should be made in the load zone as close to the bearings as possible. Radial
measurements should be made with radial bearings. Axial measurements should be made with
angular contact bearings.

Analysis techniques include the calculating of bearing defect frequencies to predict where the
vibration frequency response will appear in the frequency spectrum. Measuring and analyzing
vibration signals includes identifying sidebands and centre frequencies in the spectrum and
evaluating the spectrum and time waveform for shape, energy, and amplitude.

Load Direction

Bearing

Figure 10.116 These gear bearings are loaded in the radial direction from the gear load and
therefore the vibration signal should be measured in that direction to ensure that the sensors are
in or close to the load zone.

Bearing frequencies are defined as the following.

BPFO ball pass frequency of the outer race (outer race defects)
BPFI ball pass frequency of the inner race (inner race defects)
BSF ball spin frequency (ball or roller defects)
FTF fundamental train frequency (cage defects or improper movements)

275
Figure 10.117 Rolling element bearing defect frequencies and the equations that define them.

Figure 10.118 Approximate rolling element bearing defect frequencies.

Rolling element bearings have characteristic time waveforms and frequency spectra that
distinguish the type of fault that is present and the severity of the fault. Figures 10.119 to 10.121
show a variety of time waveforms that show the characteristic shape of the waveform for the
different types of rolling element bearing faults. Figures 10.122 to 10.124 show several different
frequency spectra of different individual rolling element bearing faults as well as several
combinations of faults.

276
Figure 10.119 Rolling element bearing defect time waveform (outer race fault relatively
consistent impact amplitudes).

Figure 10.120 Rolling element bearing defect time waveform (inner race fault variable impact
amplitudes with shaft rotation).

Figure 10.121 Rolling element bearing defect time waveform (roller fault intermittent and
unpredictable impacts).

277
Figure 10.122 Rolling element bearing defect frequency spectra (outer race fault with harmonics
upper, outer and inner race fault middle, inner race fault with sidebands lower).

Figure 10.123 Rolling element bearing defect frequency spectra (rolling element fault with
harmonics upper, 2X rolling element fault with harmonics middle, rolling element fault with
cage fault sidebands lower).

278
Figure 10.124 Rolling element bearing defect frequency spectra (cage (fundamental train) fault
upper, cage fault with harmonics middle, cage fault sidebands around another bearing fault
lower).

Rolling element bearings also have characteristic time waveforms and frequency spectra at
various stages of deterioration. There are typically four stages of deterioration. Stage one is the
pre-failure stage. This stage is when the earliest indications of failure (or deterioration) may be
detected with Spike Energy (Shock Pulse) or enveloped vibration signal processing. This stage is
when there are still only micro-cracks or micro-spalls present in the bearing components. Only
sub-surface damage is present. The bearing will be operating at normal temperature.

Figure 10.125 Frequency spectra showing responses at stage one - bearing pre-failure.

279
Stage two is the failure stage. This stage is when the interaction of the rolling elements with
slight defects begin to ring the bearing components at their natural frequencies. These
frequencies typically occur in the range of 30-32 kHz. At the end of stage two sideband
frequencies appear above and below the natural frequencies. The Spike Energy reading
increases. The bearing is developing visible flaws such as spalling on the races and/or spalling on
the elements and/or cage damage. There may be an increase in the audible noise level and/or an
increase in the bearing operating temperature.

Figure 10.126 Frequency spectra showing responses at stage two - bearing failure.

Stage three is the pending catastrophic failure stage. This stage is when rapid and ultimate failure
of the bearing is imminent. There is a significant increase in the audible noise level and the
bearing operating temperature reaches the overheated level. Rapid wear is taking place. The
bearing clearances are increasing and rotor-to-stator rub is taking place in machines with close
tolerances. There is a significant increase in the frequency responses at the primary bearing
defect frequencies and harmonics. High frequency harmonics may also appear. The Spike
Energy reading continues to increase. The number of sidebands in the frequency spectra
increase. Wear is now clearly visible and may extend around the periphery of the bearing.

Figure 10.127 Frequency spectra showing responses at stage three pending catastrophic
bearing failure.

280
Stage four is the catastrophic failure stage. Discreet bearing defect frequencies disappear and are
replaced with random broad band vibration in the form of a noise floor. Towards the end of the
bearing life even the 1X RPM level is affected (and decreases). High frequency noise floor
amplitudes and spike Energy responses may decrease. Just prior to ultimate failure the Spike
Energy level may increase again to higher levels. Figures 10.125 to 10.128 are summarized in
Figure 10.129.

Figure 10.128 Frequency spectra showing responses at stage four catastrophic bearing failure.

Figure 10.129 Frequency spectra showing responses at all stages of bearing failure.

281
Because high frequency detection (HFD) methods (also known as Amplitude Demodulation or
Enveloping) are so important in rolling element bearing fault detection and trending, a review is
provided in the following figures.

Figure 10.130 Rolling element bearing time waveform showing a dominant sinusoidal signal
(low frequency), typically due to shaft unbalance, misalignment or inner race fault moving in and
out of the load zone, with frequent impacts (high frequency) due to the transient ringing of the
bearing structure in response to the impact excitations.

Figure 10.131 Time waveform showing only the impacts (post high-pass filtering) from the
bearing, which contain the most important information.

282
Figure 10.132 Frequency spectra showing the removal of the lower frequency signals with a
high pass filter. Such a filter would typically be set to approximately 2000 Hz for bearing
analysis, but the particular frequency may vary depending on the bearing size and structure.

Figure 10.133 Frequency spectra showing the removal of the lower frequency signals with a
high pass filter.

Figure 10.134 Time waveform showing the rectified, high-pass filtered vibration signal.

283
Figure 10.135 Envelope detection resampling of the high passed signal to detect and save
only the peak signal amplitudes. This resampled signal shows the lower frequency response rate
of the ball impacts (without the ringing.),

Figure 10.136a Frequency spectra showing the rescaled final result. The defect frequency
sidebands are shifted to the low frequency range and isolated from all other response frequencies
in the original signal.

Figure 10.136b Frequency spectra showing the rescaled final result.

284
Figure 10.136c Frequency spectra showing the rescaled final result (zoomed).

Figure 10.137 Demodulated frequency spectra showing the four typical stages of bearing
deterioration.

10.3.2 Gears and Gearboxes

Because gears transmit power from one rotating shaft to another, significant forces are present
within the mating teeth. While gears are designed for robustness, the teeth do deflect under load
and then rebound when unloaded. The local stresses are high at the tooth interface and root,
which leads to fatigue damage. Proper design and perfect fabrication of gears (with perfect form
and no defects) would result in relatively low vibration levels and a long life. However, the
presence of non-perfect gears gives rise to excessive vibration.

In relatively simple gearboxes, the time waveform can be used to distinguish impacts due to
cracked, chipped, or missing teeth. The frequency spectra and cepstra are powerful tools when

285
the gearbox contains several sets of mating gears, which is most often the case. Even a
significant defect on one tooth (or even a missing tooth) often does not produce an abnormally
strong frequency spectral response at 1X. However, the defect will modulate the gear mesh
frequency (number of teeth times the shaft rotational speed) and appear as 1X sidebands of the
gear mesh frequency. That is, smaller spectral responses that appear at a distance of 1X (and
multiples of 1X for more severe gear faults) above and below the gear mesh frequency. Because
these sidebands occur at multiples of 1X and a spectral plot can become quite cluttered with
response lines, cepstral analysis is well suited to distinguish the frequency components that are
strong fault indicators.

Often, a change in the response at two times the gear mesh frequency is a good indicator of
developing gear problems. The amplitude of the gear mesh frequency, and its multiples, vary
with load. This makes it important to sample the vibration signal at the same load conditions.
When unloaded and under variable loads excessive gear backlash may also cause an increase in
the amplitude of the gear mesh frequency.

Because each gear tooth meshes with an impact, structural resonances may be excited in the
gears, shafts, and housing. Proper design of a gearbox will minimize this effect, but resonances
in gearboxes may cause accelerated gear wear and should be monitored.

Gears provide an excellent example of how machines must wear-in during early use. New gears
will have defects that are quickly worn away in the machines early life. Vibration levels will
become steady and only increase gradually later in the machines life as the gears wear out. These
gradual increases in vibration level are normal. Sudden changes in vibration levels (at gear mesh
frequency, two times gear mesh frequency, or sidebands), especially decreases, are significant. A
drop in the vibration level usually means a decrease in stiffness, and that more of the
transmission forces are being absorbed due to bending of the gear teeth. Catastrophic failure is
imminent. Premature gear failures are usually a symptom of other problems such as unbalance,
misalignment, bent shaft, looseness, improper lubrication, or contaminated lubrication.

The advantages of using gears to transmit power include the high power to size ratio possible,
the high rigidity, the no slip transmission and the accuracy. While gearboxes can be costly they
do provide a relatively low cost alternative for the amount of torque that may be transmitted.
Gearboxes can also run at high speeds. The disadvantages include the fact that they require
lubrication, precise alignment and can be quite noisy.

286
Figure 10.138 Assembly and parts of a planetary gear set.

Figure 10.139 Schematic of a planetary gear set.

287
Figure 10.140 Cross section of a planetary gear set.

Figure 10.141 Parallel shaft arrangement gear set.

288
Figure 10.142 Gear teeth and the associated terminology.

There are many different types of gears. Spur gears are used in parallel shaft boxes. They have
straight teeth parallel to gear shaft axis and contact along the entire length of each tooth. They
typically function with two pairs of teeth in contact for 1/2 the time and one pair of teeth in
contact for 1/2 the time. The maximum stress capability (maximum torque load) is limited by the
capability of the individual teeth. Unwanted variation in tooth profile (due to poor design,
manufacturing methods, deflection, etc.) occurs across the entire tooth which may cause high
tooth mesh frequency vibration.

Figure 10.143 Spur gears.

Helical gears are cylindrical gears with spiral (helical) teeth. The teeth are cut parallel and
produce a line of contact that is a slanting line. The contact starts at one end of the tooth and goes
to the other. These gears are relatively smooth running due to the averaging effect on the tooth
profile errors due to the slanting contact line. These gears have a higher stress capacity (and can
therefore carry higher loads) than spur gears. There is however, an axial force developed due to
slanting line of contact, which may cause high axial vibration. These gears typically have
relatively low radial vibration. Double helix (herring bone) are used to cancel the axial thrusts.

289
Figure 10.144 Helical gear.

Worm gears have screw thread shaped teeth. There is lots of sliding wear because of the large
contact angle and the large amount of sliding between teeth. The gear shafts are non-intersecting,
but typically perpendicular. These gears usually have high gear ratios.

Figure 10.145 Worm gear.

Bevel gears are conical in shape and have intersecting shaft axes. They have straight, axially
aligned teeth and typically have a low reduction ratio when used in perpendicular shaft drives.
Spiral bevel gears are also available. These are equivalent to helical gears.

Figure 10.146 Spiral gear.

Gear faults and defects include pitting, scuffing, tooth breakage, tooth damage, cracking and
general wear. However, the most common failure mode in a gearbox is associated with bearing
failure. Other gearbox problems include misalignment, eccentricity, and associated
manufacturing defects.

290
Figure 10.147 Spur gear with a missing tooth.

Figure 10.148 Spur gear with a chipped tooth.

Figure 10.149 Spiral gear with chipped and broken teeth.

Figure 10.150 Spur gear with local tooth wear (left) caused by local high spots on the gear
wheel (right) at the spoke locations.

291
Figure 10.151 Uneven tooth wear can be caused by gear wheel misalignment.

Figure 10.152 Visual wear patterns due to different conditions.

Gear defects may be caused by age, overloading of the gearbox, lack of proper lubrication and
contamination of lubricants, Many gearbox problems can be mitigated using established methods
such as oil particle analysis, proactive lubricant changes and other preventive methods. Material
and manufacturing defects can also lead to premature gear failures.

Oil analysis may take different forms, such as ferrographic analysis. Different types of faults
result in specific types of particles, the size and/or shape (as revealed in ferrographic analysis
to be discussed in a later chapter) can reveal the fault type.

Vibration analysis typically involves analysis of the vibrations collected perpendicular to the
shaft (radial vibrations) or parallel to the shaft (axial vibrations). There are the gearbox casing
vibrations. However, there is also the possibility of sampling the torsional vibration of the shaft
for clues to the type of fault that is present. Torsional vibrations is the oscillation of the shaft
relative to the casing at the input and/or the output shafts. This measure can be a better detection
and diagnostic tool in some cases. There are instrumentation difficulties with measuring torsional
vibration. A bump (impact) test may also be used to detect changes in natural frequencies that
have occurred due to mechanical changes in gear shape or gearbox condition.

292
The first step in conducting a gearbox vibration analysis is to obtain a drawing or sketch of the
gearbox and the internal details of the gears. This will allow for the calculation of the gear-
meshing frequencies based on the number of teeth on each gear and the shaft speed.

Table 10.5 Sample gearbox information that is important for analysis of vibration signals.

Characteristic Frequency
Input Pinion Gear Mesh Frequency 269.2 Hz
Planet Gear Mesh Frequency 19 Hz
Sun Gear Mesh Frequency 38.8 Hz
Planet Passing Frequency 1.5 Hz
Swing Pinion Gear Mesh 6.3 Hz

Figure 10.153 shows an example schematic diagram of a double-reduction gearbox including the
input, output and intermediate shaft speeds as well as the number of teeth on each gear. This
information is used to calculate the gear mesh frequencies (as shown below).

GM 1553.5 Hz
3585 RPM 26T

31T 923 RPM

101T
97T

GM 476.8 Hz 295 RPM

Figure 10.153 Schematic of a gearbox showing the number of teeth on each gear and the shaft
speeds.

Gear meshing frequency calculations.

Given input shaft speed = 3,585 RPM


Intermediate shaft speed = (3,585 RPM) [(26 T)/101 T] = 923 RPM
Output shaft speed = 923 x 31T/97T=295 RPM

293
High-speed gear mesh = 3,585 RPM x 26T = 93,210 CPM (1,553.5 Hz)
Low-speed gear mesh = 922.87 RPM x 31 T = 28,609 CPM (476.8 Hz)

The frequency range for measurement and analysis can then be determined from these numbers.
An accelerometer should be chosen that is suitable for the expected gear mesh frequency. The
measurement direction should be radial for spur gears and axial for gears that are loaded in axial
direction. Measurement locations should be as close to the gears of interest as possible. As
always the transmission path of the vibration signal needs to be considered. Most vibration
measurements are typically taken near bearing housing. And close to or in the load zone.

Because gears typically generate complex vibrations, there may be some difficulties when
interpreting gear vibrations. To further complicate matters, it is often difficult to measure
vibration close to the gear of interest. Poor signal to noise ratios are common due to other
vibration sources nearby creating noise. These vibration sources may be other gears (in perfectly
good condition, but still generating significant vibration levels), bearings, adjacent machines, and
couplings. Dramatic transient vibrations are sometimes the result of wear debris passing through
gears. It should be noted that an acoustically noisy gearbox may not indicate a faulty gearbox.
Acoustic noise increases when transmission error increases, frequency of operation increases,
tooth load increases and/or the number of gears increases.

The two factors (from the list above) that influencing gear vibrations the most (apart from gear
faults) are shaft speed and loading. Shaft speed influences amplitude and frequency at which the
responses are seen in the frequency spectrum. Loading influences the vibration signal amplitude.
Gears must be minimally loaded to transmit vibration. Backlash conditions develop when there is
not enough load to keep the teeth in contact with one another. The ideal condition for vibration
measurement is steady load and steady speed.

Gear vibration analysis may take place in the time domain, the frequency domain or the time-
frequency domain. The principal vibration frequencies are as follows.

gear shaft bearing characteristic defect frequencies (and harmonics)


rotational speed and harmonics (for all gear shafts)
gear mesh frequency (number of teeth times shaft rotational frequency)
harmonics of gear mesh
sidebands of gear mesh or harmonics (primary frequency + or - shaft speeds)
wobble of the gear (disk resonance)
tooth / shaft resonance
bearing deflections due to loading on teeth

Time domain analysis is good for detecting individual tooth faults which may result in peaks in
the time waveform at regular intervals (the location of the defect). Often searching for these
signals in the time domain is easier than looking for the associated changes in the frequency

294
spectrum because the other gear interactions will overshadow the vibration due to the individual
tooth fault.

Time domain averaging is also a good strategy for gear vibration signal analysis because it
allows defects that occur once per shaft revolution to be accentuated. It is also an efficient data
reduction method that reduces random noise. It is suitable for periodic and repetitive signals, but
a trigger is necessary to mark the start of each segment prior to averaging.

Table 10.6 Gearbox faults and symptoms

Fault Frequency Spectrum or Time Waveform


eccentric gears gear mesh gear mesh with sidebands at frequency of
eccentric gear
gear-mesh wear gear mesh gear mesh with sidebands at frequency or worn,
scored, or pitted gear(s); sometimes , 1/3,
harmonics of gear mesh
improper backlash gear mesh gear mesh with orders and sidebands at
of end float frequency of pinion or gear
broken, cracked, or natural pulses in time waveform; natural frequencies in
chipped gear teeth frequencies spectrum
gearbox distortion gear mesh gear mesh and orders in spectrum; varying gear-
and/or natural mesh amplitude in time waveform shaft
frequencies frequency plus low-amplitude orders

Gear mesh problems are typically attributed to uneven wear, improper backlash, scoring, and
eccentricity. The characteristics in the spectrum are the appearance of gear mesh frequency
response with sidebands at the frequency of the speed of the faulty shaft. Badly worn gears will
show multiples of gear mesh frequency with sidebands.

Figure 10.154 Typical Gear Spectrum.

A normal gear box (no faults or deterioration) will generate a frequency spectrum that shows
strong 1X and 2X and gear mesh frequencies. The gear mesh frequency will commonly have

295
running speed sidebands. That is the sidebands around the gear mesh frequency will be a
distance away from the gear mesh frequency equal to the shaft rotational speed (frequency). All
peaks in the spectrum will be at relatively low amplitude and no natural frequency responses will
be present. The amplitude of responses at harmonic multiples of the shaft rotating frequencies
are generally less than 1% of the amplitude of the fundamental frequency.

Figure 10.155 Time waveform and spectrum for a normal gearbox condition.

Figure 10.156 Time waveform and spectrum for a normal gearbox condition (peaks are
symmetrical, paired and equal).

296
Figure 10.157 Frequency spectra for a normal gearbox condition zoomed in on the gear mesh
frequency and sidebands separated by 1X (sideband peaks are symmetrical, paired and equal).

Figure 10.158 Frequency spectra for a gearbox experiencing changing load conditions (top
normal or low load, bottom increased load (not necessarily indicative of a developing fault)).

Figure 10.159 Frequency spectra for a gearbox experiencing changing load conditions.

297
Figure 10.160 Gear tooth wear.

Gear tooth wear is indicated by excitation of natural frequencies along with sidebands of 1X
shaft rotational frequency (the shaft containing the bad gear). Sidebands may be a better wear
indicator than gear mesh frequency as gear mesh frequency amplitudes may not change when
wear occurs. High amplitude sidebands surrounding the gear mesh frequency usually become
obvious as the wear worsens.

Figure 10.161 Frequency spectrum showing gear tooth wear.

Figure 10.162 Frequency spectrum showing gear tooth wear.

298
Figure 10.163 Frequency spectra showing gear tooth wear or excessive clearance changes
(unsymmetrical sideband spacing and amplitudes).

Figure 10.164 Frequency spectra showing gear eccentricity and backlash.

Fairly high amplitude sidebands around the gear mesh frequency suggest eccentricity, backlash
or non-parallel shafts. The gear with the problem is indicated by the spacing of the sideband
frequencies. Improper backlash normally excites gear mesh frequencies and gear natural
frequencies, both of which will have sidebands at 1X RPM. Gear mesh frequency amplitudes
will often decrease with increasing load if backlash is the problem.

Figure 10.165 Frequency spectra showing gear eccentricity and backlash.

299
Figure 10.166 Frequency spectra showing gear misalignment.

Gear misalignment almost always excites second order and higher gear mesh frequency
harmonics, which have sidebands at 1X running speed. Often the primary gear mesh frequency
will show only small amplitude, but the second and third harmonics will have mesh higher
levels. It is important to sample the vibration signal fast enough to capture all the high frequency
gear mesh frequency harmonic responses.

Figure 10.167 Frequency spectra showing gear misalignment.

300
Figure 10.168 Time waveform and frequency spectra showing a cracked or broken gear tooth.

A cracked or broken gear tooth will generate a high amplitude response at 1X running speed of
that gear. Plus it will excite the gear natural frequency with sidebands at the running speed.
These faults are best detected using the time waveform. The time interval between the extra large
impacts, indicating the missing or weakened tooth, will be equal to one per revolution of the
shaft with the faulty gear. The impact spikes (in the time waveform) will be much higher than
those of the other gear mesh impacts.

Figure 10.169 Frequency spectra showing a cracked or broken gear tooth.

Figure 10.170 Frequency spectra (zoomed in to the gear mesh frequency) showing a cracked or
broken gear tooth (asymmetrical sideband profile).

301
Figure 10.171 Frequency spectra showing gear looseness.

Some background on gears suggests that adjacent teeth on the same gear should share the same
normal to common tangent (see Figure 10.172). A single line is normal to the common tangent at
two adjacent contact points and this line passes through the pitch point. All pitch points are in the
centre of teeth. When joined they form the pitch circle.

Gear tooth shape is defined as an involute curve. This is the curve that is traced by the end of a
tight string as it is unwound from the circumference of a circle. Small errors in centre to centre
distance do not violate meshing action. Low noise and vibration levels can be expected from
gearboxes that have been well designed and manufactured.

Figure 10.172 Schematics of gears showing defining details relating to gear tooth shape.

Some definitions:

Pitch circle diameter: diameter of pitch circle


Diametrical pitch: number of gear teeth divided by the pitch circle diameter.
Circular pitch: distance between teeth on the circumference of the pitch circle
Normal pitch: distance along the normal to the common tangent between successive
tooth surfaces
Base circle diameter: diameter of circle from which involute curve is generated.

302
Gear ratio: ratio of # of teeth on each gear.

# teeth driven
gear ratio =
# teeth driver

Lead: distance of travel axially along a helical gear for one tooth to rotate through 360.
Line of action: distance along the normal to the common tangent during which one tooth
is in contact with one tooth on the other gear
Backlash: the clearance between the adjacent teeth when two teeth are in contact
Working depth: radial distance from the point of tip contact on one tooth to the tip of the
contacted tooth
Contact ratio: average number of pairs of teeth that are theoretically in contact
Addendum: difference between the pitch circle radius and the radius of the outside
diameter
Dedendum: difference between the radius of the root circle and the radius of the pitch
circle

Gearboxes typically generate high frequency vibrations as a result of the gear meshing function
of the gear. The greater the number of gear teeth the smoother is the performance of the box.
Gear mesh frequencies with sidebands at operating speeds identify wear and gearbox distortion.
Gear mesh problems are attributed to uneven wear, scoring and eccentricity. Both axial and
radial measurements can be used.

10.3.3 Belts

Belt defect frequencies are typically below the primary rotational speed of both shafts (driver and
driven). When belts are worn, loose or mismatched, they normally cause 3 or 4 multiples of belt
frequency. Often the 2X belt frequency is dominant. Amplitudes are typically unsteady and
sometimes pulse with either the driver or the driven rotational speed. On timing belt drives, wear
or pulley misalignment is indicated by high amplitudes at the timing belt frequency.

Figure 10.173 Frequency spectrum showing belt frequency and harmonics.

303
Misalignment of sheaves produces high vibration at 1X RPM, predominantly in the axial
direction. The relative amplitudes at the driver and driven RPM depends on where the data is
taken as well as on relative mass and frame stiffness. In cases involving large fans with sheave
misalignment, the highest axial vibration will be at the fan (driven) RPM.

Figure 10.174 Sheave misalignment and frequency spectrum.

Eccentric sheaves and/or unbalanced sheaves cause high vibration at 1X RPM of the sheave
shaft. The amplitude is normally highest in line with the belts, and should show up on vibration
measurements taken at both driver and driven bearings. It is sometimes possible to balance
eccentric sheaves by attaching washers to taper lock bolts. However, even if balanced, the
eccentricity will still induce vibration and irreversible fatigue stresses in the belt.

Figure 10.175 Eccentric sheave and frequency spectrum.

Belt resonance can cause high amplitude vibration if the belt natural frequency should happen to
approach or coincide with either the motor or the driven machine RPM. The amplitude is
typically highest in line with the belts and should show up in vibration signals measured on both
driver and driven bearings (see Figure 10.176).

Belt natural frequencies may develop like any other natural frequency. If an excitation occurs
close to the natural frequency, then that resonance will be excited. Like other structures, belts
have more than one natural frequency and can develop different mode shapes regardless of the
fact that they are constantly in motion (see Figure 10.177).

304
Figure 10.176 Belt natural frequency and frequency spectrum.

Figure 10.177 Belt natural frequency mode shapes.

305
10.4 Fault Diagnosis based on Specific Machine Types

10.4.1 Electric Motors

Electric motors can be divided into two groups: (1) induction motors and (2) synchronous
motors. A full description will not be given here as to the differences. Like any machine, electric
motors are subject to a full range of mechanical problems, and vibrations signals can be used to
detect and diagnose these problems. Apart from the conditions described elsewhere in this
section, there are some problems that occur only in electric motors.

Figure 10.178 Standard electric motor - labeled.

Figure 10.179 Standard electric motor - labeled.

306
Figure 10.180 Standard electric motor rotor with end ring and bearings.

Figure 10.181 Standard electric motor in use driving water pumps.

Figure 10.182 Cut-away diagram of a 2000 HP induction motor

307
Figure 10.183 Cut-away diagram of an electric motor endbell assembly.

To generate electricity it is necessary to move a coil of wire across magnetic lines of flux. As the
lines of the coil move perpendicular to the lines of magnetic flux, electricity is generated in the
coil.

Figure 10.184 Schematic of the electricity generation.

Figure 10.185 Schematic of the electricity generation one full rotation of rotor.

308
To use electricity to move a coil of wire it is required that first a moving magnetic field is
created. This is done by providing electricity to a series of coils rapped as shown in Figure
10.186. This is known as a stator the stationary part of the motor that generates the moving
magnetic field. Electricity is switched sequentially between the stator coils to effectively create a
series of magnetic fields (each stationary). The effect of these fields being generated in series is
that a magnet mounted on a rotor inside the stator will react to these magnetic fields and generate
torque on the rotor shaft.

Figure 10.186 Coils that generate magnetic fields sequentially around the stator.

Figure 10.187 Sequential magnetic fields generated by a 3 phase motor.

Figure 10.188 Stator that holds windings that create the moving magnetic field.

309
Figure 10.189 Stator windings insulated copper wire rapped in coils.

The magnetic material used for the stator core is by nature also a conductor. Voltage induced in
the core steel causes a current to flow in the core. This eddy current flow in the core raises stator
temperature and lowers motor efficiency. A laminated core offers high resistance to current flow;
hence, the eddy currents and resulting losses are reduced. The same strategy is used in high
voltage transformer cores

The rotor is the rotational element of the electric motor. A squirrel cage rotor design is
commonly used in most induction motors. This design uses a laminated slotted core in which the
conductive material for the rotor bars is placed in the slots. The rotor bars are then shorted
together by the end rings.

Figure 10.190 Standard electric motor rotor.

Squirrel cage induction motors have very few maintenance requirements, and they have a rugged
and dependable reputation. Double wound squirrel cages have a high and a low resistance
winding. This configuration combines both high starting torque and excellent constant speed
control.

310
Figure 10.191 Schematic showing magnetic field and electric current interaction.

Figure 10.193 Simple squirrel cage rotor.

Figure 10.194 Rotor bar shapes there are many.

311
Electric motor malfunctions have both mechanical effects and electrical effects the mechanical
effects include the following, which have been discussed in previously.

Mass unbalance
Looseness
Resonance
Misalignment
Eccentricity
Bearing defects
Distortion

Electric motor malfunctions that have predominantly electrical effects are listed in Table 10.7.

Table 10.7 Electric motor malfunctions (electrical)

Fault Frequency Spectrum/Time Correction/Comment


Waveform/Orbit
Shape
air-gap variation 120 Hz 120 Hz plus center armature relieving
sidebands, beating distortion on frame; eliminating
2x with 120 Hz excessive bearing clearance and/or
any other condition that causes
rotor to be off center with stator
broken rotor bars 1x 1x and sidebands replace loose or broken rotor bars
equal to (number of
poles x slip
frequency)
eccentric rotor 1x 1x, 2x/120-Hz beats may cause air-gap variation
possible
stator flexibility 120 Hz 2x/120-Hz beats stiffen stator structure
off magnetic center 1x, 2x, 3x impacting in axial remove source of axial constraint-
direction bearing thrust, coupling
stator shorts 120 Hz and 120 Hz and replace stator
harmonics harmonics

312
Stator eccentricity, shorted laminations and loose iron are all electric motor malfunctions that
may cause unbalance type response in the vibration signal. Stator problems typically generate
high vibration levels at 2X line frequency. Stator eccentricity produces an uneven stationary air
gap between the rotor and the stator which produces directional vibrations. The differential air
gap should not exceed 5% for induction motors and 10% for synchronous motors. Soft foot and
warped bases can produce an eccentric stator. Loose iron is due to stator support weakness or
looseness. Shorted stator laminations cause uneven, localized heating which can increase
significantly with operating time.

An eccentric (variable) air gap is caused by an eccentric rotor. The eccentric rotor produces a
rotating air gap between the rotor and the stator which induces pulsating vibrations (typically
between 2X line frequency and the closest running speed harmonic). This condition often
requires a detailed analysis of the frequency spectrum by zooming in on the 2X line frequency
and running speed harmonics. Eccentric rotors generate a 2X line frequency response surrounded
by pole passing frequency sidebands as well as pole passing frequency sidebands around the
running speed. Note that the pole passing frequency is equal to the motor slip frequency times
the number of poles in the motor. The slip frequency is equal to the synchronous speed of the
motor minus the rotational speed of the motor. The synchronous speed of the motor is equal to
120 times the line frequency divided by the number of poles in the motor. Typical pole passing
frequencies range from 0.3 to 2.0 Hz.

Figure 10.195 Eccentric Rotor (Variable Air Gap) frequency spectrum and time waveform.

313
Figure 10.196 Eccentric Rotor (Variable Air Gap) frequency spectrum showing pole passing
frequency sidebands around the 1X and 2X line frequency.

Broken and cracked rotor bars, broken or cracked shorting rings, bad joints between rotor bars
and sorting rings, or shorted rotor laminations will all produce high 1X running speed vibration
with pole passing frequency sidebands. In addition, cracked rotor bars will often generate pole
passing frequency sidebands around the third, fourth and fifth running speed harmonics. Loose
rotor bars are indicated by 2X line frequency sidebands surrounding the rotor bars passing
frequency and its harmonics. The rotor bar passing frequency is equal to the number of rotor bars
times the rotational speed of the rotor. This fault may also cause high levels of 2X rotor bar
passing frequency with only small amplitude at 1X rotor bar passing frequency.

Figure 10.197 Broken or Cracked Rotor Bars possible response frequency spectra.

314
Figure 10.198 Eccentric Rotor (Variable Air Gap) causing a rub between the rotor and the stator.

Phasing problems in motors (caused by loose connectors) can cause excessive vibration at 2X
line frequency which will have sidebands around it at 1/3 line frequency. Levels at 2X line
frequency can exceed 25 mm/sec (1 inch/sec) if left uncorrected. This is a difficult problem to
diagnose if the defective connector is only sporadically making contact and periodically not.

Figure 10.199 Phasing problem (loose connection) frequency spectrum.

DC motors are also common in industry. They use power electronics to convert AC to DC
waveforms with speed control.

Figure 10.200 AC line signal converted to DC with rectification.

315
DC motor problems can be detected from higher then normal amplitudes at the Silicon Control
Rectifier (SCR) firing frequency (6X line frequency) and its harmonics. These problems include
broken field windings, bad SCRs and loose connections. Other problems include loose or blown
fuses and shorted control cards. These can cause high amplitude peaks at 1X through to 5X line
frequency (60 300 Hz).

Figure 10.201 DC motor problem frequency spectrum.

Figure 10.201 DC motor ground fault frequency spectrum.

Figure 10.203 DC motor phase loss fault frequency spectrum.

316
Figure 10.204 DC motor shorts and looseness and fuse faults frequency spectrum.

Like vibration signals, motor current can be measured and analysed to reveal important
information about en electric motors condition. Motor current analysis is good for direct
detection, diagnosis, prognosis of conditions. It follows similar analysis techniques as applied in
vibration signal analysis. A flux coil or current transformers are used to measure the current
signal.

Figure 10.205 Flux coil (left) and current transformers (right).

Figure 10.206 Frequency spectrum of a current signal from a motor with a damaged rotor.

10.4.2 Fans

Fans account for a significant number of field vibration problems due to their function and
construction. Fans move air or exhaust gases that are often laden with grease, dust, sand, ash, and
other corrosive and erosive particles. Under these conditions, fans blades gain and lose material
resulting in the need to regularly rebalance. The level of balance must also be relatively fine
because fans often have large fan-blade diameters and operate at relatively high speeds. Fans are
usually mounted on spring/damper systems to help isolate vibrations, but they are also

317
constructed in a relatively flexible manner, which adds to the demands for fine balancing. Along
with fine balancing requirements, typical problems include looseness, misalignment, bent shaft,
and defective bearings.

Fans also generate a strong response at blade-pass frequency (number of blades times the shaft
rotational speed). This frequency response is present during normal operation, but it can become
elevated if the blades are hitting something, the fan housing is excessively flexible, or an
acoustical resonance is present. Acoustical resonances are relatively common where large
volumes of air are being moved through large flexible ducts and/or the fan blades are of an air-
foil design. The primary function of a fan is to move air. Typical running speeds are between 700
and 950 RPM. The rotor is usually supported on pillow blocks (journal bearings).

Figure 10.207 Typical industrial fan, drive motor and belt drive (at back).

Figure 10.208 Fan with supports at both ends. Most fans are overhung.

Figure 10.209 Fans come in many different shapes and sizes.

318
Total fan pressure is the difference between the pressure at the fan inlet and the pressure at the
fan outlet. The fan velocity pressure is the pressure resulting from the average air velocity at the
fan outlet. Fan static pressure is the total fan pressure minus the fan velocity pressure. Air power
is the work done by the fan in moving air against a constant pressure

P=pQ
where:
P = air power
p = pressure difference between the fan inlet and outlet
Q = volume of air moved per unit time

Fan efficiency is typically around 7% and is measured as an increase in air density. The
efficiency of blowers is higher and that of compressors is higher still. Fan performance
characteristics at constant speed are shown in Figure 10.210.

fan total pressure


pressure power input efficiency, %

100 surge point

80

60 fan input power

40 fan efficiency
fan static pressure
20
fan velocity pressure

Capacity (flow) cubic feet/minute

Figure 10.210 Fan efficiency as a function of capacity.

Many centrifugal fans have a volute or scroll-type casing. In these cases the flow enters axially
and leaves tangentially. The blades may be fixed or adjustable. The best performance is achieved
at the intersection of the system characteristics and the fan pressure characteristic, both of which
can be altered mechanically.

The most common fan fault is mass imbalance from the build-up (or loss) of material on blades.
Misalignment is also common and is characterised by changes around the rotational frequency
(as discussed previously). The instrumentation needed to check fan vibration levels includes
either shaft displacement measurement hardware or casing vibration sensors. A full list of fan
faults is given in Table 10.7.

Vibration monitoring of fans should be done by mounting vibration measurement transducers in


the plane(s) of least stiffness. In order to remove the effects of duct noise or other unrelated
vibrations the measured vibration signal should be bandpass filtered from approximately 0.5X
running speed to 3X running speed.

319
Table 10.7 Common fan faults.

Mass unbalance Isolator problems


Misalignment Oil whirl
Critical speeds Rolling element bearings
Resonance Soft foot
Looseness Impeller eccentricity
Aerodynamic problems Belts and pulleys

10.4.3 Pumps

There are two principal types of pumps: (1) centrifugal pumps and (2) reciprocating pumps.
Reciprocating pumps will be discussed in a later section. The sources of vibration in pumps are
widely varied. In addition to the standard mechanical problems (unbalance, misalignment, worn
bearings, etc.), problems that are particular to pumps include vane-pass frequency generating
conditions (such as starvation, impeller loose on the shaft, impeller hitting something) and
cavitation.

Starvation occurs when not enough liquid is present to fill each vane on the impeller every
revolution of the shaft. Pump starvation can be confused with unbalance. However, it can be
distinguished by the varying amplitude 1X vibration at constant speed and the reduced load on
the driving motor.

When the vanes on the impeller are striking something, the vane pass frequency (the number of
vanes times the rotational speed) is excited. Because the striking causes a force on the shaft, an
unbalance is also present. The frequency spectrum will show a response at 1X and vane pass
frequency. The time waveform will show a high frequency response (vane pass) riding on a
frequency response at 1X. The vane pass frequency is in phase with the shaft speed. If the
impeller is loose on the shaft, the vane pass frequency will be modulated by the shaft speed.

Cavitation occurs when there is sufficient negative pressure (suction) acting on the liquid in the
system that it becomes a gas (it boils). This usually takes place in localized parts of the system.
Cavitation usually occurs in a pump when the suction intake is restricted and the liquid vaporizes
when coming off the impeller. As the fluid moves past the low pressure region, the gas bubbles
collapse. If a collapsing bubble is close to a solid surface, it will aggressively erode the surface.
Cavitation may be caused by a local decrease in atmospheric pressure, an increase in fluid
temperature, an increase in fluid velocity, a pipe obstruction, or abrupt change in direction. The
vibration signal that results will have significant vibration levels at 1X with harmonics and
strong spectral responses at vane pass frequency. High frequency broadband noise is also
common. An increase in the system pressure can reduce cavitation.

Hydraulic unbalance will result if there has been poor design of suction piping (elbow close to
inlet) or poor impeller design (unsymmetrical). The vibration signal will contain high 1X axial
vibration components. Impeller unbalance is a specific form of mechanical unbalance (as
discussed previously). High 1X vibration levels will result. Pipe stresses result from inadequate

320
pipe support and cause stress on the pump casing. This may also cause misalignment. Pipe
resonances can also be excited by vane pass frequency pressure pulsations.

Diagnosis of pump problems can be improved by installing a pressure transducer in the discharge
line of the pump. The measured pressure fluctuations can be processed in the same way as
vibration signals. The frequencies measured represent the pressure fluctuations and the amplitude
is the zero-to-peak pressure change.

Centrifugal and axial machines (whether they be pumps or fans) transport fluids by converting
mechanical work into energy of the fluid in the form of pressure and velocity. Compressors
increase the energy of the compressed fluid as pressure change. The flow can be radial
(centrifugal) or axial. Pumps work with liquid fluid while fans and compressors work with gases.

The vibration spectra of these machines are characterised by a peak at the blade pass frequency
(BPF) and/or the vane pass frequency (VPF). Calculation of these frequencies is relatively
straight forward.

BPF (Blade Pass Frequency) = No. of Blades x rpm

VPF (Blade Pass Frequency) = No. of Vanes x rpm

Pumps move fluid from one point to another by adding energy. The energy added is measured as
head rather than pressure (independent of fluid specific gravity). The total work done by a pump
is called the system head. The system head is made up of the static head, the friction head and
the velocity head.

Static head is the difference in elevation between the suction and discharge locations Static head
is measured between fluid levels or from pump centre line. Some static discharge head and static
suction head are required in order to prevent cavitation (as discussed previously). Friction head is
the head required to overcome friction losses. Friction head varies with quantity of flow, pipe
size, components in the system, and fluid type. Velocity head is the kinetic energy of the fluid
and is equal to the distance the fluid mass would have to fall to obtain the same velocity.

Figure 10.211 Pumping system head.

321
Centrifugal pumps create suction at the inlet. Pumps with one impeller are called single stage
pumps. Pumps with several impellers in series are referred to as multi stage pumps. Centrifugal
pumps provide head with both pressure and velocity components.

Figure 10.212 Centrifugal pumps.

To convert velocity to pressure a volute (ever widening pump casing) is used. Most pumps
incorporate a 360 curve. With a pump working at or near design capacity there is uniform
pressure on the impeller. Operating below design capacity results in unequal forces on the
impeller and therefore a net radial movement of the impeller. In order to reduce these pressures a
pump may be designed with a heavier shaft and bearings and/or a twin volute (180 curve
creating unequal pressures acting in opposite directions). Diffusion vanes (stationary vanes
surrounding pump impeller) act to guide the fluid flow.

Figure 10.213 Centrifugal pump flow path characteristics.

At the pump best efficiency design point, the fluid discharge angle should match the angle of the
diffuser vales and flow is smooth with minimal disturbances. If the flow is decreased (too much
back pressure) or is increased (too little back pressure), the fluid flow angle no longer matches
the design flow angle, resulting in higher vibration and loss of efficiency.

322
Figure 10.214 Centrifugal pump flow versus head curve.

Figure 10.215 Axial flow pumps.

Vibration based frequency spectra representing pumps, fan and compressors will typically have
large amplitude blade (vane) pass frequency responses, which do not necessarily mean there is a
problem. Large amplitude blade pass frequency responses and harmonics can be generated in a
pump if the gap between the rotating vanes and the stationary diffusers is not kept equal all the
way around the pump casing. Blade pass frequency (BPF) and harmonics sometimes coincide
with a system natural frequency causing high vibration. High blade pass frequency vibration can
be generated if the wear ring seizes on the shaft or if welds fastening the diffusers in place fail.
High blade pass frequency vibrations can also be caused by abrupt bends in pipework (or ducts),
obstructions which disturb the flow path, or if the pump or fan rotor is positioned eccentrically
within the housing.

Figure 10.216 Pump spectra representing a normal condition.

323
Hydraulic cavitation is a phenomenon where small and largely empty cavities are generated in a
fluid, which expand to large size (due to low pressure) and then rapidly collapse, producing a
sharp sound. Cavitation occurs in pumps, propellers, impellers, and hydro-electric turbines.

Figure 10.217 Cavitation bubbles forming behind a propeller.

Figure 10.218 Cavitation bubbles forming and collapsing as a fluid moves from high to low to
high pressure.

Figure 10.219 Cavitation bubbles collapsing and generating an erosive jet.

Hydraulic cavitation may occur when there is a decrease in atmospheric pressure, an increase in
fluid temperature, an increase in fluid velocity or viscosity, the fluid must travel around pipe
obstructions, and/or there is a change in fluid direction (deviation of laminar flow). Cavitation
causes vibration at 1X RPM with harmonics up to the blade pass frequency. There may also be
high frequency noise. An increase in system pressure can reduce cavitation. When present
cavitation may sound like gravel passing through the system.

324
Figure 10.220 Cavitation spectrum.

Recirculation results from high discharge pressure accompanied by low flow. The pump flow
pushes through seals and past impeller clearance because of the high pressure. The frequency
spectrum shows sub-harmonics of operating speed. The solution to this problem is the
installation of bypass.

Hydraulic unbalance is the result of poor design of suction piping (perhaps an elbow too close to
the inlet) or poor impeller design (unsymmetrical). This fault results in high 1X RPM axial
vibration.

Interaction with volute or diffuser occurs when the pump is operating below rated capacity.
Turbulence results due to fluid interacting with the volute or the diffusion vanes at an incorrect
angle. The frequency spectrum contains high sub-synchronous axial vibrations.

Misalignment may result from thermal growth or high pressures. The frequency spectrum will
contain 2X RPM vibration in the radial or axial directions.

Figure 10.221 Frequency spectrum showing flow turbulence in a pump.

325
10.4.4 Compressors

Compressors act in much the same way as pumps, except that they are compressing some type of
gas. They come in many different sizes, but only two principal types: (1) screw-type and (2)
reciprocating compressors. Reciprocating compressors will be discussed in a later section.
Screw-type compressors have a given number of lobes or vanes on a rotor and generate a vane-
passing frequency. Screw compressors with multiple rotors can also generate strong 1X and
harmonics up to vane-pass frequency. The close tolerances involved result in relatively high
vibration levels, even when the machine is in good condition. As with pumps, signals taken from
pressure transducers in the discharge line can be useful for diagnostics.

Figure 10.222 Different types of compressors.

Blade (vane) passing frequency is extremely common in the frequency response from pumps and
compressors. The sources of excessive blade (vane) passing frequency amplitudes in the
frequency spectrum include the following.

Rotor or housing eccentricity


Non-uniform variable pitch blades
Loose, bent or misaligned housing diffuser vanes(s)
Operation at improper performance parameters
Improper damper settings in blowers
Dirty, damaged or missing filters
Inlet or discharge line restrictions
Abrupt plumbing line bends
Resonance Excitation

Most centrifugal compressors have massive casings and lightweight rotors that make seismic
vibration measurements difficult. Permanently-mounted proximity probes are preferred to
measure relative rotor vibration. Compressor faults are similar to those encountered in steam
turbines and pumps. Fault frequencies occur at or synchronous to operating speed or its
multiples. Operation is unstable below the surge limit.

326
Figure 10.223 Compressor frequency spectrum.

10.4.5 Steam and Gas Turbines

Steam and gas turbines (and high-speed compressors) require special mention because of the
high speeds and temperatures involved. Steam turbine problems are usually limited to looseness,
unbalance, misalignment, soft foot, resonance, and rubs. As discussed above, each of these
conditions has a set of characteristic vibration responses that allow for relatively straightforward
diagnosis. However, because of the high speeds, this type of machinery is usually designed to be
lighter and less rigid than other rotating machines. Excessive vibration can therefore quickly lead
to catastrophic failure. Because of this, high-speed turbines and compressors are designed to
closer tolerances than other types of machines, and extra care is taken when balancing rotors.
These machines also frequently operate above their first critical speed and sometimes between
their second and third critical speeds. At these speeds, the rotor becomes quite flexible and the
support bearings become important in that they must provide the appropriate amount of damping.

Because steam and gas turbines are supported on journal bearings, most monitoring and
diagnostics work will be based solely on proximity probe signals. While this is not a problem in
and of itself, accelerometer signals should also be taken in order to cover the higher frequencies,
which are excited by conditions such as looseness and rubs.

10.4.6 Reciprocating Machines

Reciprocating machines (gas and diesel engines, steam engines, compressors, and pumps) all
have one thing in common - a piston that moves in a reciprocating manner. These machines
generally have high overall vibration levels and particularly strong responses at 1X and
harmonics, even when in good condition. The vibrations are caused by compressed gas pressure
forces and unbalance. Vibrations at 1/2X may be present in four-stroke engines because the
camshaft rotates at one half the crankshaft speed.

Many engines operate at variable speeds, which will allow the strong forcing functions to excite
resonances of the components and the mounting structure, if it is not designed in a robust
manner. Excessive vibrations in reciprocating machines usually occur due to operational

327
problems such as misfiring, piston slap, compression leaks, and faulty fuel injection. These
problems result in elevated 1/2X vibrations, if only one cylinder is affected, and a decrease in
efficiency and power output. Gear and bearing problems may also occur in reciprocating
machines, but the characteristic defect frequencies for these faults are significantly higher.

Chapter Summary

Analysis of the vibration signal to determine the actual type of fault present will allow for
more accurate estimation of the remaining life, the replacement parts that are needed, and the
maintenance tools, personnel, and time required to repair the machinery.
A diagnostic template can be developed for the different types of faults that are common in a
given facility or plant by listing various faults that usually develop in machinery in terms of
the forcing functions that cause them and specific machine types.
Common forcing functions include unbalance, misalignment, mechanical looseness, soft
foot, rubs, resonances, oil whirl, oil whip, structural vibrations, and foundation problems.
Specific machine components that need to be monitored include damaged or worn
rollingelement bearings and gears.
Specific machine types that can be treated as common groups include pumps, fans, electric
motors, steam and gas turbines, compressors, and reciprocating machines.
In general, vibration frequencies are used to determine the location of faults in a machine
Fault diagnosis is principally conducted in the spectrum; however, the time waveform, orbit,
and phase analysis provide additional information for in-depth analysis
Spectrum analysis includes identification of orders of shaft speed; harmonics of gear,
bearing, vane-pass and nonsynchronous frequencies
Examples of nonsynchronous frequencies are bearing frequencies, beat frequencies, natural
frequencies, sidebands, centre and difference frequencies
The spectral frequency axis (horizontal axis) can be expressed in terms of CPM, Hz, or
orders
The spectral amplitude axis (vertical axis) can be expressed in rms, peak, or peak-to-peak
The vertical axis of the time waveform is expressed in peak units
Machine faults that show up at operating speed or its orders include critical speeds, mass
unbalance, misalignment, rotor bow, excessive bearing clearance or wear, structural
resonance, looseness, eccentricity, coupling lockup, and distortion
Mass unbalance occurs at the operating-speed
Critical speeds arise when operating speed, or any of its orders containing energy, is close to
or equal to a natural frequency
Misalignment can show up at the operating speed (1X), two times operating speed (2X), or
three times operating speed (3X), depending on the nature of the misalignment and the
design of the shaft , coupling, and bearings

328
Shaft bow may significantly reduce vibration at a speed at which excitation is equal to and
out-of-phase with mass unbalance
Excessive clearance and/or wear in fluid-film bearings will cause vibration similar to mass
unbalance
Structural resonance amplifies vibration
Looseness appears in the spectrum at operating speed and its orders. Fractions (e.g. 1/2X,
1/3X) may also appear
Rolling element bearing defects occur at bearing frequencies and their harmonics. Sidebands
of operating speed, fundamental train frequency, and ball spin frequency also occur,
depending on the severity of the defect
HFD methods are used to detect pulses in machine systems
Gear-mesh faults arise in the spectrum at gear mesh and its harmonics. Sidebands occur as
the condition deteriorates
Broken, cracked, or chipped gear teeth are identified as pulses in the time waveform
Eccentric gears are identified as gear mesh and sidebands at a frequency of the eccentric
speed
Electrical problems on electric motors are identified in the spectrum as sidebands of the
number of poles multiplied by slip frequency and two times line frequency and its harmonics
Broken rotor bars generate sidebands at the number of poles multiplied by slip frequency at
operating-speed vibration and its orders
Stator problems and air gap variation arise at two times line frequency and its harmonics
Common problems related to pumps result from improper flow in the system, including
recirculation (high head) and cavitation (low head)
Pump van-pass frequencies occur if internal clearances are not set correctly
Fans may exhibit blade-passing frequency is aerodynamic problems occur in the duct, fan, or
damper design

329
Chapter 11

Machine Testing
Machine tests, other than periodic monitoring, are conducted to gain information about the
design or condition of a machine. The reasons for machine tests include the following.

Acceptance
Baseline data for periodic monitoring
Design verification (damping, natural frequencies)
Fault diagnosis and condition evaluation
Balancing

11.1 Test Plans

The elements of a test plan include the following.

Description of the machine (including a sketch, design drawing and/or photo)


Test types (all that are planned and/or likely)
Data to be acquired (measurement locations, sample rates, instruments to be used)
Loads and speeds
Machine configuration (standard operating configuration and conditions or
something special for the test)

Figure 11.1 Basic sketch of motor-fan combination to be tested.

330
Figure 11.2 Internal configuration of a gearbox to be tested.

gearbox

1 1 2 3 5 generator

2 3 4

turbine
5 4 6 7 8

8
6 7

Figure 11.3 Schematic of a turbine-gearbox-generator set to be tested.

The purposes of measurements are to perform time waveform, spectrum, phase, orbital,
synchronous time and cross-channel analyses. The functions of individual measurement points
are listed below.

Triggers recorded on turbine and generator shafts for filtering and averaging and axial
phase analysis at 1X.
Turbine bearings horizontal, vertical and axial data on the turbine governor and drive
ends.
Gearbox bearings provides data for shaft and gearbox analysis (1X velocity vibration
and gear-mesh).
Generator bearings measures basic casing data for analysis of the generator.
All bearings record shaft vibrations on the turbine, generator and gear drive shaft for
orbital analysis.

331
Table 11.1 Data acquisition plan for a turbine generator set (using an eight channel data
collector). (X horizontal, Y vertical, Z Axial, T speed)

Record Measurment Recorder Channel Purpose


no. (units) (sensor location and direction)
1 2 3 4 5 6 7 8
1 Velocity 1X 1Y 1Z 2X 2Y 2Z 1T 7T Basic Turbine
Analysis
2 Velocity or 3R 3A 5R 5A 6R 6A 1T 7T Basic Gearbox
Acceleration Analysis
3 Velocity 7X 7Y 7Z 8X 8Y 8Z 1T 7T Basic Generator
Analysis
4 Displacement 1V 1H 2V 2H thrust thrust 1T 7T Turbine Shaft
(Pk to Pk) A B Vibrations
5 Displacement 7V 7H 8V 8H 3V 3H 1T 7T Generator/ Gearbox
(Pk to Pk) Shaft Analysis
6 Displacement 6V 6H 4A 4R 3A 5A 1T 7T Gearbox Shaft/
(Pk to Pk) Casing Analysis
7 Velocity 1Y 2Y 3R 6R 7Y 8Y 1T 7T Cross Sensitivity
8 Velocity 1X 2X 3A 6A 7X 8X 1T 7T Cross Sensitivity
9 Velocity 3Z 4Z 5Z 6Z 2Z 7Z 1T 7T 1X Phase Analysis

Data recording using tape recorders permits many different types of analyses to be performed
after the data has been collected. Data collectors are preferred if vibration levels are sensitive to
speed or load allowing some analysis during collection. The setup parameters of data collectors
should be selected prior to data acquisition. A two-channel data collector is required for orbit,
synchronous averaging, and cross-channel analyses.

Diagnostic tests are concerned with the goal of the plan being followed (and may vary).
Operating speed tests are conducted to obtain data for fault analysis and condition evaluation.
Impact and start-up/coast-down tests are utilized to obtain natural frequencies and critical speeds.
Acceptance tests are conducted to determine whether or not the new or repaired equipment meets
the purchase specifications. Baseline tests are used to acquire vibration data that represent
normal machine condition and operating conditions. Calibration tests are conducted to provide
information on balance weight sensitivity and phase lags in the machine.

11.2 Selection of Test Equipment

The selection of equipment depends on the goals of the plan and the equipment available. Special
transducers may be needed if low frequencies or high temperatures are involved. A tracking
analyser will be required for start-up and coast-down tests. Data collectors perform 95% of the
work including data storage.

332
11.3 Site Inspection

Site inspection and evaluation are important regardless of the type of data acquisition plan. Site
factors that may account for excessive vibration include bolts (potentially loose), foundation
(potential weak/cracked), grouting (cracked), piping (loose/unsupported), and changing thermal
conditions. Non-operating speed components of vibration should be eliminated by assessing the
environment when the equipment is not operating and by obtaining time-averaged data.

11.4 Acceptance Tests

Acceptance tests are typically conducted based on the purchase specification. These include
procedures, measurement locations, process conditions, measures and acceptable levels of
vibration. The purchase specification should include testing procedures as well as acceptable
levels of vibration similar to ISO standards. In the absence of a purchase specification, a baseline
test can be conducted and compared with general vibration standards.

11.5 Baseline Tests

Baseline tests are used to determine the nature and level of normal vibration of a machine. It
should be conducted prior to and during periodic monitoring program activity. Condition can be
observed and maintenance action initiated when baseline vibration levels change. Baseline tests
are also able to reveal design and installation problems such as resonance, critical speeds,
alignment, soft foot and distortion.

11.6 Resonance and Critical Speed Testing

These tests are carried out to obtain information about the dynamic characteristics of a machine
and its structural support and piping. The information can be used in machine diagnostics and
redesign in order to overcome chronic problems. Resonance occurs when a vibratory excitation
force is equal in frequency to a natural frequency of the system. Resonances are often artificially
excited with hammers and shakers in order to obtain an accurate estimation of the natural
frequencies.

The natural frequency of a machine or structure is governed by the design. Each machine/system
has a number of natural frequencies that can be excited by impact, random forces or harmonic
vibrating forces at or close to the natural frequency. In general, higher order natural frequencies
are not multiples of the first natural frequency. Vibration levels are amplified at resonance
frequencies. A mode shape is defined as the deflection shape assumed by a system vibrating at a
natural frequency. The different mode shapes of a system are associated with the different
natural frequencies. A mode shape consists of deflections at selected points in the system that are
determined relative to a fixed point.

Conducting a resonance test involves determining the vibrations of the structure at a number of
known points. The data collection frequency span should be selected wide enough and with
sufficient resolution. Only one impact should be made within the data acquisition time. Strike
the structure with a 4x4 timber, mallet or hammer with a soft head in the direction of the desired

333
mode (smaller hammers for smaller structures). If the desired mode is not known, strike the
structure in several directions (i.e. vertical and horizontal). Measure and record the vibration
levels at a number of reference points on the structure.

force
F

sensors
Figure 11.4 Schematic of modal (natural frequency) testing.

When conducting a resonance test the peaks in the vibration spectrum at various measurement
points indicate the natural frequencies of the structure. Some natural frequencies may not be
excited if the excitation impact occurs at a nodal point of a particular mode shape. A rule of
thumb is that the spacing between forcing frequencies and natural frequencies should be 15
percent. The structure, piping or machine should be as close as possible to its operating state.
Parts of a machine can not arbitrarily be removed and tested. For example, the natural
frequencies of a gear not mounted on its shaft differ from those when the gear is mounted.

Procedures for conducting a critical speed test require that one or more appropriate transducers to
measure the vibration be selected. Proximity probes are preferred if they are permanently
installed. Velocity and acceleration transducers should be close to the bearing(s). Wire the
vibration transducers and trigger to a tracking analyser, tape recorder or data collector. To
perform coast down test, run the machine at 10 ~ 15 % over speed and then cut the power and
allow the machine to coast down from normal operation. Process the data and identify the critical
speeds from FFT spectrum, Bod or polar plots. The natural frequency at an operating speed is
not necessarily the measured natural frequency during start-up or coast-down test.

When using the FFT analyzer/data collector the peak hold feature holds and displays the peak
values of all data after each spectrum is computed. The peak hold feature can be used to provide
data on critical speeds. The relationship between acquisition time (Ts), number of lines (N) and
frequency span (Fmax) is: Ts = N/Fmax

Polar plots show the amplitude and phase of vibration at various speeds. The tracking analyser
plots the real and the imaginary amplitudes at the various speeds. In Figure 11.5 the small loop
identifies the first critical speed of the generator (1,200 RPM). The maximum vibration
amplitude occurs at 2,100 RPM. Operating speed is 3,600 RPM.

334
Figure 11.5 Polar plot of a start-up test of a generator bearing

In summary, a tracking analyser is best for rapid run-up and coast-down tests. Vibration is
indicated in the filtered frequency band. A reference signal may be generated by a proximity
probe with notch or optical pickup with reflective tape. Peak vibration levels and phase changes
indicate critical speeds. Signal-channel analyzer/collector may be used for impact tests in either
time or frequency domains. Triggering can be free or from a hammer source. Vibration peaks
indicate resonances. During impact tests, a uniform window should be used on the analyzer.

11.7 Specifications

The purposes of preparing a specification is to procure quality equipment and services, avoid
misunderstandings, resolve differences of opinion and to establish a methodology for testing the
equipment without controversy. It is best to use existing ISO (or other) standards as guidelines
for preparing a specification.

11.8 Environment and Mounting

Mounting of the machine is often a cause of excessive vibration especially on vertical pumps.
Natural frequencies should be set away from operating speeds by pump manufactures to avoid

335
resonance. The customer is responsible for ensuring the mounting is not interfering with natural
frequencies. Sufficient bracing and support should be used in piping to assure a natural
frequency higher than pump specifications.

11.9 Presentation of Data

The presentation of data is valuable for fault analysis, condition evaluation, baseline testing.
Acceptance testing data is typically presented in a simple form involving overall levels. Spectral
data provide resolution and dynamic range sufficient to discern important frequencies and
amplitudes. The time waveform should be presented so that the data can be related to the
physical characteristics of the machine. Detailed waveform variations should be observable.
Amplitude trends may be used for long term monitoring. The span of the time waveform should
be equal to the data acquisition time from the analyzer. Orbits should not be filtered at operating
speed. High frequency filtering may be used to remove noise. The phase of operating speed
vibration to a spot on the shaft is valuable for analysis.

11.10 Reports

Reports should be written for each of the following activities:

Baseline testing
Acceptance testing
In-depth analysis including operational tests, resonance and critical speed tests
and/or environment tests.

The organization of a report should include an executive summary, an introduction, a technical


discussion, conclusions and recommendations. There should also be an appendix (for the detailed
technical data). The introduction section should describe the equipment being tested, the purpose
of the test, the approach to the test and the test equipment and techniques used. The technical
details supporting the conclusions and recommendations are presented in the technical
discussion. Conclusions and recommendations are necessary for all major findings found in the
survey or analysis. The appendix should include a description of measurement points, the last
measurement report, any trend plots, and spectral data on exceptions and alarms.

Types of reports include baseline reports. These are reports that provide a complete picture of the
condition of the equipment or faults present. They contain the normal vibration levels to the best
of the capability of the analysis. They suggest values for setting alarms. In-depth analysis may
include selected time waveforms, spectra, orbits, tests from impact or coast-down tests.

An acceptance test report should be linked to the specification. A complete analysis of the
machine may be carried out during the acceptance testing. After balancing the vibration levels
and the trial weights applied should all be detailed. Final readings should be recorded.
Conclusions and recommendations can be brief but should be inclusive.

336
Chapter Summary

A test plan should be generated prior to data acquisition on a machine acceptance tests,
baseline tests, fault analysis, condition evaluation, design, and balancing
The test plan should contain a description of the machine, the tests to be performed, the data
to be acquired, loads, speeds, machine configurations, and process conditions
The data acquisition plan should provide details about sensors including location,
measurement parameters, and process conditions
If data are processed on site, the analyzer setups must be provided, including frequency
spans, lines of resolution, range, windows, and time spans.
Sometimes multiple data acquisitions are required to obtain adequate range and resolution
A site inspection should provide details about external vibrations and machine mounting
Acceptance tests are to be listed in detail in the purchase specification of a new or required
machine.
Included in the acceptance tests are procedures, measurement locations, process condition,
measures and the way they are processed, and acceptable vibration levels
Baseline tests are conducted to establish normal operating levels of vibration when the
machine is in good operating condition
Specifications should be used to assure the procurement of quality equipment
Be realistic about acceptance levels and locating critical speeds
Good mounting environments and procedures will assure that equipment is operating
properly
Presentation and reporting of data provide quality analysis of quality data

337
Chapter 12

Trouble Shooting
Excessive machinery vibrations can result in component or structural failure, fatigue, cracking,
plastic deformation and product damage. To solve vibration problems it is necessary to identify
forcing functions and eliminate or minimize the effect.

12.1 General Steps

This chapter will discuss the following topics related to trouble shooting unknown conditions or
problems in machinery.

1. Identify the problem


2. Gather information
3. Determine possible forcing functions
4. Determine where to take data and what data collection equipment to use
5. Take vibration data
6. Analyse vibration data (and any other data available)
7. Make recommendations

12.2 Identify the Problem

Identifying a problem requires an initial inspection of the site (overview). Noise problems are
often solved using isolation and/or insulation. True vibration problems need correction.

Figure 12.1 Site plan of the inside of a large excavator.

338
Figure 12.2 Photo of a swing motor and gearbox inside excavator.

12.3 Gather Information

A sketch of the major machine components (noting modifications) is a good place to start.
Details of the major components (specifications on motors, pumps, bearings, etc.) should be
recorded. The maintenance history should be reviewed. It is often helpful to talk to operators and
maintenance staff abut the problem and the history of the machine and the particular problem.

The sketch helps locate the best positions for sampling vibration data and should include the
following.
a) All major components
b) Shaft diameters, lengths
c) Rotor dimensions, weights
d) All bearing specifications

Figure 12.3 Typical sketch of a machine setup showing all major components and potential
vibration data collection locations.

Bearing specifications should include the bearing type, number, ball or roller diameter, the
number of balls or rollers, the pitch diameter, the contact angle, and the shaft speed. If sleeve
bearings are present the type and clearance should be noted. Belts and chains require the centre
to centre distance, pulley pitch diameters, number of belts, and belt length all be recorded.

339
Coupling type, bending tolerance and coupling length should be noted. Drives, such as motors,
engines, and turbines needed to be listed and their specifications detailed. Gears present should
be defined in terms of type and layout, number of teeth, gear ratios and drive gear speed. Fans,
pumps and anything else important (nearby equipment, foundation structure, etc.) should also be
recorded.

A maintenance history should include what was the last thing done to the machine and the
maintenance frequency. This information should point to a root cause of the existing problem.
An example of what to look for in a maintenance history can be seem if the situation involving
several premature bearing failures is considered. Potential causes could be poor installation
techniques, poor lubrication, shaft bearing misalignment, overloading, electrical discharge
through bearings, a previous major failure, or some other modifications. Only by checking the
maintenance history will these potential causes be distinguished.

The operators input may include such things as the characteristics and severity of vibration at
the time of the problem, the rate of onset of the problem (sudden or gradual), the persistence of
the problem (is it continuous or intermittent (related to load, speed, temperature, time)), and any
previous similar problems (including any action taken in the past to correct a similar problem).

12.4 Determine Possible Forcing Functions

Knowing these allows later focus of frequency analysis on characteristic defect frequencies.
Possibilities include the following.

a) Rotor frequencies
- unbalance of rotor (stress relief, weight shift, material loss/gain)
- bow in rotor shaft (slow roll)
- bearing misalignment
- coupling misalignment

b) Rolling element bearings


- low start up resistance, high running friction
- speed restrictions (RPM): grease < 7,200 / shaft diameter; oil < 9,600 / shaft diameter
- characteristic defect frequencies quite distinctive

c) Sleeve bearings
- shaft rides on layer of lubricating oil in bearing journal
- may be pressure fed
- optimum efficiency dependant on shaft speed, lubricant viscosity and load.
- oil whip, oil whirl

d) Rub
- seal rubs
- loose bearing housings

340
e) Belt frequencies
- primary belt frequency - number of complete circuits per minute.
- belt length

f) Gear mesh frequencies


- rotational speed number of teeth
- offset parallel shafts
- planetary gears - speed changes on inline shafts

g) Blade pass frequency (fans, pumps)


- number of blades shaft rotational speed
- vibration created as blades pass outlet port
- high vibration levels if pump is operating far from optimum efficiency
- can induce duct (pipe) resonances

h) Resonant frequencies
- excited by forcing functions that are close to structural resonance (natural frequencies)

12.5 Determine Where to Take Data and What Data Collection Equipment to Use

The best transducers to use are dependent on the situation. The direction is also somewhat
dependant on the situation, but should start out being in all three directions (axial, vertical,
radial). An overall survey followed by more detailed investigation is the typical strategy.

Running speed is critical for baseline comparisons and identification of frequency dependant
features. The running speed is almost always a critical piece of information. Running speed can
be determined directly by observing the speed from instrumentation on the machine.
Alternatively a measurement using a tachometer or strobe light may be taken. Remember that if
possible speed should be steady during the data collection period (otherwise order tracking may
be required). The running speed can also be determined from the frequency spectrum. In general,
the first dominant peak in a frequency spectrum will the first order followed by peaks at
harmonic intervals. If a machine is driven by an induction motor at synchronous speed, look for
vibration peaks at 1800 or 3600 rpm.

12.6 Take Vibration Data

Time domain data is useful for seeing the overall amplitude and transients, (spikes,
discontinuities), as well as the character of the stationary signal (changes with time, load), and
line frequency.

Frequency domain signal (signature)


periodic components of time waveform
relative amplitudes of frequency components
harmonic relationships
precise location of frequency components

341
Within a frequency signature:
Synchronous Components
- N Speed of rotation, (N = integer)
- N = 1 : fundamental frequency

Where N = 1 to 8 (low multiples)


- unbalance
- belt pitch line vibration
- shaft/bearing misalignment
- bent shaft
- looseness
- blade pass frequency (pumps, fans)

Where N > 8 (high multiples)


- gear mesh frequency
- blade pass (compressors, turbines)
- motor slot frequency
- cavitation

Sub synchronous (less than 1 shaft speed)


- primary belt frequency
- oil whip, oil whirl
- rubs
- loose rolling element bearing in housing

Non Synchronous
- other components in machine
- other machines close by
- electrically caused vibrations

12.7 Analyse Vibration Data (and any other data available)

Estimate forcing functions present based on all the data collected and using the methods and
practices described earlier in this book.

12.8 Make recommendations

The recommendations depend on many factors. Primary among these will be the financial
considerations implicit in all maintenance decisions and discussed in earlier chapters of this
book.

342
Chapter 13

Advanced Methods of Machine Condition Monitoring


Much of the discussion in the previous chapters has highlighted the fact that many machine
defects generate distinctive vibration signals. This fact has been exploited recently with the
development of a variety of different automatic fault diagnostics techniques. The details of these
systems will not be provided here. The goal of automatic diagnostics is to augment and assist,
rather than replace, the vibration signal analyst. If characteristic defect indicators can be detected
and extracted from a vibration signal without the intervention of a signal analyst, the analyst will
have more time for other duties and will also have access to information that may not have been
uncovered through normal signal processing and analysis.

There are, however, still many situations where machine defects do not generate distinctive
vibration signals or when the vibration signals are masked by large amounts of noise or
vibrations from other machinery. In such cases, advanced diagnostic algorithms incorporating
new signal processing techniques are currently being developed and implemented. Artificial
neural networks have been found to provide an excellent basis for detecting and diagnosing
faults. Wavelet analysis and short-time Fourier transforms (STFTs) have also been shown to
effectively allow both time domain and frequency domain information to be displayed on the
same plot. This provides an opportunity to clearly see short duration transient events as well as
detect faults in machinery that is operating in nonsteady-state conditions.

13.1 Automated Machine Condition Monitoring

Automated machine condition monitoring and diagnostic tools for decision support which
complement the expertise of the analyst. Decision support in this context can be in the form of
statistical techniques, expert systems and/or smart on-line monitoring systems. These systems
and techniques capitalize on past experience (historical maintenance and vibration data) and
expertise of maintenance staff. Continuous surveillance of multiple machines is readily possible
along with integration with the larger IT picture (production, inventory, planning, etc.).
monitoring of remote equipment also becomes more practical.

Extracting condition information from machinery is argely a pattern recognition problem. The
signals from a machine are reduced to patterns or vectors. These vectors are then a digital
representation of the health of the machine. The challenge is how to classify the collected vectors
in a meaningful way.

The advantages of adopting such a strategy for condition monitoring include the fact that these
techniques are flexible. They are applicable in various applications and on many different
machines. The resources required are minimal (although some programming and/or data input
may be required) and the preferences of maintenance staff can become features of any system
developed. These systems are typically modular with the general system configuration being
easily duplicated. These systems are expandable and can therefore grow to support various levels
of sophistication. All hardware and software used are typically mainstream and easily
implemented. Communications within and outside the system is typically able to fit into existing

343
protocols (company network, paging systems etc.). In general these systems are (or should be)
easy to use and robust. They should have databases for common machine elements and
implement a transparent decision process.

Figure 13.1 Example commercial (Rockwell Automation) integrated automated monitoring


system.

Transducer
Transducer
Feature Classification

3
4 1 3
4 3 1 9
3 9 4 4
4 5 4 4 4

3 1
5 3 5
1 2 5
2
6 2 2 2
2 0 7 6 0
6 7 6
6 6 6
6
5 6 6
5 0 0

Figure 13.2 Schematic showing the signal processing and logic steps involved in a typical
automated monitoring system.

There are some issues with these systems which typically centre around the data that needs to be
collected, analyzed and stored. Sufficient storage capacity needs to exist for the storage of failure
data for forensic analysis. The overall data requirements must be practical given that it takes little
time to collect huge amounts of data. With or without automated analysis these huge amounts of
data can quickly overwhelm a system/analysis team. In some situations (where system training
examples are not available), the system must be able to detect faults with little or no previous
fault data.

344
Automated condition monitoring and fault diagnosis based on vibration signals is a truly multi-
discipline field of activity. All of the following play a role in the design and function of a useful
system.

Mechanical Engineering
Dynamics, Materials, Vibrations, Acoustics, Machine Design, Operations
Management, Maintenance Practices.

Electrical Engineering
Signal Processing, Data Acquisition, Electrical Machinery.

Computer Science and Applied Math


Artificial Intelligence, Feature Extraction, Pattern Recognition, Data Mining, Data
Fusion, Programming.

Domain Specific
Particular to a given machine.

13.1.1 Expert Systems

Expert systems have been around for a reasonable amount of time and are considered a mature
signal processing and decision making method. The method is basically the creation of sets of
rules from expert knowledge that already exists. Templates and software exist that can simplify
rule creation. The rules can be based on simple limits all the way to complex multiple parameter
relationships. These systems are typically single parameter based and use statistical analysis and
trending primarily.

The advantages of expert systems are that they have been proven to work well in well understood
applications. They also provide a transparent solution/outcome. That is, it is clearly discernable
where and how the decision came about.

The disadvantages include the fact that they can be difficult to create. Comprehensive rule bases
may be challenging to generate because of complex relationships between multiple parameters.
In particular fuzzy (relationships defined in non-numeric terms such as hot, cold, warm)
relationships are hard to codify. These systems may be difficult to apply to complex or unsteady
machinery. They also have difficulty dealing with noisy or incomplete data.

An expert rule base may be developed from input from maintenance staff, knowledge of the
behavior of machine under different operating conditions, and measured data. Rules should be
based on known system behaviour that dictates the response to condition indicators. Figure 13.3
shows a typical expert system architecture. Figure 13.4 shows the interaction of those who need
to be involved in the construction of an expert system.

Expert system rules consist of IF - THEN logical statements. These are set up as IF antecedent -
THEN consequence. These statements can be crisp or fuzzy. That is, they can be based on
numeric evaluations (IF set point temperature is above 50 C THEN change feed rate to 10

345
m/sec.). As well, they can use fuzzy statements (IF set point temperature is HIGH THEN
change feed rate to FAST), as long these statements are clearly defined.

Figure 13.3 A typical expert system architecture.

Figure 13.4 Interaction of people and systems elements to construct an expert system.

346
Figure 13.5 Hierarchical structure of expert system (example 1).

Figure 13.6 Hierarchical structure of expert system (example 2).

Pump Output vs. Swash Plate Angle


trendline y = 5.5499x - 37.427
16.00

14.00
Pump Output

12.00

10.00

8.00

6.00

4.00
8.00 8.20 8.40 8.60 8.80 9.00 9.20 9.40 9.60
Swash Plate Angle

Figure 13.7 Sample of machine data used to develop a rule (easily correlated parameters).

347
13.1.2 Fuzzy Logic based Signal Analysis

Fuzzy logic uses a known relationship to map discrete (crisp) domain values onto Fuzzy
membership values. As an example, consider the definitions below. How would one define 25C,
as hot or mild? Or a new category - warm? As well, how would one define 29C, as hot? Almost
hot? Fuzzy set membership functions (the function that defines particular membership values to
particular domain values) allow these questions to be answered and applied in a meaningful
manner. In this way the degree to which a particular temperature fits into a particular domain can
be defined. Using this method the value of 29C could potentially be defined as being 80% in the
hot classification and 20% in the mild classification.

30C - hot
20C - mild
10C - cool
0C - cold

Figure 13.8 Generic fuzzy set membership function.

Figure 13.9 Generic fuzzy set membership functions.

An example follows showing the use of fuzzy logic to automatically categorize frequency
spectra representing different rolling element bearing faults. The signal analysis involved each
fault type being represented by 15 samples (frequency spectra from 1 to 128 Hz) of that fault
type. The first step was to find the mean and standard deviation of each data set (fault type) at
each frequency (from 1 to 128 Hz). The mean values, N times the standard deviation were used
as upper and lower limits of the fuzzy membership functions.

348
Outer race Outer race fault

Inner race fault

No fault

Figure 13.10 Sample rolling element bearing frequency spectra.

Figure 13.11 Fifteen different spectra combined fault type #1.

Figure 13.12 Average spectra combined fault type #1 (2.5 Standard Deviations).

349
Values of 1 10 for N were used. As the upper and lower membership domain limits change
(proportional to N times the Standard Deviation) so did the fuzzy membership values. Both the
Pi shaped and triangular shaped membership functions were used. The Pi curve worked best
because of its similar shape to a normal distribution.

Figure 13.13 Self test classification trial (outer race fault, varying membership function limits).

Figure 13.14 Fuzzy membership results triangular curve domain function


(outer race fault vs. other fault types).

Figure 13.15 Fuzzy membership results Pi curve domain function


(outer race fault vs. other fault types).

350
13.1.3 Artificial Neural Networks

Automatically detecting and diagnosing faults in machines can be sub-divided into the tasks
shown in Figure 13.16.
Input

Sensing

Classification
Segmentation

Domain Specific
Feature Extraction

Numerical Feature
Extraction

Classification

Post Processing

Decision Support

Figure 13.16 The tasks involved in automatically detecting and diagnosing faults in machines.

Sensing is the act of measuring a physical property and collecting the electrical signal that
represents this property. Examples include temperature, vibration, acoustic emissions, pressure,
flow, current and voltage.

Segmentation is the task of gating a continuous measurement into segments for comparison.
Segmentation methods include time based (most common) and using other parameters marking
the beginning and end of each segment. Figure 13.17 shows the time based segmentation of four
different parameters.

T T
Figure 13.17 Time based segmentation of parameters.

351
Figure 13.18 Speed-based segmentation of parameters.

Domain specific feature extraction is the task of extracting features from a signal which are
likely to reveal a fault. These are typically specific to different applications and the anticipated
faults. Examples of domain specific features for vibration include time-based features (such as
RMS, Kurtosis, Peak-to-Peak, Crest Factor, autoregressive model based features (to be discussed
later)). Frequency based features include constant percentage bandwidth features and tracked
orders.

Numerical feature extraction is the task of increasing the information in a raw feature vector
through the removal of redundant or dependant features, by reducing the size of the feature
vector and/or by revealing features that provide separation between classes of data. Common
methods include manual inspection, principal component analysis (PCA), non-linear principal
component analysis, independent component analysis (ICA), and Fischer discriminant analysis
(among many).

The classification task involves the mapping of the feature vectors by condition. This step
requires training using example data. Figure 13.19 shows an example of a linear boundary for the
classification of 2-D data.

Figure 13.19 An example of a linear boundary for the classification of 2-D data.

352
Because example fault data is required the central aim of designing a classifier to correctly
classify novel input has not yet be achieved. The issue is one of generalization. In a real life
system, we usually have to tolerate some sort of error.

The decision tree classification method is a hierarchy based method. It is used to break down a
complex multivariate decision making process into separate sets of simpler individual decisions.

Figure 13.20 Decision tree classification method.

Supervised neural net classifiers are methods that are loosely based on the function of the brain.
They are trained on massive amounts of example data (in a similar way to the way the human
brain learns). It is widely applied to condition monitoring applications. It involves training a
classifier with examples of data representing every machine state (including faulty and normal
conditions) that it must recognize during operation.

Several benefits are often attributed to the application of neural networks to pattern
classification. There is no need for an explicit rule base. They are able to learn from experience.
They can perform complex tasks without programming or a predefined rule-base (as is the case
with expert systems and decision trees). They are able to deal with large data sets. Typically their
performance tends to increase as the number of measurements increases. They can model
complex, multidimensional data and they can deal with noisy or incomplete data.

The disadvantages of using supervised artificial neural networks are the fact that they are
dependant on a comprehensive fault data base. They are not likely to deliver repeatable results.
They are sometimes difficult to optimise. They must be treated somewhat as a Black Box. That
is it is usually difficult to determine exactly the internal behaviour that results are based on.

13.1.4 Support Vector Machines

Support vector machines (SMV) are kernel based learning algorithms which were first
introduced in the latter half of the 1960s. It was not until the early 1990s, when cheap high
performance computing power became widely available, that the techniques used for SVMs

353
began to emerge and become practical. What makes this technique particularly attractive is that
the approach is systematic, reproducible and properly motivated by statistical learning theory.
Unlike neural networks, the results of SVMs can be investigated and reproduced easily. SVMs
have been applied to a wide variety of classification tasks including linear regression, multi class
classification and recently novelty detection.

SVMs can be thought of as creating a linear boundary between the two classes of data. In the
case where a linear boundary is inappropriate (for instance for data that is not linearly separable)
the SVM can map the input vector into a high dimensional feature space. By choosing a non-
linear mapping a priori, the SVM constructs an optimal separating hyper-plane in this higher
dimensional space. Different techniques (such as Nearest Neighbour, Bayes, Least Squares) can
be used within SVM.

Figure 13.20 Simple Support Vector Machine geometric representation of a functional separation
between two classes (classification boundary with margin).

Figure 13.21 More complex classification relationships can be simplified by transforming the
representation to higher dimensions and separating the classes with a hyperplane.

354
Figure 13.22 Multiple classes shown with a 3-D hyperplane.

13.1.5 Novelty Detection

Novelty detection is a method that involves a model built on normal data. New data is then
compared to the model. Degrees of novelty are defined as a novelty score. A predefined
threshold identifies faults (classes).

Feature Novelty Novelty Threshold Diagnostic


Vector Detector Score
Determination Result

(Training Stage Only)

Adaptation Data A Priori


Algorithm Model Model

Figure 13.23 Novelty detection process.

The advantages of novelty detection are that it is not dependant on knowledge of previous faults.
It is able to detect known and unknown (previously unseen) faults. It offers many of the
advantages of previous ANN approaches. The primary disadvantage is that it is unable to
diagnose fault types without example data.

Neural network based novelty detection is a pragmatic approach addressing the data issue
mentioned above. It becomes a two class classification problem.

355
Novelty detector training involves the network being trained to model and reconstruct a single
class of data. During novelty detector operation new data is introduced to the fixed feature
network. The class of data is judged on the ability of the network to reconstruct the sample data.
The reconstruction success is of reconstruction is judged based on the Euclidian distance
between the input and the output of the network. High novelty scores indicate possible imminent
failure (data that has significantly changed from the original no-fault data).

Figure 13.24 Novelty detection network.

Approaches to novelty detection include parametric statistical approaches (Gaussian mixture


models, hypothesis testing, hidden Markov models). Non-parametric statistical approaches
include the K-Nearest Neighbor method, clustering and Parzan windows. Neural network based
approaches include multi-layer perceptrons (MLP), Kohonen self organizing maps, radial basis
functions, and auto-associative neural networks. Support vector machine based approaches
include the Schllkopf approach and the Tax and Duin approach.

Applications include analysing mammograms, fraud detection, radar tracking and of course
condition based maintenance.

13.1.6 Automated Machine Condition Monitoring A Case Study

A case study is presented in which a machine that normally operates in an unsteady state is
analyzed. This case represents a relatively large group of machinery that operates in a range of
normal modes. This can have a profound influence on measurement parameters and make
condition monitoring difficult. Unaccounted for or unexpected mode shifts can result in false
alarms.

Factors that influence operational modes in machines include the duty (speed, loading, operator,
time of day), environmental characteristics (ambient temperature, season), and the machine age.
Some possible approaches to this situation are considered below.

1. Ignore: The classification system attempts to model all modes. This approach will typically
result in a generalized model with reduced fault detection sensitivity. Alternatively there is an
increased risk of false alarms. Generally, ignoring the problem will not make it go away.

356
2. Data Normalization: This strategy attempts to identify the governing parameters of the
system. These parameters are then linked to the associated modes of operation. A
classification system can than be designed to recognize the modes of operation. When data is
collected from a known mode the appropriate optimal classifier can be dispatch to deal with
this data. Note that all modes must be represented in the training data. This is the approach
that is used in the following case study.

This case study reports on the computational approach of an autonomous condition monitoring
system for a large hydraulic system. The project involved experimental development of the
system followed by the practical implementation.

The application was a double roll crusher at Syncrude Canada Limited, Fort McMurray, Alberta.
The specific subsystem was the apron feeder hydraulic drive system for the crusher.

Figure 13.25 Crusher (KRUPP) being fed by haul trucks.

Figure 13.26 Crusher apron feeder (chute top, crusher rolls bottom).

357
Figure 13.27 Hydraulic (constant displacement) drive motor for the apron feeder.

Figure 13.28 Electric motors and hydraulic (variable displacement) pumps for the apron feeder.

The apron feeder hydraulic system is made up of three 9-piston variable displacement hydraulic
pumps. These pumps are driven by constant speed electric motors. The pumps are in a parallel
configuration and discharge into a common manifold. The apron feeder speed is controlled by
the variable displacement from these pumps. The pump swash plate angles regulate
displacements of pumps which in turn regulates the speed of the hydraulic motors driving the
apron feeder.

358
Figure 13.29 Schematic of overall material handling system. A lengthy pump failure could lead
to depletion of the surge pile and interruption of the feed stream.

The justification for a dedicated condition monitoring system for this machine is based on the
following. The maintenance history of these pumps shows that the apron feeder hydraulic drives
have been subject to various sporadic and unpredictable failures. The pumps are located in a
relatively remote location meaning that access to the crusher for maintenance staff is difficult,
especially in winter or wet conditions. Condition monitoring via manual data collection and plant
information system (already in place) yielded little insight into the causes of the failures.

Due to the serial nature of the operation, a failure in the crusher lasting more than a short time
will result in a production stoppage on that line until the problem is rectified. Due to the highly
irregular nature of the ore load on the conveyor (trucks dump at irregular intervals) and the stop
and go nature of its duty (apron feeder operates only when there is ore in the hopper chute), the
hydraulic system is subject impulsive shock loading and several different modes of normal
operation. The effect of this duty is twofold. Firstly, it results in unpredictable premature
failures. The second effect is that this impulsive loading renders a simple limit based condition
monitoring system useless.

System requirements in this case require that it should capitalize on past experience and expertise
of maintenance staff. It should be modular and expandable to be able to accommodate additional
machines. It must be able to detect faults with little or no previous fault data. All hardware and
software used must be mainstream and easily implemented.

The proposed solution was to develop a dedicated PC based on-line condition monitoring system
incorporating an expert system and artificial neural network based novelty detector.

359
HYDRAULIC ELECTRICAL ROOM
SYSTEM PENTIUM III
ANALOG
SIGNALS PC
ANALOG
TRANSDUCERS
RS232 Cable
PC
-Case Drain x3
-Pump Output x3
Analog Matlab
-Swashplate x3 I/O Neural
-Charge Pressure
-Drive Speed Network Network
Interface Multifunction LabVIEW
I/O Data
PROGRAMMABLE
Acquisition
LOGIC Card
CONTROLLER

Ethernet Card
HYDRAULIC
or MODEM
PUMP#1
ACCELEROMTERS
Signal Conditioning

HYDRAULIC Shielded
SH100100
PUMP#2
ACCELEROMTERS
Connector Cable
Block
HYDRAULIC
PUMP#3
ACCELEROMTERS

Client PC Company LAN or


Imminent
Phone Line
VIBRATION SIGNALS failure report

Web enabled mobile


device

Figure 13.30 Schematic of the overall monitoring system hardware configuration.

The advantages of using a PC based system in this case included that it would be relatively
inexpensive (when compared to dedicated stand alone systems commercial systems). The
hardware and software were more easily supported by the company IT department. The system
can be easily upgraded and expanded. It provided relatively inexpensive and reliable data
storage. The PC platform allows many separate software types to work in unison. It provided
adequate computing power for advanced signal processing and computing applications. It also
provided familiar interfaces for operators. Finally, it was easily integrated into the existing
company network.

The monitoring system software was used the Windows NT operating system (for stability).
National Instruments LabVIEW was used for data acquisition hardware control, expert system
routines and user interface development. MATLAB was used for neural network development
and execution of neural network routines. The expert system rulebase was developed from
maintenance staff input, knowledge of the behavior of the machines and measured data.

360
Table 13.1 Instrumented measurement parameters.

Parameter Transducer
Pump Vibration (9 total, 1 per pump axis x 3 pumps) ICP Piezoelectric Accelerometers
Pump Output Flow (3) Inline Flowmeters
Swash Plate Angles (3) Potentiometer
Motor Current Draw (3) Inline
Apron Feeder Speed (1) RVDT via PLC
System Charge Pressure (1) Inline Pressure Transducer
Manifold Pressure (1) Inline Pressure Transducer
Pump Case Drain Flow (3) Inline Flowmeters

Figure 13.31 Instrumented pump.

Figure 13.32 The monitoring system.

361
The expert system rulebase development began with sample machine data representing easily
correlated parameters. Since the swash plate angles for the three pumps are linked together, they
have identical loading and the motors driving them are always at synchronous speed. The current
draw for the pumps should be identical (within an allowable tolerance). If one current goes out
of the tolerance, it could signify a failure.

Pump Output vs. Swash Plate Angle


trendline y = 5.5499x - 37.427
16.00

14.00

Pump Output
12.00

10.00

8.00

6.00

4.00
8.00 8.20 8.40 8.60 8.80 9.00 9.20 9.40 9.60
Swash Plate Angle

Figure 13.33 Pump output versus swash plate angle.

Simultanious Motor Current Measurements


DRC7 Apron Feeder (normal operation)
35.00
Current (Amps @ 600VAC

30.00
25.00
Motor 11
20.00
Motor 12
15.00
Motor 13
10.00
5.00
0.00
15
22
29
36
43
50
57
64
71
78
85
92
99
1
8

Sample Number

Figure 13.34 Motor current versus time.

The neural network based novelty detector needed to be used due to fact that there was no failure
data to work with.

Several challenges were encountered. These included the fact that multiple operating states were
typical. This wide range of normal operating behaviour made it difficult for a single network to
model without compromising sensitivity. Transient behaviour was also common. Spikes and
transient behavior lead to high novelty scores and false alarms. This was likely during mode
changes and shock loading of hydraulic system.

The proposed solution included using multiple operating modes. A single governing parameter
was found to influence modes of operation, namely the swash plate angle. An expert system was
designed to dispatch an appropriate specialized network depending on the mode of operation (the
swash plate angle). This led to better modeling. The implementation involved sorting the data by
governing parameter. A histogram was used to sort data while ensuring a minimum amount was
required for effective training. The system automatically decides how many modes to model and
the corresponding networks to create and train.

362
Figure 13.35 Training data segmented into modes based on swash plate angle.

The monitoring system diagnostic routine was a combination of expert system and neural
network based novelty detector. The novelty detector output was input for the expert system.

DATA
ACQUISITION Slowly Changing Raw Vibration
ROUTINE Parameters from Signal from High
FieldPoint Module Speed DAQ Board

DIAGNOSTIC Crest Frequency RMS


ROUTINE Factor Spectrum Values

Redundant Feature
data Vector
removed

Vectors

Vectors sorted into groups


of common operating

Autoencoder Autoencoder Autoencoder


Mode 1 Mode 2 Mode N

Expert System
Inference
Engine

User Interface And Diagnostic


Communications Result
Routine

Figure 13.36 Diagnostic routine data flow.

363
The problem of transients reducing the ability of the system to perform fault discrimination was
addressed by designing the system to recognize the difference between gradually developing
faults and momentary transient behaviour. Only persistent behavior occurring during stable
operation was considered for fault detection purposes. The implementation of this strategy used a
moving average transient filter to detect transient behavior. A novelty score and duration were
then passed on to the expert system.

Novelty Detector Response to Unfiltered New


Running Data

450
Relative Novelty Score

400
350
300
250
200
150
100
50
0
1 101 201 301 401 501 601
Consecutive Samples

Figure 13.37 Novelty score clearly distinguishing transients.

The performance of the novelty detector was validated on test data. Known faults were simply
caught by the expert system. Progressively developing unknown faults were not described in
expert rule-base and were therefore overlooked by expert system. These data sets resulted in high
novelty scores from the novelty detector, indicating a faulty condition was developing or already
present.

Novlety Detection Errors on Progressive Simulated


Faults

80.0
70.0 Progressive Fault 1
Mean Square Error

60.0
(Novelty Score)

50.0 Progressive Fault 2


40.0
Mean Fault Free
30.0 Error
20.0 Above + 1 stdv
10.0
0.0
1

13

17

21

25

29

Fault Progression

Figure 13.38 Novelty detector output from progressive faults (simulated data).

364
Figure 13.39 Novelty detection using auto-encoders feature vector interface.

Figure 13.40 Novelty detection using auto-encoders sensor output interface.

365
Figure 13.41 Novelty detection using auto-encoders monitoring parameter interface.

Figure 13.42 Novelty detection using auto-encoders vibration alarm interface.

Based on preliminary results and offline testing the system was able to detect known and
unknown (previously unseen) faults. Novelty detection proved to be an effective approach, if the
proper steps are taken during implementation. To be truly useful future work must address
integration of the system into the overall decision support systems at the company. This
approach merits further investigation, refinement and development.

366
13.2 Model based Spectral Estimation

As outlined in the previous sections automatic diagnostic techniques do exist and implementation
is difficult, but not impossible. The most straight forward example of this type of analysis is the
application of amplitude limits on FFT based frequency spectra within different frequency bands
(bandwidth acceptance limits).

Amplitude

Frequency

Figure 13.43 Constant percentage bandwidth acceptance limits can be used for automatic fault
detection and diagnosis.

Mathematical models can also provide descriptions of system responses and be useful for fault
detection and diagnosis. They have the added advantage of being readily automated. Changes in
the model are sensitive to changes in the system (faults). Model types that have been explored
for use in machine condition monitoring applications include the following.

Auto-Regressive (AR) models


Auto-Regressive Moving Average (ARMA) models
Minimum Variance (MV)
Prony Models

Model based frequency spectra are also useful for vibration signal analysis. Because the model
used as the basis of the frequency spectrum calculation is sensitive to the changes in the systems
under consideration, the spectra reflect this sensitivity. Another advantage to using this method
to extract frequency information from a vibration signal is that the length of vibration signal
needed to generate a frequency spectrum is considerably shorter than for FFT based techniques.
If data collection time or the amount of signal to work with are limited these methods become
extremely useful. The only disadvantage is that the process is more computationally expensive.
However, with inexpensive and powerful computing readily available, this is currently not a
serious concern.

There are three steps in the process of calculating a frequency spectrum from a vibration signal
using model based (parametric) methods. Here steps are listed below.

367
1. Selection of an appropriate model type.
2. Calculation of the model parameters and determination of the optimum model order (size).
3. Calculation of the spectral estimate.

Selection of an appropriate model type may at first seem challenging as there are many different
types of models. However, the general form of the data that is under consideration will typically
lead to a few optimum choices.

Figure 13.44 Model types and their advantages and disadvantages.

368
Figure 13.44 (contd) Model types and their advantages and disadvantages.

One method of quickly selecting an appropriate model type is to consider the data
autocorrelation data. Figure 13.45 shows two typical auto-regressive (AR) process
autocorrelation functions. If the autocorrelation function from a sample signal is similar to the
shapes shown in Figure 13.45 it is likely that an auto-regressive model will provide a suitable
representation of that data (system).

Figure 13.46 shows the autocorrelation function from data sampled from a rolling element
bearing. While it is not exactly the same as the shapes shown in Figure 13.45, it is similer and
suggests that the underlying process can be considered as autocorrelated. That is, recorded data
points are related in a predictable way to data that came previously. An auto-regressive model is
suitable for representing this data (system).

369
Figure 13.45 Sample autocorrelation functions (theoretical left, from sampled sinusoidal
waveform right).
Correlation

Time

Figure 13.46 Sample autocorrelation function from sampled rolling element bearing vibration
data.

Equation 13.1 shows the standard auto-regressive equation, where all data points are related to
previous data points by some weighting factor. This equation can significantly compress all the
features that represent the data used to build the model, and therefore the system that generated
the data. The equation has the additional use of being able to regenerate synthetic data that
represents how the system would respond.

X (t ) a0 X (t 1) a1 X (t 2) a2 X (t 3) n(t ) 13.1

Model parameters represent a weighted function (series of terms) that, when used as a filter with
pure noise (random dynamic data) will generate the original time series used to make the model.
The equation should contain all the valuable information required to reproduce the original
signal (data compression). It should also contain fault classification information.

A detailed description of the calculation of model parameters will not be presented here. A brief
overview follows the references listed below. These references provide descriptions of the most
common model calculating algorithms.

370
Yule Walker Method

Yule, G.U., On a Method of Investigating Periodicities in Distributed Series with Special


Reference to Wolfers Sunspot Numbers, Transactions of the Royal Statistical Society of
London, Series A, Vol.226, p267-298, July 1927.

Walker, G., On Periodicity in Series of Related Terms, Transactions of the Royal Statistical
Society of London, Series A, Vol.231, p518-532, 1931

Levinson Durbin Algorithm

Levinson, N., The Wiener (root mean square) Error Criterion in Filter Design and Prediction,
Journal of Mathematics and Physics, Vol.25, p261-278, 1947.

Durbin, J., The Fitting of Time Series Models, The International Institute of Statistical Review,
Vol.28, p223-244, 1960.

Wiggins, R.A. and E.A. Robinson, Recursive Solutionto the Multichannel Filtering Problem,
Journal of Geophysical Research, Vol.70, No.8, p1885-1891, 1965.

Forward Linear Prediction

Kay, S.M., Modern Septral Estimation: Theory and Application, Prentice-Hall, Englewood
Cliffs, New Jersey, USA, 1988.

Kay S.M. and S.L. Marple, Spectral Analysis: A Modern Perspective, Proceedings of the
IEEE, Vol.69, No.11, p1380-1419, November 1981.

Morf, M., B. Dickinson, T. Kailath and A. Vieira, Efficient Solution of Covariance Equations
for Linear Prediction, IEEE Transactions on Acoustics, Speech and Signal Processing,
Vol.ASSP-25, p429-433, October 1977.

Forward-Backward Linear Prediction

Marple, S.L., A New Autoregressive Spectrum Analysis Algorithm, IEEE Transactions on


Acoustics, Speech and Signal Processing, Vol.ASSP-28, p441-454, August 1980.

Burg Method

Burg, J.P., Maximum Entropy Spectrum Analysis, Proceedings of the 37th Meeting of the
Society of Exploration Geophysicists, Oklahoma City, Oklahoma, USA, October 1967.

Once the type of model to use has been decided, the basic procedure for the calculation of model
parameters is as follows. The covariance function derived from a given data set (time series) is
used to generate a set of model parameters. Some random data is filtered with the new model and
an approximation of the original data set is generated. The original time series data used to build

371
the model is compared to the time series data generated from the model. The model parameters
are then adjusted in some way to reduce the error between the model based time series data and
the original data. This process is repeated until the error is suitably small. The final result will be
a model that suitably (within acceptable error limits) captures the system behavior.

A critical step not mentioned specifically above is the determine of the optimum model order.
That is, the size of the model (the number of parameters that are used in the model). A small
number of parameters (small model order) is efficient, but may not suitably capture all the
important system characteristics. A large model order will do a better job of defining the system
characteristics, but it will not be computationally efficient. A balance is required. Several
estimation criteria for optimum AR model order selection exist and are listed in Figure 13.47.

These functions give an estimate of the goodness of fit that models will have to data as a function
of the model order. Figure 13.48 shows one of these functions decreasing gradually as the model
order if increased. Again, the strategy is to select a model order that is most efficiently represents
all the important systems characteristics. This figure shows that there is considerable
improvement as the model order grows (decrease in loss function) in the early stages (up to
model order 10). The rate of decrease slows above that until a model order of about 50 is
reached. The loss function decreases dramatically above model order 50 and then the rate of
decrease levels off again. This suggests that there is something in the sample data that is well
represented by a model order above 50, but that there may be no reason to have model orders
above 50. Unfortunately, this is a trail and error process, but only needs to be done once for a
given system.

Figure 13.47 Estimation criteria for optimum AR model order selection.

372
Figure 13.48 Final Prediction Error (FPE) Loss Function for Rolling Element Bearing Data

Figure 13.49 shows a series of frequency spectra calculated using increasing auto-regressive
model orders. The spectra generated from low model order models lack detail. The spectra
calculated from models with higher model orders show more detail. It is a somewhat subjective
judgment as to the optimum model order in any particular case. One could argue that in this case
a model order of about 60 would be sufficient, while not being excessively large.

Figure 13.49 Rolling element bearing outer race fault AR spectra (increasing model orders)

Equation 13.2 shows the formula that is used to calculate an AR frequency spectrum using an
AR model.

2t
F AR ( f ) 2 13.2
p
1 a k exp j 2 fk t
k 0

373
Figure 13.50 AR model based spectral estimate sine waves in noise (left) FFT-based spectral
estimate sine waves in noise (right).

An example of the use of AR model based spectra calculated from models of different orders
follows.

Figure 13.51 Vibration signal from an outer race fault on a rolling element bearing (raw vibration
signal upper left, vibration signal after high pass filtering upper right, vibration signal after
rectification lower).

374
Figure 13.52 AR frequency spectrum of outer race fault (model order 20 left, model order 40 -
right).

Figure 13.53 AR frequency spectrum of outer race fault (model order 60)

Figure 13.54 AR frequency spectrum no fault.

375
Figure 13.55 AR frequency spectrum rolling element fault.

Figure 13.56 AR frequency spectrum inner race fault.

As outlined above parametric models (such as the AR model) can be used as fault detection and
diagnostic tools. They facilitate the generation of useful frequency spectra from relatively short
vibration signals. However, they also provide an opportunity to apply automated techniques that
use the model, rather than the vibration signal or frequency spectrum, for fault detection and
diagnosis. An example of how this can work is described below using nearest neighbour
classification.

Different time series (vibration signals) can represent different conditions (faults), but these are
often difficult to distinguish. When converted to models they become easier to distinguish or
group into sets with similar characteristics primarily because they models themselves have
similar characteristic (such as being all of the same order) that allow easier classification. The
difference between two sets can be defined as in equation 13.3.

376
f 0 x
I ( f0 , fm ) f 0 x log dx 13.3
f m x
Where fo and fm are the probability density functions of two different variables. When xo and xm
are multidimensional, normally distributed variables, with mean values o and m and the
covariance matrices o and m, then,

2 I ( f 0 , f m ) log
m
tr 1
tr 1
0 m 0 m n 13.4
0
m 0 m

Where: |A| is the determinant of matrix A, tr(A) is the trace of matrix A, A-1 is the inverse of
matrix A, and A is the transpose of matrix A.

If only a sample of data is available, (exact probability density functions are not known) then an
approximation can be made using equation 13.5.

1 1

2 d ( x 0 , x m ) log m

tr tr 0 m 0 m n 13.5
m m
0
0

Where the ^ represents estimated values based on the sample data. Given that each sample time
series has a corresponding AR model, a dissimilarity number can be determined, equation 13.6
can be written.

log m2 1 p0 pm
2d (x 0
,x m
) 2 a 0 i a m k C 0 k i n 13.6
0
2
m i0 k 0

Where j2 is the sample covariance, aj is the AR model parameter, pj is the AR model order, and
Co is the estimated covariance function.

x t k x x t x
nk
1
C j k j j
13.7
n t0

n
1
x
n
x t ;
t 1
j
k 0 ,1, , p j 13.8

Knowing 2d(x(o), x(m)) (the dissimilarity between x(o) and x(m)) we can determine the probability of
misclassification of a sample as


Pe exp d x o , x m 13.9

377
The probability of fault existence is then defined as the likelihood of a fault being present when
comparing new samples with samples known to represent fault-free conditions and represented
as equation 13.10.

P fe 100 1 Pe 13.10

Experimental results using the same data as above are shown in Figures 13.57 to 13.59.

Figure 13.57 Statistical distance measure between signals. Same signals (on the diagonal)
generate small distances (due to small differences). Note: NOF No Fault, ORF Outer Race
Fault, REF Rolling Element Fault, IRF Inner Race Fault.

Figure 13.58 Probability of fault existence (Pfe) measure between signals. Same signals (on the
diagonal) generate small Pfe (due to small differences).

Figure 13.59 Probability of fault existence (Pfe) measure between signals. Same signals (left hand
column) generate small Pfe (due to small differences).

378
The above illustrates a good example of a trending and classification parameter that can
distinguish between fault-free conditions and various types of faults, as well as distinguish
between each fault type. However, this procedure needs known fault data. Another concern is
what happens when faults are poorly distinguishable (early stages) or if the data is noisy. The
following example is one that addresses these issues.

This is an example of supervised classification. The user defines the specifics of classification (#
of classes, etc.). Some prior knowledge of the system and signals is required. The tool that will
be used in known as inductive inference classification. The data is reduced to a common form,
which removes redundant/unneeded data. Classification based on the length of description of a
data set is then possible.

An example of how this works is shown when a group of people are classified using as few
parameters as possible. Physical attributes such as sex, height, weight, hair colour, eye colour,
etc. may be used for classification. By randomly dividing the sample into groups, then shifting
members between groups we can zero in on the shortest (optimum encoded) description of all the
groups. That is, everyone ends up in the group that best describes them. Descriptions are
typically based on sample statistics.

Figure 13.60 Data to be classified as shown in true (correct) classes.

Figure 13.61 Data description length estimation equations.

379
Figure 13.62 Estimated data description lengths for various classifications.

Experimental Results Low Speed Rolling Element Bearing

REF
NOF

COM1
ORF ORF & REF

IRF COM2
ORF, REF
& IRF

Figure 13.62 Experimental results for a low speed rolling element bearing.

Figure 13.63 Classification results for all fault types.

380
Figure 13.64 AR model based frequency spectra from low speed rolling element bearing
gradually deepening outer race fault.

Figure 13.65 Estimated data description lengths vs. gradual development of an outer race fault.

Figure 13.66 Estimated data description lengths vs. gradual development of an outer race fault.
(1 NOF, 2 ORF, 3 REF, 4 IRF, 5 COM1, 6 COM2)

381
Figure 13.67 Flowchart of procedural steps.

The following is another example of experimental results (cutting tool deterioration tests)
showing how automated signal classification can work. There were five tests. This first four are
accelerated wear rate tests and the final (the fifth) was a regular wear rate test.

Figure 13.68 Feed rate and speed conditions

382
Figure 13.69 AR frequency spectra for test # 1.

Figure 13.70 AR frequency spectra for test # 2.

Figure 13.71 AR frequency spectra for test # 3.

383
Figure 13.72 AR frequency spectra for test # 4.

Figure 13.73 Classification results accelerated wear rate test data.

Figure 13.74 AR frequency spectra for test # 5 (normal wear rate).

384
Figure 13.75 Classification results normal wear rate test data.

Figure 13.76 Estimated data description lengths versus gradual deterioration. Baseline minimal
wear. (1 test #1, 2 test #2, 3 test #3, 4 test #4).

Figure 13.77 Estimated data description lengths versus gradual deterioration. Baseline
advanced wear. (1 test #1, 2 test #2, 3 test #3, 4 test #4).

385
Figure 13.78 Estimated data description lengths versus gradual deterioration. Baseline minimal
wear. (test #5 - normal wear rate).

13.3 Minimum Variance based Spectral Estimation

Minimum Variance based spectral estimation is valuable because it is the only spectral
estimation method that requires no prior knowledge of the measured signal to yield separate
spectral estimates for both periodic and random signals.

The goal of machinery maintenance programs is to develop an accurate physical understanding


of the machinery. A significant portion of that understanding comes from analyzing the dynamic
signals that are measured from that machinery. This implies that the signal processing methods
used must be appropriate for the types of measured signals. Here we will focus on signals from
constant speed machines.

During constant speed operation, rotating and reciprocating machinery are expected to generate
mixed signals that consist of a combination of periodic and random signals. Unbalance,
misalignment, impacts, looseness, etc. can generate periodic signals. Fourier tells us that a
periodic signal can be expressed as a sum of sinusoids that are all periodic over the same
interval. Periodic signals are therefore associated with line spectra (spectra with very distinct
lines at specific frequencies) because all the power in the signal is located at the frequencies of
the sinusoidal components that combine to make up the overall periodic signal.

Random signals can be generated by sources such as flow, rubbing (friction) and combustion.
There are two basic varieties of random vibration signals associated with constant speed
machinery. They are time invariant (wide sense stationary) and periodically time dependent
(cyclostationary). Using sound as an example, a time invariant signal will always sound the
same, whereas the sound from a cyclostationary signal will change periodically as a function of
time. For example, the periodic movement of a piston in an internal combustion engine will
result in periodically modulated random signals due to the rubbing of the piston on the cylinder
walls and the flow of gases into and out of the cylinder.

The minimum variance (MV) approach to spectral estimation is quite different from the
conventional FFT approach. The MV spectrum is based on using a signals auto-correlation

386
sequence as opposed to being directly applied to the time waveform that is used by an FFT. To
explain the MV method, two examples will be presented. The first will focus on single channel
issues and the second will consider multi-channel issues.

Consider the frequency response of a resonant system (natural frequency = 256Hz, damping ratio
= 0.01) exited by white noise with an added sinusoidal (320Hz) at the output.

320Hz Sinusoid
+
Resonant System Output
White Noise +
n = 256Hz
= 0.01

Figure 13.79 Data generation schematic for MV example.

10

0
Power (dB)

-10

-20

0 150 300 450 600


Frequency

Figure 13.80 Random signal frequency response.

20

0
Power (dB)

-20

-40

0 150 300 450 600


Frequency

Figure 13.81 MV method applied to output (10, 40,160, 320th order).

The use of the 320th order MV spectrum means that the auto-correlations are used up to the 320th
order lag. Multiple data sets are not required to produce the sequence of spectral estimates. The
spectral estimates are obtained from a single auto-correlation sequence that was obtained from a
single data set. Note that the sequence of spectra presented are monotonically decreasing

387
downward, with the 10th order at the top. This is the first key property associated with any MV
spectrum based analysis. Higher order spectra are always below lower order spectra when
calculated from the same data set.

For MV spectra the existence of a sinusoid is indicated by the convergence of the monotonic
sequence of spectra at a specific frequency. In this case the spectra converge at 320Hz. By
comparison, the spectra do not converge in the vicinity of the system resonance at 256Hz,
indicating a random signal rather than a sinusoidal one.

No Convergence Convergence

20

0
Power (dB)

-20

-40

0 150 300 450 600


Frequency

Figure 13.82 Convergence of the monotonic sequence of spectra at a specific frequency.

The level at which the spectra converge for a particular frequency is known as the power of the
sinusoid. In this case the amplitude of the sinusoid (in the time domain) is 5.221, therefore the
following is true.
A2 5.2212
Power 10 log 10 log 11.34db
2 2

The advantage to using a sequence of MV spectra can be demonstrated by considering the MV


spectra at single model orders. Consider the peak at 256 Hz for the 160th order MV spectrum.

20
Power (dB)

-20

-40

0 150 300 450 600


Frequency

Figure 13.83 The peaks at 256 Hz and 320 Hz for the 160th order MV spectrum.

388
The large broad peak at 256 Hz is similar to the large broad peak at 320Hz for the 40th order MV
spectrum. Hence, as with standard spectral estimators, when using only a single spectral
estimate, one could mistakenly conclude that there are two sinusoidal frequencies in the data. It
is only by considering the convergence pattern for a sequence of MV spectra that it is apparent
which peaks correspond to periodic signal components.

Close inspection of the MV spectra as the order increases shows that the spectra are approaching
drops of 3dB for every doubling of the model order at pure random signal component
frequencies. This not only demonstrates the rate of convergence, but also qualitatively shows
how noise is being driven out of the spectra. In essence, the MV spectra are working to drive out
half the noise in the original signal for each doubling of the model order until a sinusoid is found.

This illustrates the very practical approach to model order selection possible with MV spectral
estimation. Random signal components can be driven out to a desired point where any remaining
undiagnosed, or underlying, periodic signal components are considered insignificant due to their
small power. As a result, the rate of convergence of the MV spectra is obviously tied to the local
signal to noise ratio (SNR). The higher the local SNR, the more quickly convergence occurs
because fewer orders are required to uncover the underlying sinusoids.

At this point it is important to note two things.

1. All power/amplitude estimates are taken based on peak values only. There is never any
summation or integration under the peak.

Hence, the MV spectrum as described here has no leakage or associated windowing issues. The
observation window may always be taken to be rectangular. Because there are no leakage issues,
the multi-channel MV spectrum are not generally susceptible to coherence dropout.

2. MV spectral estimates are robust with respect to perturbations in frequency.

Since all MV spectra are defined over the frequency continuum, as opposed to specific bin
frequencies, any perturbation in frequency of a periodic signal component will result in a
corresponding perturbation in the spectral peak frequency. The peak remains essentially constant
otherwise.

Since the convergence properties of the MV spectral estimate are so well understood, in practice
a convergence test can automatically scan for and identify periodic signal component frequencies
and then accurately estimate the limit points for the converging spectra that correspond to the
various periodic signal component powers. For this example, the estimated sinusoidal amplitude
is 5.221, which is equal to the actual sinusoidal amplitude. In practice the limit values can be
automatically placed into a conventional line spectrum without having the user consider the
sequence of MV spectra. It is important to note that along with the limit estimate on the
sinusoidal power, a PSD estimate for the associated random signal is also obtained.

389
A weighted MV spectrum is when a weighting of between 0.0 and 1.0 is applied to the
correlations. This forces the weighted MV models to converge to a weighted estimate of the
random signals PSD. Once again monotonic convergence is evident, except that now the periodic
signal influence is gradually removed from the spectra. In addition, note that the convergence
occurs at all frequencies, as expected, since the continuous spectrum exists everywhere.

30

Power (dB) 20

10

0 150 300 450 600


Frequency

Figure 13.84 Weighted MV method applied to output (10, 40,160, 320th order).

Multi channel issues will be discussed with the aid of a two channel (periodic plus random)
signal system. The system will be used to demonstrate the multi channel MV spectral estimation
(both weighted and unweighted) properties along with a brief comparison to the FFT. Note that
to ensure optimum performance of the FFT, this example uses 100 data records each consisting
of 4096 computer generated data points where the periodic signals all exist at bin centered
frequencies.

160Hz Sinusoid Input

+ + Output
Channel
Input Resonant
System +
White 384Hz Sinusoidal
+
+ Output
Input Channel
+

160Hz Sinusoid Input

Figure 13.85 Details of the periodic signal structure.

The key issue in considering multi-channel (in this case three vibration signal sources inside the
one machine and being recorded on one accelerometer channel) MV analysis is the concept of an
eigenvalue spectrum. Relative to the user, an eigenvalue spectrum is qualitatively identical to an
autospectrum (auto-correlation based spectrum). In fact, for the single channel case, the
eigenvalue spectrum and the autospectrum are identical. They look similar and there is an
eigenvalue spectrum and an autospectrum for each channel. For the multi channel case, these

390
spectra are computationally different and the eigenvalue spectra have an additional special
significance.

0 0

Power (dB)
Power (dB)

-10 -10

-20 -20
0 100 200 300 400 500 0 100 200 300 400 500
Frequency Frequency
Channel 1 Channel 2

Figure 13.86 System frequency response (channel 1 and 2).

1 200
Degrees
MSC

0.5 0

1.0 -200
0 100 200 300 400 500 0 100 200 300 400 500
Frequency Frequency
Magnitude Squared Coherence Phase

Figure 13.87 System Frequency Response (coherence and phase).

In a complex machine, multiple signal sources can exist at the same frequency. By using the
multi channel eigenvalue spectrum, each source can be identified. Each channel will have an
eigenvalue spectrum.

Both eigenvalue spectra 1 and 2 in Figure 13.88 converge at 160Hz, but only eigenvalue
spectrum 1 converges at 384Hz. This illustrates the advantage of using eigenvalue spectra. Each
sinusoidal source causes a convergent eigenvalue at its freqeueny. Based on these spectral results
we know that there is only one source at 384Hz and at least two sources at 160Hz.

To determine the number of different sources at 160Hz, we would need to add another channel.
Adding a third channel would add a third eigenvalue spectrum. Then, because there are only two
sources at 160Hz in out sample data, the third eigenvalue would not converge (similar to the
second eigenvalue at 384Hz). It is important to note that the purpose of the sequence of the MV
spectra is to identify sinusoidal sources.

391
20 10
Power (dB)

Power (dB)
10
-10

-10 -40
0 100 200 300 400 500 0 100 200 300 400 500
Frequency Frequency
Eigenvalue 1 Eigenvalue 2

Figure 13.88 MV spectra (orders 10, 40, 160, 320).

Once the sources are identified, the limits and their frequencies for the sequence of eigenvalue
spectra can be determined, along with accurate estimates. What if there are more sources than
channels? Say three sources at a given frequency and only two channels. The third source is then
distributed over the other two. The final estimates of power will not be affected.

Note also that while the peaks at 384Hz in eigenvalue spectra #2 do not converge, they are there
and relatively sharp. These peaks are caused by crosstalk, which is a common problem in
parametric spectral estimation methods. By using a sequence of spectra, the crosstalk effect can
be removed (with no convergence the peaks are ignored).

28 20
Power (dB)

Power (dB)

22
0

18

12 -20
0 100 200 300 400 500 0 100 200 300 400 500
Frequency Frequency
Eigenvalue 1 Eigenvalue 2

Figure 13.89 Weighted MV spectra (orders 10, 40, 160, 320, weighting 0.96).

The influence of the sinusoids is reduced with continued weighting of higher order models.

It is important to remember that the MV random spectrum contains the spectral information for
all non-periodic signals. In other words, it does not reject non-stationary random periodic signals,
but instead models them as wide sense stationary signals. This is typically not a major issue

392
when data is taken at nominally steady state operating conditions on rotating machinery. In many
cases the most significant consequence is that the modulation information associated with
cyclostationary random signals is lost. This information is usually not critical, however it can be
estimated using time-frequency methods if needed. In contrast the MV sinusoidal spectrum seeks
to reject all non-periodic signals within the limit of its application.

The MV sinusoidal spectrum converges to the periodic signal line spectrum for any physically
realizable background random signal process (regardless of the colour of the noise or non-
stationarity of the signal).

While a sample plot will not be shown here, an FFT based auto-spectrum (with bin centered
periodic signal frequencies, a large record size, and a large number of records) would be
generally comparable to the relevant MV spectrum. However, the MV spectrum would have
superior accuracy and smoothness (lower variability). These characteristics have potentially
larger benefits for common multi-channel applications such as dynamic balancing, where
accurate phase and power estimates are crucial.

Bias percentage (needed when beginning a MV spectral analysis) in used to account for the
requirement that the random signal component should have a positive PSD at all frequencies.
This results in two problems. First, a random signal does not possess energy simultaneously at all
frequencies. A true continuous spectrum is based on a long time average. Second, for the data
used, regions may exist that do not have sufficient measurable energy. As the model order
increases, this causes numerical instability.

Because MV polynomial models are based on inverting matrices, the energy at a specified
frequency cannot be zero (and should not be close to zero) since dividing by a small number will
yield a very large number (potentially infinite). The solution to this situation is that white noise
can be progressively added to stablize these regions. This process in controlled by the bias
percentage.

The bias percentage is the percentage of the total original signal power that is progressively
added to the auto-correlation sequence in the form of a white noise correlation sequence to
eliminate the previously mentioned numerical instabilities. If the bias percentage is too small, the
computation slows unacceptably. Bias must be added and tested repeatedly until sufficient bias
(white noise) is achieved for numerical stability. If the bias percentage is too large, the white
noise can unacceptably corrupt spectral estimated in other regions. Therefore, bias should be
kept as small as possible. A setting of between 0.001 and 0.1 is recommended.

In summary, the minimum variance based decomposition spectrum derives its power in a novel
way by using the monotonic convergence properties of a sequence of spectra obtained from a
fixed data set. As a result of these properties, and based on the previous discussion and
examples, a number of unique properties regarding this technique can be listed. MV spectra can
be used to develop distinct and accurate spectral estimates for both periodic and random signals
without any prior knowledge regarding the measured signal. Hence, results can be displayed in a
true line spectrum or a power spectral density plot, both of which are easy to interpret. The
technique can identify periodic signals embedded in an unknown, arbitrary, and possibly non-

393
stationary random signal environment. There is no leakage with respect to power estimation on
periodic signals. For multi-channel situations, the MV spectra are not susceptible to coherence
dropout or crosstalk. There are no windowing issues in MV spectral analysis. All data is
rectangularly windowed. Because the MV spectra are defined on the frequency continuum, the
estimates are robust with respect to perturbations in frequency.

13.4 Time-Frequency Analysis

A signal can be represented in several ways depending on the purpose of analysis. The time
domain is a plot of magnitude versus time. The advantages of using time domain as an analysis
tool include the fact that harmonic motions and parameters such as amplitude, period, and
frequency can be obtained from the waveform. It is readily available on most data collectors and
analyzers. The disadvantages include the fact that for complex motions the time waveform can
not typically give much information. The display is in two dimensions (amplitude versus time) in
the time domain and no frequency information is given directly except in harmonic motion.

The frequency domain is a display of magnitude versus frequency after performing the Fourier
transform. The advantages to this presentation format are that it reveals the spectral components
of a signal. It is a well established technique in machine condition monitoring and also available
on most data collectors and analyzers. The disadvantages are that it is suitable for periodic signal
analysis only. It shows data in only two dimensions (amplitude and frequency).

The time-frequency domain data presentation format is the display of a signal after performing
various transforms (Short Time Fourier Transform, wavelet, etc.). The advantages are that it is
suitable for analyzing non-stationary signals because both time and frequency information are
displayed. Three dimensional information (amplitude, time and frequency) is available with
close correlation. Higher resolution and localization may also be achieved. The disadvantages
include the relatively time consuming computations required and the fact that it is not typically
available in portable data analyzers.

Stationary signals are signals where the frequency does not change with time. They can be
signals are when the frequency content evolves with time. These signals need time-frequency
analysis techniques. Examples are biomedical and speech signals.

A signal in domain can be expanded or projected into several subspaces. The expansion can be
realized using a linear combination of elementary functions n.

s an n 13.11
n

The expansion coefficients an depict the signal behaviour which can be calculated as


an s, n st n t dt 13.12

Types of expansion include orthogonal expansion and non-orthogonal expansion.

394
Figure 13.90 Orthogonal expansion.

Figure 13.91 Non-orthogonal expansion.

Signal expansion is fundamental to signal processing. Orthogonal expansion is typically


preferred. Through expansion, a signal can be projected into definite physical domains (time and
frequency). The selection of elementary functions forms the basis of different time-frequency
techniques (STFT, wavelet).

Frequency Time-frequency

Time

Figure 13.92 Time-frequency representations.

395
Time-frequency representations involve shifting and scaling. A shift in time by t0 results in
multiplication by a phase factor in the frequency domain. Scaling in the time domain leads to
inverse scaling in the frequency domain. Equation 13.13 defines shifting. Equation 13.14 defines
scaling.

13.13

13.14

Figure 13.92 Scaling by different factors in time domain.

The temporal variance and the frequency variance must satisfy the Heisenberg uncertainty
principle. There is a minimum area of the Heisenberg box, as shown in Figure 13.93. It implies
that the optimum localization cannot be reached in both the time and frequency domain at the
same time.

Figure 13.93 The Heisenberg box.

Applications of time-frequency analysis include speech mechanism understanding, biomedical


engineering (brain functions through EEG, ECG and MRI image analysis), machinery signal
analysis and diagnosis, signal detection in noise, radar signal detection, image processing, and
economic data analysis.

396
PSD

Time waveform

Figure 13.94 The word Hood spoken by a five-year old boy.

Time waveform Spectrogram

Figure 13.95 Vibration impulse response of a beam.


Frequency (kHz)

No knock Knock

Figure 13.96 Engine knock detection.

397
Normal spur gear One broken teeth

Figure 13.97 Gearbox fault diagnosis Wigner-Ville distribution.

Fourier transform T-F transform

Figure 13.98 Radar image of an aircraft.

13.5 Short Time Fourier Transforms (STFT)

The definition of a continuous short time Fourier transform is shown in equation 13.15 below.

13.15

x(t) the signal


(t) short window function
t, - time and frequency indices

The definition of a discrete short time Fourier transform is shown in equation 13.16 below.

13.16
T, - time and frequency steps

398
The short time Fourier transform can be interpretation in several ways. Which basically means
that it can be used for different purposes. The first of these interpretations is to use the STFT to
calculate a series of spectra that represent the changing frequency content of a time waveform as
a function of time. This requires at time t, the transform signal is s(t) multiplied by a running
short time window function (-t). Following the windowing, the Fourier transform is performed.
By moving the window and repeating the previous two steps, a series of localized spectra can be
obtained. Summing the spectra gives the STFT spectrum (in the time-frequency domain).

Alternatively the local spectrum can be calculated specifically.

Figure 13.99 Local spectrum can be calculated specifically.

The local spectrum can also be calculated repeatedly using Fourier transforms.

Figure 13.100 Local spectra can be calculated repeatedly.

The STFT can also be considered as a filtering process.

STFT maps can also be constructed by combining signals from the time domain to the time-
frequency domain. However the inverse process may not be possible. STFT is the subset of the
entire two-dimensional function B(t,) shown in Figure 13.102.

399
Figure 13.101 Local spectra can be considered a filtering process.

Figure 13.102 STFT maps

STFTs are linear. That is the signal x(t) = x1(t)+x2(t), or STFTx(t,) = STFT1(t,) +
STFT2(t,). There is the usual window selection trade-off. A small time window, (t), has good
time resolution but poor frequency resolution. A large time window, (t), has poor time
resolution but good frequency resolution.

The resolution in time and frequency of the STFT depends on the choice of the window in time
and frequency. According to the uncertainty principle, no best window is available in both the
time and frequency domain. Similar to Fourier transform, windows like rectangle and Hanning
can be used. Once a window is selected, the resolution is determined across the t-f plane.

The spectrogram (the time-frequency plot resulting from a STFT) is concentrated at (0,0), with a
center of s(t). The contours of the spectrogram are ellipses. The smaller the ellipse, the better the
concentration of the STFT. The minimum area is reached when = and the area is 2

400
Figure 13.103 STFT spectrogram (basic simplified presentation).

The MATLAB function specgram is able to perform STFT and calculate the spectrogram. The
syntax is as follows.

specgram(x, nfft, fs, window, numoverlap)


x: the discrete signal
nfft: number of FFT samples
fs: sampling frequency (Hz)
window: rectangle, Hanning, Hamming, etc.
numoverlap: number of overlaps

As an example the following is a generated two second signal, sampled at 10,000


samples/second whose instantaneous frequency is a triangle function of time.

fs = 10000; % sampling rate


t = 0:1/fs:2; % time duration
x = vco(sawtooth(2*pi*t,0.75), [0.1 0.4]*fs, fs); % signal
specgram(x, 512, fs, kaiser(256,5), 220)
5000

4500

4000

3500

3000
Frequency

2500

2000

1500

1000

500

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
Time

Figure 13.104 The spectrogram from the above example.

401
In LabVIEW a graphical programming language is used to program data acquisition and signal
processing. The graphical syntax is shown below.

Figure 13.105 LabVIEW graphical syntax for STFT.

The input parameters (shown on the left) are as follows.

X is the time waveform


time increment is the number of samples to shift the sliding window
window length is the actual length of the selected window
window selector determines the type of analysis window (0-rectangle, 1- Blackman, 2-
Hamming, 3- Hanning)

The output parameters (shown on the right) are as follows.

STFT Spectrogram {X} is a 2D array that describes the time waveform energy
distribution in the joint time-frequency domain
The number of rows (time axis) is equal to the number of elements in the time waveform
divided by time increment
The number of columns (frequency axis) in is equal to (window length/2 + 1)
error returns any error or warning

Figure 13.106 LabVIEW STFT interface and example.

402
Figure 13.107 A linear chirp signal and its STFT spectrogram.

Figure 13.108 Two linear chirps and their spectrogram.

Figure 13.109 A displacement signal and its STFT spectrogram.

403
A pump case study is now presented with sensors (two accelerometers left side, one optical
speed sensor-right side).

Figure 13.110 A pump case study (two accelerometers - left, one optical speed sensor - right).

Figure 13.111 The measured vibration signal.

Figure 13.112 The power spectrum shows the energy is a function of rotational frequency.

404
Figure 13.113 STFT spectrogram at constant speed.

In summary, the STFT (a.k.a. windowed Fourier transform) performs a series of FFTs over
selected segments of the signal. The transform result is represented as spectrogram which is the
square of STFT. The STFT can be interpreted in several ways. The choice of the window could
improve the resolution in time and frequency domain. Commonly used windows are rectangle,
Hanning, Hamming, Gaussian, Blackman and Kaiser. One a window size is determined and the
resolution across the time-frequency plane is fixed. According to the uncertainty principle, no
best resolution in both the time and frequency domain can be achieved simultaneously. In a
two-dimensional representation of STFT spectrogram, the horizontal corresponds to the time (or
rotation) information while the vertical axis displays the frequency information. The amplitude
can be represented in colour. Compared with other T-F transforms, STFT has a fixed window
size (i.e. fixed resolution).

13.6 Wigner-Ville Distributions (WVD)

The Wigner distribution was developed by Eugene P. Wigner for the area of quantum mechanics
in 1932. It was adopted to signal analysis by J. Ville in 1947. The WVD is a bilinear (non-linear)
transform. The resolution of the WVD is generally better than the STFT. WVD does not have
window effects.

Limitations of the WVD include cross-term calculations that may artificially give rise to
negative energy. Aliasing effects may distort the spectrum such that a high frequency
component may be incorrectly identified as a low frequency component.

The continuous definition of auto-WVD (one signal) is shown in equation 13.15.

13.15

where, s(t) is the time signal, s*(t) is the conjugate of s(t), t is the time index, is the frequency
index, and is the integral variable.

405
The definition of the continuous cross-WVD (one signal) is shown in equation 13.16.

13.16

where s(t) and g(t) are the time signals, g*(t) is the conjugate of g(t), t is the time index, is the
frequency index, and is the integral variable.

The auto-WVD can be viewed as the Fourier transform of the auto-correlation function R(t,) at
time t. (see equation 13.17).

13.17

The WVD of a signal s(t) at a particular time-frequency plane (t,) and depends not only on s(t)
but the signal far away from time t (i.e. ->). Therefore the localization of WVD is limited.

From the definition of the cross-WVD, the following applies.

13.18

For auto-WVD, the above relationship reduces to

13.19

This means the auto-WVD is a real-valued transform. The Gaussian-type signal is expressed as

13.20

The auto-WVD is then

13.21

13.22

The WVD of a Gaussian signal is centred at origin (0,0). A larger leads the less spread in the
time domain but larger spread in the frequency domain. The area of the contour ellipse for the

406
amplitude of e-1 is (the smaller the area, the higher the resolution). Compared with the STFT,
the resolution of the WVD is twice as higher as that of the STFT spectrogram.

Properties of the WVD include the fact that they are time-shift invariant. That is, if the signal s(t)
is shifted in the time domain to s0(t-t0), the WVD is shifted as well. WVDs0(t,) = WVDs(t-t0,).
Frequency modulation is also invariant. If the signal s(t) is modulated in the frequency domain,
s0(t)=s(t)exp{j0t}, then WVDs0(t,) = WVDs(t, - 0). The integration of WVD along the
frequency axis is equal to the signal power in the time domain (instantaneous energy). The
integration of WVD along the time axis is equal to the signal energy spectral density in the
frequency domain. The integration of the WVD across the time-frequency plane is equal to the
energy possessed by the original signal s(t). Mean instantaneous frequency (MIF) is defined as
the average frequency at a particular time. The relationship between MIF and the phase is s(t) =
A(t)exp{jt}, <>t = (t).

One of the limitations of the WVD is the generation of interference terms (or cross terms) when
analyzing multiple signals. The amplitude of the interference term may be twice as large as the
auto-WVDs. The interference terms may be interpreted as independent T-F components leading
to false judgements as to the contant of the signal. The interference terms prevent the WVD from
being widely used in real applications.

Suppose s(t)=s1(t) + s2(t). The WVD is then

13.23

In addition to the two auto-WVDs, a cross term WVDs1,s2(t,) is generated. The amplitude of the
cross-term is twice as large as that of the auto-WVDs. This is due to the bilinear nature of the
WVD.

Example I: Two sinusoidal signals: s(t) = exp(j1t) + exp(j 2t). The WVD is then

13.24

The first term contains the sum of two auto-WVDs at 1 and 2. The second term is the
interference term with =(1 + 2)/2, d = 1 - 2. The cross-term is in the middle of the two
auto-terms. The average of the cross-term is zero.

Figure 13.114 WVD spectrum of two sinusoids.

407
Example II: Two Gaussian signals: The signals are

13.25

The WVD is then

13.26

The first two terms are the non-negative auto-WVDs. The last term is the cross-term centred at
(t, ), midway between the auto-WVDs. t = (t1 + t2)/2, td = t1 - t2 and = (1 + 2)/2, d =
(1 - 2). The cross-term oscillates in both the time and frequency domain.

Figure 13.115 WVD spectrum of two Gausian signals.

To suppress the presence of the cross-terms, smoothing is used. Smoothing is basically a low-
pass filter that lets in the smoothed auto-WVDs while filtering out the cross-terms. The new
version is called smoothed WVD (SWVD). Smoothing suppresses the cross-terms but also
reduces the resolution.

13.27

For a discrete signal s[n], 0n<N, the discrete WVD is:


N 1
p p i 2kp
WVD[n, k ] s(n 2 )s (n 2 ) exp
p N
*

N

13.28

In a real calculation, the discrete signal s[n] needs to be interpolated (padded) with additional
zeros so as to make s[n] the size of 2N. Other fast computing versions are also available.

408
To suppress the interference terms generated by traditional WVD, many modified versions have
been put forward. In essence, almost all the modified versions focus on designing the appropriate
filters (t,). The Fourier transform of the filter is called the kernel function, denoted by (,).
Commonly used distributions include the following. These will not be discussed in detail here.

Choi-Williams distribution (CWD)


Cone-shaped distribution
Page distribution (PD)
Generalized WVD
Pseudo-WVD (PWVD)
Smooth pseudo-WVD (SPWVD)

Applications of the use of WVD in machine condition monitoring include gear damage
detection.

Figure 13.116 Gear damage detection undamaged gear.

409
Figure 13.117 Gear damage detection damaged gear.

The WVD result of an overhanging rotor vibration is shown below. The speed in this case is 660
900 RPM (11 -15 Hz). The 1st natural frequency is 800 RPM (13.3 Hz).

Figure 13.118 Schematic of overhanging rotor.

410
Figure 13.118 Rotor vibration signals and FFT.

Cross-term

Figure 13.119 The WVD result of the rotor signal.

411
Figure 13.120 The smoothed WVD result of the rotor signal.

In summary, WVD is based the Fourier transform of the auto-correlation function of a signal.
Compared with STFT, the WVD possesses many useful properties for signal processing
purposes. The resolution of the WVD is higher than that of the STFT. The major deficiency of
the WVD is the cross-term interference. The cross-terms are due to the nonlinear transform itself.
The cross-terms reflect the correlation of the auto-terms. To suppress the cross-terms, low-pass
filter smoothing may be applied. There are a variety of modified versions of the WVD which are
effective in suppressing the cross-term interference.

13.7 Wavelets

In conventional time-frequency methods (STFT, WVD), the resolution across the whole time-
frequency plane is fixed. The resolution is determined by the selected time window (rectangle,
Hanning, etc.). In the analysis of complex signals, different resolutions in the time-frequency
plane may be required.

The Wavelet transform (WT) was developed as a modification and alternative to the STFT. The
WT employs a series of time windows which are dilated and translated from an elementary
function. The elementary function is called the mother wavelet. The dilated (scaled) and
translated (time-shifted) functions are known as wavelets. When analyzing the low-frequency
information, the WT employs a long time window so as to obtain more precise low-frequency
information. When analyzing the high-frequency bands, the WT applies shorter time windows in
an effort to a complete picture of the high-frequency band components.

Figure 13.121 The STFT has a fixed resolution in both the time and frequency domain (equally
spaced tilings)

412
Figure 13.122 The WT has flexible resolutions in the time and frequency domain (unequally
spaced tilings)

It is important to note that although the WT has better resolution, it must obey the uncertainty
principle. The uncertainty principle dictates that a minimum area exists in all the time-frequency
windows. This implies the resolutions of WT can not go arbitrarily high.

STFT with a uniform bandwidth

WT increases bandwidth with frequency

Figure 13.123 Frequency bandwidth for STFT and WT.

The continuous wavelet transform (CWT) is defined as the sum over all time of the signal
multiplied by dilated and translated versions of the wavelet function (t) (i.e. mother wavelet).
The mother wavelet (t) is square integratible with a zero average. The wavelets are obtained by
dilating (a) and translating (b) the mother wavelet.

The term continuous in the CWT implies that the CWT can operate at every dilation (scale),
from that of the original signal up to some maximum scale. The CWT is also continuous in terms
of shifting. During computation, the analyzing wavelet is shifted smoothly over the full domain
of the analyzed function.

Dilating (scaling) means stretching a mother wavelet. The smaller the scale factor, the more
compressed the wavelet. Translating (shifting) means delaying (or hastening) its onset.

413
Figure 13.124 Dilating (scaling) means stretching a mother wavelet.

Figure 13.125 Translating (shifting) means delaying (or hastening) its onset

Although the CWT has to be computed digitally, the computation itself is not called the DWT.
The CWT is a redundant transform and requires a significant amount of computation time if
every possible scale is to be calculated. The DWT chooses only a subset of scales and positions
at which to make the transform. The DWT provides sufficient information both for analysis and
synthesis with a significant reduction in the computation time.

Figure 13.126 The decomposition process (h-low, g-high filters).

414
Procedures of DWT calculation include passing a discrete signal s(t) to both the low- and high-
pass filters. While filtering, a down-sampling is used which produces only half of the original
data from each filter. The decomposition process halves the time resolution while doubles the
frequency resolution. Repeat the above decomposition until the level where only two samples are
left from each filter.

The WT decomposes a signal into a series of wavelets (continuous) or coefficients (discrete).


The inverse process is to reconstruct the original signal from the decomposed wavelets or
coefficients. The reconstruction is also called the inverse wavelet transform (IWT). The WT
process starts by selecting appropriate wavelets for particular purposes. The design of wavelets
should satisfy several properties such as vanishing moments, support and regularity. Over the
years, several families of wavelets have been developed and proven successful for different
purposes.

The Harr wavelet was the first and simplest. It is discontinuous and resembles a step function.
Daubechies wavelets are compactly supported orthonormal wavelets.

Figure 13.127 Daubechies wavelets.

Figure 13.128 Symlets are similar to Daubechies in properties but more symmetrical.

Figure 13.128 Mexican hat wavelets are roportional to the second derivative of the Gaussian
probability density function.

415
Wavelets can be used for frequency identification. Figure 13.129 shows the determination of
pure sinusoidal signal components. The slow sine is identified in the approximation a4. The
medium sine is found in detail d4. The rapid sine is found in detail d1.

Figure 13.129 Frequency identification using wavelets.

Wavelets can also be used for signal de-noising. Figure 13.130 shows a Doppler-shifted sinusoid
with some added noise. Performing wavelet decomposition and then thresholding the transform
coefficients followed by restructuring the signal by thresholded transform coefficients can be
used to extract the signal.

Figure 13.130 Signal de-noising (approximation and details).

Figure 13.131 The de-noising result.

416
The following is a rolling element bearing fault diagnosis case study. The bearings are three
double-row self-aligning ball bearings (FAG 1204). The radial load is 500 N. The shaft rotation
speed is 1602 RPM or 26.7 Hz. The surface failures on the bearing are on the inner race, the
outer race and on one of the rolling elements.

FAG 1204 SKF 3207X FAG 1204


bearing test bearing
Figure 13.132 The test rig.

(a) small, (b) medium, (c) large


Figure 13.133 Inner race defects.

(a) small, (b) medium, (c) large


Figure 13.134 Outer race defects.

(a) small, (b) medium, (c) large

Figure 13.135 Rolling element defects.

417
Table 13.1 Characteristics of the tested bearings.

Table 13.2 Bearing characteristic frequencies.


Frequency (Hz) SKF 1204E FAG 1204
Shaft 26.7 26.7
Inner race defect 193.58 190
Outer race defect 126.82 130.39
Ball defect 120.34 135.98

(a) normal, (b) small, (c) medium, (d) large


Figure 13.136 Vibration spectra of inner race fault (190 Hz).

(a) normal, (b) small, (c) medium, (d) large


Figure 13.137 Vibration spectra of outer race fault (130 Hz).

418
(a) normal, (b) small, (c) medium, (d) large
Figure 13.138 Vibration spectra of ball fault (136 Hz).

From a comparison of the frequency the spectra the following conclusions can be made. Spectral
analysis is generally inadequate in identifying the bearing defects. The only exception is the
medium inner race defect (190 Hz). In the case of ball damage (136 Hz), there is even a decrease
in frequency response amplitude at the fault frequency. The amplitudes of both normal and faulty
bearings are comparable at fault frequencies.

Envelop spectrum Wavelet spectrum


Figure 13.139 Envelope and WT spectra of the normal bearing.

(a) small, (b) medium, (c) large


Figure 13.140 Envelope and WT spectra of the inner race fault.

419
(a) small, (b) medium, (c) large
Figure 13.141 Envelope and WT spectra of outer race fault.

(a) small, (b) medium, (c) large


Figure 13.142 Envelope and WT spectra of ball fault.

The previous figures allow the following brief conclusions to be made. The envelope spectrum
gives a clearer interpretation of the inner race defect (190 Hz).The envelope spectrum proves
insensitive to the small damage affecting the bearing outer race (130 Hz). The WT spectra are
sensitive to bearing faults in most cases.

In summary the following can be concluded of wavelet transforms (WT). Wavelet transforms are
seen as an alternative to Fourier transforms and STFTs. One interpretation of WT is that it breaks
up a signal into shifted and scaled versions of the original (or mother) wavelet. The WT employs
a flexible window in the time-frequency plane giving different resolutions. For the analysis of
low frequency band, a wider time window is used giving better frequency but poor time
resolution. For the analysis of high frequency band, a narrower time window is used giving
better time but poor frequency resolution. According to the uncertainty principle, there is a
minimum area in all the windows. The uncertainty principle dictates that the resolution can not
be obtained arbitrarily high enough in both domains simultaneously.

420
Two basic forms of WT are the CWT and the DWT. The CWT is a redundant transform and
continuous in terms of scales and shifting. The DWT is NOT the digital implementation of the
CWT. The DWT is performed on selected scales and positions usually based on powers of 2. The
DWT is faster and more efficient in computation. The original signal can be reconstructed from
both the CWT wavelets and the DWT coefficients. Several wavelet families can be chosen for
different analysis purposes. WT has found its applications in a variety of signal processing fields

13.8 Independent Component Analysis

Blind source separation (BSS) attempts to separate a mixture of signals into different sources
Blind implies the sources are not observable and that there is no exact prior information available
about the mixture. Independent Component Analysis (ICA) is one of the most computationally
efficient BSS methods. BSS and ICA are terms that are often used interchangeably. The ICA
assumes the linear mixture from a number of statistically independent sources. The objective of
the ICA is to separate the statistically independent sources from the mixed signals.

Mathematical modelling of two microphone recordings can be described as follows.


x1 a11s1 a12 s2

x2 a21s1 a22 s2

where x1, x2 are the recorded mixed signals (known), s1, s2 are speech signals (unknown) and a11
~ a22 are unknown parameters. The goal is to estimate s1 and s2.

The mixing process can be written as follows.


X AS
X [ x1 (t ),, xn (t )]T

where X is the observation of mixtures, A is the mixing matrix ( n-by-n) and S is the independent
source vector.
S [ s1 (t ),, sn (t )]T

The de-mixing process can be written as follows.


Y WX

Y [ y1 (t ), , yn (t )]T

X [ x1 (t ),, xn (t )]T
where Y is an estimation of source vector S, W is the de-mixing matrix (n-by-n) and X is the
mixture.

Independent component analysis requires that the sources be statistically independent. That is

421
E y1 y2 E y1 E y2

where E is the expectation parameter and y1, y2 are the variables. Independence also implies un-
correlation among variables. ICA is to give the uncorrelated estimates of the independent
sources.

Mixing A Demixing W
Observation Output
Source vector y
vector s vector x

Unknown process x As y Wx WAs

Figure 13.143 The mixing and de-mixing process.

Pre-processing involves centering and whitening. Centering removes the mean value from
the observation and helps to simplify the ICA algorithm. Whitening involves transforming the
covariance matrix into a unity matrix by eigenvalue decomposition (EVD).

E{xxT } I
The number of sensors required is normally larger than the number of sources. A small number
of sensors may result in incomplete separation. Super- or sub-Gaussian distributions can be
accommodated, but not Gaussian distribution (inseparable). The separation results and sources
are equivalent in the statistical sense (independent). No direct physical relationships (amplitude,
energy) exist between sources and estimations. The estimations keep the statistical characteristics
of the sources (i.e. shape and distribution). The order of estimations may be arbitrary.

Algorithms used for ICA include the Maximum likelihood method which separates the sources
using the maximum likelihood method. The Infomax principle is derived from neural networks
viewpoint and maximizes the output entropy of network outputs. The FastICA maximizes the
non-Gaussianity of the sources. ICA finds application in a wide range of fields. For example ICA
is used for signal extraction. medical engineering (EEG, ECG, MEG, fMRI, TMJ data
separation), image processing and feature extraction, telecommunications, financial market
analysis, and mechanical engineering.

Figure 13.144 Two-channel original signals.

422
Figure 13.145 The observed mixtures (measurements).

Figure 13.146 Estimations from the mixtures.

An example of ICA (BSS) can be seen in the following diesel engine noise separation.
Recordings were taken by microphones around a diesel engine. ICA performs the separation of
the recorded acoustic signals. Time frequency analysis was applied to post-process the separated
components. Several engine related sources could be identified.

Table 13.3 Typical engine noise sources versus crank angle.

Events
Cylinder #3 Cylinder #1 Cylinder #2 Cylinder #4
Combustion 0

Inlet Valve Open -13


(IVO)
Inlet Valve +39
Close(IVC)
Exhaust Valve -51
Open (EVO)
Exhaust Valve +13
Close (EVC)
Piston Slap (PS) +5

Fuel Injection (FI) -10

423
(a) The original signals (b) The estimated signals
4 4

IC #8 IC #7 IC #6 IC #5 IC #4 IC #3 IC #2 IC #1
#1
0 0
-4 -4
4 4

#2
0 0
-4 -4
4 4
#3
0 0
-4 -4
4 4
#4

0 0
-4 -4
4 4
#5

0 0
-4 -4
4 4
#6

0 0
-4 -4
4 4
#7

0 0
-4 -4
4 4
#8

0 0
-4 -4
-180 0 180 360 520 -180 0 180 360 520
Crank angle (degree) Crank angle (degree)

Figure 13.147 Noise recordings and ICA separation.

Figure 13.148 Time-frequency representation of the ICA representing the combustion in cylinder
#1 (Crank Angle: 0).

424
Figure 13.149 Time-frequency representation of the ICA representing the piston slap in cylinder
#1 (Crank Angle: +5).

Figure 13.150 Time-frequency representation of the ICA representing the fuel injection in
cylinder #1 (Crank Angle: -8~-10).

425
Figure 13.151 Time-frequency representation of the ICA representing the inlet valve opening in
cylinder #4 (Crank Angle: -13).

Figure 13.152 Time-frequency representation of the ICA representing the inlet valve closing #2
(Crank Angle: +39).

426
Figure 13.153 Time-frequency representation of the ICA representing the exhaust valve opening
in cylinder #3 (Crank Angle: -51).

In summary, BSS and ICA are novel signal separation methods. They attempt to recover the
latent sources from the measured mixtures. Sources are assumed to be statistically independent.
The mixing process is linear. Certain limitations restrict the wide applications of the BSSand
ICA. BSS and ICA are effective in extracting multi-channel observations in which independent
sources exist. The separation results may be further processed using time-frequency analysis and
wavelet transform. Applications of BSS and ICA can be found in many different areas such as
biomedical engineering, speech and image processing and finial applications.

427
Chapter 14

Non-Vibration based Machine Condition Monitoring Techniques


When deciding to employ machine condition montoring and diagnostic methods and strategies in
whatever form, consider first the Costs versus Benefits. All of the following items can be
considered as having an effect on the costs of the operation.

Purchase Prices and Training Skill Level Needed


Data Collection Costs Maintenance Costs
Data Storage/Retrieval Costs Convenience of use
Ease of Calibration, Repeatability Range of uses
Degree of forewarning Unexpected failure avoidance
Elimination of chronic faults Improvements in design
Improved customer relations Provision of reliability and maintenance data
Personnel viewpoint Improved competitiveness (less expensive, higher
quality product)

14.1 Visual Monitoring

Direct visual monitoring methods include the following. These (and the indirect methods listed
farther below) are relatively self-explanatory and will not be discussed further here.

human senses
microscopes, borescopes, stroboscopes.
extremely sensitive
often erratic
requires long experience
inexpensive (unless equipment is remote)
difficult to pass on skill
distractions are a factor
basically a breakdown strategy

Indirect visual monitoring methods include the following.

photographs, videos
radiographs
thermo graphs
x-rays

428
14.2 Performance Monitoring

Performance monitoring techniques are often already available, require some reasonable level of
experience and have widely variable sensitivity. The primary parameters that can be used include
pressure versus flow, temperature differences, cycle efficiencies and compression ratios.
Secondary performance parameters include mean effective pressure, fuel consumption, steam
consumption, exhaust gas composition, and charge air pressure.

Performance monitoring strategies include the variation in production rate, the variation in
product quality, input/output ratios, power in, power out, and efficiency.

14.3 Temperature Monitoring

Applications of temperature monitoring allow the condition of a wide range of machines and
components to be checked. These include the condition of lubricant and insulation as well as
operating conditions. This method is relatively inexpensive, can be implemented on-line, but is
not useful for detection of incipient damage.

Techniques are varied and include softening cones/wax/paint, thermography (to be discussed in
more detail later in this chapter), bimetallic strips, thermistors, thermocouples, vapour pressure in
bulbs, and mercury in glass.

The advantages include the fat that it is relatively inexpensive to employ and relatively low skill
levels are required.

Different types of heat transfer need to be considered. Conduction is the transfer of heat through
a material. Convection is the exchange of heat across a fluid/solid interface (forced and free
modes). Heat radiation can be absorbed, emitted, transmitted and reflected.

14.3.1 Thermography (thermal mapping)

Infrared (IR) sensors can be used for thermal mapping of a surface. They are now less expensive
and more reliable than even a few years ago. They can also be coupled with signal processing
software, which makes their output easier to analyze and store.

The main advantages of thermography are that it is non-intrusive, the sensor can be relatively
remote from the source, it is fast and measures temperature at the surface.

Applications where thermography is particularly useful include the following.

where the target is in motion


when the target is electrically charged (high voltage equipment)
if the target is fragile, particularly small or large
the target is remote (difficult to reach)
the target temperature is changing quickly

429
the target is destructive to thermocouples (jarring, burning, erosion)
if multiple measurements are required in a short time

Things to consider when using thremopgraphy include the following.

material type and surface conditions of the target


target size
temperature range
reflected sources that may be heating the surface
transmitted sources that may be heating the surface
ambient conditions (sun)
calibration

Figure 14.1 Single point thermography sensor.

Figure 14.2 Mapping thermography sensor.

430
Figure 14.3 Thermography cameras come in many shapes and sizes.

Figure 14.3 Thermographic image of a face. Why are the eyes different temperatures?

Figure 14.4 Heat sink with semiconductor mounted to it.

431
Figure 14.5 Electric motors.

Figure 14.6 Couplings, bearings, shafts.

432
Figure 14.7 Refractory deterioration, chimney stacks, pipelines.

Figure 14.8 Electrical equipment.

Figure 14.9 Images show poor electrical connection.

433
Week 1 Week 2 Week 3

Week 4 Week 5

Figure 14.10 Gearbox monitoring.

Figure 14.11 Severity criteria for mechanical components.

Figure 14.12 Turbine oil coolers - identical operating conditions


(oil flow or water flow problem in turbine #2).

14.4 Acoustic Emission (AE)

Acoustic emission (AE) is the elastic wave generated by the release of energy internally stored in
a structure. The surface detection of these internal stress waves is referred to as acoustic

434
emission. It has nothing to do with acoustic sounds or acoustic noise. The frequency range of
measurements is 100 2000 kHz. There are two types of acoustic emission. These are
continuous (steady state, random) and burst (transient, single decaying sinusoids due to
resonances in structure and/or transducer) type.

AE sources include dislocation movements in metal crystal structures, phase transformations and
crack formation and/or extension as well as friction in cracks. Dislocation movements are a
displacement of a line imperfection through the crystal lattice of the structure of a metal. These
AE events are of very small amplitude. Many sequentially may cause a continuous type signal.
Phase transformations are when some metals are stressed they undergo changes to their crystal
lattice structure. These events are at least 100 times the amplitude of dislocations. They typically
occur in single bursts. Crack formation and extension results in new surfaces being formed.
Strain energy is released during these events. This energy is partly transformed into an AE
signal. It is a burst type signal and is often at a high rate and of the same amplitude as phase
transformations. Friction in cracks causes sudden sliding which releases burst type AE signals.
These signals are useful for detecting and locating cracks.

AE Peak Amplitude
AE count
AE Event

Threshold Level

Continuous Background Noise


Burst Emission Emission

Figure 14.13 Typical AE signal (continuous and bust events).


-3
x 10
3

1
Magnitude (Volts)

-1

-2

-3
0 0.2 0.4 0.6 0.8 1
Time (Seconds) -3
x 10

Figure 14.14 Typical AE signal (continuous).

435
-3
x 10
3

Magnitude (Volts)
0

-1

-2

-3
0 0.2 0.4 0.6 0.8 1
Time (Seconds) -3
x 10

Figure 14.15 Typical AE signal (burst)

AE signal propagation depends on signal source strength material properties, geometry and
external strain rates. Symmetrical waves are generated in isotropic, homogenous, ideally elastic
mediums. Surfaces and discontinuities on surfaces create reflections and surface waves.
Transducers are a special type (piezoelectric crystal based) and require special mounting. They
operate in the 100 kHz to +2 MHz frequency range. There are broad band and narrow band
(resonance type) accelerometer based sensors available A laser vibrometer can also be
configured to measure AE signals.

AE signal analysis typically requires a high speed digital oscilloscope. For purely continuous
signals measuring the RMS value with a voltmeter is sufficient. For burst signals a counter (total
sum and/or rate) is required. If using a tape recorder it needs to be able to collect high speed
(high frequency data. Transient recorder may be used as peak counters and in this way detect (ad
count burst AE events. Standard frequency analysis is then done on the recorded time waveform.

Applications for SAE include general plant machinery condition monitoring, partial discharge
detection and location in high voltage electrical equipment, crack location, and materials testing
(AE signature prior to failure).

Some advantages of using AE include the fact that it is good for remote detection and location of
flaws, one (or a few measurements) will cover an entire structure, the sensors are quick to set-up,
the sensors have high sensitivity, limited access to machinery and structures is required,
relatively low loads are required, and it can be used to forecast failure loads.

The disadvantages include the fat that the structure must be loaded (at least somewhat), AE
signals are highly material specific, noise can be a problem (electrical and mechanical), it has
limited location accuracy, limited type of flaw information is available and the signal analysis
interpretation may be difficult.

436
Figure 14.16 AE used for leak detection.

Figure 14.17 Acoustic Emission best measurement locations.

14.5 Oil Quality Analysis

14.5.1 Lubrication Fundamentals

Friction and wear reduction (separates parts moving relative to one another) is the main reason
(benefit) of using lubrication. There are also other benefits such as heat control, contamination
control, reduced chemical attach, and the transfer of energy (in hydraulic systems).

437
Figure 14.17 The role of lubrication.

14.5.2 Base Stock Considerations

Base stock properties largely determine the lubricant performance. Most base oils are mineral
oils (refined from crude oil). Fractional distillation sorts hydrocarbon molecules by size and type.
Other refining involves solvent extraction, hydrogen processing, acid refining, clay refining, and
solvent de-waxing.

Table 14.1 Key physical properties of new base oils.

Property Why Important How Determined ASTM No.


Viscosity Defines base oil viscous Gravity flow capilary D-445
grade viscometer
Viscosity Index Defines viscosity- Viscosity variance between D-2270
temperature relationship 40C and 100C (indexed)

Specific Gravity Defines density of oil Hydrometer D-1298


relative to water
Flash Point Defines high-temp Flash point tester (temp at D-92/D-93
volatility and which flash surface flame is
flammability properties achieved)

Pour Point Defines low-temp oil Gravity flow in test jar D-97/IP-15
fluidity behaviour

438
Table 14.2 Typical base oil properties.

Property ASTM Paraffinic Oil Naphthenic Aromatic Oil


Method Oil
Viscosity @ D-445 40 40 36
40C
(Pas )
Viscosity @ D-445 6.2 5 4
100C
(Pas )
Viscosity D-2270 100 0 -185
Index
Specific D-287 0.8628 0.9194 0.9826
Gravity
Flash Point C D-92 229 174 160
Pour Point C D-97 -15 -30 -24
% Paraffinic D-3238 66% 45% 23%
% Naphthenic D-3238 32% 41% 36%
% Aromatic D-3238 2% 14% 41%

14.5.3 Mineral versus Synthetic Oils

Potential benefits of using synthetic base oil include increased oxidative life, improved lubricity,
fire resistance, better thermal resistance, and extended drain intervals.

Potential limiting factors of using synthetic base oil include high purchase cost, seal or coating
incompatibility, potential toxicity, high disposal cost, possible incompatibility with mineral oil.

14.5.4 Hydrostatic, Elasto-Hydrodynamic and Boundary Lubrication

Hydrostatic lubrication needed is when surface-to-surface contact is likely due to high loads or
contamination. It is also called thin-film lubrication. A wedge of lubricating oil is created
hydrodynamically. The performance depends on machine speed, load, interacting surface shape,
and oil viscosity. Hydrostatic lubrication is not available at start-up and coast-down (low speeds
will not form the thin film wedge.

439
Figure 14.18 Hydrostatic lubrication.

Elasto-hydrodynamic lubrication takes place in rolling element bearings and between gears. The
small contact area creates very high local pressure. The liquid oil may change into a solid state
temporarily, then change back to liquid state after pressure is removed.

Boundary lubrication is the lubricant sticking to the surface of the machine components and
acting to separate moving parts. It occurs when component surface roughness is high, there are
frequent stops and starts, shock loading conditions, high static loads, or low speeds. Additives
and/or solid lubricants may help.

14.5.5 Additives and Their Functions

Additives enhance existing base oil properties. They suppress undesirable base oil properties.
They also impart new properties to base oils.

Antioxidants or oxidation inhibitors reduce oil reaction with oxygen at high temperatures. These
additives act to reduce aeration (air mixed into oil). Because oxidation rates increase with
increasing temperature, it helps to keep machine temperatures as low as possible. The oxidation
rates typically double with every 10 C increase in temperature. These additives also reduce
water in oil and reduce metal catalysts (copper, lead, iron).

Oxidation results in increased viscosity, increased acidity, increased specific gravity, darkened
colour, varnished component surfaces, and sludge accumulation. These effects cannot be
stopped, only delayed.

Rust inhibitors are additives that form a barrier to protect steel (iron) from water.

Dispersants act to envelop sludge and soot particles to stop them from joining together into
larger particles that would degrade the lubricant. Detergents cleanse high temperature surfaces
(pistons, rings, valves, etc.).

440
Figure 14.19 Rust inhibitor.

Figure 14.20 Dispersants and detergents.

Anti-wear and extreme pressure additives are also called anti-scuff additives. Under high
pressure these additives react with component surfaces to form soap-like oxide films that
enhance lubricity at the boundary contact between surfaces.

Figure 14.21 Anti-Wear and Extreme Pressure Additives

441
14.5.6 Oil Quality Parameters

Oil quality parameters that can be measured and give an indication of how well an oil is able to
perform its designed function are viscosity, acidity, flashpoint, dilution (fuel leakage into oil),
total solids in oil, and water in oil.

14.5.7 Water in Oil

Dissolved water is not good for oil The of water that can dissolve in an oil depends on the oil
base stock, the additives present, the contaminant load and the temperature. Emulsified water is
water in the form of small globules that remain suspended. Free water is water that settles to
tank/sump bottom.

Water enters a system from several sources and through several different ways. Atmospheric
water is free water from the working environment (rain, sprays, coolant) that enters through
seals, vents, hatches or in new oil. Condensation occurs during frequent machine starts and stops
when the temperature increases and decreases often. Water can dissolve in oil and then condense
out. There could also be coolant leakage (if the coolant is water).

The effect of water contamination on the oil is that is reacts with additives to form precipitates
and aggressive chemical by-products. These then may act as catalysts to oxidation. Micro-
organisms can live in water and feed on oil. They often decompose to form acids and cause filter
clogging.

The effect of water contamination on a machine is that it interferes with lubrication by


weakening the strength of the oil film. Water causes rust on iron and steel. Water also increases
the corrosive effect of acids.

Water contamination can be controlled by using settling tanks. These are quiet areas where water
can settle out of oil. Heating is sometimes applied to encourage separation of water and oil.
Centrifugal separators accelerate the settling process. Vacuum distillation can be used to lower
pressure and therefore lower the water boiling temperature. Water is separated from oil without
damaging oil with high temperatures. Polymeric filters impregnated with a super absorbent
polymer can remove water from oil.

Figure 14.22 Calcium hydride test kit.

442
Figure 14.23 On-line impedance-type moisture sensor.

Figure 14.24 On-line headspace moisture sensor.

Figure 14.25 Single-channel infrared moisture sensor.

Figure 14.26 Desiccant air breathers.

443
14.5.8 Fuel in Oil

Fuel in oil occurs primarily in crankcase applications. Fuel dilutes oil during engine operation.
Fuel in oil may be caused by delayed oil changes, improper operation, and malfunction. Fuel
enters a system during piston blowby. Blowby is due to incomplete or ineffective combustion,
excessive idling, lugging, defective fuel injection spray pattern, or improper fuel/air ratio.

Fuel contamination affects oil by causing premature oxidation (fuel oxidizes easily), loss of
viscosity (thinning of oil), diluting oil additives, and sulfur build-up which increases the risk of
corrosion. The effect on the machine of fuel contamination is to increase wear due to loss of
viscosity and additive dilution. increase corrosion due to sulfuric acid generation, and increase
the risk of fire and explosion (due to the presence of highly flammable fuel in the oil). Fuel
contamination can only be controlled if it is stopped at its source. Once contaminated with fuel
the oil must be replaced.

14.5.9 Soot in Oil

Soot is a natural by-product of combustion. Poor combustion leads to soot accumulation. Soot
enters a system during piston blowby. Soot in blowby is caused by low compression (resulting in
poor combustion), high fuel/air ratio (resulting in poor combustion), cold air temperatures,
lugging, and excessive idling.

Soot contamination affects the oil by causing the dispersant additive to loose effectiveness, anti-
wear additives to loose effectiveness and to increased viscosity. The effect on the machine of
soot contamination is to cause filter plugging, increased abrasive wear and deposits of sludge
that could cause flow blockages. Controlling soot contamination can be achieved by improving
fuel combustion and general operation of an engine as well as by having low flow and slip-
stream filters and centrifuges to remove soot from oil.

14.5.10 Glycol in Oil

Glycol contamination is particularly common in crankcase applications, but may affect any
system that is using a glycol/water mix for cooling. Glycol enters a system through defective
seals, if cavitation and erosion of components has changed the shape of these components,
corrosion and a damaged cooler core.

Glycol contamination affects oil by causing gels and emulsions to form. It increases viscosity,
increases oxidation, forms acids and oil balls (glycol droplets collect oil into balls). The effect
of glycol contamination on a machine is to increase wear by reducing the lubricating quality of
the oil due to increased viscosity, increasing corrosion by creating acids, and causing filter
failure by clogging filters. Controlling glycol contamination can only be done by controlling the
ingestion of glycol into the system. Once it is in the system glycol cannot be effectively removed
from oil.

14.6 Oil Sampling

The methods for sampling oil include the following.

444
Drain Port Sampling
Drain Port Vacuum Sampling
Drain Line tap Sampling
Portable Off-Line Sampling
Dedicated Off-Line Sampling
Probe-On Vacuum Sampling
Drop Tube Vacuum Sampling

Pressurized line sampling methods include the following.

Low Pressure Tap Sampling


Portable Tap Sampling
Low Pressure Ball Valve Sampling
High Pressure Sampling

In general poor oil samples lead to poor decisions based on oil analysis. Oil properties that may
be considered to be unaffected by sample location include: viscosity, pH (neutralization number),
oxidation, sulfation, nitration, and additive levels. Oil properties that should be considered as
being affected by sample location include: particle count, moisture level, and wear levels.

14.6.1 General Guidelines for Drawing Samples

Sample while operating under normal operating and environmental conditions. Sample from
zones of active circulation. Sample upstream of filters and downstream of components. Use
sampling hardware that minimizes interference with operations and allows samples to be drawn
from the same place each time. Flush sample ports thoroughly before collecting samples. Use
clean bottles and tubing. Sample at the proper frequency. Record the sampling time. Analyse the
oil immediately. Educate staff about the importance of high-quality oil samples. Train staff on
effective sampling procedures.

14.6.2 General Safety Considerations

Conform to all safety requirements of the plant. Check oil temperature prior to sampling. If oil
temperature is above 120F (50C) use proper protective gloves and clothing. Wear protective
eyewear. Wear latex gloves to protect hands. Wash any skin that comes into contact with oil with
a high quality industrial cleaner and plenty of water. Seek medical attention if oil gets into the
eyes or is ingested.

14.6.3 Sources of Error in Data

Samples taken under cold (machine not running) conditions. Drain port sampling without
flushing. Variable sampling methods and locations. Contaminated sampling hardware.

445
Insufficient dead space flushing. Sampling just after an oil change. Cross contamination between
two oil samples. Delays in sending samples to the testing laboratory.

14.6.4 Recommended Oil Sampling Frequencies

Table 14.3 Recommended oil Sampling frequencies.

Machine Type Hours


Diesel Engines off highway 150
Transmissions, differentials, final drives 300
Hydraulics mobile equipment 200
Gas turbines industrial 500
Steam turbines 500
Air/gas compressors 500
Chillers 500
Gear boxes high speed/duty 300
Gear boxes low speed/duty 1000
Bearings journal and rolling element 500
Aviation reciprocating engines 25-50
Aviation gas turbines 100
Aviation gear boxes 100-200
Aviation hydraulics 100-200

14.6.5 Information to be Supplied with Oil Sample

Machine I.D.
Sample point
Date sampled
Running conditions
Hour meter reading (machine usage)
Last oil change
Last addition of oil (and amount)
Last major service, repair or overhaul
Last filter change
Operator reported machine/oil abnormalities
Requested test on oil sample

446
14.6.6 Drain Port Sampling

As this sample is typically taken from the bottom of the tank, there is a high possibility of
obtaining a high particle count due to sedimentation.

Figure 14.27 Drain Port Sampling.

Table 14.4 Drain Port Sampling.

Lubrication System or Machine Type Drain Port Sampling


Bath-Lub Plain Bearings Less Desirable
Ring Lub Plain Bearings Less Desirable
Inverted Bottle Bath Oiler Less Desirable
Vented Bottle Bath Oiler N/A
Splash Lub Gearing Less Desirable
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Less Desirable
Hydraulic System Less Desirable
Screw Compressor Less Desirable
Large Circulating System Less Desirable
Crankcase Lub Less Desirable

447
14.6.7 Drain Port Vacuum Sampling

Sample is taken from a permanent sampling fixture. An example of this type of sampling fixture
is a Minimess. The sample can be taken from a strategic location to avoid unrepresentative
amounts of water or debris. There is no need to remove a drain plug, vent or fill port each time.

Figure 14.28 Drain Port Vacuum Sampling.

Table 14.5 Drain Port Vacuum Sampling.

Lubrication System or Machine Type Drain Port Vacuum Sampling


Bath Lubricated Plain Bearings Best
Ring Lubricated Plain Bearings Best
Inverted Bottle Bath Oiler Best
Vented Bottle Bath Oiler N/A
Splash Lubricated Gearing Best
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Best
Hydraulic System Less Desirable
Screw Compressor Good
Large Circulating System Less Desirable
Crankcase Lubrication Good

448
14.6.8 Drain Line Tap Sampling

Figure 14.29 Drain Line Tap Sampling.

Table 14.6 Drain Line Tap Sampling.

Lubrication System or Machine Type Drain Line Tap Sampling


Bath Lubricated Plain Bearings N/A
Ring Lubricated Plain Bearings N/A
Inverted Bottle Bath Oiler N/A
Vented Bottle Bath Oiler Best
Splash Lubricated Gearing N/A
Circulating System Dry Sump Best
Circulating System Wet Sump N/A
Hydraulic System Best
Screw Compressor N/A
Large Circulating System Best
Crankcase Lubrication N/A

449
14.6.9 Drain Line Vacuum Sampling

Sample comes from a common line. Good representation of machine overall condition.

Figure 14.30 Drain Line Vacuum Sampling.

Table 14.7 Drain Line Vacuum Sampling.

Lubrication System or Machine Type Drain Line Vacuum Sampling


Bath Lubricated Plain Bearings N/A
Ring Lubricated Plain Bearings N/A
Inverted Bottle Bath Oiler N/A
Vented Bottle Bath Oiler Best
Splash Lubricated Gearing N/A
Circulating System Dry Sump Best
Circulating System Wet Sump N/A
Hydraulic System Best
Screw Compressor N/A
Large Circulating System Best
Crankcase Lubrication N/A

450
14.6.10 Portable Off-Line Sampling

Good for machines without return line sampling (gearboxes, engines, wet-sump lubrication
systems). After sample is taken the oil remaining in the machine can be filtered or conditioned
by portable apparatus.

Figure 14.31 Portable Off-Line Sampling.

Table 14.8 Portable Off-Line Sampling.

Lubrication System or Machine Type Portable Off-Line Sampling


Bath Lubricated Plain Bearings N/A
Ring Lubricated Plain Bearings N/A
Inverted Bottle Bath Oiler N/A
Vented Bottle Bath Oiler N/A
Splash Lubricated Gearing Good
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Good
Hydraulic System Less Desirable
Screw Compressor N/A
Large Circulating System Less Desirable
Crankcase Lubrication N/A

451
14.6.11 Dedicated Off-Line Sampling

Figure 14.32 Dedicated Off-Line Sampling.

Table 14.9 Dedicated Off-Line Sampling.

Lubrication System or Machine Type Dedicated Off-Line Sampling


Bath Lubricated Plain Bearings N/A
Ring Lubricated Plain Bearings N/A
Inverted Bottle Bath Oiler N/A
Vented Bottle Bath Oiler N/A
Splash Lubricated Gearing Best
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Best
Hydraulic System Less Desirable
Screw Compressor N/A
Large Circulating System Less Desirable
Crankcase Lubrication N/A

452
14.6.12 Probe-On Vacuum Sampling

Requires permanent sampling port. Good for sampling tanks, sumps and static containers.
Sample should be taken from strategic location. No need to remove/replace drain plug or vent.
Requires sampling with sample bottle inside a plastic bag to reduce risk of contamination.

Figure 14.33 Probe-On Vacuum Sampling.

Table 14.10 Probe-On Vacuum Sampling.

Lubrication System or Machine Type Probe-On Vacuum Sampling


Bath Lubricated Plain Bearings Best
Ring Lubricated Plain Bearings Best
Inverted Bottle Bath Oiler Best
Vented Bottle Bath Oiler N/A
Splash Lubricated Gearing Best
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Best
Hydraulic System Less Desirable
Screw Compressor Good
Large Circulating System Less Desirable
Crankcase Lubrication Good

453
14.6.13 Drop-Tube Vacuum Sampling

Common method for sampling tanks and sumps. Access to tank from fill port. No permanently
installed hardware needed.

Figure 14.34 Drop-Tube Vacuum Sampling.

Table 14.11 Drop-Tube Vacuum Sampling.

Lubrication System or Machine Type Drop-Tube Vacuum Sampling


Bath Lubricated Plain Bearings Avoid
Ring Lubricated Plain Bearings Avoid
Inverted Bottle Bath Oiler Avoid
Vented Bottle Bath Oiler Less Desirable
Splash Lubricated Gearing Avoid
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Less Desirable
Hydraulic System Less Desirable
Screw Compressor N/A
Large Circulating System Less Desirable
Crankcase Lubrication Less Desirable

454
14.6.14 Pressurized Line Sampling Low Pressure Tap Sampling

Can be used to detect contamination from breather vents and new oil additions. Good for
detecting wear generated by the oil pump. For systems without pressure-side, in-line filters, this
method can help assess the cleanliness of the oil being fed to the machine components. Provides
background for comparing return line samples. For systems with in-line filters this provides a
background for comparing filter performance. Good for wet-sump circulating systems that do not
allow drain line sampling. Requires sampling with sample bottle inside a plastic bag to reduce
risk of contamination.

Figure 14.35 Pressurized Line Sampling Low Pressure Tap Sampling.

Table 14.12 Pressurized Line Sampling Low Pressure Tap Sampling.

Lubrication System or Machine Type Pressurized Line Sampling


Bath Lubricated Plain Bearings N/A
Ring Lubricated Plain Bearings N/A
Inverted Bottle Bath Oiler N/A
Vented Bottle Bath Oiler N/A
Splash Lubricated Gearing Best
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump N/A
Hydraulic System Less Desirable
Screw Compressor Avoid
Large Circulating System Less Desirable
Crankcase Lubrication Avoid

455
14.6.15 Pressurized Line Sampling Portable Tap Sampling

Requires sampling with sample bottle inside a plastic bag to reduce risk of contamination.

Figure 14.36 Pressurized Line Sampling Portable Tap Sampling.

Table 14.13 Pressurized Line Sampling Portable Tap Sampling.

Lubrication System or Machine Type Pressurized Line Sampling

Bath Lubricated Plain Bearings N/A


Ring Lubricated Plain Bearings N/A
Inverted Bottle Bath Oiler N/A
Vented Bottle Bath Oiler N/A
Splash Lubricated Gearing N/A
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Best
Hydraulic System Best
Screw Compressor Best
Large Circulating System Best
Crankcase Lubrication Best

456
14.6.16 Pressurized Line Sampling Low Pressure Ball Valve Sampling

Figure 14.37 Pressurized Line Sampling Low Pressure Ball Valve Sampling.

Table 14.14 Pressurized Line Sampling Low Pressure Ball Valve Sampling.

Lubrication System or Machine Type Pressurized Line Sampling


Bath Lubricated Plain Bearings N/A
Ring Lubricated Plain Bearings N/A
Inverted Bottle Bath Oiler N/A
Vented Bottle Bath Oiler N/A
Splash Lubricated Gearing N/A
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Best
Hydraulic System Best
Screw Compressor Best
Large Circulating System Best
Crankcase Lubrication Best

457
14.6.17 Pressurized Line Sampling High Pressure Sampling

Figure 14.38 Pressurized Line Sampling High Pressure Sampling.

Table 14.15 Pressurized Line Sampling High Pressure Sampling.

Lubrication System or Machine Type Pressurized Line Sampling


Bath Lubricated Plain Bearings N/A
Ring Lubricated Plain Bearings N/A
Inverted Bottle Bath Oiler N/A
Vented Bottle Bath Oiler N/A
Splash Lubricated Gearing N/A
Circulating System Dry Sump Less Desirable
Circulating System Wet Sump Best
Hydraulic System Best
Screw Compressor Best
Large Circulating System Less Desirable
Crankcase Lubrication Best

14.7 Wear Particle Analysis

Wear particle analysis is good for direct detection, diagnosis, prognosis in many systems. It
requires ease of access to the oil supply. Skill requirements are high. Care must be taken during
sampling. On-line systems are possible.

Damage to a system depends on particle size, shape, hardness and chemistry. Particles enter from
vents and breathers, ineffective or damaged seals, new oil, and filters that are full, damaged or

458
defective. Wear particals can alter oxidation rates, strip oil of additives and increase oil viscosity
(many small particles in suspension). Wear particles can also increase wear rates in a machine,
reduce oil film thickness, damage or cut away component material, increase local loads between
mating surfaces, and cause erosive wear (at high velocity).

14.7.1 Wear Particle Generation (Types of Particles)

Rolling wear occurs when small contacting areas are exposed to extremely high pressures and
these pressures cause sub-surface fatigue cracks to develop and worsen until small particle break
free of the surface. Rolling element bearings and gears commonly display this type of wear.

Figure 14.39 Rolling wear due to fatigue spalling in a gear.

Figure 14.40 Surface fatigue causing spalling (pitting).

Sliding wear is also known as scuffing wear and is caused by metal-to-metal contact (no
lubricant film present). It happens between gear teeth, on pistons, in hydraulic systems, and in
loose bearings. Parallel striations are evident on the particle surface. This type of wear may be
caused by lubricant loss, overloading, or the wrong oil. Particles are typically 5 m to 5000 m
(15 - 50 m usual) in size.

459
Figure 14.41 Scuffing (sliding wear) on a gear.

Figure 14.42 Severe sliding wear (striations on particle).

Cutting wear is caused by abrasive contact between two surfaces. During two body wear
(metal/metal) the softer metal is cut. During three body wear (metal/contaminant/metal) the
harder metal is cut. The wear particles are curls of metal like lathe swarf. The particles are
typically 5 m to 5000 m (10-85 m usual) in size. This type of wear particle is generated in
situations where there has been poor machining of parts, contaminants in the oil, or
misalignment.

Figure 14.43 Two-body and three-body cutting wear.

460
Figure 14.44 Cutting wear (x400 magnification).

Laminar wear is due to sub-surface stressing in rolling element bearings. It may also be seen in
reworked wear particles that have traveled through a machine several times. These particles are
flat, thin flakes of metal. These particles are 5 m to 200 m (15 - 65 m usual) in size. This
type of wear is caused by bearing spalling and poor filtration.

Figure 14.45 Wear particles.

Figure 14.46 Copper alloy particle before run-in (left) and after run-in (right).

461
Rubbing wear is considered normal wear. It is made up of small flakes of metal. The particles are
typically less than 10 m in size and occur throughout the life of a machine.

Fatigue chunks result from serious sub-surface stressing in gears or on shafts. The chucks are
quite large (up to 5000 m). They occur during periods of severe machine stress. In rolling
element bearings, chunks are usually rolled flat or broken into smaller sizes (see laminar wear).

Spheres are caused by friction reworking of other wear particles. These particles are small ball-
bearing like spheres. They are typically 5 m to 30 m (5 - 15 m usual) in size. They occur
only during periods of extreme friction (high heat).

Figure 14.47 Spheres (500x optical).

Oxides are caused by the heating of wear particles. These particles can be any shape. The size of
these particles is typically from 5 m - 5000 m (15 - 50 m). They occur during periods of high
temperature, typically due to the wrong oil being used or machine overloading, or no oil.

14.7.2 Controlling Particle Contamination - Filters

Filter stability over time is a function of temperature, pressure fluctuations, and vibrations. Filter
capacity (amount of contaminant filter can remove) and filter efficiency are both critical factors.

Figure 14.48 Filter efficiency.

462
Particle separation methods (filters) come on several different types. Cellulose fiber media filters
are good for larger particles. Micro-fiberglass media filters are superior to cellulose fiber filters.
Centrifugal separators and electro-static separators can also be used.

14.7.3 Origins of Particles

Environmental origins of wear particles (contaminants) come from :


dust contamination
moisture contamination
fumes
careless maintenance
improper repairs
malicious acts
biological contamination

Generated wear particles come from:


standard wear modes
mechanical interference
material incompatibility
surface fatigue

Figure 14.49 Many wear particles come from bearings and gears (severely worn cylindrical roller
bearing left, severely worn spur gear middle, severely worn spur gear right).

Figure 14.50 Severe wear on bearing raceways.

463
14.7.4 Trending using Wear Particles

Oil quality can be used as a trending parameter and indicate machine condition.

Figure 14.51 Metal concentrations in oil as a trending parameter.

Figure 14.52 Oil viscosity as a trending parameter.

Figure 14.53 Oil additives as trending parameters.

464
Figure 14.54 Wear Rate vs. Particle Angularity

Table 14.16 Wear and contamination particles numbers (from ISO 4406 (1999) oil cleanliness
coding).

465
Figure 14.55 Typical progression of wear particle generation

Figure 14.56 Typical wear particle generation profile for small particles

Figure 14.57 Typical wear particle generation profile for large particles

466
Figure 14.58 Gear system operating regimes as a function of speed and load

14.7.5 Collection Techniques

Magnetic plugs - Ferrous only wear particles


Filters - wear particles and contaminants
Nephelometer - all particles and mixed fluids
Lab tests - oil quality, wear and contamination particles and fluids
Can also be applied to grease

Figure 14.59 Magnetic Plug.

Figure 14.60 Fatigue chunks.

467
Figure 14.61 Chunks, flakes and fines.

14.7.6 Particle Counting

Particle counters work by passing a diluted oil sample through a device that shines a light or
laser through the sample. Any blockages that are detected as the sample passed through are
counted as particles. All particles are counted, regardless of what they are. Particle counters
typically give size distribution information.

Figure 14.62 Particle counter.

14.7.7 Visual Analysis Methods - Analytical Ferrography

Analytical ferrography is when a diluted sample of oil is passed over a magnetic separation plate.
The ferrous particles (but only the ferrous particles) are captured by the magnet. Rotary particle
depositors are also common. The collected samples of partiles are then viewed under an optical
microscope and the particles are identified by their colour, size, and shape.

468
Figure 14.63 Particles, descriptions and likely type.

Wear debris image capture systems can now be quite sophisticated. The first ferrograph was
developed in 1971. It consisted of a pump, magnet, and a slide upon which the particles were
deposited. An excellent description of this technique is given by Roylance, B.J., Ferrography
Then and Now. Tribology International Vol. 38 Issue 10, Oct 2005 pp. 857-862. Analytical
Ferrography can show amounts and sizes of case hardened steel, low alloy steel, and medium
alloy steel abnormal gear and bearing wear particles up to 120 microns in size.

Figure 14.64 Wear debris image capture system: 1) optical microscope with CCD digital camera
2) additional color monitor 3) computer 4) display 5) color printer.

469
Table 14.16 potential information available from different oil analysis methods.

Method Size range Quantity Size Distribution Morphology Composition


Spectrometric <10
Particle Counting 1 -150
Magnetic Plug 100 -1000
Ferrograph 1 -100

Figure 14.65 Ferrogram maker.

Figure 14.65 Low and high alloy steel gear and bearing wear particles (120 m max.) 200X
ferrogram.

470
Figure 14.66 Wear particles from ferrograms.

Filter debris analysis can also be used to investigate wear particles in oil. The filter separation is
dependant on the filter pore size of course. Filtergrams are similar to ferrograms, except that they
separate out all the particles, not only the iron particles. The filter separates out the particles and
then the filter paper is made transparent through a clarification process. The filter and wear
particles are then viewed under an optical microscope.

A scanning electron microscope (SEM) may also be used. They see all particles, but sample
analysis can be expensive.

Figure 14.67 Soot particle under traditional microscope (left) and under a SEM (right).

14.7.8 Spectrametric Oil Analysis

Elemental analysis can detect up to 24 elements, measuring less than 5 m, that can be present in
used oil due to wear, contamination or additives.

471
Appendix A: Fault Tables

Table A.1 Forcing Function Faults

Spectral Spectrum,
Fault Correction
Frequency Time Waveform
Critical speeds 1x, 2x, 3x, etc. Amplified vibration due to Tune natural frequency
proximity of operating speed or change operating
to natural frequency speed
Mass unbalance 1x Distinct 1x with much lower Field or shop balancing
values of 2x, 3x; vertical/
horizontal 90 out of phase
Misalignment 1x, 2x Distinct 1x with equal or Perform hot and/or
occasionally 3x higher values of 2x, 3x; cold alignment
vertical/horizontal in phase
Shaft bow 1x Dropout of vibration around Heating or peening to
critical speed in Bod plot; straighten rotor (allow
1x constant with speed rotor to float axially)
Fluid-film bearing 1x, sub- High1x, 1/2x, sometimes or Replace bearing
wear and excessive harmonics, orders; cannot be balanced
clearance orders
Resonance 1x, 2x, 3x, etc. High balance sensitivity, high- Change structural
amplitude vibration at order of natural frequency
operating speed
Looseness 1x 2x, 3x, etc., High 1x with lower-level Shim and tighten bolts
1/2x may occur orders, large1/2 order, low to obtain rigidity
axial vibration
Eccentricity 1x High 1x Machine journal for
concentricity
Coupling lockup 1x, 2x, 3x, etc. 1x with high 2x similar to Replace coupling or
misalignment; start and stops remove sludge
may yield different vibration
patterns
Thermal variability 1x 1x has varying phase angle and Compromise balance
amplitude with load
Distortion 1x and orders 1x from preload of bearings, Relieve soft foot
2x line frequency, air gap on
motor
Impact-excited Nx Excitation of natural frequency Remove source of
natural frequency at multiple of operating speed pulses or impacts
Sheave eccentricity 1x Spectral content depends on Replace sheave
pitch line profile.
Horizontal/vertical phase

472
Table A.2 Rolling Element Bearing Defects

Spectral Spectrum,
Fault Comment
Frequency Time Waveform
Outer race defect BPFO and Multiples of BPFO with Remaining life sensitive
multiple sidebands of RPS, FTF, or to speed and load
BSF
Inner race defect BPFI and Multiples of BPFI with Remaining life sensitive
multiple sidebands of RPS, FTF, or to speed and load
BSF
Cage defect FTF and FTF and multiples or Worst defect; failure
multiples sidebands on BPFO or BPFI imminent
Ball defect BSF or FTF May be sidebands of BPFO or Limited time to failure
and multiples BPFI
Excessive Natural Natural frequencies modulated Excessive clearance
Internal clearance frequencies by RPS from wear

Table A.3 Gearbox Faults

Spectral Spectrum,
Fault Comments
Frequency Time Waveform
Eccentric gears Gear mesh Gear mesh with sidebands at
frequency of eccentric gear
Gear-mesh wear Gear mesh and Gear mesh with sidebands at
harmonics frequency of worn, scored, or
pitted gear(s); sometime 1/2,
1/3, 1/4 harmonics of gear
mesh
Improper backlash Gear mesh and Gear mesh with orders and
of end float harmonics sidebands at frequency of
pinion or gear
Broken, cracked, Natural Pulses in time waveform;
or chipped gear frequencies natural frequencies in
teeth spectrum
Gearbox distortion Gear mesh Gear mesh and orders in
and/or natural spectrum; varying gear-mesh
frequencies amplitude in time waveform
shaft frequency plus low-
amplitude orders
Common factor Fractional gear Fractional gear mesh
wear mesh and/or frequencies dependent on the
natural common factor(s)
frequencies

473
Table A.4 Fan and Blower Faults

Spectral Time Waveform/


Fault Correction
Frequency Spectrum
Structural 1x, blade Focused energy at frequency Change natural
resonance pass (BP) of resonance frequency or operating
speed or isolators
Critical speeds 1x, blade Focused energy at frequency Change natural
pass (BP) of critical speed frequency or operating
speed
Structural cracks in 1x Non-repeatable phase on Repair wheel
impeller wheel balancing
Loose wheel shaft 1x and Erratic phase Repair shaft
multiples
Aerodynamic 1x, BP, random High amplitude BP or high Change damper
problems noise noise flow, random noise and position or redesign
blade pass ducting
Pedestal stiffness, 1x Large horizontal-vertical Stiffen pedestal
asymmetry Amplitude difference
Belts and pulleys 1x, 2x, belt Belt pulses, 1x Replace pulley or belt
Frequency Vertical-horizontal in phase
Acoustical duct 1x, blade Vibration focused at Change ducting or
resonance pass frequency one fan speed fan speed or number of
blades
Isolators 1x and Orders Impact like waveform and Replace isolators if
spectrum or focused energy worn out redesign if
resonant
Impeller or pulley 1x High amplitude 1x Replace impeller or
eccentricity pulley
Impeller Cracks 1x, 2x High amplitude 1x, critical Difficult to balance
speed at 1/2 x, non-repeatable new or repaired wheel
data needed
Rubs Factional Sub-harmonics or orders of Change clearance or
Frequencies, operating speed reduce source or rub
1x and Orders excitation
Surge Natural Low frequency pulse with ring Change operation to get
Frequencies down on fan curve avoid
flow reversal
Belts Belt Frequency Vertical and horizontal in Replace belts
And Orders phase. Pulses in time
waveform when defect passes
pulley
Blower Pulsation No. of Lobes High amplitude of pulsation Add pulsation damper
xRPM frequency and multiples in discharge pipe

474
Table A.5 Pump Faults

Spectral Spectrum,
Fault Correction
Frequency Time Waveform
Recirculation 1x, vane pass Noisy time waveform and Increase flow through
spectrum elevated baseline pump
Cavitation 1x, vane pass Noisy data and acoustics Increase head on the
pump
Incorrect Assembly Vane pass Elevated vane pass frequency Reassemble casing
Structural 1x, 2x or vane Focused energy at resonant Alter natural frequency
Resonance pass frequency
Shaft Resonance 1x, 2x or Focused energy at shaft Alter natural frequency
(Critical Speeds) vane pass critical speed or fine tune balance
Impeller Deflection 1x Focused energy at 1x Increase size of shaft or
bearing stiffness
Torsional Vane pass, Focused energy at vane pass Change natural or
Resonance resonance frequency forcing frequency
Trapped Foreign 1x Increased vibration at 1x Unclog pump
Material
Excessive Wear 1x, critical Focused energy at 1x due to Return internal
Ring clearance speed critical speed components to
specification
Impeller-Diffuser 1x, vane pass Vibration at 1x, vane pass, Alter gaps to optimize
Gaps and random noise efficiency and vibration
Improper Inlet 1x, random Noisy time and frequency Change inlet to increase
Conditions noise straight section length
10 pipe diameters
Piping Structural 1x, 2x, Focused energy Change natural or
Resonance vane pass forcing frequencies
Accoustical 1x, or Focused energy at natural Change piping path or
Resonance (Piping) vane pass frequency vane pass frequency
Foundations 1x, 2x, or 1x and orders Improve foundation
vane pass design or construction

475
Table A.6 Pumps and Piping

Spectral Spectrum,
Fault Correction
Frequency Time Waveform
Structural 1x or Vibration focused at RPM or Change natural
resonance/piping vane pass vane pass frequency/speed
and
vertical pump
Acoustical piping Vane pass Vibration focused at vane Change piping length or
resonance pass install damper
Impeller 1x or Vibration focused at 1x Install new impellers
eccentricity vane pass RPM
Excessive impeller- 1x, Vane-passing frequencies Change clearances
diffuser clearance vane pass and multiples
Recirculation Random Random noise on data Change operation to
noise best efficiency point
(BEP)

Cavitation Random noise, Random noise on data Change operation to


vane pass BEP
Excessive wear ring 1x Strong 1x May alter critical speed,
clearance damping and stability

476
Table A.7 Motor-Generator Fault Analysis

Spectral Spectrum,
Fault Correction
Frequency Time Waveform
Uneven air-gap 2 x LF Two times line frequency Air gap eccentricity
beats with two times should be less than 5%
operating speed on two pole of nominal radial
motor clearance
Slot passing n x S/P x LF High frequency that is a Improve air-gap
excitation 1 multiple of line frequency eccentricity
Magnetic centre Natural Impacting (in axial direction) Restore shaft axial
frequencies induced natural frequencies position to magnetic
center
Eccentric stator 2 x LF Two times line frequency Repair stator
plus number of poles x slip
sidebands
Stator winding 2 x LF Pulses at 120 Hz modulated Rewind the stator
faults and multiples by operating speed
Stator resonance Natural High amplitude vibrations Move natural frequency
Frequencies multiples of line frequency from forcing frequency
Stator 2 x LF High amplitude vibration at 2 Align electrical unit to
misalignment x LF driver
Broken rotor bars Operating Operating speed and Rewind/repair rotor
and/or end shorting speed multiples plus sidebands at
rings number of poles x slip
frequency
Loose rotor bars Operating Operating speed vibration Rewind rotor
speed phase angle not repeatable
Eccentric rotor Operating Operating speed with Rewind rotor
speed sidebands at number of poles
and x slip frequency, rotor bar
2 x LF passing frequency with 2 x
LF sidebands
Shorted rotor Operating Rotor bow which may cause Machine to remove
laminations speed 1x vibration shorted areas
Mass unbalance, Operating 2 x LF sidebands at number Correct the mechanical
coupling lockup Speed and of poles x slip frequency fault
and eccentric 2 x LF
bearing journal

477
References
Alfredson, R.J. 1982. A computer based system for condition monitoring, Symposium on
Reliability of Large Machines, The Institute of Engineers Australia, Sydney, pp. 39 46.
Boto, P.A., Detection of bearing damage by shock pulse measurement, Ball Bearing J., 5, 167
176, 1979.
Braun, S. 1986. Mechanical Signature Analysis: Theory and Applications, Academic Press,
London.
Courrech, J. 1985. New techniques for fault diagnostics in rolling element bearings, In
Proceedings of the 40th Meeting of the Mechanical Failure Prevention Group, National Bureau
of Standards, Gaithersburg, MD, pp. 47 54.
Eisenmann, R.C. Sr. and Eisenmann, R.C. Jr. 1998. Machinery Malfunction: Diagnosis and
Correction, Prentice Hall, Newark.
Eschman, P. 1985. Ball and Roller Bearings, 2nd ed., Wiley, New York.
Goldman, S. 1999. Vibration Spectrum Analysis, Industrial Press, New York.
Grove, R.C. 1979. An investigation of advanced prognostic analysis techniques, Paper
USARTL-TR-79-10. Northrup Research and Technology Center, Palos Verdes, CA.
Lin, J., Zuo, M.J., and Fyfe, K.R., Mechanical fault detection based on the wavelet de-noising
technique, ASME J. Vib. Acoust., 126, 9 16, 2004.
Lyon, R.H. 1987. Machinery Noise and Diagnostics, Butterworth, Boston.
Mathew, J., Machine condition monitoring using vibration analysis, J. Aust. Acoust. Soc., 15, 7
21, 1987.
Mathew, J. and Alfredson, R.J., The condition monitoring of rolling element bearings using
vibration analysis, ASME J. Vib. Acoust. Stress Reliab. Des., 106, 447 457, 1984.
McFadden, P.D. and Smith, J.D., A model for the vibration produced by a single point defect in a
rolling element bearing, J. Sound Vib., 96, 69 77, 1984.
McFadden, P.D. and Smith, J.D., A signal processing technique for detecting local defects in a
gear from the average of the vibration, Proc. IMechE, 199, 99 112, 1985.
Mechefske, C.K., Fault detection and diagnosis in low speed rolling element bearings using
inductive inference classification, Mech. Syst. Signal Process., 9, 275 282, 1995.
Mechefske, C.K., Objective machinery fault diagnosis using fuzzy logic, Mech. Syst. Signal
Process., 12, 855 864, 1998.
Mechefske, C.K. and Liu, L., Fault detection and diagnosis in variable speed machines, Int. J.
Condition Monit. Diagn. Eng. Manage., 5, 29 42, 2001.
Mechefske, C.K. and Mathew, J., Fault detection in low speed rolling element bearings, part I:
the use of parametric spectra, Mech. Syst. Signal Process., 6, 297 308, 1992a.
Mechefske, C.K. and Mathew, J., Fault detection in low speed rolling element bearings, part II:
the use of nearest neighbour classification, Mech. Syst. Signal Process., 6, 309 317, 1992b.

478
Mitchell, J.S. 1981. An Introduction to Machinery Analysis and Monitoring, Penwell, Los
Angeles.
Mobley, R.K. 1990. An Introduction to Predictive Maintenance, Van Nostrand Reinhold, New
York.
Moubray, J. 1997. Reliability Centered Maintenance, Industrial Press, New York.
Randall, R.B. 1981. A new method of modeling gear faults, ASME Annual Congress. Paper No.
81-Set-10.
Randall, R.B. 1987. Frequency Analysis, Bruel & Kjaer, Copenhagen.
Randall, R.B. 2001. Bearing diagnostics in helicopter gearboxes, Condition Monitoring and
Diagnostic Engineering Management Conference, Manchester, U.K., August 2001.
Rao, B.K.N. 1996. The Handbook of Condition Monitoring, Elsevier, Oxford.
Reeves, C.W. 1999. The Vibration Monitoring Handbook, Coxmoor, Oxford.
Smith, J.D. 1983. Gears and Their Vibrations, Marcel Dekker, New York.
Taylor, J.I. 1994. The Vibration Analysis Handbook, Vibration Consultants, Tampa.
Timusk, M. and Mechefske, C.K. 2002. Applying neural network based novelty detection to
industrial machinery, 6th International Conference on Knowledge-Based Intelligent Information
Engineering Systems, Crema, Italy, November 2002.
Wowk, V. 1991. Machinery Vibration: Measurements and Analysis, McGraw-Hill, New York.

479

Das könnte Ihnen auch gefallen