Download as pdf or txt
Download as pdf or txt
You are on page 1of 495

Maintenance

and
Reliability
Best Practices

SECOND EDITION

Ramesh Gulati, CMRP, CRE, P.E.

Foreword by
Terrence O’Hanlon, CMRP

Industrial Press, Inc.


New York
A full catalog record for this book will be available from the Library of Congress

ISBN 978-0-8311-3434-1

Industrial Press, Inc.


989 Avenue of the Americas
New York, NY 10018

Sponsoring Editor: John Carleo


Interior Text and Cover Design: Janet Romano
Developmental Editor: Robert Weinstein

Copyright © 2013 by Industrial Press Inc., New York. Printed in the


United States of America. All rights reserved. This book, or any parts thereof,
may not be reproduced, stored in a retrieval system, or transmitted
in any form without the permission of the publisher.

10 9 8 7 6 5 4 3 2 1
Preface and Acknowledgements
to the Second Edition

The first edition of this book came out about three years ago;
since then, it has gone through several printings. The book has become the
most widely read by maintenance, operations, reliability, and safety pro-
fessionals and has been used as the textbook for maintenance and reliabil-
ity curricula in many colleges and universities. In 2011, the book received
the first prize (Gold Award) in the RGVA book competition, in the main-
tenance and reliability category, at MARTS/Chicago.
I want to thank you, my readers, professional friends, and col-
leagues at AEDC and in Jacobs for the book’s success and continued sup-
port. This second edition is the result of your valuable feedback which I
have received through emails, letters, comments, and personal discussions
at many conferences and professional society meetings.
I would like to thank David Hurst, Walt Bishop, Vijay Narain,
Sheila Sullivan, Scott Bartlett, Bart Jones, Lynn Moran, and many of other
my colleagues and management at AEDC/ATA – Jacobs for their contin-
ued support (and also for reviewing the manuscript). I would be unjust if
I didn’t acknowledge and give special thanks to Christopher Mears, a very
young and energetic engineer. Christopher is now head of the continuous
improvement section at AEDC; he has transformed himself into a true
reliability and maintenance expert in just a few years. Christopher and I
spent many hours reviewing and critiquing the manuscript to improve the
quality of the final product. I’m also thankful to Terrence O’Hanlon of
Reliabilityweb for his continued support and encouragement and for writ-
ing a very special foreword for this book.
I don’t want to miss the opportunity to thank John Carleo, Janet
Romano, and Robert Weinstein of Industrial Press, my publisher, for their
editorial support and having patience with my constantly changing sched-
ule. And finally, of course, my wife, Prabha and daughter, Sona for their
continued support, without which I wouldn’t have finished this valuable
work.

Ramesh Gulati

xv
Table of Contents

Foreword to the Second Edition xiii


Preface and Acknowledgements to the Second Edition xv
Preface to the First Edition xvii
Acknowledgements to the First Edition xix

Chapter 1 Introducing Best Practices 1


1.1 Introduction: What Is a Best Practice? 2
1.2 Key Terms and Definitions 3
1.3 What Do Best Practices Have to Do with Maintenance
and Reliability? 4
1.4 Examples of Maintenance and Reliability Benchmarks 6
1.5 Basic Test on Maintenance and Reliability Knowledge 9
1.6 Summary 17
1.7 Self Assessment Questions 17
1.8 References and Suggested Reading 18

Chapter 2 Culture and Leadership 19


2.1 Introduction 20
2.2 Key Terms and Definitions 21
2.3 Leadership and Organizational Culture 22
2.4 Strategic Framework: Vision, Mission, and Goals 28
2.5 Change Management 36
2.6 Reliability Culture 38
2.7 Measures of Performance 42
2.8 Summary 45
2.9 Self Assessment Questions 46
2.10 References and Suggested Reading 47

vii
viii

Chapter 3 Understanding Maintenance 49


3.1 Introduction 50
3.2 Key Terms and Definitions 51
3.3 Maintenance Approaches 53
3.4 Other Maintenance Practices 59
3.5 Maintenance Management System: CMMS 63
3.6 Maintenance Quality 74
3.7 Maintenance Assessment and Improvement 76
3.8 Summary 78
3.9 Self Assessment Questions 79
3.10 References and Suggested Reading 80

Chapter 4 Work Management: Planning and Scheduling 81


4.1 Introduction 82
4.2 Key Terms and Definitions 85
4.3 Work Flow and Roles 87
4.4 Work Classification and Prioritization 91
4.5 Planning Process 98
4.6 Scheduling Process 105
4.7 Turnarounds and Shutdowns 108
4.8 Measures of Performance 113
4.9 Summary 114
4.10 Self Assessment Questions 114
4.11 References and Suggested Reading 115

Chapter 5 Materials, Parts, and Inventory Management 117


5.1 Introduction 118
5.2 Key Terms and Definitions 120
5.3 Types of Inventory 122
5.4 Physical Layout and Storage Equipment 128
ix

5.5 Optimizing Tools and Techniques 134


5.6 Measures of Performance 145
5.7 Summary 147
5.8 Self Assessment Questions 148
5.9 References and Suggested Reading 149

Chapter 6 Measuring and Designing for Reliability and


Maintainability 151
6.1 Introduction 152
6.2 Key Terms and Definitions 156
6.3 Defining and Measuring Reliability and Other Terms 157
6.4 Designing and Building for Maintenance and Reliability 176
6.5 Summary 184
6.6 Self Assessment Questions 185
6.7 References and Suggested Reading 186

Chapter 7 Operator Driven Reliability 187


7.1 Introduction 188
7.2 Key Terms and Definitions 190
7.3 The Role of Operations 193
7.4 Total Productive Maintenance (TPM) 194
7.5 Workplace Organization: 5S 203
7.6 Overall Equipment Effectiveness (OEE) 208
7.7 Measures of Performance 213
7.8 Summary 214
7.9 Self Assessment Questions 217
7.10 References and Suggested Reading 217

Chapter 8 Maintenance Optimization 219


8.1 Introduction 220
8.2 Key Terms and Definitions 221
x

8.3 Understanding Failures and Maintenance Strategies 224


8.4 Maintenance Strategy — RCM 225
8.5 Maintenance Strategy — CBM 245
8.6 Other Maintenance Strategies 274
8.7 Summary 280
8.8 Self Assessment Questions 281
8.9 References and Suggested Reading 282

Chapter 9 Managing Performance 283


9.1 Introduction 284
9.2 Key Terms and Definitions 286
9.3 Identifying Performance Measures 288
9.4 Data Collection and Data Quality 297
9.5 Benchmarking and Benchmarks 298
9.6 Summary 305
9.7 Self Assessment Questions 306
9.8 References and Suggested Reading 307

Chapter 10 Workforce Management 309


10.1 Introduction 310
10.2 Key Terms and Definitions 313
10.3 Employee Life Cycle 314
10.4 Understanding the Generation Gap 317
10.5 Communication Skills 323
10.6 People Development 329
10.7 Resource Management and Organization Structure 338
10.8 Measures of Performance 348
10.9 Summary 348
10.10 Self Assessment Questions 349
10.11 References and Suggested Reading 349
xi

Chapter 11 Maintenance Analysis and Improvement Tools 351


11.1 Introduction 352
11.2 Key Terms and Definitions 353
11.3 Maintenance Root Cause Analysis Tools 357
11.4 Six Sigma and Quality Maintenance Tools 381
11.5 Lean Maintenance Tools 387
11.6 Other Analysis and Improvement Tools 394
11.7 Summary 398
11.8 Self Assessment Questions 399
11.9 References and Suggested Reading 399

Chapter 12 Current Trends and Practices 401


12.1 Introduction 402
12.2 Key Terms and Definitions 402
12.3 Sustainability, Energy Management, and the
Green Initiative 405
12.4 Personnel, Facility, and Arc Flash Safety 417
12.5 Risk Management 427
12.6 Corrosion Control 434
12.7 Systems Engineering and Configuration
Management 437
12.8 Standards and Standardization 441
12.9 Summary 447
12.10 Self Assessment Questions 449
12.11 References and Suggested Reading 450

Appendix for Chapter 1 451


Best Practices Q–A: Answer Key and Explanation 451

Index 467
Chapter 1
Introducing Best Practices

I have not failed. I have found 10,000 ways that won’t work.
— Thomas Edison

1.1 Introduction: What Is a Best Practice?


1.2 Key Terms and Definitions
1.3 What Do Best Practices Have to Do with Maintenance and
Reliability?
1.4 Examples of Maintenance and Reliability Benchmarks
1.5 Basic Test on Maintenance and Reliability Knowledge
1.6 Summary
1.7 Self Assessment Questions
1.8 References and Suggested Reading

After reading this chapter, you will be able to understand:

• What are best practices and why do we care?


• What do best practices have to do with maintenance and
reliability?
• Key Maintenance and Reliability (M&R) terms and bench
mark examples

In addition, you will also be able to assess your knowledge about the
basics of Maintenance and Reliability by taking a short test.

1
2 Chapter 1

1.1 Introduction: What Is a Best Practice?

The notion of a best practice is not new. Frederick Taylor, the father
of modern management, said nearly 100 years ago, “Among the various
methods and implements used in each element of each trade, there is
always one method and one implement which is quicker and better than
any of the rest.” In recent times, this viewpoint has come to be known as
the “one best way” or “best practice.”
“Best practice” is an idea which asserts that there is a technique,
method, or process that is more effective at delivering a desired outcome
than any other technique, method, or process. The idea is that with this
technique, a project or an activity such as maintenance can be completed
with fewer problems and unforeseen complications. Simply, we can say
that a technique, method, or process may be deemed a “best practice”
when it produces superior results. A best practice is typically a document-
ed practice used by the most respected, competitive, and profitable organ-
izations. When implemented appropriately, it should improve perform-
ance and efficiency in a specific area. We also need to understand that
“best practice” is a relative term. To some it may be a routine or a stan-
dard practice; but to others, it may be a best practice because a current
practice or method is not effective in producing the desired results.
History is filled with examples of people who were unwilling to
accept or adopt the industry standard as the best way to do anything. The
enormous technological changes since the Industrial Revolution bear wit-
ness to this fact. For example, at one time horses were considered the best
form of transportation, even after “horseless carriages” were invented.
Today, most people drive a gasoline or diesel vehicle — all improvements
on the original horseless carriage. Yet concerns over oil costs, supplies,
and global warming are driving the next group of transportation improve-
ments.
In the 1968 Summer Olympics, a young athlete named Dick Fosbury
revolutionized the high-jumping technique. Using an approach that
became known as the Fosbury Flop, he won the gold medal by going over
the bar back-first instead of head-first. Had he relied on “standard prac-
tice,” as did all of his fellow competitors, he probably would not have won
the event. Instead, by ignoring standard practice, he raised the perform-
ance bar — literally — for everyone. The purpose of any standard is to
provide a kind of reference. Therefore, that standard must be, “what is
possible?” and not, “what is somebody else doing?”
Introducing Best Practices 3

In real-world applications, a best practice is a very useful concept.


Despite the need to improve on processes as times change and things
evolve, a best practice is considered by some simply to be a business
buzzword used to describe the process of developing and following a stan-
dard way of doing activities that any organization can use or implement to
get better results. Implementing best practices in the area of maintenance
and reliability can help an organization to:

• Increase output with the same set of assets


• Reduce the need for capital replacements
• Reduce maintenance cost per unit
• Reduce total cost per unit of output
• Improve performance — cost, productivity, and safety
• Increase competitiveness
• Increase market share

A best practice tends to spread throughout an industry after success


has been demonstrated. However, demonstrated best practices can be slow
to implement, even within an organization. According to the American
Productivity and Quality Center, the three main barriers to adoption of a
best practice are a lack of:

1. Knowledge about current best practices


2. Motivation to make changes for their adoption
3. Knowledge and skills required to do so

The objective of this book is to provide knowledge of best practices


in the areas of maintenance and reliability, and to implement best practices
effectively. In later chapters, we will be discussing what we can do to
eliminate these barriers to create a sustainable reliability culture in an
organization.

1.2 Key Terms and Definitions


Asset
An electronic or mechanical hardware component or device, a
software product, or a manufacturing system or process.

Benchmark
Process of identifying, sharing, and using knowledge and best
practices. It focuses on how to improve any given business
4 Chapter 1

process by exploiting topnotch approaches rather than merely


measuring the best performance.

Best Practices
Technique, methods, or processes that are more effective at deliv-
ering a desired outcome than any other techniques, methods, or
processes. These are usually documented practices used by the
most respected, competitive, and profitable organizations.

Maintenance
The act of maintaining, or the work of keeping an asset in proper
operating condition.

Reliability
The probability that an asset or item will perform its intended
functions for a specific period of time under stated conditions. It
is usually expressed as a percentage and calculated using Mean
Time Between Failures (MTBF).

1.3 What Do Best Practices Have to Do with


Maintenance and Reliability?

In any organization, assets are needed to produce products or provide


services. An item or asset, as defined here, could be an electronic or
mechanical hardware component or device, a software product, or a man-

Figure 1.1 Asset Performance


Introducing Best Practices 5

ufacturing system or process. The objective of performing better mainte-


nance and improving reliability of assets in an organization is to ensure
that the assets are available to perform required functions, when needed,
in a cost-effective manner. The performance of an asset is based on three
factors (see Figure 1.1):

1. Inherent reliability — how it was designed?


2. Operating environment — how it will be operated?
3. Maintenance plan — how it will be maintained?

Usually assets are designed with a certain level of reliability. This


designed-in (or built-in) reliability is the result of individual components’
reliability and the way they are configured. This level of reliability is
called inherent reliability. We cannot change or improve the reliability of
an asset after it has been installed without replacing or modifying it with
better and improved components (with the exception of redesigning it).
The second factor, the operating environment of the asset, considers
operating conditions under which the asset has to operate along with the
operator’s skills. Several studies have indicated that 40% or more failures
are the result of operational errors. Organizations need to ensure that oper-
ators are appropriately educated and trained in operating these assets with-
out causing operational errors that lead to failures. In fact, operators
should be the first line of defense in monitoring the asset’s performance
and any abnormal conditions, and in initiating timely corrective actions.
The third factor is a maintenance plan that defines how the asset will
be maintained. The objective of a good maintenance plan is to sustain
asset reliability and to improve its availability. The plan should include the
necessary maintenance and service-type actions needed to detect potential
failures before they lead to unscheduled downtime.
So what do best practices have to do with these principles of mainte-
nance and reliability? Throughout the many years of the maintenance and
reliability industry, good and bad practices have been identified. These
good and bad practices have been briefed at international conferences,
discussed in person and over the airwaves, and written in magazines,
books, websites, and blogs. The best of these practices are now becoming
more accepted and published throughout local, national, and international
industries, becoming the benchmarks that companies seek to achieve.
Throughout this book, we will be discussing these factors and what best
practices can be used to improve asset performance.
6 Chapter 1

1.4 Examples of Maintenance and Reliability Benchmarks

What are the best practices in the maintenance and reliability (M&R)
area and how could those be implemented to get better results? M&R best
practices are practices that have been demonstrated by organizations who
are leaders in their industry. These companies are the quality producers
with very competitive costs, usually the lowest in their industry. A few
examples (by no means an exhaustive list) of maintenance and reliability
best practice benchmarks along with their related typical world class val-
ues are listed in Figure 1.2.

Figure 1.2 Comparing Best Practices Benchmark


Best of the Best Performance Measure Best Practice Typical
Benchmark World Class
Maintenance Cost as a percent of RAV
(RAV — Replacement Asset Value) 2–9% 2.0–3.5%

Maintenance - Material Cost


as a percent of RAV 1–4% 0.25–1.25%

Schedule Compliance 40–90% > 90%

Percent (%) Planned work 30–90% > 85%

Production Breakdown Losses 2–12% 1–2%

Parts Stock- out Rate 2–10% 1–2%

The first value shown in Figure 1.2, Maintenance Cost as a


Percentage of Replacement Asset Value (RAV), is a maintenance perform-
ance measure that is used extensively as a benchmark for evaluating best
practices. We can immediately identify the cost differences between com-
panies that are Typical and World Class or Best in Class. However, typi-
cal companies are likely to spend more to build-up their maintenance and
reliability program. Then, once they have achieved a desired level of reli-
ability and availability, they should be able to reduce the maintenance cost
by continuing to apply the best practices.
We need to be very diligent in using these benchmarks for compara-
tive purposes because definitions of these benchmarks can vary from one
Introducing Best Practices 7

organization to another. It is significantly important to ensure that the


terms used by the two organizations have the same meaning. The perform-
ance measures in Figure 1.2 have been included to demonstrate this.

1. Maintenance Cost as a percent of RAV: This measure is calculated


as maintenance cost divided by the replacement asset value. In this
benchmark, two factors must be defined in order to ensure a compar
ison is accurate:
a. Maintenance cost. This factor is the cost of maintenance for a
plant or facility; it includes maintenance labor, maintenance materials,
contractors used to perform maintenance work, capital maintenance, and
the cost of all projects to replace worn out assets.
b. Replacement Asset Value (RAV). This number typically comes
from the engineering or company’s insurance carrier and not from
accounting. It is not the book value considered for accounting purposes.
Instead, it is the current replacement cost of all assets for an industrial
facility. This measure should include the cost of removing old assets and
the cost of installing new ones.

2. Maintenance Material Cost as a Percent of RAV: This benchmark


is very similar to the previous measure and calculated simply as
maintenance material cost divided by the replacement asset value. In
most organizations, the material cost is easier to obtain from the
Computerized Maintenance Management System (CMMS) or orga-
nization’s financial system. To ensure a comparison is accurate, we
must ensure that the maintenance cost includes all maintenance
material purchased for all assets in a plant, including maintenance
storeroom parts and material, parts and material used by contractors
on maintenance, and capital maintenance work.
For example, organization “A” typically has their Maintenance
Material Cost as 2% of RAV and organization “B,” which has
applied best practices, typically has their Maintenance Material
Cost as 0.5% of RAV. This comparison indicates that organization
“A” is spending four times more for maintenance material when
compared to organization “B.”

Caution: Organizations need to understand that neither maintenance


cost nor maintenance material cost can be reduced in a sustainable man-
ner without the application of best practices. Many organizations attempt
to focus solely on maintenance cost reduction, but this approach is usual-
8 Chapter 1

ly misdirected until they have reviewed and improved their processes, and
applied best practices.

3. Schedule Compliance: This measure is the ratio of maintenance


labor hours consumed for the jobs or tasks completed (which were on
an approved schedule) divided by the total maintenance labor hours
available during that period. Some organizations also track the num-
ber of jobs/tasks completed which were on an approved schedule ver-
sus the total jobs/tasks on a schedule.

Maintenance Schedule. The maintenance schedule identifies


jobs/tasks to be completed and approved in the previous week or at least
three days in advance. It should cover 100% of maintenance labor.
For example, organization “A” is typical; their schedule compliance
is 40%. Organization “B” has applied best practices and has a schedule
compliance of 80%. This comparison indicates that organization “B” is
actually getting twice as much scheduled work out of its maintenance staff
as organization “A.” When schedule compliance is high, we usually find
that organizations also have high uptime and asset utilization rates. There
is a direct correlation between them.

4. Percent of Planned Work: This measure calculates the percent of


maintenance work orders where all parts, material, specifications,
procedures, tools, etc., have been defined prior to scheduling the
work. This best practice is a key to long-term success of any success-
ful maintenance organization.
For example, a typical organization “A” has a percent of planned
work measured at 30% whereas organization “B” has a percent of
planned work measured at 90%. This comparison indicates that the
organization “B” is proactively planning three times more work and
will typically have high uptime and asset utilization rate. Their main-
tenance cost is also low because unplanned work costs more to exe-
cute.

5. Production/Operations Breakdowns Losses: This number becomes


small and insignificant as best practices are applied and become a nor-
mal way of life. One important issue which directly impacts this
benchmark is that all personnel from the executive level to production
operator must be responsible for the plant’s assets. The organization
management must support the journey to excellence in implementing
Introducing Best Practices 9

best practices. Operators must see assets as something they own. The
only way this transformation can occur is through education and
empowerment.

6. Parts Stock-out Rate: This measure is based on the number of times


a maintenance craft person visits the storeroom to get the parts need
ed versus when parts are supposed to be in the storeroom, but is not
available in stock.

In working with many organizations over the years, we’ve noted that
benchmarking is not an easy process, particularly when there are no stan-
dard definitions of terms to benchmark. For example, RAV (replacement
asset value) may not have same meaning to Organizations A and B. Both
of them may have different definitions. This problem has been a major
challenge in M&R-related benchmarking initiatives. The Society for
Maintenance and Reliability Professionals (SMRP) has taken the lead
toward standardized maintenance and reliability terms, definitions, and
metrics. The author—along with Bob Baldwin, retired editor of
Maintenance Technology Magazine, and Jerry Kahn of Siemens—has
also published the manual Professional’s Guide to Maintenance and
Reliability Terminology to standardize the M&R terms.
When measuring performance against known benchmarks of best
practices, we will find that all benchmarks are interconnected and interde-
pendent. This is why an organization must have a clearly defined group of
maintenance and reliability processes to implement best practices.
Tailoring a best practice to suit your needs and working environment is
essential for its successful and effective implementation.
So far, we have discussed just a few examples of best practices and
their benchmarks. Throughout this book, we will be discussing practices
which may be standard, good, or best depending upon where you stand in
your journey for excellence in maintenance and reliability.

1.5 Basic Test on Maintenance and Reliability Knowledge

Many maintenance and reliability practitioners have not been success-


ful in implementing best practices. It is usually due to a lack of or limited
understanding of the best practices or not getting adequate management
support. Take the test on the following pages to assess your knowledge of
M&R best practices. This test will help you understand where you stand
in your reliability journey.
10 Chapter 1

Once you complete the test, go to the Appendix and score yourself
appropriately. Try not to guess when answering any of the questions. If
you are uncertain, skip the question and review the text later; otherwise,
your results may give you a false sense of how well you know “best prac-
tices” when it comes to maintenance and reliability.

1. Best Practices are practices that are defined and applied by an


organization to improve their operation. These practices may or may
not be proven, but results are found to be acceptable.
a. True
b. False

2. Maintainability is measured by PM schedule compliance.


a. True
b. False

3. All maintenance personnel’s time should be covered by work orders.


a. True
b. False

4. Operations and Maintenance must work as a team to achieve


improved OEE.
a. True
b. False

5. Best practices would indicate that 90% or more of all maintenance


work is planned.
a. True
b. False

6. 100% of PM and PdM tasks should be developed using FMEA /


RCM methodology.
a. True
b. False

7. Utilization of assets in a world-class facility should be about 85%.


a. True
b. False
Introducing Best Practices 11

8. 100% of maintenance personnel’s (craft) time should be scheduled.


a. True
b. False

9. Time-based PMs should be less than 20% of all PMs.


a. True
b. False

10. The 10% rule of PM is applied on critical assets.


a. True
b. False

11. Most emergency work orders should be written by production/


operations.
a. True
b. False

12. It should be common practice for Operators to perform PMs.


a. True
b. False

13. The P–F interval can be applied to visual inspections.


a. True
b. False

14. Understanding of a P-F curve should help in optimizing PM


frequency.
a. True
b. False

15. The best method of measuring the Reliability of an asset is by count-


ing downtime events.
a. True
b. False

16. The primary purpose of scheduling is to coordinate maintenance jobs


for greatest utilization of the maintenance resources.
a. True
b. False
12 Chapter 1

17. What percentage of your assets should be ranked critical based on


the risk to business?
a. Less than 20%
b. 20–50%
c. over 50%

18. Vibration monitoring can detect uniform impeller wear.


a. True
b. False

19. Understanding the known and likely causes of failures can help
design a maintenance strategy for an asset to prevent or predict
failure.
a. True
b. False

20. Reliability can be improved easily after a maintenance plan has been
put into operation.
a. True
b. False

21. What percentage of maintenance work should be proactive?


a. 100%
b. 85% or more
c. 50%

22. MTBF is measured by operating time divided by the number of


failures of an asset.
a. True
b. False

23. Maintenance cost will decrease as reliability increases.


a. True
b. False

24. The “F” on the P–F Interval indicates that equipment is still
functioning.
a. True
b. False
Introducing Best Practices 13

25. A rule of thumb is that, on average, an experienced planner can plan


work for how many craft people?
a. 5
b. 15
c. 25 or more

26. Which of the following is not a primary objective for implementing a


Planning process?
a. Reduce reactive work
b. Prevent delays during the maintenance process
c. Mesh the production schedule and the maintenance schedule

27. The best method of measuring the reliability of an asset is by:


a. MTTR
b. MTBF
c. Both

28. With the exception of emergency work orders, Planning and


Scheduling will benefit all maintenance work.
a. True
b. False

29. Leading KPIs predict results.


a. True
b. False

30. The 6th S in the 6 S (also called 5 S plus) process stands for safety.
a. True
b. False

31. RCM stands for:


a. Regimented Centers of Maintenance
b. Reliability Centered Maintenance
c. Reliable Centers of Maintenance (uses best practices)

32. The objective of RCM is to preserve functions.


a. True
b. False
14 Chapter 1

33. An MRO storeroom shouldn’t be stocking parts for emergencies.


a. True
b. False

34. The inventory turnover ratio for MRO store should be:
a. Less than 2
b. Between 4–6
c. Over 6
35. PM compliance is a _________ KPI.
a. Laggimg
b. Lagging or Leading
c. Leading

36. Quality is one key component of OEE.


a. True
b. False

37. Reliability and Maintainability can only be designed in.


a. True
b. False

38. Creating a reliability culture from a reactive mode can be accom-


plished in a short period of time if enough resources are made avail-
able.
a. True
b. False

39. Karl Fischer’s Coulometric Titration Method is an effective tech-


nique to determine the metallic content (in PPM) in an oil sample.
a. True
b. False

40. An IR thermography window is an effective method to satisfy NFPA


70E arc flash requirements.
a. True
b. False

41. FMEA is applicable only to assets currently in use.


a. True
b. False
Introducing Best Practices 15

42. RCM methodology can’t be used effectively on new systems being


designed.
a. True
b. False

43. Properly training the M&R workforce can increase asset and plant
availability.
a. True
b. False

44. TPM is a type of maintenance performed by the operators.


a. True
b. False

45. Lagging KPIs are the results of a process.


a. True
b. False

46. EOQ improves the inventory turn ratio.


a. True
b. False

47. New incoming oil from the supplier is always clean and ready to be used.
a. True
b. False

48. Which phase of the asset life cycle has the highest cost?
a. Design
b. Acquisition
c. O & M

49. Most of the maintenance costs become fixed


a. After installation
b. During operations
c. During design

50. RCM provides best results when used


a. During Operation /Production
b. During Design / Development
c. After an asset has failed or keeps failing
16 Chapter 1

51. How soon we can restore an asset is measured by:


a. MTBF
b. MTTR
c. MTBMA
d. None of the above

52. Availability is a function of:


a. MTBF
b. MTTR
c. Uptime
d. Uptime and downtime

53. The failure rate of a component / asset can be calculated by


knowing:
a. Number of failures
b. MTBF
c. MTTR
d. Uptime

54. The biggest benefit of a Failure Modes and Effects Analysis occurs
during:
a. Operations phase
b. Maintenance phase
c. Design phase
d. None of the above

55. PM schedule compliance should be equal or greater than 95%.


a. True
b. False

Please go to the Appendix 1 to check the best answers to these ques-


tions. If your correct answers are:

50 or more: You have an excellent M&R knowledge, but continue


to learn and enhance your knowledge.
41–49: You have good M&R knowledge, but there is a
potential for improvement.
40 or less: You have significant opportunities to improve.
Introducing Best Practices 17

1.6 Summary

A best practice is a technique or methodology that is found to be the


most effective and has consistently shown to achieve superior results.
When implemented appropriately, a best practice should improve per-
formance and efficiency in a specific area. We need to understand that best
practice is a relative term. A practice may be routine or standard to some,
but to others, it may be a best practice because their current practice or
method is not effective in producing the desired results.
A best practice is often not what everyone else is doing, but is what is
possible to achieve. It requires persuasive techniques that rely more on
appeals rather than force. A best practice usually requires a change in
process; therefore, it needs to be accepted by all parties/stakeholders for
successful implementation.
A best practice tends to spread throughout an industry after success
has been demonstrated. However, demonstrated best practices can often
be slow to implement, even within an organization. According to the
American Productivity and Quality Center, the three main barriers to
adoption of a best practice are a lack of:

1. Knowledge about current best practices


2. Motivation to make changes for their adoption
3. Knowledge and skills required to make changes

The work force needs to have the knowledge of best practices in the
area of maintenance and reliability to implement them effectively. A com-
mitment to using the best practices in the M&R field and utilizing all the
knowledge and technology at one’s disposal ensures success.

1.7 Self Assessment Questions

Q1.1 Define best practice. What are the barriers to implementing


best practices?

Q1.2 What are keys factors that impact the performance of plant
machinery?

Q1.3 Why has reliability become a competitive advantage in today’s


environment?
18 Chapter 1

Q1.4 Identify five key performance measures in the area of mainte


nance and reliability. Elaborate on each element of these
measures. What are typical “world class benchmark values”
for these performance measures?

Q1.5 Define what makes a benchmark “World Class.” Discuss,


using specific examples.

Q1.6 Define planned work and identify its benefits. What is a


typical world class benchmark number?

1.8 References and Suggested Reading

IW — “Industry Week,” Monthly Publications. Penton Media


Publications, 2006–2008.
Gulati R., Kahn and Baldwin. “The Professional’s Guide to
Maintenance and Reliability Terminology.” Reliabilityweb
Publication, 2010.
Mitchell, John S. Physical Asset Management Handbook, 4th ed. Clarion
Technical Publishing, 2007.
Moore, Ron. Making Common Sense to Common Practice. 3rd ed.
Butterworth Heinemann, 2004.
SMRP* and IMC**. Annual Conference Proceedings. 2003–2010.
www.wikipedia.com.

* SMRP — Society for Maintenance & Reliability Professionals


** IMC — International Maintenance Conference
Chapter 2
Culture and Leadership

Effective leadership is putting first things first.


Effective management is discipline, carrying it out.
—- Stephen Covey

2.1 Introduction
2.2 Key Terms and Definitions
2.3 Leadership and Organizational Culture
2.4 Strategic Framework: Vision, Mission, and Goals
2.5 Change Management
2.6 Reliability Culture
2.7 Measures of Performance
2.8 Summary
2.9 Self Assessment Questions
2.10 References and Suggested Reading

After reading this chapter, you will be able to understand:

• Organizational culture
• Leadership and its role
• Vision, mission, and goals
• Reliability culture
• Change management and role of change agents

19
20 Chapter 2

2.1 Introduction

Successful implementation of a new practice, small or large, is a chal-


lenge for any organization. The implementation requires enthusiasm
rather than distrust or fear from the individuals who will be impacted by
the change. Guiding, nurturing, and shepherding the workforce are the
skills needed to ensure that changes are received and implemented with a
positive attitude. Usually, how the workforce perceives these changes as
well as their beliefs, values, attitudes, and expectations, are a few of the
factors that need to be evaluated and considered in developing a “change”
implementation plan.
Leadership plays a key role in enabling this process by providing both
vision and resources. Creating a reliability culture conducive to change is
a long journey. It is not just the maintenance workforce that needs to
change their thinking, but also others in the organization (operations, pro-
duction, design, stores, information technology, etc.). All need to be work-
ing together as a team to create a sustainable reliability culture.
The vision of where the organization wants to be is an important ele-
ment of creating a reliability culture. Steven Covey, a leading motivation-
al author, emphasizes the importance of mission, vision, and goals when
he talks about “beginning with the end in mind” in his famous best seller,
7 Habits of Highly Effective People. According to a study of a Stanford
Management professor reported in The Wall Street Journal, organizations
using mission and vision statements successfully outperform those that do
not (with the key being these organizations actually using their mission
and vision statements). When we visualize, we are able to materialize and
convert our vision into goals and then into reality.
All organizations want to improve their processes in order to become
efficient and effective. For a maintenance and reliability (M&R) organi-
zational culture, progress requires not only reducing but also eliminating
failures while optimizing and educating the workforce to perform their
tasks effectively.
These efforts could require significant changes in work practices,
processes, and organization structure. Eventually these changes become
part of their daily work habits in the organization and lead to a positive
reliability culture. But for these changes to truly become part of a reliabil-
ity culture, an effective change management process must be executed.
Culture and Leadership 21

2.2 Key Terms and Definitions

Change Management
The process of bringing planned change to an organization.
Change management usually means leading an organization
through a series of steps to meet a defined goal. Synonymous
with management of change (MOC).

Culture
A common set of values, beliefs, attitudes, perceptions, and
accepted behaviors shared by individuals within an organization.

Cultural Change
A major shift in the attitudes, norms, sentiments, beliefs, values,
operating principles, and behavior of an organization.

Leadership
An essential organizational role that must establish a clear vision,
communicate that vision to those in the organization, and provide
direction, resources, and knowledge necessary to achieve goals
and accomplish the vision. It may require coordinating and bal-
ancing the conflicting interests of all members or stakeholders.

Mission
An organization’s purpose.

Mission Statement
A broad declaration of the basic, unique purpose, and scope of
operations that distinguishes the organization from others of its
type.

Organizational Culture
The beliefs and values that defines how people interpret experi-
ences and how they behave, both individually and in groups.

Strategy
An action plan that sets the direction for the coordinated use of
resources through programs, projects, policies, and procedures,
as well as organizational design and the establishment of per-
formance standards.
22 Chapter 2

Vision
The achievable dream of what an organization or a person wants
to do and where it wants to go.

Vision Statement
An overarching statement of the way an organization wants to be;
an ideal state of being at a future point.

2.3 Leadership and Organizational Culture

What Is Organizational Culture?

Culture refers to an organization’s values, beliefs, and behaviors. In


general, it is the beliefs and values which define how people interpret
experiences and behave, both individually and in groups. Culture is both
a cause and a consequence of the way people behave. Cultural statements
become operationalized when leaders articulate and publish the values of
their organization. They provide the pattern for how employees should
behave. Organizations with strong work cultures—including reliability
cultures—achieve higher results because employees sustain focus both on
what to do and how to do it.
Behavior and success are key enablers in creating the culture. There
is circular flow of mutual causation among organizational behavior, suc-
cess, and culture, as shown in Figure 2.1. When a change is accepted by
the team members, it changes their behavior. They do tasks differently as
required by the new change. Then, if this change allows work to get done
more easily, they can see some success. This success makes them accept
the change and leads to changing habits or routine work methods.
Eventually it becomes culture to do the tasks the new way (Figure 2.2).

Leadership’s Role in Creating and/or Sustaining Organizational


Culture
Leaders at both the corporate and plant levels must keenly understand
the impact reliability has on the bottom-line performance of the organiza-
tion. The valuation of an asset-dependent organization is significantly
affected by the effectiveness with which that asset is managed. Leadership
is a key enabler in creating an environment to implement reliability strate-
gies, which helps in fostering a “reliability” culture in the long run.
Culture and Leadership 23

Figure 2.1 Organizational Culture and Success Flow

Figure 2.2 Change and Culture


24 Chapter 2

True leadership is uncommon in today’s society and organizations


because it is not genuinely understood. Furthermore, it has been misinter-
preted, according to James McGregor Burns, author of the book,
Leadership. He wrote in this landmark book that:

Leadership is leaders inducing followers to act for


certain goals that represent the values and the motiva-
tions — the wants and needs, the aspirations and expec-
tations — of both leaders and followers. And the genius
of leadership lies in the manner in which leaders see and
act their own and their followers’ values and motivations.

Recently General (ret.) Colin Powell said that “Leadership is the art
of accomplishing more than the science of management says possible”. In
fact, observes author Oren Harari, in The Leadership Secrets of Colin
Powell, aspiring business leaders would do well to adopt Powell’s style.
Powell captured his leadership approach within 18 lessons:

Lesson 1: Being responsible sometimes means pissing people off.

Lesson 2: The day soldiers stop bringing you their problems is the day
you have stopped leading them. They have either lost confidence that
you can help them or concluded that you do not care. Either case is a
failure of leadership.

Lesson 3: Don’t be buffaloed by experts and elites. Experts often pos-


sess more data than judgment. Elites can become so inbred that they
produce hemophiliacs who bleed to death as soon as they are nicked by
the real world.

Lesson 4: Don’t be afraid to challenge the pros, even in their own back-
yard.

Lesson 5: Never neglect details. When everyone’s mind is dulled or dis-


tracted the leader must be doubly vigilant.

Lesson 6: You don’t know what you can get away with until you try.

Lesson 7: Keep looking below surface appearances. Don’t shrink from


doing so (just) because you might not like what you find.
Culture and Leadership 25

Lesson 8: Organization doesn’t really accomplish anything. Plans don’t


accomplish anything, either. Theories of management don’t much mat-
ter. Endeavors succeed or fail because of the people involved. Only by
attracting the best people will you accomplish great deeds.

Lesson 9: Organization charts and fancy titles count for next to nothing.

Lesson 10: Never let your ego get so close to your position that when
your position goes, your ego goes with it.

Lesson 11: Fit no stereotypes. Don’t chase the latest management fads.
The situation dictates which approach best accomplishes the team’s mis-
sion.

Lesson 12: Perpetual optimism is a force multiplier.

Lesson 13: Powell’s Rules for Picking People: Look for intelligence
and judgment and, most critically, a capacity to anticipate, to see around
corners. Also look for loyalty, integrity, a high energy drive, a balanced
ego, and the drive to get things done.

Lesson 14: Great leaders are almost always great simplifiers, who can
cut through argument, debate, and doubt, to offer a solution everybody
can understand.

Lesson 15: Part I: Use the formula P = 40-to-70, in which P stands for
the probability of success and the numbers indicate the percentage of
information acquired. Part II: Once the information is in the 40-to-70
range, go with your gut.

Lesson 16: The commander in the field is always right and the rear ech-
elon is wrong, unless proved otherwise.

Lesson 17: Have fun in your command. Don’t always run at a break-
neck pace. Take leave when you’ve earned it. Spend time with your
families. Corollary: Surround yourself with people who take their work
seriously, but not themselves; those who work hard and play hard.

Lesson 18: Command is lonely.


26 Chapter 2

Harari, a management professor at the University of San Francisco,


says Powell engages in “the kind of practical, mission- and people-based
leadership that ... has translated into performance excellence and compet-
itive success.”
Powell doesn’t use bluster to inspire his troops. He is polite and “is
not interested in intimidating people. He is convinced that frightened peo-
ple don’t take initiative or responsibility and that their organizations suf-
fer as a result.”
On the other hand, Harari says Powell doesn’t mind making people
angry—particularly in pursuit of organizational excellence. “Being
responsible sometimes means pissing people off,” Harari quotes Powell,
who believes that good leaders must defy the status quo.
Harari closes The Leadership Secrets of Colin Powell with this obser-
vation: “Leadership is not rank, privilege, titles, or money. It is responsi-
bility.”
Donald Phillips, the author of Lincoln on Leadership—Executive
Strategies for Tough Times, points out the following of Lincoln’s princi-
ples on leadership.

• Get out of the office and circulate among the troops.


• Build strong alliances.
• Persuade rather than coerce.
• Honesty and integrity are the best policies.
• Never act out of vengeance.
• Have courage to accept unjust criticism.
• Be decisive.
• Lead by being led.
• Set goals and be results-oriented.
• Encourage innovations.
• Preach a VISION and continually reaffirm it.

In modern days, Lincoln’s principle of “Get out of the office and cir-
culate among the troops” is known to us as Management by Wandering
Around (MBWA), as dubbed by Tom Peters and Robert Waterman in
their 1982 book, In Search of Excellence. The principle has also been
referred to us by other names and phrases, such as “roving leadership,”
“being in touch,” and “get out of the ivory tower.” It is simply the
process of getting out of the office and interacting with people. Peters
and Nancy Austin, in A Passion for Excellence, define MBWA as “the
technologies of the obvious.”
Culture and Leadership 27

Leaders emerge in every life situation to guide others along a partic-


ular path of change and toward a final destination point. Effective leader-
ship is not easy. History has shown us that the responsibilities and hazards
of leadership are as great as its rewards. Recent studies in the field of lead-
ership recognize and stress the need for building strong interpersonal rela-
tionships and bonds. In their book, Leaders, Warren Bennis and Burt
Nanus note that “leadership establishes trust, leaders pay attention; they
have the ability to trust others even if the risk seems great.”
The first dictionary definition of “leader” describes a primary shoot of
a plant, the main artery through which the organism lives and thrives. In
much the same way, organizations prosper or die as the result of their
leader’s ability to embody and communicate the organization’s vision.
How the M&R leader influences others very much dictates the health of
the M&R department and, ultimately, the entire organization.
Effective visions, according to Tom Peters, are inspiring. They should
be “clear and challenging — and about excellence.” It consists of a con-
cise statement or picture of where the organization and its people are

Figure 2.3 Leadership Attributes Survey Results


28 Chapter 2

heading and why they should be proud of it. An effective vision empow-
ers people and prepares for the future while also having roots in the past.
Leadership creates vision and energizes people to make organizations
and people successful. Figure 2.3 shows the results of a survey ranking
five key attributes of leadership.

1. Charisma
2. Competence
3. Communication
4. Energizing people
5. Vision (in creating)

Leadership plays an important role in creating vision and energizing


the workforce, as reported by this survey. One of the important factors
impeding success in an organization is a lack of or not enough leadership
support to implement changes. It has been found that successful leaders
share a number of qualities needed to improve their processes. They sup-
port:

• Creating the organization’s vision and mission


• Ensuring resource availability
• Empowering line managers with authority and accountability
• Ensuring both individual and group goals are aligned with the
organization’s vision and goals
• Viewing training as an investment in developing the workforce
rather than an unnecessary expense
• Aligning and integrating all changes and process improvements
(the best practices) towards meeting the organization’s overall
objectives.

2.4 Strategic Framework: Vision, Mission, and Goals

Both the organization and its people need to establish a strategic


framework for significant success. This framework, which is illustrated in
Figure 2.4, consists of:

• A vision for the future — Where are we and the organization


going?
• A mission that defines what we are planning to do — What will
be accomplished?
Culture and Leadership 29

• Values that shape our actions — Why are we going in this


direction?
• Strategies that zero in on key success approaches — How will we
get there?
• Goals and action plans to guide our daily, weekly, and monthly
actions — When will we get there?

An organization’s success and our personal success depend on how


well we define and live by each of these important concepts. In fact, it has
been found that organizations whose employees understand the mission
and goals enjoy a 29 percent greater financial return than other organiza-
tions, as reported by the Watson Wyatt Work Study.
The “Workplace 2000” Employee Insight Survey reported that U.S.
workers want their work to make a difference, but 75% do not think their
company’s mission statement has become the way they do business.

Vision
A vision statement is a short, succinct, and inspiring declaration of
what the organization intends to become or to achieve at some point in the
future. Vision refers to the category of intentions that are broad, all-inclu-
sive, and forward-thinking. It is the image that a business must have of its
goals before it sets out to reach them. It describes aspirations for the
future, without specifying the means that will be used to achieve those
desired ends.
Corporate success depends on the vision articulated by the organiza-
tional leaders and management. For a vision to have any impact on the
employees of an organization, it has to be conveyed in a dramatic and

Figure 2.4 A Strategic Framework for Success


30 Chapter 2

enduring way. The most effective visions are those that inspire, asking
employees for their best and communicating that constantly. A vision
statement is a pronouncement about what an organization wants to
become. It should resonate with all members of the organization and help
them feel proud, excited, and part of something much bigger than them-
selves. A vision statement should stretch the organization’s capabilities
and image of itself. It gives shape and direction to the organization’s
future. Visions range in length from a couple of words to several pages.
Sample vision statements are given in Figure 2.5.

Organization Vision Statement

“Bringing to the world a portfolio of beverage brands that


The Coca Cola Company anticipate and satisfy peoples’ desires and needs. Also,
being a great place to work where people are inspired to
be best they can be.”

“To become the world’s leading consumer company for


The Ford Motor Company automotive products and services.”

Kraft Foods “Helping People Around the World Eat and Live Better.”

“To give ordinary folk the chance to buy the same things
Wal-Mart
as rich people.”

“FedEx is committed to our People – Service – Profit


Philosophy. We will produce outstanding financial returns
Federal Express by providing totally reliable, competitively superior, global,
air-ground transportation of high quality goods and
documents that require rapid, time-certain delivery.”

Harley-Davidson “To fulfill dreams through the experiences of motorcycling.”

The Society for “To become the global organization known for providing
Maintenance & Reliability competitive advantage through improved physical asset
Professional’s (SMRP) management.”

“To be a premiere resource for advancing Physical Asset


Management (modified 07/14/10).”

Figure 2.5 Sample Vision Statements

Warren Bennis, a noted writer on leadership, says, “To choose a direc-


tion, leaders must have developed a mental image of the possible and
desirable future state of the organization. This image, which we call a
vision, may be as vague as a dream or as precise as a goal or a mission
statement”.
SMRP is a leading, not-for-profit, maintenance and reliability socie-
ty; it provides opportunities for its members to exchange best practices
and other educational and networking opportunities through conferences
or workshops. It has taken the initiative to standardize maintenance defi-
nitions including metrics. The certification arm of SMRP has a vision to
Culture and Leadership 31

Figure 2.6 Sample Vision Picture

have at least one Certified Maintenance & Reliability Professional


(CMRP) in every plant in the world.
Sometimes a picture portraying the vision can convey the message
very effectively. Figure 2.6 is a picture displaying the intent of an organi-
zation’s vision statement. This organization developed its vision picture in
the mid-1990s to become the best M&R organization. The buses (people)
are lined-up to see their successful M&R program. They have come a long
way from a reactive mode to a proactive, reliability-based culture; their
journey to excellence continues in spite of several changes in upper man-
agement of the M&R organization.

Examples of some maintenance vision statements include:

1. Maintenance Vision Statement


To leverage a highly-skilled workforce and employ effective
maintenance strategies in order to position XYZ organization as
the leading manufacturer across the industry.

2. Maintenance Vision Statement


A world-class maintenance system with a standardized approach
to plan, execute, track, and analyze maintenance and production
processes.
32 Chapter 2

The vision must convey the essence of how the organization desires
to accomplish feats that prove to be big, exciting and compelling.

Mission Statements
Mission and vision statements are very different. A mission statement
is an organization’s vision translated into written form. It’s the leader’s
view of the direction and purpose of the organization. For many corporate
leaders, it is a vital element in any attempt to motivate employees and to
give them a sense of priorities.
A mission statement should be a short and concise statement of goals
and priorities. In turn, goals are specific objectives that relate to specific
time periods and are stated in terms of facts. The primary goal of any busi-
ness is to increase stakeholder value. The most important stakeholders are
shareholders who own the business, employees who work for the busi-
ness, and clients or customers who purchase products or services from the
business.
The mission should answer four questions:

1. What do we do? What is the purpose of the organization?


2. How do we do it? What’s unique about the organization?
3. For whom do we do it? Who are our customers and
stakeholders?
4. What are our values and beliefs?

What do we do? This question should not be answered in terms of


what is physically delivered to customers, but instead by the real and psy-
chological needs that are fulfilled when customers buy our products or
services. Customers make purchase decisions for many reasons, including
economical, logistical, and emotional factors.
How do we do it? This question captures the more technical elements
of the business. Our answer should encompass the physical product or
service, how it is sold and delivered to customers, and how it fits with the
need that the customer fulfills with its purchase.
For whom do we do it? The answer to this question is also vital as it
will help us to focus our efforts. Anybody who uses our products or serv-
ices is our customer. It could be a person or system next in the production
line that takes what we build or provides a service. In a broader sense, it
could include stakeholders for whom we ultimately do it.
Culture and Leadership 33

What are our values and beliefs? Values and beliefs guide our plans,
decisions, and actions. Values become real when we demonstrate them in
the way we act and the way we insist that others behave. In forward-look-
ing and energized organizations, values are the real boss. They drive and
keep the workforce moving in the right direction.

Three Main Benefits Attributed to Mission Statements

1. They help companies focus their strategy by defining some


boundaries within which to operate.
2. They define the dimensions along which an organization’s
performance is measured and judged.
3. They suggest standards for individual ethical behavior.

In his book, First Things First, Steven Covey points out that mission
statements are often not taken seriously in organizations because they are
developed by top executives, with no buy-in at the lower levels. But it’s a
pretty safe assumption that there probably is buy-in when we develop our
own mission statements.
First Things First is actually about time management, but Covey and
his co-authors use the personal mission statement as an important princi-
ple. The idea is that if we live by a statement of what’s really important to
us, we can make better time-management decisions. The authors ask,
“Why worry about saving minutes when we might be wasting years?”
A mission statement may be valuable, but how in the world do we go
about crafting one? As one way to develop a mission statement, Covey
talks about visualizing your 80th birthday or 50th wedding anniversary
and imagining what all our friends and family would say about you. A
somewhat more morbid, but effective approach is writing your own obit-
uary.
Can we visualize what it would be like if there were no asset failures
or if production met their schedule without any overtime for one month,
even three months, or if there was not a single midnight call for three
months and we were able to sleep without worrying?

Developing a Mission Statement

• The mission statement should describe the overall purpose of the


organization.
34 Chapter 2

• If the organization elects to develop a vision statement before


developing the mission statement, ask “Why does the image, the
vision exist — what is its purpose?” This purpose is often the
same as the mission.
• When wording the mission statement, consider the organization’s
products, services, markets, values, concern for public image, and
maybe priorities of activities for survival.
• Ensure that wording of the mission is such that management and
employees can infer some order of priorities in how products and
services are delivered.

Some example mission statements related to M&R are:

A Physical Plant Mission Statement


The Physical Plant Department is a service organization whose main
purpose and goal is to provide the best possible facilities and climate in
which to support the instruction, learning programs, and public services of
the university. We strive to make our customers feel nurtured, inspired,
and uplifted by the excellence of our service and the caring concern of our
service providers.

Mission Statement of a Maintenance Department


To manage the business of maintenance in a manner that facilitates a
production effort which yields high quality products, low operating costs,
utilizes all resources, and involves production and maintenance employ-
ees working together towards a common goal.

Mission Statement of a Maintenance Organization


To maximize equipment performance, fostering an environment of
ownership and pride, through a structured approach to predictive and pre-
ventive maintenance.

Society of Maintenance & Reliability Professionals (SMRP’s)


Mission Statement

• Facilitate information exchange through a structured network of


maintenance and reliability professionals
• Support maintenance and reliability as an integral part of
business management
• Present a collective voice on maintenance and reliability issues
and advance innovative maintenance and reliability practices
Culture and Leadership 35

• Promote and support maintenance and reliability education for


people, production, and quality processes to improve the work
environment

If the environment changes, mission statements may require revisions


to include additional or different needs being fulfilled, delivery systems,
or customer groups. With this in mind, vision and mission statements
should be revisited periodically to determine whether modifications are
desirable to more accurately reflect the current environment or strategic
direction.
SMRP recently (July 2010) reviewed and updated their mission state-
ment. It says now:

“To develop and promote leaders in Reliability and


Physical Asset Management”.

Corporate Strategy
Strategy is a very broad term which commonly describes any think-
ing that looks at the bigger picture. Successful organizations are those
that focus their efforts strategically. To meet and exceed customer satis-
faction, the business team needs to follow an overall organizational
strategy. A successful strategy adds value for the targeted customers
over the long run by consistently meeting their needs better than the
competition does.
Strategy is the way in which an organization orients itself towards the
market in which it operates and towards the other companies in the mar-
ketplace against which it competes. It is a plan based on the mission an
organization formulates to gain a sustainable advantage over the competi-
tion.

The objectives must be:

• Focused on a result, not an activity


• Consistent
• Specific
• Measurable
• Related to time
• Attainable
36 Chapter 2

Setting Goals
The major outcome of strategic planning, after gathering all neces-
sary information, is the setting of goals for the organization based on its
vision and mission statement. A goal is a long-range aim for a specific
period. It must be specific and realistic. Long-range goals set through
strategic planning are translated into activities that will ensure reaching
the goal through operational planning. Examples of an M&R goal include
achieving 90% PM compliance or reducing the overtime to less than 5%
in a specified time period.

2.5 Change Management

No matter how many goals are set or how grand the vision, an organ-
ization can go only as far as the organizational culture will allow it. Any
organization that will succeed in a culture change must have some form
of a change management process. Even if the organization does not have
a formal (written-down) change management process, a process will exist
in that organization. However, the success rate of these changes will prob-
ably be limited by the charisma of the leadership or the vision of its team.
Changing a culture from reactive thinking to proactive follows a sim-
ilar process to what was mentioned in Section 2.2 in answering the ques-
tion “What Is Organizational Culture?” It has to be shown why preventive
and condition based approaches are better than reactive work. Making
people change what they do or how they think takes time. After all, it took
them a long time to build their current habit. People do certain things cer-
tain ways. In turn, when we ask them to do things differently or ask them
to buy into our plan (vision), we are taking them out of their comfort zone.
We must have very convincing reasons for people to change; we must
inspire them to accept change. These reasons would help greatly in the
implementation process. We have found that implementing changes in
small steps or in a small pilot area helps the process. Figure 2.7 shows the
culture change process.
Organizational change management takes a structured approach to
change, helping executive management, business units, and individual
employees make the transition from the current state to the desired future
state. The goal is to assist employees in assimilating change: to minimize
the disruption of expectations and loss of control that can easily lead to
resistance on the part of those who must actually change.
Culture and Leadership 37

Figure 2.7 Culture Change Process

Two key elements of any successful effort to change the culture are:

1. Influencing the behavior to change


2. Overcoming resistance to change

To influence the behavior to change, the following actions are


suggested:
• Increase understanding (i.e., why the change is needed and how it
relates to vision).
• Set goals and expectations.
• Establish a process for praise and recognition.
• Define and clarify roles.
• Establish and standardize process and procedures.
• Create discipline, develop tenacity, and be persistent.

To overcome resistance to change, the following actions are


suggested:
• Listen and communicate.
• Create awareness.
• Educate and train to create understanding.
• Get team members involved and let them see some success.
• Empower team members to improve and tailor the process —
change if needed.
38 Chapter 2

Role of a Change Agent


The change agent is another important person who helps implement
changes successfully. Change agents have the clout, conviction, charisma,
and resourcefulness to make things happen and keep others engaged in
implementing change. They usually have a number of skills including:

• They understand the organization’s politics to get work accom-


plished, but do not participate in the work.
• They have a good understanding of processes, the improvements
needed, and interface issues, including any financial impact of the
change.
• They are keen analyzers who can persuade and defend changes to
every level of the organization
• They speak several organizational languages or have certain back
grounds or traits — marketing, finance, operations, engineering,
etc.
• They have a passion for improvement; in essence, they bring order
out of chaos.

The change agent is a very important role in an organization that is


trying to implement a reliability culture. This role should be filled by a
senior management person who has the respect and trust of the people.

2.6 Reliability Culture

What Is Reliability Culture?


Can we define reliability culture? What does it mean if someone says
that organization XYZ has a reliability culture? Is there a metric to meas-
ure it? The problem is that culture — any kind of culture — is a “touchy-
feely” concept that is difficult to define and specifically measure.
We can see the impact of a reliability culture in the final outcome or
services provided by the organization. Reliability improves the bottom
line of an organization because it can meet the customer’s needs on time
and cost effectively. In a reliability-based organization:

• Assets are reliable and available as and when needed — High


UPTIME
• Assets are functioning and producing as designed
• Maintenance costs are reasonable (at optimum level)
• Plant operates safely and reliably
Culture and Leadership 39

The importance of reliability and implementing best M&R practices


are discussed at the highest level of the organization. Most organizations
talk about RCM/Reliability, but it’s treated as the “program of the month”
and loses its emphasis over time. Changing an existing culture of “run-to-
failure” or little/no PM program to a sustainable reliability culture takes
many years and consistent management support and resources.
In a reliability culture, prevention of failures becomes an emphasis at
every organizational level. The entire workforce is focused on asset relia-
bility. The workforce — operators, maintainers, engineers — think and act
to ensure:

• Assets are available to produce when needed


• Assets are maintained at a reasonable cost
• An optimized maintenance plan (RCM/CBM based)
• An effective facility maintenance plan — 80/20 principal applied
to prioritize the work. Most of the work is planned and scheduled.

If an asset fails, it gets fixed quickly, the root cause is determined, and
action is taken to prevent future failures. Facility/asset reliability analysis
is performed on a regular basis to increase uptime. The workforce is
trained and taught to practice reliability-based concepts and best practices
on a continuous basis.

Reliability Culture — Creating and Sustaining


Let us look into a real plant scenario where an asset breakdown/fail-
ure has occurred.

Operations reported that “Valve P-139” would not close. An


Operations workaround was used to divert the process temporarily.
The breakdown was reported to the Maintenance Department with
an urgent request in the CMMS/EAM system to fix the valve.

The following events happened:

• Maintenance dispatched a mechanic to evaluate and fix the


valve.
• Mechanic noticed “a burning smell” upon arrival, and suspected
the electric motor on a hydraulic pump had burned up. He called
an electrician to help.
• Electrician determined that the motor had failed. He asked his
supervisor to find a replacement motor.
40 Chapter 2

• Supervisor called the Storekeeper, who found that no spare motor


was available.
• Supervisor called Operations to report that the motor had failed
and would take a couple of days to repair. Operations demanded
the repair immediately, so the supervisor called the Plant Engineer
to help locate a spare motor.
• Plant Engineer and Supervisor found the same type of motor on a
similar system not being used. Supervisor sent another crew to
remove this motor while the first crew removed the failed motor.
• Maintenance replaced the motor and adjusted linkages due to
sluggish operation. The valve was released to Operations.
• The work order was closed with comment “valve was fixed.”
• Operations were so happy with a four-hour repair time (rather than
two days) that they sent an e-mail thanking the maintenance crew
for a job well done.

What kind of culture does this plant have? What kind of message is
being delivered to the workforce? It appears that this organization has a
reactive culture. Fixing things are recognized and appreciated.
Now, let us look into another plant with the same breakdown scenario,
but where the sequence of events happened a little differently:

• Maintenance Supervisor/Scheduler visited the site and assessed


the failure, finding that the valve linkage was tight and dry, along
with a failed electric motor on a hydraulic system.
• Supervisor/Scheduler assigned a mechanic and an electrician, and
requested both a 6-month chronological history report and a rec-
ommended parts list. He also alerted the plant engineer of the
problem.
• Electrician determined that the motor had failed (burned). The
overload relays didn’t function properly. Mechanic found that
linkage was tight due to inadequate lubrication.
• The repair history (attached to the Work Order) showed the fol-
lowing problem — a few months ago: Problem with valve
closing. Mechanic had adjusted and greased the linkage. The
hydraulic pressure on the system had been raised from 1500 psi to
1800 psi to make the actuator and linkage work smoothly.
• Repair plan included replacement of the motor and overload
relays, restoration of hydraulic pressure to system design, and
Culture and Leadership 41

greasing/ adjustment of linkage. A spare motor was available as a


part of the repairable program.
• Work was completed as planned. Operator was supporting the
repair and helped in testing the system. The valve returned to
operation.
• The WO was closed and repair details documented.
• Operations were pleased with a two-hour repair. The maintenance
manager personally thanked the maintenance crew for a job well
done and for finding the root cause. He then asked them for a plan
of further action needed to improve the reliability for review in 10
days.

Now let’s consider what happened in this plant. It seems like this plant
was doing fairly well. A lot of things went well during this repair and in
the follow-up actions suggested by the Maintenance Manager. But is
everything going as well as it could? Is the CMMS/EAM system provid-
ing the data we need to make the right decisions? What kind of message
is being delivered to the workforce? What kind of culture is in this plant?
In this plant, the CMMS/EAM system has provided the information
to help make the right decisions. The maintenance manager is emphasiz-
ing failure prevention. It’s a proactive culture — a step in the right direc-
tion.
Now let us look into another plant, a similar type of situation, but
where events happened a little differently. In this case, the plant opera-
tions (Operator) noted that on Valve # P–139:

• Motor: Current data on operations panel indicated a higher-than-


normal current. The visual inspection and site visit indicated that
the valve actuator was running sluggish. Maintenance was alerted
by the operator.
• Maintenance evaluated the situation with the help of the operator
and planned the repair on a scheduled downtime period.
• The repair was completed and there was no unscheduled down
time. All repairs were documented in the CMMS/EAM system for
asset history.
• PM tasks were reviewed and root cause analysis performed. Based
on this analysis, PM tasks were updated. A work order to redesign
the linkage based on root cause analysis was also issued to design
/ engineering.
42 Chapter 2

• Operators were thanked for watching the asset/system closely

What happened in this plant? What kind of culture does this organiza-
tion have? In this plant, “Failure” was identified and addressed before it
happened. Additionally,

• Operations and Maintenance worked together as a team


• System provided the “warning” data
• Process was designed to make it happen

In this organization, the reliability/maintenance leaders have done


their work. They provided the right tools, trained both operators and main-
tenance, and created the right culture — a reliability culture.

2.7 Measures of Performance

Leadership is one of the most prominent aspects of the organization-


al context. However, measures of leadership effectiveness are difficult to
quantify. Leadership is a critical catalyst for driving change in the organ-
ization to add value. To measure the impact (effectiveness) of leadership,
we need to understand to what extent it influences and drives perform-
ance. Leadership is basically the capacity of someone to bring about
change; in turn, it allows us to leverage it for greater organizational per-
formance.
An accepted model for measuring leadership comes from James M.
Kouzes and Barry Z. Posner, authors of the book “The Leadership
Challenge: How to Keep Getting Extraordinary Things Done in Organ-
izations”. Kouzes and Posner have devised the so-called Leadership
Practices Inventory (LPI). The LPI Measurement Model uses a series of
questions to assess leadership effectiveness. People under designated
leaders are observers, evaluating leaders on a series of qualities, such as:

• Discusses future trends regarding how they (the leaders) can


change their work
• Provides positive feedback on accomplishments
• Follows through on promises
• Treats others with respect
• Solicits feedback and opinions from others

Leaders are also required to assess their leadership based on several


behaviors, such as:
Culture and Leadership 43

• Sets a good personal example


• Actively listens to other viewpoints
• Supports others in their decisions
• Willing to take certain risks and experiment

Performance measures in Leadership can be classified into four cate-


gories: People Initiatives, Collaboration Initiatives, Organizational
Initiatives, and Professional Success Indicators. People Initiatives
describe the measures which leader-managers take to enhance employee
engagement. It includes factors like Reduced Turnover, Improved Safety,
Succession Planning, and Employee Effectiveness.
Collaboration Initiatives include KPIs (key process indicators) like
Collaboration with stakeholders and shareholders, Information Sharing,
Problem Solving Time, and Consensus Building Exercises. It indicates the
factors which describe the initiatives taken by the leader for improving
collaboration and flow of information in the organization.
Organizational Initiatives take into consideration indicators such as
Goal Achievement, Key Processes, Change Management, and Evaluation
and Assessment. These factors help define the future direction of the
organization. Professional Success indicators include Qualification Level,
Experience Level, Successful Projects Undertaken, and Industry Contacts
of a leader-manager. They indicate the power, influence, and knowledge
of the leader.
Tacy Byham and Audrey Smith of Development Dimensional
Institute (DDI) identified four distinct attributes of a leader in their
research paper “Optimizing Your Leadership Pipeline for: STRATEGIC
LEADERS.” They state that promising leaders / executives exhibit a dis-
tinctive mix of skills / competencies, knowledge, experience, and person-
ality attributes.

Competencies
Typical critical skills/competencies for executives include:

• Demonstrating an understanding of the competitive global busi-


ness environment as well as an awareness of economic, social, and
political trends that impact the organization’s global strategy.
• Establishing and committing to a long-term business direction
based on an analysis of systemic information and consideration of
resources, market drivers, and organizational values.
• Leading change, from identifying opportunities for improvement
44 Chapter 2

and supporting a better approach for addressing work process


issues to advocating for organizational transformation.
• Entrepreneurship, which is demonstrated by actively using knowl
edge to create or seize new business opportunities.
• Vividly communicating a compelling view of the future state in a
way that helps others understand how business outcomes will be
different when the vision and values become a reality.

Knowledge
Depending on their role, senior leaders need:

• Background on how to build a long-range strategic plan.


• Intimate understanding of customers’ requirements, markets, and
key competitors.
• Working knowledge of talent management functions, including
compensation, employee training and development, performance
management and measurement, recruitment, selection, promotion,
succession management, union contracts, and bargaining.
• Knowledge of product design and management, research and
development, distribution, and supply chain management.
• Keen insight into how to best create relationships with boards of
directors, analysts, and investors.

Experience
Before assuming leadership roles, leaders might require experience
in:
• Creating a corporate culture.
• Managing a diversity of functional areas.
• Having profit/loss responsibility for a business.
• Global leadership assignments.
• Managing a large operation.
• Playing a key role in a joint venture or merger.

Personal Attributes
Some of the key attributes successful leaders should demonstrate
include:

• Ambition, initiative, competitiveness, and leadership potential.


• Inquisitiveness, imagination, gregariousness, and a need for
social interaction.
Culture and Leadership 45

2.8 Summary

Corporate success depends on the vision articulated by the organiza-


tional leaders and management. For a vision to impact the employees of
an organization, it has to be conveyed in a dramatic and enduring way.
The most effective visions are those that inspire, usually asking employ-
ees for the best, the most, or the greatest. A vision explains where the
organization wants to be and is an important element of creating a relia-
bility culture.
Culture refers to an organization’s values, beliefs, and behaviors. In
general, this culture is concerned with the beliefs and values of people
based on their interpretation of experiences and how they behave, both
individually and in groups. Cultural statements become operationalized
when leaders articulate and publish the values of their organization which
provide patterns for how employees should behave. Organizations with
strong reliability cultures achieve higher results because employees sus-
tain focus both on what to do and how to do it.
When we visualize, we are able to materialize; to convert our vision
into goals and then into reality. Leadership is an enabler in this process by
providing direction and resources to make this happen.
Creating a sustainable reliability culture is a long journey that requires
sound change management practices. Many organizations get impatient
and stop supporting reliability change or implementing best practices in
the organization; they fail to understand it takes many years to create a
sustainable reliability culture.
To sustain a reliability culture, the reliability/maintenance leaders in
an organization must continue to provide the right tools, training, and edu-
cation to both the operators and maintainers together as a team. They need
to ensure that the workforce is always current on:

• Knowledge — of best practices


• Teamwork — to assure communication and understanding
• Focus — on the right goals for business success
• Planning — to create a roadmap for knowing where they are and
where they want to be
• Processes — documentation, adherence, and discipline
• Measurements — to provide feedback and control and to ensure
that they, the leadership, continue to support the continuous
improvement environment and are creating a conducive, sustain-
able culture
46 Chapter 2

2.9 Self Assessment Questions

Q2.1 What are the key attributes of a leader?

Q2.2 Why is vision important?

Q2.3 Define MBWA. Why is it considered one of the key leadership


practices?

Q2.4 Define an organizational culture.

Q2.5 What are the key benefits of having a mission statement?

Q2.6 Why are mission and vision statements important for an


organization?

Q2.7 Define reliability culture.

Q2.8 Why is change management an important part of creating the


right reliability culture?

Q2.9 Define the role of a change agent. Who is best qualified to


perform this role?

Q2.10 State the differences between a manager and a leader.


Culture and Leadership 47

2.10 References and Suggested Reading

Bennis, Warren and Burt Nanus. Leaders. HarperCollins Publishers,


1985.
Burns, James McGregor. Leadership. Harper Perennial Modern Classics,
1982.
Covey, Stephen R. 7 Habits of Highly Effective People. Simon &
Schuster, Fireside Edition, 1990.
Covey, Stephen R. First Things First. Simon & Schuster, 1994.
Harari, Oren. Leadership Secrets of Colin Powell. McGraw-Hill, July
2003.
Kouzes, James M. and Barry Z. Posner. The Leadership Challenge,4th
Edition. Jossey-Bass, August, 2008.
Peters, Tom and Nancy Austin. A Passion for Excellence. Grand Central
Publishing, 1989.
Peters, Tom and Robert Waterman. In Search of Excellence: Lessons from
America’s Best Run Companies. Warner Books, 1982.
Phillips, Donald. Lincoln on Leadership: Executive Strategies for Tough
Times. Warner Business Books, 1992.
Schwartz, David, PhD. The Magic of Thinking BIG. Simon & Schuster,
Fireside Edition, 2007.
Thomas, Stephen. Improving Maintenance & Reliability through Cultural
Change. Industrial Press, 2005.
Chapter 3
Understanding Maintenance

“Your system is perfectly designed to give you the results that you get.”
-- W. Edwards Deming

3.1 Introduction
3.2 Key Terms and Definitions
3.3 Maintenance Approaches
3.4 Others Maintenance Practices
3.5 Maintenance Management System: CMMS
3.6 Maintenance Quality
3.7 Maintenance Assessment and Improvement
3.8 Summary
3.9 Self Assessment Questions
3.10 References and Suggested Reading

After reading this chapter, you will be able to understand:

• Why do maintenance?
• The objective of maintenance
• Benefits of maintenance
• Types of maintenance approaches
• Purpose of CMMS/EAM
• Maintenance quality challenges
• Importance of assessing your maintenance program regularly

49
50 Chapter 3

3.1 Introduction

What Is Maintenance and Why Is It Important?


Maintenance is concerned with keeping an asset in good working con-
dition so that the asset may be used to its full productive capacity. The
maintenance function includes both upkeep and repairs. The dictionary
defines maintenance as “the work of keeping something in proper condi-
tion” A broader definition is:

• Keep in ‘designed’ or an acceptable condition;


• Keep from losing partial or full functional capabilities;
• Preserve, protect.

This definition implies that the term maintenance includes tasks per-
formed to prevent failures and tasks performed to restore the asset to its
original condition.
However, the new paradigm of maintenance is related to capacity
assurance. With proper maintenance, the capacity of an asset can be real-
ized at the designed level. For example, the designed capacity of produc-
tion equipment of x units per hour could be realized only if the equipment
is operated without considerable downtime for repairs.
An acceptable capacity level is a target capacity level set by manage-
ment. This level cannot be any more than the designed capacity. Consider
production equipment that is designed to make 500 units per hour at a
maintenance cost of $150 per hour. If the equipment is down 10% of the
time at this level of maintenance, the production level will be reduced to
450 units per hour. However, if the maintenance department, working
with the production department together as a team, can find a way to
reduce the downtime from 10% to 5% at a slightly increased maintenance
cost/hour, this reduction will increase the output by another 25 units/hour.
Therefore, it is conceivable that management would be able to justify the
increased maintenance cost. Thus capacity could be increased closer to
designed capacity by reducing downtime.
Unfortunately, literature related to maintenance practices over the past
few decades indicates that most companies did not commit the necessary
resources to maintain assets in proper working order. Rather, assets were
allowed to fail; then whatever resources needed were committed to repair
or replace the failed asset or components. In fact, maintenance function
was viewed as the necessary evil and did not receive the attention it
deserved.
Understanding Maintenance 51

However, in the last few years, this practice has changed dramatical-
ly. The corporate world has begun recognizing the reality that mainte-
nance does add value. It is very encouraging to see that maintenance is
moving from so-called “backroom” operations to the corporate board
room. A case in point —in the 2006 annual report to investment brokers
on the Wall Street, the CEO of Eastman Chemical included a couple of
slides in his presentation related to maintenance and reliability stressing
the company’s strategy of increasing equipment availability by commit-
ting adequate resources for maintenance.

3.2 Key Terms and Definitions


Asset
An asset is defined as something that has potential or actual value
to an organization; the physical resources of an organization,
such as equipment, machines, mobile fleet, systems, or their parts
and components, including software that performs a specific
function or provide a service; sometimes also referred to as phys-
ical assets.

Component
An item or subassembly of an asset, usually modular and replace-
able, sometimes serialized depending on the criticality of its
application; interchangeable with other standard components
such as belt of a conveyor, motor of a pump unit, or a bearing.

Computerized Maintenance Management System / Enterprise


Asset Management (CMMS / EAM)
A software system that keeps record and tracks all maintenances
activities, e.g., maintenance work orders, PM schedules, PM
masters, material parts, work plans, and asset history. Usually it
is integrated with support systems such as inventory control, pur-
chasing, accounting, manufacturing, and controls maintenance
and warehouse activities.

Failure Mode and Effects Analysis (FMEA)


A technique to examine an asset, process, or design to determine
potential ways it can fail and the potential effects(consequences);
and subsequently identify appropriate mitigation tasks for highest
priority risks.
52 Chapter 3

Maintenance, Backlog
Maintenance tasks those are essential to repair or prevent equip-
ment failures that have not been completed yet.

Maintenance, Capital Project (CPM)


Major repairs, e.g., overhauls and turnaround projects, valued
over a certain threshold are sometime treated as capital projects
for tax purposes. If these projects are essential to restoring the
asset back to the designed capacity — not to add additional capa-
bilities, they should be treated as maintenance costs.

Maintenance, Condition Based (CBM)


Maintenance based on the actual condition (health) of an asset as
determined from non-invasive measurements and tests. CBM
allows preventive and corrective actions to be optimized by
avoiding traditional calendar or run-time directed maintenance
tasks. The terms Condition Based Maintenance (CBM), and
Predictive Maintenance (PdM) are used interchangeably

Maintenance, Corrective (CM)


Repair actions initiated as a result of observed or measured con-
ditions of an asset after or before the functional failure.

Maintenance, Operator Based (OBM)


OBM involves operators performing some basic maintenance
activities. Operator-based maintenance is a cost-effective practice
to perform minor routine, and recurring maintenance tasks by the
operators to keep the asset working efficiently for its intended
purpose.

Maintenance, Predictive (PdM)


A maintenance strategy based on the actual condition (health) of
an asset as determined from non-invasive measurements and
tests. PdM allows preventive and corrective actions to be opti-
mized by avoiding traditional calendar or run-time directed main-
tenance tasks. The condition of equipment could be measured
using condition monitoring, statistical process control, or equip-
ment performance, or through the use of the human senses. The
terms Predictive Maintenance (PdM) and Condition Based
Maintenance (CBM) are used interchangeably
Understanding Maintenance 53

Maintenance, Preventive (PM)


A maintenance strategy based on inspection, component replace-
ment, and overhauling at a fixed interval, regardless of its condi-
tion at the time. Usually, scheduled inspections are performed to
assess the condition of an asset. Replacing service items, e.g., fil-
ters; oils, and belts and lubricating parts are a few examples of
PM tasks. PM inspection may require another work order to
repair other discrepancies found during the PM.

Maintenance, Reactive (RM)


Maintenance repair work done as an immediate response to an
asset failure, normally without planning and unscheduled.
Synonymous with breakdown and emergency maintenance.

Maintenance, Run-to-Failure (RTF)


A maintenance strategy (policy) for assets where the cost and
impact of failure is less than the cost of preventive actions. It is a
deliberate decision, based on economical effectiveness, not to
perform PM but let asset run to fail.

Proactive (Maintenance)Work
The sum of all maintenance work that is completed to avoid fail-
ures or to identify defects that could lead to failures (failure find-
ing). It includes routine preventive and predictive maintenance
activities and work tasks identified from them.

Reliability
The probability that an asset or item will perform its intended
functions for a specific period of time under stated conditions.

Reliability Centered Maintenance (RCM)


A systematic, disciplined process, for establishing the appropriate
maintenance plan for an asset/system to minimize the probability
of failures. The process ensures safety, system function, and mis-
sion compliance.

3.3 Maintenance Approaches

Many organizations have many different approaches (or some may


call practices) when it comes to their maintenance programs. All
54 Chapter 3

approaches have at their basis the requirement to keep their facility’s


assets at whatever capacity level is necessary for their current operational
needs. Some of these maintenance programs are more structured than oth-
ers—some maintenance programs are based on an RCM analysis, and
some organizations even develop an annual or even multi-year mainte-
nance program plan to guide their maintenance decisions strategically and
tactically. The truth is that an organization will have a maintenance pro-
gram whether they admit it or not; their program will simply be more cost-
ly than it has to be since they will live in a reactive maintenance state.
All assets requires some form of care—maintenance, for example.
Belts and chains require adjustment, alignment of shafts such as pump-
motor or blower-motor shafts need to be properly maintained, filters need
to be changed at regular intervals, proper lubrication on rotating machin-
ery is required, and so on. In some cases, certain components need
replacement after a specified number of hours of operations, e.g., a pump
bearing on a hydraulic system to ensure that the system lasts through its
design life. Anytime we fail to perform maintenance activities, we may be
shortening the operating life of the asset. Over the past few decades, many
cost-effective approaches have been developed to insure an asset reaches
or exceeds its design life. Instead of waiting for assets to fail and then fix
them, maintenance actions are performed to keep assets in good working
condition to provide continuous service.

Why Have a Structured Maintenance Program


The most important reason to have a maintenance program with a
structured approach is to ensure that assets don’t fail prematurely, that
they keep producing or providing service as intended. Maintenance pro-
grams should improve production capacity and reduce overall facility
costs by:

• Reducing production downtime — the result of fewer asset


failures.
• Increasing life expectancy of assets, thereby eliminating prema-
ture replacement of machinery and asset.
• Reducing overtime costs and providing more economical use of
maintenance personnel due to working on a scheduled basis,
instead of an unscheduled basis, to repair failures.
• Reducing cost of repairs by reducing secondary failures. When
parts fail in service, they usually damage other parts.
Understanding Maintenance 55

• Reducing product rejects, rework, and scrap due to better overall


asset condition.
• Identifying assets with excessive maintenance costs, indicating the
need for corrective maintenance, operator training, or replacement
of obsolete assets.
• Improving safety and quality conditions.

A structured maintenance program can have different philosophies,


approaches, and practices embedded within the program. The basic phi-
losophy is really only two-fold: do some form of maintenance to an asset
to prevent failure or allow the asset to be run-to-failure. The basic
approaches to maintenance can be grouped into four major categories:
Condition Based Maintenance (also known as Predictive Maintenance),
Preventive Maintenance, Proactive Maintenance, and Corrective
Maintenance. A brief description of each of these approaches is discussed
on the following pages with more details on each covered within chapters
4 and 8.

Condition Based Maintenance (CBM)


Condition Based Maintenance (CBM), also known as Predictive
Maintenance (PdM), attempts to evaluate the condition of an asset by per-
forming periodic or continuous asset monitoring. The ultimate goal of
CBM is to identify proactive maintenance actions to be performed at a
scheduled point in time when the maintenance activity is most cost effec-
tive and before the asset fails in service. The “predictive” component
stems from the goal of predicting the future trend of the asset’s condition.
This approach uses principles of statistical process control, trend analysis,
and preselected thresholds to determine at what point in the future main-
tenance activities should be scheduled.
CBM inspections typically are performed while the asset is operating,
thereby minimizing disruption of normal system operations. Adoption of
CBM/PdM in the maintenance of an asset can result in substantial cost
savings and higher system reliability.
There are a number of different CBM / PdM technologies that can be
used to evaluate assets condition. A few of the more common technolo-
gies (or data) are:

• Vibration analysis
• Infrared (IR) thermography
• Acoustic / Ultrasonic — sound level measurements
56 Chapter 3

• Oil analysis
• Electrical — amperage plus other data
• Shock Pulse Method (SPM)
• Partial discharge & Corona detection
• Operational performance data — pressure, temperature, flow
rates, etc.

Basically, in the CBM approach, the maintenance need is based on the


actual condition of the machine rather than on some preset schedule.
Activities such as changing oil are based on a predetermined schedule
(time), like calendar time or asset runtime. For example, most of us
change the oil in our cars every 3,000–5,000 miles driven. This is effec-
tively basing the oil change needs on asset runtime. No concern is given
to the actual condition and performance capability of the oil. It is changed
because it is time to change it. This methodology would be analogous to
a preventive maintenance task.
On the other hand, if we ignore the vehicle runtime and have the oil
analyzed at some regular period to determine its actual condition and
lubrication properties, then we may be able to extend the oil change until
the car has been driven 10,000 miles, or maybe even more.
This is the advantage of utilizing condition based maintenance. CBM
is used to define needed maintenance tasks based on quantified asset con-
ditions or performance data. The advantages of CBM are many. A well-
established CBM program will eliminate or reduce asset failures cost-
effectively. It will also help to schedule maintenance activities to mini-
mize overtime cost. In addition, we will be able to minimize inventory and
order parts, as required, well ahead of time to support the downstream
maintenance needs.
Past studies have shown that a well-implemented CBM program can
provide an average savings of 10% (7–15%) over a program utilizing pre-
ventive maintenance (PM) alone. These savings could easily exceed
30–40% if there is not a good PM program in place. In fact, independent
surveys and technical papers presented at the International Maintenance
Conferences 1999–2002 combined with the author’s own experience indi-
cate the following industrial average savings resulting from a good estab-
lished condition based maintenance program:

• Reduction in maintenance costs: 15–30%


• Reduction in downtime: 20–40%
• Increase in production: 15–25%
Understanding Maintenance 57

On the down side, starting a full-blown CBM program utilizing all of


the previously mentioned technologies can be quite expensive. Some
technology’s test equipment may cost in excess of $40,000. In addition,
training plant personnel to utilize PdM technologies effectively will
require considerable funding as well. This is one reason to have an RCM-
basis for choosing where to apply which CBM technology; it helps to
determine the test equipment purchase that can provide the most “bang for
your buck.” Program development will require an understanding of pre-
dictive maintenance and a firm commitment to make the program work by
all facility organizations and management.
How the CBM team should be organized is another issue. We have
found that a centralized dedicated team is a good way to start a program.
This approach helps in standardizing testing methods and practices.
The CBM approach consists of scheduling maintenance activities
only when equipment or operational conditions warrant — by periodical-
ly or continuously monitoring the machinery for excessive vibration, tem-
perature, noise, etc. When the condition gets to a level that has been pre-
determined to be unacceptable, the asset is shut down. The asset is then
repaired or has damaged components replaced in order to prevent more
costly failures from occurring. This approach works very well if person-
nel have adequate knowledge, skills, and time to perform the CBM work.
In addition, the company must allow asset repairs to be scheduled in an
orderly manner. The approach provides some lead-time to purchase mate-
rials for the necessary repairs, reducing the need for a high parts invento-
ry. Because maintenance work is only performed when it is needed, there
is likely to be an increase in production capacity.

Preventive Maintenance (PM)


As stated previously, a CBM approach is the preferred approach if
your organization can handle the expense of implementing this approach.
However, a PM approach is the next best thing and maybe the only
approach with certain types of assets. Additionally, regulatory require-
ments may force some level of PM to be performed (e.g., crane inspec-
tions).
Preventive Maintenance requires that maintenance or pro-
duction/operation personnel pay regular visits to monitor the condition of
an asset in a facility. The basic objective of PM visits is to take a look at
the asset to determine if there are any telltale signs of failure or imminent
failure. Also, depending on the type of the asset, a checklist or a procedure
with task details indicating what to check or what data to take may be
58 Chapter 3

used, e.g., change filter, adjust drive belts, and take bearing clearance
data. The observers also document the abnormalities and other findings.
These abnormalities should be corrected before they turn into failures for
a PM program to add any value.
These PM inspections can be based on either calendar time or asset
runtime. If CBM is not being performed on a particular piece of equip-
ment, or if CBM cannot detect a particular failure, then the next best
approach is a runtime-based PM program, but only for equipment and
failure modes that have a time basis. If a calendar time-based PM program
is all that really adds value, then that approach is still better than a run-to-
failure strategy. The exception to this is when an analysis has been per-
formed that indicates the most cost-effective strategy is run-to-failure
because the total cost of maintenance is less than the corrective mainte-
nance necessary for this run-to-failure strategy (assuming that there is no
safety impact to this run-to-failure strategy).
The objective of preventive maintenance can be summarized as fol-
lows:

• Maintain assets and facilities in satisfactory operating condition


by providing for systematic inspection, detection, and correction
of incipient failures before they develop into major failure.
• Maintenance, including tests, measurements, adjustments, and
parts replacement, performed specifically to prevent failure from
occurring.
• Record asset health condition for analysis which leads to the
development of corrective tasks.

Proactive Maintenance
Proactive Maintenance is one of those terms used to mean different
things to different people. The term can be a controversial one. Some con-
sider CBM and PM approaches to be proactive —these approaches do
take a proactive approach as opposed to simply reacting to equipment fail-
ure. In some cases, tasks that are generated based on what is found during
CBM and PM tasks (including work identified as a result of root cause
and failure analysis) are considered proactive. In some organizations,
proactive maintenance is calculated as a ratio of all maintenance work
minus unscheduled corrective maintenance, divided by all maintenance
work. Another definition is that anything on the maintenance schedule is
proactive — that is, any maintenance work that has been identified in
advance and is planned and scheduled. These last definitions make better
Understanding Maintenance 59

sense. Therefore, proactive maintenance can be defined as all work tasks


completed to avoid failures or to identify defects that could lead to fail-
ures.

Corrective Maintenance (CM)


Corrective Maintenance is another term used in different ways. CM is
an action initiated as a result of an asset’s observed or measured condition
before or after functional failure. CM work can be further classified into
Scheduled, Major Repairs/Projects, and Reactive.
When an asset breaks down, it fails to perform its intended function
and disrupts scheduled operation. This functional loss, partial or total,
may result in defective parts, speed reduction, reduced output, and unsafe
conditions. For example, a wear or slight damage on a pump impeller,
which reduces output, is a function reduction failure. Full functional fail-
ure that shuts down the asset is called function-disruption failure.
Function-disruption or reduction failures that are not given due attention
will soon develop into asset stoppage if not acted on.
Many abnormalities such as cracks, deformations, slacks, leakages,
corrosions, erosions, scratches, excessive heats, noises, and vibrations are
the indicators of imminent troubles. Sometimes these abnormalities are
neglected because of the insignificance or the perception that such abnor-
malities will not contribute to any major breakdowns. The tendency to
overlook such minor abnormalities soon may grow and contribute to seri-
ous catastrophic failures. It is not uncommon to receive queries from pro-
duction staff in response to a “high temperature or vibration condition”
about how long we can continue running.
It has been observed that a high percentage of the failures occur dur-
ing startups and shutdowns. However, asset failure could also be due to
poor maintenance. Causes that go unnoticed are called “hidden abnormal-
ities.” The key to achieving zero failures is to uncover and rectify these
hidden abnormalities before failure actually occurs.
In many organizations, CM is called repair maintenance; it is conduct-
ed to correct deficiencies and to make the asset work again after it has
failed or stopped working.

3.4 Other Maintenance Practices

In addition to maintenance approaches that are used to form a more


structured maintenance program, maintenance practices are also used to
define a company’s program. One of the key maintenance practices used
60 Chapter 3

by companies to effectively execute their maintenance program is related


to operator-based maintenance, including the use of operators in design-
ing for maintenance and reliability.

Operator-Based Maintenance (OBM)


Unlike what is typically assumed, the operator is actually one of the
most important members of the maintenance team. Well-informed,
trained, and responsible operators will ensure that assets are being kept in
good working order.
Operators are the first line of defense against unplanned asset down-
time. OBM assumes that the operators who are in daily contact with the
assets can use their knowledge and skills to predict and prevent break-
downs and other losses.
The main objective of an Operator’s Maintenance program (aka
autonomous maintenance program) is to equip operators with the follow-
ing asset-related skills:

• Ability to detect abnormalities


• Ability to correct minor abnormalities and restore function, if they
can
• Ability to set optimal asset conditions
• Ability to maintain optimal equipment conditions

Autonomous maintenance is one of the basic pillars of Total


Productive Maintenance (TPM). TPM is a Japanese maintenance philoso-
phy which involves operators performing some basic maintenance activi-
ties. The operators learn the maintenance skills they need through a train-
ing program. They then perform the following tasks:

• Conduct general inspection.


• Keep assets clean and all areas accessible.
• Identify and eliminate problem sources.
• Support and create cleaning and lubricating standards and proce-
dures.
• Standardize through visual workplace management.
• Implement autonomous asset management.
• Perform minor maintenance and service items, e.g., replacing
filters, lubricating, and changing oil.
• Work with the maintenance team to repair what they are unable to
perform.
Understanding Maintenance 61

The operators use the following four sensory tools to identify prob-
lem areas, then either fix them or get help to get the problems repaired
before they turn into major failures.
a. Look for any abnormalities — clean, in place, accessible.
b. Listen for abnormal noises, vibrations, leaks.
c. Feel for abnormal hot or cold surfaces.
d. Smell abnormal burning or unusual odors.

The following suggested measures could help in achieving that goal:

a) Operator Involvement
Operators can detect any abnormalities and symptoms at an early
stage and get them corrected before they turn into major failures. O&M
personnel can ensure that all the assets are properly secured and bolted.
The support structures — piping, hoses, guards, etc. — are not loose and
vibrating. These should be properly fastened.

b) Cleaning
Cleaning leads to inspection and timely detection of any incipient fail-
ures like cracks and damaged belts. Dirt and dust conceal small cracks and
leaks. If an asset is clean, we could assess if things are not working right,
e.g., leaking, rubbing, and bolt loosening, which may be an indication of
an incipient failure.
Keep assets and the surrounding area clean. A clean asset creates a
good feeling and improves employee safety and morale.

c) Lubricating
Lubrication helps to slow down wear and tear. Check if components
are being lubricated properly with the correct type of lubricants and that
oil is being changed at the proper frequencies. Don’t over-lubricate; use
the right amount. Ultrasonic guns can be used to ensure the required
amount of lubricant is used. Apply 5S plus or 6S practices to have a lubri-
cation plan, with pictures identifying all lube points and the type of lubri-
cant to be used.

d) Operating Procedures
All operating procedures available at the site should be current. Are
these procedures easily understood? Do operators know how to shut down
or provide lockout / tag out for the asset safely in case of an emergency?
Do they know what operating parameters — pressures, temperature,
62 Chapter 3

trip/alarm settings, etc., — to watch? Make sure that operators and other
support personnel have a good understanding of the answers to these
questions. It is a good practice and very desirable to have these operating
instructions laminated and attached to the asset.

e) Maintenance Procedures
Be sure that maintenance / repair procedures are current when used.
Maintenance personnel should have the right tools available to perform
maintenance correctly and effectively. Having a current procedure is an
ISO principle.
When an asset is ready to be repaired, all items identified in the work
plan should be staged at the asset site for craft personnel to execute their
work in the most effective and efficient manner. Specialized tools should
be kept at or near the asset with proper markings.
It is a good practice to laminate the procedures, drawings, part list,
wiring diagrams, logic diagrams, etc., and make them available at or near
each asset location.

f) Operating Conditions
All assets are designed to operate under specific conditions. Check
that assets are operating in the correct environment and are not being mis-
used, i.e., overloaded or unsafely used. If they are not being operated in
their designed environment, (e.g., they are being used at much higher
level of speed than normal use), take steps to see that appropriate safety
precautions are being taken and all concerned personnel are aware of the
risks involved.

g) Workforce Skills
Ensure that the workforce, operators, maintainers, and support staff
are all properly trained and have the right skill sets to operate and main-
tain the asset effectively. Although ignorance and lack of skill, etc., can be
overcome easily by proper training, the human attitude and mindset
towards asset failure is somewhat difficult to handle. It takes a lot of effort
and time to create the right culture.

h) Repair Documentation
Repair documentation — what we did, with some details — is very
important when performing an analysis. We often see entries such as
“Pump broke – repaired” or “Mechanical seal replaced.” Such entries
merely help in maintaining failure statistics, but not in failure analysis.
Understanding Maintenance 63

The challenge is usually how to make data input easy for our crafts
personnel. For a good reliability analysis, we need to have quality data to
understand how the asset was found before and after the failure, what
actions were taken to repair, parts used, time taken to repair, etc.

i) Designing for Reliability and Maintenance


If the asset is being modified or replaced, make sure that the operators
and maintainers are involved with design reviews and are part of the
improvement team. The asset should be designed with high reliability and
ease of maintenance features. This best practice will be discussed in more
detail in later Chapter 6.
TPM maintenance practice will be discussed in more detail in
Chapter 7.

3.5 Maintenance Management System: CMMS

A maintenance management system is an essential tool for all main-


tenance organizations. It can help to improve a maintenance department’s
efficiency and effectiveness and, ultimately, to get more out of assets by
streamlining critical workflows, work identification, work task planning,
scheduling, and reporting. Two types of systems are available. One type
of system is an enterprise-wide collection of modular applications such as
asset management, material resource planning, finance, and human
resources. These applications or systems interface with each other seam-
lessly and can work effectively across many locations and plants. Most of
these systems, first developed in the mid-1990s and throughout the fol-
lowing decade, are known as Enterprise Asset Management (EAM) sys-
tems. Examples of these types of systems are those from companies such
as J.D. Edwards, IFS, Oracle, PeopleSoft, and SAP. They can be expen-
sive to install as well as keep up-to-date.
Other types of systems are standalone applications related to mainte-
nance management. They can be interfaced with other enterprise-wide
systems such as Finance or Human Resource systems. These systems are
called Computerized Maintenance Management Systems (CMMS). The
CMMS name was coined in the late 1970s and 1980s when PM programs
were automated using computers. New CMMS systems have a lot more
capabilities and functionalities; they are easier to use compared to some
EAM systems. Examples of these systems are assetPoint, Champs,
DataStream, eMaint, EPEC, Ivara, Maximo / IBM, Mapcon, mPulse, and
Synergen/SPL/Oracle WAM. Over a hundred systems are available in the
64 Chapter 3

market, starting from $1,000 to over $250,000 depending upon the num-
ber of users or the size of the plant. Most of these systems are now web-
based. Basically, now there are no major differences in the way both types
of systems function so the terms CMMS and EAM are often used inter-
changeably.
CMMS / EAM systems should have the following capabilities, al-
though they are not limited to them:

1. Asset / Equipment History


2. Asset description and specifications
3. Asset Register
4. CM results of PM and CBM findings
5. Configuration Management
6. Contractor Work Management
7. Critical asset identifications
8. Drawing and technical document management
9. EPA/OSHA permits
10. FMEAs and RCM analysis history
11. Hierarchy management
12. Inventory / spares management
13. Materials – MRO Stores Management
14. MTBF and MTTR data by assets and asset types
15. Non-recurring work — failures / breakdowns
16. Pending work — Backlog
17. People Management
18. PM and CBM/PdM Work Procedures
19. PM Optimization including Reliability Analysis
20. Pressure vessel certifications
21. Recurring type work — PM, PdM/CBM,
22. Reporting — Standard and Specialized Reports
23. Timekeeping
24. Training management
25. Warrantee management
26. Work Order routing
27. Work close-out and feedback
28. Work estimating data tables and link to other resources
29. Work Identification
30. Work Order (WO) Management
31. Work planning
32. Work scheduling and resource balancing
Understanding Maintenance 65

Work Order (WO) Management Module


One of the most useful management tools in a CMMS package is
work order (WO) management or workflow process. The workflow
engine allows automatic routing of data through an optimized process,
including configurable approvals, notification, and automated transac-
tions based on user-defined business rules.
A standard workflow for routing work orders to the appropriate
approver, depending on estimated total labor and material dollars, can be
established. Furthermore, organizations can establish work limit rules by
teams and approval type for customizing authorization schemes. The sys-
tem can be set up to request the next level of authorization when the actu-
al dollar expenditure logged exceeds a user-defined percentage. Thus,
work orders or projects can be monitored for significant overruns. The
system also can be configured to allow only certain people to approve
emergency work orders.
Organizations should be able to establish elaborate business rules, if
desired. For example, a work order of a certain type and dollar value is
sequentially routed to two approvers. If the first approver doesn’t approve
the work order within a certain time period, the supervisor is notified by
email/pager. It also can designate alternate approvers under certain condi-
tions, such as when the approver is on vacation. Other features for work
order management may include:

1. Ability to create multi-step WO; for similar work, WO can be saved


as a template for future work.
2. Ability to build work standards for labor estimating.
3. Easy to input data including graphical user interface having a similar
look and feel as Microsoft office software.
4. Ability to provide status for a given workflow item directly from a
table or dashboard.
5. Ability to provide statistics such as volume of transactions that went
through a given time period, or the average time to complete a specif-
ic work activity.
6. Entering standard times for work activities to predict how long a
process should take, and to report on actual versus standard comple-
tion time.
7. Making activities mandatory or optional, depending on the character-
istics of the work type (e.g., skip the approval step if a work order is
urgent).
66 Chapter 3

PM and CBM /PdM Module


PM and Condition-Based Monitoring is another important module for
a CMMS. Some of the features for optimizing the workflow are multiple
PM triggers; schedule flexibility that accounts for seasonality, multiple
formats, zoom, and simulation; and condition monitoring for user-defined
data. Another helpful feature is task shadowing. This feature allows skip-
ping a weekly PM routine if the short-term schedule has an upcoming
monthly routine that includes the same weekly tasks.
Information from data collection systems, such as barcode-based time
reporting, Supervisory Control and Data Acquisition (SCADA) system,
and Human Machine Interface (HMI) can automatically feed into CMMS
the condition of assets and the use of maintenance labor and material. If a
variance is detected, it can be explained via drill-down to the source data.
The condition-based maintenance functionality in a CMMS can be used
to establish the control limits that trigger actions, such as issuing a work
order or paging a technician, thereby increasing workflow efficiency and
effectiveness. Most managers find it increasingly difficult to control ris-
ing maintenance costs because of inadequate or outdated procedures. The
CMMS can identify such procedures that are consuming large resources
and need reviews or updates.

Scheduling Module
Scheduling is an area where different CMMS packages provide sig-
nificant capabilities. CMMS should provide a schedule to match the work
demand for maintenance — open work orders with the labor resource
availability. Some systems compare the work backlog with a listing of
available hours, all similarly sorted and filtered. Some system displays
this data in graphs to help in workload balancing. A good way to display
this data can be a bar graph in the top half of the screen and the lists of
work orders in the bottom half.
Some CMMS packages increased their level of sophistication by
seamless linkage to home-grown or third-party project management soft-
ware. This gives users access to comprehensive features such as critical
path analysis, Gantt charting, and resource utilization optimization.
Probably the most exciting breakthrough in scheduling functionality
is the ability to perform “what-if” analysis. By playing with variables such
as estimated duration of work, work order priority, and labor availability,
the maintenance scheduler can fine-tune the schedule without having to
make a permanent change in the source data. Only after the scheduler and
craft supervisors are satisfied with the schedule, the data is frozen and the
Understanding Maintenance 67

source data updated.

Productivity and User-Centered Design


One of the most important trends seen in CMMS is the improvement
in user-centered design or usability during recent years. For those CMMS
packages that rewrote their software to become Web-based, the new,
improved user interfaces have become more user-friendly. To compare
prospective CMMS vendors, some users have even developed several sce-
narios to assess how many screens it takes to complete a given series of
tasks and over what time. Some compare how many screens or clicks are
typically needed to get the information they need. The Web-based soft-
ware packages offer much improvement over older systems, with features
such as:
1. Search toolbar
2. Bookmarks
3. Favorites
4. History pull-down, which provides a listing of screens that were
visited in the past, in chronological order, including hyperlinks
5. Back and Forward buttons to move through the last viewed
screens
6. URL toolbar, which allows the user to key in any screen address
or Web site reference (like a “go to” feature)

Another key trend in user-centered design has been the flexibility in


customizing the application to the varying needs of individuals, or tailor-
ing it to different roles such as maintenance planner, scheduler, supervi-
sor, craftsperson, and stock keeper. This trend has been widely accepted
by users, as it decreases training time, simplifies execution of day-to-day
processes, improves accuracy and speed of data entry, and facilitates
extraction of relevant information that lead to better and faster decision
making. Examples of customization capability are:
1. Security access that defines who has the access to certain fields,
screens, menus, etc., and whether data is read-only or even visi-
ble on-screen
2. Screen layouts including what fields are viewed on which screen
or tab, field labels, size and shape of each field, field position, col-
ors, tab labels and content, size and position of columnar data,
and default values
3. Language that a package displays, as well as currency used
4. Start-up or main menu, i.e., what menu options, shortcuts, report
68 Chapter 3

highlights, KPIs, dashboard elements, alarms, drill-downs, notifi-


cations, and so on that users wants on their home page, and in
what level of detail
5. Reports or searches that are customized by the user in terms of fil-
ters and sort criteria, as well as how they appear on screen and are
printed
6. Forms and templates that can make data entry easier
7. Help and error messages

Incremental gains in CMMS features and functions have made many


packages better at handling the specialized requirements of particular
industries and facilities.

Data Analysis and Reporting


Users are slowly learning the incredible power of the CMMS to take
raw data and turn it into information and knowledge that can improve
maintenance effectiveness dramatically through use of analysis tools. It’s
not enough to collect data and report on it. A CMMS can take the data and
then by using analysis tools convert them to helpful information such as
Top 12 Problem Assets (or, as some call it, “bad actors”). These problem-
atic assets can be based on dollar value or on the number of downtime
events experienced in a specific area last month, last quarter, or last year;
they can be shown graphically in ascending order (Pareto Chart format) or
in some other manner. This analysis allows users to uncover the biggest
problems first. We can then drill down on root causes such as faulty spare
parts or material from a given manufacturer or supplier, design issues,
workmanship issues indicating inadequate training of maintainer, or oper-
ator errors.
Asset management is an area where good reporting and analysis are
critical. Equipment history reports on actual labor, planned labor, materi-
al, and other costs are sometimes available. The more advanced features
include tracking maintenance costs in accordance with user-defined statis-
tics, and tracking and analysis of equipment status, problems, causes,
actions, and delay codes.
Other reporting features that help optimize work are analysis of asset
availability and performance, mean-time-between-failure (MTBF), and
drill-down capability to determine the root cause of downtime.
Some CMMS packages show MTBF and average corrective and pre-
ventive cost in a graph. Other features include depreciation schedule, cap-
ital cost budgeting for major repairs and replacements, and remaining
Understanding Maintenance 69

asset lifespan. Repair vs. replace analysis can be shown graphically,


where the cost of a new asset exceeds the historical trend cost of repair-
ing it, all based on a user-defined amortization period and inflation rate.
Many CMMS vendors have been trying to improve their failure
analysis capability in response to the ever-increasing interest in failure
modes and effects analysis (FMEA), reliability-centered maintenance
(RCM), and root cause analysis (RCA). Asset-intensive organizations are
finding it painfully slow and complex to implement these advanced tech-
niques. When these techniques are used on critical assets and systems, the
potential payback is enormous.
Another group of analysis tools for which demand is steadily growing
includes costing and budgeting tools such as life-cycle costing. Capturing
costs associated with an asset, from its procurement to its disposition,
gives management greater insight into the total cost of ownership or eco-
nomic life of various assets and asset classes. The benefits of tracking life-
cycle costs are many, including:

• Comparing the cost of various offerings of the same asset type


before e.g., comparing, say, a Caterpillar versus a Toyota forklift
with the same or similar specifications)
• Understanding the trade-off between asset performance and the
total cost of ownership (e.g., deciding how long it is economical to
hold onto a specific asset and when is it economical to replace it)
• Forecasting cost of assets based on the life-cycle cost profile of
similar assets
• Becoming aware of the costs in order to control them

Life-cycle cost analysis can be quite complex, especially for facilities


or infrastructure that require monitoring and assessment of the asset’s con-
dition and rate of deterioration. The multi-year considerations include fac-
tors such as discount rates used in the net present value (NPV) calcula-
tions. Important assumptions about how the business, market and product
alternatives will change over time.
A CMMS can help in tracking life-cycle costs by accumulating rele-
vant labor, material, contract, and overhead costs associated with a given
asset — even if the asset is moved, sent outside for repair, shows sign of
deterioration faster or slower than expected, or either depreciates or
appreciates in value. Costs other than maintenance costs must be consid-
ered; these include installation costs, operating costs, and risk abatement
(e.g., health, safety, and environmental impact).
70 Chapter 3

Mobile Technology
The popularity of mobile technology continues to rise as more users
realize its power. Meanwhile, the telecommunications networks continue
to expand their geographic reach and their ability to handle interference.
Handheld devices are also improving in terms of functionality and afford-
ability. Much of the functionality of a desktop terminal can be put in the
hands of a mobile user, including uploading and downloading work order
and spare parts inventory information, accessing equipment history and
reports, and even viewing or redlining drawings and maps. The mobile
technology is one of the most important trends being adopted in the
CMMS industry, just as the BlackBerry, Smartphones and IPads took the
business world to a whole new level.

System Affordability
The need for and use of a CMMS is not specific to any one industry
or type of application. Any organization using assets to make products or
providing services is a potential candidate for a CMMS.
Computerized systems are becoming more attractive as more mainte-
nance personnel have become computer literate and prices of hardware
and software have dropped significantly. These factors make a CMMS an
attractive option for even smaller plants. CMMS packages are available in
modular format. In other words, organizations don’t have to buy all the
modules and options. For example, smaller plants can purchase only the
asset, PM, and work order modules to start. They can add other modules
later on. Also, many CMMS programs are designed with scaled-down
functionalities for smaller plants. These programs are fully functional and
relatively inexpensive. However, organizations must determine if a
CMMS is beneficial to their operations and have buy-in from all stake
holders.
Workforce average age and continuity of the organization’s knowl-
edge base is another important issue to consider. How much information
will leave the company when a key maintenance employee retires? Years
of critical information can be lost the moment that employee walks out the
door.

Barriers to CMMS Acquisition


Opposing the CMMS acquisition are the internal roadblocks that
stand in the way of the system purchase, particularly in smaller organiza-
tions. The following list can help overcome barriers associated with
acquiring a CMMS:
Understanding Maintenance 71

1. Organization is too small for a system. — This attitude suggests a


basic lack of understanding of the true benefits and functions of a CMMS.
A CMMS that is ideally matched to the organization’s needs should pay
for itself, even for very small plants. There are many plants with just a few
maintenance technicians successfully using a CMMS. A CMMS can help
record and maintain the equipment histories that will be the basis for
future repair vs. replacement decisions and associated justifications. An
accurate and complete history can also describe how the job was execut-
ed last time, thereby, saving time associated with a job or task redesign.

2. The project payback or savings is inadequate. — Maintenance


organization must do a thorough job of determining benefits and savings
in order to show real ROI of a well-chosen CMMS.

3. MIS or IT doesn’t give CMMS high enough priority. — This lack


of sufficient support is a very common situation. If MIS supports the proj-
ect, its chances of success are increased greatly. A CMMS is complicated
to many decision makers. Helping MIS understand the importance of
CMMS should be a primary goal of the maintenance department.

4. MIS and maintenance speak different technology languages. —


Being able to translate the technological and business benefits effectively
can go a long way toward overcoming this roadblock. With MIS support,
it is easier to convince others.

5. Participants fail to reach consensus. — When the parties involved


disagree on either the need for a CMMS or the features required in a
CMMS, it becomes difficult to gain approval for CMMS funds. There
must be an internal champion who is empowered to select a team and act
on results.

Selecting the Right CMMS


Selecting the right CMMS is crucial to a successful implementation.
Some suggested guidelines are discussed below.

System Features
There are numerous features that the system should include. One such
feature is flexibility. The CMMS should be flexible in terms of allowing
users to enter information pertaining to your organization. It should also
accommodate both present and future needs. Organizations should also be
72 Chapter 3

aware of the system’s limitations. Before buying a particular system, one


should check the limitations of the system. For example, if the number of
records in the database increases significantly in the future, the system’s
searching and reporting capabilities should not be slowed down.
Another feature to consider is the system’s interfacing capabilities.
The CMMS should be capable of interfacing with other information sys-
tems. Self-sufficiency is something else to consider. Programs should be
capable of direct, full use without needing to consult a manual or other
outside sources. The on-screen instructions should explain what the pro-
gram will do and how to use it. Other features to consider include the sys-
tem’s security, data security, modifications, user customizable screens,
and user customizable reports.

Ease of Use
The CMMS should be easy to learn and should come with training
aids and documentation. It should also be easy to use. The package should
be icon and menu driven, contain input screens to enter information in an
orderly manner, and provide error handling and context-sensitive help.

Vendor Support
Consider the qualifications of the potential CMMS vendors.
Obviously we want a vendor who is both knowledgeable and experienced
when it comes to CMMS. Also consider the vendor’s financial strength. A
CMMS project is an investment in time, resources, and money. Therefore,
the vendor must be established. Ask about references, delivery, payment
options, source code, and warranty.
Also, investigate the level of vendor support for training. Whether this
training is provided at their facility or on site, this small investment can
save a great deal of money and frustration in the long run. Other factors
to consider include the vendor’s system support, upgrade policy, and over-
all system cost. Select the vendor that provides the best combination of
characteristics for your particular situation.
The bottom line is that there a need for a CMMS for maintenance no
matter how small or large plant is. We should be aware of the barriers and
be well prepared to face them during the justification process. We can
avoid failure by looking at why so many installations have failed and
making the right selection for application for the organization.
Understanding Maintenance 73

Why So Many CMMS Projects Fail


Many CMMS projects fail to reach their full potential. The following
are some of the reasons:
1. Selecting the wrong CMMS for your application
2. Employee turnover
3. Lack of adequate training during implementation
4. Employee resistance
5. Being locked into restrictive hardware/software
6. Inadequate supplier support for the CMMS
7. High expectations and quick return on investment
8. Internal Politics — Financial or IT — heads the CMMS/EAM
implementation team

Note: Ideally, senior M&R professionals should lead the project


because they understand the need more than anybody else. Keep IT and
Finance employees on the team to get their continued support.
In general, incremental gains in CMMS features and functions have
made many packages better at handling the myriad and specialized
requirements of particular industries and facilities. Many organizations
are simply looking for a CMMS package that meets their unique needs,
and a vendor that understands their industry. Some examples are as fol-
lows:
1. Pharmaceutical organizations require a CMMS that has an elec-
tronic signature capability to comply with FDA 21 CFR Part 11.
2. Municipalities are looking for a CMMS package with sophisticat-
ed linear asset functionality for handling utilities, underground
water and wastewater networks, and so on. They need capability
to be in compliance with GASB 34 requirements.
3. Organizations with considerable mobile equipment are looking
for fleet management functionality, such as compliance with the
vehicle maintenance reporting standards (VMRS) coding struc-
ture or the American Trucking Association’s standard listing of
components.
4. Pipeline companies need inspection and risk assessment features.
5. Chemical, oil and gas, and nuclear plants need sophisticated safe-
ty-related functionality such as lockout/tag out.
6. Third-party service providers want features such as contract man-
agement, third-party billing, help desk, and dispatch.
7. Government departments are keen on sophisticated budgeting
capability including encumbrance accounting.
74 Chapter 3

8. Asset-intensive companies experiencing considerable capital


expansion can save millions of dollars if the CMMS can help inte-
grate and follow the complete life cycle of asset-related data from
engineering design to deployment to maintenance.

An effective CMMS has proven to be an invaluable tool for many


plants and facilities. Various internal obstacles to acquiring a CMMS also
confront businesses. When deciding to acquire a CMMS, take the follow-
ing steps:
a. Form a team.
b. Identify problems with the existing system.
c. Define the objectives, features, and benefits of a CMMS.
d. Conduct a financial analysis.

And identify a project manager, preferably a senior M&R profession-


al, not an IT or financial person, to lead the project.
New CMMS packages are so feature-rich that most users can hope to
exploit only a small percentage of the functionality they buy. It has been
reported that most of the users utilize less than 50% of the capabilities and
functionalities of the CMMS they have. The true differentiation is how the
CMMS is implemented. Organizations should set quantifiable goals and
objectives, re-engineer processes in light of those goals, configure the
CMMS to optimize the new processes, and change behavior across organ-
ization to embrace these changes.

3.6 Maintenance Quality

It is said that “Accidents do not happen, they are caused”. The same
is true for asset failure. Assets fail due to basically two reasons: poor
design and human error. Our negligence, ignorance, and attitude are the
prime factors of human errors. Several studies have indicated that over 70
percent of failures are caused by human errors such as overloading, oper-
ational errors, ignoring failure symptoms and not repairing an asset when
it needs to be taken care, and the skill level of our work force. There is
usually a human factor behind most asset failures. Because most failures
are caused and do not happen independently, they are preventable.
If a survey is taken among operations and maintenance personnel as
to whether there can be zero failures, the overwhelming answer will be
zero failures are theoretically possible, but impossible in an actual work
environment. Yes, zero failures are difficult to achieve, but they may not
Understanding Maintenance 75

be impossible. If all concerned operations and maintenance personnel set


a goal of zero failures and diligently work toward that goal, it is attainable.
However, total commitment is needed from all involved, from top man-
agement to supervisors and down to the operator and maintainer level.
What we need to do is implement some good and best practices, as well
as strict adherence to the procedures.

Quality of Maintenance Work


All maintenance work involves some risk. Here, the risk refers to the
potential for inducing defects of various types while performing the main-
tenance tasks. In other words, human errors made during the PM, CBM,
and CM tasks eventually may lead to additional failures of the asset on
which the maintenance was performed.
For example, a review of the data from the power plants that exam-
ined the frequency and duration of forced outages after a planned mainte-
nance outage reinforces this risk. The analysis of data showed that in 55%
of the cases, unplanned maintenance outages were caused by errors com-
mitted during a recent maintenance outage. Most of the time these failures
occur very soon after the maintenance is performed. Typically, the follow-
ing errors or damages may occur during PMs and other types of mainte-
nance work.
Damage to the asset receiving the PM task may include such things
as:
• Damage during the performance of an inspection, repair adjust-
ment, or installation of a replacement part.
• Installing material/ part that is defective
• Incorrectly installing a replacement part, or incorrectly
reassembling
• Reintroducing infant mortality by installing new parts which have
not been tested
• Damage due to an error in reinstalling asset into its original
location
• Damage to an adjacent asset or component during a maintenance
task

A quality maintenance program requires trained and motivated main-


tenance personnel. To create high quality and motivated personnel, the
following measures are suggested:
a. Provide training in maintenance best practices and procedures for
maintenance on specific assets.
76 Chapter 3

b. Provide appropriate tools to perform the tasks effectively.


c. Get them involved in performing FMEA and RCA/RCFA, and in
developing maintenance procedures.
d. Follow up to assure quality performance and to show everyone
that management does care for quality work.
e. Publicize reduced costs with improved uptime, which is the result
of effective maintenance practices.

3.7 Maintenance Assessment and Improvement

Maintenance Key Performance Indicators (KPI)


It is often said that “what gets measured gets done” and “If we can’t
measure it, we can’t improve it.” KPIs, also called metrics, are an impor-
tant management tool to measure performance and help us make improve-
ment actions. However, too much emphasis on performance indicators, or
on the wrong indicators, may not be the right approach. The selected indi-
cators shouldn’t be easy to manipulate just to “feel good.” The following
criteria are recommended for selecting the best KPI / metrics:

• Should encourage the right behavior


• Should be difficult to manipulate
• Should be easy to measure — data collection and reporting

Some key maintenance metrics, with some benchmark data, are listed in
Figure 3.1.

Figure 3.1 Maintenance Benchmarks


Understanding Maintenance 77

Other maintenance metrics to consider, depending on the maturity of


the maintenance program, include:

• PM and CBM effectiveness, or the number of hours of corrective


work identified by PM and CBM work divided by hours spent on
PM and CBM inspections. The PM and CBM should be able to
identify 1/2 – 2 hours of corrective maintenance work to every one
hour of PM and CBM performed; otherwise, the frequency or
condition parameters should be reviewed or adjusted.
• PM and CBM schedule adherence. It should approximate 90% or
more.
• Percent of maintenance labor dedicated to performing PM and
CBM inspections should be more than 50 percent. The rule of
thumb is:
• PM — Time based 15–20%
• CBM 30–40% — The distribution may vary depending on type
of asset and industry.

Maintenance Task Optimization


Maintenance effectiveness can be improved by optimizing the main-
tenance work tasks (content) and by effective task execution through the
utilization of the many tools available to us. The maintenance tasks —
e.g., PM, CBM work instructions, and repair plans — must cover what
needs to be done. These tasks can be optimized by using tools and tech-
niques such as FMEA, RCM, and predictive technologies. These tools and
techniques can help to optimize the content of the work tasks to be accom-
plished.

Establishing a Successful Maintenance Program


Scheduling and execution are the keys to a successful maintenance
program. Maintenance programs should be automated by using
Computerized Maintenance Management Systems (CMMS) or Enterprise
Asset Management (EAM) systems. In addition, a monitoring process
should be established to ensure a 90% or better schedule compliance and
quality of work performed.
In addition, create a “living maintenance program” consisting of the
following key elements:

• Continually review processes, procedures, and tasks for applicab-


ility, effectiveness, and interval frequency; these should be opti-
78 Chapter 3

mized as required. Get the right people both in operation and


maintenance involved in the review process.
• Standardize procedures and maintain consistency on asset and
components.
• Identify and execute mandated tasks to ensure regulatory
compliance.
• Apply and integrate new predictive technologies where
effective.
• Ensure task instructions cover lockout / tag out procedures and all
safety requirements.
• Ensure operation and maintenance personnel understand the
importance of PM practice and provide feedback for improving
PM instructions and procedures.

3.8 Summary

Maintenance prevents an asset or item from failing and repairs it after


it has failed. However, the new paradigm for maintenance is capacity
assurance, meaning that maintenance assures asset capacity as designed
or to an acceptable level.
Maintenance approaches – practices can be classified in the following
categories:

• Condition-based Maintenance (CBM)


• Preventive Maintenance (PM)
• Time (Calendar) based maintenance (TBM)
• Run-based maintenance (RBM)
• Operator-based maintenance (OBM)
• Proactive maintenance
• Corrective Maintenance (CM)
• CM — Planned and Scheduled
• CM — Major Repairs / Projects (Planned and Scheduled)
• CM — Reactive (Breakdowns / Emergency)

Computerized Maintenance Management Systems are essential data-


based decision making tools for managing asset. A CMMS or EAM sup-
ports a maintenance department to ensure that assets and systems operate
efficiently, and to minimize downtime. They help to improve maintenance
effectiveness in any organization.
Understanding Maintenance 79

All maintenance tasks involve some risk of introducing defects of var-


ious types while performing the maintenance tasks. In other words, errors
committed during the PM, CBM, and CM tasks eventually may lead to
additional failures of the asset on which the maintenance was performed.
A maintenance quality program requires trained and motivated mainte-
nance personnel.
Selecting the right performance indicators to measure maintenance
performance is critical and important to implementing best practices. The
indicators should encourage the right behavior; they should be difficult to
manipulate just to have “feel good” results. Finally, they should be easy
to collect and report.
Maintenance cost and asset availability can be improved by optimiz-
ing the maintenance work tasks (content) and by effectively executing
tasks through the utilization of tools available to us. Maintenance tasks
such as PM/CBM work instructions and repair plans must cover what
needs to be done. These tasks can be optimized by using tools and tech-
niques such as RCM, FMEA, predictive technologies, and Six Sigma.
These tools and techniques help to optimize the content of the work tasks
to be accomplished. The execution of maintenance tasks can also be opti-
mized by using other tools and techniques such as planning and schedul-
ing. These tools and techniques can help to utilize maintenance resources
effectively.

3.9 Self Assessment Questions

Q3.1 Define maintenance and its role.

Q3.2 What are the different categories of maintenance work?

Q3.3 What can equipment operators do to support maintenance?

Q3.4 Why would an organization support operators getting involved


in maintenance?

Q3.5 Why would an organization need to have a CMMS? What is


the difference between a CMMS and an EAM?

Q3.6 List five maintenance metrics and discuss why they are
important.
80 Chapter 3

Q3.7 Name five PdM technologies. Discuss how they can help
reduce maintenance costs.

Q3.8 Define Proactive Maintenance.

Q3.9 What are the benefits of a structured maintenance program?

Q3.10 How can a CMMS/EAM system help improve maintenance


productivity?

3.10 References and Suggested Reading

Levitt, Joel. Handbook of Maintenance Management. Industrial Press,


1997.
Mitchell, John. Physical Asset Management Handbook, 4th Edition.
Clarion Publishing, 2006.
Narayan, V. Effective Maintenance Management. Industrial Press, 2004.
Nyman, Don. Maintenance Management training notes. Seminars,
1994–96.
Bagadia, Kris, David Burger, et al. Miscellaneous technical papers and
reviews on CMMS/EAM.
Plant Services, Maintenance Technology, and Reliabilityweb.com.
Chapter 4
Work Management:
Planning and Scheduling

“A goal without a plan is just a wish”


Antoine de Saint-Exupery

4.1 Introduction
4.2 Key Terms and Definitions
4.3 Work Flow and Roles
4.4 Work Classification and Prioritization
4.5 Planning Process
4.6 Scheduling Process
4.7 Turnarounds and Shutdowns
4.8 Measures of Performance
4.9 Summary
4.10 Self Assessment Questions
4.11 References and Suggested Reading

81
82 Chapter 4

After reading this chapter, you will be able to understand:

• The basic work flow process


• The role of planners, schedulers, and others in managing work
• Work classification and prioritization
• The importance of backlog management
• Why planning is necessary
• The planning process
• Why scheduling is necessary
• The scheduling process
• Turnaround management

4.1 Introduction

Maintenance tasks should be performed efficiently to ensure plant


capacity is sustained cost effectively. In the previous chapters, we dis-
cussed developing the proper maintenance tasks to keep our assets work-
ing. In order to reduce overall operations and maintenance costs, these
tasks must be executed efficiently and effectively. Basically this is
achieved by eliminating or minimizing avoidable delays and wait time.
Imagine yourself repairing a leaky faucet or dishwasher at home. Your
spouse has asked you repeatedly to fix it. Finally you find the time to take
on this assignment. Can you recall the number of times you went back and
forth to the garage, to the tool box, or to the hardware store to acquire the
correct-sized tool, washer, or seal? It probably took about four or more
hours to finish this task.
Imagine again, a couple of months later, a similar type of problem
occurred. This time you are not available and your spouse calls a plumber.
The plumber comes in and assesses the problem, goes back to the truck,
gets the right tools and parts, corrects the problem, and leaves in 40–45
minutes. Does this sound familiar? Maybe if you had the right tools and
right parts, and better instructions, your task would have taken less than 2
hours instead of 4 hours? The point here is, proper work planning with the
right tools, parts, and instructions can save time and avoid wasteful activ-
ities.
Figure 4.1 shows a job without planning (sometimes called “on–the-
run” planning) and one with proper planning. Figure 4.1a shows a disor-
ganized work activity with frequent work interruptions and restarts, evi-
dence of inadequate planning.
Work Management: Planning and Scheduling 83

Figure 4.1 Impact of Planning

The frequent work interruptions encountered could be due to a lack of


availability of the right parts or tools, or improper work instructions. A
well-planned job with upfront planning and no interruptions is shown in
Figure 4.1b. Planned and scheduled jobs take substantially less time than
unplanned jobs.
For many years, industry experts have pointed to the low productivi-
ty levels in maintenance departments of many companies around the
world. Several studies and survey results reported at major Maintenance
and Reliability conferences such as IMC (International Maintenance
Conference) and SMRP have indicated that maintenance craft productiv-
ity varies anywhere from 30–50%, or 3–4 hours of average productive
time for an 8-hour shift. Some call this productive time “wrench time,”
during which maintenance craft personnel actually spend their efforts
repairing the assets, as opposed to walking to the store to get the right
tools, receiving unclear instructions, waiting for other craft to arrive or
release of the asset from operations, and other wasteful activities.
In general, every hour invested in work planning saves 1–3 hours in
work execution. Abraham Lincoln once said “If I had eight hours to cut a
tree, I’d spend six hours in sharpening the axe.”
There are, of course, some managers who say they would be thrilled
to hear that their maintenance craft workers are sitting idle most of the
time waiting for breakdowns to happen. The “Maytag repairman” image
of a maintenance department should not be compared to a fire department,
where fewer fires to battle are better. A maintenance department can be far
more productive in so many ways, becoming proactive instead of
84 Chapter 4

responding to emergencies like fire departments. Maintenance depart-


ments should be performing preventive and condition-based maintenance
tasks, participating in process improvement projects, and working on cap-
ital improvement initiatives. Maintenance workers can upgrade their
skills, train others, and educate operators to run the assets properly to min-
imize errors. In essence, good planning and scheduling avoids delays and
minimizes wait time, other wasteful activities, and non-productive work.
Planning and scheduling (P&S) is a disciplined approach both for uti-
lizing maintenance resources effectively and for executing maintenance
tasks such as PM/CBM or corrective maintenance tasks efficiently. This is
accomplished through:

• Defining and clarifying the right work


• Prioritizing work
• Developing the work sequence and steps to complete the task
• Identifying necessary tools, materials, and skills sets
• Assuring on-schedule availability of materials and assets
• Scheduling the work to be done with agreement from
production on scheduled time
• Ensuring details of completed work are documented in CMMS

A work plan is the key deliverable of the planning process. This


product is where the largest gains in productivity can be made. In some
organizations, a single person provides both planning and scheduling
functions. In larger organizations, these functions are often split, allow-
ing additional resources for each role.
To move from reactive to proactive maintenance, at least 80% of the
work should be planned on a weekly basis. Compliance to this work
schedule should be at least 90%.
In this chapter, we will discuss the “what and how” of planning and
scheduling maintenance tasks so that they can be executed effectively as
well as some associated topics (work flow and roles, work categorization
and priority, turnarounds, etc.). Some of the key terms we will be using
follow.
Work Management: Planning and Scheduling 85

4.2 Key Terms and Definitions

Bill of Material (BOM)


A list of materials needed to complete a particular assembly or
fabrication job. The BOM can also be a listing of items neces-
sary to support the operations and maintenance of an asset or
component.

CMMS / EAM (Computerized Maintenance Management


System / Enterprise Asset Management)
A software system that keeps record and tracks all maintenance
activities, e.g., maintenance work orders, PM schedules, PM
masters, material parts, work plans, and asset history. Usually it
is integrated with support systems such as inventory control,
purchasing, accounting, manufacturing, and controls mainte-
nance and warehouse activities.

Coordinators
Individuals who oversee the execution of all work within a facil-
ity, including maintenance. They are accountable to the asset or
process owner for insuring that the asset or process is available to
perform its function in a safe and efficient manner and to help pri-
oritize the work according to the operational needs.

Planned Work
Work that has gone through a formal planning process to identi-
fy labor, materials, tools, work sequence, safety requirements,
etc., to perform that work effectively. This information is assem-
bled into a job plan or work package and is communicated to craft
workers prior to the start of the work.

Planner
A dedicated role with the single function of planning work tasks
and activities.

Planning
The process of determining the resources and method needed
including safety precautions, tools, skills, and time necessary to
perform maintenance work efficiently and effectively. Planning is
86 Chapter 4

different from scheduling. In short, planning defines what and


how, whereas scheduling defines who and when.

Preventive Maintenance (PM) Schedule Compliance


The number of PM work orders (or labor hours) completed,
including PdM/CBM, divided by the total number of PM work
orders (or labor hours) scheduled during a specific time period.

Schedule Compliance
A measure of adherence to the schedule. It is calculated by the
number of scheduled jobs (or scheduled labor hours) actually
accomplished during the period covered by an approved
daily/weekly schedule, expressed as a percentage.

Scheduled Work
The work that has been identified in advance and is logged in a
schedule so that it may be accomplished in a timely manner based
upon its criticality.

Schedulers
Individuals who establish daily, weekly, monthly, and/or rolling
yearly maintenance work schedules of executable work in their
facility. The schedule includes who will perform and when the
work will be performed. The schedule is developed in concert
with the maintenance craft supervisor and operations.

Scheduling
The process of determining which jobs get worked on, when, and
by whom based on the priority, the resources, and asset availabil-
ity. The scheduling process should take place before the job is
executed. In short, scheduling defines when and who execute the
work tasks.

Turnaround
Planned shutdown of equipment, production line, or process unit
to clean, change catalyst, and make repairs, etc., after a normal
run. Duration is usually in days or weeks; it is the elapsed time
between unit shutdown and putting the unit on-stream/online
again
Work Management: Planning and Scheduling 87

Work Order (WO)


Paper or electronic document specifying the work needed on an
asset. A work order is a unique control document that comprehen-
sively describes the job to be done, including a formal requisition
for maintenance, authorization, and charge codes.

Work Order Parts Kitting


The collection and staging of parts required for each individual
work order. This step is usually accomplished in a plant’s store-
room within the maintenance shop. Each kit is identified by a
number or label so that it can be staged or delivered to the right
maintenance crew.

Work Plan
An information packet, sometime called job or work package,
provided to the worker; it contains job specific requirements such
as task descriptions sequenced in steps; job specific instructions;
and safety permits/ procedures, drawings, materials, and tools
required to perform the job effectively.

4.3 Work Flow and Roles

Figure 4.2 illustrates a simple maintenance work flow process. There


are three types of work:

• PM work including CBM/PdM


• CM — New work resulting from PM / CBM activities
• CM — Breakdown/emergency work (Reactive)

Preventive Maintenance (PM) work should have already been


planned, therefore going directly to scheduling. Corrective Maintenance
(CM)—breakdown/emergency work—can be executed while bypassing
the planning process, and sometime even the scheduling process based on
its urgency, depending on whether there is enough time to plan this type
of work. The new CM work identified from PM tasks, including CBM
activities, should be planned and scheduled before it is executed.
88 Chapter 4

Figure 4.2 Simple Work Flow Process

Figure 4.3 illustrates the work flow and key players in the mainte-
nance work flow process. The following are the key players in this
process:

• Coordinator—Asset / Resource
• Planner
• Scheduler
• Configuration Specialist/Systems Engineer
• Craft Supervisor
• Work Performer

In addition, other players such as maintenance / systems engineers


and MRO-material personnel play supportive roles in the work flow
process.
Initially, the required or requested work task gets routed to an asset /
resource coordinator. This person represents the asset owner and may
work for maintenance or operations. The coordinator helps to prioritize
the work, insuring required resources are in the budget, and to schedule
asset outages, if necessary. The coordinator forwards the work task to a
planner, scheduler, or directly to the craft supervisor or maintenance crew,
depending on the task’s priority and planning needs. For example, PM-
type work, which should already be planned, could go directly to the
maintenance scheduler. The coordinator may also work with the mainte-
nance engineer or configuration management personnel for any technical
help or if a configuration change request is needed.
Work Management: Planning and Scheduling 89
Figure 4.3 Work Flow with Role Process
90 Chapter 4

As the work order (WO) gets routed from one stage to another, a WO
status is assigned based on what’s being done to that WO. Figure 4.4 is a
suggested list of Work Order Status Codes. In addition, work type, as sug-
gested in Figure 4.5, is also assigned by the coordinator or the
planner/scheduler. It is a good practice to code the work orders to help
analyze the data for improvements. More will be said about work order
classification in the next section.
Maintenance planners plan the job and create a work plan or job pack-
age that consists of what work needs to be done; how it will be done; what
materials, tools, or special equipment are needed; estimated time; and
skills required. The planners need to identify long delivery items and work
with stores and purchasing personnel to insure timely delivery. Planners
may need to work with maintenance / systems engineers and craft super-
visors for technical support to insure that the work plan is feasible with
sufficient technical details.
Maintenance schedulers—in working with the craft supervisor, coor-
dinator, and other support staff—develop weekly, monthly, and rolling
annual long-range plans to execute maintenance work. They are more
concerned with when the job should be executed in order to optimize the
available resources with the work at hand.

Figure 4.4 Work Order Status Codes


Work Management: Planning and Scheduling 91

Figure 4.5 Work Type Category Codes

Craft supervisors take the weekly schedule and assign who will do the
job on a daily basis. In addition, they also review work plans from an exe-
cution point of view and recommend necessary changes in work plans to
the planner and the scheduler. It is also their responsibility to ensure that
the high work quality is maintained and details of work completed are
documented properly in the system.
Figure 4.6 illustrates a work flow process with its key elements; it
includes an example of a productivity report based on delay hours
reported.

4.4 Work Classification and Prioritization

Maintenance Work Task Classifications


Maintenance work tasks can be classified in two major categories:
Preventive and Corrective.
92 Chapter 4
Figure 4.6 Work Flow Process.
Work Management: Planning and Scheduling 93

Preventive Maintenance (PM)


• Time (Calendar)-based maintenance (TBM) (age related)
• Run-based maintenance (RBM) (usage related)
• Condition-based maintenance (CBM aka Predictive)
(health related)
• Operator-based maintenance (OBM aka Autonomous
Maintenance, a pillar of TPM) (operations related)

Corrective Maintenance (CM)


• CM Routine work resulted from PMs: Planned and Scheduled
• CM Major Repairs/Projects: Planned and Scheduled
• CM Reactive:Unplanned/Unscheduled (aka Breakdown/
Emergency)

Preventive Maintenance (PM)


Preventive maintenance refers to a series of actions that are performed
on an asset on schedule. That schedule may be either calendar time-based
or machine operations dependent (i.e., runtime or the number of machine
cycles). These actions are designed to detect, preclude, or mitigate degra-
dation of a system and its components. PM includes cleaning, adjusting,
and lubricating, as well as minor component replacement, to extend the
life of assets and facilities. The goal of a preventive maintenance approach
is to minimize system and component degradation and thus sustain or
extend the useful life of the asset. Assets within your facilities should not
be allowed to run to the breaking point unless a run-to-failure strategy has
been selected for that specific asset. The PM work can be further classi-
fied into four categories:
PM—Calendar-Based
PM—Calendar-Based Maintenance (Time-Based Maintenance or
TBM) is typically performed based on the calendar time. Maintenance
personnel schedule periodical visits to an asset based on fixed time inter-
vals, for example, every three or six months. Although better than no PM
at all, calendar-based PMs are not the optimal way to run PM programs.
They may result in too much time being spent on an asset. Numerous vis-
its to assets with “no data – abnormalities found” can be regarded as wast-
ed maintenance dollars. If this happens, the PM periodicity should be
reevaluated and adjusted. Nevertheless, time-based PMs are a good
approach for assets having fixed operating schedule such as 24/7 or 80
hours/week operation.
94 Chapter 4

PM—Run-Based
PM—Run-Based Maintenance (RBM) is typically the next step up
from calendar-based maintenance. It involves performing PMs based on
asset cycles or runtime. Intuitively, this approach makes sense. An asset
does not have to be checked repeatedly if it has not been used. Generally
speaking for some failure modes, it is the actual operation of the asset that
wears it down, so it makes sense to check the asset after it has been work-
ing for a specified amount of time to cause some wear. It may be neces-
sary either to adjust or replace the component.
PM—Condition-Based
PM—Condition-Based Maintenance (CBM), also known as
Predictive Maintenance (PdM), attempts to evaluate the condition of an
asset by performing periodic or continuous asset monitoring. This
approach is the next level up from runtime-based maintenance. The ulti-
mate goal of CBM is to perform maintenance at a scheduled point in time
when the maintenance activity is most cost effective yet before the asset
fails in-service. The “predictive” component stems from the goal of pre-
dicting the future trend of the asset’s condition. This approach uses prin-
ciples of statistical process control and trend analysis to determine at what
point in the future maintenance activities will be appropriate and cost
effective.

PM—Operator-Based
PM—Operator-Based Maintenance (OBM) uses the fact that opera-
tors are often the first line of defense against unplanned asset downtime.
OBM assumes that the operators who are in daily contact with the assets
can use their knowledge and skills to predict and prevent breakdowns and
other losses. OBM is synonymous with autonomous maintenance, one of
the basic pillars of Total Productive Maintenance (TPM). TPM is a
Japanese maintenance philosophy that involves operators performing
some basic maintenance activities. The operators learn the maintenance
skills they need through the training program and use those skills on a
daily basis during operations.

Corrective Maintenance (CM)


CM, sometimes called repair, is performed to correct the deficiencies
found during PM and CBM assessment; it restores the asset in good work-
ing condition after it has failed or stopped working. CM is also an action
Work Management: Planning and Scheduling 95

initiated as a result of an asset’s observed or measured condition before or


after the functional failure. The CM work can be further classified into
three categories:
CM—Scheduled
CM—Scheduled is a repair activity performed to mitigate potential
asset failure or correct deficiencies found during PM and CBM tasks. It
brings an asset to its designed capacity or to an acceptable level in a
planned way. This work should be planned and scheduled.
CM—Major Repairs / Projects (Planned and Scheduled)
In many organizations, all major repairs or improvement work valued
over a certain threshold—e.g., overhauls and turnaround projects—are
treated as capital projects for tax purposes. If these projects are to bring
the asset back to the designed capacity, not to add additional capabilities,
they should be treated as corrective maintenance. In that case, they should
always be planned and scheduled.
CM—Reactive (Unscheduled) aka Breakdowns / Emergency
Corrective Maintenance—Reactive (Unscheduled) is basically repair-
ing of the assets after they fail. This work is also known as breakdown or
failure repair work. Most of the time, completing this work interferes with
the regular weekly schedule. Unscheduled work costs much more than
planned and scheduled work.

Some maintenance professionals classify maintenance in the follow-


ing categories: PM, CBM/PdM, Proactive Work resulting from PM and
CBM/PdM, and CM—Reactive (Breakdowns / Emergency). It really does
not matter how we classify them as long as maintenance management sys-
tems can provide us data in the format to help us to make the right deci-
sions. Our objective is to reduce reactive breakdowns and then adjust or
increase PM and CBM work accordingly.
Sometimes we try to mix maintenance work types with how we
respond to get the work done. For example, is emergency work really
CM—Unplanned/Unscheduled or is it reactive work that needs to be done
now? In some organizations, the breakdown work is called urgent main-
tenance, but could be done within 48 hours. Some regular work, also
sometimes called routine work, may need to be completed in 5 or 7 days.
These examples are not the work type, but just how we respond to get it
done.
96 Chapter 4

Sometime a decision is made to take no actions or make no efforts to


maintain the asset as the original equipment manufacturer (OEM) origi-
nally intended. Therefore, no PM program is established for that particu-
lar asset. This maintenance strategy, called Run-to-Failure (RTF), should
be applied only after a risk to the business has been analyzed and its cost
effectiveness determined. In reality, this work should not be considered
failure or reactive work because we made the decision in advance not to
perform any PM or CBM based on economic justification.
All maintenance work-tasks are needed to be documented and clas-
sified in CMMS/EAM system as described earlier and then, to optimize
resources, all work-tasks are prioritized per the organization’s priority
system for execution. Every organization should, if don’t have one;
establish a work-tasks order priority system.

Job Priority
Priority codes allow ranking of work orders to get work accom-
plished in order of importance. Too many organizations neglect the ben-
efits of a clearly-defined prioritization system. Organizational discipline
that comes through communication, education, and management support
is key to the correct usage of priority codes.
Many organizations have more than one prioritization systems; how-
ever, most of them have been found to be ineffective. The drawbacks of
not clearly defining the priorities include:

• Wasted maintenance man-hours on tasks of low relative


importance
• Critical tasks being lost in the maintenance backlog
• Dissatisfied operations customers
• Lack of faith in the effectiveness of the maintenance
delivery functions

A disciplined method of prioritization will eliminate tasks being done


on a whim and instead allow work to proceed according to its true impact
on the overall operations of the plant. It will also allow the maintenance
delivery function to be executed in a far more effective manner.

Priority System Guidelines


The system needs to cater to the following requirements equally and
provide a universal method of coding all works orders.
Work Management: Planning and Scheduling 97

• Plant-wide asset priorities, allowing for better plant-wide


utilization of resources
• Operations requirements
• Improvement projects

Accurate prioritization covers two distinct decision-making process-


es. These are:
• Asset criticality
• Impact of task or work to be done on overall operations

The original priority of the work orders needs to be set by the origi-
nator of the work order and should be validated by the coordinator. The
work originator is the most qualified to make an initial assessment of asset
criticality and impact of the work. Listings of major assets and their crit-
icality will help in decision-making for final priority ranking. Lower crit-
icality items or areas will then be easier to recognize. The following cri-
teria can be used to assign asset criticality and work impact (if not correct-
ed), which can then be used to make an objective assessment of overall
job priority.

Asset Criticality
Criticality # Description

5 Critical safety-related items and protective devices


4 Critical to continued production of primary product
3 Ancillary (support) system to main production process
2 Stand-by unit in a critical system
1 Other ancillary assets

Work Impact, If Not Corrected


Work Impact # Description

5 Immediate threat to safety of people and/or plant


4 Limiting operations ability to meet its primary goals
3 Creating hazardous situations for people or machinery,
although not an immediate threat
2 Will affect operations after some time, not immediately
1 Improve the efficiency of the operation process
98 Chapter 4

Work Priority = Asset Criticality X Work Impact


WO #1:
Asset Criticality of 5 and Work Impact of 4 gives an over
all job priority of 20.

WO #2:
Asset Criticality of 4 and Work Impact of 4 gives an over
all job priority of 16.

In this case, WO #1 will have the higher priority when compared to


WO #2. The combination of the criticality and impact of the work can be
cross-referenced to give a relative weight to each task when compared to
all other work.

Backlog Management
The combination of work classification and job priority allow an
organization to make sense out of their maintenance backlog. A mainte-
nance backlog is very simply the essential maintenance tasks to repair or
prevent equipment failures that have not been completed yet. By classify-
ing these maintenance tasks into different categories and then prioritizing
within those categories, maintenance backlogs can be developed from an
overall organizational perspective or within smaller organizational groups
or categories (e.g., PM, CM).
Why manage your backlog at all? Why not just work whatever main-
tenance tasks come due? The more toward proactive maintenance that an
organization moves, the more likely it is, at least in the beginning, that the
organization will identify more maintenance tasks than can possibly be
addressed within that immediate time period (typically that week).
Therefore, to keep from addressing the low priority tasks or the categories
of work that will not have the largest impact to the overall reliability of
the organization, a backlog management system must be developed. Then
the most effective approach to the backlog management system requires
appropriate work classification and priority.

4.5 Planning Process

“Poor Planning Leads to Poor Performance”


Author: Unknown, but a wise person
Work Management: Planning and Scheduling 99

Basics of Planning
Planning defines what work will be accomplished and how.
Scheduling identifies when the work will be completed and who will do
it. Planning and scheduling are dependent on one another to be effective.
However, planning is the first step. The ultimate goal of the planning
process is to identify and prepare a maintenance craft person with the
tools and resources to accomplish this work in a timely and efficient man-
ner. In other words, planning provides maintenance craft workers with
everything they need to complete the task efficiently.
Many maintenance engineers and managers consider planning to be
nothing more than job estimating and work scheduling. This is not true.
Planning is the key enabler in reducing waste and non productive time,
thereby improving productivity of the maintenance workforce. Many
organizations have started considering planning to be an important func-
tion.
However, they realize that proper planning is not an easy task to do.
It takes time to do it right. The time needed to plan a job properly can be
considerable, but it has a high rate of return. It has been documented by
many studies including Doc Palmer—noted author of Maintenance
Planning and Scheduling Handbook—and the author’s own experience
document that proper planning can save 1–3 times the resources in job
execution. If a maintenance job is repeatable, as most are, then it is essen-
tial to plan the work properly because it will have a much higher rate of
return.
Consider a maintenance shop AB where most of the work is per-
formed on a reactive basis. The shop has no planner or scheduler on the
staff. It has:
• 20 maintenance craft personnel
• 0 planner/schedulers
• 1 supervisor
• Estimated wrench time = 30%

The estimated productive work available (or performed) for AB per


week
= 20 people X 40 hours/week X 0.30
= 240 man-hours /week

Now, consider another maintenance shop XY that has a proactive cul-


ture and has demonstrated a wrench time of 55%. This shop has the fol-
100 Chapter 4

lowing staff:

• 18 maintenance craft personnel


• 2 planners / schedulers
• 1 supervisor
• Estimated wrench time = 55%

The estimated productive work available (or performed) for XY per


week
= 18 people X 40 hours/week X 0.55
= 396 man-hours /week

The XY shop has performed 156 hours (396 – 240) of additional work
with the same number of personnel as AB shop. This equates to a 65%
increase in resources or 13 more people on the staff.
But as mentioned earlier, planning requires more than just changing
personnel from a craft function to a planner/scheduler function. They
must have the skills and experience to understand the different types of
work and the various details that will need to be organized and assembled
for that specific task (skills and resources, steps and procedures, parts and
tools).

Understanding Work
The work to be performed needs to be clearly understood. If the scope
of the work has not been defined clearly, the maintenance planner must
talk to the requester, visit the job site, and identify what steps, procedures,
specifications, and tools are required to perform the job correctly. If the
job is too large or complicated, it may have to be broken down into small-
er sub-tasks for ease of estimating and planning.

Resource Required and Skill Levels


The skill level of the person required to perform the work must be
identified with the estimated hours. The job may include one highly-
skilled craft person and one or more low-to-mid-level skilled maintenance
technicians. Many times, maintenance professionals believe that it is dif-
ficult to estimate the time required to perform a specific job, especially if
the skills of the maintenance staff range from very low to very high with
everyone theoretically at the same pay grade and position.
Therefore, planners must have good knowledge of workforce capabil-
ities and the environment. The skill of the maintenance workforce and
Work Management: Planning and Scheduling 101

basic understanding and knowledge of their trade and plant assets will
determine the level of detailed steps and work instructions required in the
planning process. Highly-skilled workforces may not need detailed
instructions. Job estimating can become easier and potentially more accu-
rate when the jobs are broken down into smaller elements. Long and com-
plex jobs can be difficult to estimate as a whole.
A job standards database such as Means Standards or other standard
benchmarks can be used to estimate jobs. It is a good practice to build a
labor standards library for specific jobs, e.g., removing/installing motors,
5-50HP, 100-500 HP, replacing brake shoes on an overhead crane or fork-
lift, or aligning a pump–motor unit. Predetermined motion times, time
studies, and slotting techniques can be used to develop good estimates if
tasks are repetitive in nature. An estimate should include work content,
travel time, and personal and fatigue allowances.
The following are essential to good estimating practices:

• Familiarity with jobs and plant assets


• Comparing jobs against benchmarks
• Be cautious when using historical data as they may have
built-in delays
• Don’t try to be 100% accurate

It is a good practice for the planner to be a former senior craftsperson


or a craft supervisor who has been given training in job estimating.

Steps and Procedures


Steps and procedures must be developed with specifications identified
to ensure high work quality. The work instructions to disassemble or
assemble a complex component should be clear with sketches and draw-
ing as needed. They should include steps at which data such as bearing
clearances or temperature readings should be recorded. Human error caus-
es more failures of assets than any other type of error in an organization.

Parts and Tools


Materials, including parts and kit lists, must be identified in order to
have the parts available on-site before the job is scheduled. Special tools
need to be identified in order to insure the work is completed without
delays. For example, does the maintenance person need a torque wrench
1 — FMEA is a tool to identify failure modes; it is discussed in more details in Chapter 11.
102 Chapter 4

to tighten a bolt instead of a box end wrench? Furthermore, the torque


wrench is of no value if the torque value is not known. Inadequate infor-
mation may lead to a number of self-induced failures. The objective is to
reduce the likelihood that an error could occur by using the wrong part or
the potential for a maintenance person to stop work to locate the right
tools required for the job. A planned job template is shown in Figure 4.7.
It is a good practice to have a planning check list to ensure that all the
steps and documentation have been prepared or arranged. Figure 4.8
shows an example of a planner’s checklist.

Symptoms of Ineffective Planning


The following are some symptoms of ineffective planning:
• Maintenance people standing around waiting on parts
• High rework
• Poor work performance
• High stockout in the storeroom
• Planners being used to expedite parts
• Maintenance personnel arriving at the job site and waiting for
the asset / system to be shut down (wait is over 15min.)
• Frequent trips to storeroom by maintenance personnel
• Production downtime always more than estimated

Enhancing Planning Capabilities


Planning capabilities can be enhanced by ensuring the following:
• Employee involvement and roles. Educate all maintenance stake-
holders from plant manager to the maintenance technician in P&S
process to ensure all players understand their role.
• Planners may require additional assistance in developing effec-
tive work plans. It is recommended that a senior maintenance
technician may be assigned to the maintenance planners for a few
hours each day. This will help in developing better work plans.
Rotating other personnel such as craft supervisors and senior craft
personnel in planning support jobs is a good practice. It helps
them to understand why planning is important and how it func-
tions.
• Maintenance planners must have a library of information includ-
ing equipment manuals, drawings, specifications, and specific
equipment manufacturer manuals and other documentation.
• Planners shouldn’t perform additional duties such as a temporary
Work Management: Planning and Scheduling 103

Typical Job Template forPlanning a


Maintenance Work (Example)

Work Order Requirement: Replace Electric Motor (10 HP)

Job Time Standard: 2 hours - duration 4 Man-hours (2

craft person x 2 hour)

Multi-craft technician or Electrician

Craft Type:

Parts Required: Part# 11111 Motor, Electric; Location: 22-11-XX

Optional Parts:

(These parts are not required but could be needed if they are worn out)

Part# 2222 Coupling, Flex; Location: 11-00-YY

Part# 3311 Bolts, Coupling (9-16 x 3); Location: Free Bin, Shop

Special Tools: None

Procedure:
Step 1: Lock out / tag out (see attached procedure for details).
Step 2: Disconnect motor, mark/label wires.
Step 3: Unbolt coupling, inspect coupling and remove motor bolts.
Step 4: Remove motor using jib crane available.
Step 5: Install new motor (check motor is rotating freely).
Step 6: Bolt motor and check for soft foot. – record and correct any soft
foot findings.
Step 7: Install coupling, bolt motor (torque bolts to xx ft. lbs) and align them
using dial gauge or laser within acceptable range +/- 0.xxx (organiza-
tion standard).
Step 8: Remove lock out / tag out.
Step 9: Connect the motor and check for right rotation.
Step 10: Test run.
Step 11: Clean up and return asset to service.
Step 12: Close out the work order in CMMS detailing what was done.

Figure 4.7 A Planned Job Template


104 Chapter 4
Figure 4.8 A Planner’s Check List
Work Management: Planning and Scheduling 105

or relief supervisor or a safety or environmental representative.


The planner is not a secretary or clerk.

Additionally, planners shouldn’t expedite parts for breakdowns or


problems. Their responsibility is to insure that future work is planned
properly so it can be executed effectively. This also insures that they do
not get wrapped up in the day-to-day operations and maintenance issues.

• Planners should have technical and hands-on experience as a


maintenance technician or craftsman.
• Planned work package, should be reviewed by a craft supervisor
to validate that the work package is doable as planned before
scheduling.

4.6 Scheduling Process

Understanding Scheduling Basics


Scheduling insures that resources—personnel, material and the asset
on which the job is to performed—will be available for maintenance at a
specified time and place. Scheduling is a joint maintenance and operations
activity in which maintenance agrees to make resources available at a spe-
cific time when the asset can be also made available by the operations.
Jobs should be scheduled to have the least impact on normal operations.
Once a job has been planned, its status is moved to “Ready to
Schedule.” Now the job will go to the scheduler, who works with opera-
tions and maintenance supervision to develop a schedule that optimizes
operations needs with the availability and capacity of the maintenance
resources. Organizations use different strategies for scheduling plans. For
example, some use monthly, weekly, and daily schedules whereas others
use only weekly schedules. Many organizations also maintain a rolling
quarterly and yearly schedules. Yearly schedule are usually high-level
schedules providing visibility of major outage and turnaround plans.
Figure 4.9 shows examples of an organization’s integrated scheduling
plan structure.
Schedules are built by assigning dates as requested by the requester.
Some jobs need to be re-prioritized to attend to the most pressing prob-
lems first. Thereafter, the large majority of the available time remaining in
the schedule is filled with jobs that are selected in accordance with man-
agement’s priority, or other important criteria. Preventive maintenance
jobs should be given high priority; they need to be scheduled to meet their
106 Chapter 4

Figure 4.9 Integrated Scheduling Plan

due dates.
Once a job is on the schedule, the materials list should go to the MRO
store for parts kitting and material staging before the specified schedule
date. In many organizations, the CMMS / EAM system does this work
automatically. In addition, the job work package will be delivered or made
available to the individuals who will execute the job.
When the scheduled time for the job arrives, the maintenance person-
nel will have everything they need for the job:

1. A work permit to execute the job


2. Asset ready to be released by operation personnel
a. Ready for lock- and tag-out measures
b. The system already flushed or cleaned if necessary
3. Material / parts on hand including specified tools and material
handling equipment (or at site)
4. Right maintenance personnel with proper safety measures —
appropriate personal protection equipment (PPE)
Work Management: Planning and Scheduling 107

There should be no delays when the maintenance personnel arrive at


the job site. They should only have to complete the permits and set their
own locks of the asset before starting the job. Ideally the job should
progress without any hitch; however, there will be some issues. The plan-
ner / scheduler should be available to answer any job-related questions
and the craft supervisor needs to insure the quality of work.
Doc Palmer, a noted authority in the area of Maintenance Planning
and Scheduling, cites six basic scheduling principles:

1. Job plans providing number of persons required, lowest required


craft skill level, craft work hours per skill, and job duration
information are necessary for effective scheduling.
2. Weekly and daily schedules must be adhered to as closely as
possible. Proper priorities must be placed on new work orders to
prevent undue interruption of these schedules.
3. A scheduler develops a one-week schedule for each crew based
on craft hours available, forecast that shows highest skill avail-
able, job priorities, and information from the job plans.
Consideration is also made of multiple jobs on the same equip-
ment or system and of proactive and reactive work available
4. The one-week schedule assigns work for every available work
hour. The schedule allows for emergencies and high priority,
reactive jobs by scheduling a significant amount of work on eas-
ily interrupted tasks. Preference is given to completing higher
priority work by under-utilizing available skill levels over com-
pleting lower priority work.
5. The crew supervisor develops a daily schedule one day in
advance using current job progress, the one-week schedule, and
new high priority, reactive jobs as a guide. The crew supervisor
matches personnel skills and tasks. The crew supervisor handles
the current day’s work and problem even to rescheduling the
entire crew for emergencies.
6. Wrench time is the primary measure of work force efficiency
and of planning and scheduling effectiveness. Work that is
planned before assignment reduces unnecessary delays during
jobs and work that is scheduled reduces delays between jobs.
Schedule compliance is the measure of adherence to the one-
week schedule and its effectiveness.
108 Chapter 4

4.7 Turnarounds and Shutdowns

A major downtime that “just happens” can be disastrous for a plant. A


planned shutdown can provide maintenance organizations an opportunity
to identify and address major potential problems or failures in a timely
manner to improve plant safety and efficiency. Usually a system or a
process is shut down until the requested and specified work is completed
and then restarted, thus “turning around” the process/plant. Examples of
this type of work can be relining a large furnace, overhauling and upgrad-
ing an assembly system, replacing turbine or compressor blades, cleaning
and upgrading a chemical reactor, or replacing process tanks. In a produc-
tion facility, a turnaround usually consists of combinations of investment
projects, maintenance projects or overhauls, and typical maintenance
activities such as PMs or corrective maintenance activities that require the
plant to be removed from service.
All of the major heavy metal and process industries—steel mills,
refining, petrochemicals, power generation, pulp & paper, etc.—have
their own nomenclature for their maintenance projects. These are called
turnarounds, maintenance shutdowns, planned outages, or just mainte-
nance repair projects.
Shutdowns for scheduled major maintenance work and large capital
investments are the most expensive and time-consuming of maintenance

Capital Project Turnaround

Well defined & static-drawings Loosely defined, dynamic - changes as


Scope
available inspections made

Planning & Can be planned and scheduled Planning & scheduling can’t be finalized
Scheduling in well advance until scope is approved

Safety Requires shift and daily basis due to


Fixed, weekly or monthly basis scope fluctuations
permits

Manpower Fixed, usually don’t change Variable, changes a lot during execution
staffing much due to scope fluctuations

Schedule
Weekly or bi-monthly Shift and daily basis
update

Figure 4.10 Capital Projects Vs. Turnaround Maintenance


Work Management: Planning and Scheduling 109

projects because of the loss of production and the expense of the turn-
around itself. They can be complex, especially in terms of shared
resources; as the complexity increases, they become more costly and dif-
ficult to manage. Scheduled shutdowns usually are of a short duration and
high intensity. They can consume an equivalent cost of a yearly mainte-
nance budget in just a few weeks. They also require the greatest percent-
age of the yearly process outage days. Controlling turnaround costs and
duration represents a challenge.
A shutdown always has a negative financial impact. This negative
impact is due to both loss of production revenue and a major cash outlay
for the shutdown expenses. The positive side is not as obvious; therefore,
it is often overlooked. The positive impacts are an increase in asset relia-
bility, continued production integrity, investment in infrastructure, and a
reduction in the risk of unscheduled outages or catastrophic failure.
Scope management is one of the major challenges in a turnaround.
The scope will change, sometimes dramatically, and it will impact the
schedule. Typically, scope is developed based on information gathered
from operating parameters, capital investments, preventive maintenance
actions, and predictive tools. Sometimes, we don’t have a good under-
standing of the scope until an asset or system is opened for inspection. As
an asset is opened, cleaned, and inspected, the extent of required repairs
can be determined and planned.
There are distinct differences between turnaround maintenance work
and capital projects. Work scope is well defined in capital projects; how-
ever, in turnarounds, scope is dynamic and fluctuates a lot. Figure 4.10 list
major differences between capital projects and turnarounds.
Identifying and appointing a Turnaround Planner well in advance,
maybe six to eight months, is a good practice. This planner helps to devel-
op the scope, integrate the full scope of work including resources, and
assure readiness for execution of the turnaround. Similarly, identifying
and appointing a Turnaround Manager well in advance, maybe three to
four months, is also a good practice. The Turnaround Manager should
have the delegated authority to lead the turnaround effort to a successful
conclusion. In some organizations, new turnaround managers and plan-
ners get appointed just after completion of the last turnaround, as an ongo-
ing process to begin planning for the next turnaround. Lessons learned
from the previous turnaround are then transferred to the planning and exe-
cution of the next turnaround.
110 Chapter 4

The following is a suggested checklist for a turnaround manager:

• Identify the rough scope of the work and resources required,


specifically who will be planning, scheduling, and supervising the
work.
• Scope finalization—Work with key players to identify the scope
as soon as possible. As a minimum, freeze the scope four weeks
before the start of a turnaround, depending on the size and com-
plexity of the turnaround. There will be changes. Accommodate
them as they arise within the contingency allowances of the turn-
around. Significant additions that exceed contingency plans
require revisiting the total scope of the work and authorization of
changes by the stakeholders.
• Work planning—Plan the work and prepare job packages with the
help of planners and craftsmen who are familiar with the work /
area.
• Ensure the work plans have been reviewed by the assigned craft
supervisors from an execution point of view.
• Ensure all drawings, repair instructions, and required materials
have been identified and updated, and that their availability has
been validated. Check that arrangements have been made to stage
the material at proper location.
• Check that special tools and lifting devices (e.g., forklifts, mobile
crane of right capacity) have been arranged and will be available
at site on the scheduled day. Make sure that lift plans, equipment
capacity, and condition for service have been validated prior to
scheduled lifts.
• Work scheduling—Break large work into smaller work tasks and
then schedule them based on resource availability and duration of
shutdown. Schedule all work to be completed in 90% of the
approved duration. Leave 10% time as a contingency.
• Identify “critical path” tasks that can impact overall schedule and
focus your attention to them.
• Make sure all material, tools, cranes, etc. have been arranged to
be delivered at least one day before the start of shutdown.
• Ensure all the necessary permits have been procured and the lock-
out and tag–out plans have been arranged to provide for safe, effi-
cient access to the scheduled work.
• Establish a communication system. How is work accomplished?
Work Management: Planning and Scheduling 111

Once problems have been encountered or uncovered and correc-


tive action taken, how will this information be communicated and
how will feedback be provided in a timely manner? For large and
critical tasks, communication may be necessary on every shift.
Arrange to meet face-to-face with task leaders, planners, and
schedulers on a daily basis for schedule execution and on a week-
ly basis to review progress and change in direction, if needed. The
schedule is intended to accomplish the overall goals, while main-
taining enough flexibility to accomplish minor changes.
• Arrange a face-to-face meeting with all your key players, includ-
ing operations personnel, to discuss the goal and schedule of this
shutdown. Make sure they understand the cost of this undertaking
and impact of delays. Emphasize safety and quality of work. This
meeting should be held a few days before the start of the shut-
down. At a minimum, the Operations personnel should be includ-
ed in the weekly progress reviews.
• On the first day of the shutdown, make sure all safety measures
are taken in shutting down the system and that appropriate per-
sonal protection equipment (PPE) are used. All lock-out and tag-
out should be completed properly.

Attention to the following items may be required and appropriate


corrective action planned:

• Barricades. These should be considered to restrict the movement


into or the presence of people in restricted areas where overhead
lifting, high voltage, radiography, and hazardous materials may
be present. Ensure that proper safety signs are displayed in appro-
priate areas.
• Dust Control Management. A large shutdown can also be the
source of excessive dust, depending upon the area and work to be
accomplished. Make necessary arrangements to control the dust.
• Emergency Showers and Eye Baths. Make certain that emer-
gency showers and eye baths are available at the right locations.
• Liquid and Solid Waste Handling. Certain cleaning operations
may create liquid and solid waste which can be handled within the
in-plant industrial sewer system. Other wastes—including
asbestos, spent chemicals, sandblast media—may create materials
that require special handling, disposal, and access limitations.
112 Chapter 4

Unanticipated disturbance or creations of hazardous materials are


show-stoppers often overlooked during the planning process.
Develop and review policies for spill control, and containment
and disposal of hazardous material including potential handling
problems
• Noise Control. Some repair steps may generate excessive levels
of noise. These operations need to be identified and corrective
action taken that may require use of ear plugs and posting of
“High Noise” area.
• Scaffolding Control. During a shutdown, scaffolding is often
moved from one location to another. Ensure all scaffolding from
different sources is properly marked and color coded if necessary.
Portable, motorized lift devices add significant flexibility to any
scaffold plan for large turnaround.
• Ensure that “return-to-service” is well planned. What are the
critical items that needed to be insured before the system can be
released for start-up? Involve Operations personnel in developing
and executing the return-to-service plan, integrating their stan-
dard operating procedures with special concerns involving new or
modified equipment. Make sure those critical items are OK. The
right sequence of operation and energizing electrical devices safe-
ly is very important.

Holding a Post-Turnaround Meeting is one of the last important tasks


for the turnaround manager. The area of turnaround planning that is most
often underestimated is the area of lessons learned. Assuming the
Manager and Planners keep good meeting notes or logs during the plan-
ning and execution phases, these notes provide excellent sources of les-
sons learned and process improvements for future turnarounds.
The time to collect this information is throughout the entire duration
of planning and executing the turnaround. The lessons learned are com-
piled and reviewed with the turnaround team and stakeholders in a post-
turnaround session. The purpose of the meeting is to discuss what worked
and what did not work in the turnaround process while the memories of
the turnaround are still fresh in everyone’s mind. Recommendations from
the team are then woven into the process for future reference and imple-
mentation.
Work Management: Planning and Scheduling 113

4.8 Measures of Performance

The planning and scheduling processes, like other processes, need to


be measured and evaluated to make improvements. A few examples of
performance measures and benchmark data include:
1. Percentage of planned work. This measure is the percentage of
all jobs that have been planned. It assumes that all parts, proce-
dures, specifications, tools, drawings, etc., have been identified
before the job is scheduled. The benchmark is 90%.
2. Percentage of schedule compliance. This measure is the per-
centage of work accomplished that is agreed upon or on the week-
ly schedule. The benchmark is 85% or better.
3. Percentage of time that kits (materials and parts) are deliv-
ered on time. This measure is calculated as the number of times
the kits (material and parts) were delivered on time divided by the
total number of kits delivered. This measure affects the planner’s
ability to plan jobs properly. Expediting parts adds unnecessary
and wasteful cost to the P&S process.
4. Percentage of time the right part (s) is delivered. As part of the
planning process, planners and schedulers should have the confi-
dence that a specific vendor will deliver the right part when
required. Otherwise this problem could create a delay in perform-
ing the work. The benchmark is 99% or higher.
5. Percentage of work generated from a formal work PM/CBM.
Most work should come from identifying the degradation of a
component or asset far enough in advance of any PM/CBM tasks
that the job can be planned and scheduled properly, thus minimiz-
ing unexpected delays and production loss.
6. Percent Rework. This measure is the percent of work orders
requiring rework. Each organization needs to define what rework
means to them. It may differ from one organization to another.
Examples of rework include revisiting an asset to fix something
within 7, 15, or 30 days of a PM or major repair work performed.
The benchmark number is less than 2%.
7. Backlog. This measure shows how much work is ready to be per-
formed. This measure is important to keep maintenance effective-
ly utilized. A 4–6 week’s backlog of work is a good benchmark.
114 Chapter 4

4.9 Summary

Planning and scheduling have the highest potential impact on timely


and effective accomplishment of maintenance work. The planning and
scheduling functions are where all maintenance activities are coordinated.
Although planning and scheduling are closely related, they are two dis-
tinct functions.

• Planning: what and how


• Scheduling: when and who

Planning is what and how to do the job. It’s an advanced preparation


of a work task so that it can be executed in an efficient and effective man-
ner some time in future. It involves detailed analysis to determine and
describe the work to be performed, task sequence, and identification of
required resources—including skills, crew size, man-hours, spare parts
and other service materials, special tools, and any lifting device or equip-
ment needed. It also includes identification of special lock-out and tag-out
or any special permit required before the start of the task.
Scheduling is when and who is going to do the job. It’s a process by
which resources are allocated to a specific job based on operational
requirements and resources availability.
Planning and scheduling eliminates or minimizes the waiting time and
delays. When maintenance personnel have to return to the store room
numerous times to locate the required parts, or to locate a specific tool, it
delays the work execution and adds additional cost to the job. Poor plan-
ning and scheduling lead to poor utilization of maintenance resources.

4.10 Self Assessment Questions

Q4.1 Draw a workflow chart to show work from a request to


completion.

Q4.2 Explain each role as shown in the workflow chart from Q 4.1.

Q4.3 What is the purpose of a job priority system?


Work Management: Planning and Scheduling 115

Q4.4 Why do we need to manage maintenance backlog? What


is a good benchmark?

Q4.5 What are the symptoms of ineffective planning?

Q4.6 Should planners help schedulers or craft supervisors during an


emergency? If yes, explain.

Q4.7 Who are key players in scheduling process? Explain their


roles.

Q4.8 What are the key differences between planning and scheduling
processes?

Q4.9 Discuss work types and the benefits of work classifications.

Q4.10 What are the key differences between capital projects and
turnarounds?

4.11 References and Suggested Reading

Kister, Timothy and Bruce Hawkins. Maintenance Planning and


Scheduling Handbook. Elsevier Science & Technology, 2006.
Levitt, Joel. Handbook of Maintenance Management. Industrial Press,
1997.
Nyman, Don. Maintenance Management training notes. Seminars,
1994–96.
Nyman, Don and Joel Levitt. Maintenance Planning, Scheduling and
Coordination. Industrial Press, 2001.
Palmer, Doc. Maintenance Planning and Scheduling Handbook, 2nd
Edition. McGraw-Hill, 2005.
Chapter 5
Materials, Parts, and Inventory
Management

Almost all quality improvement comes via simplification


of design, material, manufacturing layout,
processes, and procedures.
Tom Peters

5.1 Introduction
5.2 Key Terms and Definitions
5.3 Types of Inventory
5.4 Physical Layout and Storage Equipment
5.5 Optimizing Tools and Techniques
5.6 Measures of Performance
5.7 Summary
5.8 Self Assessment Questions
5.9 References and Suggested Reading

After reading this chapter, you will be able to understand:

• Maintenance store operations


• Types of inventory
117
118 Chapter 5

• Tools and techniques to optimize inventory


• How to ensure availability of parts and materials on time
• Effective storeroom layout and storage equipment

5.1 Introduction

Maintenance storerooms play an important role in supporting the


maintenance function. The objective is to provide the right spares, service
parts, and supplies at the right time in the right quantities. If the right part
is not available when needed, the repairs will have to be delayed. Any
delay in restoring a failed asset will increase the maintenance and opera-
tions costs. Thus, a storeroom may be considered a very important enabler
in reducing the maintenance cost.
It is not uncommon to see maintenance technicians spending a con-
siderable amount of their time (as much as 20–30% in a shift) hunting for
the right parts. To provide the best possible support for the maintenance
technicians, a reasonable amount of spare parts and materials must be
available in stock. Readily available spare parts will enable emergency
repairs on a timely basis. Availability of routine adequate supplies such as
lubricating oil, gaskets, etc., will facilitate the performance of scheduled
routine maintenance. Items that are very expensive and not routinely
stocked may be purchased when needed to reduce inventory carrying cost.
In many manufacturing and support facilities, the budget for spare
parts can be a significant percentage of the total maintenance budget. This
level may be justified because of the fact that non-availability of spare
parts could substantially increase the cost of taking care of failures. It is
inconceivable and impractical for a maintenance department to carry all
the required spares in stock. This is prohibitively expensive. Therefore,
managing inventory of spare parts, supplies, and tools is a very important
function in maintenance and reliability. Usually quantitative decision
techniques for determining when and what to buy are used. An overview
of quantitative techniques available for reducing inventory costs is pre-
sented in this chapter.
In practice, application of these techniques has produced the follow-
ing results:

• 20% reduction in the workload of maintenance planners


• 30% reduction in the number of purchase orders for
replenishment parts
Materials, Parts, and Inventory Management 119

• 40% reduction in manually-prepared direct purchase requisitions


• 30% reduction in maintenance storeroom inventories
• 20% reduction in total maintenance costs

A maintenance storeroom, also simply called a storeroom, is respon-


sible for the following functions:

• Provide the right spare parts, supplies, and tools


• Deliver the needed items to the right location at the right time.

These major responsibilities of a maintenance storeroom may be met


with good advanced planning based on best practices. However, for cer-
tain parts, these expectations could be unrealistic due to cost, unexpected-
ly high failure rates, and high lead time. Maintenance, engineering, pur-
chasing, and management must work together in developing a plan to
determine the most economical stocking levels for critical items. Some
items have small or negligible lead times and can be bought with very lit-
tle lost time; as such, these items will not likely need to be stocked.
The right time to decide what parts and material should be stocked,
and in what quantity, is before placing an asset or system in service. The
manufacturer of the assets and systems typically provides a recommend-
ed list of spare parts as well as a preventive maintenance program based
on a Failure Modes and Effects Analysis1 (FMEA). The failure modes and
frequency of failures should optimize the spares list and provide a good
estimate of what and how many are required to be stocked during a spec-
ified period.
Placing an order has costs because the process requires writing the
specification and identifying the potential sources. After the sources are
identified, bids may be solicited and a qualified vendor selected. These
functions, which cost money, are called the ordering costs. The cost of
stocking an item and holding it in inventory could be as much as 30% of
item cost per year. The last few sections cover different techniques that
can be applied to reduce inventory costs.

1 — FMEA is a tool to identify failure modes; it is discussed in more


detail in Chapter 11.
120 Chapter 5

5.2 Key Terms and Definitions

Bill of Materials (BOM)


A list of materials needed to complete a particular assembly or
fabrication job. The BOM can also be a listing of items necessary
to support the operations and maintenance of an asset or compo-
nent.

Commonly Used Parts


A combination of standard replacement parts and hardware items
that may be used on multiple assets.

CMMS / EAM (Computerized Maintenance Management


System / Enterprise Asset Management)
A software system that keeps record and tracks all maintenance
activities, e.g., maintenance work orders, PM schedules, PM
masters, material parts, work plans, and asset history. Usually it
is integrated with support systems such as inventory control, pur-
chasing, accounting, manufacturing, and controls maintenance
and warehouse activities.

Emergency Spares / Parts


Replacement parts required for critical assets and equipment that
are kept in reserve in anticipation of outages caused by man-
made or natural disasters. The demand for these parts is unpre-
dictable. Usually their cost is high and they have long lead times
to procure. Not having these parts in stock may result in extend-
ed downtime and major production loss. Sometime these spare
parts are called insurance spares.

Inventory Turnover Ratio (or Inventory Turns):


This ratio indicates how often an inventory turns over during the
course of the year. Because inventories are the least liquid form
of an asset, a high inventory turnover ratio is generally positive.
However, in the case of inventory, the inventory turnover ratio is
usually low, less than two.
Materials, Parts, and Inventory Management 121

Inventory turns = Inventory issued in a year / Average inventory

Average inventory = (Beginning inventory + Ending inventory) / 2

Just-in-Time Inventory (JIT)


A method of inventory management in which small shipments of
stock are delivered as soon as they are needed. JIT minimizes stock-
ing levels.

MRO
Maintenance, Repair, and Operations. Sometime “O” is referred to
as Overhaul.

Store
Maintenance, repair, and operations store; it stocks all the material
and spare parts required to support maintenance and operations.

Service (Self-Service) Stock


Commonly-used parts and maintenance supplies kept nearby in high
maintenance areas or outside the storeroom. Withdrawal of this
stock requires no requisition or paperwork. Sometimes referred to as
Dime Store.

Spare Parts
Replacement items found on a bill of material and/or in a CMMS;
inventory management system that may or may not be kept in inven-
tory to prevent excessive downtime in case of a breakdown.

Stock Keeping Unit (SKU)


An inventory management term for individual stock items carried in
inventory, with assigned inventory numbers.

Stratification
A technique that separates data gathered from a variety of sources so
that a pattern can be seen.
122 Chapter 5

5.3 Types of Inventory

The traditional vocabulary definition defines inventory as the quanti-


ty of goods or material on hand. All inventories are not alike. For exam-
ple, retail or consumer inventory includes TVs, clothing, cars, and gro-
ceries whereas production or operations inventory includes pumps,
motors, steering wheel assemblies, valves, steel plates, and spare parts
that are vital to plant operations.
Inventories in a production process are often divided into four cate-
gories: 1) finished goods, 2) work-in-process, 3) raw materials, and 4)
maintenance and operating items such as spare parts and operating sup-
plies, including consumables. The spare parts, consumable items, and
other materials that are required to keep assets operating in a plant are the
focus of this chapter.
The maintenance inventory meets emergency, short-term, and long-
term maintenance requirements to keep the assets operating. Inventory is
a hedge against the unknown. If we knew exactly when a part was
required, we wouldn’t need to carry it in stock. We would simply buy the
part and have it delivered exactly when needed. This view sounds good in
theory, but because we don’t know exactly when we’ll need that part, we
have to carry it. Thus, inventory is sometimes called “buffer stock against
use.”
Inventory also protects us from the uncertainties of delivery. If we
knew exactly when a supplier would deliver our order, we would never
need to have inventory to cover for erratic delivery schedules. Suppliers
have problems, too. Thus, inventory is sometimes called “buffer stock
against delivery.”
Buffer stock, also called safety stock or level, can and should be kept
to a minimum by applying practices and techniques discussed later in this
chapter.

Inventory Classifications
Inventories can be classified into three major categories based on their
usage rate:

1. Active inventory
2. Infrequently used inventory
3. Rarely used inventory
Materials, Parts, and Inventory Management 123

Active Inventory (AI)


Active Inventory includes items that are used frequently enough that
future demand can be predicted with good accuracy. If an item or part is
used at least once a month, it is considered an active inventory item.
Active items are:

• Smaller spare parts, e.g., standard bearings, oil seals


• Commodity or supply items, e.g., safety gloves, bathroom supplies
• Items that have generally high demand each month
• Predictable future demand

Infrequently Used Inventory (IUI)


These are the items which are infrequently used, usually less than 10
times per year, but the demand still can be predicted with some accuracy.

Rarely Used Inventory (RUI)


These are items that fall into the category of “Must Have.” These
parts are almost impossible to obtain or lead time to acquire them is so
long that it often seems like we can’t get them. They sit on the shelves,
and there is little we can do about it. The vast majority of store items fall
into this category. An analysis of 100 stores indicated that 50% or more
items had no usage during the past two years. Yet, most of these items
must be on hand when needed.
A typical Inventory profile of more than 100 plants is shown in
Figure 5.1. Over 80% of the items in a typical store can be classified as
infrequently and rarely used items, as shown by the first bar in the figure.

120

100

80
Percent

Active items
60 Infrequently/rarely
used items
40

20

0
# of items value $ # of transcations
Items

Figure 5.1 Materials, Parts, and Inventory Management


124 Chapter 5

In order to reduce the costs of RUI items, some organizations have


started to team up with other organizations in their area to share high
value RUI items such as large motors, valves, and transformers.
The middle and right bars in Figure 5.1 represent the percentage of
inventory value and the percentage of transactions for active items vs.
infrequently and rarely-used items.

Analysis (Inventory Stratification)


Analysis, sometimes called inventory stratification, is another tech-
nique used to classify and optimize inventory levels. In this technique,
inventory is classified based on an item’s value and usage rate. This clas-
sification system is used to distinguish between the trivial many and the
vital few. In fact, this classification system reflects the Pareto principle.
Most of the items in classification A are one-of-a-kind parts with long
delivery, high cost, and low demand. They may cost over $500/unit to as
much as $100,000 or more—for example, a large 10,000 HP electric
motor required for a critical operation. Items in this classification are usu-
ally most critical. Their demand is difficult to predict and unavailability
can cause long downtime/shutdown. These items are needed in order to
have good inventory control. It has been found that the number of inven-
tory items in this category usually range from 10 to 20% of all items, aver-
aging 15%. Their cost, however, range from 60 to 80% of total inventory
cost. They can be compared to Rarely Used Inventory (RUI) discussed
earlier.

100
90
Percentage of Dollar Value

80
70
60
50
40 A
30
20
10 B
C
0
0 10 20 30 40 50 60 70 80 90 100
Percentage of Inventory Items

Figure 5.2 Analysis


Matericals, Parts, and Inventory Management 125

Items in the B classification are standard parts that may be stored in


vendor’s warehouses and made available by a local distributor in a few
days to a few weeks. Usually these items are mid-to-high cost, possibly
$100/unit or more. Items in this category are less critical and infrequent-
ly used. Their future demand can be predicted with some effort.
It has been found that in the B category, the number of items usually
ranges from 20 to 35%, averaging 25% of all inventory items. Their cost
ranges from 15 to 25% of all costs. They can be compared to Infrequently
Used Inventory (IUI) discussed earlier.
Figure 5.2 shows a typical inventory analysis and value stratification.
Most of the items in the C classification are standard parts—consum-
able or commodity items that can delivered by the vendor on a regular
schedule or made available by local distributors in a few hours or a cou-
ple of days. Usually they cost less than $100/unit. Items in this category
are actively used; their future demand can be accurately predicted and
may not need inventory control.
It has been found that in category C, the number of items ranges from
55 to 75%, averaging 65% of all items. Their cost ranges from 5 to 15%
of all costs. They can be compared to Active Inventory (AI) discussed ear-
lier.
A typical inventory stratification example is shown in Figures 5.3 and
5.4. Figure 5.3 shows 21 items with their unit cost and their demand rate
for the current and last three years. Figure 5.4 then shows each item’s clas-
sification category and current cost as well as the percent of items in each
category and their costs. The objective of this analysis is to move items
from category A to B and from B to C in order to minimize inventory
costs.
For criteria used to classify items in this example, see Figure 5.3:

A items = Value over $1,000 and usage rate less than 6/year
B Items = Value of $100–999 and usage rate over 6/year
C items = Value below $100 and usage rate over 12/year

The criteria should be tailored to meet your needs, environment, and


type of inventory.
Figure 5.4 shows the data after analysis. Again, our objective is to
review item cost and usage (or demand) on a regular basis in order to
reduce the number of items to stock in the store without impacting the
maintenance needs.
126 Chapter 5
Annual Past 3 Total
ABC Demand years Current Current
Part # Part Description Unit Cost $ Classification Past Year Demand Stock Cost
10001 Bearing, roller xxxxxx $85.00 C 50 100 29 $2,465.00
10002 Gloves, Safety $15.00 C 120 400 60 $900.00
10003 Oil Seal, xxxxx $6.50 C 40 100 38 $247.00
10004 Slings, wire rop #abc $370.00 B 5 10 4 $1,480.00
10005 Slings, wire rop #xyz $850.00 B 3 5 4 $3,400.00
10006 Wire rope Crane, #xxxx $1,450.00 A 0 1 2 $2,900.00
10007 Bearing, Spespherical xxxx $180.00 B 8 6 6 $1,080.00
10008 “O” rings - misc sizes kit $1.90 C 200 680 80 $152.00
10009 Hydraulic Cyclinder Repair kit $48.00 C 18 40 20 $960.00
10010 Hydraulic Cyclinder #xxxx $860.00 B 6 10 4 $3,440.00
10011 Motor, elec, zzzzzHP $8,400.00 A 0 1 2 $16,800.00
10012 Motor, elec, xxxxxHP $48,000.00 A 1 2 1 $48,400.00
10013 Valve, servo #xxxxx $1,250.00 A 4 3 4 $5,000.00
10014 Utility Supply misc. $1.30 C 600 2000 340 $442.00
10015 Gearbox #xxxxxx $2,600.00 A 1 0 2 $5,200.00
10016 Pump-motor unit xxxx $180.00 B 15 40 18 $3,240.00
10017 Pump Hyd xxxx $680.00 B 4 10 5 $3,400.00
10018 Bearing, friction roller #xxxx $120.00 B 24 90 20 $2,400.00
10019 Misc. Fittings $3.20 C 200 1000 160 $512.00
10020 Card, Circuit board xxxx $110.00 B 60 100 30 $3,300.00
10021 wire, electrical roll, misc $105.00 B 14 20 18 $1,890.00
Total 847 $107,208.00

Figures 5.3: Inventory costs.


Annual Demand Past 3 Year Current Total Current
Classification Part # Part Description Unit Cost $ Past year Demand Stock Stock Sub Total COST # of ITEMS % ITEMS

A 10006 Wire rope Crane # xxxx $1,450.00 0 1 2 $2,900.00

A 10011 Motor, elec, zzzzzHP $8,400.00 0 1 2 $16,800.00

A 10012 Motor, elec, xxxxxHP $48,000.00 1 2 1 $48,000.00

A 10013 Valve, servo # xxxxx $1,250.00 4 3 4 $5,000.00

A 10015 Gearbox # xxxxxxx $2,600.00 1 0 2 $5,200.00 $77,900.00 73% 11 1%

B 10016 Pump-motor unit xxxx $180.00 15 40 18 $3,240.00

Materials, Parts, and Inventory Management 127


B 10017 Pump Hyd xxxx $680.00 4 10 5 $3,400.00

B 10018 Beaming, friction roller # xxxx $120.00 24 90 20 $2,400.00

B 10020 Card, Circuit board xxxx $110.00 60 100 30 $3,300.00

B 10021 Wire, electrical roll, misc $105.00 14 20 18 $1,890.00

B 10004 Slings, wire rope #abc $370.00 5 10 4 $1,480.00

B 10005 Slings, wire rope #xyz $850.00 3 5 4 $3,400.00

B 10007 Bearing, Spepherical xxxx $180.00 8 6 6 $1,080.00

B 10010 Hydraulic Cyclinder # xxxx $850.00 6 10 4 $3,440.00 $23,630.00 22% 109 13%

C 10019 Misc. Fittings $3.20 200 1000 160 $512.00

C 10001 Bearing, roller xxxxxx $85.00 50 100 29 $2,465.00

C 10002 Gloves, Safety $15.00 120 400 60 $900.00

C 10003 Oil Seal, xxxxx $6.50 40 100 38 $247.00

C 10008 “O” rings - misc sizes kit $1.90 200 680 80 $152.00

C 10009 Hydraulic Cyclinder Repair kit $48.00 18 40 20 $960.00

C 10014 Utility Supply misc. $1.30 600 2000 340 $442.00 $5,678.00 5% 727 86%

Totals 847 $107,208.00 $107,208.00 847

Figure 5.4 ABC results.


128 Chapter 5

Data in this table indicate the following: 11 items in Category A with


73% of the total cost; 10 parts, 109 items in Category B with 22 % of the
total cost; and 7 parts, but 727 items in Category C with only 6% of the
total cost. Obviously, efforts attributed towards A items should be greater
than that for C items. Items in Category A should be reviewed frequently
whereas Category C can be reviewed with lesser frequency.
Another type of inventory—hidden stock—covers those items that
mechanics stash under conveyors and stairwells, inside cabinets, and in
toolboxes. This is the material called “lost” each year when physical
inventory is done. It’s a real problem because the condition of those parts
is unknown when the mechanic finally uses them. If the parts are bad, a
costly second downtime period may be needed to fix the asset correctly.
Usually organizations that have high hidden stocks have a reactive cul-
ture.
You may also have items in store which can’t be classified. They are
dead stock. They may be spares for assets that were removed long ago.
Can this dead stock be sent back to the supplier or one of the customers?
Other solutions are to sell it to a surplus operator or for scrap, or just trash
it. Dead stock still takes space to store and the organization has to pay
inventory taxes too. In some states, businesses have to pay property taxes
that cover inventory. Remember too that it costs, on average, 25% per year
to store the material.

5.4 Physical Layout and Storage Equipment

The physical layout of the store is an important factor in gaining pro-


ductivity. Two issues are involved in this decision: the location of the
maintenance store itself and the location of parts and material within the
store.
The maintenance store should be located as close as possible to where
work is performed—near the assets. Most of the time, a maintenance store
becomes a hub of maintenance activities. The physical layout of the store-
room should be planned for efficient material flow. To ensure that the
store room is run efficiently and effectively, the layout should consider the
following:
• The store room should be separated from main plant operations
either by walls or with a secured cage. The secured area is
required to discourage pilferage of tools and expensive items.
Many organizations have started using access cards issued to
each employee. The card provides controlled access to the store-
Materials, Parts, and Inventory Management 129

Figure 5.5 Disorganized and Well-Maintained Storerooms

room as well as the tool crib. Nevertheless, organizations need


to make sure that any material issued gets charged to the right
asset and project.
• The parts-materials area should be sized and equipped appropri-
ately for the type of parts-material to be stored. Keep heavy
parts low, close to or on the floor.
• Parts that are slow movers should be stored in the back of the
storeroom, and fast movers in the front for easy and fast access.
Consumable items and low-value items such as bolts, nuts, fit-
tings, filters, and gaskets that may be needed for frequent main-
tenance tasks should be located near the front of the storeroom
or outside the storeroom for easy access.
• Oil supplies should be kept away from the main storage area.
Any oil supply area needs to be designed to meet all fire and
environmental safety requirements.
• Each storage location and parts storage bin should be properly
labeled.
• The storage area should be free of clutter and debris to ensure
personnel can move around to access the parts easily. There
must be sufficient lighting in the area so that store personnel and
maintenance technicians can easily see and count the parts.
• Like a retail store, the storeroom will receive returns. Sufficient
space should be available to handle returns. A smooth process
for accepting and accounting for these returns must be imple-
mented.
130 Chapter 5

Figure 5.6 Material Flow Involving and Stock Supplies

Two different storerooms are shown in Figure 5.5. On the left is a dis-
organized store room and on the right is clearly a well-maintained store-
room. It will be difficult to find an item in the one on the left. A typical
material flow in a facility with a storeroom is shown in Figure 5.6. When
designing the storeroom, ensure that material flow is smooth and reduces
travel and procurement time.

Parts—Material Storage and Retrieval System


There are many different storage and retrieval methods that can be
used to handle parts and materials in the storeroom. Use of each method
will depend on the characteristics of the part and its demand.

Storage Equipment
Parts storage equipment can generally be broken down into two main
categories: man to part and part to man. The first category, man to part,
will be most familiar to personnel and consists of storage standbys like
pallet racks, shelving, and bin storage. In this arrangement, we go to the
part to pick it. This arrangement is very common in small stores.
Materials, Parts, and Inventory Management 131

In the part to man arrangement, the part comes to us. With the advent
of system-directed storage, and particularly when integrating with produc-
tion and distribution storage, part to man systems—such as horizontal and
vertical carousels, and Automated Storage and Retrieval System
(AS/RS)—have become viable. They may offer significant improvements
in parts storage efficiency.

Man to Part Man to part storage systems are the mainstay of parts stor-
age. Initially cheaper than automated part to man storage systems, they
can provide dense part storage. Of the two types, man to part is easier to
manage manually. On the downside, this type of storage (by itself) does
not provide part check-in / check-out control and inventory tracking,
which can lead to lower inventory storage accuracy. The three major types
of man to part storage are described below:

Shelves/Bins Installed at some level in virtually all parts storage


areas, shelving and bin storage is perhaps the most common type of part
storage. It is most appropriate for smaller, slower-moving parts not
accessed on a regular basis. Available in numerous configurations and
styles, shelving and bin storage will always have a place in parts storage
methodology.

Pallet Rack The big brother to shelving and bin storage, pallet rack is
the second most common type of parts storage. It is used primarily for
parts that are too big or too heavy for shelving. Pallet rack storage has the
common advantage of having a low initial installation cost and virtually
no maintenance; it is very configurable. Negatives include a lower storage
utilization density than shelving or modular drawers. Furthermore, either
rack decking or actual pallets are required for storage on the rack beams.

Modular Drawers Modular drawer storage consists of lockable stor-


age cabinets containing multiple custom-divided drawers that closely
match the specific part/tool configuration requirements. Particularly well
suited for small part storage and tool storage, modular drawer storage can
provide very high-density, secure storage. Best utilized for very slow
moving parts or as a dedicated location for secure tool storage, the cus-
tom-configured nature of this type of storage makes it less suited for con-
stant access and random part storage. Modular storage cabinets are gener-
ally more expensive than standard shelving. Some level of systematic
132 Chapter 5

Figure 5.7 Horizontal Carousels and Vertical Carousels.

tracking, rather than simple paper records, is often required to manage


large numbers of modular storage cabinets. However, the extreme storage
density and security of modular drawer storage makes it a consideration
for use in all storage strategies.

Part to Man Part to man storage systems are usually automated storage
devices that offer several advantages over standard man to part methods.
These advantages include controlled access that provides more part pro-
tection and security, check-in / check-out processes that aid in access
supervision and tracking, and ease of access to a greater vertical dimen-
sion. This last feature often results in more effective storage density per
square foot of floor space.
One of the major disadvantages of part to man systems is high initial
cost. Automated storage systems are more difficult to reconfigure than
more traditional storage methods; they also have an on-going mainte-
nance cost associated with their use. Regardless of configurability and
upkeep concerns, the high initial cost of these systems has been most
responsible for the relatively low numbers of automated storage equip-
ment in use for stores. With the advent of integrated systems, and the
resulting ability to combine stores with production and inventory stores,
this investment cost is not specific to maintenance and can be spread
across other department budgets. This ability to spread investment costs
across departments has resulted in an increase in the use of automated
storage systems and warrants their inclusion in any discussion of planned
storage. The three major types of part to man storage are described below:
Materials, Parts, and Inventory Management 133

Horizontal Carousels Horizontal carousels consist of multiple sec-


tions of shelving (often called “bins”) mounted on a revolving track sys-
tem (Figure 5.7, left side). Control for these systems can be manually or
systematically directed. Often the least expensive of the automated sys-
tems, horizontal carousels are becoming increasingly common for both
general inventory and parts storage.

Vertical Carousels Similar to horizontal carousels, vertical carousels


consist of shelving layers (often called “pans” or “trays”) mounted on a
vertically-revolving track system (Figure 5.7, right side). Generally con-
structed with a solid metal enclosure, vertical carousels provide a very
secure environment for high-value parts storage.

Automated Storage and Retrieval Systems (AS/RS) An Automated


Storage and Retrieval System (AS/RS) is a combination of equipment and
controls that moves (handle), stores and retrieves materials as needed with
precision, accuracy and speed under a defined degree of automation
(Figure 5.8). Systems vary from relatively simple, manually-controlled
order-picking machines operating in small storage structures to extremely
large, computer-controlled storage/retrieval systems totally integrated into
a manufacturing and distribution process.
AS/RSs are categorized into three main types: single-masted, double-
masted, and man-aboard. Most are supported on a track and ceiling guid-
ed at the top by guide rails or channels to ensure accurate vertical align-
ment, although some are suspended from the ceiling. The ‘shuttles’ that
make up the system travel between fixed storage shelves to deposit or
retrieve a requested load (ranging from a single book in a library system
to a several ton pallet of goods in a warehouse system). As well as mov-
ing along the ground, the shuttles are able to telescope up to the necessary
height to reach the load, and can store or retrieve loads that are several
positions deep in the shelving.
To provide a method for accomplishing throughput to and from the
AS/RS and the supporting transportation system, stations are provided to
precisely position inbound and outbound loads for pickup and delivery by
the crane.
Automated Storage and Retrieval Systems are typically used in appli-
cations where: there is a very high volume of loads being moved into and
out of storage; storage density is important because of space constraints;
no value adding content is present in this process; or accuracy is critical
because of potential expensive damages to the load.
134 Chapter 5

Figure 5.8 Automated Storage and Retrieval System

No matter which storage, retrieval, or parts identification technology


is used, the important issue is that the parts–material usage history must
be analyzed to determine the movement. Each alternative will have a dif-
ferent payback as labor and productivity savings offset the capital invest-
ment.

5.5 Optimizing Tools and Techniques

Computerized Inventory Control System


Most Computerized Maintenance Management Systems (CMMS)
and Enterprise Asset Management (EAM) systems have a built-in inven-
tory management system. Each item is recorded in the CMMS system
when it is purchased or issued on a work order and gets charged to spec-
ified equipment.
Parts are assigned locations in the inventory management system of
the CMMS/EAM, and a physical inventory verifies quantity and location
CMMS/EAM systems can usually generate a physical inventory form
sorted by location, bin, and part number, description, etc.
A process should be set to evaluate parts usage and lead time, and
reviewed periodically to adjust minimum/maximum quantities and the
Materials, Parts, and Inventory Management 135

Economic Order Quantity (EOQ). Economic order quantity may be con-


sidered as the order quantity that will minimize the total inventory costs.
Note the total inventory costs consist of ordering costs and inventory car-
rying/holding costs. The real power of an inventory control system
embedded in a CMMS is its ability to capture and analyze usage data both
to apply EOQ and pay greater attention to a classification. These applica-
tions in turn enable the company to optimize inventory cost.

“Shelf Life” Management and PM Plan for Stored Items


Components and sub-assemblies such as blowers, motors, motor-
gearbox units, and bearings need to be appropriately lubricated for stor-
age and may also require rotation of shafts to reduce damage to the bear-
ings at specified intervals. It has been found that improper storage can
damage parts and reduce their life. Similarly, rubber and chemical mate-
rials, e.g., o-rings, oil seals, cylinder cups, and adhesives, have limited
shelf life. All of these types of materials should be identified in the CMMS
system with a PM or shelf-life management program.

Inventory Accuracy
Achieving a high level of inventory accuracy is a critical factor in the
success of storeroom operations. Accurate inventory is defined as the
actual quantity and types of parts in the right location in the storeroom
matching exactly what is shown on the inventory system in the
CMMS/EAM system. If a part, quantity, or location is not correct when
matched against the system, then that location is counted as an error.
Some limited variance can be tolerated in the case of certain supplies such
as nuts and bolts, as they can be considered consumable items.
Inventory accuracy is important for several reasons. The conse-
quences of inaccurate inventory are:

• If the part is not found in the location indicated in CMMS records,


the repair cannot be completed on time, thus delaying the asset
availability for operations.
• An out-of-stock condition can occur because parts will not be
ordered on time if the actual quantity is lower than the system
record.
• If the system record number is lower than the actual inventory
record, then the parts will be flagged for reorder by the system,
even if not required, resulting in unnecessary inventory.
136 Chapter 5

• Maintenance and operations personnel will lose confidence in the


inventory system, CMMS, and in stores management. This
situation can encourage hiding stock items in technician’s tool
boxes or floor cabinets.

Because inventory accuracy is very critical for the maintenance store,


it is important that a process is established to ensure that high inventory
accuracy of 95% or better is maintained. This means that 95% of the time,
a part or material is found in the right location and that the quantity in the
bin matches with the system inventory number.
Achieving a high level of inventory accuracy requires ensuring:

• All parts–materials received against a purchased order should be


recorded in an inventory system/CMMS.
• Additional information regarding parts — specific data such as
manufacturer’s number, serial number, lot size, cost, and shelf life
—should be recorded in the system.
• All parts–materials issued to a work order should be recorded
accurately along with the employee name, number, equipment,
and projects.
• All parts–materials not used after a repair or PM should be
returned and recorded in the system and put back into the correct
location.

Figure 5.9
Kitting Bins
Materials, Parts, and Inventory Management 137

A process should be set-up to perform cycle and location counts on a


regular basis. This count can be daily, weekly, monthly, or yearly depend-
ing on the size of the store, value, and other factors such as current accu-
racy level. The store personnel can be assigned a number of bins/locations
to be counted on a daily basis or a pre-assigned schedule to cover all items
in the store in, for example, six months or one year. In some stores, specif-
ically consumer warehouses, where a large number of items such as books
and CDs are kitted and shipped on an everyday basis, a weekend count by
special part-time employees is performed to ensure inventory is accurate
at the start of the week.

Parts Kitting Process


One of the functions of the storeroom is to provide parts, materials,
tools, and consumable supplies for the technicians to perform PM and
repair tasks. The storeroom can build PM or repair part kits in advance of
the scheduled PM tasks. The CMMS/EAM system should send a PM or
repair schedule with a materials request to the storeroom in advance to
hold parts inventory and to provide parts at the right location on the sched-
uled day.
The parts listed on the PM work order are picked from the storage
locations and are placed into one of the kit bins. These kit bin locations
are an extension of the storage locations, and the inventory control system
in CMMS will track these kit locations with a staged status. When pick-
ing is completed, the entire cart is moved to a kit holding area, and
scanned into the hold area location. The area maintenance supervisor or
scheduler, or in some cases the technician, is informed of the kit status and
its staged location. On the scheduled day to perform the PM or repair, the
technician picks up the kit and communicates with the inventory/CMMS
system for any changes. In fact, in some organizations, the kit and other
material are delivered on site, near the asset before the repair task is start-
ed. Figure 5.9 shows examples of kitting bins on mobile carts.

Total Inventory Costs


Operations and maintenance are the customers of a store. Customers
usually perceive quality service as the availability of goods, parts, materi-
als, and tools when they want them. A store must have sufficient invento-
ry to provide high-quality customer service. On the other hand, high
inventory levels require investments. Inventory costs consist of:
138 Chapter 5

• Carrying cost: the cost of holding an item in the store


• Ordering cost: the cost of replenishing the inventory
• Stockout cost: loss of sales or production when an asset cannot be
repaired and made available to produce due to a part stockout.

Economic Order Quantity (EOQ)


Economic order quantity (EOQ) analysis is one of the techniques
which could be used to optimize inventory levels by ordering the “right”
quantity at a specific time interval in order to minimize inventory cost but
still meet customer needs.
EOQ helps optimize order quantity that will minimize the total inven-
tory cost. EOQ is essentially an accounting formula that determines the
point at which the combination of order costs and inventory carrying costs
are the least. The result is the most cost effective quantity to order.
Although EOQ may not apply to every inventory situation, most
organizations will find it beneficial in at least some aspect of their opera-
tion. Anytime we have repetitive purchasing of items such as bearings, fil-
ters, and motors, EOQ should be considered. EOQ is generally recom-
mended in operations where demand is relatively steady. Still, items with
demand variability such as seasonality can still use the model by going to
shorter time periods for the EOQ calculation. We have to make sure that
usage and carrying costs are based on the same time period.
To determine the most cost-effective quantities of an item, we will
need to use the EOQ formula. The basic Economic Order Quantity (EOQ)
formula is:

where
D = Demand / Usage in units per year
S = Ordering cost per order
H = Inventory carrying cost per unit per year

The calculation itself is fairly simple. However, the task of determin-


ing the correct cost data inputs to accurately represent inventory and oper-
ations can be a bit of a project. Exaggerated order costs and carrying costs
are common mistakes made in EOQ calculations.
Materials, Parts, and Inventory Management 139

Annual Demand
The number of units of an item used per year may be explained as the
annual usage.

Ordering Cost
Also known as purchase cost, this is the sum of the fixed costs that are
incurred each time an item is ordered. These costs are not associated with
the quantity ordered, but primarily with physical activities required to
process the order. There is a big variation in this cost. We have found that
an order cost varying $20–200 per order depending upon factors such as
organization size. Usually order cost includes the cost to enter the pur-
chase order or requisition, any approval steps, the cost to process the
receipt, incoming inspection, invoice processing, and vendor payment. In
some cases, a portion of the inbound freight may also be included in order
cost. These costs are associated with the frequency of the orders and not
the quantities ordered.

Carrying Cost
Also called holding cost, carrying cost is the cost associated with hav-
ing inventory on hand. It includes the cost of space to hold and service the

Figure 5.10 EOQ and Stocking Levels


140 Chapter 5

items. Usually this cost varies between 20–30% of the item’s value on an
annual basis.
Figure 5.10 graphically portrays the concept of a typical EOQ and
stocking levels. The illustration assumes a constant demand–consump-
tion, failure rate, and constant lead time. In real practice, demands are not
always constant and often reorder cycle changes with time.
The next two examples demonstrate how EOQ and total inventory
costs may be computed.

Example #1
A plant buys lubricating oil in 55-gallon drums and its usage rate is an
average of 132 drums of oil in a year. What would be an optimal order
quantity and how many orders per year will be required? What would be
the additional total cost of ordering and holding these drums in store if
ordering cost is increased by $10/order? Plant data indicates that:

a) Preparing an order and receiving the material cost $60/order


b) The oil drum carrying (holding) cost is 22% per year. The average
cost of a 55-gallon oil drum is $500.

Solution:
The economic order quantity (EOQ) is:

Where
Annual usage D = 132
Ordering cost S = $60
Annual carrying cost H = 22% of item cost = 0.22 x $500 = $110

Average number of oil drums on hand = = 6 drums

Numbers of orders / year = 11/year


Materials, Parts, and Inventory Management 141

Total annual cost (TC) of ordering and holding oil drums in inventory

Now, if the cost of ordering is increased by $10 to $70 per order, the
new EOQ and total cost (TC) can be calculated as follows:

Total annual cost (TC) of ordering and holding oil drums,

An increase of order cost by $10 has resulted in an increase of total


cost by $105, from $1320/year to $1425/year.

Example #2
A plant maintenance department consumes an average of 10 pairs of
safety gloves per day. The plant operates 300 days per year. The storage
and handling cost is $3 per pair and it costs $25 to process an order.

a) What would be an optimal order quantity as well as the total cost


of ordering and carrying this item?
b) If carrying cost increases by $0.50 per pair, what would be the new
EOQ and total cost of carrying this item in store?

Solution:

a) Given,
D = Annual demand = 10 x 300 = 3000 pairs
S = Ordering cost = $25/order
H = Carrying cost = $3/pair
142 Chapter 5

Total annual cost of ordering and holding safety gloves (TC)

If we change order quantity to 250/order, the new TC

If we change order quantity to 200/order, the new TC

An order quantity of 200 or 250 would give us the same total cost.
Therefore, we could go with an EOQ of 200 gloves/order.

b) If the carrying cost is increased to $3.50/pair


D = Annual demand = 10 x 300 = 3000
S = Ordering cost = $25
H = Carrying cost = $3.50/pair

Total annual cost of ordering and holding safety gloves (TC)


Materials, Parts, and Inventory Management 143

If we change order quantity to 200/order, the new TC

With increased carrying cost from $3.00 to $3.50 per glove, the new
EOQ is still 200, but the total cost of ordering and carrying safety gloves
would increase to $725 per year.

New Technologies
New technology such as bar codes, Radio Frequency Identification
Device (RFID), and handheld data collectors similar to those used in
supermarkets or FedEx/ inventory systems could effectively help improve
productivity of storeroom operations. The introduction of bar coding, auto
ID (identification) systems, and now RFID technology into storerooms
has resulted in a significant contribution to storeroom productivity, inven-
tory accuracy, and error elimination. Use of this new technology in store-
rooms is a best practice.

Automated ID Technology
No discussion of parts storage would be complete without some dis-
cussion of automated ID technology. Although commonly used in distri-
bution operations for years, the use of automated ID in —tied mainly to
CMMS use—is only now beginning to increase. Whereas a stand-alone
maintenance system (particularly a smaller one) may function well with
manual entry and tracking of parts, integration with manufacturing and
distribution parts storage systems will almost certainly warrant the invest-
ment in and use of some form of automated ID technology. The discus-
sion below covers the two most common automated ID technologies: bar
coding and radio frequency identification (RFID).

Bar Code A barcode is an array of parallel bars and spaces arranged


according to a particular symbology that allows automated scanning
devices to read them. In use for over three decades, bar codes are now
familiar and commonplace in distribution and retail operations. However,
144 Chapter 5

only in recent years have maintenance organizations begun aggressively


implementing barcode systems in stores. When combined with a system-
atic storage process, the use of barcodes can virtually eliminate misiden-
tified parts and the selection of incorrect parts, and greatly increase the
efficiency of reusable parts and equipment tracking. Automated bar code
tracking is a baseline enabler of systematic ID tracking and is a prerequi-
site for any effective storage strategy. Organizations not currently using
barcodes for stores should investigate their use in the near future.
Overall benefits of a bar code solution include:

• More accurate and timely information


• Faster service
• Easier work for employees
• Lower labor costs
• Higher productivity than manual counting and recording
• More accurate inventory information

Many maintenance practices have been discussed in this chapter.


Most of them are good or best practices depending upon where you are in
your journey to maintenance excellence. All of them, when implemented
and tailored to suit your environment, can provide better inventory con-
trol and service, and more efficient maintenance and purchasing activities.
For example, when an organization redesigned its maintenance operations
to capture all labor and material costs used on specific assets, it assigned
a bar-coded metal tag to each piece of asset. This eliminated all labor and
material charges against the wrong job and resulted in hundreds of hours
saved because employees no longer had to fill out paperwork and key
them into the system. This change also provided for 100 percent data cap-
ture. The organization used the information to develop a cost-effective
preventive maintenance program.

Radio Frequency Identification (RFID) RFID is an automated iden-


tification and data collection technology that uses radio frequency waves
to transfer data between a reader (interrogator) and items that have affixed
tags (transponders). Unlike bar codes, which are familiar and well known
to most people, RFID is still in limited use in industry, and even less so in
maintenance operations. RFID is similar to bar coding in many aspects, as
both use tags and labels affixed to the part for identification, and both use
special readers to read the tag and/or label data. The major difference in
Materials, Parts, and Inventory Management 145

the two is that RFID uses radio waves to read the tag data, whereas bar-
code readers use light waves (laser scanners).
Although still in early adoption, RFID offers several distinct advan-
tages over traditional bar coding, including:
• No line of sight required
• Dynamic tag read/write capability
• Simultaneous reading and identification of multiple tags
• Tolerance of harsh environments

Many organizations have piloted RFID programs, and an increasing


number are actively using RFID technology in stores. Some areas of doc-
umented savings include:
• Reduced inventory control and provisioning costs
• Accurate configuration control and repair history
• Part installation and removal time tracking
• Accurate and efficient parts tracking
• Reduced parts receiving costs
• Elimination of data entry errors
• Improved parts traceability
• Reduced risk of unapproved parts

All maintenance organizations, even those currently without any sys-


tem control, should investigate the use of RFID in both their parts stores
and directly on assets to replace so-called Brass Tag identification. The
current rapid adoption of RFID in industry may soon allow it to overtake
bar codes as the new industry standard for automatic parts or assets iden-
tification.

5.6 Measures of Performance

Several performance indicators measure the efficiency of storeroom


operations. Some of these key indicators are:

1. Percentage of inactive inventory. Number of inactive items


(SKU) / total items issued.

2. Percentage of classification. The inventory items are divided


into three categories (A, B, and C), according to a criterion
established by the organization such as revenue generation, or
146 Chapter 5

value. Typically, A items represent 20 percent in terms of quanti-


ty and 75 to 80 percent in terms of the value. Also called usage
value analysis.

3. Inventory Variance (Inaccuracy). Difference between the


actual number, amount, or volume of an inventory item and the
balance shown in the inventory records. Such differences are
summarized in the variance report that is prepared to record and
resolve inventory control problems.

4. Service Level. Inventory level at which demand for an item can


be met from the on-hand stock. Usually, expressed as percent-
age of order satisfied.

5. Percentage inventory cost to plant value. The total cost of


inventory divided by the total plant replacement value.

6. Inventory Shrinkage Rate. Cost of material / items lost


through deterioration, obsolescence, pilferage, theft, and/or
waste divided by the total inventory cost.

7. Percentage Vendor Managed Inventory (VMI). Inventory


replenishment arrangement whereby the supplier (vendor) moni-
tors the customer’s inventory with own employees or receives
stock information from the customer. The vendor then refills the
stock automatically, without the customer initiating purchase
orders. The items managed by VMI divided by the total items in
inventory.

8. Inventory growth rate in number of items and suppliers.


This measure evaluates how many more items have been added
to the inventory list. The objective is to reduce the number of
items by standardization or managed by the vendors /suppliers
in their storage and delivered on time when we need it.
Similarly, we also track number of suppliers / vendors.

9. Percentage of stock-outs. Number of stock outs / total items


issued. Could be trended to include past years.
Materials, Parts, and Inventory Management 147

10. Inventory turnover ratio. Number of times an organization’s


investment in inventory is recouped during an accounting peri-
od. It is calculated by value of issued inventory divided by an
average inventory value in an accounting period.

It is a good practice to track these performance measures on a regu-


lar basis to evaluate the improvements that have been made or to set-up
improvement goals.

5.7 Summary

Continually changing business pressures are forcing maintenance


departments to review their operational processes and find ways to run
leaner, faster, and more efficient than ever before. Unless stores are inte-
grated with purchasing, operations, and material planning, any optimiza-
tion of maintenance strategy will be suboptimal, often resulting in reduc-
tion of overall organization efficiency. Managing inventory / material and
parts storage effectively is a key strategy that can’t be overlooked.
Efficient storage principles, such as storing pre-kitted parts and storing
parts close to the point-of-use, can greatly improve maintenance store per-
formance. The parts storage equipment that best fits with the overall stor-
age strategy of the organization can then be selected to meet maintenance
and operations needs. The sharing of storage equipment often results in
the ability to justify storage automation, which can lead to efficiency gains
not only in maintenance, but also across the organization.
Almost all scheduled maintenance requires specified materials and
parts to accomplish the task. Advance planning for maintenance process-
es and pre-kitting of the parts can greatly increase the efficiency of main-
tenance and reduce the wait time for maintenance personnel.
Parts held in inventory needed solely in the event of an unscheduled
asset failure are critical spares. Planned maintenance activities use service
parts on a regular schedule. The part usage is fundamentally different
between the two categories, and so the storage strategy of each should
vary as well. A major aspect of spares storage is balancing availability
against the cost of storage. Tracking the usage of all parts by user, task,
location, etc., allows reporting and analyzing of specific usage patterns.
Particularly with consumables, the very act of tracking usage will cause
awareness, and overall usage will decrease.
148 Chapter 5

Tools such as analysis and EOQ should be used to optimize invento-


ry. The benefits of optimizing materials and spares inventory include
reduction in inventory costs, elimination or reduction of craft wait time,
and reduction in stock returns. The decision of what spares to stock should
not be based on vendors’ recommendations, but instead on FMEA/RCM
analysis, the stocking costs, the lead time to procure, and impact on oper-
ations if spares are not in stock.
Holding all critical parts in inventory can result in very high storage
expenses. Consider partnering with local industry and vendors for sharing
some critical spares. Eliminating idle inventory is possible by negotiating
delivery controls and establishing vendor trust.
Focus on implementing the following best practices to improve store
effectiveness:

• Create a culture to emphasize that the storeroom is a service


provider and its objective is to provide the right part–material at
the right location at the right time.
• Ensure inventory accuracy.
• Perform daily/weekly cycle counting as a part of routine store
Room operations.
• Use auto ID to streamline data entry and reduce errors.
• Build PM / repair kits in advance.
• Establish shelf life and PM program for stored items.
• Ensure all parts–materials get charged to the correct asset.
• Establish KPIs to measure and track performance.

5.8 Self Assessment Questions

Q5.1 Inventories into a plant are generally classified in what


categories?

Q5.2 What is meant by ABC classification as related to inventory?

Q5.3 Discuss how the cost of inventory can be optimized.

Q5.4 How you will organize a store room? Discuss the key features
of a small store room you have been asked to design.
Materials, Parts, and Inventory Management 149

Q5.5 Why is inventory accuracy important? What will you do to


improve it?

Q5.6 What are key factors used in calculating EOQ?

Q5.7 Explain the benefits of using RFID technology to label stock


items – material?

Q5.8 Explain inventory turnover ratio. What are the benefits of


tracking this ratio?

Q5.9 Identify three key performance measures that can be used to


manage MRO store effectively.

Q5.10 What is meant by shelf life? What should be done to improve it?

5.9 References and Suggested Reading

Blanchard, Benjamin. Logistics Engineering & Management, 6th Edition.


Prentice Hall, 2003.
Brown, Michael. Managing Maintenance Storeroom. Wiley
Publishing, 2004.
Orsburn, Douglas K. Spares Management Handbook. McGraw Hill,
1991.
Conference Proceedings. Reliabilityweb.com, 2005-2010. .
Chapter 6
Measuring and Designing for
Reliability and Maintainability

Insanity is doing the same thing over and over again


and expecting different results.
-- Albert Einstein

6.1 Introduction
6.2 Key Terms and Definitions
6.3 Defining and Measuring Reliability and Other Terms
6.4 Designing and Building for Maintenance and Reliability
6.5 Summary
6.6 Self Assessment Questions
6.7 References and Suggested Reading

After reading this chapter, you will be able to:

• Understand reliability and why it is important


• Calculate reliability, availability, and maintainability
• Measure and specify reliability
• Review designs for reliability
• Explain the impact of O&M costs on an asset’s life cycle cost
151
152 Chapter 6

6.1 Introduction

Asset reliability is an important focal point for many organizations.


It’s a source of competitive advantage for many visionary companies. It is
the central theme for maintenance departments trying to improve their
bottom line. To some, reliability identifies the right work and is synony-
mous with reliability centered maintenance (RCM). Reliability is not just
RCM, however; it has a much broader meaning. Understanding the term
reliability and how it differs from maintenance is key to establishing a
successful program for improving reliability in any organization. In this
chapter, we will define key terms related to reliability and discuss impor-
tant factors that will help realize higher reliability of assets and plants.

What and Why Reliability?


Reliability is a broad term that focuses on the ability of an asset to per-
form its intended function to support manufacturing or to provide a serv-
ice. Many books written about reliability tend to focus on Reliability
Centered Maintenance (RCM). Reliability is not just RCM. RCM is a
proactive methodology utilizing reliability principles for identifying the
right work to be done to maintain an asset in a desired condition so that it
can keep performing its intended function. In fact, RCM is basically a PM
optimizing tool to define the “right” maintenance actions. In its most
effective and widely-accepted form, it consists of seven structured steps
for building a maintenance program for a specific asset. When organiza-
tions first try to improve reliability, they label this undertaking as RCM,
but RCM really differs from a reliability improvement initiative. Details
of the RCM process will be discussed in chapter 8.
Improving asset reliability is important to the success of any organi-
zation, particularly to its operation and maintenance activities. To do this,
we need to understand both reliability and maintenance, and how they’re
interrelated. Reliability is the ability of an asset to perform a required
function under a stated set of conditions for a stated period of time, called
mission time. Three key elements of asset reliability are the asset function,
the conditions under which the asset operates, and mission time. The term
reliable assets means that the equipment and plant are available as and
when needed, and they will perform their intended function over a prede-
termined period without failure. Reliability is a design attribute and
should be “designed in” when an asset is designed, built, and installed.
Measuring amd Designing for Reliability and Maintainability 153

On the other hand, maintenance is an act of maintaining, or the work


of keeping an asset in proper operating condition. It may consist of per-
forming maintenance inspection and repair to keep assets operating in a
safe manner to produce or provide designed capabilities. These actions
can be preventive maintenance (PM) and corrective maintenance (CM)
actions. So, maintenance keeps assets in an acceptable working condition,
prevents them from failing, and, if they fail, brings them back to their
operational level effectively and as quickly as possible.
Maintainability is another term we need to understand with reliabili-
ty. It is another design attribute which goes hand in hand with reliability.
It reflects the ease of maintenance. The objective of maintainability is to
insure maintenance tasks can be performed easily, safely, and effectively.
Reliability and maintainability attributes are usually designed into the
asset to minimize maintenance needs by using reliable components, sim-
pler replacement, and easier inspections.
With these definitions, the differences start to become clear.
Reliability is designed in and is a strategic task. Maintenance keeps assets
functioning and is a tactical task. Maintenance does not improve reliabil-
ity, it just sustains it. Improving reliability requires redesign or replace-
ment with better and reliable components. Improving reliability needs a
new thinking — a new paradigm. Rather than asking how to restore the
capability of a failed asset efficiently and effectively, we need to ask what
we can do proactively to guarantee that the asset does not fail within the
context of meeting the business needs of the overall operation.
A challenge in this transition is the belief that we should strive to max-
imize asset reliability. However, it has been found that insuring 100% reli-
ability — although a great goal —often results in high acquisition costs
and may require a high level of maintenance to sustain high reliability. It
may not be a cost-effective strategy and may not be affordable. We need
to define an asset’s or plant’s reliability requirements in the context of
supporting the underlying business needs. Then, we inevitably realize that
we may need a different reliability and an affordable maintenance pro-
gram.
As shown in Figure 6.1, we need to find the right level of reliability
required to give us the optimum total cost. This graph illustrates the pro-
duction or use cost, which is operations and downtime cost verses the
reliability (and maintenance) cost.
154 Chapter 6

Total cost
Optimum
level
Cost

Reliability

Production & use costs

Reliability (availability)

Figure 6.1 Reliability/ Availability Economics

Why is Reliability Important?


Asset reliability is an important attribute for several reasons, includ-
ing:

• Customer Satisfaction. Reliable assets will perform to meet the


customer’s needs on time and every time. An unreliable asset will
negatively affect the customer’s satisfaction severely. Thus, high
reliability is a mandatory requirement for customer satisfaction.
• Reputation. An organization’s reputation is very closely related
to the reliability of their services. The more reliable that plant
assets are, the more likely the organization is to have a favorable
reputation.
• O & M Costs. Poor asset performance will cost more to operate
and maintain.
• Repeat Business. Reliable assets and plant will insure that cus-
tomer’s needs are being met in a timely manner. Customer satis-
faction will bring repeat business and also have a positive impact
on future business.
• Competitive Advantage. Many leading and visionary companies
have begun achieving high reliability / availability of their plants
and assets. As a result of their greater emphasis on plant reliabil-
ity improvement programs, they gain an advantage over their
competition.
Measuring amd Designing for Reliability and Maintainability 155

Reliability vs. Quality Control


In a manufacturing process, quality control (QC) is concerned with
how the process is meeting specifications to guarantee consistent product
quality. Its objective is to see that both an asset and its components are
manufactured and assembled with high quality standards and meet the
designed specifications. Thus, QC is a snapshot of the manufacturing
process’ quality program at a specific time. Reliability is usually con-
cerned with failures after an asset has been put in operation for its whole
life. The QC of manufacturing processes for building assets makes an
essential contribution to the reliability of an asset — it can be considered
as an integral part of an overall reliability program.
The same way that a chain is only as strong as its weakest link, an
asset is only as good as the inherent reliability of the asset, and the quali-
ty of the manufacturing process used to build or assemble this asset. Even
though an asset may have a reliable design, its reliability may still be
unsatisfactory when the asset is built and installed or used in the field. The
reason for this low reliability may be that the asset or its components were
poorly built. This could be the result of a substandard manufacturing
process to build the asset. For example, cold solder joints could pass ini-
tial testing at the manufacturer, but fail in the field as the result of thermal
cycling or vibration. This type of failure does not occur due to poor
design, but as a result of an inferior manufacturing process.
Usually assets are designed with a level of reliability based on the
effective use of reliable components and their configurations. Some com-
ponents may be working in series and others in parallel arrangements to
provide the desired overall reliability. This level of reliability is called
inherent reliability. After the asset has been installed, the reliability of an
asset cannot be changed without redesigning or replacing it with better
and improved components. However, asset availability can be improved
by repairing or replacing bad components before they fail, and by imple-
menting a good reliability-based PM plan.
Evaluating and finding ways to attain high asset reliability are key
aspects of reliability engineering. There are a number of practices we can
apply to improve the reliability of assets. We will be discussing these
practices to improve reliability later in this chapter, as well as in other
chapters.
156 Chapter 6

6.2 Key Terms and Definitions

Availability (A)
The probability that an asset is capable of performing its intend-
ed function satisfactorily, when needed, in a stated environment.
Availability is a function of reliability and maintainability.

Failure
Failure is the inability of an asset / component to meet its
expected performance. It does not require the asset to be inoper-
able. The failure could also mean reduced speed, or not meeting
operational or quality requirements.

Failure Rate
The number of failures of an asset over a period of time. Failure
rate is considered constant over the useful life of an asset. It is
normally expressed as the number of failures per unit time.
Denoted by Lambda (λ), failure rate is the inverse of Mean
Time Between Failure (MTBF).

Maintainability (M)
The ease and speed with which a maintenance activity can be
carried out on an asset. Maintainability is a function of equip-
ment design and usually is measured by MTTR.

Mean Time Between Failures (MTBF)


MTBF is a basic measure of asset reliability. It is calculated by
dividing total operating time of the asset by the number of fail-
ures over some period of time. MTBF is the inverse of failure
rate (λ).

Mean Time to Repair (MTTR)


MTTR is the average time needed to restore an asset to its full
operational condition upon a failure. It is calculated by dividing
total repair time of the asset by the number of failures over some
period of time. It is a basic measure of maintainability.
Measuring amd Designing for Reliability and Maintainability 157

Reliability (R)
The probability that an asset or item will perform its intended
functions for a specific period of time under stated conditions. It
is usually expressed as a percentage and measured by the mean
time between failures (MTBF).

Reliability Centered Maintenance (RCM)


A systematic and structured process to develop an efficient and
effective maintenance plan for an asset to minimize the proba-
bility of failures. The process insures safety and mission com-
pliance.

Uptime
Uptime is the time during which an asset or system is either
fully operational or is ready to perform its intended function. It
is the opposite of downtime.

6.3 Defining and Measuring Reliability and Other Terms

There are two types of assets: repairable and non-repairable.


Assets or components that can be repaired when they fail are called
repairable, e.g., compressors, hydraulic systems, pumps, motors, and
valves. Reliability of these repairable systems is characterized by the term
MTBF (Mean Time Between Failure).
Assets or components that can’t be repaired when they fail are called
non-repairable, e.g., bulbs, rocket motors, and circuit boards. Some com-
ponents such as integrated circuit boards could be repaired, but the repair
work will cost more than the replacement cost of a new component.
Therefore, they’re considered non-repairable. Reliability of non-
repairable systems is characterized by the term MTTF (Mean Time to
Failure).

Reliability, Maintainability and Availability


Reliability (R), as defined in military standard (MIL-STD-721C), is
“the probability that an item will perform its intended function for a spe-
cific interval under stated conditions.”
As defined here, an item or asset could be an electronic or mechani-
cal hardware product, software, or a manufacturing process. The reliabil-
ity is usually measured by MTBF and calculated by dividing operating

157
158 Chapter 6

time by the number of failures. Suppose an asset was in operation for 2000
hours (or for 12 months) and during this period there were 10 failures. The
MTBF for this asset is:

MTBF = 2000 hours / 10 failures = 200 hours per failure


or
12 months / 10 failures = 1.2 months per failure

A larger MTBF generally indicates a more reliable asset or component.


Maintainability (M), is the measure of an item’s or asset’s ability to be
retained in or restored to a specified condition when maintenance is per-
formed by personnel having specified skill levels, using prescribed proce-
dures and resources at each stage of maintenance and repair.
Maintainability is usually expressed in hours by Mean Time to Repair
(MTTR), or sometimes by Mean Downtime (MDT). MTTR is the average
time to repair assets. It is pure repair time (called by some wrench time).
In contrast, MDT is the total time the asset is down, which includes repair
time plus additional waiting delays.
In simple terms, maintainability usually refers to those features of
assets, components, or total systems that contribute to the ease of mainte-
nance and repair. A lower MTTR generally indicates easier maintenance
and repair.
Figures 6.2 a, b, and c show trends of MTBF and MTTR data in hours.
The baseline should be based on at least one year of data, dependent on
your operations (could require as much as three years of data for assets

Figure 6.2a Trending of MTBF Data


Measuring amd Designing for Reliability and Maintainability 159

with minimal operating time). This type of trend line is essential for track-
ing impact of improvements. Figure 6.2a shows MTBF trend data, which
is increasing. This trend is a good one.
Figure 6.2b shows MTTR trend data, which is increasing. It is going
in the wrong direction. We need to evaluate why MTTR is increasing by
asking: Do we have the right set of skills in our work force? Do we iden-
tify and provide the right materials, tools, and work instructions? What
can we do to reverse the trend?

Figure 6.2b Trending of MTTR (I) Data

Figure 6.2c Trending of MTTR (D) Data


159
160 Chapter 6

Figure 6.2c shows MTTR trend data, which is decreasing. In this case,
the trending is in the right direction. To continue this trend, we need to ask
the questions: What caused this to happen? What changes did we make?
Trending of this type of data can help to improve the decision process.

Availability
Availability (A) is a function of reliability and maintainability of the
asset. It is measured by the degree to which an item or asset is in an oper-
able and committable state at the start of the mission when the mission is
called at an unspecified (random) time.
In simple terms, the availability may be stated as the probability that
an asset will be in operating condition when needed. Mathematically, the
availability is defined:

MTBF Uptime
Availability (A) = =
MTBF+MTTR Uptime + Downtime

The availability defined above is usually referred to as inherent avail-


ability (A ). It is the designer’s best possible option.
i
In reality, actual availability will be lower than inherent availability as
the asset will be down due to preventive and corrective maintenance
actions. Another term, Operational Availability (Ao), considers both pre-
ventive and corrective maintenance and includes all delays — administra-
tive, materials and tools, travel, information gathering, etc. — that keep
the asset unavailable. Achieved Availability (Aa) includes preventive
maintenance, but not delays for getting materials and tools, information,
etc.
Naturally, the designer or the manufacturer of the asset should be
responsible for inherent or achieved availability. The user of the asset
should be interested in operational availability. The inherent availability
will be degraded as we use the asset and it can never be improved upon
without changes to the hardware and software. Availability can be
improved by increasing reliability and maintainability. Trade-off studies
should be performed to evaluate cost effectiveness of increasing MTBF
(reliability) or decreasing MTTR (maintainability). For the sake of sim-
plicity and to reduce confusion, we will be using the term Availability in
this book to represent inherent availability.
The standard for availability is about 95%, meaning that the asset is
available for 9.5 hours out of 10. This is based on general industry expec-
Measuring amd Designing for Reliability and Maintainability 161

tations. In some cases, if assets are not very critical, the standard may be
lower. But in case of critical assets such as aero-engines or assets involved
with 24-7 operations, the standard may require 99% or higher availability.
In general, the cost to achieve availability above 95% increases expo-
nentially. Therefore, we need to perform operational analysis to justify
high availability requirements, particularly if it’s over 97 percent.

The Bathtub Curve and Reliability Distribution


The bathtub curve seen in Figure 6.3 is widely used in reliability engi-
neering, although the general concept is also applicable to people as well.
The curve describes a particular form of the hazard function which com-
prises three parts:

• The first part is a decreasing failure rate, known as early failures


or infant mortality. It’s similar to our childhood.
• The second part is a constant failure rate, known as random
failures. It’s similar to our adult life.
• The third part is an increasing failure rate, known as wear-out
failures. It’s similar to our old age.

The bathtub curve is generated by mapping the rate of early infant


mortality failures when first introduced, the rate of low random failures
with constant failure rate during its useful life, and finally the rate of wear
out failures as the asset approaches its design lifetime.

The Bathtub Curve


Hypothetical Failure Rate versus Time
Increased Failure Rate

Infant Mortallity End of Life Wear-Out


Decreasing Failure Rate Increasing Failure Rate

Normal Life (Useful Life)


Low “Constant” Failure Rate
Time

Figure 6.3 The Bathtub Curve


162 Chapter 6

Figure 6.4 Failure Patterns

In less technical terms, in the early life of an asset adhering to the


bathtub curve, the failure rate is high. However, it quickly decreases as
defective components are identified and discarded, and early sources of
potential failure, including installation errors, are eliminated. In the mid-
life of an asset, the failure rate is generally low and constant. In the late
life of the asset, the failure rate increases, as age and wear take its toll.
The airline industry and the U.S. Navy performed studies in the
1960s and 1970s to have a better understanding of asset failures. These
studies showed that these types of assets remain fairly close, even though
they do not all follow the bathtub curve failure concept exactly. All assets
followed a constant or slightly increasing failure rate for most of their life.
Some didn’t follow early mortality rate and some didn’t have a wear out
region either. Figure 6.4 shows a series of these failure patterns based on
original study data. The failure patterns are categorized into two groups
— age-related and random. Fewer than 20% of failures follow an age
degradation pattern; the remaining follow a random pattern with constant
failure rate.

Reliability Failure Distribution


The exponential distribution is one of the most common distributions
used to describe the reliability of an asset or a component in a system. It
models an asset or component with the constant failure rate, or the flat
section of the bathtub curve. Most of the assets, consumer or industrial,
follow the constant failure rate for their useful life, so exponential distri-
bution is widely used to estimate the reliability. The basic equation for
Measuring amd Designing for Reliability and Maintainability 163

estimating reliability, R(t),is


-λt
R =e
(t)
where
λ (lambda) = Failure rate = 1/MTBF
t = mission time, in cycles, hours, miles, etc.

(Note: e is base of the natural logarithm = 2.71828)

Calculating Reliability and Availability

Example 1
A hydraulic system, which supports a machining center, has operat-
ed 3600 hours in the last two years. The plant’s CMMS system indicated
that there were 12 failures during this period. What is the reliability of this
hydraulic system if it is required to operate for 20 hours or for 100 hours?

MTBF = operating time / # of failures = 3600 / 12 = 300 hours


Failure rate = 1 / MTBF = 1/ 300 = 0.003334 failures / hour

Reliability for 20 hours of operations,


(-λt)
R (t) = e

– (0.003334) (20)
R (20) = e = 93.55%

Reliability for 100 hours of operations,


– (0.003334) (100)
R (100) = e = 71.65%

For 100 hours of operation, the reliability of the hydraulic system is


71.65%. This means that there is a 71.65% probability that the hydraulic
system will operate without a failure. If we need to operate the system for
only 20 hours, however, the probability of failure-free operations will
increase to 93.55%.
Now, let us suppose that there is a need to operate this hydraulic sys-
tem for 100 hours to meet a key customer’s needs and the current reliabil-
164 Chapter 6

ity of 71.65% is not acceptable. The system needs to have 95% or better
assurance (probability) to meet the customer’s need.
To have reliability requirements of 95% for 100 hours of mission
time, we need to calculate a new failure rate, λ. We use the reliability
equation,

Required Reliability = 0.95 = R (100) = e (-λ x 100)

Solving this equation, gives us

(λ x 100) = 0.05
100 λ = 0.05

Thus,

Failure rate λ = 0.0005 or MTBF = 2000 hours

This indicates that the failure rate needs to be dropped from 0.00334
(or an MTBF of 300 hours) to a new failure rate of 0.0005 (or an MTBF
of 2000 hours). If we consider the same 3600 operating hours, then the
number of failures needs to be reduced from 12 to 1.8. A root cause fail-
ure or FMEA analysis needs to be performed on this hydraulic system to
identify unreliable components. Some components may need to be re-
designed or replaced to achieve the new MTBF of 2000 hours.

Example 2
A plant’s air compressor system operated for 1000 hours last year. The
plant’s CMMS system provided the following data on this system:

Operating time = 1000 hours


Number of failures, random = 10
Total hours of repair time = 50 hours

What’s the availability and reliability of this compressor system if we


have to operate this unit for 10, 20, or 100 hours? Figure 6.5 shows the
failure data and Figure 6.6 shows repair time data for those failures.
Figure 6.5 shows that the first failure happened at 100 hours of oper-
ation, the second at 152 hours of operation, and so forth. Figure 6.6
shows that the first failure happened at 100 hours of operation and took
2 hours to repair; the second failure happened at 152 hours of operation
Measuring amd Designing for Reliability and Maintainability 165
Figure 6.5 Compressor Failure Data

Figure 6.6
Compressor
Failure and
Repair Time
Data
166 Chapter 6

and took 6 hours to repair; and so forth. The total repair time for 10 fail-
ures is 50 hours.

Calculating MTBF and Failure Rate


Operating Time 1000 hours
MTBF = = = 100 hours
# of Failures 10 failures
This indicates that the average time between failures is 100 hours.
1 1
Failure Rate (λ - Lambda) = = = 0.01 Failures /Hour
MTBF 100

Calculating MTTR and Repair Rate

Calculating Availability
Earlier, we calculated,

Mean Time Between Failures (MTBF) = 100 hours


Mean Time To Repair (MTTR) = 5 hours

Then,

or

This means that the asset is available 95% of the time and is down for 5%
of the time for repair.
Measuring amd Designing for Reliability and Maintainability 167

Calculating Reliability
As calculated earlier for the compressor unit,

MTBF = 100 hours


1 1
Failure Rate λ (FR) = = = 0.01 Failures / Hour
MTBF 100

Reliability R (t) = e-λt

If t = time = 10 hours and λ = 0.01, then

Reliability (R10) = e–λt = e–(.01)(10) = e–(.1) = 0.90 or 90%

If t = 20 hours and λ = 0.01, then

Reliability (R20) = e –λt = e –(.01)(20) = e –(0.2) = 0.81 or 81.8%

If t = time = 100 hours and λ = 0.01, then

Reliability (R100) = e –λt = e –(.01)(100) = e –(1) = 0.3678 or 36.78%

So, for this compressor system with MTBF of 100 hours,


Reliability for 10 hours of operation = 90%
Reliability for 20 hours of operation = 82%
Reliability for 100 hours of operation = 37%

This data indicates that the reliability of the air compressor unit in this
example is 90% for 10 hours of operation. However, reliability drops to
37% if we decide to operate the unit for 100 hours. For 20 hours of oper-
ation, reliability is 82%. If this level of reliability is not acceptable, then
we need to perform root cause failure or FMEA analysis to determine
what component needs to be redesigned or changed to reduce the number
of failures, thereby increasing reliability.

Reliability Block Diagram (RBD)


The failure logic of an asset, components, or a group of assets and
components called a system can be shown as a reliability block diagram
168 Chapter 6

(RBD). This diagram shows logical connections among the system’s com-
ponents and assets. The RBD is not necessarily the same as a schematic
diagram of the system’s functional layout. The system is usually made of
several components and assets which may be in series, parallel, or combi-
nation configurations to provide us the designed (inherent) reliability. The
RBD analysis consists of reducing the system to simple series and paral-
lel component and asset blocks which can be analyzed using the mathe-
matical formulas.
Figure 6.7 shows a simple diagram, using two independent compo-
nents and assets to form a system in series.

Figure 6.7 An Example of Series System

The reliability of a system with multiple components in series is cal-


culated by multiplying individual component reliabilities,

R = R x R x R x R x…..R
sys 1 2 3 4 n,
And the reliability of system R as shown in Figure 6.7
sys12
Rsys = R1 x R2
12
or –(λ1 + λ2)t
Rsys =e
12

where λ is failure rate and t is the mission time.

Figure 6.8
An Example of a Parallel System
Measuring amd Designing for Reliability and Maintainability 169

Active Redundancy or Parallel System


The RBD for the simplest redundant system is shown in Figure 6.8.
This system is composed of two independent components and assets
with reliability of R and R
3 4

The reliability of a parallel system as shown is often written as

R = 1 – (1 – R ) (1 – R )
sys34 3 4

or
R = R3+ R4– (R3x R4)
sys34

or –(λ3)t –( λ4)t –(λ3 + λ4)t


R =e +e –e
sys34
In this arrangement, the reliability of the system, Rsys34 is equal to
the probability of component 3 or 4 surviving. It simply means that one of
the components is needed to operate the system and the other component
is in active state and available if the first one fails. Therefore, the reliabil-
ity of the whole system in parallel configuration is much higher than in
series configuration. The components in parallel improve system reliabil-
ity whereas components in series lower system reliability.
Standby redundancy is achieved when, in a redundant system, the
spare component is not in an active mode continuously, but gets switched
on only when the primary component fails. In standby mode, the resultant
reliability is a little higher in comparison to active mode. However, the
assumption is made that switching is done without failure or without any
delay. The reliability of a two component system in standby mode is:

+λte λ
–(λ)t –( )t
Rsys-standby = e

Example 3
In a two-component parallel system with a failure rate of 0.1 /hour of
each component, what would be the active and standby reliability of the
system for one hour of operation?
170 Chapter 6

In this example, λ = λ = 0.1 and t = 1 hour. Then active


3 4
reliability

–(λ3)t –(λ4)t –(λ3 + λ4)t


R =e +e –e
active
or

= R + R – (R R )
3 4 3 4
Note that
–(λ )t –(0.1)1
R =R =e 3 =e = 0.09048
3 4
and therefore

= 0.9048 + 0.9048 – 0.8187 = 0.9909

–(λ)t –(λ)t
Rstandby = e +λte or = R + λ t R
= 0.9048 + (0.1 x 1 x 0.9048) = 0.9953

m – out – of – n Reliability Models


In real application, there will be many components and assets in series
and parallel arrangements, depending upon design requirements. For
example, Figure 6.9 shows a typical system comprised of 13 components,
or individual assets, arranged in a combination of series and parallel con-
figuration. The system reliability can be determined by calculating first
the individual component and asset reliability, then the system’s subsys-
tems, and finally the system as a whole. Reliability of some of the subsys-
tems that are in parallel arrangements can be calculated using m-out-of-n
reliability formulas. This means how many m legs (components in series)
are necessary out of n legs for the system to operate properly. In Figure
6.9, subsystem B has three legs, but we need only one to operate the sys-
tem. Similarly, subsystem C has three components and assets in parallel,
and we need two to operate.
A simple approach for calculating the reliability of m-out-of-n sys-
tems is utilizing the binomial distribution and the relationship (R + Q)n =
1, where R is reliability, Q is unreliability, and n is the number of ele-
ments. Figure 6.10 shows the formula for 2-, 3-, and 4-element systems to
calculate system reliability.
Measuring amd Designing for Reliability and Maintainability 171

The reliability of the system, as shown in Figure 6.9, is

R =R xR xR xR xR
sys 1 2 B 10 C

R is an active parallel configuration which is equal to


B
R = 1 – [1– (R x R x R )] [1 – (R x R x R )] (1 – R )
B 3 4 5 6 7 8 9

Figure 6.9 An Example of Multiple Component System RBD

m - # of
Working
element Overall System Reliability
1 out of 2 R2 + 2R (1-R) = 1 - (1-R)2
2 out of 2 R2

1 out of 3 R3 + 3 R2 (1-R) + 3R(1-R)2 = 1 - (1-R)3


2 out of 3 R3 + 3 R2 (1-R)
3 out of 3 R3

1 out of 4 R4 + 4 R3 (1-R) + 6R2 (1-R)2 + 4 R (1-R)3 = 1 - (1 -R)4


2 out of 4 R4 + 4 R3 (1-R) + 6 R2 (1-R)2
3 out of 4 R4 + 4 R3 (1-R)
4 out of 4 R4
Assumption: each element/component’s reliability is the same (R)

Figure 6.10 System Reliability for m-out-of-n components


172 Chapter 6

= 1 – [1 – (0.85 x 0.95 x 0.9)] [1 – (0.85 x 0.9 x 0.95)] (1 – 0.9)


= 1 – (1 – 0.7267) (1 – 0.7267) (0.1)
= 1 – 0.00746
= 0.9925

Rc is another active parallel configuration, but it requires


2 (m) out-of-3(n) to operate. We could also use the binomial distribution, from
Figure 6.10 to calculate Rc
3 2
R = R + 3R (1 – R)
c
3 2
= (0.8) + 3 (0.8) (1 – 0.8)

= (0.512) + (0.384)

= 0.896
The whole system reliability

R = 0.9 x 0.95 x 0.9925 x 0.95 x 0.896


sys
= 0.7223

Example 4
Figure 6.11 shows a compressor drive system consisting of seven compo-
nents, motors, gear boxes, the compressor itself, electrical controls, and lube
oil system, with failure rates based on four years of data. All components are
assumed to be in series arrangement. Figure 6.12 shows the same compressor
system in RBD format.

The total system failure rate based on the last 4 years of data,

λX1=1.3742x10-4+ 0+ 0+ 6.871x10-5+ 4.1226x10-4+ 6.871x10-4+


1.3742x10-4
= 1.443 x 10 -3 = 0.001443
Measuring amd Designing for Reliability and Maintainability 173

Figures 6.11 Compressor X1 system with major components


and failure rates

Therefore,

MTBF of total system = 1/λ = 1/0.001443 = 693 hours

Based on 16 hours of operation, the reliability of each component is


calculated and shown in the reliability block diagram in Figure 6.12. The
system reliability

R =R xR xR xR xR xR xR
sys 1 2 3 4 5 6 7

Figures 6.12 Reliability Block Diagram for Compressor X1 System


174 Chapter 6

Substituting individual reliability, the system reliability

R = 0.9978 x 1 x 1 x 0.9989 x 0.9934 x 0.9891 x 0.9978 = 0.9772


sys
Thus the reliability of the compressor unit is 97.72% based on 16
hours of operation. Now, let us assume that there are two compressor sys-
tems X1 and X2 in the facility, as shown in Figure 6.13, with reliability of

X1 = 0.9772

X2 = 0.8545

Figures 6.13 Two compressors system arrangement

These reliability levels are based on 16 hours, 2-shift operation sce-


narios. Let us also assume that most of the time, say 85%, we need only
one compressor to meet our production needs. The second compressor
will work as active standby. However, for 15 percent of the time, we may
need both of the compressor units. During that time, both compressors
will be in series arrangement.
The reliability of one unit (85% of time needing only one
compressor)

=R +R – (R xR )
X1 X2 X1 X2
= 0.9772 + 0.8545 – (0.9772 x 0.8545)
= 1.8317 – 0.8350
= 0.9967 or 99.67%
Measuring amd Designing for Reliability and Maintainability 175
Figure 6.14 Example of a Reliability Block Diagram at Plant Systems
176 Chapter 6

The reliability of two units (15% of time needing both compressors)

=R xR
X1 X2
= 0.9772 x 0.8545
= 0.8350

So, when there is need for only one compressor unit, we are 98% reli-
able. However, when there is need for both compressors, we are only 85%
reliable to meet the customer’s needs. This level may be acceptable. If not,
we may need to redesign or replace some of the components in compres-
sor X2 to make it more reliable.
Similarly, a reliability block diagram for a process, a manufacturing
line, or a plant could be developed, as shown in Figure 6.14. This type of
reliability block diagram can provide the information needed to improve
the reliability of systems in the plant.

6.4 Designing and Building for Maintenance and Reliability

Asset Life Cycle Cost


Life cycle costs (LCC) are all costs expected during the life of an
asset. The term refers to all costs associated with acquisition and owner-
ship, specifically operations and maintenance, of the asset over its full
life, including disposal. Figure 6.15 shows a typical asset life cycle chart.
The total cost during the life of an asset includes:

• Acquisition Cost
• Design and Development
• Demonstration and Validation (mostly applicable to one-of-a-
kind, unique systems)
• Build and Installation (including commissioning)

• Operations and Maintenance (O&M)


• Operating Cost (including energy and supplies)
• Maintenance Cost
• PM
• CM

• Disposal
Measuring amd Designing for Reliability and Maintainability 177

Figures 6.15 Asset Life Cycle

Based on several studies reported, the distribution of estimated LCC


is as follows:

For a Typical
DoD System* Industrial

Design and Development 10 – 20 % 5 – 10 %


Production / Fabrication /
Installation 20 – 30 % 10 – 20%
Operations and Maintenance (O&M) 50 – 70 % 65 – 85%
Disposal <5% <5%

* Assurance Technologies Principles and Practices by Raheja and Allocco

A graph showing the typical cost commitment and expenditures for


the life of an asset is shown in Figure 6.16, as reported by Paul Barringer,
a leading reliability expert. It is clear from the figure that the O&M cost
is on average about 80% of the total life cycle cost of the asset. It is obvi-
ously important that we need to minimize operations and maintenance
(O&M) costs. As shown in the chart, the major portion of the O&M cost
becomes fixed during early design and development phase of the asset.
There are ample opportunities to reduce the LCC during the design, build-
ing, and installation of the asset.
178 Chapter 6

Figure 6.16 Cost Commitment and Expenditures during


an Asset Life Cycle

Assets should be designed so that they can be operated and main-


tained easily with minimum operations and maintenance needs. As dis-
cussed earlier in this chapter, reliability and maintainability are design
attributes; they should be designed in, rather than added later.
To have reliable and easy-to-maintain assets, we need to insure that
asset owners, including operators, are involved in developing the require-
ments as well as in reviewing the final design. In designing for reliability
and maintainability, attention must focus on:

• Reliability requirements and specifications


• Designing for reliability and maintainability
• Proper component selection and configuration to guarantee
required reliability and availability
• Review design for maintainability
• Logistics support — maintenance plan and documentation to
reduce MTTR
• Reducing the operations and maintenance costs

Reliability Requirements and Specifications


In order to develop a reliable asset, there must be good reliability
requirements and specifications. These specifications should address
most, if not all, of the conditions in which the asset has to operate, includ-
ing mission time, usage limitations, and operating environment. In many
Measuring amd Designing for Reliability and Maintainability 179

instances, developing these specifications will require a detailed description


of how the asset is expected to perform from a reliability perspective. Use
of a single metric, such as MTBF, as the sole reliability metric is inadequate.
Even worse is the specification that an asset will be “no worse” than the
existing or earlier model. An ambiguous reliability specification leaves a
great deal of room for error, resulting in poorly-understood design require-
ments and an unreliable asset in the field.
Of course, there may be situations in which an organization lacks the
reliability background or history to properly define specifications for asset
reliability. In these instances, an analysis of existing data from previous or
similar assets may be necessary. If enough information exists to character-
ize the reliability performance of a similar asset, it should be a relatively
simple matter to transform this historical reliability data into specifications
for the desired reliability performance of the new asset.
Indeed, the financial concerns will have to be taken into account when
formulating reliability specifications. What reliability can we afford? How
many failures can we live with? Do we need to have zero failures? Zero fail-
ures is a great goal, but can we justify the cost in achieving it? A proper bal-
ance of financial constraints and realistic asset reliability performance
expectations are necessary to develop a detailed and balanced reliability
specification.

Key Elements of Reliability Specifications

• Probability of successful performance


• Function (mission) to be performed
• Usage time (mission time)
• Operating conditions
• Environment
• Skill of operators / maintainers

An example of reliability requirements for an automotive system* con-


sists of an engine, a starter motor, and a battery.

There shall be a 90% probability (of success) that the cranking


speed is more than 85 rpm after 10 seconds of cranking (mission) at
— 20oF of (environment) for a period of 10 years or 100,000 miles
(time). The reliability shall be demonstrated at 95% confidence.

*Assurance Technologies Principles and Practices by Raheja and Allocco


180 Chapter 6

Let us take another example of a manufacturing cell / system that


needs to produce xyz product, at a rate of ## / hour or day, at a quality
level of Qx. The reliability-related requirements can be developed using
operational data and some assumptions. A suggested approach is:

• Define operating environment / duty cycle, i.e.,


• 20 hours/day and 250 days/year or 5000 hours/year
• Expected number of failures < 5/yr (This is an assumption —
what can we afford? Can we live with fewer than 5
failures/year?)

• Reliability and Maintainability requirements (based on above data


and assumptions)
• MTBF = 5000/5 = 1000 hours; FR = 1/1000 = 0.001
failures/hr
• Estimated repair time or MTTR can be calculated based on the
following assumptions
• 3 failures @ < 2 hours = 6 hours
• 1 failure @ < 10 hours = 10 hours
• 1 failure @ < 24 hours = 24 hours

Therefore, the required MTTR = 40/5 = 8 hours

• Reliability and Availability requirements:


• Reliability for 20 hours/day operation
• R = e –(0.001x20) = 98%
20
• Availability = MTBF / (MTBF+MTTR) = 1000/1008 = 99%

• Required (desired) operating costs


• 2 man-hour / hour of operation (currently is 3 man-hour / hour)
• Energy plus other utility cost
• 20% less than current (current usage is 2 MW plus other)
• Maintenance cost (preventive and corrective)
• 2% or less of Replacement Asset Value (currently 2.7%
increasing by 0.2% per year)

Based on the calculations and data above, we can specify the following
requirements for this new system we are procuring.
Measuring amd Designing for Reliability and Maintainability 181

• MTBF of 1000 hours or FR = 0.001 failures/hr


• MTTR of 8 hours

Or we can ask reliability of 98% for 20 hours of operations/day and


availability of 99%. Similarly, we can specify that total operating cost and
maintenance costs may not exceed some number or percent of system
replacement value. However, these numbers should be validated by sys-
tem designers/builder by performing an FMEA.
In addition, the requirements and specifications should include:

• Display of asset performance data — such as early warnings


• Current, Temperature, Pressure, etc.
• Other operating / asset condition data
• Diagnostic display — pinpointing problem areas
• Use of modular and standard components
• Use of redundant parts / components to increase reliability
• Minimize special tools — parts
• Operations and maintenance training material
• FMEA / RCM-based maintenance plan
• Maximum use of CBM technologies
• Basis of spares recommendations
• Life Cycle cost analysis
• O&M cost estimates

Reliability Approach in Design


It has been found that as much as 60% of failures and safety issues can
be prevented by making changes in design. Assets must be:

• Designed for fault tolerance


• Designed to fail safely
• Designed with early warning of the failure to the user
182 Chapter 6

• Designed with a built-in diagnostic system to identify fault


location
• Designed to eliminate all or critical failure modes cost
effectively, if possible

The following analyses are recommended to be performed during the


design phase — from conceptual design to final design.

• Reliability Analysis
• Lowers asset and system failures over the long term
• System reliability depends on robustness of design, as well as
quality and reliability of its components

• Maintainability Analysis
• Minimizes downtime — reduces repair time
• Reduces maintenance costs

• System Safety and Hazard Analysis


• Identifies, eliminates, or reduces safety-related risks through
out its life cycle

• Human Factors Engineering Analysis


• Prevents human-induced errors or mishaps
• Mitigates risks to humans due to interface errors

• Logistics Analysis
• Reduces field support cost resulting from poor quality, relia-
bility, maintainability, and safety
• Insures availability of all documentation, including PM plan,
spares, and training needs

The following checklist is recommended as a guide to review the


design and make sure that it adequately addresses reliability, maintainabil-
ity, and safety issues.

Design Reviews Checklist


• Are reliability, maintainability, availability, and safety analysis
performed?
• Is Failure Modes and Effects Analysis (FMEA) performed dur-
ing the design — at preliminary design reviews (PDR) and criti-
cal design reviews (CDR)?
Measuring amd Designing for Reliability and Maintainability 183

• Can fault-free analysis be used to improve the design?


• Is fault-tolerant design considered?
• Are components interchangeabilities analyzed?
• Is modular design considered?
• Are redundancies considered to achieve desired reliability?
• Has the design been critiqued for human errors?
• Are designers familiar with the human engineering guidelines?
• Is Reliability-Centered Maintenance (RCM) considered in
design?
• Is a throwaway type of design considered instead of repair (e.g.,
light bulbs)?
• Has built-in testing and diagnostics been considered?
• Are self-monitoring and self-checking desirable?
• Are components and assets easily accessible for repair?
• Are corrosion-related failures analyzed?
• Do components need corrosion protection?
• Is zero-failure design economically feasible?
• Is damage detection design needed?
• Is software reliability specified and considered in design?
• Is fault-isolation capability needed?
• Do electronic circuits have adequate clearances between them?
• Are software logic concerns independently reviewed?
• Has software coding been thoroughly reviewed?
• Is self-healing design feasible or required?
• Are redundancies considered for software?
• Are the switches for backup devices reliable? Do they need
maintenance?
• Are protective devices such as fuses, sprinklers, and relief
valves reliable?
• Does the asset need to withstand earthquakes and unusual loads?
If yes, are design changes adequate?
• Can manufacturing/fabrication or maintenance personnel intro-
duce any defects? Can they be prevented by design?
• Can the operator introduce wrong inputs — wrong switching or
overloads, etc.? If so, can the asset be designed to switch to a
fail-safe mode?
• Can a single component cause the failure of a critical function?
If yes, can it be redesigned?
184 Chapter 6

• Are there unusual environments not already considered? If haz-


ardous material is being used, how will it be contained or han-
dled safely?
• Is crack growth and damage tolerance analysis required?
• Are safety margins adequate?
• Are inspection provisions made for detecting cracks, damage,
and flaws?
• Are production tests planned and reviewed?
• How will reliability be verified and/or validated?

6.5 Summary

Improving reliability is essential to the success of any organization,


particularly to its operation and maintenance. Understanding reliability
and maintenance and how they’re interrelated are the basis for reducing
the life cycle costs of assets and plant.
Reliability focuses on the ability of an asset to perform its intended
function of supporting manufacturing a product or providing a service.
Reliability terminates with a failure — i.e., when unreliability occurs.
Unreliability results in high cost to the organization. The high cost of
unreliability motivates an engineering solution to control and reduce
costs.
Maintenance is an act of maintaining, or the work of keeping the
asset in proper operational condition. It may consist of performing main-
tenance inspection and repair to keep assets operating in a safe manner to
produce or provide designed capabilities. Thus, maintenance keeps assets
in an acceptable working condition, prevents them from failing, and, if
they fail, brings them back to their operational level effectively and as
quickly as needed.
Reliability should be designed in. It is a strategic task. In contrast,
maintenance keeps the asset functioning and is a tactical task. The relia-
bility and maintainability attributes are usually designed into the product
or asset. These attributes minimize maintenance needs by using reliable
components, simpler replacements, and easier inspections. Reliability is
measured by MTBF, which is the inverse of failure rate. Maintainability
— the ease of maintenance — is measured by MTTR.
It has been found that the Operations and Maintenance (O&M) costs
are about 80% or more of the total life cycle cost of an asset. It is impor-
tant to minimize O&M costs. The majority of the O&M costs to be
Measuring amd Designing for Reliability and Maintainability 185

incurred in the future are set during the design and development phase of
the asset. Therefore, we must adequately address reliability, maintainabil-
ity, and safety aspects of the system in order to reduce the overall life
cycle cost of the assets during the design and building of the assets.

6.6 Self Assessment Questions

Q6.1 Define reliability and maintainability.

Q6.2 What’s the difference between maintenance and


maintainability?

Q6.3 If an asset is operating at 70% reliability, what do we need to


do to get 90% reliability? Assume assets will be required to
operate for 100 hours.

Q6.4 If an asset has a failure rate of 0.001failures/hour, what would


be the reliability for 100 hours of operations?

Q6.5 What would be the availability of an asset if its failure rate is


0.0001failures/hour and average repair time is 10 hours?

Q6.6 What would be the availability of a plant system if it is up for


100 hours and down for 10 hours?

Q6.7 If an asset’s MTBF is 1000 hours and MTTR is 10 hours,


what would be its availability and reliability for 100 hours of
operations?

Q6.8 Define availability. What strategies can be used to improve it?

Q6.9 What is the impact of O&M cost on the total life cycle cost of
an asset?

Q6.10 What approaches could we apply during the design phase of


an asset to improve its reliability ?
186 Chapter 6

6.7 References and Suggested Reading

Paul Barringer. www.Barringer1.com


Blanchard, B.S., D. Verma, and E.L. Peterson. Maintainability: A Key to
Effective Serviceability and Maintenance Management. John
Wiley and Sons, 1995.
Ebeling, C.E. An Introduction to Reliability and Maintainability
Engineering. McGraw Hill, 1996.
Narayan, V. Effective Maintenance Management. Industrial Press, 2004.
O’Conner, Patrick D.T. Practical Reliability Engineering. John Wiley
and Sons, 1985.
Raheja, Dev and Michael Allocco. Assurance Technologies Principles
and Practices, 2nd edition. Wiley-Interscience, 2006
Chapter 7
Operator Driven Reliability

Reliability cannot be driven by the maintenance organization. It must be


driven by the operations … and led from the top.
-- Charles Bailey

7.1 Introduction
7.2 Key Terms and Definitions
7.3 The Role of Operations
7.4 Total Productive Maintenance (TPM)
7.5 Workplace Organization: 5S
7.6 Overall Equipment Effectiveness (OEE)
7.7 Measures of Performance
7.8 Summary
7.9 Self Assessment Questions
7.10 References and Suggested Reading

After reading this chapter, you will be able to understand:

• The role of operators in sustaining and improving reliability


• Total Productive Maintenance (TPM) and its implementation
• Overall Equipment Effectiveness (OEE)
• Workplace design
• Implementation of the 5S program to optimize productivity
187
188 Chapter 7

7.1 Introduction

In Chapter 1, we discussed the objective of the maintenance and reli-


ability organization: to insure that the assets are available to produce
quality products and to provide quality service in a cost-effective manner
when needed. The performance of an asset is based on three factors:

• Reliability (inherent) — how was it designed?


• Maintenance plan — how will it be maintained?
• Operating environment — in what environment and with what
methods will it be operated?

The reliability and maintenance plan factors are discussed in several


other chapters. The third factor, the operating environment, will be dis-
cussed in this chapter. This factor includes the skills of the operators and
the operating conditions under which the asset performs. Several studies
have indicated that more than 40% of failures are the direct result of oper-
ational errors or unsuitable operating conditions. These failures, and those
created by the asset itself due to faulty inherent design can be minimized
or eliminated if operators have a good understanding of the asset and the
manner in which it is operated affects the overall performance. Operators
must feel responsible for the proper operation of the assets under their
control. They live and breathe with them. They can sense if something is
wrong or abnormal about that asset’s operation or condition.
For example, consider the car you drive to work every day. If it does
not sound right or starts rough, if the brakes are making a noise, or if the
car takes extra distance to stop, you — as the operator — will be first to
notice the abnormal condition of the car. As the primary operator of the
car (asset), you know that something is wrong as you are driving (operat-
ing). You then take corrective actions (fix it yourself or get it repaired at
a service location).
Similarly, the operator of an asset can sense if there is something
abnormal or out of the ordinary with the asset. Often these incipient prob-
lems can be corrected cost-effectively by the operators themselves or with
timely help from the maintainers. However, if incipient problems are not
corrected in time, they may result in bigger failures costing many times
more to fix. In fact, operators should be the first line of defense in watch-
ing abnormal conditions of an asset and initiating corrective actions. But,
many organizations haven’t been able to involve them successfully in
maintenance because our work culture has been undergoing changes over
Operator Driven Reliability 189

the last several decades.


There are two primary reasons for this type of work culture:

1. Division of operations and maintenance labor


2. Historical reward system

A clear division of labor exists in the workforce today. The production


department operates the assets and the maintenance department fixes
when they break. Maintenance is about restoring assets to an optimum
operational state. For a maintenance team that has historically defined
itself as the “fix-it” guys, a paradigm shift to a culture of reliability chal-
lenges their self-preservation. They think, “If assets aren’t failing, the
value of their contribution goes unnoticed and who will value their pres-
ence?”
Likewise, operators just want to operate the assets without any regard
to their maintenance needs and proper operating conditions. They some-
times have trouble seeing the overall picture. They can help in reducing
the number of failures by getting involved, taking proactive actions,
catching failures in early stages, and working with maintenance in a time-
ly fashion to get them attended. Thus, there is a need for responsible own-
ership from both sides.
For decades, we have used a reward system that has created a mis-
aligned culture. Design teams are rewarded for achieving functional capa-
bility at the lowest cost, not really concerned about the downstream prob-
lems for operations and maintenance and the true life-cycle cost of own-
ership of the asset. Production teams are rewarded when they beat a pro-
duction number, regardless of any real demand for the product and with-
out any concern for the effect their actions have on asset health.
Maintenance teams have typically been rewarded for fixing asset fail-
ures and not improving reliability or availability. They get extra pay for
coming in at inconvenient times when the asset is broken and get “well
done!” from management when they fix it. If we are rewarded for failures,
why would we want reliability? Who would step up and volunteer for a
15–20% pay cut for reduced overtime?
People don’t pay as much attention to what their managers say com-
pared to what they actually do. If management says they want reliability
— no failures or minimum failures — but they keep paying for failures,
we will continue to get failures. This culture needs to be changed and
improved.
190 Chapter 7

Total Productive Maintenance (TPM) is a practice originated in Japan


that addresses problems between an organization’s maintenance and other
departments, primarily production/operations. Most organizations are
structured with maintenance on one side and operations on the other side.
Although all departments have the same goal — to be a productive unit in
the organization — the organizational lines frequently get in the way,
causing delays and production stoppages. TPM helps to reduce or elimi-
nate some of these issues.
An effective workplace design includes a targeted list of activities and
also promotes organization and efficiency within a workspace. These
activities are known as five S (5S).
The maintenance and production departments need to work together
in a cohesive manner to deliver a high-quality product in a waste-free,
cost-effective manner. Virtually every major management philosophy and
methodology in practice today recognizes and fosters the integral relation-
ship between the maintenance and production/operations departments.
Just-In-Time (JIT) and Lean Manufacturing methods would not be possi-
ble without high levels of asset reliability and availability, driven by
active operator involvement in the maintenance process.
In this chapter, we will discuss the role of operations in sustaining and
improving the reliability and availability of the assets and systems while
lowering overall cost at the same time — in other words, an operator-
based reliability.

7.2 Key Terms and Definitions

Availability (OEE related)


Availability is defined as the percentage of actual time that an
asset has operated (uptime) compared to how long it was
scheduled to operate.

Five S (5S)
5S is a structured program to achieve organization-wide clean-
liness and standardization in the workplace. A well-organized
workplace results in a safer, more efficient, and productive
operation. It consists of five elements: Sort, Set in Order,
Shine, Standardize, and Sustain.
Operator Driven Reliability 191

Operator Driven Reliability (ODR)


Equipment operator helps improve reliability by identifying
potential equipment problems and failures early. The operator
then fixes them if minor or if major, gets them repaired with
the help of maintenance in a planned manner. ODR is also
called Operator-Based Reliability (OBR) or Operator-Based
Maintenance (OBM)

Overall Equipment Effectiveness (OEE)


OEE is a measure of equipment or process effectiveness based
on actual availability, performance, and quality of product or
output. It is calculated by multiplying these three factors and
then expressed as a percentage. The objective of OEE is to
identify sources of waste and inefficiencies that reduce avail-
ability, performance, and quality (defects). Corrective actions
can then be taken to improve the process.

Total Effective Equipment Performance (TEEP)


TEEP is a measure of overall asset or process effectiveness. It
is based on four factors: utilization, availability, performance,
and quality. TEEP is calculated by multiplying these four fac-
tors and is then expressed as a percentage. It can also be calcu-
lated by multiplying the utilization rate with OEE.

The objective of TEEP is to measure how well an organization


creates lasting value from its assets.

Total Productive Maintenance (TPM)


A maintenance strategy that originated in Japan; it emphasizes
operations and maintenance cooperation. Its goal includes zero
defects, zero accidents, zero breakdowns, and an effective
workplace design to reduce overall operations and mainte-
nance costs.

Utilization Rate
The percentage of time an asset is scheduled to operate divided
by the total available time (which could be 24 hours a day, 365
days a year, etc,).
192 Chapter 7

Visual Workplace
A visual workplace uses visual displays to relay information to
employees and guide their actions. The workplace is setup with
signs, labels, color-coded markings, etc., so that anyone unfamiliar
with the assets or process can readily identify what is going on,
understand the process, and know both what is being done correct-
ly and what is out of place.

Figure 7.1 lists Japanese words and definitions that have a special indus-
trial meaning. Some of these words are becoming part of our routine work-
place terminology.

Figure 7.1 M&R Related Japanese Words & Definition


Operator Driven Reliability 193

7.3 The Role of Operations

The concept of Operator Driven Reliability (ODR) — sometimes


called Operator-Based Reliability or OBR — is an integral part of an
overall proactive maintenance strategy.
The objective of ODR is to help keep plants running better, longer,
cost-effectively, and competitively by reducing unplanned downtime and
increasing uptime of production processes and associated assets.
By proactively identifying problems, operators can eliminate or
reduce failures, thereby increasing reliability. Owning and operating
assets constitutes one of the biggest factors in the total cost in a plant.
Reducing that cost by increasing asset uptime can generate additional
profits without any additional expenditures.
Under the ODR concept, operators perform basic maintenance activ-
ities beyond their classic operator duties. They take responsibility to
observe and record the asset’s overall health by checking for leaks and
noises, monitoring temperature, vibration, and any abnormal asset/system
conditions. In some cases, operators correct the minor deficiencies they
find. They perform tasks such as cleaning, minor adjustments, lubrication,
and simple preventive and corrective tasks traditionally handled by the
maintenance technician. These tasks represent a departure from their tra-
ditional role as just an equipment operator. ODR encourages production
to interact with maintenance and other departments as a team to reduce the
number of failures, thus improving plant-wide asset reliability.
In most cases, the original equipment manufacturers (OEM) recom-
mend how equipment should be operated. OEM recommendations are
sometimes made without any understanding or appreciation for the
process or environment in which the asset is actually operated. The OEM
also has very little if any knowledge of the operator’s skill sets. Usually
they also require operators and maintainers to do a lot more than what
really is needed to preserve the asset’s functions. This may result in too
many PM tasks and unnecessary inspections as well as time wasted col-
lecting irrelevant information.
With operators taking responsibility for identifying problems, the
probability of detecting early asset failures rises exponentially. This
improvement can contribute to increased asset reliability at much lower
cost due to earlier fault detection.
In many organizations, maintenance and operations departments
function virtually independently of each other, effectively divorced with
separate agendas. Such situations do not turn out well for organizations
194 Chapter 7

striving to improve productivity and profitability. ODR can serve as a


bridge to those achievements by fostering and promoting internal dia-
logue and offering a cost-effective way to improve asset reliability. ODR
can encourage a culture that will not tolerate failures; it can maximize
cross-functional teamwork and identify many previously hidden opportu-
nities for continuous improvements.
The ODR concept is similar to another maintenance strategy known
as Total Productive Maintenance (TPM) which was developed in Japan in
the early 1960s. TPM will be discussed in detail in the next section.

7.4 Total Productive Maintenance (TPM)

Total Productive Maintenance (TPM) is a team-based asset manage-


ment strategy that emphasizes cooperation between operations and main-
tenance departments with a goal of zero defects, zero breakdowns, zero
accidents, and an effective workplace design.
TPM seeks to engage all levels of an organization with their different
functions in an effort to maximize the overall effectiveness of production
assets. This helps to bring improvements in existing processes and asset
availability by reducing mistakes and accidents. Traditionally, the mainte-
nance department manages the plant’s maintenance programs, but TPM
seeks to involve employees in all departments — including production
and maintenance at all levels from the plant floor to senior executives —
to insure effective asset operation.
TPM is based around the following principles:

• Improving asset and equipment effectiveness


• Autonomous maintenance by operators
• Servicing, adjustments, and minor repairs
• Planned maintenance by maintenance department
• Training to improve operation and maintenance skills
• Better workplace design including standardization of
procedures and cleanliness

TPM is an innovative Japanese concept. The origin of TPM can be


traced back to the 1960s when preventive maintenance was introduced in
Japan. Nippondenso, a Toyota part supplier, was the first company to
introduce a plant-wide preventive maintenance program in 1960.
However later with the increasing automation of assets and systems at
Nippondenso, the maintenance program needed to perform additional
Operator Driven Reliability 195

work; more maintenance personnel were required. Management decided


that the routine maintenance of equipment would be carried out by the
equipment operators themselves. The maintenance group took the respon-
sibility for major repairs and essential maintenance tasks. This practice
was later called autonomous maintenance, which became one of the pil-
lars of TPM.
In this model, after establishing a preventive maintenance program,
Nippondenso also added the autonomous maintenance work to be done by
the equipment operators. The maintenance department concentrated on
equipment PMs and process modifications for improving reliability.
These improved modifications were also incorporated into new equip-
ment designs leading to the reduction or prevention of maintenance work.
Thus preventive maintenance, along with maintenance prevention and
design improvement, gave birth to Productive Maintenance. The aim of
productive maintenance was to maximize plant and equipment effective-
ness to achieve the optimum life cycle cost for production equipment.
Nippondenso was also involved in forming quality circles, which
facilitated the employee’s participation in improving quality of their prod-
ucts. All Nippondenso employees got involved in implementing
Productive Maintenance. Because all (or the total) plant employees partic-
ipated in implementing productive maintenance, Seiichi Nakajima of the
Japanese Institute of Plant Engineers (JIPE), who led this effort, named
the concept Total Productive Maintenance (TPM). Based on these devel-
opments, Nippondenso was awarded the distinguished plant prize for
developing and implementing TPM. Thus Nippondenso of the Toyota
group became the first company to obtain the TPM certification.
The TPM program closely resembles the popular Total Quality
Management (TQM) program. Many of the tools used in TQM, such as
employee empowerment, benchmarking, and documentation, are also
used to implement and optimize TPM. The following shows the similari-
ties between the two programs.

a) Total commitment to the program by upper level management.


b) Employees must be empowered to initiate corrective action
c) A long-term strategy is required as it may take a long time to
implement programs and make them a part of the routine, ongoing
process
d) Cultural change — new mindsets are required

Right from the start, TPM requires effective leadership and involve-
196 Chapter 7

ment of all employees from a craft person to senior managers. That is


part of the meaning of “total” in Total Productive Maintenance. TPM
holds people accountable for performing highly specified (or special-
ized) work and improved equipment performance. Without management
support, TPM will become a program of the month, and will die. Many
of today’s business leaders have risen through the ranks when mainte-
nance was responsible only for fixing equipment and not for preventing
failures. Viewing maintenance as a non-value-added support function
often leads to severe cost-cutting measures. In turn, this step results in
higher costs due to decreased equipment effectiveness.

TPM Objectives and Benefits


The objectives of TPM are to:

1. Achieve zero defects, zero breakdowns, and zero accidents in all


functional areas of the organization.
2. Involve people at all levels of the organization.

These goals are accomplished by involving all employees in small


group activities that identify both the causes of failures and opportunities
for plant and equipment modifications. They are also accomplished by
adopting the life cycle approach for improving the overall performance of
production equipment.

Benefits of TPM
1. Increased productivity
2. Reduced manufacturing cost
3. Reduction in customer complaints.
4. Satisfy the customer’s needs by 100%
• Delivering the right quantity
• At the right time
• With best, required quality
5. Reduced safety incidents and environmental concerns.

In addition, TPM creates a positive work culture and environment to:

• Build a higher level of confidence among the employees


• Keep the workplace clean, neat, and attractive
• Favorable / positive attitude of the operators and maintainers
• Deployment of a new concept in all areas of the organization
• Share knowledge and experience
Operator Driven Reliability 197

The employees are empowered and gain a real sense of owning the
assets they operate.

TPM Pillars
TPM consists of eight pillars of activities that impact all areas of the
organization. These pillars are:

Pillar #1 Autonomous Maintenance (Jishu Hozen)


This pillar is geared towards developing operators to be able to take
care of small maintenance tasks. This ability in turn frees up the skilled
maintenance people to spend time on higher value-added activities and
technical repairs, instead of fire-fighting, breakdown maintenance. Under
this concept, operators are responsible for upkeep of their equipment to
prevent them from deterioration-induced breakdowns.
Implementing autonomous maintenance requires not only a change in
organization culture, but also a heavy investment in training. Operators
who have always said “That’s not my job — call maintenance,” must now
acquire a sense of ownership. They must also acquire the skills to proper-
ly implement their new accountability. Operators will now keep the equip-
ment clean, lubricated, and secure. Minor repairs and adjustments also
become part of the operator’s responsibility. Operators need to be trained
to inspect, measure, and continuously diagnose and fix problems.
In carrying out these responsibilities, operators need to learn more
about their equipment and become better equipped to detect problems
early. Therefore, they take corrective action at the right time and become
key members of the team to improve equipment effectiveness.
Management must promote a working environment that fosters posi-
tive change. This change can be accomplished through their increased
involvement in the program, by providing physical surroundings con-
ducive to work. Positive change also requires monetary support for the
TPM program and for implementation of the ideas it generates.
The goals of autonomous maintenance are:

• Uninterrupted operation of equipment


• Operators to operate and maintain the equipment
• Elimination of the defects and potential failures at source quickly
• Involvement of all employees to solve problems through active
participation
198 Chapter 7

5S, also known as workplace organization, is an important practice.


To some authors, it is a part of the autonomous maintenance pillar. To oth-
ers it is a separate pillar by itself. But 5S is a key element of operator-driv-
en reliability and a foundation of TPM. We will discuss 5S in more detail
later in this chapter.
Overall equipment effectiveness (OEE) is a key metric in determining
how well equipment is performing with regards to losses. OEE measures
equipment effectiveness in terms of availability, performance, and prod-
uct quality. Details of OEE will be discussed in Section 7.6.

Pillar #2 Focused Improvement — Kaizen (Kobetsu)


This pillar is aimed at reducing losses in the workplace to improve
operational efficiencies. Kai’ means change, and Zen means good. Kaizen
is a collection of small improvements that produces amazing results when
carried out on a continual basis; it involves all employees in a group or
team. The goals of Kaizen improvement are:

• Zero losses — identify and eliminate losses


• Remove unsafe conditions
• Improve effectiveness of all equipment
• Reduce O&M costs

The following are six major losses that can become a focus of kaizen
teams to improve effectiveness:

1. Breakdown losses
2. Setup and adjustment losses
3. Idling and minor stoppage losses
4. Speed losses
5. Defective product losses (quality) and rework
6. Equipment design losses

Breakdown losses Breakdown losses are equipment-related losses


resulting from failures, breakdowns, and repairs. Costs can include down-
time and lost production opportunity, labor, and material costs.
Setup and adjustment losses These losses result from equipment
setups and adjustments that occur during product changeovers, shift
changes, or other changes in operating conditions.
Minor stoppage losses These losses are the result of short but fre-
quent production stoppages from zero to few minutes in length (less than
Operator Driven Reliability 199

5–6 minutes). They are usually difficult to record. As a result, these loss-
es are usually hidden from production reports. These are built into
machine capabilities, but provide substantial opportunities for improving
production efficiencies.
Speed losses Sometime equipment must be slowed down to prevent
quality defects or minor stoppages, resulting in production losses. In most
cases, this loss is not recorded because the equipment continues to oper-
ate.
Quality defect losses These losses result from out-of-spec produc-
tion and defects due to equipment malfunction or poor performance, lead-
ing to output which must be reworked or scrapped as waste.
Equipment design losses These losses are typical of heavy wear and
tear on equipment due to “non-robust” design, which reduces their
durable and productive life span. Such designs lead to more frequent
equipment modifications and capital improvements.
By using a detailed and thorough analysis, equipment design losses
are reduced or eliminated in a systematic manner using tools such as
Pareto, 5-Why Analysis, and Failure Modes and Effects Analysis
(FMEA). The use of such tools are not limited to production areas, but can
be employed in administrative and service areas as well to eliminate loss-
es or waste. These and other improvement tools are discussed in more
detail in Chapter 11.

Pillar #3 Planned Maintenance


Planned maintenance and improvement are carried out by the mainte-
nance department. These planned maintenance tasks are usually beyond
the scope of the autonomous maintenance program. They require special
skills, significant disassembly, special measuring techniques and tools,
etc. As equipment operators improve their skills, the maintenance group
performs fewer and fewer planned maintenance activities and starts focus-
ing their efforts instead on improvements that are designed to reduce the
maintenance requirements of the equipment, thus reducing overall main-
tenance work.

Pillar #4 Quality Maintenance


This pillar focuses on eliminating product quality related to non-con-
formances in a systematic manner. As operators gain understanding of
how various equipment components affect the product quality, they begin
to eliminate the current quality issues and then prepare to tackle potential
quality concerns. At this point, operators start making the transition from
200 Chapter 7

a reactive to proactive approach, that is, from quality control to quality


assurance.
The aim is to delight customers by providing the highest quality prod-
ucts while at the same time achieving defect-free production.

Pillar #5 Training and Development


Under this pillar we assess technical training needs, determine the
current status of skill sets, and establish a training plan based on the gap
analysis. The goal is to have a multi-skilled workforce and to create a
cadre of experts for supporting all aspects of TPM.
The purpose of providing training is to upgrade the skill set of opera-
tors. It is not sufficient to have only “Know–How” skills. They must also
“Know–Why.” Appropriate training can improve their skill sets to per-
form root cause analysis and other tasks required to improve equipment
effectiveness and reduce costs.

Pillar #6 Design and Early Equipment Management


In this pillar, the lessons of successes and failures of TPM activities
are incorporated into the design of new equipment and products. The goal
is to produce almost perfect equipment and better quality products by tak-
ing care of inefficiencies and safety issues during the design, build, and
commissioning processes.
Maintenance Prevention (MP), which is Design and Early Equipment
Management, involves discovering weak points in the currently-used
equipment and feeding this information back to the equipment design
engineers. It is similar to design for manufacturability. MP design takes
the following factors into consideration:

1. Ease of autonomous maintenance (operator maintenance)


2. Ease of operation
3. Ease of maintenance — improving maintainability
4. Improving quality
5. Safety

Early equipment management is a system for dealing with problems


that surface during the commissioning and start-up of new equipment.
During this period, engineering personnel from production and mainte-
nance / reliability must correct problems caused by poor selection of
materials at the design stage and by errors occurring during fabrication
and installation of the equipment.
Operator Driven Reliability 201

Pillar #7 Office Improvement


In this pillar, the objective is both to eliminate efficiency losses in the
office and service areas and to implement tools such as 5S in order to cre-
ate an organized and efficient office environment. This can be aimed at
logistics, scheduling, HR, accounting, and other areas of the plant admin-
istrative support.

Pillar #8 Safety, Health, and Environment


In this pillar, the focus is on the 100% elimination of accidents as well
as on employee health and environmental concerns. The focus is to create
a safe workplace and surroundings with the following goals:

• Zero accidents
• Zero health concerns (damage)
• Zero environmental incidents

Many consider this pillar the base of all the pillars.

Implementing TPM
Many successful organizations usually follow an implementation
plan that includes the following 10 steps:

Step 1: Announcement of TPM.


Top management needs to create an environment that will support
the introduction of TPM. Without the support of management, skepti-
cism and resistance will kill the initiative.

Step 2: Launch a formal education program.


This program informs and educates everyone in the organization
about TPM activities, benefits, and the importance of contribution from
everyone.

Step 3: Create an organizational support structure.


This group will promote, coordinate, and sustain team-based TPM
activities. It needs to include members from every level of the organiza-
tion — from management to the shop floor. This structure will promote
communication and will guarantee everyone is working toward the same
goals.
202 Chapter 7

Step 4: Establish basic TPM policies and quantifiable goals.


Analyze the existing conditions, and then establish TPM policies
and set attainable and realistic goals.

Step 5: Outline a detailed master deployment plan.


This plan will identify what resources will be needed as well as
when they will be needed for training, equipment restoration and
improvements, maintenance management systems, and new technolo-
gies.

Step 6: TPM kick-off.


TPM implementation will begin at this stage.

Step 7: Improve the effectiveness of each piece of equipment.


Operations and Maintenance Kaizen teams will analyze each piece
of equipment and implement necessary improvements on a continuing
basis.
• Step 7a: Develop an autonomous maintenance program for
operators. Operators will routinely clean, inspect, and perform
minor maintenance tasks that will help to stabilize and improve
equipment conditions.
• Step 7b: Develop a planned or preventive maintenance pro-
gram. Create a schedule for preventive maintenance on each
piece of equipment.
• Step 7c: Identify losses / waste and implement reduction
plan. Create Kaizen teams to eliminate or reduce waste.

Step 8: Conduct training to improve operation and maintenance skills.


The maintenance department will take on the role of trainers or guides
and provide training, advice, and equipment information to the operators
(Kaizen teams).

Step 9: Develop an early equipment management program.


Lessons learned in operations and maintenance should be communi-
cated to the design process of new equipment development. Reliability
and maintainability should be built into the new design.

Step 10: Continuous improvement.


As in any lean initiative, the organization needs to develop a continu-
ous improvement mindset.
Operator Driven Reliability 203

7.5 Workplace Organization: 5S

Five S (5S) is a technique to reduce waste and optimize productivity


by maintaining an orderly workplace and using visual cues to achieve
more consistent operational results. 5S promotes a cleaner environment
and a better organized workplace.
It is a structured program to achieve total organization-wide cleanli-
ness and standardization in the workplace. A well-organized workplace
results in a safer, more efficient, and more productive operation. It boosts
the morale of the employees, promoting a sense of pride in their work, and
a responsible ownership of their equipment.
5S was invented in Japan and stands for five Japanese words that start
with the letter S: Seiri, Seiton, Seiso, Seiketsu, and Shitsuke. An equiva-
lent set of five S words in English has been adopted by many to preserve
the 5S acronym in Japanese. These are:

S1 Sort (Seiri)
S2 Set-in-Order (Seiton)
S3 Shine (Seiso)
S4 Standardize (Seiketsu)
S5 Sustain (Shitsuke)

S1 Sort (Seiri)
Sort is the first step in making a work area tidy. It refers to the act of
throwing away all unwanted, unnecessary, and unrelated materials in the
workplace and freeing up additional space. This step makes it easier for
operators and maintainers to find the things they need. This step requires
keeping only what is necessary. Materials, tools, equipment, and supplies
that are not frequently used should be moved to a separate, common-stor-
age area. Items that are not used should be discarded. Don’t keep things
around just because they might be used someday.
As a result of the sorting process, we will eliminate (or repair) broken
equipment and tools. Obsolete fixtures, molds, jigs, scrap material, waste,
and other unused items and materials are discarded.
People involved in Sort must not feel sorry about having to throw
away things. The idea is to insure that everything left in the workplace is
related to work. Even the number of necessary items in the workplace
must be kept to its absolute minimum. Because of the Sort concept, the
simplification of tasks, effective use of space, and careful purchase of
items will follow.
204 Chapter 7

S2 Set-in-Order (Seiton)
Set-in-order (also, sometimes known as Straighten) or orderliness is
the second step and is all about efficiency. It requires organizing, arrang-
ing, and identifying everything in a work area. Everything is given an
assigned place so that it can be accessed or retrieved quickly, as well as
returned to that same place quickly. If everyone has quick access to spe-
cific items or materials, work flow becomes efficient, and the worker
becomes productive. The correct place, position, or holder for every tool,
item, or material must be chosen carefully in relation to how the work will
be performed and who will use which items. Every single item must be
allocated its own place for safekeeping. Each location must be labeled for
easy identification of its purpose.
Commonly-used tools should be readily available. Properly label stor-
age areas, cabinets, and shelves. Clean and paint floors to make it easier
to spot dirt, waste materials, and dropped parts and tools. Outline areas on
the floor to identify work areas, movement lanes, storage areas, finished
product areas, etc. Put shadows on tool boards, making it easy to quickly
see where each tool belongs.
In an office environment, provide bookshelves for frequently used
manuals, books, and catalogs. Label the shelves and books so that they are
easy to identify and return to their proper place.
Again, the objective in this step is to have a place for everything and
everything in its place, with everything properly identified and labeled.
Many M&R professionals have started calling these practices of using
labels, color-coded markings, etc., a visual workplace. This practice helps
operators and anyone unfamiliar with the asset or process to readily iden-
tify what is going on, understand the process, and know what is to be done
correctly and what is out of place. A visual workplace uses visual displays
to relay information to operators and other employees, and to guide their
actions.
Figure 7.2a indicates safe operating parameters for oil and pressure
levels and also when to tighten the chain or belt. Figure 7.2b displays an
organized tool box and provides examples of labels applied to a switch
box and a Danger Area. Figure 7.2c demonstrates the ASME standard
color code scheme suggested for piping. Figure 7.2d shows examples of
color-coded pipes and hoses.
Operator Driven Reliability 205

Figure 7.2a Safe oil and pressure levels and chain tightening

Figure 7.2b Organized tool box and labeling


206 Chapter 7
Figure 7.2c ASME Standard Color Codes
Operator Driven Reliability 207

Figure 7.2d Examples of Color-Coded Pipes and Hoses

S3 Shine (Seiso)
Shine is all about cleanliness and housekeeping. The Seiso principle
says that everyone is a janitor. The step consists of cleaning up the work-
place and giving it a shine. Cleaning must be done by everyone in the
organization, from operators to managers. It would be a good idea to have
every area of the workplace assigned to a person or group of persons for
cleaning. Everyone should see the workplace through the eyes of a visitor
— always wondering if it is clean enough to make a good impression.
While cleaning, it’s easy to inspect the equipment, machines, tools,
and supplies we work with. Regular cleaning and inspection makes it easy
to spot lubricant leaks, equipment misalignment, breakage, missing tools,
and low levels of supplies. Problems can be identified and fixed when
they are small. If these minor problems are not addressed while small,
they could lead to equipment failure, unplanned outages, or long, unpro-
ductive waits while new supplies are delivered.
When done on a regular, frequent basis, cleaning and inspecting gen-
erally will not take a lot of time. In the long run, they will most likely save
time.

S4 Standardize (Seiketsu)
The fourth step is to simplify and standardize. Seiketsu translates to
standards for all operational activities, including cleanliness. It consists of
208 Chapter 7

defining the standards by which personnel must measure and maintain


operating and maintaining standards such as lubrication plans, filter
change out instructions, or measures of cleanliness. Visual management is
an important ingredient of Seiketsu. Color-coding and standardized col-
oration of surroundings are used for easier visual identification of anom-
alies in the surroundings. Employees are trained to detect abnormalities
using their five senses and to correct such abnormalities immediately.
One of the hardest steps is avoiding old work habits. It’s easy to slip
back into what we have been doing for years. That’s what everyone is
familiar with. It feels comfortable.
The good practices developed in earlier steps should be standardized
and made easy to accomplish. As we learn more, update and modify the
standards to make the process simpler and easier.

S5 Sustain (Shitsuke)
The final step is to sustain the gain by continuing education, training,
and maintaining the standards. In fact, Shitsuke means discipline. It pro-
motes commitment to maintaining orderliness and to practicing the first
four steps as a way of life. The emphasis of Shitsuke is elimination of bad
habits and constant practice of good ones.
Continue to educate people about maintaining standards. When there
are changes such as new equipment, new products, and new work rules
that will affect the 5S program, adjustments will be needed to accommo-
date those changes, to modify changes in the standards, and to provide
training that addresses those changes.
If your organization is planning to implement Lean Manufacturing,
5S is one of the first activities that need to be carried out once Lean has
been adopted.
Some organizations have added a sixth S to emphasize safety in their
program, calling the program 5S Plus or 6S.

7.6 Overall Equipment Effectiveness (OEE)


Overall Equipment Effectiveness (OEE) is a key metric used in TPM
and Lean Manufacturing programs to measure the effectiveness of TPM
and other initiatives. It provides an overall framework for measuring pro-
duction efficiency. OEE is the traditional and most widely used metric to
measure equipment and asset productivity based on actual availability,
performance efficiency, and product quality. However, true equipment
productivity is measured by Total Effective Equipment Performance
Operator Driven Reliability 209

(TEEP), which is based on 24 hours per day and 365 days per year oper-
ations. TEEP also considers equipment utilization.
OEE and TEEP measure the overall utilization of assets and equip-
ment for manufacturing operations, directly indicating the gap between
actual and ideal performance. OEE quantifies how well a manufacturing
unit performs relative to its designed capacity, during the periods when it
is scheduled to run. TEEP measures how well an organization creates
value from its assets by effective utilization based on 24 hours per day,
365 days per year availability.
OEE and TEEP are calculated as:

OEE = Availability X Performance X Quality


TEEP = Utilization X Availability X Performance X Quality
TEEP = Utilization X OEE

OEE breaks the performance of an asset into three separate but meas-
urable elements: availability, performance, and quality. Each element
points to an aspect of the process that can be targeted for improvement.
OEE may be applied to any individual asset or to a process. It is unlikely
that any manufacturing process can run at 100% OEE. Many manufactur-
ers benchmark their industry to set a challenging target; 85% is not
uncommon.
Figure 7.3 illustrates the concept of OEE and TEEP and how differ-
ent production losses impacts productivity.

Figure 7.3 OEE and TEEP


210 Chapter 7

Calculating OEE

OEE = Availability x Performance x Quality

Example 7.1
A given asset, a machining center, experiences the following:
Availability of asset = 88.0%
Asset Performance = 93.0%
Quality it produces = 95.0%
OEE = 88% (Availability) X 93% (Performance) X 95%
(Quality) = 77.7%

Calculating TEEP
TEEP = Utilization X Availability X Performance X Quality

Example 7.2
Whereas OEE measures effectiveness based on scheduled hours,
TEEP measures effectiveness against 24 hours per day, 365 days per
year operation. In the example above, suppose this same asset — the
machining center — operates 20 hours a day, 300 days in a year.
OEE of machinating center (calculated above) = 77.7%
Machining Center Utilization
= (20 hours X 300 days) / (24 hours X 365 days) = 68.5%
TEEP = 68.5% (Utilization) X 77.7% (OEE) = 53.2%

Example 7.3
A six-station hammer assembly machine shows the following opera-
tional data from the CMMS and operational log of an assembly machine.

Machine Cycle Time (design): 1 unit/minute


Scheduled Time:
2 shifts/day X 10 hours/shift X 250 days/year = 5,000 hours/year

Note that 5 days of production were cancelled in the year, including


4 days due to lack of wood handles (raw material) and 1 day due to loss
of electric power resulting from a winter storm.
Operator Driven Reliability 211

Scheduled Downtime:
5 PMs, each at 1000 operating hours, each requiring 2 people at 8
hours each of downtime (16 man hours per PM)

1 PM at 5000 operating hours, requiring 4 people each work-


ing 10 hours of downtime (40 man hours per PM)

50 weekly checks by operators, 30 minutes each

10 planned repairs requiring a total of 55 hours of downtime

50 models set ups, 2 hours each

Unscheduled Downtime:
8 failures resulting in 50 hours of downtime (132 man hours of fail-
ure repair work) 22 setups and tooling changes resulting in 20 hours
of downtime

Performance Losses:
Minor stoppages / jams (Less than 5 min each) 750 instances per
year — average 3.2 minute each

During winter (about 120 days in year), the system runs slower in
the morning for 30 minutes. This increases the machine cycle time
from 1 unit/minute to 1 unit / 1.5 minutes during this period.

Quality Losses:
On average, every hour the assembly unit produces 57 good quality
units, 2 units needing some repair, and 1 unit scrapped.

Calculate OEE and TEEP for this assembly system.

Solution for Example 7.3:

Asset Utilization
Ideally, total hours available for production
= 365 days X 24 hours/day = 8,760 hours/year

Idle hours = Hours the asset doesn’t run due to lack of demand or
factors beyond the control of asset/plant.
212 Chapter 7

= (4 hours/day X 250 days) + [24 hours X (365 – 250) days] + (24


hours X 5 days)
= 1,000 + 2760 + 120
= 3880 hours

Gross hours of scheduled production


= Total available hours – Idle hours = 8760 – 3880
= 4880 hours /year

Asset utilization rate


= Gross hours of scheduled production / Total available hours
= 4880 / 8760 = 55.7%

Asset Availability
Asset availability (%) = Uptime x 100 / (Uptime + Downtime)

Uptime hours = Gross hours of scheduled production – Downtime


hours

Downtime hours = Scheduled and unscheduled downtime hours


Scheduled downtime hours = (5 x 8) + 10 + (50 x 0.5) + 55 + (50 x
2) = 230 hours

Unscheduled downtime hours = 50 + 20 = 70 hours


Total downtime hours = 230 + 70 = 300 hours

Uptime hours = 4880 – 300 = 4580 hours

Asset availability (%) = 4580 / (4580 + 300) = 93.9%

Asset Performance
Asset performance – Efficiency in %
= Actual production rate / Designed (best) Production rate x 100

Designed production rate or Cycle time = 1 minute /unit (60 units


per hour)

Performance losses
Minor stoppages (Hours/year) = 750 X 3.2 min = 40 hours
Speed losses (machine running slow) = 120 x 0.5 x [(1.5 - 1) / 1.5]
Operator Driven Reliability 213

= 20 hours/year Total Performance losses (Hours/year) = Minor


stoppages + Speed losses = 40 + 20 = 60 hours

Performance Efficiency %
= (Uptime hours – Performance losses) / Uptime hours
= (4580-60) / 4580 = 98.7%

Quality Losses
The quality portion of the OEE represents the good units produced
as a percentage of the total units. The quality performance is a pure
measurement of process yield that is designed to exclude the effects
of availability and performance.

On average, the hammer assembly machine produces 57 good units


per hour out of 60. The other 3 units are usually reworked or
scraped.

Quality Performance (%) = 57 / 60 = 95%

OEE for hammer assembly machine


= Availability X Performance X Quality
= 93.9% X 98.7% X 95% = 88.0%

TEEP for hammer assembly machine


= Utilization X OEE
= 55.7% X 88.0% = 49.0%

7.7 Measures of Performance

Several measures and performance indicators indicate the efficiency


of the operations. These include:

• Downtime as % of scheduled hours (operating time)


• Downtime due to operational issues
• Downtime due to maintenance
• Downtime due to other issues, e.g., design, act of God, etc..
• Total Downtime
214 Chapter 7

• 5 S audit results
• % of assets covered by 5 S plus principles
• Asset — area cleanliness / housekeeping
• Asset condition — visual inspection
• Color coded labels of piping, hoses, valves, etc.
• Check lists — PMs instructions attached to the asset
• Required tools properly placed and labeled

• Asset Performance — throughput


• Output as % of designed / demonstrated capacity

• Operators – % operational personnel involved with asset


improvement projects
• RCM analysis support
• Design reviews
• Kaizen activities

• Percent operators qualified / certified to operate assets and


understand their role
(one of the roles is to support maintenance)

• Percent assets ready — delivered on time to maintenance per


agreed schedule

• OEE – Overall Equipment Effectiveness trend


• OEE – Availability X Performance X Quality

• TEEP — Total Effective Equipment Performance


• TEEP — Utilization X OEE

It is a good practice to track these performance measures on a regu-


lar basis to evaluate the improvements that have been made or to set up
improvement goals.

7.8 Summary

Reliable equipment operating at the lowest possible cost is an essential


enabler of an organization’s profit. Assets and plants are often the single
largest investment; it would make sense that asset reliability should be as
Operator Driven Reliability 215

important to the organization as are environment, health, quality, and safety.


But asset reliability has not received its due emphasis in the past. Operator
Driven Reliability (ODR) and Total Productive Maintenance (TPM) strate-
gies encourage participation of all employees, specifically operators. These
strategies provide a framework for policies, procedures, and structure to have
assets available and operated at the lowest cost possible.
By making equipment more efficient, TPM is focused on keeping assets
functioning optimally and minimizing equipment breakdowns and associat-
ed waste. Autonomous maintenance, a key pillar of TPM, seeks to eliminate
major losses that can result from faulty equipment or operation by involv-
ing operators in maintenance of equipment they operate. Under the TPM
concept, equipment operators become owners of their assets. Working
closely with maintenance, they take care of all details that will preserve the
assets in the best possible condition.
TPM has eight pillars of activity, with an ultimate goal of zero break-
downs and zero accidents. These pillars are:

1. Autonomous maintenance
2. Focused improvement — Kaizen
3. Planned maintenance
4. Quality maintenance
5. Training and development
6. Design and early equipment management
7. Office improvement
8. Safety, health, and environment

Under autonomous maintenance, which is a key pillar, operators clean


and lubricate the equipment and execute the recommended maintenance
plan. They are empowered to modify the program according to the real
needs and personal observations. The operator has access to the manufac-
turer specifications and the support of the maintenance technicians.
The operators also become responsible for small adjustments, checking
for parts that become loose, and fixing them, as well as reporting small
details like noises, vibrations, or temperature changes while operating the
equipment.
An important factor in the success of the TPM program is the pride that
operators experience from the optimal condition in which their equipment is
preserved. A great deal of this improved effectiveness comes from the moti-
vation given to the employees through adequate training and education.
216 Chapter 7

TPM is an organization-wide equipment improvement strategy, not a


maintenance improvement strategy. It requires a systematic focus on elim-
inating equipment-related losses. It is not a program to just clean and paint
equipment to look good. TPM demands and encourages the involvement
of all employees, not merely involving operators in performing some ele-
ments of maintenance.
5S — a visual workplace system — has five elements: sort, set in
order, shine, standardize, and sustain. These elements are the most funda-
mental and often overlooked aspects in continuous improvement initia-
tives. 5S is a structured program. If properly implemented, it can achieve
total organization-wide cleanliness and standardization in the workplace.
A well-organized workplace results in a safer, more efficient, and more
productive operation. It boosts the morale of the employees, promoting a
sense of pride in their work and ownership of their responsibilities.
Overall Equipment Effectiveness (OEE) is a key metric that quanti-
fies how well an asset or a manufacturing process performs relative to its
designed capacity, during the periods when it is scheduled to run. It is cal-
culated by multiplying asset availability, performance, and quality of
products it produces. Another TPM related metric is Total Equipment
Effectiveness Performance (TEEP) which measures how well an organi-
zation creates value from its assets based on 24 hours per day, 365 days
per year availability.
The benefits of TPM are:

• Safer workplace
• Employee empowerment and improved morale
• Increased production / output
• No or minimum defects
• No or minimum breakdowns
• No or fewer short stoppages
• Decreased waste
• Decreased O&M costs

Total Productive Maintenance thrives on the spirit of teamwork. It


has a long-range outlook and may take years to implement. It works not
only in the manufacturing industry, but also in the service industry, con-
struction, building maintenance, and other industrial situations.
Operator Driven Reliability 217

7.8 Self Assessment Questions

Q7.1 Explain operator-driven reliability. Why is the operator’s


involvement important in maintenance?

Q7.2 Define TPM. What are TPM’s various elements — the pillars
of TPM?

Q7.3 How do we implement TPM?

Q7.4 Define OEE. How do we measure it?

Q7.5 What is the difference between OEE and TEEP?

Q7.6 Explain 5S. What benefits do we derive from implementing 5S?

Q7.7 What is the difference between 5S plus or 6S?

Q7.8 Explain what is meant by the visual workplace?

Q7.9 Explain Muda, Mura, and Muri?

Q7.10 What are the benefits of standardizing?

7.9 References and Suggested Reading

Hansen, Robert C. Overall Equipment Effectiveness (OEE). Industrial


Press, 2005.
Hartmann, Edward. Successfully Installing TPM. TPM Press, 1992.
Maggard, Bill. TPM that Works. TPM Press, 1992.
Productivity Press Development Team, OEE for Operators: Overall
Equipment Effectiveness. Productivity Press, 1999.
Williamson, Robert M. Visual Systems for Improving Equipment
Effectiveness, Lean Equipment Management, and other articles.
Strategic Works System, www.swspitcrew.com
5S for Operators: 5 Pillars of the Visual Workplace.
Productivity Press.
Chapter 8
Maintenance Optimization

“Innovative practices combined with true empowerment produce


phenomenal results.”
Captain Michael Abrashoff, Former Commanding Officer,
USS Benfold, Author, It’s Your Ship

8.1 Introduction
8.2 Key Terms and Definitions
8.3 Understanding Failures and Maintenance Strategies
8.4 Maintenance Strategy — RCM
8.5 Maintenance Strategy — CBM
8.6 Other Maintenance Strategies
8.7 Summary
8.8 Self Assessment Questions
8.9 References and Suggested Reading

After reading this chapter, you will be able to understand:

• What is a failure?
• What is RCM?
• What does it take to implement RCM effectively?
• What CBM technologies are available?
• What are the different maintenance strategies?
• How can you integrate PM and CBM into RCM methodology
• When would RTF be a good maintenance strategy?
219
220 Chapter 8

8.1 Introduction
Maintenance has entered the heart of many organizational activities
due to its vital role in the areas of environment preservation, productivity,
quality, system reliability, regulatory compliance, safety, and profitability.
With this new paradigm, new challenges and opportunities are being pre-
sented to maintenance and operations professionals. Central to mainte-
nance is a process called Reliability Centered Maintenance, or RCM.
RCM helps determine how assets can continue to do what their users
require in certain operating contexts. RCM analysis provides a structured
framework for analyzing the functions and potential failures of assets such
as airplanes, manufacturing lines, compressors or turbines, telecommuni-
cation systems, etc. RCM was developed in the commercial aviation
industry in the late 1960s to optimize maintenance and operations activi-
ties. RCM strategy (or some call it process), can help in developing an
effective maintenance plan by selecting appropriate strategies such as PM,
CBM, or RTF.
Preventive maintenance (PM) is the planned maintenance of assets
designed to improve asset life and avoid unscheduled maintenance activ-
ity. PM includes cleaning, adjusting, and lubricating, as well as minor
component replacement, to extend the life of assets and facilities.
Condition-based maintenance (CBM) is another maintenance opti-
mizing strategy. CBM attempts to evaluate the condition of assets by per-
forming periodic or continuous condition monitoring. The ultimate goal
of CBM is to perform maintenance at a scheduled point in time when the
maintenance activity is most cost-effective and before the asset loses opti-
mum performance.
Recent developments in technologies have allowed instrumentation of
assets to provide us information regarding its health. Together with better
tools for analyzing condition data, today’s maintenance personnel are bet-
ter able to decide the right time to perform maintenance on assets. Ideally,
CBM allows maintenance personnel to do the right things — minimizing
asset downtime, time spent on maintenance, and spare parts cost. CBM
uses real-time data to prioritize and optimize resources.
Although many would not consider it to be a maintenance-optimizing
strategy, Run-to-Failure (RTF) can be a viable and economical choice for
certain equipment. Selecting RTF should be a deliberate choice because it
will lead to unplanned downtime and increased corrective maintenance
cost for the specific equipment selected for this maintenance strategy.
However, if the facility and personnel risk is low, RTF may be the most
cost-effective strategy for an organization’s overall maintenance program.
Maintenance Optimization 221

The key to optimizing your facility’s maintenance program is to make


the best choice for each piece of the equipment as well as the facility as a
whole, although the best application of RCM is during design and devel-
opment of equipment to eliminate or mitigate effects of failure modes.

8.2 Key Terms and Definitions

Age Exploration
An iterative process used to optimize preventive maintenance
(PM) intervals.

Condition-Based (or Predictive) Maintenance (CBM / PdM)


Maintenance based on the actual condition (health) of assets
obtained from in-place, non-invasive measurements and tests.

Condition-Directed (CD) Tasks


Tasks directly aimed at detecting the onset of a failure or failure
symptom.

Corona (Partial Discharge)


The term corona is used as a generic name for any electrical
discharges that take place in an energized electrical insulation
as the result of accelerated ionization under the influence of the
electric field in the insulation. It is defined as a type of local-
ized discharge resulting from transient gaseous ionization in an
insulation system when the voltage stress exceeds a critical
value.

Critical Asset
Assets that have been evaluated and classified as critical due to
their potential impact on safety, environment, quality, produc-
tion/operations, and maintenance if failed.

Emissivity
A fundamental property of a material, emissivity is the ratio of
the rate of radiant energy emission at a given wavelength from
a body with an optical smooth surface, as a consequence of its
temperature only, to the corresponding rate of emission from a
black body at the same temperature and wavelength.
222 Chapter 8

Failure
Failure is the inability of an asset / component to meet its
expected performance.

Failure Cause
The reason something went wrong.

Failure Effect (Consequences)


What happens when a failure mode occurs; its consequences.

Failure-Finding (FF) Tasks


A scheduled task that seeks to determine if a hidden failure has
occurred or is about to occur.

Failure Mode
An event that causes a functional failure; the manner of failure

Failure Mode Effect and Analysis (FMEA)


A technique to examine an asset, process, or design to deter-
mine potential ways it can fail and the potential effects; and
subsequently identify appropriate mitigation tasks for highest
priority risks.

Ferrography
An analytical method of assessing machine health by quantify-
ing and examining ferrous wear particles suspended in the
lubricant or hydraulic fluid.

Functional Failure
A state in which an asset / system is unable to perform a specif-
ic function to a level of performance that is acceptable to its
user.

Hidden Failure
A failure mode that will not become evident to a person or the
operating crew under normal circumstances.
Maintenance Optimization 223

Operating Context
The environment in which an asset is expected to be used.

P–F Interval
The interval between the point at which a potential failure
becomes detectable and the point at which it degrades into a
functional failure. It is also sometime called lead time to fail-
ure.

Potential Failure
A condition that indicates a functional failure is either about to
occur or in the process of occurring.

Prognosis
A forecast or prediction of outcome such as how long this asset
or component will last or remaining life left.

Reliability Centered Maintenance (RCM)


A systematic and structured process to develop an efficient and
effective maintenance plan for an asset to minimize the proba-
bility of failures. The process insures safety and mission com-
pliance.

Run–to-Failure (RTF)
A maintenance strategy (policy) for assets where the cost and
impact of failure is less than the cost of preventive actions. It is
a deliberate decision based on economical effectiveness.

Time-Directed (TD) Tasks


Tasks directly aimed at failure prevention and performed based
on time — whether calendar-time or run-time.

Viscosity
Measurement of a fluid’s resistance to flow. It is also often
referred to as the structural strength of liquid. Viscosity is criti-
cal to oil film control and is a key indicator of condition related
to the oil and the machine.
224 Chapter 8

8.3 Understanding Failures and Maintenance Strategies


Figure 8.1 illustrates the period when a failure initiates and eventual-
ly becomes a functional failure that leads to a complete asset breakdown.
Assets perform very well in Zone A. However, somewhere in that region
— due to a lack of or reduction in lubricant supply, human error, defect in
material, or some other reason —a failure is initiated at the end of Zone
A. This defect may be in the form of a small crack or debris stuck in the
lubricant or in the valve assembly, etc. It continues to grow in Zone B,
increasing the asset’s failure potential, though still unnoticed. At Point P
at the beginning of Zone C, this defect becomes a Potential Failure. Then
at Point F, at the end of Zone C, this potential failure creates a functional
failure, and a function of the asset stops working.
However, the asset may continue to operate at a reduced capacity or
functionality. By now, there will be some visual and/or measurable evi-
dence of a functional failure. Eventually, at Point B, the asset completely
shuts down.
The time interval between points P and F is called the P–F interval.
In theory, the PM or on-condition tasks interval should be less than the
P–F interval time to catch potential failures and correct them in time.
However, we don’t have good information about where points P or F are
in time. Analysis of condition and operating data can help to estimate their

Figure 8.1 Understanding Failure


Maintenance Optimization 225

(time) location. Our discussion assumes that these points are fixed in time,
yet this is not the case in practice. They may vary based on the nature of
the defects and the environment. Our goal is to catch any defects before
they shut us down.
The best strategy is to find a defect or any abnormal condition in
Zone B as soon as possible, utilizing condition-based tasks. RBM and/or
PM can be used to identify the sources of these defects and correct them
in their early stages.
Traditional thinking has been that the goal of preventive mainte-
nance (PM) is to preserve assets. On the surface, it makes sense, but the
problem is in that mindset. In fact, that thinking has been proven to be
flawed at its core. The blind quest to preserve assets has produced many
problems, such as being overly conservative with any maintenance
actions that could cause damage due to intrusive actions, thereby increas-
ing the chances of human error. Other flaws include both thinking that all
failures are equal and performing maintenance simply because there is an
opportunity to do so.
In the last few decades many initiatives have been developed in
cost reduction, resource optimization, and bottom line focus of any action
we take. The mentality of preserving assets quickly consumed resources,
put maintenance plans behind schedule, and overwhelmed the most expe-
rienced maintenance personnel. Worse, this mentality sometimes caused
maintenance actions to become totally reactive.
The development of a Reliability Centered Maintenance approach
has provided a fresh perspective in which the purpose of maintenance is
not to preserve assets for the sake of the assets themselves, but rather to
preserve asset functions. At first, this might be a difficult concept to accept
because it is contrary to our ingrained mindset that the sole purpose of
preventive maintenance is preserving equipment operation. But in fact, in
order to develop an effective maintenance strategy, we need to know what
the expected output is and the functions that the asset supports — that is,
the real purpose of having the asset.

8.4 Maintenance Strategy — RCM

Reliability-Centered Maintenance, often known as RCM, is a process


to ensure that assets continue to do what their users require in their pres-
ent operating context. It is a structured process to develop an efficient and
effective maintenance plan for an asset to minimize the probability of fail-
ures.
226 Chapter 8

RCM is generally used to achieve improvements in all aspects of the


asset management, such as the establishment of safe minimum or opti-
mum levels of maintenance, changes to operating procedures, and estab-
lishment of an effective maintenance plan. Successful implementation of
RCM will promote cost effectiveness, asset uptime, and a better under-
standing of the level of risk that the organization is presently managing.
It has been demonstrated that the best application of RCM is during
design and development phases of the assets to eliminate or mitigate
effects of failure modes.

RCM History and Development


Reliability Centered Maintenance (RCM) is a systematic approach for
developing new maintenance requirements where one does not exist, and
optimizing an existing maintenance program. In both cases, the end result
of the RCM analysis is a maintenance program composed of tasks that
represent the most technically correct and cost-effective approach to
maintaining asset/component operability. This operability in turn lends
itself to improved system reliability and plant availability. Another impor-
tant result of an RCM program is a documented technical basis for every
maintenance program decision.
In the late 1960s as Boeing’s 747 jumbo jet was becoming a reality,
all owners/operators of the aircraft were required to provide a PM pro-
gram to FAA for approval in order to get certified for operation. No air-
craft can be sold without this type of certification. The recognized size of
the 747, about three times as many passengers as the 707 or DC-8, and its
many technological advances in structure and avionics led the FAA to take
the position at first that the preventive maintenance on the 747 would be
very extensive. In fact, airlines thought that they may not able to operate
this aircraft in a profitable manner with that requirement.
This development essentially led the commercial aircraft industry to
undertake a complete re-evaluation of their preventive maintenance strat-
egy. Bill Mentzer, Tom Matteson, Stan Nowland, and Harold Heap of
United Airlines led the effort. What resulted was an entirely new approach
that employed a decision-tree process for ranking PM tasks that were nec-
essary to preserve critical aircraft functions during flights. This new tech-
nique was defined and explained in Maintenance Steering Group 1
(MSG1) for the 747 and was subsequently approved by the FAA.
With MSG-1 success, its principles were applied to other aircrafts
such as DC-10, MD-80/90, and Boeing 757 / 777; to Navy P-3 and S3;
Maintenance Optimization 227

and to Air Force F-4J aircrafts under a contract with the U.S. Department
of Defense (DOD). In 1975, DOD directed that the MSG concept be
labeled Reliability–Centered Maintenance (RCM) and be applied to all
major military systems. In 1978, United Airlines produced the initial
RCM “bible” under DOD contract.
RCM development has been an evolutionary process. Over 40 years
have passed since its inception during which RCM has become a mature
process. However, industry has yet to fully embrace the RCM methodol-
ogy in spite of its proven track record. In recent years, Anthony (Mac)
Smith and Jack Nicholas have been leaders in creating increased RCM
awareness. Examples discussed in this section are the result of work per-
formed by Mac Smith, Glen Hinchcliffe and the author in optimizing PMs
utilizing RCM methodology.

The RCM Principles


There are four principles that define and characterize RCM, and set it
apart from any other PM planning process.

Principle 1: The primary objective of RCM is to preserve system


function.
This principle is one of the most important and perhaps the most dif-
ficult to accept because it is contrary to our ingrained notion that PMs are
performed to preserve equipment operation. By addressing system func-
tion, we want to know what the expected output should be, and also
understand that preserving that output (function) is our primary task at
hand.

Principle 2: Identify failure modes that can defeat the functions.


Because the primary objective is to preserve system function, the loss
of function is the next item of consideration. Functional failures come in
many sizes and shapes; they are not always as simple as, “we have it or
we don’t.” For example, the loss of fluid boundary integrity in a pumping
system illustrates this point. A system loss of fluid can be 1) a very minor
leak that may be qualitatively defined as a drip; 2) a fluid loss that can be
defined as a design basis leak — that is, any loss beyond a certain flow
value will produce a negative effect on system function, but not necessar-
ily total loss; or 3) a total loss of boundary integrity, which can be defined
as a catastrophic loss of fluid and loss of function. In this example, a sin-
gle function — preserve fluid integrity — led to three functional failures.
228 Chapter 8

The key point of Principle 2 is to identify the specific failure modes in a


specific component that can potentially produce those unwanted function-
al failures.

Principle 3: Prioritize function needs (failures modes).


All functions are not equally important. A systematic approach is
taken to prioritize all functional failures and failure modes using a priori-
ty assignment rationale.

Principle 4: Select applicable and effective tasks.


Each potential PM or CBM task must be judged as being applicable
and effective. Applicable means that if the task is performed, it will
accomplish one of three reasons of doing PM or CBM:

1. Prevent or mitigate failure.


2. Detect onset of a failure.
3. Discover a hidden failure.

Effective means that we are sure that this task will be useful and we
are willing to spend resources to do it. In addition, RCM recognizes the
following:

Design Limitations. The objective of RCM is to maintain the inher-


ent reliability of system function. A maintenance program can only main-
tain the level of reliability inherent in the system design; no amount of
maintenance can overcome poor design. This makes it imperative that
maintenance knowledge be fed back to designers to improve the next
design. RCM recognizes that there is a difference between perceived
design life (what the designer thinks the life of the system is) and actual
design life. RCM explores this through the Age Exploration (AE) process.

RCM Is Driven by Safety First, then Economics. Safety must be


maintained at any cost; it always comes first in any maintenance task.
Hence, the cost of maintaining safe working conditions is not calculated
as a cost of RCM. Once safety on the job is ensured, RCM assigns costs
to all other activities.

Elements of RCM
The SAE JA1011 standard describes the minimum criteria to which a
process must comply to be called RCM. An RCM Process answers the fol-
Maintenance Optimization 229

lowing seven essential questions:

1. What are the functions and associated desired standards of


performance of the asset in its present operating context
(functions)?
2. In what ways can the asset fail to fulfill its functions (functional
failures)?
3. What causes each functional failure (failure modes)?
4. What happens when each failure occurs (failure effects)?
5. In what way does each failure matter (failure consequences)?
6. What should be done to predict or prevent each failure (proactive
tasks and task intervals)?
7. What should be done if a suitable proactive task cannot be found
(default actions)?

Unlike some other maintenance planning approaches, RCM results


in all of the following tangible actionable options:

• Maintenance task schedules, which can include:


• Time Directed (TD) tasks, (Calendar/run time based PMs)
• Condition Directed (CD) tasks, (CBM/PdM tasks)
• Failure Finding (FF) tasks (operator supported tasks)
• Run-to-Failure (RTF) tasks (economical decision based)
• Revised operating procedures for the operators of the assets,
which might include service-type tasks such as changing filters,
taking oil samples, and recording operating parameters
• A list of recommended changes to the design of the asset that
would be needed if a desired performance is to be achieved

RCM shifts the emphasis of maintenance from the idea that all fail-
ures are bad and must be prevented, to a broad understanding of the pur-
pose of maintenance. It seeks the most effective strategy that focuses on
the performance of the organization. It might include not doing something
about a failure or letting failures happen. The RCM approach encourages
us to think of more encompassing ways of managing failures.

RCM Analysis Process


Although RCM has a great deal of variation in its application, most
procedures include some or all of the following nine steps:
230 Chapter 8

1. System selection and information collection


2. System boundary definition
3. System description and functional block diagram
4. System functions and functional failures
5. Failure mode and effects analysis (FMEA)
6. Logic (decision) tree analysis (LTA)
7. Selection of maintenance tasks
8. Task packaging and implementation
9. Making the program a living one — continuous improvements

Step 1: System Selection and information Collection The purpose


of Step 1 is to assure that the RCM team has sufficiently evaluated their
area to know which systems are the problems or so-called bad actors. The
team can use Pareto analysis (80/20 rule) to determine the list of prob-
lems, using criteria of highest total maintenance costs (CM+PM), down-
time hours, and number of corrective actions. Identifying these systems
defines the dimensions of the RCM effort that will provide the greatest
Return–on–Investment. Figure 8.2 lists a plant’s asset data — failure fre-
quency, downtime, and maintenance costs — to help us decide which
assets are the right candidates for RCM analysis. The first few assets list-
ed in Figure 8.2 are good candidates for RCM efforts.

Figure 8.2 Plant Failure and Cost Data by Assets


Maintenance Optimization 231

Selection of RCM team members is a key element in executing a suc-


cessful RCM program. The team should include the following:

• System operator (craft)


• System maintainer (craft – mechanical / electrical / controls)
• Operations / Production engineer
• Systems / Maintenance engineer (mechanical / electrical)
• CBM/PdM specialist or technician
• Facilitator

The use of facilitators is recommended to support RCM efforts. They


ensure that the RCM analysis is carried out at the proper level, no impor-
tant items are overlooked, and the results of the analysis are properly
recorded. Facilitators also manage issues among the team members, help-
ing them reach consensus in an orderly fashion, retaining the members’
commitment, and keeping them engaged.
Another objective of Step 1 is to collect information that will be
required by the team as they perform the system analysis. This informa-

Figure 8.3 System Boundary Definition


232 Chapter 8

tion includes schematics, piping and instrumentation diagrams (P&ID),


vendor manuals, specification and system descriptions, operating instruc-
tions, and maintenance history.

Step 2: System Boundary Definition After a system has been select-


ed, the next step is to define its boundaries — understanding the system as
a whole and its functional sub-systems. This step assures that there are no
overlaps or gaps between adjacent systems. We need to have a clear record
for future reference on exactly what was defined within the system. In
addition, we must specify the boundaries in precise terms; a key portion of
analysis depends on defining exactly what is crossing the boundaries, both
“incoming – IN” and “outgoing – OUT” interfaces respectively.
An example of a system boundary is shown in Figure 8.3

Step 3: System Description and Functional Block Diagram Step 3


requires identifying and documenting the essential details of the system.
It includes the following information:

• System description
• Functional block diagram
• IN / OUT interfaces
• System work breakdown structure
• Equipment /component history

A well-documented system description will record an accurate base-


line definition of the system as it existed at the time of the RCM analysis.
Various design and operational changes can occur over time. Therefore,
the system must be baselined to identify where PM task revision might be
required in the future. Frequently it has been found that team members
and analysts may have only a superficial knowledge of the system.
Recording a detailed system functional description narrative will assure
that the team has a comprehensive review of the system.
In addition, documenting the following information can be helpful in
analyzing the data later on:

• Redundancy Features: Equipment / component redundancy,


alternate mode operations, design margins, and operator
workaround capabilities
• Protection Features: A list of devices that are intended to prevent
personnel injury or secondary system damages when an unexpect-
Maintenance Optimization 233

ed component failure occurs; it may include items such as inhibit


or permissive signals, alarms, logic, and isolation
• Key Control Features: An overview of how the system is
controlled; also briefly highlighting features such as automatic vs.
manual, central vs. local, and various combinations of the above
as they may apply

The next item in Step 3 is to develop a Functional Block Diagram


(FBD), which is a top-level representation of the major interfaces between

Figure 8.4 Functional Block Diagram


234 Chapter 8

a selected system and adjacent systems. Figure 8.4 illustrates an FBD with
functional interfaces including sub-systems.
In an actual team setting, it is desirable to have a discussion first
regarding various possibilities that should be considered in creating func-
tional sub-systems. When the FBD is finalized, it will show a decision on
the use of functional subsystems as well as the final representation of the
IN / OUT interfaces.
Listing all components as part of the System Work Breakdown
Structure (SWBS) is very desirable. The SWBS is the compilation of the
line items list for the system. SWBS is a system hierarchy listing parent-
child relationships. In most cases, the SWBS should be what’s in the
CMMS for the system being analyzed. In older plants, where the reference
sources could be out of date, the RCM team should perform a system walk
down to assure accuracy in the final SWBS. This practice is a good one,
even if the system is well documented, to help the team familiarize itself
with the system.
The last item in Step 3 is to collect historical system data. It will be
beneficial for the analysis team to have a history of the past 2–5 years of
component and system failure events. This data should come from correc-
tive maintenance reports or from the CMMS system. Unfortunately, it is
not uncommon to find a scarcity of useful failure event information. In
many plants, the history kept is of very poor quality. Most of the time, the
repair history will simply state “Repaired pump” or “Fixed pump.”
Improving data quality is a challenge for many organizations. If a good
failure history is not available, the team can work together to develop a
list of failure events over the last few years. This list of failures would
help in performing the FMEA analysis in Step 5.

Step 4: System Functions and Functional Failures Because the


ultimate goal of RCM is “to preserve system function,” it is incumbent
upon the RCM team to define a complete list of system functions and
functional failures. Therefore, in Step 4, system functions and functional
failures are documented.
The function statement should describe what the system does — its
functions. For example, a correct function might be “Maintain a flow of
1000 GPM at header 25,” but not “provide a 1000 GPM centrifugal Pump
for discharge at header 25.” Another example would be “maintain lube oil
temperature 110°F. In theory, we should be able to stand outside of a
selected system boundary, with no knowledge of the SWBS for the sys-
tem, and define the functions by simply what is leaving the system (OUT
Maintenance Optimization 235

interfaces).
The next step is to specify how much of each function can be lost, i.e.,
functional failures. Most functions have more than one loss condition if

Figure 8.5 Functions / Functional Failures

we have done a good job with the system description. For example, the
loss condition can range from total loss and varying levels of partial loss
which have different levels of plant consequences (and thus priority) to
failure to start on demand, etc. The ultimate objective of an RCM analy-
sis is to prevent these functional failures and thereby preserve function. In
Step 7, this objective will lead to the selection of preventive maintenance
tasks that will successfully avoid the really serious functional failures.
236 Chapter 8

Step 5: Failure Mode and Effects Analysis (FMEA) Step 5, Failure


Mode and Effects Analysis, is the heart of the RCM process. FMEA has
been used traditionally to improve system design and is now being used
effectively for failure analysis that is critical to preserve system function.
By developing the functional failure – equipment matrix, Step 5 con-
siders for the first time the connection between function and hardware.
This matrix lists functional failures from Step 4 as the horizontal elements
and the SWBS from Step 3 as the vertical elements. The team’s job at this
point is to ascertain from experience whether each intersection between
the components and functional failure contains the making of some mal-
function that could lead to a functional failure. The team completes the
matrix by considering each component’s status against all functional fail-
ures, moving vertically down the component list one at a time. After the
entire matrix has been completed, it will produce a pattern of Xs that
essentially constitutes a road map to guide to a more detailed analysis.
The next step in this process is to perform the FMEA (which is dis-
cussed in Chapter 11), considering each component and functional fail-

Figure 8.6 Example of FMEA


Maintenance Optimization 237

ures as shown in Figure 8.6. FMEA addresses the second RCM principle,
to “determine the specific component failures that could lead to one or
more of the functional failures.” These are the failures which defeat func-
tions and become the focus of the team’s attention.
In reviewing failure modes, teams can use the following guidelines in
accepting, rejecting, or putting aside for later considerations:

• Probable Failure Mode. Could this failure mode occur at least


once in the life of the equipment / plant? If yes, it is retained. If
no, it is considered a rare event and is dropped out from further
consideration.
• Implausible Failure Mode. Does this failure mode defy the nat-
ural laws of physics — is it one that just could not ever happen?
There are usually few, if any, hypothesized failure modes in this
category. But if one arises, label it as “Implausible” and drop it
from further consideration.
• Maintainable Failure Mode. Certain failure modes clearly can
pass the above two tests, but in the practical sense would never be
a condition where a preventive action would be feasible. Doing
preventive maintenance on a printed circuit board full of IC chips
is one example where the practical maintenance approach is to
replace it when (and if) it fails.
• Human Error Causes. If the only way this failure mode could
happen is the result of an unfortunate (but likely) human error, we
note as such for the record. But we drop it from further consider-
ation because we really cannot schedule a preventive action to
preclude such random and uncontrollable occurrences. If the
hypothesized human error problem is important, we could consid-
er this condition later in the evaluation of other forms of correc-
tive or mitigating action such as redesign through control logic.

For each failure mode retained for analysis, the team then decides on
its one or two most likely failure causes. A failure cause is, by definition,
a 1–3 word description of why the failure occurred. We limit our judg-
ments to root causes. If the failure mode can occur only due to another
previous failure somewhere in the system or plant, then this is considered
a consequential cause.
Each failure mode retained is now evaluated as to its local effect.
What can it do to the component; what can it do to the system functions;
how can it impact the system / plant output? If safety issues are raised,
238 Chapter 8

they too can become part of the recorded effect. In the failure effects
analysis, assume a single failure scenario. Also allow all facets of redun-
dancy to be employed in arriving at statements of failure effect. Thus,
many single failure modes can have no effect at the plant or system level,
in which case, designate the failure mode as low priority and do not pass
them to Step 6, Logic Tree Analysis. If there is either a system or plant
effect, the failure mode is passed on to Step 6 for further priority evalua-
tion. Those failure modes considered as low priority here are assigned as
candidates for run-to-failure (RTF) and are given a second review in Step
7 for final RTF decision.
RTF does not imply that this component or asset is unimportant;
instead components that are designated as RTF have no significant conse-
quence as the result of a failure. It does not matter if failed components
are restored immediately as long as they are repaired to an operable status
in a timely manner.

Figure 8.7 Logic Tree Analysis Structure


Maintenance Optimization 239

Step 6: Logic (Decision) Tree Analysis (LTA) In Step 6, because of


the fact that not all failures are equally important, we need to screen our
information further to focus on what really counts.
The Logic Tree shown in Figure 8.7 poses three simple questions that
require either a Yes or No answer. The result is that each failure mode will
ultimately be assigned an importance designator that will constitute a nat-
ural ordering of the priority that we should address in allocating our
resources. The following coding can be used to label the failure modes:

A — Top priority item


B — Second and next significant item
C —A low priority item and may likely be a non-PM item to
consider
D — RTF item

The three questions are:


1. Under normal conditions, do the operations know that something
has occurred?
Question 1 looks at operator knowledge that something is not nor-
mal, given the occurrence of the failure mode. It is not necessary
that operators know exactly what failure mode has occurred. They
may pinpoint the exact failure instantly. If they sense an abnor-
mality, they will look to find out what is wrong. This mode is an
evident failure mode. If the operators have no clue whatsoever as
to the occurrence of the abnormality, the failure mode is hidden,
and receives the label “D” at this point.

2. Does this failure mode cause a safety problem?


The failure mode is then carried to Question 2 regarding possible
safety or environment issues. An answer of Yes to Question 2
picks up an A label on the failure mode, a rating which raises the
failure mode to the highest importance level in the LTA.

3. Does this failure mode result in a full or partial outage of the


plant?
Finally, Question 3 inquires as to whether the failure mode could
lead to a plant outage (or production downtime). An answer of Yes
here results in a B label; No, by default, results in a C label. A C
designates the failure mode as one of little functional signifi-
cance.
240 Chapter 8

Thus, every failure mode passed to Step 6 receives one of the follow-
ing labels or categories: A, B, C, D, or any combination of these. Any fail-
ure mode that contains an A in its label is a top priority item; a B is the
second and next significant priority item; a C is essentially a low priority
item that, in the very practical sense, is probably a non-issue in allocating
preventive maintenance resources. All C and D/C failure modes are good
candidates for RTF. Primary attention will be placed on the A and B labels,
which are addressed in Step 7.

Step 7: Selection of Maintenance Tasks The fourth RCM principle,


“select applicable and effective maintenance tasks for the high priority
failure modes,” is addressed in this step.
In Step 7, all team knowledge is applied to determine the most appli-
cable and cost effective tasks that will eliminate, mitigate, or warn us of
the failure modes and causes that we assigned to each component or piece
of equipment in Step 5. The team revisits those failure modes they initial-
ly believed did not impact the functioning of system, and re-evaluates
them. Finally, the team compares the new PM program to the old one, see-
ing where the program has been improved and optimized. Step 7 is com-
prised of three steps that are discussed in this section:

• Task selection
• Sanity check
• Task comparison

In task selection, the following questions are addressed:


• Is the age reliability relationship for this failure known? If yes,
• Are there any applicable Time directed (TD) tasks? If yes,
• Specify those tasks.
• Are there any Condition Directed (CD) tasks? If yes,
• Specify those tasks.
• Is there a “D” category of failure modes? If yes,
• Are there any applicable Failure Finding (FF) Tasks? If yes,
• Specify those tasks.
• Can any of these tasks be ineffective? If no,
• Finalize tasks above as TD, CD, and FF tasks.
• If any tasks may not be effective, then can design modifications
eliminate failure mode or effect? If yes,
• Request design modifications.
Maintenance Optimization 241

Task Selection is a key item to be discussed in this step. It is very


important that the team members put their past biases aside at this junc-
ture and develop a creative and free-flowing thought process to put forth
the best possible ideas for candidate PM tasks — even if some of their
suggestions may sound a bit off-the-wall at first blush. It might also be
useful to get the help of predictive maintenance specialists if they are not
part of the team.
A final aspect of the task selection process is to revisit the failure
modes that are designated as RTF candidates. This is part of the sanity
check, to insure all task selections are appropriate. We need to examine
other non-function related consequences that could cause us to reverse the
RTF decision for reasons such as high cost, regulatory difficulties and vio-
lations, the likelihood of secondary failure damage, warranty and insur-
ance factors, or hidden failure conditions. The team can elect to drop the
RTF decision in favor of a PM task if they believe that the potential con-
sequences of the failure mode are severe.
The last item in the RCM system analysis process before proceeding
to PM task implementation is Task Comparison. Now the team lays out
what they have recommended for an RCM-based program versus the cur-
rent PM program. This is the first time in the entire process that the team
will deliberately examine the current PM task structure in detail.
The difficulty in performing task comparison stems from the fact that
the RCM-based PM tasks were developed at the failure mode level of
analysis detail whereas the current PM tasks were identified at the com-
ponent level. Hence, analysts must use their experience and judgment to
fit the current PM tasks into a structure that is comparing PM programs at
the failure mode level. This can be somewhat difficult at times and may
require careful review.

Step 8: Task Packaging and Implementation Step 8, task packag-


ing and implementation, is a crucial step for realizing the benefits of RCM
analysis. Usually this step is very difficult to accomplish successfully. In
fact, the majority of RCM failures happen during this step and analysis
results are put aside on the shelf. However, if team members have been
selected from all critical areas and they have been participating diligently,
implementation will go smoothly and will be successful.
The final implementation action is to write task procedures that com-
municate analysis results to the actionable instructions to the operations
and maintenance teams including CBM/PdM technicians. If the work is
multi-disciplined (multi-craft skills), it may require writing separate
242 Chapter 8

instructions for each craft group, depending upon union contract require-
ments. However, the coordination between the craft should be part of each

PM Work Instruction Example

1. Work Instruction Title Compressor #xx — area XY


2. Task Instructions # PM XXXXX
3. Task Interval 6 Month / 1000 hours of operations
4. Priority Low, medium, high, or X, based on organi
zation’s priority scheme
5. Estimated Hours with Skill Mech. — 20, Elect — 10, Total = 30
6. Actual Hours Mech. — 18, Elect – 10, Labor – 4, Total = 34
(Inputted after the job has been completed)
7. Component Name and ID #
8. Contact # Planner #; Systems Engineer #
9. Special safety instructions Lock out / Tag Out details and
clearance permits
10. Special Tools requirements
11. Material Handling support
requirements
12. Task Objectives
13. Task detailed — steps
14. Spare / parts required
15. As found condition list
16. Work performed
17. Post maintenance test
measurement data
18. Other observations

Figure 8.8 Sample of PM Work Instruction

Notes: Could the effectiveness of this PM be improved? No / Yes, and how?


Maintenance Optimization 243

instruction. Nevertheless, it is always beneficial and effective to have


multi-skill crews handle multi-disciplined work. In most cases, these
instructions will be kept in the CMMS and will be issued per established
schedule or based on CBM data. Figure 8.8 shows a list of items that can
be part of good PM work instructions.
Step 9: Making the Program a Living One — Continuous
Improvements RCM execution is not a one-time event. It’s a journey.
RCM is a paradigm shift in how maintenance is perceived and executed.
An RCM-based maintenance program needs to be reviewed and updated
on a continuous basis. A living RCM program consists of:

• Validation of existing program — maintenance decisions made


are appropriate
• Reviewing current failure history and evaluate maintenance
tasks and their effectiveness
• Making adjustments in maintenance program if needed

A living RCM program assures continual improvement and cost-


effective operation and maintenance in the organization. We also need to
establish some effective metrics to know where the program stands.

Other RCM Processes


There are many derivatives of RCM such as RCM++, RCM cost,
RCM turbo, RCM backfit, RCM streamline, VRCM, Abbreviated, and
Experience-Based. All of these derivatives help perform RCM cost effec-
tively. Most of them take some shortcuts—cutting some steps, consider-
ing only a limited number of failure modes, or automating the process
using software to reduce the time taken to complete the analysis. In addi-
tion, RCM software programs are also available from JMS software,
Isograph, ReliaSoft, Relex, and others. These programs can help to reduce
the time taken to perform RCM analyses.

RCM Benefits
• Reliability. The primary goal of RCM is to improve asset relia-
bility and availability cost-effectively. This improvement comes
through constant reappraisal of the existing maintenance program
and improved communication between maintenance supervisors
and managers, operations personnel, maintenance mechanics,
planners, designers, and equipment manufacturers. This improved
communication creates a feedback loop from the maintenance
craft in the field all the way to the equipment manufacturers.
244 Chapter 8

• Cost. Due to the initial investment required to obtain the techno-


logical tools, training, and equipment condition baselines, a new
RCM program typically results in a short-term increase in main-
tenance costs. The increase is relatively short-lived. The cost of
reactive maintenance decreases as failures are prevented and pre-
ventive maintenance tasks are replaced by condition monitoring.
The net effect is a reduction of reactive maintenance and a reduc-
tion in total maintenance costs.
• Documentation. One of the key benefits of an RCM analysis is
understanding and documentation of operations and maintenance
key features, failures modes, basis of PM tasks, related drawings
and manuals, etc. This documentation can be good training mate-
rial for new O&M personnel.
• Equipment/Parts Replacement. Another benefit of RCM is that
it obtains the maximum use from the equipment or system. With
RCM, equipment replacement is based on equipment condition,
not on the calendar. This condition-based approach to mainte-
nance extends the life of the facility and its equipment.
• Efficiency/Productivity. Safety is the primary concern of RCM.
The second most important concern is cost effectiveness, which
takes into consideration the priority or mission criticality and then
matches a level of cost appropriate to that priority. The flexibility
of the RCM approach to maintenance ensures that the proper type
of maintenance is performed when it is needed. Maintenance that
is not cost effective is identified and not performed.

In summary, the multi-faceted RCM approach promotes the most effi-


cient use of resources. The equipment is maintained as required by its
characteristics and the consequences of its failures.

Impact of RCM on a Facility’s Life Cycle


RCM must be a consideration throughout the life cycle of a facility if
it is to achieve maximum effectiveness. The four major phases of a facil-
ity’s life cycle are:

1. Planning (Concept)
2. Design and Build
3. Operations and Maintenance
4. Disposal
Maintenance Optimization 245

It has been documented in many studies that about 80% or more of a


facility’s life cycle cost is fixed during the planning, design and build
phases. The subsequent phases fix the remaining 20% or so of the life-
cycle cost. Thus, the decision to institute RCM at a facility, including con-
dition monitoring, will have a major impact on the life-cycle cost of the
facility. This decision is best made during the planning and design phas-
es. As RCM decisions are made later in the life cycle, it becomes more
difficult to achieve the maximum possible benefit from the RCM pro-
gram.
Although relatively small impact on the overall life-cycle cost, a bal-
anced RCM program is still capable of achieving savings of 10–30% in a
facility’s annual maintenance budget during the O&M phase.

8.5 Maintenance Strategy — CBM

CBM / PdM Technologies


The start of Condition-Based Maintenance (CBM) — also called
Predictive Maintenance or PdM — may have been when a mechanic first
put an ear to the handle of a screwdriver, touched the other end to a
machine, and pronounced that it sounded like a bearing was going bad.
We have come a long way since then with a variety of technologies for
analyzing what’s happening inside the asset. However, the need for a
knowledgeable, experienced person to use the technology has not
changed. Today, as in the beginning, successful predictive maintenance is
a combination of man and technology.
Recent advances in technology have made CBM a reality — the ready
availability of inexpensive computing power to gather, store, and analyze
the data that makes CBM possible. By some counts, there are more than
30 technologies being used for condition-based maintenance. Others
might argue that many of these are simply variations of each other. This
section discusses some of the most-used CBM/PdM technologies.
Any condition-based maintenance program can be characterized by a
combination of three phases:

• Surveillance — monitoring machinery condition to detect


incipient problems
• Diagnosis / Prognosis — isolating the cause of the problem and
developing a corrective action plan based on its condition and
remaining life.
• Remedy — performing corrective action
246 Chapter 8

Consistent, accurate data gathering is essential to all three phases.


Analysis of data is where the knowledge and experience of mainte-
nance personnel becomes most important. It normally requires extensive
training not only in the analysis techniques, but also in the use of the par-
ticular hardware and software employed.

Condition Monitoring and Data Collection


Condition monitoring uses primarily non-intrusive testing tech-
niques, visual inspections, and performance data to assess machinery con-
dition. It replaces arbitrarily timed maintenance tasks with maintenance
scheduled only when warranted by asset condition. Continuing analysis of
the asset condition allows planning and scheduling of maintenance or
repairs in advance of catastrophic or functional failure.

Data Collection
Asset condition data is collected basically in two ways:

1. Spot readings — route based with portable instruments


2. Permanently installed data acquisition equipment for continuous
online data collection

Generally, taking spot readings provides sufficient information for


making informed decisions regarding maintenance of assets. The degrada-
tion of facility assets is usually not so rapid as to require the “up to the
second” reporting that a permanent data acquisition system produces. The
maintenance technician or a CMMS can usually keep a log of these spot
readings and develop trends from these logs.
Permanent condition monitoring equipment is more expensive to
install, and the databases created do cost money to analyze and maintain.
Typically, permanent data collection systems are installed only on critical
and expensive assets and systems used in production processes. If they go
down, it can cost the facility “money by the minute” when it is not oper-
ating.
A variety of technologies are available to assess the condition of sys-
tems and equipment, and to determine the most effective time for sched-
uled maintenance. Some of the key technologies covering the basic theo-
ry of how the technology operates, the purpose of applying the technolo-
gy, and acceptable applications are discussed in this section.
The data collected is used in one of the following ways to determine
the condition of the asset and to identify the precursors of a failure:
Maintenance Optimization 247

• Trend Analysis. This method reviews data to see if an asset is on


an obvious and immediate “downward slide” toward failure. It
includes recognizing the changes in data as compared to earlier
data or baseline data on similar assets.
• Pattern Recognition. This method reviews data to recognize any
causal relationships between certain events and asset failure. For
example, we might notice that after asset X is used in a certain
production run, component Y fails due to stresses unique to that
run. The method identifies deviations from established patterns.
• Correlation Analysis. This approach compares data from multi-
ple sources, related technologies, or different analysts.
• Tests against Limits and Ranges. These tests set alarm limits
and check if they are exceeded.
• Statistical Process Analysis. This analysis uses statistical tech-
niques to identify deviations from the norm.

If published failure data on a certain asset or component exists, then


we can compare failure data collected onsite with the published data to
verify or disprove the data.
Several CBM technologies are available to assess the condition of an
asset or system. In some instances, several technologies are used together
to provide a more accurate picture of the asset condition. For example, to
obtain the total picture of a cooling water system, a CBM effort may need
to collect the following data:

Flow Rates. Water flow is measured using precision, non-intrusive flow


detectors.
Temperature. Differential temperature is measured to determine heat
transfer coefficients and to indicate possible tube fouling of heat exchang-
ers.
Pressure. Differential pressures across the pump and piping are meas-
ured to determine pump performance and to determine the condition of
the tubes.
Electrical. Online and Offline testing is used to assess the condition of
the motor.
Vibration. Vibration monitoring is used to assess the condition of rotat-
ing equipment, specifically compressors, pumps and motors. Additionally,
structural problems can be identified through resonance and model test-
ing.
248 Chapter 8

Ultrasonic Testing. Pipe wall thickness is measured to determine erosion


and corrosion degradation and also leaking pipes.
Airborne Ultrasonic Testing. Airborne ultrasonic indicates air leaking
from control system piping and pumps.
Lubricant Analysis. Oil condition and wear particle analysis are used to
identify problems with the lubricant, and to correlate those problems with
vibration when wear particle concentrations exceed pre-established limits.
Infrared Thermography. Thermography scans motor control system
and electrical distribution junction boxes for high-temperature conditions.
High temperature is indicative of loose connections, shorts, or failing con-
ductor insulation. Piping insulation is checked for porosities. High tem-
peratures are indicative of failed/failing areas in the pipe insulation.

Vibration Analysis
Vibration monitoring might be considered the “grandfather” of condi-
tion / predictive maintenance, and it provides the foundation for most
facilities’ CBM programs.
Vibration usually indicates trouble in the machine. Machine and struc-
tures vibrate in response to one or more pulsating forces that may be due
to imbalance, misalignment, etc. The magnitude of vibration is dependent
on the force and properties of the system, both of which may depend on
speed.
There are four fundamental characteristics of vibration: frequency,
period, amplitude, and phase. Frequency is the number of cycles per unit
time and is expressed in the number of cycles per minute (CPM) or cycles
per second (Hz). The period is the time required to complete one cycle of
vibration, the reciprocal of frequency. The amplitude is the maximum
value of vibration at a given location of the machine. Phase is the time
relationship between vibrations of the same frequency and is measured in
degrees.
The three key measures used to evaluate the magnitude of vibra-
tions are:

• Displacement
• Velocity
• Acceleration

The units and descriptions of these measures are shown in Figure 8.9.
Displacement measurement is dominant at low frequency and is
caused by stresses in flexible members of the machine. It is typically
Maintenance Optimization 249

Figure 8.9 Vibration Measures

expressed in mils peak-to-peak. Displacement is a good measure for low


frequency vibration, usually less than 20 Hz. Velocity is the time-rate
change of displacement. It is dependent upon both displacement and fre-
quency. It is related to fatigue characteristics of the machine. Greater dis-
placement and frequency of vibration relates directly to a greater severity
of machine vibration at the measured location. Velocity is generally used
to evaluate machine condition in the frequency range of 10–1,000 Hz. The
acceleration is the dominant measure at higher frequencies that exceed
1,000 Hz. Acceleration is proportional to the force on machine compo-
nents such as gears and couplings. Both velocity and acceleration are
expressed in PEAK.

Velocity and acceleration are calculated by the following formulas:


Velocity (V) = 2 π f d
2
Acceleration = 2 π f V = (2 π f) d
where
f = frequency in cycles per second
d = peak displacement

Monitoring the vibration of facility machinery can provide direct cor-


relation between the mechanical condition and recorded vibration data of
each machine in the plant. This data can identify specific degrading
machine components or the failure mode of plant assets before serious
damage occurs.
250 Chapter 8

Vibration monitoring and trending works on the premise that every


machine has a naturally correct vibration signature. This signature can be
measured when the machine is in good working order, and subsequent
measurements can be compared with what is considered the norm. As the
component wears or ages, the vibration spectra change. Analyzing the
changes identifies components that require further monitoring, repair, or
replacement.
With a few exceptions, mechanical troubles in a rotating machine
cause vibration. Common problems that produce vibration are:

• Imbalance of rotating parts


• Misalignment of couplings and bearings
• Bent shafts
• Worn, eccentric, or damaged parts
• Bad drive belts and chains
• Damaged / bad bearings
• Looseness
• Rubbing
• Aerodynamic and other forces

Under conditions of dynamic stress, displacement alone may be a bet-


ter indication of severity, especially when the asset components exhibit
the property of brittleness — the tendency to break or snap when stressed
beyond a given limit. Consider a slowly rotating machine that operates at
60 RPM, and that exhibits vibration of 20 mils peak-to-peak displacement
caused by rotor imbalance. In terms of vibration velocity, 20 mils at 60
CPM (1 Hz) is only 0.0628 in/sec [V = 2(3.14)(1)(0.02/2) = 0.0628]. This
level would be considered good for general machinery and little cause for
immediate concern. However, keep in mind that the bearing of this
machine is being deflected 20 mils. Under these conditions, fatigue may
occur due to stress (resulting from the displacement) rather than due to
fatigue (caused by the velocity of displacement).
Generally, the most useful presentation of vibration data is a graph
showing vibration velocity (expressed in inches/second) on the vertical
axis and frequency on the horizontal axis. By analyzing this data, a trained
vibration technician can ascertain what kinds of problems exist. A trained
technician can learn to read vibration signatures and to interpret what the
different peaks in the different frequency ranges indicate. For example,
when analyzing a 3580 RPM pump motor, a peak at 3580 RPM generally
indicates some kind of mass imbalance. A peak at 7160 RPM (two times
Maintenance Optimization 251

the rotational frequency) generally indicates a bent shaft.


All rotating machinery will exhibit a certain degree of vibration. The
question then becomes “How much is too much?” There are no realistic
figures for selecting a vibration limit, which, if exceeded, will result in
immediate machinery failure. The events surrounding the development of
a mechanical failure are too complex to set reliable limits. However, there
are some general guidelines that have been developed over the years that
can serve as general indication of the condition of a piece of machinery.
Some of the vibration equipment manufacturing and supplier companies
can provide these guidelines and lessons learned.
Figure 8.10 lists the forcing frequencies associated with machines as
a guideline for possible fault sources.

Figure 8.10 Forcing Frequencies Associated with Machines


252 Chapter 8

Types of vibration analysis

• Broadband trending provides a broadband or overall value that


represents the total vibration of the machine at the specific meas-
urement point where the data was acquired. It does not provide
information on the individual frequency components or machine
dynamics that created the measured value. Collected data is com-
pared either to a baseline reading taken when the machine was
new (or sometimes data from a new, duplicate machine) or to
vibration severity charts to determine the relative condition of the
machine.
• Narrowband trending monitors the total energy for a specific
bandwidth of vibration frequencies and is thus more specific.
Narrowband analysis utilizes frequencies that represent specific
machine components or failure modes. A narrowband vibration
analysis can provide several weeks or months of warning of
impending failure. Different vibration frequencies predict differ-
ent potential failures.
• Signature analysis provides visual representation of each fre-
quency component generated by a machine. With appropriate
training and experience, plant personnel can use vibration signa-
tures to determine the specific maintenance required on the
machine being studied.

When setting up a vibration monitoring program that uses handheld


vibration instrumentation, it is necessary to ensure that the measurements
are taken consistently. A slight variation in the location where a measure-
ment is taken on machinery can significantly alter its accuracy. This
becomes an issue especially when several technicians take measurements
at different times on the same machinery.
If applied by a trained professional, vibration monitoring can yield
information regarding: wear, imbalance, misalignment, mechanical loose-
ness, bearing damage, belt flaws, bent shaft, sheave and pulley flaws, gear
damage, flow turbulence, cavitations, structural resonance, and material
fatigue.
Detection Interval/Amount of Data Collected The frequency of
data collection depends on machine type and failure category. Typically,
spot readings of facility assets with handheld vibration monitoring equip-
ment once per month, per quarter, or per 500 operating hours usually pro-
vides sufficient warning of impending problems. Facility rotating equip-
Maintenance Optimization 253

ment, e.g., fans and pumps, usually do not deteriorate fast enough to war-
rant continual real time data collection. However, critical and expensive
assets may warrant having real-time, continuous data collection system.
Spectrum Analysis and Waveform Analysis Spectrum analysis is the
most commonly-employed analysis method for machinery diagnostics. In
this type of analysis, the vibration technician focuses on analyzing specif-
ic “slices” of the vibration data taken over a certain range of CPM.
Spectrum analysis can be used to identify the majority of all rotating equip-
ment failures (due to mechanical degradation) before failure. Waveform
analysis, or time domain analysis, is another extremely valuable analytical
tool. Although not used as regularly as spectrum analysis, the waveform
often helps the analyst more correctly diagnose the problem.
Shock Pulse Analysis This type of analysis is used to detect impacts
caused by contact between the surfaces of the ball or roller and the race-
way during rotation of anti-friction bearings. The magnitude of these puls-
es depends on the surface condition and the angular velocity of the bear-
ing (RPM and diameter). Spike energy is similar in theory to shock pulse.
Alignment Misalignment of shafted equipment will not only cause
equipment malfunctions or breakdowns; it may also be an indicator of
other problems. Checking and adjusting alignment used to be a very slow
procedure. The advent of laser alignment systems has reduced labor time
by more than half and increased accuracy significantly.
Laser alignment is a natural compliment to vibration analysis.
Properly aligning shafts eliminates one of the major causes of vibration in
rotating machines and also drastically extends bearing life. For the mini-
mal amount of work involved, the payback is great.
Vibration Equipment For permanent data collection, vibration
analysis systems include microprocessor-based data collectors, vibration
transducers, equipment-mounted sound discs, and a host personal com-
puter with software for analyzing trends, establishing alert and alarm
points, and assisting in diagnostics. Portable handheld data collectors con-
sist of a data collection device about the size of a palm-top computer and
a magnetized sensing device.
The effectiveness of vibration monitoring depends on sensor mount-
ing, signal resolution, machine complexity, data collection techniques,
and the ability of the analyst. This last factor, the ability of the analyst, is
probably the most important aspect of establishing an effective vibration
monitoring program. The analyst must be someone who possesses a thor-
ough understanding of vibration theory and the extensive field experience
necessary to make the correct diagnosis of the acquired vibration data.
254 Chapter 8

Infrared Thermography
As one of the most versatile condition-based maintenance technolo-
gies available, infrared thermography is used to study everything from
individual components of assets to plant systems, roofs, and even entire
buildings.
Infrared inspections can be qualitative or quantitative. Qualitative
inspection concerns relative differences, hot and cold spots, and devia-
tions from normal or expected temperatures. Quantitative inspection con-
cerns accurate measurement of the temperature of the target. One must be
careful not to put too much emphasis on the quantitative side of infrared
because temperature-based sensors are better for accurate temperature
measurements.
Infrared instruments include an optical system to collect radiant ener-
gy from the object and focus it, a detector to convert the focused energy
pattern to an electrical signal, and an electronic system to amplify the
detector output signal and process it into a form that can be displayed.
Most instruments include the ability to produce an image that can be dis-
played and recorded. These thermographs, as the images are called, can be
interpreted directly by the eye or analyzed by computer to produce addi-
tional detailed information. Mid-wave range instruments detect infrared in
the 2–5 micron range; long-wave range instruments detect the 8–14
micron range. High-end systems can isolate readings for separate points,
calculate average readings for a defined area, produce temperature traces
along a line, and make isothermal images showing thermal contours.
It is essential that infrared studies be conducted by technicians who
are trained in the operation of the equipment and interpretation of the
imagery. Variables that can destroy the accuracy and repeatability of ther-
mal data, for example, must be compensated for each time data is
acquired.
Infrared Thermography (IRT) cameras are non-contact, line-of-sight,
thermal measurement and imaging systems. Because IRT is a non-contact
technique, it is especially attractive for identifying hot and cold spots in
energized electrical equipment, large surface areas such as boilers and
building roofs, and other areas where “stand-off” temperature measure-
ment is necessary. Instruments that perform this function detect electro-
magnetic energy in the short wave (3–5 microns) and long wave (8–15
microns) bands of the electromagnetic spectrum.
Because of the varied inspections (electrical, mechanical, and struc-
tural) encountered, the short wave instrument is the best choice for facil-
ity inspections. However, the short wave instrument is more sensitive than
Maintenance Optimization 255

long wave to solar reflections. Sunlight reflected from shiny surfaces may
make those surfaces appear to be “hotter” than the adjacent surfaces when
they really are not. IRT instruments–cameras are portable, usually sensi-
tive to within 0.20oC over a range of temperatures from –100 to +3000oC,
and accurate within +/–3 percent. In addition, the instrument can store
images for later analysis.
IRT inspections attempt to accurately measure the temperature of the
item of interest. To perform an inspection requires knowledge and under-
standing of the relationship of temperature and radiant power, reflection,
emittance, and environmental factors, as well as the limitations of the
detection instrument. This knowledge must be applied in a methodical
manner to control the imaging system properly and to obtain accurate
temperature measurements.
The qualitative inspections are significantly less time-consuming
because the thermographer is not concerned with highly-accurate temper-
ature measurement. In qualitative inspections, the thermographer obtains
accurate temperature differences (ΔT) between like components. For
example, a typical motor control center will supply three-phase power,
through a circuit breaker and controller, to a motor. Ideally, current flow
through the three-phase circuit should be uniform so the components
within the circuit should have similar temperatures. Any uneven heating,
perhaps due to dirty or loose connections, would quickly be identified
with the IRT imaging system.
IRT can be used very effectively to identify degrading conditions in
facilities’ electrical systems such as transformers, motor control centers,
switchgear, substations, switch yards, or power lines. In mechanical sys-
tems, IRT can identify blocked flow conditions in heat exchangers, con-
densers, transformer cooling radiators, and pipes. IRT can also be used to
verify fluid level in large containers such as fuel storage tanks. IRT can
identify insulation system degradation in building walls and roofs, as well
as refractory in boilers and furnaces. Temperature monitoring, infrared
thermography in particular, is a reliable technique for finding the mois-
ture-induced temperature effects that characterize roof leaks, and for
determining the thermal efficiency of heat exchangers, boilers, building
envelopes, etc.
Deep-probe temperature analysis can detect buried pipe energy loss
and leakage by examining the temperature of the surrounding soil. This
technique can be used to quantify ground energy losses of pipes. IRT can
also be used as a damage control tool to locate mishaps such as fires and
leaks.
256 Chapter 8

Thermography is limited to line of sight. Errors can be introduced due


to color of material, material geometry, and by environmental factors such
as solar heating and wind effects. Emissivity is a key concern which can
introduce 5–20% error in measurements. Shiny or highly polished sur-
faces can be very tricky to measure; even dull smooth metal surfaces may
not be equally emissive in all directions. Be especially careful where sur-
faces are highly curved. If the emissivity is very low, some polished sur-
faces have an emissivity of 0.2 or less, and then an accurate reading is
unlikely to be possible. Many low cost Infrared cameras have a fixed
emissivity (usually around 0.95), which is a reasonable value in normal
circumstances.
Because IRT images are complex and difficult to measure and ana-
lyze, training is required to obtain and interpret accurate and repeatable
thermal data and to interpret the data. With adequate training and certifi-
cation, electrical/mechanical technicians and engineers can perform this
technique.
Training is available through infrared imaging system manufacturers
and vendors. Also, the American Society of Non-destructive Testing
(ASNT) has established guidelines for nondestructive testing (NDT) ther-
mographer certification. These guidelines, intended for use in nondestruc-
tive testing, may be used as guidelines for thermography in CBM if appro-
priately applied.

Ultrasonic Testing
Ultrasonic testing is extremely useful in the diagnosis of mechanical
and electrical problems. Testing instruments are usually portable hand
held devices. Their electronic circuitry converts a narrow band of ultra-
sound (between 20 and 100 kHz) into the audible range so that a user can
recognize the qualitative sounds of operating equipment through head-
phones. Intensity of signal strength is also displayed on the instrument.
Ultrasonic instruments–scanners are most often used to detect gas, liquid,
or vacuum leaks.
Ultrasonic detectors are somewhat limited in their use. For example,
they may help identify the presence of suspicious vibrations within a
machine, but they are generally not sufficient for isolating the sources or
causes of those vibrations.
On the plus side, ultrasonic monitoring is easy, it requires minimal
training, and the instruments are inexpensive. Airborne ultrasonic devices
are highly sensitive listening “guns” (similar in size to the radar speed
guns used by police at speed traps). They provide a convenient, non-intru-
Maintenance Optimization 257

sive means of assessing asset condition. Airborne ultrasonic monitoring is


especially easy and useful in testing remote electrical equipment, as well
as shielded electrical equipment, e.g., connections inside switchgear and
panels. In the case of high voltage insulator failures, airborne ultrasonic
devices can often detect faults earlier than infrared thermography can.
Except for severe cases where a current path to ground was established,
infrared thermography would not detect high-voltage insulation failures
because the corona or tracking typically produces little or no heat.
Airborne ultrasonic devices can also detect the noise caused by loose con-
nections as they vibrate inside of panels.
Airborne ultrasonic devices operate in the frequency range from 20 to
100 kHz and translate the high frequency signal to a signal within the
audible human range. This allows the operator to hear changes in noise
levels associated with air leaks, corona discharges, and other high fre-
quency events. For example, a maintenance technician could use ultrason-
ic equipment to “hear” a bearing ring and surrounding housing resonating
at the resonant frequency. Once detected, a maintenance technician could
then proceed to find the cause of the problem; insufficient lubrication or
minor bearing material defects would be the likely cause of this malfunc-
tion.
Some of the most common applications of ultrasound detection are:

• Leak detection in pressure and vacuum systems (e.g., boiler, heat


exchanger, condensers, chillers, vacuum furnaces, specialty gas
systems)
• Bearing inspection
• Steam trap inspection
• Pump cavitations
• Detection of corona in electrical switchgear
• Valve analysis
• Integrity of seals and gaskets in tanks and pipe systems

All operating equipment and most leakage problems produce a broad


range of sound. The high frequency ultrasonic components of these
sounds are extremely short wave in nature, and a short wave signal tends
to be fairly directional. It is therefore easy to isolate these signals from
background noises and detect their exact location. In addition, as subtle
changes begin to occur in mechanical equipment, the nature of ultrasound
allows these potential warning signals to be detected early — before actu-
al failure.
258 Chapter 8

Leak Detection in Mechanical Systems Ultrasound is a very versa-


tile technique that detects the sound of a leak. When a fluid (liquid or gas)
leaks, it moves from the high pressure side through the leak site to the low
pressure side, where it expands rapidly and produces a turbulent flow.
This turbulence has strong ultrasonic components. The intensity of the
ultrasonic signal falls off rapidly from the source, allowing the exact spot
of a leak to be located.
Generally gas leak detection is easy to apply. An area should be
scanned while listening for a distinct rushing sound. With continued sen-
sitivity adjustments, the leak area is scanned until the loudest point is
heard.
Some instruments include a rubber focusing probe that narrows the
area of reception so that a small emission can be pinpointed. The rubber
focusing probe is also an excellent tool for confirming the location of a
leak. This is done by pressing it against the surface of the suspected area
to determine if the sound of the leak remains consistent. If it decreases in
volume, the leak is elsewhere.
Vacuum leaks may be located in the same manner; the only difference
being that the turbulence will occur within the vacuum chamber. For this
reason, the intensity of the sound will be less than that of a pressurized
leak. Though it is most effective with low-mid to gross leaks, the ease of
ultrasound detection makes it useful for most vacuum leak problems.
Liquid leaks are usually determined through valves and steam traps,
although some successes have been reported in locating water leaks from
pressurized pipes buried underground. A product can be checked for leak-
age if it produces some turbulence as it leaks.
Valves are usually checked for leakage with the contact probe on the
downstream side. This is accomplished by first touching the upstream side
and adjusting the sensitivity to read about 50% of scale. The downstream
side is then touched and the sound intensity is compared. If the signal is
lower than upstream, the valve is considered closed; if it is louder than
upstream and is accompanied by a typical rushing sound, it is considered
to be leaking.
Steam traps are also inspected easily with ultrasonic translators.
During the steam trap operation and while observing the meter, trap con-
ditions can be interpreted. The speed and simplicity of this type of test can
allow every trap in a plant to be routinely inspected.
Leaking tubes in heat exchangers and condensers as well as boiler
casing leaks are detectable with ultrasonic translators. In most power
plants, the problem of condenser in-leakage is a major concern.
Condenser fittings are often routinely inspected using the leak detection
Maintenance Optimization 259

method previously described. If a leak is suspected in a condenser tube


bundle, it is possible to locate the leak by putting a condenser at partial
load and opening up a water box of a suspected tube bundle. After the tube
sheet is cleared of debris, the tube sheet is scanned.
Bearing Inspections Ultrasonic inspection and monitoring of bear-
ings is a reliable method for detecting incipient bearing failure. The ultra-
sonic warning appears prior to a rise in temperature or an increase in driv-
ing torque. Ultrasonic inspection of a bearing is useful in recognizing the
beginning of fatigue failure, brinnelling of bearing surfaces, or flooding of
(or lack) of lubricant. In ball bearings, as the metal in the raceway, roller,
or bearing balls begins to fatigue, a subtle deformation begins to occur.
This deforming of the metal will produce an increase in the emission of
ultrasonic sound waves.
It is observed that as the lubricant film reduces, the sound level will
increase. A rise of about 8 dB over baseline accompanied by a uniform
rushing sound will indicate lack of lubrication. When lubricating, add just
enough to return the reading to baseline. Some lubricants will need time
to run in order to cover the bearing surfaces uniformly.
One of the most frequent causes for bearing failure is over-lubrica-
tion. The excess stress of lubricant often breaks bearing seals or causes a
buildup of heat, which can create stress and deformity. To avoid over-
lubrication, do not lubricate if the baseline reading and baseline sound
quality is maintained. When lubricating, use just enough lubricant to bring
the ultrasonic reading to baseline. Recently new grease guns have become
available in the market with built-in ultrasonic systems that can provide
the appropriate amount of grease. These are very practical and easy to use
devices.

Detection in Electrical Systems


Three types of high voltage electrical problems detectable with ultra-
sound are:
• Arcing. An arc occurs when electricity flows through space.
Lightning is a good example.
• Corona. When voltage on an electrical conductor, such as an
antenna or high voltage transmission line, exceeds threshold
value, the air around it begins to ionize to form a blue or purple
glow.
• Tracking. Often referred to as “baby arcing,” electricity follows
the path of damaged insulation, using surrounding dirt, debris, and
moisture as the conductive medium.
260 Chapter 8

Theoretically, ultrasonic detection can be used in low, medium, and


high voltage systems; however, applications have only normally been
used with medium and high voltage systems. When electricity escapes in
high voltage lines or when it jumps across a gap in an electrical connec-
tion, it disturbs the air molecules around it and generates ultrasound.
Often this sound will be perceived as a crackling or frying sound. In other
situations, it will be heard as a buzzing sound. In substations and distribu-
tion systems, components such as insulators, transformers, cable,
switchgear, bus bars, relays, contractors, junction boxes, and bushings
may be tested.
Ultrasonic testing is often used for evaluation at voltages exceeding
2,000 volts, especially in enclosed switchgear. This is especially useful in
identifying corona problems. In enclosed switchgear, the frequency of
detection of corona greatly exceeds the frequency of serious faults identi-
fied by infrared. It is recommended that both tests be used with enclosed
switchgear. When testing electric equipment, be sure to follow safety pro-
cedures.
The method for detecting electric arc and corona leakage is similar to
that discussed in mechanical leak detection. Instead of listening for a rush-
ing sound, a user will listen for a crackling or buzzing sound. Determining
whether a problem exists is relatively simple. By comparing sound quali-
ty and sound levels among similar equipment, the problem will become
easy to identify, even though the sound itself will differ somewhat as it
resonates through various types and sizes of equipment.
On lower voltage systems, a quick scan of bus bars will often pick up
a loose connection. Checking junction boxes can reveal arcing. As with
leak detection, the closer one gets to the leak site, the louder the signal. If
power lines are to be inspected and the signal does not appear to be
intense enough to be detectable from the ground, a parabolic reflector
which is an ultrasonic waveform concentrator, will increase the detection
distance of the system and provide pinpoint detection.

Lubricant (Oil) and Wear Particle Analysis


The objective of oil analysis is to determine:

• An asset’s mechanical wear condition


• Lubricant condition
• If the lubricant has become contaminated

A wide variety of tests can provide information regarding one or more


Maintenance Optimization 261

of these areas. The three areas are not unrelated. Changes in lubricant con-
dition and contamination, if not corrected, will lead to machine wear.
Lubricant Condition Bad lubricating oil is either discarded or recon-
ditioned through filtering or by replacing additives. Analyzing the oil to
determine the lubricant condition is, therefore, driven by costs. Small
machines with small oil reservoirs have the oil changed on an operating
time basis. An automobile is the most common example of time-based
lubricating oil maintenance. In this example, the costs to replace the auto-
mobile oil change (which includes the replacement oil, labor to change the
oil, and disposal costs) are lower than the cost to analyze the oil, e.g., the
cost of sample materials, labor to collect the sample, and the analysis. In
the case of automobile oil, time-based replacement is cheaper than analy-
sis due to competition and the economies of scale that have been created
to meet the consumer need for replacing automobile oil.
In an industrial set-up, lubricating oil can become contaminated due
to the machine’s operating environment, improper filling procedures, or
through the mixing of different lubricants in the same machine. If a
machine is “topped off” with oil frequently, we should periodically send
the oil out for analysis to check the machine for any serious problems.
The full benefit of oil analysis can be achieved only by taking fre-
quent samples and trending the data for each asset in the program. The
length of the sampling intervals varies with different types of equipment
and operating conditions. Based on the results of the analyses, lubricants
can be changed or upgraded to meet the specific operating requirements.
It cannot be overemphasized that the sampling technique is critical to
meaningful oil analysis. Sampling locations must be carefully selected to
provide a representative sample and sampling conditions should be uni-
form so that accurate comparisons can be made.
Standard Analytical Test Types Lubricating oil and hydraulic fluid
analysis should proceed from simple, subjective techniques such as visu-
al and odor examination through more sophisticated techniques. The more
sophisticated tests should be performed when conditions indicate the need
for additional information and based on asset criticality.
Visual and Odor Simple inspections can be performed weekly by the
equipment operator to look at and smell the lubricating oil. A visual
inspection looks for changes in color, haziness or cloudiness, and parti-
cles. This test is very subjective, but can be an indicator of recent water or
dirt contamination and advancing oxidation. A small sample of fresh
lubricating oil in a sealed, clear bottle can be kept on hand for visual com-
parison. A burned smell may indicate oxidation of the oil. Other odors
262 Chapter 8

could indicate contamination. Odor is more subjective than the visual


inspection because people’s sensitivity to smell varies, and there is no
effective way to compare the odor between samples. The operator must
also be careful not to introduce dirt into the system when taking a sample.
Viscosity is one of the most important properties of lubricating oil; it
is often referred to as the structural strength of liquid. The analysis con-
sists of comparing a sample of oil from an asset to a sample of unused oil
to determine if thinning or thickening of the oil has occurred during use.
Viscosity is critical to oil film control and is a key indicator of machine
oil condition.
Viscosity is a measure of oil’s resistance to flow at a specified temper-
ature. A change (increase or decrease) in viscosity over time indicates
changes in the lubricant condition, or it may indicate lubricant contamina-
tion. Viscosity can be tested using portable equipment, or it can be tested
more accurately in a laboratory using the ASTM D445 standard. Viscosity
is measured in centistoke (cSt) at 40°C, and minimum and maximum val-
ues are identified by the ISO grade.
Water (Moisture) Test Water is generally referred to as a chemical
contaminant when suspended in oils. Its effects in bearings, gearing, and
hydraulic components can be very destructive. Like particles, control
must be established to minimize water accumulation to the oil and
machines.
Water in lubricating oil and hydraulic fluid contributes to the corro-
sion and formation of acids. Small amounts of water (less than 0.1 per-
cent) can be dissolved in oil and can be detected using the crackle test or
infrared spectroscopy (minimum detectable is approximately 500 ppm),
the ASTM D95 distillation method (minimum 100 ppm). If greater than
0.1 percent water is suspended or emulsified in the oil, the oil will appear
cloudy or hazy. Free water in oil collects in the bottom of oil reservoirs
and can be found by draining them from the bottom of the reservoir.
Using a titration process with a Karl Fischer method, low levels of
water can be detected and quantified. The volumetric titration test uses
ASTM D1744, and Coulometric titration uses ASTM D4928. This test is
useful when accepting new oil.
Wear Particle Count High particle counts indicate that machinery
may be wearing abnormally or that failures could be caused by blocked
orifices. Particle count tests are especially important in hydraulic systems.
The wear particle test emphasizes the detection and analysis of cur-
rent machine anomalies — the symptoms of failure. The oil serves as the
messenger of information on the health of the machine. When a machine
Maintenance Optimization 263

component is experiencing some level of failure such as rubbing, it will


shed particles in the oil. The presence of abnormal level of wear particles,
their size, shape, color, orientation, and elements define the cause, source,
and severity of the condition.
Total Acid Number (TAN) Total acid number (TAN) is a measure of
the amount of acid or acid-like materials in oil. It is an indicator of the
lubricating oil condition and is monitored relative to the TAN of new oil.
In some systems, the TAN will also be used to indicate acid contamina-
tion. TAN is measured in milligrams of potassium hydroxide (KOH) per
gram of oil (mg KOH/g). KOH is used in a titration process and the end
point is indicated by color change (ASTM D974) or electrical conductiv-
ity change (ASTM D664).
Total Base Number (TBN) Total base number (TBN) indicates oil’s
ability to neutralize acidity. Low TBN is often an indicator that the wrong
oil is being used for the application, intervals between oil changes are too
long, oil has been overheated, or a high-sulfur fuel is being used.
Spectrometric Metals Analysis Also known as emission spec-
troscopy, this test examines the light (spectrum) emitted from the sample
during testing, and identifies approximately 21 metals. Metals are catego-
rized as wear, contaminate, or additive metals. The procedure identifies
both soluble metal and metal particles.
Infrared Spectroscopy This test is also known as infrared analysis,
infrared absorption spectroscopy or spectrophotometry, and Fourier
Transform Infrared (FTIR) spectroscopy. The technique examines the
infrared wavelength that is absorbed by the oil sample. The test is used to
identify nonmetallic contamination and lubricant conditions (e.g., oxida-
tion, anti-oxidant, other additive depletion). Connecting a computer
expert system with known oil spectrums can produce highly accurate
diagnosis of small changes in the oil condition.
Analytical Ferrography Analytical ferrography is often initiated
based on changes in Direct Reading (DR) indicating increases in metal or
particle counts. DR Ferrograph quantitatively measures the concentration
of ferrous wear particles in lubricating or hydraulic oil.
Analytical ferrography is qualitative and requires visual examination
and identification of wear particles. Properties and features of the wear
debris are inventoried and categorized. This includes size, shape, texture,
color, light effect, density, surface oxide, etc.
This analysis is sometimes performed on a regular basis on expensive
or critical machines. The test process is labor intensive and involves the
264 Chapter 9

preparation of a sample and examination under magnification. Results


vary with the analyst’s capability, but the procedure can provide detailed
information regarding wear: e.g., wear type (rubbing, sliding, cutting),
color, particle types (oxide, corrosive, crystalline), and other nonferrous
particles. This detailed information can be critical in finding the root cause
of wear problems.
Foaming Some oils may have anti-foam agents added to improve the
lubrication capability in specific applications such as gear boxes or mix-
ers. ASTM test D892 can be used to test the oil’s foam characteristics. The
test blows air through a sample of the oil and measures the foam volume.
Rust Prevention Some systems are susceptible to water contamina-
tion due to equipment location or the system operating environment. In
those cases, the lubricating oil or hydraulic fluid may be fortified with an
inhibitor to prevent rust. The effectiveness of rust prevention can be test-
ed using ASTM D665 (or ASTM D3603).
Rotating Bomb Oxidation Test (RBOT) Also known as the Rotary
Bomb Oxidation Test, ASTM D 2272 is used to estimate oxidation stabil-
ity and the remaining useful life of oil. The test simulates aging, identify-
ing when rapid oxidation takes place, and indicating that anti-oxidants
have been depleted. The test is not a one-time test; it must be performed
over time, starting with a baseline test of the new oil. Subsequent tests are
necessary to develop the trend line. Because of the high cost and the mul-
tiple tests required, this test is usually only performed on large volume
reservoirs or expensive oil.
Sampling and Frequency Oil samples must be collected safely and
in a manner that will not introduce dirt and other contaminates into the
machine or system, or into the sample. It may be necessary to install per-
manent sample valves in some lubricating systems. The oil sample should
be representative of the oil being circulated in the machine. The sample
should, therefore, be collected from a mid-point in reservoirs and
upstream of the filter in circulating systems. Clean sample collection bot-
tles and tubing must be used to collect the sample. Oil sample pumps for
extracting oil from reservoirs are used to avoid contamination. Samples
must be collected from the same point in the system to ensure consisten-
cy in the test analysis. Therefore, the maintenance procedure must provide
detailed direction on where and how to collect samples. It is a good prac-
tice to have operators that use the equipment collect the samples.
Typically, lubricating oil analysis should be performed on a quarterly,
yearly, or 500 / 2000 hours of operations basis for most of the assets. The
analysis schedule should be adjusted based on asset usage, criticality, and
Maintenance Optimization 265

cost. Analyze more frequently for machines that are indicating emerging
problems and less frequently for machines that operate under the same
conditions and are not run on a continuous basis. A new baseline analysis
is recommended following machine repair or oil change out.
Grease is usually not analyzed on a regular basis. Although most of
the testing that is done on oil can also be done on grease, getting a repre-
sentative sample is usually difficult. The machine may have to be disas-
sembled in order to get a good sample that is a homogeneous mixture of
the grease, contaminants, and wear.
Oil Contamination Program A concern common to all machines
with lubricating oil systems is keeping dirt and moisture out of the system.
Common components of dirt, such as silica, are abrasive and naturally
promote wear of contact surfaces. In hydraulic systems, particles can
block and abrade the close tolerances of moving parts. Water in oil pro-
motes oxidation and reacts with additives to degrade the performance of
the lubrication system. Ideally, there would be no dirt or moisture in the
lubricant; this, of course, is not possible. The lubricant analysis program
must therefore monitor and control contaminants.
Oil analysis is a reliable predictive maintenance tool and is very effec-
tive in detecting contaminants in oil that are a result of ingressed dirt or
internal wear debris generated by the effect of machine degradation and
wear. An increase in contaminant levels accelerates the wear-out process
of all components in industrial machine applications. Oil cleanliness is
measured by ISO standard, ISO 4406. For each numerical increase in ISO
contaminant level, the amount of contaminants in the oil almost doubles.
If the standard is a 16/14/11, then the increase in contaminants in the oil
for a 22/21/17 is about 64 times dirtier than the standard.
Contaminants in oils can be prevented. Good filtration on the return
side of hydraulic power units will help in taking out dirt and other
ingressed particles. Usually 3-micron filtration with a 200 beta ratio is the
standard set for most machinery.
Large systems with filters will have steady-state levels of contami-
nants. Increases in contaminates indicate breakdown in the system’s
integrity (leaks in seals, doors, heat exchangers, etc.) or degradation of the
filter. Use of “Air Breathers” in gearboxes and hydraulic system tanks is
a best practice to control contaminants. Unfiltered systems can exhibit
steady increases during operation. Operators can perform a weekly visual
and odor check of lubricating systems and provide a first alert of contam-
ination. Some bearing lubricating systems may have such a small amount
of oil that a weekly check may not be cost effective.
266 Chapter 8

A basic oil contamination control program can be implemented in


three steps:

1. Establish the target fluid cleanliness levels for each machine fluid
system.
2. Select and install filtration equipment (or upgrade current filter
rating) and contaminant exclusion techniques to achieve target
cleanliness levels.
3. Monitor fluid cleanliness at regular intervals to achieve target
cleanliness levels.

It is a good practice to establish a quality control program for incom-


ing oil. Set up a minimum oil cleanliness standard (ISO) for all oils, new
or old, before they are used in the machines.

Electrical Condition Monitoring


Electrical condition monitoring encompasses several technologies
and techniques used to provide a comprehensive system evaluation.
Electrical equipment represents a major portion of a facility’s capital
investment. From the power distribution system to electric motors, effi-
cient operation of the electrical systems is crucial to maintaining opera-
tional capability of a facility.
Monitoring key electrical parameters provides the information to
detect and correct electrical faults such as high resistance connections,
phase imbalance, and insulation breakdown. Because faults in electrical
systems are seldom visible, these faults are costly due to increased elec-
trical usage and increased safety concerns. They involve life cycle cost
issues due to premature replacement of expensive assets. According to the
Electric Power Research Institute (EPRI), voltage imbalances of as little
as 5% in motor power circuits result in a 50% reduction in motor life
expectancy and efficiency in a three-phase AC motors. A 2.5% increase in
motor temperatures can be generated by the same 5% voltage imbalance
accelerating insulation degradation.
Monitoring intervals of several weeks to several months for various
technologies will provide sufficient condition information to warn of
degrading equipment condition. Specific expectations of the length of
warning provided should be factored into developing monitoring intervals
for specific technologies.
Until fairly recently, predictive maintenance technologies for motors
were limited to vibration testing, high-voltage surge testing for winding
Maintenance Optimization 267

faults, meg-Ohm and high-potential tests for insulation resistance-to-


ground, and voltage and current tests for testing phase balance. Many of
these tests still have their place in plant maintenance, but several of them
are dangerous or harmful when tests are conducted with motors in place.
New technologies allow for portable, safe, and trendable tests that can
be used for more accurate troubleshooting or identifying problem areas.
Each of these technologies has its strengths and weaknesses. But as part
of a CBM program, they can accurately detect potential faults and avoid
costly downtime.
Electrical equipment evaluation can be divided into two categories:
online monitoring/testing and offline testing. Online monitoring/testing is
the measuring of any aspect of electrical equipment while it is in service
and operating. Offline testing is usually done by inducing voltage or cur-
rent into equipment and taking electrical measurements, which requires
that the equipment be de-energized and completely isolated from its nor-
mal circuit. Both are very valuable in evaluating an electrical system; in
most cases, they detect different types of faults or potential problems,
meaning much of online monitoring/testing cannot replace offline testing
and vice versa.
Online monitoring is continuous whereas online testing takes meas-
urements at periodic intervals. Some examples of online monitoring
include dissolved gas analysis for transformers, temperature monitoring
for motors and transformers, power quality, and partial discharge. An
example of online testing is current signature analysis, which can detect
broken rotor bars and air gap eccentricity in squirrel cage motors. Current
signature analysis is also capable of detecting mechanical imbalances,
gear mesh problems, broken fan blades, bearing issues, and any other type
of problem that results in torque pulsations that result in changes to the
current draw. Currently, some of the mechanical faults such as bearing
damage are more clearly detected through vibration analysis. All online
testing could be converted to continuous monitoring, but the amount of
data acquired and cost of online testing equipment makes portable devices
and periodic testing more feasible in some cases.
Much of offline testing often involves evaluating the equipment’s
insulation. Other tests, such as low ohm resistance measuring, will inspect
other aspects of the equipment such as high resistance connections. There
are a wide array of tests available that use AC signals, DC signals, and
varying frequency signals. The most common maintenance insulation
tests are DC insulation resistance tests, step voltage tests, high potential
tests, dissipation/power factor tests, and the transformer turns ratio test.
268 Chapter 8

Note that insulation tests are sensitive to environmental factors such as


temperature, humidity, and contamination or cleanliness of the insulation.
Insulation testing is standard industry practice and crucial in determining
the condition of the electrical insulation.
Several of the technologies outlined below are also effective when
used for acceptance testing and certification for new systems.
Motor Current Readings Clamp-on ammeter attachments provide
the capability of taking actual current draw information while the equip-
ment is operating. On three-phase equipment, comparison of current
draws can reveal phase imbalance conditions.
Motor Current Signature Analysis (MCSA) MCSA is performed by
taking current data and analyzing it using Fourier transform analysis. The
primary purpose of the test is rotor bar fault detection, but it is also useful
for detecting rotor faults and power quality problems as well as other
motor and load defects in later stages of failure. Motors must be energized
and loaded during tests.
MCSA is a proven method of detecting the presence of broken or
cracked rotor bars or high resistance connections in end rings. Motor cur-
rent spectrums in both time and frequency domains are collected with a
clamp-on ammeter and Fast Fourier Transform (FFT) analyzer. Rotor bar
problems will appear as side-bands around the power supply line frequen-
cy. MCSA evaluates the amplitude of the side bands that occur about the
line frequency.
AC High Potential Testing (HiPot) HiPot testing applies a voltage
equal to twice the operating voltage plus 1000 volts to motor windings to
test the insulation system. This is typically a “go/no-go” test. Industry
practice calls for HiPot tests on new and rewound motors. This test stress-
es the insulation systems and can induce premature failures in marginal
motors. Due to this possibility, AC HiPot is not recommended as a rou-
tinely repeated condition monitoring technique, but as an acceptance test.
Surge Comparison Testing This testing uses high-voltage pulses to
detect winding faults. Only experienced operators should conduct these
tests because of the potentially harmful effects of high voltage impressed
on used windings and cables. There are also challenges with testing
assembled motors due to rotor effects on the motor circuit. The motor
being tested must be de-energized with controls disconnected.
Surge Testing uses equipment based on two capacitors and an oscillo-
scope to determine the condition of motor windings. This is a comparative
test evaluating the difference in readings of identical voltage pulses
Maintenance Optimization 269

applied to two windings simultaneously. This test is primarily an accept-


ance go/no-go test. Data are provided as a comparison of waveforms
between two phases indicating the relative condition of the two phases
with regard to short circuits. The readings for a particular motor can be
trended, but the repeated stress of the insulation system is not recom-
mended.
Conductor Complex Impedance The total resistance of a conductor
is the sum of its resistance, capacitive impedance, and inductive imped-
ance. Accurate measurement of the conductor impedance allows minor
degradations in a motor to be detected and addressed prior to motor fail-
ure. The condition of the insulation system can be determined by measur-
ing the capacitance between each phase and ground. The presence of
moisture (or other conducting substance) will form a capacitor with the
conductor being one plate, the insulation being the dielectric, and the con-
taminate forming the second plate. Maintaining proper phase balance is
imperative to efficient operation and realizing the full lifetime of electri-
cal equipment.
Megohmmeter Testing A handheld meter is used to measure the
insulation resistance phase-to-phase or phase-to-ground of an electric cir-
cuit. Readings must be temperature-corrected to trend the information, as
the winding temperatures affect the test results.
Polarization Index is another term which gives an idea of the cleanli-
ness of the motor or generator windings. It is a ratio of the Insulation
Resistance Measured for 10 minutes to the insulation resistance value
measured after 1 minute. Because it is a ratio; it does not have any units.
Polarization Ratio = Insulation Resistance after 10 minutes /
Insulation Resistance after 1 minute
The recommended minimum value of the polarization index for AC
and DC motors and generators is 2.0. Machines having windings with a
lower index are less likely to be suited for operation. The procedure for
determining the polarization index is covered in detail by IEEE Standard
No. 43.
Time Domain Reflectometry In this test, a voltage spike is sent
through a conductor. Each discontinuity in the conductor path generates a
reflected pulse. The reflected pulse and time difference between initial
pulse and reception of the reflected pulse indicate the location of the dis-
continuity.
Radio Frequency (RF) Monitoring RF monitoring is used to detect
arcs caused by broken windings in generators. It consists of establishing
270 Chapter 8

RF background levels and the amplitude trend over a narrow frequency


band.
Power Factor and Harmonic Distortion Maintaining optimum
power factor maximizes the efficient use of electrical power. The power
factor is the ratio of real power to reactive power usage. Dual channel
data-loggers are used to determine the phase relationship between voltage
and current, then to calculate the power factor. If this process detects a low
power factor, subsequent engineering analysis will be required to devise a
means of improving power system power factor.
Application of Various Technologies to Electrical Assets Specific
electrical assets that can be monitored by CBM technologies are:

• Electrical Distribution Cabling. Megohmmeter, Time Domain


Reflectometry, HiPot, Infrared Thermography(IRT), and Airborne
Ultrasonic.
• Electrical Distribution Switchgear and Controllers. Timing, Visual
Inspection, IRT, and Airborne Ultrasonic
• Electrical Distribution Transformer. Oil Analysis, Turns Ratio,
Power Factor, and Harmonic Distortion
• Electrical Motors. Current Draw, Motor Current Spectrum
Analysis, Motor Circuit Analysis, Megohmmeter, HiPot, Surge
Test, Conductor Complex Impedance, Starting Current, and Coast-
Down Time
• Generators. Megohmmeter, RF, and Coast-Down Time

Technologies Limitations
The technologies discussed earlier can be divided into two categories:

• Energized. These technologies can safely provide information on


energized systems and require the system be energized and
operational. They include IRT, Ultrasonics, Motor Current
Readings, Starting Current, Motor Current Spectrum Analysis, RF,
Power Factor, and Harmonic Distortion.
• De-Energized. These technologies require the circuit to be de-
energized for safe usage. They include Surge Testing, HiPot
Testing, Time Domain Reflectometry (TDR), Megohmmeter,
Motor Circuit Analysis, Transformer Oil Analysis, Turns Ratio,
and Conductor Complex Impedance.
Maintenance Optimization 271

Each technology will require specific initial conditions to be set prior


to conducting the test. For instance, prior to an IRT survey, typical equip-
ment powered through the switchboard should be running to bring the dis-
tribution equipment to normal operating temperatures. Higher load accen-
tuates problem areas. Conducting the survey at low load conditions may
allow a problem to remain undetected.
HiPot and surge testing should be performed with caution. The high
voltage applied during these tests may induce premature failure of the
units being tested. For that reason, these tests normally are performed only
for acceptance testing, not for condition monitoring.

Other Miscellaneous Non-Destructive Testing


Non-Destructive Testing (NDT) evaluates material properties and
quality of expensive components or assemblies without damaging the
product or its function. Typically, NDT has been associated with the weld-
ing of large high-stress components such as pressure vessels and structur-
al supports. Process plants such as refineries or chemical plants use NDT
techniques to ensure integrity of pressure boundaries for systems process-
ing volatile substances. The following are various NDT techniques.
Radiography Radiography is performed to detect sub-surface
defects. Radiography or X-ray is one of the most powerful NDT tech-
niques available in industry. Depending on the strength of the radiation
source, radiography can provide a clear representation (radiograph) of dis-
continuities or inclusions in material several inches thick. X-ray or
gamma ray sensitive film is placed on one surface of the material to be
examined. The radiation source is positioned on the opposite side of the
piece. The source may be either a natural gamma emitter or a powered X-
ray emitter. The source is accurately aligned to ensure the proper exposure
angle through the material. When all preparations and safety precautions
are complete, the radiation source is energized or unshielded.
Radiography, though a versatile tool, is limited by the potential health
risks. Use of radiography usually requires the piece be moved to a special
shielded area, or that personnel be evacuated from the vicinity to avoid
exposure to the powerful radiation source required to penetrate several
inches of dense material.
Ultrasonic Testing (Imaging) Ultrasonic imaging provides detection
of deep sub-surface defects. Ultrasonic inspection of welds and base
material is often an alternative or complementary NDT technique to radi-
ography. Though more dependent on the skill of the operator, ultrasonic
272 Chapter 8

inspection does not produce the harmful radiation experienced with radi-
ography. Ultrasonic inspection is based on the difference in the wave
reflecting properties of defects and the surrounding material. An ultrason-
ic signal is applied through a transducer into the material being inspected.
The speed and intensity with which the signal is transmitted or reflected
to a transducer provides a graphic representation of defects or discontinu-
ities within the material.
Due to the time and effort involved in surface preparation and testing,
ultrasonic inspections are often conducted on representative samples of
materials subjected to high stress levels, high corrosion areas, and large
welds.
Magnetic Particle Testing (MPT) Magnetic Particle Testing uses
magnetic particle detection of shallow sub-surface defects. It is a very
useful technique for localized inspections of weld areas and specific areas
of high stress or fatigue loading. MPT provides the ability to locate shal-
low sub-surface defects. Two electrodes are placed several inches apart on
the surface of the material to be inspected. An electric current is passed
between the electrodes producing magnetic lines. While the current is
applied, iron ink or powder is sprinkled in the area of interest. The iron
aligns with the lines of flux. Any defect in the area of interest will cause
distortions in the lines of magnetic flux, which will be visible through the
alignment of the powder. Surface preparation is important because the
powder is sprinkled directly onto the metal surface and major surface
defects will interfere with sub-surface defect indications. Also, good elec-
trode contact and placement is important to ensure consistent strength in
the lines of magnetic flux.
A major advantage of MPT is its portability and speed of testing. The
handheld electrodes allow the orientation of the test to be changed in sec-
onds. This allows for inspection of defects in multiple axes of orientation.
Multiple sites can be inspected quickly without interrupting work in the
vicinity. The equipment is portable and is preferred for onsite or in-place
applications. The results of MPT inspections are recordable with high
quality photographs.
Hydrostatic Testing Hydrostatic Testing is another NDT method for
detecting defects that completely penetrate pressure boundaries.
Hydrostatic Testing is typically conducted prior to the delivery or opera-
tion of completed systems or subsystems that act as pressure boundaries.
During the hydrostatic test, the system to be tested is filled with water or
the operating fluid. The system is then sealed and the pressure is increased
to approximately 1.5 times operating pressure.
Maintenance Optimization 273

This pressure is held for a defined period. During the test, inspections
are conducted to find visible leaks as well as monitor pressure drop and
make-up water additions. If the pressure drop is out of specification, any
leaks must be located and repaired. The principle of hydrostatic testing
can also be used with compressed gases. This type of test is typically
called an air drop test and is often used to test the integrity of high pres-
sure air or gas systems.
Eddy Current Testing Eddy current testing is used to detect surface
and shallow subsurface defects. Also known as electromagnetic induction
testing, eddy current testing provides a portable and consistent method for
detecting surface and shallow subsurface defects. This technique provides
the capability of inspecting metal components quickly for defects or
homogeneity. By applying rapidly varying AC signals through coils near
the surface of the test material, eddy currents are induced into conducting
materials. Any discontinuity that affects the material’s electrical conduc-
tivity or magnetic permeability will influence the results of this test.
Component geometry must also be taken into account when analyzing
results from this test.

Why Have a CBM Program


Condition Based Maintenance can:

• Warn of most problems in time to minimize unexpected failure,


the risk and consequences of collateral damage, and adverse
impact on safety, operations, and the environment. It will reduce
the number of preemptive corrective actions.
• Increase equipment utilization and life; minimize disruption to
mission and schedule. It will decrease asset and process down-
time, resulting in increased availability.
• Reduce maintenance costs — both parts and labor.
• Reduce a significant amount of calendar / run-based preventive
maintenance.
• Minimize cost and hazard to an asset that result from unnecessary
overhauls, disassemblies, and PM inspections.
• Increase the likelihood that components operate to optimum life-
time. In some cases, replacement prior to end-of-life is more effi-
cient for meeting operational requirements and optimum cost.
• Reduce requirements for emergency spare parts.
• Increase awareness of asset condition.
274 Chapter 8

• Provide vital information for continuous improvement, work, and


logistic planning.
• Improve worker safety.
• Increase energy savings.

However, CBM cannot:

• Eliminate defects and problems, or stop assets from deteriorating.


• Eliminate all preventive maintenance.
• Reliably and effectively warn of fatigue failures.
• Reduce personnel or produce a major decrease in lifetime
maintenance costs without a commitment to eliminating defects
and chronic problems.

CBM is not a “silver bullet.” Some potential failures, such as fatigue,


or uniform wear on a blower fan are not easily detected with condition
measurements. In other cases, sensors may not be able to survive in the
environment; measurements to assess condition may be overly difficult
and may require major asset modifications.

8.6 Other Maintenance Strategies

Maintenance Tasks to Prevent Failures


An asset has a predefined life expectancy based upon how it has been
designed. The design life of most assets requires periodic maintenance.
For example, belts and chains require adjustment; alignment of shafts
such as pump-motor shafts need to be properly maintained; filters need to
be changed at regular intervals; proper lubrication on rotating machinery
is required; and so on. In some cases, certain components need replace-
ment after a specified number of hours of operations, e.g., a pump bear-
ing on a hydraulic system to ensure that the system lasts through its design
life. Anytime we fail to perform maintenance activities, we may be short-
ening the operating life of the asset. Over the past 40 years, many cost-
effective approaches have been developed to insure an asset reaches or
exceeds its design life. Instead of waiting for assets to fail and then fixing
them, maintenance actions are performed to keep assets in good working
condition to provide continuous service.
When an asset breaks down, it fails to perform its intended function
and disrupts scheduled operation. This functional loss — partial or total
— may result in defective parts, speed reduction, reduced output, and
Maintenance Optimization 275

unsafe conditions. For example, a wear or slight damage on a pump


impeller, which reduces output, is a function reduction failure. Function-
disruption or reduction failures that are not given due attention will soon
develop into asset stoppage if not acted on.
Many abnormalities such as cracks, deformations, slacks, leaks, cor-
rosion, erosion, scratches, excessive heat, noise, and vibration, are indica-
tors of imminent troubles. Sometimes these abnormalities are neglected
because of the insignificance or the perception that such abnormalities
will not contribute to any major breakdowns. The tendency to overlook
such minor abnormalities may contribute to serious catastrophic failures.
It is not uncommon to receive queries from production staff in response to
a “high temperature or vibration condition” about how long they can con-
tinue running.
It has been observed that a high percentage of failures occur during
startups and shutdowns. However, asset failure could also be due to poor
maintenance. Causes that go unnoticed are termed as “hidden abnormali-
ties.” The key to achieving zero failures is to uncover and remedy these
hidden abnormalities before failure actually occurs. This is the fundamen-
tal concept of maintenance, specifically Preventive and Condition/
Predictive Maintenance.

Preventive Maintenance (PM)


Preventive maintenance refers to a series of actions that are performed
on an asset on schedule. That schedule may be either based on time or
based upon machine runtime or the number of machine cycles. These
actions are designed to detect, preclude, or mitigate degradation of a sys-
tem and its components. The goal of a preventive maintenance approach
is to minimize system and component degradation and thus sustain or
extend the useful life of the asset.
Preventive maintenance is the planned maintenance of assets
designed to improve asset life and avoid unscheduled maintenance activ-
ity. PM includes cleaning, adjusting, and lubricating, as well as minor
component replacement, to extend the life of assets and facilities. Its pur-
pose is to minimize failures. Neither assets nor facilities should be
allowed to go to the breaking point unless we have selected a run-to-fail-
ure strategy for that specific asset. In its simplest form, preventive main-
tenance can be compared to the service schedule for an automobile. The
amount of preventive maintenance needed at a facility varies greatly. It
can range from walkthrough inspections of assets and facilities to meas-
uring bearing clearances, checking pump and motor alignment, etc., while
276 Chapter 8

noting other deficiencies for later corrections.


The objective of preventive maintenance can be summarized as fol-
lows:

• Maintain assets and facilities in satisfactory operating condition


by providing for systematic inspection, detection, and correction
of incipient failures either before they occur or before they devel-
op into a major failure.
• Perform maintenance, including tests, measurements, adjust-
ments, and parts replacement, specifically to prevent failure from
occurring.
• Record asset health condition for analysis, which leads to the
development of corrective tasks.

Preventive maintenance is typically performed based upon calendar


time. Maintenance personnel schedule periodic visits to an asset based on
fixed time intervals, for example, every three or six months. Although bet-
ter than no PM at all, time-based PMs are not the optimal way to run PM
programs. They may result in too much time being spent on an asset.
Numerous visits to assets with “no data – abnormalities found” can be
regarded as wasted maintenance dollars. If this happens, the PM periodic-
ity should be reevaluated and adjusted. Nevertheless, time-based PMs are
a good approach for assets having fixed operating schedule such as 24/7
or 80 hours/week operation.
Typically, the next step up from time-based PMs is performing PMs
based upon asset cycles or runtime (operations-based PM, not the same as
operator-based PM). Intuitively, this approach makes sense. An asset does
not have to be checked repeatedly if it has not been used. Generally speak-
ing, it is the actual operation of the asset that wears it down, so it makes
sense to check the asset after it has been working for a specified amount
of time to cause some wear. It may be necessary either to adjust or replace
the component.

PM Myths and Practices


If we were to conduct a survey among the maintenance professionals
to ascertain how their PM came about or the basis of their program, the
responses would probably fail to provide definitive and meaningful infor-
mation. Most existing PM programs cannot be traced to their origins. For
those that can, most are unlikely to make sense. The following reasons are
usually the ones given for a PM program:
Maintenance Optimization 277

• OEM Recommendations. The vendor / equipment supplier says


“do this.” The problem with this argument is that the vendor’s
recommendations are mainly based on their judgment. But the
vendor frequently does not know how this equipment will be
used. The equipment may be designed for steady-state operations,
but the real application could be highly cyclical. Moreover, ven-
dors want us to check everything because it doesn’t cost them
anything.
• Experience. This is the most common answer given to justify
current PM tasks. “It has been done this way for years, so it must
be good.”
• Failure Prevention. The belief that all failures can be prevented
suggests that an overhaul task can help reduce the failures —
without understanding the mechanism of failure. But the overex-
tended use of overhauls can be counterproductive. It can create
failures that were not present before the overhaul. Some items
wear out with age, but many items don’t. In the absence of wear
out or aging mechanisms, correcting a problem that does not exist
is a waste of money.
• Brute Force. “More is always better.” Thus, if it is physically
possible to do something on equipment that appears to have PM
characteristics, then it must be a good thing to do. This belief
leads to over-lubrication, or cleaning when equipment shouldn’t
even be touched, part replacement when the installed part is fine,
etc.
• Regulations. Many products and services come under the gover-
nance of some sort of regulatory agency such as EPA, FDA,
OSHA, NRC, or a Public Utilities Commission. In their well-
meaning ways, these regulations can mandate PM actions that are
potentially counterproductive to their objectives.

Rather than using the preceding reasons, a good PM program should be


based on a FMEA / RCM analysis. In addition, there is a risk of poor
workmanship in performing PM tasks. Typically, risk may include:

• Damage to the asset receiving the PM — damage during inspec-


tion, repair, adjustment, or installation of a replacement part or
material that is defective
• Incorrect reassembly or wrong-installation of parts
278 Chapter 8

• Infant mortality of replaced parts or material


• Damage to adjacent equipment/machinery during a PM task

Traditional thinking has been that the goal of preventive maintenance


(PM) is to preserve assets. On the surface, it makes sense, but the prob-
lem is in that mindset. In fact, that thinking has been proven to be flawed
at its core. The blind quest to preserve assets has produced many prob-
lems, such as being overly conservative with any maintenance actions that
could cause damage due to intrusive actions, thereby increasing the
chances of human error. Other flaws include both thinking that all failures
are equal and performing maintenance simply because there is an oppor-
tunity to do so.

The 10 Percent Rule of PM


A PM plan must be executed per schedule. The best practice is to use
the 10 percent rule of PM — a time-based PM must be accomplished
within 10 percent of the time frequency to remain within compliance.
Many organizations use a “PM compliance” metric as a measurement of
their maintenance department’s performance. If an asset is on a 30-day
PM schedule, it should be executed within +/– 3 days of its due date; oth-
erwise it is out of compliance. This rule should apply to all PMs, but we
must ensure that, at a minimum, critical assets are being maintained prop-
erly at the right time, within 10 percent of time frequency.
Organizations that have implemented the 10 percent rule have been
found to have increased reliability of the assets due to a consistent and dis-
ciplined approach.

If we are performing Preventive Maintenance on an asset that


continues to fail, we are in a reactive maintenance mode.
The PM plan should be reviewed and adjusted.

Run-to-Failure (RTF)

RTF is a maintenance strategy where the organization decides to


allow specific assets/systems to fail without any PM or CBM performed
against them. This strategy is not the same as reactive maintenance. (In
reactive maintenance, an organization does not have a structured mainte-
nance program, which would include elements of PM, CBM, and RTF
spread throughout the facility, with each asset/system having its own spe-
Maintenance Optimization 279

cific maintenance strategy.) For assets where the cost and impact of fail-
ure is less than the cost of preventive (PM and CBM) actions, RTF may
be an appropriate maintenance strategy. It is a deliberate decision based
on economical effectiveness.
Many times, we do consider and accept RTF for specific non-critical
assets or components. However, we usually fail to document this fact in
our CMMS. It is imperative that we document that RTF was chosen on
purpose and what the criteria or basis was for this decision. Additionally,
we must have a plan to repair the failure, if and when it happens. An
example is a spare parts program for the specifically-selected RTF
assets/components that allow for minimal downtime for these.
Documentation minimizes the excitement when RTF failure occurs. We
need to understand that it was a deliberate economical decision not to
have a PM program for that asset.
This maintenance strategy can be a valid stand-alone maintenance
strategy, especially in today’s budget constrained environment. However,
the RTF strategy needs more discussion throughout the maintenance com-
munity regarding how to establish it more formally as an effective piece
of your overall maintenance strategy. Such a discussion will allow this
strategy to stretch beyond its near-reactive tendencies by some organiza-
tions.

Final Thoughts on Maintenance Optimization


Finally, it is suggested that all assets or, at a minimum, critical assets
should have an asset management strategy established and documented.
An asset management strategy should have the following:

• Maintenance strategies selected and why (basis)


• List by subassembly or components
• Asset hierarchy structure identified
• Suggested spare parts needed — when and how many (basis)
• List of PM and CBM actions/tasks/routes
• Major repair plans
• Operating guidelines or procedures
• List of key troubleshooting procedures
• Any specific qualification/certification needs for maintenance
personnel
280 Chapter 8

8.7 Summary
Reliability-Centered Maintenance, often known as RCM, is a main-
tenance improvement approach focused on identifying and establishing
the operational, maintenance, and design improvement strategies that will
manage the risks of asset failure most effectively. The technical standard
SAE JA1011 has established evaluation criteria for RCM, which specifies
that RCM address, at a minimum, the following seven questions:

1. What is the asset or component supposed to do? (functions)


2. In what ways can it fail to provide the required functions?
(functional failures)
3. What are the events that cause each failure? (failure modes)
4. What happens when each failure occurs? (failure effect)
5. Why does the failure matter? (failure consequences)
6. What task can be performed proactively to prevent, or to diminish
to a satisfactory degree, the consequences of the failure?
7. What must be done if a suitable preventive task cannot be
found?

Thus, RCM is a process that determines what must be done to ensure


that assets continue to do what their users need them for in a certain oper-
ating context. RCM analysis provides a structured framework for analyz-
ing the functions and potential failures of assets. RCM is a maintenance /
PM plan optimizing strategy. However, RCM analysis provides the max-
imum benefits during the asset’s life.
Condition Based Maintenance (CBM) is a process that determines
what must be done to ensure that assets continue to function cost-effec-
tively in the desired manner based on actual operating environment. CBM
is based on using real-time data to assess the condition of the assets utiliz-
ing predictive maintenance technologies. The data and its analysis help us
to make better decisions to optimize maintenance resources. CBM will
determine the equipment’s health, and act only when maintenance is actu-
ally necessary.
Condition Based Maintenance endeavors to predict impending fail-
ure based on actual operating data instead of relying on traditional
Preventive Maintenance, generally eliminating unnecessary maintenance
performed. Thus, CBM is another maintenance optimizing strategy. In
fact, when it is used with RCM in establishing maintenance tasks, it pro-
duces a much better return on investment.
Maintenance Optimization 281

Preventive Maintenance (PM) is the basic asset (maintenance) strat-


egy that many organizations use to begin their formal maintenance pro-
gram. PM is probably the first program that most maintenance experts use
to establish a maintenance program in any organization. Most PM pro-
grams are either calendar time-based or operations runtime-based.
Run-to-failure (RTF) is another economically valid strategy for
specifically identified assets/systems. This strategy must be deliberately
selected for non-critical assets only. It should be documented and planned
with the right level of support, such as spare parts.

8.8 Self Assessment Questions

Q8.1 What is RCM? How did it get its start? Tell a little about RCM’s
history.

Q8.2 Which Standards Development Organization (SDO) developed


RCM Standard JA1011?

Q8.3 Describe the 4 principles of RCM. What is the key objective of


RCM analysis?

Q8.4 During what phases of asset development do we get the maxi-


mum benefit of a RCM analysis? Why?

Q8.5 Describe the 9-step RCM analysis process.

Q8.6 Which type of failure mode is not evident to asset operator?

Q8.7 What are the benefits of RCM?

Q8.8 What is meant by CBM and PdM? What methods are used to
perform these?

Q8.9 What is the difference between diagnostic and prognostic


analysis?

Q8.10 What is velocity analysis? With which CBM technology is it


associated? What does a peak at twice rotational speed indi-
cate?
282 Chapter 8

8.9 References and Suggested Reading

Corio, Mario R. and Lynn P. Constantini. Frequency and Severity of


Forced Outage Immediately Following Planned or Maintenance
Outages. North American Electric Reliability Council. May
1989.
Gulati, Kahn & Baldwin. The Professional’s Guide to Maintenance and
Reliability Terminology. Reliabilityweb Publication, 2010.
Murphy, Thomas J. and Allan A. Reinstra. HEAR MORE. Reliabilityweb
Publication, 2009.
Moubray, John. Reliability-Centered Maintenance. Industrial
Press, Inc., 1997.
Nicholas, Jack, and R. Keith Young. Predictive Maintenance
Management. Reliabilityweb Publication, 2007.
SAE JA1011. Evaluation Criteria for Reliability-Centered Maintenance
(RCM) Processes. August 1999.
SAE JA1012. A Guide to the Reliability-Centered Maintenance (RCM)
Standard. January 2002.
Smith, Anthony and Glenn R. Hinchcliffe. RCM — Gateway to World
Class Maintenance. Elsevier Inc, 2003.
Smith, A.M. (AMS Associate), Glen Hinchcliffe (G&S Associate), and
Ramesh Gulati. RCM Training Manual. Sverdrup, Inc., 1998.
Smith, A. M. Reliability Centered Maintenance. McGraw Hill, New
York, 1993. ISBN 0-07-059046-X
Chapter 9
Managing Performance

“You cannot manage something you cannot control,


and you cannot control something you cannot measure.”
— Peter Drucker

9.1 Introduction
9.2 Key Terms and Definitions
9.3 Identifying Performance Measures
9.4 Data Collection and Data Quality
9.5 Benchmarking and Benchmarks
9.6 Summary
9.7 Self Assessment Questions
9.8 References and Suggested Reading

After reading this chapter, you will be able to understand:

• What to measure and why to measure performance


• Differences between lagging and leading indicators
• Key performance indicators
• A balanced scorecard
• The challenges of data collection
• The importance of data quality and integrity
• Benchmarks and benchmarking

283
284 Chapter 9

9.1 Introduction

An organization must measure and analyze its performance if it is to


make the improvements needed for staying in business in a competitive
market place. The performance measures must be derived and aligned
with the organization’s goals and strategies of the business. They should
be centered on the critical information and data related to the key business
processes and outputs, and should be focused on improving results.
Data needed for process improvement and performance measurement
includes information about products and services, asset performance, cost
of operations, and maintenance. Data can be easy or difficult to collect,
and emphasis must be placed on data quality. This data is analyzed to
determine trends, cause and effects, and the underlying reasons for certain
results that may not be evident without analysis. Data are also used to
serve a variety of purposes, such as planning, projections, performance
reviews, and operations improvements, and for comparing an organiza-
tion’s performance with the “best practices” benchmarks.
A key component of an improvement process involves the creation
and use of performance indicators, also known as metrics. These metrics
are measurable characteristics of products, services, and processes related
to the business. They are used by an organization to track and improve its
performance. Metrics should be selected to best represent the factors that
lead to improved operations, including maintenance and customer satis-
faction. A comprehensive set of measures or metrics tied to the business
activities and customers should be based on long- and short-term goals of
the organization. Metrics need to be constantly reviewed and aligned with
the new or updated goals of the organization and become part of its strate-
gic plan.
Metrics based on the priorities of the strategic plan make lasting
improvements to the key business drivers of the organization. Processes
are then designed to collect information relevant to these metrics and
reduce them to numerical form for easy dissemination and analysis.
The value of metrics is in their ability to provide a factual basis in the
following areas:
• Strategic feedback to show the present status of the organization
from various perspectives
• Diagnostic feedback of various processes to guide improvements
on a continuous basis
Managing Performance 285

• Trends in performance over time as the metrics are tracked


• Feedback around the measurement methods themselves in order
to track the correct metrics

In most businesses, success is easily measured by looking at the bot-


tom line—the profit. But what’s the bottom line for maintenance as a busi-
ness function? To better understand how to evaluate maintenance business
performance, it’s helpful to examine how businesses generate profit. In
simple terms, businesses generate profit by selling goods and services and
by minimizing their costs. Obviously, revenues generated from sales must
exceed the costs.
Customers generally demand value. Key components of value are:
timeliness, quality, price, and return on investment (ROI). Therefore, met-
rics for maintenance and reliability should reflect how an organization is
providing value to its customers in terms of timely maintenance (avail-
ability of assets), quality of service (minimum rework), controlling costs,
etc. Thus, maintenance as a business function must develop its internal
metrics to evaluate its performance in terms of these parameters.

The Benefits of Performance Measurement

Accountability
Well-designed performance measures document progress toward
achievement of goals and objectives, thereby motivating and catalyzing
organizations to fulfill their obligations to their employees, stakeholders,
and customers.

Resources / Budget Justification


Because it ties activities to results, performance measurement
becomes a long-term planning tool to justify proper resource or budget
allocation.

Ownership and Teamwork


By providing a clear direction for concentrating efforts in a particular
functional area, performance measurement provides more employees par-
ticipation in problem solving, goal setting, and process improvements. It
helps set priorities and promotes collaboration among departments and
business areas.
286 Chapter 9

Communication—A Common Language


Achievement of goals through metrics can enhance employee under-
standing and support of management strategies and decisions. They also
give employees a common language to communicate, alert them to poten-
tial problem areas, and encourage them to share knowledge. Therefore,
performance metrics, if properly designed and implemented, enhance pro-
ductivity and reduce cost.

9.2 Key Terms and Definitions

Benchmark
A standard measurement or reference that forms the basis for
comparison; this performance level is recognized as the stan-
dard of excellence for a specific business process.

Benchmarking
American Productivity and Quality Council (APQC) defines
benchmarking as the process of identifying, learning, and adapt-
ing outstanding practices and processes from any organization,
anywhere in the world, to help an organization improve its per-
formance. Benchmarking gathers the tacit knowledge—the
know-how, judgments, and enablers.

Benchmarking Gap
The difference in performance between the benchmark for a
particular activity and the level of other organizations. It is the
measured performance advantage of the benchmark organiza-
tion over other organizations.

Best-in-Class
Outstanding process performance within an industry; words
used as synonyms are best practice and best-of-breed.

Best Practice
A method or technique that has been found to be the most effec-
tive and has consistently achieved superior results compared to
results achieved with other means, e.g., current practices while
minimizing the use of an organization’s resources. This practice
becomes a benchmark.
Managing Performance 287

Generic Benchmarking
Benchmarking process that compares a particular business func-
tion or process with other organizations, independent of their
industries.

Goals
An observable and measurable end result having objectives that
will be achieved within a more or less fixed time frame. Goals
indicate the strategic direction of an organization.

Internal Benchmarking
Benchmarking process that is performed within an organization
by comparing similar business units or business processes.

Metric
A metric is a standard measure to assess performance in a spe-
cific area. Metrics are at the heart of a good, customer-focused
process management system and any program directed at con-
tinuous improvement.

Networking
A practice of building up informal relationships with people,
with a common set of values and interests that can help both par-
ties professionally.

Objective
The set of results to be achieved that will deploy a vision
into reality.

Performance
The results of activities of an organization over a given peri-
od of time.

Performance Measurement
A quantifiable indicator used to assess how well an organization
or business is achieving its desired objectives.

Vision
The achievable dream of what an organization wants to do and
where it wants to go.
288 Chapter 9

World-Class
Practices that are ranked by the customers and industry experts
to be among the best of the best. An exemplary performance
achieved independent of industry, function, or location.

9.3 Identifying Performance Measures

It is often said that “what gets measured gets done.” Getting things
done, through people, is what management is all about. Measuring things
that get done and the results of their effort is an essential part of success-
ful management. But too much emphasis on measurements or the wrong
kinds of measurements may not be in the best interest of the organization.
A few vital indicators which are important for evaluating process per-
formance are called KPIs—Key Performance Indicators. KPIs are an
important management tool; they measure business performance, includ-
ing maintenance. There only are a few “hard” measures of maintenance
output and the metrics that are commonly used are often easy to manipu-
late. Maintenance and operations KPIs must be integrated to make them
effective and balanced. There are three other criteria that should be con-
sidered when deciding what aspects of maintenance to measure:
1. The performance measures should encourage the right behavior.
2. They should be difficult to manipulate to “look good.”
3. They should not require a lot of effort to measure.

Some metrics may encourage people to do things that we do not want.


A common metric is “adherence to weekly work schedule” for mainte-
nance work. It’s easy to achieve a high adherence to schedule by schedul-
ing less work, through overestimating work orders. However, what we
really want is higher productivity, which can often be achieved by chal-
lenging people and scheduling more work, but with better work estimat-
ing. Thus the wrong measurement may work against us. Like “adherence
to schedule,” some other metrics, such as time spent on Preventive
Maintenance (PM) work, percent rework, and percent emergency work
are easy to manipulate.
The KPIs which are truly relevant and satisfy the criteria listed above
should only be considered only for implementation. A good example
comes from an organization trying to improve turnaround and shutdown
planning, where a new target of completing all planning work two or three
weeks in advance of a shutdown has been set and agreed upon. All of its
Managing Performance 289

shutdown work orders have a specific code. Therefore, a simple report


from the CMMS listing all purchase requisitions against work orders for
a specific shutdown that were originated less than two or three weeks in
advance will provide a very useful measure to evaluate planning perform-
ance. This metric supports the right behavior, is unlikely to be manipulat-
ed, and is easy to measure. It will also provide information on where to
take action and when to recognize good planning efforts.
Metrics such as the one in this example are of immense value when
measuring the success of efforts to implement better practices and to
change the behavior. These metrics may in turn be discontinued when the
new and improved practices become a habit.

Metrics Development Process


The first step in developing metrics is to involve the people who are
responsible for the work to be measured. They are the most knowledge-
able about the work. Once these people are identified and involved, it is
necessary to:

1. Identify critical work processes and customer requirements.


2. Identify critical results desired and align them to customer
requirements.
3. Develop measurements for the critical work processes or critical
results.
4. Establish performance goals, standards, or benchmarks.

A SMART test can be used to ensure the quality of a particular per-


formance metric. Here, the letters in SMART represent:

S = Specific. Be clear and focused to avoid misinterpretation. The


metric should include measurement assumptions and
definitions and should be easily interpreted.
M = Measurable. The metric can be quantified and compared to
other data. It should allow for meaningful statistical
analysis. Avoid “yes/no” measures except in limited cases
such as start-up or systems-in-place situations.
A = Attainable. The metric is achievable, reasonable, and credible
under conditions expected.
R = Realistic. It fits into the organization’s constraints and is cost-
effective.
T = Timely. It’s do-able, data is available within the time needed.
290 Chapter 9

Figure 9.1 lists a few examples of key maintenance and reliability


related metrics.

Figure 9.1 List of Maintenance and Reliability Metrics.


Managing Performance 291

Leading and Lagging Indicators


A simple way to determine if a metric is leading or lagging is to ask
the question, “Does the metric allow us to look into the process, or are we
outside of the process looking at the results?” Leading indicators are for-
ward looking and help manage the performance of an asset, system, or
process, whereas lagging indicators tell how well we have managed.
Process measures are leading indicators. They offer an indication of
task performance with a lead time to manage for successful results. For
example, a leading maintenance process indicator will measure how
proactive the planning or scheduling function has been in preparing pre-
ventive and condition-based maintenance work packages or to monitor
the percentage of PM / CBM inspections completed per schedule. If peo-
ple are doing all the right things, then the expectation is that improved
results will follow. The leading process indicators are typically more
immediate than lagging measures of results. We must manage by the lead-
ing indicators. Some examples of M&R-related leading metrics are:

• % Schedule compliance
• % Planned work
• % PM / CBM work compliance (completed on time)
• Work order cycle time
• % Rework
• Planner to craft workers ratio

Lagging indicators are results that occur after the fact. They monitor
the output of a process. They measure the results of how well we have
managed an asset, process, or overall maintenance business. Some exam-
ples of M&R-related lagging metrics are:

• Maintenance cost as % of RAV


• Return on Net Assets (RONA)
• Asset Availability
• MTBF
• OEE
• Maintenance Training Man-hours or $

Figure 9.2 on the next page shows a hierarchical model of lagging


and leading indicators.
292 Chapter 9

Figure 9.2 Leading and Lagging KPI Model

On a cautionary note, an indicator could be either leading or lagging.


For example, PM/CBM work compliance is a lagging indicator—the
result of how much PM/CBM work is completed—when viewed in the
context of work execution. However, when viewed as an indicator of asset
reliability, PM/CBM compliance is a leading indicator of the reliability
process. Higher PM/CBM work compliance predicts or very likely leads
to improved asset reliability. Similarly, improved asset reliability will lead
to reduced maintenance costs, which is a lagging indicator of the mainte-
nance process.
Whether leading or lagging, metrics should be used to provide infor-
mation on where the process is working well and where it isn’t. In doing
so, these metrics help build on successes and lead to making process
changes where unfavorable trends are developing.

Balanced Scorecard
Most of the time, we measure what’s important from the financial and
productivity prospective of a process or an organization. The balanced
scorecard suggests that we view the process or an organization from four
perspectives. We should also develop metrics, collect data, and analyze
the data relative to each of these perspectives to balance out any bias.
The Balanced Scorecard is a strategic management approach devel-
oped in the early 1990s by Dr. Robert Kaplan of Harvard Business School,
Managing Performance 293

Figure 9.3 The Balanced Scorecard

and Dr. David Norton. As the authors describe the approach,


“The balanced scorecard retains traditional financial measures. But
financial measures tell the story of past events, an adequate story for
industrial age companies for which investments in long-term capabilities
and customer relationships were not critical for success. These financial
measures are inadequate, however, for guiding and evaluating the journey
that organizations must make to create future value through investment in
customers, suppliers, employees, processes, technology, and innovation.”

The balanced scorecard (Figure 9.3) identifies four perspectives from


which to view a process or an organization. These are:

• Learning and Growth Perspective


• Business Process Perspective
• Customer Perspective
• Financial Perspective
294 Chapter 9

The balanced scorecard is a strategic planning and management sys-


tem that is used extensively in business and industry, government, and
nonprofit organizations worldwide. It helps to align business activities to
the vision and strategy of the organization, improve internal and external
communications, and monitor organization performance against strategic
goals. It provides a balanced view of organizational performance.
The balanced scorecard has evolved from its early use as a simple per-
formance measurement framework to a full strategic planning and man-
agement system. The “new” balanced scorecard transforms an organiza-
tion’s strategic plan from an attractive but passive document into the
marching orders for the organization on a daily basis. It provides a frame-
work that not only provides performance measurements, but helps organ-
izations to identify what should be done and measured. It enables execu-
tives to truly execute their strategies.

The Learning and Growth Perspective


This perspective includes employee training and corporate cultural
attitudes related to both individual and corporate self-improvement. In a
knowledge-worker organization, people—the only repository of knowl-
edge—are the main resource. In the current climate of rapid technological
change, it is becoming necessary for knowledge workers to be in a con-
tinuous learning mode. Government agencies often find themselves
unable to hire new technical workers. At the same time, there is a decline
in training of existing employees. This is a leading indicator of a “brain
drain” that must be reversed. Metrics can be put into place to guide man-
agers in focusing training resources where they can help the most.
Kaplan and Norton emphasize that “learning” is more than “training.”
It also includes things like mentors and tutors within the organization, as
well as ease of communication among workers that allows them to readi-
ly get help on a problem when it is needed. It also includes technological
tools; what the Baldrige criteria calls “high performance work systems.”
In the maintenance area, these tools include the use of new technologies,
e.g., Ultrasonic, Infrared Thermography, Motor Current Analysis, and
applying RCM in new designs.
Maintenance & Reliability (M&R) related examples of this perspec-
tive are:
• Hours (or dollars) spent on training per person, e.g., 80
hours/person in a given year
• Percent of training hours per total, e.g., 5% in year 2009
Managing Performance 295

• Number of technical papers presented or written/$M of M&R


budget
• Percent of employees certified in Condition Based Maintenance
(CBM) technologies or Certified Maintenance Reliability
Professionals (CMRP)
• Percent of work orders created by CBM/Predictive Maintenance
(PdM) technology
• Percent of CBM tasks in overall Preventive Maintenance (PM)
program
• Percent of FMEA/RCM processes applied to new designs

The Business Process Perspective


This perspective refers to internal business processes. Metrics based
on this perspective allow the managers to know how well their business is
running, and whether its products and services conform to customer
requirements (the mission). These metrics have to be carefully designed
by those who know these processes most intimately. Usually with mis-
sions unique to each organization, these metrics are developed by the
organizations themselves without the help of outside consultants.
In addition to the strategic management process, two kinds of busi-
ness processes may be identified: a) mission-oriented processes, and b)
support-oriented processes. Many processes in government, such as DoD
/ NASA, are mission-oriented processes, and have many unique problems
in measuring them. The support processes are more repetitive in nature.
Hence, they are easier to measure and benchmark using generic metrics.
Maintenance & Reliability (M&R) related examples of this perspec-
tive include:
• PM Backlog—Percent or Number of Tasks
• Percent Schedule Compliance
• Percent Rework
• Percent Reliability (or MTBF)—Asset / System
• Percent Material Delivered or Available on Time

The Customer Perspective


Recent management philosophy has shown an increasing realization
of the importance of customer focus and customer satisfaction in any busi-
ness. These are leading indicators: if customers are not satisfied, they will
eventually find other suppliers that will meet their needs. Poor perform-
ance from this perspective is thus a leading indicator of future decline,
296 Chapter 9

even though the current financial picture may look good. For maintenance
organizations, their customers are the operations. If they are not happy
with the service due to increasing failure rate and downtime, they could
outsource the maintenance.
In developing metrics for satisfaction, customers should be analyzed
in terms of the kinds of customers and processes for which the organiza-
tion is providing a product or service to those groups. M&R related exam-
ples of this perspective include:
• Percent downtime
• Percent availability
• Percent delivery on time (asset/system back to operation as
promised)
• Customer satisfaction with the services maintenance provides,
such as turn-around (no cost overruns, worked right the first time,
asset operates at 100% performance, etc.)

The Financial Perspective


Kaplan and Norton do not disregard the traditional need for financial
data. Timely and accurate financial data will always be a priority, and
managers will do whatever necessary to provide it. In fact, often there is
more than enough handling and processing of financial data. With the
implementation of a corporate database, it is hoped that more of the pro-
cessing can be centralized and automated. But the point is the current
emphasis on financials leads to an unbalanced situation with regard to
other perspectives.
There are two general types of measures that affect the financial out-
come of a business: effectiveness and efficiency. An organization may be
effective in safely producing a good product on-time, but can be out-
flanked by a more efficient competitor. The reverse is also true, i.e., pro-
ducing in an efficient manner but without the quality expectations.
Generally, the effective measures are mastered first, followed by efficien-
cy measures. It is the old struggle between quality and production.
There is perhaps a need to include additional financial-related data,
such as risk assessment and cost-benefit data, in this category. M&R relat-
ed examples of this perspective are:
• Maintenance cost per unit of product or service provided
• Maintenance cost as percent of Replacement Asset Value (RAV)
• Inventory turns (of MRO store)
• MRO store Inventory as a percent of RAV
• Maintenance cost / HP installed
Managing Performance 297

9.4 Data Collection and Data Quality

Another key challenge with a performance measurement system is


data collection and availability of quality data on a timely basis. Data is
the key ingredient in performance measurement. Major factors in estab-
lishing a performance measurement system are:
a. Cost of data collection
b. Data quality
c. Data completeness
d. Extrapolation from partial coverage
e. Matching performance measures to their purpose
f. Understanding extraneous influences in the data
g. Timeliness of data for measures
h. Use of measures in allocation of funding
i. Responsibility for measures, and limited control over the process
j. Benchmarking and targets

Evidently, an efficient and effective data collection system is needed


to ensure availability of quality data. A data collection system should:
1. Identify what data needs to be collected and how much; the pop-
ulation from which the data will come; and the length of time over
which to collect the data.
2. Identify the charts and graphs to be used, the frequency of chart-
ing, various types of comparisons to be made, and the methodol-
ogy for data calculation.
3. Identify the characteristics of the data to be collected. (Attribute
data are items that can be counted—variable data are items that
can be measured.)
4. Identify if existing data sources can be utilized or new data
sources need to be created for new or updated measure of per-
formance. All data sources need to be credible and cost effective.

How good are the metrics? The following questions may serve as a
checklist to determine the quality of metrics and to develop a plan for
improvement:
• Do the metrics make sense? Are they objectively measurable?
• Are they accepted by and meaningful to the customer?
• Have those who are responsible for the performance being meas-
ured been fully involved in the development of this metric?
298 Chapter 9

• Does the metric focus on effectiveness and/or efficiency of the


system being measured?
• Do they tell how well goals and objectives are being met?
• Are they simple, understandable, logical, and repeatable?
• Are the metrics challenging but at the same time attainable?
• Can the results be trended? Does the trend give useful
management information?
• Can data be collected economically?
• Are they available timely?
• Are they sensitive? (Does any small change in the process get
reflected in the metric?)
• How do they compare with existing metrics?
• Do they form a complete set—a balanced scorecard (e.g., ade-
quately covering the areas of learning and growth, internal busi-
ness process, financial, and customer satisfaction)?
• Do they reinforce the desired behavior—today and in the long
haul?
• Are the metrics current (living) and changeable? (Do they
change as the business changes?)

9.5 Benchmarking and Benchmarks

What Is Benchmarking?
Benchmarking is the process of identifying, sharing, and using knowl-
edge and best practices. It focuses on how to improve any given business
process by exploiting topnotch approaches rather than merely measuring
the best performance. Finding, studying, and implementing best practices
provide the greatest opportunity for gaining a strategic, operational, and
financial advantage.
Informally, benchmarking could be defined as the practice of being
modest enough to admit that others are better at something, and wise
enough to try to learn how to match, and even surpass them.
Benchmarking is commonly misperceived as simply number crunching,
site briefings and industrial tourism, copying, or spying. It should not be
taken as a quick and easy process. Rather, benchmarking should be con-
sidered an ongoing process as a part of continuous improvement.
Benchmarking initiatives help blend continuous improvement initiatives
and breakthrough improvements into a single change management sys-
tem. Although benchmarking readily integrates with strategic initiatives
Managing Performance 299

such as continuous improvement, re-engineering, and total quality man-


agement, it is also a discrete process that delivers value to the organiza-
tion on its own.

Types of Benchmarking
Generally there are two types of benchmarking activities. They
include:
1. Internal
2. External
a. Similar industry
b. Best Practice

Internal Benchmarking
Internal benchmarking typically involves different processes or
departments within a plant or organization. This type of benchmarking has
some advantages such as ease of data collection and comparison—some
of the enablers such as employee’s skill level and culture would be gener-
ally similar. However, the major disadvantage of internal benchmarking is
that it is unlikely to result in any major breakthrough in improvements.

External Benchmarking
External benchmarking is performed outside of an organization and
compares similar business processes or best in any industry.
Similar industry benchmarking uses external partners in a similar
industry or with similar processes; it shares their practices and data. This
process may be difficult in some industries, but many organizations are
open to share non-proprietary information. This type of benchmarking ini-
tiative usually focuses on meeting a numerical standard rather than
improving any specific business process. Small or incremental improve-
ments have been observed in this type of benchmarking.
Best Practices benchmarking focuses on finding the best or leader in
a specific process and partnering with them to compare their practices and
data.

Benchmarking Methodology
One of the essential elements of a successful benchmarking initiative
is to follow a standardized process. Choosing an optimal benchmarking
partner requires a deep understanding of the process being studied and of
the benchmarking process itself. Such understanding is also needed to
300 Chapter 9

properly adapt best practices and implement changes to each organiza-


tion’s unique culture. Simply stated, the best practice needs to be tailored
to meet an organization’s culture if it is to be implemented successfully.
This dynamic process often involves finding and collecting internal
knowledge and best practices, sharing and understanding those practices
so they can be used, and adapting and applying those best practices in new
and existing situations to enhance performance levels.
The following steps are recommended for successfully implementing
a benchmarking initiative:

1. Conduct internal analysis.


2. Compare data with available benchmarks.
3. Identify gaps in a specific area.
4. Set objectives and define scope.
5. Identify benchmarking partners.
6. Gather information.
a. Research and develop questionnaire.
b. Plan benchmarking visits.
7. Distill the learning—compile results.
8. Select practice to implement.
9. Develop plan and implement improvements—tailored practice.
10.Review progress and make changes if necessary.

Challenges in Benchmarking: The Code of Conduct


Benchmarking can create potential problems, ranging from simple
misunderstandings to serious legal problems. To minimize the likelihood
of these types of difficulties, it is strongly recommend that the benchmark-
ing teams follow a simple Code-of-Conduct.

Legal
Don’t enter into discussions or act in any way that could be construed
as illegal, either for you or your partner. Potential illegal activities could
include a simple act of discussing costs or prices, if that discussion could
lead to allegations of price fixing or market rigging.

Be Open
Early in your discussion, it helps to fully disclose your level of expec-
tations with regard to the exchange.
Managing Performance 301

Confidentiality
Treat the information you receive from your partners with the same
degree of care that you would for information that is proprietary to your
organization. You may want to consider entering a non-disclosure agree-
ment with your benchmarking partner.

Use of Information
Don’t use the benchmarking information you receive from a partner
for any purpose other than what you have agreed to.

The Golden Rule of Benchmarking


Treat any benchmarking partners and their information the way you’d
like them to treat you and your information.

Lack of Standardized Definitions


One of the challenges in the M&R benchmarking process is the
absence of standardized definitions of M&R terms, including metrics. We
have found from our own experience that during the benchmarking
process, the benchmarking partners usually spend considerable time
learning to understand each other’s terms, including metrics, as well as
what data goes in to satisfy that specific metric. To overcome this chal-
lenge, the Society of Maintenance & Reliability Professionals (SMRP),
has taken the initiative to define and standardize M&R terms. The SMRP
team has undertaken a very rigorous and time-consuming development
process to standardize maintenance and reliability-related terms, and to
obtain feedback from the M&R community to ensure their validity.
There are about two hundred terms which have been defined and
standardized by the SMRP’s best practices team. However, to fill the
needs of a larger M&R community, a new publication called “The
Professional’s Guide to Maintenance and Reliability Terminology” by
Gulati, Kahn and Baldwin was recently released. It is a very comprehen-
sive list of over 3000 definitions and acronyms in the M&R field, includ-
ing project management and quality.

Society for Maintenance & Reliability Professionals (SMRP) Initiative


SMRP’s effort is being carried out by its Best Practices committee.
The committee has been developing definitions for key M&R perform-
ance metrics. Through group consensus and an extensive review by sub-
ject matter experts, including the use of web-based surveys, these metrics
302 Chapter 9

are becoming industry standards. As such, they can be used in benchmark-


ing processes and when searching for best practices. This would help to
create a common language in M&R field which is badly needed now.
The development process used by the SMRP Best Practices commit-
tee is a six-step process:
1. Selection of key metrics
2. Preparation of the metric descriptions
3. Review and consensus by the committee
4. Review and feedback by subject matter experts
5. Final review and editing by the committee
6. Publication

A template was also developed by the best practices / metrics team to


provide a consistent method of describing each metric. The basic elements
of each metric are:
• Title: The name of the metric
• Definition: A concise definition of the metric in easily under
standable terms
• Objectives: What the metric is designed to measure or report
• Formula: Mathematical equation used to calculate the metrics
• Component Definitions: Clear definitions of each of the terms
that are utilized in the metric formula
• Qualifications: Guidance on when to apply and not apply the
metrics
• Sample Calculation: A sample calculation utilizing the formula
with realistic values

Visit SMRP’s website at www.smrp.org to view its current list of the


metrics and to obtain additional information regarding best practices and
metrics. Figure 9.4 is a list of metrics developed by the SMRP’s Best
Practices team.

Benchmarks
A benchmark refers to a measure of best practice performance where-
as benchmarking refers to the actual search for the best practices.
Emphasis is then given to how we can apply the process to achieve supe-
rior results. Thus, a benchmark is a standard, or a set of standards, used as
a point of reference for evaluating performance or quality level.
Benchmarks may be drawn from an organization’s own experience, from
# Metric SMRP Pillar # Metric SMRP Pillar
1 Actual Cost to Planning Estimates #5 - Work Management 36 Planner to Craft Ratio #5 - Work Management
2 Availability #2 - Process Reliability 37 Planning Variance Index #5 - Work Management
3 Condition Based Maintenance Cost #5 - Work Management 38 PM & PcM (CBM) Compliance #5 - Work Management
4 Condition Based Maintenance Hour #5 - Work Management 39 PM & PcM (CBM) Effectiveness #5 - Work Management
5 Continuous Improvement Man Hours #5 - Work Management 40 PM & PcM (CBM) Work Order Backlog #5 - Work Management
6 Contractor Manpower #5 - Work Management 41 PM & PcM (CBM) Work Order Overdue #5 - Work Management
7 Corrective Maintenance Cost #5 - Work Management 42 PM & PcM (CBM) Yield #5 - Work Management
8 Corrective Maintenance Hours #5 - Work Management 43 Preventive Maintenance (PM) Cost #5 - Work Management
9 Craft Workers on Shift #5 - Work Management 44 Preventive Maintenance (PM) Hour #5 - Work Management
10 Emergency Purchase Orders #5 - Work Management 45 Proactive Work #5 - Work Management
11 Idle Time #2 - Process Reliability 46 Ratio of Indirect to Direct Maintenance Personnel #5 - Work Management
Figure 9.4
12 Inactive Stock #5 - Work Management 47 RAV $ per Maintenance Craft Head Count #1 - Business Management
List of 13 Indirect Maintenance Personnel Cost #5 - Work Management 48 Reactive Work #5 - Work Management
SMRP 14 Indirect Stock #5 - Work Management 49 Ready Backlog #5 - Work Management
15 External Maintenance Personnel Cost #5 - Work Management 50 Schedule Compliance #5 - Work Management
Metrics
16 Maintenance Cost as % of Asset Replacement Value (RAV) #1 - Business Management 51 Schedule Compliance - Work Orders #5 - Work Management
17 Maintenance Cost per Unit of Production #1 - Business Management 52 Schedule Downtime #3 - Equipment Reliability
18 Maintenance Margin #1 - Business Management 53 Standing Work Orders #5 - Work Management
19 Maintenance Material Cost #5 - Work Management 54 Stock Outs #5 - Work Management
20 Maintenance Shutdown Cost #5 - Work Management 55 Store Value as % of RAV #1 - Business Management

Managing Performance 303


21 Maintenance Training Cost as % of Total #4 - People Shifts 56 Storeroom Records #5 - Work Management
22 Maintenance Training Hours as % of Total #4 - People Shifts 57 Storeroom Transactions #5 - Work Management
23 Maintenance Training ROI #4 - People Shifts 58 Stores Inventory Turns #5 - Work Management
24 MDT #3 - Equipment Reliability 59 Supervision to Craft Ratio #5 - Work Management
25 MTRF #3 - Equipment Reliability 60 Systems Covered by Criticality Analysis #3 - Equipment Reliability
26 MTRM #3 - Equipment Reliability 61 Total Downtime #3 - Equipment Reliability
27 MTTF #3 - Equipment Reliability 62 Total Operating Time Ratio #2 - Process Reliability
28 MTTR #3 - Equipment Reliability 63 Unplanned Work #5 - Work Management
29 OEE #2 - Process Reliability 64 Unscheduled Downtime #3 - Equipment Reliability
30 OEE + 24x7 #2 - Process Reliability 65 Uptime #2 - Process Reliability
31 Overtime Maintenance Cost #5 - Work Management 66 Utilization Rate #2 - Process Reliability
32 Overtime Maintenance Hours #5 - Work Management 67 Vendor Management Stock #5 - Work Management
33 Planned Backlog #5 - Work Management 68 Work Order Aging #5 - Work Management
34 Planned Work #5 - Work Management 69 Work Order Cycle #5 - Work Management
35 Planner Effectiveness #5 - Work Management 70 Wrench Time #5 - Work Management
304 Chapter 9

Figure 9.5 Maintenance and Reliability Best Practices Key Benchmarks


Managing Performance 305

the experience of others in the industry, or from regulatory requirements


such as those from Environmental Protection Agency (EPA) or
Occupational Safety and Health Agency (OSHA).
If we were to benchmark world conquest, what objective measure
would we use to compare Julius Caesar to Alexander the Great, or
Genghis Khan to Napoleon? Which of them was the epitome, and why?
We do the same thing in business. Which organization has the best PdM
program? Who provides the most responsive customer service depart-
ment? Who is the best in planning/scheduling? What about the leanest
manufacturing operation? And how do we quantify that standard?
Figure 9.5 lists some of the key maintenance and reliability best prac-
tices benchmarks.

9.6 Summary

Performance measurement is a means of assessing progress against


stated goals and objectives in a way that is quantifiable and unbiased. It
brings with it an emphasis on objectivity, consistency, fairness, and
responsiveness. At the same time, it functions as a reliable indicator of an
organization’s health. Its impact on an organization can be both immedi-
ate and far-reaching.
Performance measurement asks the question, “What does success
really mean?” It views accomplishment in terms of outputs and outcomes,
and it requires us to examine how operational processes are linked to
organizational goals. If implemented properly, performance measures—
metrics—are evaluated not on the basis of the amount of money that is
spent or the types of activities performed, but on whether the organization
has produced real, tangible results.
The real objective of metrics should be to change the behavior so that
people do the right things. The secondary objective is to determine the
health of the process or assets being monitored. A metric is nothing more
than a standard measure to assess performance in a particular area.
However there is an imperative need to ensure that the right things are
being measured.
Performance indicators can be leading or lagging. The purpose of
using these indicators is to measure the performance of the process or
asset and to help identify where the process is working well and where it
is not. Leading and lagging indicators provide information so that positive
trends can be reinforced and unfavorable trends can be corrected through
306 Chapter 9

process changes. Leading indicators measure the process and predict


changes and future trends. Lagging indicators measure results and con-
firm long-term trends. Whether an indicator is a leading or lagging indi-
cator depends on where in the process the indicator is applied. Lagging
indicators of one process component can be a leading indicator of anoth-
er process component.
A benchmark is a measure of best practice performance.
Benchmarking refers to the search for the best practices that yields the
benchmark performance with emphasis on how can we implement the
best practice to achieve superior results.
Finally, developing a performance measurement system that pro-
vides feedback relative to an organization’s goals and supports it in
achieving these goals efficiently and effectively is essential. A successful
performance measurement system should:

• Comprise a balanced set of a limited vital few measures


• Produce timely and useful reports at a reasonable cost based on
quality accurate data
• Disseminate and display information that is easily shared, under-
stood, and used by all in the organization
• Help to manage and improve processes and document
achievements
• Support an organization’s core values and its relationship with
customers, suppliers, and stakeholders

9.7 Self Assessment Questions

Q9.1 Why do we need a performance measurement system? What


are the benefits of such a system ?

Q9.2 What are the benefits of benchmarking?

Q9.3 Explain what is meant by a “World Class” benchmark?

Q9.4 What are the key attributes of a metric?

Q9.5 Explain leading and lagging metrics.

Q9.6 What types of metrics show results?


Managing Performance 307

Q9.7 Explain the Balanced Scorecard model.

Q9.8 Explain the different types of benchmarking. What are the


benefits of external benchmarking?

Q9.9 Discuss data collection and quality issues. How can we


improve data quality?

Q9.10 List five metrics that can be used to measure overall plant
level performance of maintenance activities. Discuss the rea-
son for your selection.

9.8 References and Suggested Reading

Camp, Robert. Benchmarking: The Search for Industry Best Practices


That Lead to Superior Performance. Productivity Press, 1989,
2006.
Kaplan, Robert. The Balanced Scorecard: Translating Strategy into
Action. Harvard Business Press, 1996.
Mitchell, John. Physical Asset Management Handbook. 4th edition.
Clarion Technical Publishing, 2007.
Wireman, Terry. Benchmarking: Best Practices. Industrial Press, 2005.
Wireman, Terry. Developing Performance Indicators for Managing
Maintenance. 2nd Edition. Industrial Press, 2005.
Chapter 10
Workforce Management

What I hear, I forget. What I see, I remember. What I do,


I understand.
— Kung Fu Tzu (Confucius)

10.1 Introduction
10.2 Key Terms and Definitions
10.3 Employee Life Cycle
10.4 Understanding the Generation Gap
10.5 Communication Skills
10.6 People Development
10.7 Resource Management and Organization Structure
10.8 Measures of Performance
10.9 Summary
10.10 Self Assessment Questions
10.11 References and Suggested Reading

After reading this chapter, you will be able to understand:

• Deming’s view to improve organizational effectiveness


• Employee life cycles
• Generation gaps and the aging workforce
• Communication issues

309
310 Chapter 10

• People development-related issues


• Training types and benchmarks
• Diversity challenges
• Challenges in managing the workforce, including organization
structures and outsourcing

10.1 Introduction

People make it happen. They get things done. We may have great
plans and the best processes, but if we don’t have the people available
with the right skills, these plans and processes can’t be implemented or
carried out effectively. Developing people—the workforce—and empow-
ering them to give their best is key to defining the difference between an
ordinary company and a great organization. In addition, the right process-
es and technology must be in place to nurture and harness the potential of
human capital.
Maintenance and reliability processes are no different than any other
processes in the workplace. Organizations that are considered to be the
“Best of the Best” or “World Class” use many of the same key principles.
Dr. W. Edwards Deming, a world-renowned expert in the field of quality,
gave us the following 14 principles, which any organization can use to
improve its effectiveness.

1. Create constancy of purpose towards improvement.


Replace short-term reaction with long-term planning. Aim to become
competitive and stay in business.

2. Adopt the new philosophy.


We are in a new economic age. Management must awaken to the chal-
lenge, must learn their responsibilities, and take on leadership for change
rather than merely expect the workforce to do so.

3. Cease dependence on inspection.


If variation is reduced, there is no need to inspect manufactured items
for defects, because there won’t be any. Build quality into the product in
the first place. Inspection is not a value-added activity.
Workforce Management 311

4. End the practice of awarding business on the basis of price tag


(minimum cost).
Move towards a single supplier for any one item. Multiple suppliers
mean variation between feedstocks. Build a long-term relationship and
trust with suppliers.

5. Improve constantly and forever.


Strive constantly to improve the system of production and service, to
improve quality and productivity, and thus constantly to decrease costs.

6. Institute training on the job.


If the people—the workforce—are not adequately prepared and
trained to do the job right, they will introduce variation and defects. New
skills are required to keep up with changes in materials, methods, prod-
ucts, and services.

7. Institute leadership.
There is a distinction between leadership and mere supervision. The
latter is quota and target based. The aim of supervision should be to help
people, machines, and gadgets to perform a better job.

8. Drive out fear.


Deming sees management by fear as counter-productive in the long
term because it prevents people from acting in the organization’s best
interests.

9. Break down barriers between departments.


All in the organization working in research, design, sales, opera-
tions/production, and maintenance must work as a team, to foresee oper-
ations/production problems and customer satisfaction issues that may be
encountered with the product or service. The organization should build the
concept of the internal customer that each department serves—not the
management, but the other departments that receive its outputs.

10. Eliminate slogans and exhortations.


It’s not people who make most mistakes; it’s the process they are
working within. Eliminate the use of slogans, posters, and exhortations for
the workforce, demanding zero defects or zero breakdowns and new lev-
els of productivity, without providing improved methods. Harassing the
312 Chapter 10

workforce without improving the processes they use is counter-produc-


tive.

11. Eliminate arbitrary numerical targets.


Eliminate work standards that prescribe quotas for the work force and
numerical goals for people in management. Substitute aids and helpful
leadership in order to achieve continual improvement of quality and pro-
ductivity.

12. Permit pride of workmanship.


Remove the barriers that rob people of the pride of workmanship. The
responsibility of managers, supervisors, and foremen must be changed
from sheer numbers to quality. Allow employees to see their end product.
Doing so creates ownership and helps to generate new ideas for improve-
ment.

13. Institute education and self-improvement.


Institute a vigorous program of education, training, and encouraging
self improvement for everyone. What an organization needs is not just
good people; it needs people who are improving with education.
Advances in competitive position in the market place will have their roots
in knowledge.

14. The transformation is everyone’s job.


Create a structure in top management that will push every day on the
preceding points. Take action in order to accomplish the transformation.
Support is not enough; action is required. Put everybody in the organiza-
tion to work to accomplish the transformation.

Most of Dr. Deming’s principles relate to the workforce, including


management and their role in acquiring, preparing, and educating the
workforce as well as improving the processes to get productivity gains.
These points can be applied to any organization, small to large, and to any
industry (services, manufacturing, etc.).
In this chapter, we will be discussing people—the workforce. What do
we need to do to hire the right people, prepare them with the right skill
sets, and then retain them to perform their job effectively? This topic is a
very important challenge in today’s global economy and demographical-
ly diverse workforce.
Workforce Management 313

10.2 Key Terms and Definitions

Baby boomers
The generation born following World War II. They are idealists
in nature and usually very loyal to their organization. They feel
a sense of belonging and dedication based on their history. They
are motivated by power and prestige, learning opportunities, and
long-term benefits.

Certification
The result of meeting the established criteria set by an accredit-
ing or certificate-granting organization.

Communication
Effective transfer of information from one party to another;
exchanging information between individuals through a common
system of symbols, signs, or behavior.

Diversity (Workplace)
Workplace diversity refers to the variety of differences between
people in an organization. Diversity encompasses race, gender,
ethnic group, age, personality, cognitive style, tenure, organiza-
tional function, education, background, and more.

Employee Involvement
A practice within an organization whereby employees regularly
participate in making decisions on how their work areas operate,
including making suggestions for improvement, planning, goal
setting, and monitoring performance.

Employee Life Cycle


The entire useful life of an employee from hiring to retiring.
During this time, a person learns new skills and becomes a pro-
ductive employee until retirement.

Generation Xers
People born primarily from the mid-1960s through the 1970s.
They are very ambitious and independent in nature and strive to
balance the competing demands of work, family, and personal
life.
314 Chapter 10

Generation Yers
People born after 1980, also known as the millennial generation.
They are technologically savvy with a positive, can-do attitude.

Job Task Analysis


An analysis of the various elements pertaining to skill require-
ments, training and experience, mental and physical demands,
working conditions, hazards exposure, and requirements of
responsibilities of performance for a particular job.

Organization Structure (Chart)


Diagram or graphic representation of an organization that
shows, to varying degrees, the functions, responsibilities, peo-
ple, authority, and relationships among these.

Qualified People
People who because of their knowledge, training, qualifications,
certification, or experience, are competent to perform the duties
of their job.

Succession Planning
A key element of the workforce development process. It identi-
fies and prepares suitable employees through mentoring, train-
ing, and job rotation to replace key players in the organization.

10.3 Employee Life Cycle

An employee life cycle covers the steps employees go through from


the time they are hired into an organization until they leave. Human
Resource professionals often focus their attention on the steps in this
process in hopes of making an impact on the organization’s bottom line.
Normally, their goal is to reduce the organization’s cost per employee
hired, which theoretically is a good thing. However, this should not be the
only goal because they aren’t the ones who can make them stay willingly
and be productive with the organization. The managers for whom newly-
hired employees are going to work can make the real difference in
employees wanting to stay and their adding value to the organization.
Workforce Management 315

On a practical day-to-day basis, employees don’t really work for an


organization; they work for somebody—the boss. To the extent that there
can be good bosses, they can keep employees motivated, happy, and pro-
ductive, thereby reducing the costs associated with employee turnover. In
turn, managers can make their employees’ jobs easier and increase value
to the organization.
One important job of a manager is to ensure that employees under-
stand the key purpose of the organization. For example, from day one, all
Disney employees are told that they (Disney) create happiness in their
guests. Also, Disney calls their customers their guests. Guest is a better
word and has a deeper meaning than customer. It suggests more intimacy,
closeness to that person. These kind of practices help instill the right cul-
ture in an employee from their first day.
Employees are one of an organization’s largest expenses these days.
Unlike other major capital costs such as buildings, machinery, and tech-
nology, human capital is highly volatile. Managers are placed into key
positions to reduce that volatility by reducing the overall life cycle cost of
employees in the organization. This life cycle consists of four steps:

• Hire
• Inspire
• Admire
• Retire

Hire
This first step is probably the most important. It is important to hire
the best people we can find. This is not a time to be cheap. The cost of
replacing a bad hire far exceeds the marginal additional cost of salary or
expenses needed to hire the best person in the first place. Hire talent, not
just trainable skills. Skills can be taught to a talented employee.
Make your organization a place people want to come to and work for.
An organization’s culture can be a powerful recruiting tool. Ensure that
the new hire understands the goals your department or organization wants
to achieve.

Inspire
Once we have recruited the best employees to come to work on our
team, the hard part begins. We have to inspire them to perform to their
best abilities. We have to challenge and motivate them. That is where we
316 Chapter 10

will get their best effort and creativity that will help the organization
excel.
Make them welcome. Make them feel like part of the team from day
one. Set goals for them that are hard, but achievable. Be a leader, not just
a manager. It is a good practice to assign a mentor to help them get
acquainted with the new environment.
Mentoring is another good practice to implement during early stages
for new employees to help them adjust to a new environment. It is a good
practice to assign a senior person in the area to mentor new employees.

Admire
Once we have hired the best employees and have challenged and
motivated them, our job to inspire them continues. The biggest mistake is
made when a manager ignores them. As soon as we start to slack off, their
satisfaction and motivation decreases. If we don’t do something, they will
become disenchanted and will leave. They will become part of the
“employee turnover” statistics that need to be avoided.
We want TGIM (thank goodness, it’s Monday) employees, not TGIF
(thank goodness, it’s Friday) ones.
Give employees positive feedback as much as possible, even if it’s
just a few good words. Provide appropriate rewards and recognition for
jobs well done. Provide them additional training to develop new skill sets.

Retire
The time when somebody retires after a long service is when we know
that we have been successful. When employees see the organization as the
employer of choice, they will come and work for us. When they recognize
us as a good boss and a real leader, they will stay around. As long as we
continue to inspire, motivate, and challenge them, they will continue to
contribute at the high levels the organization needs in order to beat the
competition. They will be long-term employees; even staying with the
organization until they retire. They will refer other quality employees,
including their relatives. Organizations will attract and retain second and
even third generation loyal employees.
Along the way, the organization will have had some of the most cre-
ative and productive employees with the lowest employee costs in the
market.
Workforce Management 317

10.4 Understanding the Generation Gap

There is a shortage of skilled industrial workers, the people who oper-


ate and maintain sophisticated assets/systems on the plant floor, as well as
engineering and management professionals. This shortage poses a serious
threat to our industrial competitiveness. The shortage lies in the demo-
graphics as most veterans have left the workplace and baby boomers are
getting close to their retirement. Predominately four generations of work-
ers co-exist in today’s workplace. The following list reflects William
Strauss and Neil Howe’s organization from their book Generations:

Silent Born 1925–1942


Baby Boom Born 1943–1960
Generation X Born 1961–1981
Generation Y Born 1982–2001

The differences between the generations create many challenges in


the workplace. These challenges, which can be both negative and positive,
often relate to variations in perspective and goals as a result of genera-
tional differences. This area gets further complicated because of the age
differences between managers and employees. Organizations can’t
assume that people of varying ages will understand each other or have the
same perspectives and goals. In order to be successful, there is a need to
understand and value the generational differences and perspectives and
turn those negatives into positives.
Each group’s talent and most common differences are discussed in the
following sections:

Silent
The Silent generation is adaptive in nature and has a very intense
sense of loyalty and dedication. They are also known as Traditionalist.
Many of them have been puppets, paupers, or pirates. As a result, they
have a wealth of knowledge to contribute to any organization. Some in
this group are not leaving the workforce yet. It is important to note that
some can’t leave due to financial reasons. Others love their work, are still
in good health, and want to continue to contribute to the organization.
The work ethic of the Silent Generation is built on commitment,
responsibility, and conformity as tickets to success. Most came of age too
late to be the heroes of World War II, yet too early to participate in the
318 Chapter 10

youthful rebelliousness of the 1960s “Consciousness Revolution.” On the


job, they are not likely to “rock the boat,” break the rules, or disrespect
authority. Tempered by war, a command and control approach comes nat-
urally.
The challenge for leaders is how to empower Silent employees—to
find out what motivates them and then tailor the job to each individual’s
needs. They still want respect for contribution and longevity. It’s also cru-
cial that leaders don’t let them think that their time has come, that they
should go. They have a loyalty and commitment that is increasingly hard
to find in younger employees.
It would be very beneficial for the organization to team up a seasoned
Silent employee with a Generation Xer or Yer on a project to foster knowl-
edge transfer both ways. Veterans bring the institutional and historical
experience whereas Xers and Yers often bring the technological and inno-
vational savvy. Also, Silent employees can be trained in new technology.
Some leaders believe that training them on a new technology or process
is a bad investment due to their age. This type of thinking doesn’t add
value to the organization. Due to their hard-working attitudes, they can be
just as easily trainable and probably more accepting of this new training
than other generations.

Baby Boomers
Boomers, idealists in nature, have always been seen as loyal to their
organizations. They feel a sense of belonging and dedication based on
their history. The Baby Boom Generation changed the physical and psy-
chological landscape forever. As products of “the Wonder Years,” they
were influenced by the can-do optimism of John F. Kennedy and the hope
of the post-World War II dream. But the intense social and political
upheaval of Vietnam, assassinations, and civil rights, led them to rebel
against conformity and to carve a perfectionist lifestyle based on person-
al values and spiritual growth. They welcome team-based work, especial-
ly as an anti-authoritarian declaration to “The Silents” ahead of them, but
they can become very political when their turf is threatened. Rocked by
years of reorganizing, reengineering, and relentless change, they now long
to stabilize their careers.
One of the most common complaints Boomers make about Gen Xers
and Gen Yers is that “they don’t have the same work ethic.” This does not
mean that they are not hardworking employees, but it does mean that they
place a different value and priority on work.
Workforce Management 319

Boomers are motivated by:


• Power and prestige. Boomers are often traditionalists, and
perks of the position matter. They want titles and authority com-
mensurate with responsibility.
• Networking and Learning. Boomers want to participate in
associations and conventions that keep them professionally con-
nected to their peers. Boomers are motivated by working togeth-
er on professional projects in affiliation with others like them.
• Long-Term Benefits. Boomers are interested in compensation
that is more long term, such as profit sharing and health care
benefits including long-term care.

Generation (Gen) Xers


In general, Gen Xers are those born during the 1960s and 1970s. This
generation, therefore, includes a range of employees in their thirties, for-
ties, and early fifties.
The Gen Xers, often-maligned, so named because no one could settle
on an accurate definition, are characterized by an economic and psycho-
logical “survivor” mentality. They grew up very quickly amid rising
divorce rates, latchkeys, violence, and low expectations. They entered the
job market in the wake of the Boomers, only to be confronted with new
terms like “downsizing” and “RIFs” (reductions-in-force) as the economy
plunged into recession. It’s hardly surprising then that they tend to be
skeptical toward authority and cautious in their commitments. Their self-
reliance has led them, in unprecedented numbers, to embrace “free
agency” over organizational loyalty. Ambitious and independent, they’re
now striving to balance the competing demands of work, family, and per-
sonal life.
Gen Xers bring new challenges as well as new ideas into the labor
market. They want and demand benefits such as stock option plans, health
care insurance, and time off (paid vacation, sick days, personal leave
days). They tend to be less motivated by promises of overtime pay and
more motivated by personal satisfaction with their jobs. The number one
benefit for Generation Xers is training/development opportunities to
enhance/further their careers. They want to grow in their jobs and learn
new skills.
Gen Xers do not anticipate staying with one job or company through-
out their entire career. They have seen their parents laid off. Many of them
have grown up in divorced family situations. They expect to change jobs
320 Chapter 10

as they seek employment that offers them both better benefits and more
opportunity for professional growth as well as personal fulfillment. They
want, and expect, their employers to hear what they have to say. They
have an interest in understanding the “big picture” for the organization
and how this influences their employment and growth. They are less like-
ly to accept a “because I said so” attitude from a supervisor.
Some of the ways to motivate Gen Xers to maximize productivity:

Take time to be personal.


Thank an employee for doing a good job in person, in writing, or both.
Listen to what employees have to say, both in a one-on-one situation
and in a group meeting.

Encourage growth.
Provide feedback on the employee’s performance. Make sure the
employee understands expectations. Involve them in the decision-
making process whenever possible.

Provide Training.
Pay for employees to attend workshops and seminars. Provide on-site
classes where employees can learn new skills or improve upon old
ones. Challenge them.

Motivate through Rewards and Promotions.


Recognize employees who have done an outstanding job by giving an
unexpected reward, such as a day off or a free dinner for employees
and their family at a nice restaurant.

Encourage Teamwork.
Provide opportunities for collaboration and teamwork. This genera-
tion “fuels their fire” through teamwork.

Create Ownership.
Help employees understand how the business operates. They need to
experience a sense of ownership. Encourage this by providing them
with information about new products or services, advertising cam-
paigns, strategies for competing, etc. Let each employee see how he
or she fits into the plan and how meeting their goals contribute to
meeting the organization’s objectives.
Workforce Management 321

Build morale.
Have an open work environment; encourage initiative, and welcome
new ideas. Don’t be afraid to spend a few dollars for such things as
free coffee or tea for employees, or ordering a meal for employees
who have to work overtime. Take time to speak with an employee’s
spouse or family when you meet them and let them know you appre-
ciate the employee. Gen Xers look for more than just fair pay; they
need and want personal acknowledgment and job satisfaction.

Generation (Gen) Yers


In general, Gen Yers are those born after 1980. They are also known
as Echo-Boom or Millennial generation. This generation, therefore,
includes a range of employees from those approaching their late 20s and
all the way to older teenagers just entering the job market. Coming of age
during a shift toward virtue and values, they’re attracted to organizations
whose missions speak to a purpose greater than a bottom line. They’re
technologically savvy with a positive, can-do attitude that says, “I’m here
to make a difference.” And they will.
To motivate Gen Yers, consider the following suggestions:

Work Schedule Flexibility.


Provide Gen Yers flexibility in when and where work is done. Gen
Yers resist what they see as rigid workday starting times. They do not
understand why coming to work 15–30 minutes late is viewed by
Boomers as irresponsible behavior.

Change and Challenges.


Gen Yers are interested in change and challenge. They will leave a
higher-paying good job for the opportunity to experience something new.
They do not see their careers as needing to be linear. They stay in a par-
ticular job often no more than two to three years.
Do not interpret their rebellious nature as negative. Caution!
Sometime they need to vent out; let them do it. Do not take it personally.
This generation will challenge and change much of what we need to
change.
Although Gen Xers and Gen Yers are motivated by different things,
both age groups need the following:
• Frequent communication, including being told the “why” and
not just the “what” of projects and priorities.
322 Chapter 10

• To be included, and not just in what affects them most directly.


• To have fun at work.

In order for an organization to be truly successful, all co-existing


generations in the workplace need to understand and value each other,
even when their perspectives and goals are different.
As Silents, Baby Boomers, Gen Xers, and Gen Yers intersect in the
workplace, their attitudes, ethics, values, and behaviors inevitably col-
lide. Nearly 70% of participants in a recent Web poll say they’re experi-
encing a “generational rift.”
Organizations need to look beyond this clash of the generations for
ways to leverage multi-generational perspectives to their benefit. If we
take the time to master a few tools and strategies for communicating
across generations, we’ll be better positioned to tap the best that each
generation as well as each individual brings to the workplace. Here are
some suggestions:
• It’s not what we say, but how we say it. Generational clashes often
stem from miscommunications in tone or style. The Silents, for
example, are aware that they might be technologically chal-
lenged; empathy is a better strategy than decision. The younger
generations, in general, might have shorter attention spans than
their seniors, so they may prefer verbal training to reading docu-
ments.
• Understand the different generational motivations. Gen Xers may
seem to be less driven, and Baby Boomers managing Gen Xers
should know that money usually isn’t the motivating force. It’s
quality of life. We should look for ways to support Gen Xers’ bal-
anced lifestyle.
• Look beyond appearances. Benefit from diverse opinions. Keep
an open mind about attitudes. Adapt style to the realities of
today’s workplace.

We’re living through profound changes in the business world.


Traversing this generational landscape, bolstered by new learning and
respect for differing ideals about the workplace, will get the job done bet-
ter and faster. Look for what unites us with our peers. We will be better
prepared to welcome the generation that comes next.
Each generation has complained about those in other groups. So the
fact that there are differences in the generations is nothing new. What is
Workforce Management 323

new today is the magnitude of the differences. It is time to understand and


value this diversity so that we can benefit from it. Failing to do this can
result in failure for everyone.

10.5 Communication Skills

Why Communication Is Important


People in organizations typically spend over 75% of their time in
interpersonal situations to get work done. It is no surprise to find that poor
communications are at the root of a large number of organizational prob-
lems. Effective communication is an essential component of organization-
al success whether it is at the interpersonal, inter-group, intra-group, orga-
nizational, or external level.
Effective communication is all about conveying messages to other
people clearly and unambiguously. It’s also about receiving information
that others send to us, with as little distortion as possible.
Expressing our feelings, thoughts, and opinions clearly and effective-
ly is only half of the communication process needed for interpersonal
effectiveness. The other half is listening and understanding what others
communicate to us. When we decide to communicate with another person,
we do it to fulfill a need. In deciding to communicate, we select the
method or code which we believe will effectively deliver the message to
the other person. The code used to send the message can be either verbal
or nonverbal. When the other person receives the coded message, they go
through the process of decoding or interpreting it into understanding and
meaning.
Effective communication exists between two people when the receiv-
er interprets and understands the sender’s message in the same way the
sender intended it. By successfully getting the message across, we convey
our thoughts and ideas effectively. When not successful, the thoughts and
ideas that we actually send do not necessarily reflect what we think, caus-
ing a communications breakdown and creating roadblocks that stand in
the way of our and the organization’s success—both personally and pro-
fessionally.

Communication Process
Problems with communication can be at every stage of the communi-
cation process (see Figure 10.1):
• Sender (source)
324 Chapter 10

• Encoding
• Channel—media
• Decoding
• Receiver
• Feedback

At each stage, there is the potential for misunderstanding and confu-


sion. To be effective communicators and to get our point across without
misunderstanding and confusion, our goal should be to reduce the fre-
quency of problems at each stage of this process, with clear, concise,
accurate, well-planned communications. The key elements of communi-
cation process as shown in Fig. 10.1 are:

Figure 10.1 Communication process

• Sender. We, the senders, the source of the message, need to be


clear about why we’re communicating, and what we want to com-
municate. In addition, the senders need to ensure that the informa-
tion being communicated is useful and accurate.
• Message. The message is the information that we want to
communicate.
• Encoding. This is the process of transferring the information we
want to communicate into a form that can be sent and correctly
decoded at the other end. Successful encoding depends partly on
our ability to convey information clearly and simply, but also to
anticipate and eliminate sources of confusion. For example, cul-
tural issues, wrong assumptions, and missing information can
interfere with clear communication. Failure to understand whom
we are communicating with will result in delivering messages
that are misunderstood.
Workforce Management 325

• Channel. We convey our message through the channel or media.


Channels can be verbal, including face-to-face meetings, tele-
phone, and videoconferencing. Written channels include letters,
emails, memos, and reports.
• Decoding. In order to understand the message correctly, we must
decode it, which involves taking time to read the message careful-
ly or listening actively. Just as confusion can arise from encoding
errors, it can also arise from decoding errors. This is particularly
the case if the decoder doesn’t have enough knowledge to under-
stand the message. Listening is a very important skill which will
be discussed in more detail in the next section.
• Receiver. The receiver is the person for whom the message was
intended. Caution: To be a successful communicator, we should
consider how the receiver will react to the message.
• Feedback. Feedback is what the receiver sends back to us as ver-
bal and nonverbal reactions to our communicated message. We
need to pay close attention to this feedback, as it is the only thing
that can give us assurance the audience has understood our mes-
sage. If we find that there has been a misunderstanding, at least
now we have the opportunity to send the message a second time
to clarify the issue. Part of the feedback process involves under-
standing and predicting how the other person will react. We need
to understand ways that we respond to feedback, especially
threatening feedback. Feedback provides the basic need to
improve, and to be accurate. It can be reinforcing, if given prop-
erly. It is crucial that we realize how critical feedback can be; it
should always be appreciated.

Importance of Listening

“We were given two ears but only one mouth,


that we may hear the more and speak the less.”
—Greek philosopher Xeno

Listening is one of the most important skills we can have. How well
we listen has a major impact on our job effectiveness, and on the quality
of our relationships with others. We listen to:
• Obtain information
• Understand
326 Chapter 10

• Enjoy
• Learn

There are three basic listening modes:


• Competitive Listening. This happens when we are more interest-
ed in promoting our own point of view than in understanding or
exploring someone else’s view. We listen either for openings to
take the floor, or for flaws or weak points we can attack. We don’t
pay any attention to the “sender,” but instead are formulating our
rebuttal plan.
• Passive or Attentive Listening. In this mode, we are genuinely
interested in hearing and understanding the other person’s point
of view. We are attentive and passively listen.
• Active Listening. In active listening, we are genuinely interested
in understanding what the other person is thinking, feeling, or
wanting, and what the message means. We actively check out our
understanding before we respond with our own new message.
Listening is a skill that we can all benefit from improving. By
becoming better listeners, we will improve our productivity, as
well as our ability to influence, persuade, and negotiate. What’s
more, we’ll avoid conflict and misunderstandings—all necessary
for workplace success.

Becoming an Active Listener


There are five elements of active listening. They all help us ensure
that we hear the other person, and that the other person knows we are
hearing what they are saying.

1. Pay attention.
Give the speaker undivided attention and acknowledge the message
.
• Look at the speaker directly.
• Put aside distracting thoughts.
• “Listen” to the speaker’s body language.
• Refrain from side conversations and avoid being distracted by
environmental factors.

2. Show that we are listening.


Use body language and gestures to convey that we are listening.
Workforce Management 327

• Nod occasionally.
• Smile and use other facial expressions.
• Encourage the speaker to continue with small verbal comments
like “yes” and “uh huh.”

3. Provide feedback.
Our personal filters, assumptions, judgments, and beliefs can distort
what we hear. As a listener, our role is to understand what is being said.
This may require us to reflect what is being said and ask questions.
• Reflect what has been said by paraphrasing. “What I’m hearing
is…” and “Sounds like you are saying…” are great ways to reflect
back.
• Ask questions to clarify certain points. “What do you mean when
you say…?” or “Is this what you mean?”

4. Defer judgment.
Interrupting is a waste of time. It frustrates the speaker and limits full
understanding of the message. If you don’t agree with the speaker’s view,
wait till you hear the full story or till the end during the question and
answer period to bring your viewpoint. It’s the speaker who is speaking.

• Allow the speaker to finish.


• Don’t interrupt with counterarguments.

5. Avoid negative mannerisms, but respond.


Everyone has a mannerism. If a mannerism is encouraging and brings
positive response, do it often. Unfortunately, some mannerisms are nega-
tive or distracting. These should be avoided.

• Tapping a pencil /pen or playing with a rubber band or other


objects
• Continually looking at the watch or clock
• Reading book or reports or texting/emailing
• Displaying arrogance or lack of interest

Active listening is a model for respect and understanding. It takes a


lot of concentration and determination to be an active listener. Old habits
are hard to break; it requires a lot of self-discipline to do it.
328 Chapter 10

Effective Team Meetings


Meetings are wonderful tools for generating ideas, expanding on
thoughts and managing group activity. But this face-to-face contact with
team members and colleagues can easily fail without adequate preparation
and leadership. Effective communication is an enabler to make meetings
successful.
To ensure everyone involved has the opportunity to provide their
input, establish meeting rules—where, when, and how long. This will
allow all participants the time needed to adequately prepare for the meet-
ing.
Once a meeting time and place have been chosen, ensure that the
leader or coordinator is available for questions that may arise as partici-
pants prepare for the meeting. As the meeting leader, make a meeting
agenda, complete with detailed notes.
In these notes, outline the goal and proposed structure of the meeting,
and share this with the participants. This will allow all involved to prepare
and to come to the meeting ready to work together to meet the goals at
hand.
The success of the meeting depends largely on the skills displayed by
the meeting leader. To ensure the meeting is successful, the leader should:

• Issue an agenda.
• Start the discussion and encourage active participation.
• Work to keep the meeting at a comfortable pace—not moving too
fast or too slow.
• Summarize the discussion and the recommendations at the end of
each logical section.
• Ensure all participants receive minutes promptly.

Managing Meetings
Choosing the right participants is key to the success of any meeting.
Make sure all participants can contribute; choose good decision-makers
and problem-solvers. Try to keep the number of participants to a maxi-
mum of 12, preferably fewer. Make sure the people with the necessary
information for the items listed in the meeting agenda are the ones who
are invited.
As a meeting leader, work diligently to ensure everyone’s thoughts
and ideas are heard by guiding the meeting so that there is a free flow of
debate with no individual dominating and no extensive discussions
between two people. As time dwindles for each item on the distributed
Workforce Management 329

agenda, it will be necessary to stop the discussion, then quickly summa-


rize the debate on that agenda item, and move on the next item on the
agenda.
When an agenda item is resolved or action is agreed upon, make it
clear who in the meeting will be responsible for this. In an effort to bypass
confusion and misunderstandings, summarize the action to be taken and
include this in the meeting’s minutes.
Meetings are notorious for eating up people’s time. Here are some
ways of ensuring that time is not wasted in meetings:

• Start on time.
• Don’t recap what you’ve covered if someone comes in late. It
sends the message that it is OK to be late for meetings, and it
wastes everyone else’s valuable time.
• State a finish time for the meeting and don’t over-run. If needed
to over-run, ensure everyone is available to stay for a stated peri-
od.
• To help stick to the stated finish time, arrange your agenda in
order of importance so that if you have to omit or rush items at the
end to make the finish time, you don’t omit or skimp on impor-
tant items.
• Finish the meeting before the stated finish time if you have
achieved everything you need to.

Minutes record the decisions of the meeting and the actions agreed.
They provide a record of the meeting and, more important, they provide a
review document for use at the next meeting so that progress can be meas-
ured—this makes them a useful disciplining technique as individuals’ per-
formance and non-performance of agreed actions is given high visibility.

10.6 People Development

Developing people within an organization should be viewed as an


investment. Then it becomes easy to understand the importance of main-
taining this key resource so that it keeps performing at optimum level. If
we know the cost of replacing an employee, then it is easy to conclude that
getting the most from employees just makes good business sense.
But this is much easier said than done, particularly with so many
diverse behaviors, motivations, and desires. How do we develop our peo-
330 Chapter 10

ple resources? There are several ways to help increase personal satisfac-
tion and productivity that will benefit the employees and the organization.
Are people performing their jobs to the best of their abilities? What
additional training would allow particular employees to do their job bet-
ter? (Answers may include personal development, skills training, or both.)
Can job rotations or on-the-job training (OJT) with experienced cowork-
ers enhance employees’ skills and awareness of how their work fits into
overall organization’s goals? Could their tasks be automated, allowing
people to grow in other areas? These are some of the questions we need
to consider while developing a plan to enhance knowledge and skill sets
of our people. Job task analysis is one of the techniques that could be used
to identify specific training needs, ensuring that employees have appropri-
ate knowledge and skill sets to do their jobs effectively.

Job Task Analysis (JTA)


Job task analysis is the foundation of a successful training program.
Before employees can be trained, we must identify what they need to
learn. Job analysis is a process that determines in detail the particular
duties and requirements for a given job as well as the relative importance
of these duties.
A job is a collection of tasks and responsibilities that are assigned to
an employee. A task is typically defined as a unit of work, that is, a set of
activities needed to produce some result, e.g., changing belts, repairing a
pump, delivering or expediting material, performing FMEAs, or sorting
the mail. Complex positions in the organization may include a large num-
ber of tasks, which are sometimes referred to as functions. Job descrip-
tions are lists of the general tasks, or functions, and responsibilities of a
position. Typically, they also include to whom the position reports, speci-
fications such as the qualifications needed by the person in the job, salary
range for the position, etc.
Job descriptions are usually developed by conducting a job analysis,
which includes examining the tasks and sequences of tasks necessary to
perform the job. The analysis looks at the areas of knowledge and skills
needed for the job. A role is the set of responsibilities or expected results
associated with a job. A job usually includes several roles. An important
aspect of job analysis is that the analysis is conducted for the job func-
tions, not the person. Although job analysis data may be collected from
incumbents through interviews or questionnaires, the product of the
analysis is a description or specifications of the job, not a description of
the person.
Workforce Management 331

The purpose of job analysis is to establish and document the skills


required for performing the job effectively. This can help in employment
procedures such as selection of employee, compensation, performance
appraisal, and training.

New Hire Selection Process/Procedures


Job Analysis can be used in the selection process to identify or develop:
• Job duties that should be included in advertisements of vacant
positions
• Appropriate salary level for the position to help determine what
salary should be offered to a candidate
• Minimum requirements (education and experience) for screening
applicants
• Interview questions
• Selection instruments (e.g., written tests; oral tests; job
simulations)
• Applicant appraisal and evaluation forms
• Orientation materials for applicants and new hires

Performance Review
Job analysis can be used in the performance review to identify or
develop:
• Goals and objectives
• Performance standards
• Evaluation criteria
• Job duties to be evaluated
• Time period between evaluations

Determining Training Needs


Job analysis can be used to identify required knowledge and skills and
to develop appropriate training, including:
• Training content
• Assessment tests to measure effectiveness of training
• Equipment to be used in delivering the training
• Methods of training (e.g., small group, computer-based, video,
classroom)

Several methods may be used individually or in combination to per-


form job analysis. These include:
332 Chapter 10

• Review of job classification systems


• Incumbent interviews and logs
• Observations
• Area supervisor interviews
• Expert panels
• Structured questionnaires
• Task inventories
• Check lists

A typical method of job analysis is to give the incumbent a simple


questionnaire to identify job duties, responsibilities, equipment used,
work relationships, and work environment. The completed questionnaire
is then used to assist the job analyst, who conducts an interview with the
incumbents. Next, a draft of the identified job duties, responsibilities,
work environment, asset and tools knowledge, and relationships is
reviewed with the supervisor for accuracy. The job analyst then prepares
a job description and job specifications. In case of a new position, the job
analyst prepares the document with the help of the requester and area
supervisor. Information collected by the job analyst can be grouped into
the following five categories:

Duties and Tasks


The basic unit of a job is the performance of specific tasks and duties.
Information to be collected about these items may include: frequency,
duration, effort, skill, complexity, equipment, and standards.

Environment
The work environment may include unpleasant conditions such as
temperature extremes, offensive odors, and physical limitations (or con-
straints) that can hinder the job performance. There may also be definite
risks such as noxious fumes, radioactive exposures such as X-rays, and
dangerous explosives.

Tools and Equipment


Some duties and tasks are performed using specific equipment and
tools. Equipment may include protective clothing. These items need to be
specified in a job analysis.
Workforce Management 333

Relationships
Is supervision required or not? What kind of interaction does this job
require—internally with fellow employees or externally with others out-
side the organization?

Requirements
Knowledge, skills, and abilities (KSAs) are required to perform the
job. Although an incumbent may have higher KSAs than those required
for the job, a job analysis typically states only the minimum requirements
to perform the job.

Skills Development Training


The skill level of the maintenance personnel in most organizations
today is well below what industry would classify as acceptable. The liter-
acy level of many maintenance personnel is becoming a challenge. New
entry-level employees have shown a noticeable drop in basic math and
reading skills. They have also shown a lack of interest in Operations and
Maintenance related work. Moreover, many organizations have eliminat-
ed their apprenticeship programs so now they are unable to fill their own
positions with qualified employees. These factors are creating a great
demand for employees with acceptable skill levels.
Today’s assets and systems are increasingly complex. They require an
educated and skilled workforce to operate and maintain them effectively.
It has been determined by several studies that today’s craft person requires
a minimum educational level of 12 years with additional vocational train-
ing to meet the work requirements.
Numerous studies have also shown that 70–80% of equipment fail-
ures are self-induced; most of those are a result of human error. All human
error failures can’t be blamed on education or skills, but they do make the
problem worse.
The U.S. Department of Education funded a survey by the Bureau of
Census to determine how training impacts productivity. The survey
revealed that increasing an individual’s educational level by 10% increas-
es the productivity by 8.6%. Several other research studies have indicated
similar results. Educated and skilled workforce improves productivity and
reduces human-induced errors.
A well-developed training program based on job task analysis and
maintenance skills assessment can provide the solution to inadequate
maintenance skills availability. The training must be focused to produce
334 Chapter 10

results as quickly as possible and must also meet an organization’s long-


term goals. Maintenance training, when developed and implemented
properly, can help organizations save money, increase productivity and
product quality, and improve employee morale.
The training curriculum should include at a minimum:
a) Regulatory and safety requirements, e.g. OSHA, EPA, FDA
b) Technical
i) Asset/System—operations and maintenance
To provide basic understanding of how an asset or system
operates and how it interfaces with utilities and other assets.
It may include minimum maintenance needs or operator-
required maintenance.
ii) Specific repair techniques and technology
To provide new repair technique or technology-related train-
ing, e.g., hydraulic servo valves, vibration, ultrasonic, and
laser alignment
iii) Professional development
RCM, FMEA, 6-Sigma tools, blueprint reading, etc.
c) Organization specific, e.g., process related—how to write work
orders or requests for material, company diversity policy, etc.

In addition, the organizations may provide remedial enhancement


educational courses to improve basic reading and math capabilities. This
could be done economically through local community colleges.
It is also a good practice to track employee training records by estab-
lishing a training database. This database should include:
• Training required with due date
• Training completed by date with hours credited
• Any mandatory, regulatory, or qualification training required,
e.g., waste water operators, welding, or special equipment opera-
tions or maintenance
• Skill assessment information
• Certifications achieved, etc.

Training Resources and Benchmark


How much money should we spend on training; how much is good
enough? Many organizations face this dilemma. Some managers believe
that money spent on training has a very low or poor return on investment.
In reality, this is not so because benefits and results cannot be realized in
the short term. Training metrics are lagging indicators and it takes a while
Workforce Management 335

to see the results. Figure 10.2 lists typical benchmark data, based on
reviewing several benchmark studies and my personal discussion with
many M&R and Training Managers.

Figure 10.2 Training Benchmarks

In a recent survey of 984 North American manufacturing plants, the


Manufacturing Performance Group (MPI) found:

Training Levels (more than 40 hours/year/person)


United States 18%
Canada 17%
Mexico 29%

Only 18% of manufacturers in the United States spend more than 40


hours per person annually on training, compared to 29% of Mexican man-
ufacturers.

Certification and Qualification


Earlier in this chapter we discussed investing in our people to educate
and train them. Organizations want their people to have a good under-
standing and appropriate skills in the M&R field in order to help them to
become more efficient and effective. How do we know that they, the
employees, have the required knowledge? How do we assess that knowl-
edge? Did they comprehend the proper use of M&R tools and best prac-
tices during training sessions? Certification is a means to assess the
required knowledge or set of skills in a specific field.
Why do organizations need to get their people certified? What value
does a certification provide? These are the questions raised by many
organizations. One answer is “Can we afford not to get certified?”
336 Chapter 10

According to the U.S. Department of Labor, Bureau of Labor Statistics


Outlook Handbook 2008/9 edition, “Many employers regard certifications as
the industry standard.” Certification measures and evaluates an employee’s
understanding of a body of knowledge in a specific area with a standard
knowledge that has been established by an appropriate industrial or academ-
ic body. Most of the certifying organizations are either professional societies
or educational institutions. Some of them have their certifying process com-
ply with and approved by the American National Standard Institute (ANSI)
and International Standards Organizations (ISO).
Certifications in the M&R area such as CMRP (Certified Maintenance
& Reliability Professional) or CMRT (Certified Maintenance &
Reliability Technician) by SMRP (www.smrp.org), CVA—Certified
Vibration Analyst by Vibration Institute (www.vibinst.org), or MLA
(Machinery Lubrication Analyst) by ICML (www.lubcouncil.org) indicate
that a successful applicant has the following attributes:

• Demonstrated knowledge in specific M&R field—concept and


implementation
• Equipped with the skills to perform effectively in a specific area

Employers of certified employees should be confident that they have


individuals who have proven themselves, possess the skill sets necessary
for success, and have met a specific certification standard. The same can
be said of new certified employees being hired.
Certifications can be grouped into four major classifications:

1. Asset/System Level
2. M&R Technologies
3. M&R Professionals/Managers
4. Plant/Facility Level

Asset/System Level
Assets and systems are getting much more complex. Many organiza-
tions have started to qualify operators and maintainers for critical and
complex assets. Organizations need to assure that people who are going to
operate and maintain have the appropriate skills. Operators and maintain-
ers are required to go through a training curriculum specifically design to
educate key aspects of that asset and then test to ensure that they have
comprehended the knowledge. Usually this type of certification or quali-
fication requirements are developed and administered within the organi-
Workforce Management 337

zations. In some cases where public assets are involved—such as the


operation of water treatment plants, boilers, etc.—such certification is
provided by an outside agency.

M&R Technologies
Several technology and general maintenance-related certifications
are available such as oil analysis, machinery vibration analysis, infrared
(IR) thermography, ultrasonic testing, motor current testing, hydraulics,
and pressure vessels. These certifications are valuable to both employees
and organizations. They test knowledge in their respective area of main-
tenance technologies. Training for most of these certifications is provided
by the key suppliers of these technologies or related professional soci-
eties, who also administer the test.

M&R Professionals/Managers
In this category, maintenance and reliability engineers, capital project
engineers, designers, managers, and other professionals working in the
M&R field are tested for their broad knowledge of M&R.
Two key certifications are available in this category. One is Certified
Plant Engineer (CPE)/Certified Maintenance Manager (CMM) offered by
the Association of Facility Engineers (AFE). The other is Certified
Maintenance and Reliability Professional (CMRP) conducted by the
Society of Maintenance and Reliability Professionals Certifying
Organization (SMRPCO). AFE and some industrial training companies
provide the necessary training for the test.
The CMRP certification process is accredited. According to require-
ments, SMRPCO is prohibited from providing any specific training. They
want to keep the certification—testing process away from trainers.
However, SMRPCO has a study guide that identifies M&R body of
knowledge, available in hard copy format from SMRP headquarters. This
guide can be downloaded from their website www.smrp.org/certification
free of cost.
CMRP certification was initiated in 2000 and is now recognized as the
standard of M&R certification by many organizations worldwide. The
certification process evaluates an individual’s skills in the five pillars of
knowledge defined by SMRP: Business and Management, Manufacturing
/Operations Processes Reliability, Equipment Reliability, Organization &
Leadership (People Skills), and Work Management. Many organizations
have now started using the CMRP certification to assess their employees’
knowledge and then to develop appropriate training programs that help
338 Chapter 10

their employees improve their skills.


Another certification available for reliability engineers, called
Certified Reliability Engineers (CRE), is conducted by the American
Society for Quality (ASQ). deals with the quantitative and analytical
skills employed by reliability engineers to manage reliability and risk.
This certification is more technically inclined towards the product relia-
bility professional.

Plant/Facility Level
There are no plant certifications available. However, two organiza-
tions recognize best plants/facilities. One is the North American
Maintenance Excellence (NAME) Award given to a plant based on
NAME-established criteria. The other is the “Best Plant” award conduct-
ed by the publishers of Industry Week. This award is not specific to M&R,
but recognizes overall aspects of manufacturing plant operations, e.g.,
quality, productivity, meeting customer needs on a timely basis, and
inventory levels.

10.7 Resource Management and Organization Structure

Developing people and managing them to be productive is a key fac-


tor in operating a successful business. The success of workforce manage-
ment hinges on implementation of the following core, though interde-
pendent goals.

a. Aligning the workforce with the business strategy


b. Attracting, developing, and retaining key talent
c. Managing diversity
d. Designing best organization structure for integration of M&R
functions
e. Succession planning
f. Developing a leadership culture
g. Establishing and maintaining a learning environment
h. Creating a flexible work environment

Aligning the Workforce with the Business Strategy


Employees need to know why the organization is in business. Involve
employees in creating or confirming the organization’s mission.
Crystallize the organization’s reason for being and the direction it plans to
take into the future. People support what they help to create.
Workforce Management 339

Having the right people in the organization has a huge influence on


successfully executing the business strategy. Elevating workforce man-
agement to the level of a strategic function and aligning it with an organi-
zation’s overall strategic direction can provide a focused approach to peo-
ple development.

Attracting, Developing, and Retaining Key Talent


Individuals are still likely to shop around for better opportunities if
they are not nurtured on the job. A well-defined process for developing
people means enhancing career and succession planning programs to
show employees that the organization is committed to their long-term suc-
cess. It also means demonstrating a desire for strategic investment in
career development programs. Five primary reasons why people leave or
change jobs are listed below. A plan should be in place to minimize the
impact of these factors.
1. I don’t fit in here. This perspective is a corporate culture issue in
most cases. Employees are also concerned with the organization’s
reputation, the physical conditions of comfort, convenience, safe-
ty, and the clarity of mission.
2. They wouldn’t miss me. Even though managers do value
employees, they don’t tell them often enough. If people don’t feel
important, they’re not motivated to stay. No one wants to be a
commodity, easily replaced by someone off the street. They’ll
leave for a position where they’re appreciated.
3. I don’t get the support to get my job done. Most of the time,
people really do want to do a good job. When they’re frustrated
by too many rules, red tape, or incompetent supervisors or co-
workers, employees look for other opportunities.
4. Lack of opportunity for advancement. People want to learn, to
sharpen their skills, and to pick-up new knowledge and skills.
They want to improve their capacity to perform a wide variety of
jobs. They desire better training opportunities and development.
If people can’t find the growth opportunities within one organiza-
tion, they’ll seek another employer where they can grow.
5. Compensation is the last reason people leave. Employees want
fair compensation, but the first four aspects must be strong. If
they’re not, but compensation is high, you’ll hear people say “you
can’t pay me enough to stay here.”
340 Chapter 10

Many organizations are either trying to lure valued retirees back to


work or entice older workers to stay on the job longer. However, an organ-
ization needs to capture the knowledge that retiring baby boomers threat-
en to take with them. Mentoring is an effective way of transferring that
knowledge. Assign young employees with experienced baby boomers and
let them mentor and transfer their knowledge.
Today’s workforce also likes to have stretch goals. It may require
abandoning some old ways of doing things to optimize employees’
options, and thus maximize chances of retaining nurtured talent. We need
to make job opportunities visible across the organization, as well as invest
in retraining and re-assignment of responsibilities. Of course, the right
compensation plan is as crucial as career advancement opportunities for
all high-performing employees.

Managing Diversity
Over 80% of new entrants into the workforce today are women and
minorities. This changing workforce is one of the major challenges facing
businesses. Organizations that recognize the need to fully develop all
members of the workforce demographic are forced to manage diversity.
Definitions of “diversity” range from narrow to very broad. Narrow
definitions tend to reflect Equal Employment Opportunity law, and define
diversity in terms of race, gender, ethnicity, age, national origin, religion,
and disability. Broad definitions may include sexual orientation, values,
personality characteristics, education, language, physical appearance,
marital status, lifestyle, beliefs, and background characteristics.
In the near future the labor market will become more and more of a
seller’s market. The shrinking workforce and the shortage of skilled labor
will force employers to compete to attract and retain all available employ-
ees, including previously under-represented groups. These demographic
changes have led many organizations to begin changing their cultures in
order to value and manage diversity better.
According to 2004 U.S. Bureau of the Census projections (comparing
2020 to 2000), the percentage of workers aged 20–44 will decline from
36.9% to 32.3%; the number of workers aged 45–64 will increase from
22.1% to 24.9%; and the number of workers aged 65–84 will increase
from 10.9% to 14.1%.
Such gray-haired demographics aren’t limited to the United States
either. The number of workers aged 20–44 will decrease and workers aged
45–84 will increase in other countries such as the United Kingdom,
Germany, Japan, and China.
Workforce Management 341

Many boomers say they plan to balance work and leisure in retire-
ment. They don’t plan to stop working at age 65, instead opting for a
“working retirement.” The reasons are both financial and personal.
Several studies have found that many corporate policies hinder efforts
to adapt to the aging workforce. Still, most organizations say they hire for
ability and willingness to work. A few employers say they’re “hiring wis-
dom” when hiring older workers. Gray-haired workers are viewed as reli-
able, settled, compassionate, and honest. Some organizations have set up
a Casual Worker Program that allows them to hire or reemploy workers
who would receive limited benefits and no pension.

Designing Best Organization Structure for Integration of M&R


Functions
Throughout the industry, organizations are being flattened, intermedi-
ate layers of management are being removed, and the overall size of M&R
departments are being reduced. This is the reality of today’s competitive
nature of the business.
In this environment, how organizations are structured to provide key
M&R functions is important for effective use of its people. There are three
types of organization structures which could be used to set up an M&R
organization:
• Centralized
• De-centralized
• Hybrid

Any time we start discussing organization structure and skills, the first
question that arises is which type of organization structure would be more
beneficial: a centralized or decentralized structure (see Figure 10.3).
Typically, a centralized organization structure is characterized by greater
specialization and standardization. The decentralized structure provides
stronger ownership and responsiveness.
Many experts have advocated shifting from one structure to another,
centralized to decentralized, and vice versa. The primary reason for this
recommendation is not that one type is necessarily better than the other,
but to bring about change. Shaking things up is necessary to break old, tra-
ditional practices. A hybrid structure, containing the best characteristics of
both, centralized and decentralized, is often the optimal solution.
In a hybrid organization, an individual production unit or area may
have their own maintenance technician assigned to take care of area
breakdowns and minor repairs. Operators become valuable resources as
342 Chapter 10

Centralized Decentraliz ed
Advantages Disadvantages Advantages Disadvantages
Standardiz ed practices Less responsive to Str ong ownership Dif ficu lt to buil d
individu al units / specialized skills
departme nt
Enterprise focu s — Lack of ownership Very responsive to Dif ficu lt to prioritize by
objective s aligne d with individu al area facil ity/department
organiza tion
Effici ent us e of resources Sub -optimum use of tools

Easy to bui ld spec ialize d


skills

Figure 10.3 Centralized and Decentralized Organizations Comparison

troubleshooting experts, or as system turnover or return-to-service techni-


cians. In some cases, they provide extra hands in the repair process. The
large PMs, major repairs, and specialized technical tasks such as PdM/
and critical alignments are performed by the centralized specialized staff.
In designing the organization structure, it is important to address the
following M&R functions:
• PM program development
• Execution of PM and CM tasks
• Planning and scheduling
• Material—spares including tools availability
• CBM/PdM and specialized skills, e.g., laser alignment
• Failure elimination and reliability planning
• Designing for reliability
• Resource management and budgeting
• Workforce development

One example of an M&R organization structure is shown in Figure


10.4.

Outsourcing Maintenance
We are under constantly increasing pressure to reduce our labor cost.
If there are some constraints put on the number of employees we can have,
under those circumstances we can evaluate outsourcing or shared servic-
es for some specific skills in order to obtain a cost-effective solution.
Large variations in workload may lead to poor utilization of resources
and overstaffing. Outsourcing some of the maintenance activities is a
Workforce Management 343

Figure 10.4 Typical M&R Organization Structure

good practice. During scheduled shutdowns and major outages, it may be


economical to contract work outside the organization. It can be very cost-
effective not only to contract the resources for executing the work but also
to have this group plan and schedule major outages. However, periodic
shutdowns—for example, a critical rotating machine—can be managed
better by our own shutdown planners.
Augmenting maintenance staff with engineering students (Intern and
Co-op) during summer is another cost effective good practice to follow.
Some of the maintenance and reliability data collection, analysis, and
CBM-related work can easily be accomplished by these interns. Some
university groups, such as the University of Tennessee’s Reliability and
Maintenance Center (RMC), equip their interns with appropriate mainte-
nance and reliability basics before they intern or co-op. Each student
intern is required to attend one week of so called “R&M boot camp.” This
boot camp is taught by the University staff as well as industry subject mat-
ter experts. The author had the privilege to teach this group for several
years. One other benefit of these individuals is that after gaining experi-
ence with your organization, they can be valuable full-time employees
upon graduation.
344 Chapter 10

Some organizations have established a small reliability and mainte-


nance staff core group at the corporate level. They are subject matter
expert internal consultants who can help standardize and implement reli-
ability-related policies and procedures and best practices corporate-wide.
However, this group needs to be very careful and diplomatic when deal-
ing with small organizations. They should create a good relationship in
working with them to help implement some of the best practices without
acting simply like an enforcer of corporate policies.

Succession Planning
Succession planning is another key element of the workforce devel-
opment process. It is the process of identifying and preparing suitable
employees—through mentoring, training, and job rotation—to replace
key players in the organization. It involves having management periodi-
cally review their key personnel and those in the next lower level to deter-
mine several backups for key positions. This is important because it often
takes years of grooming to develop effective managers and leaders.
A careful and considered action plan ensures the least possible dis-
ruption to the person’s responsibilities and to the organization’s effec-
tiveness if a key player suddenly becomes unavailable to perform his or
her role. Some sudden actions that may take place include:

• Sudden or unexpected inability or unwillingness to continue their


role within the organization
• Acceptance of a position from another organization or external
opportunity which will terminate or lessen their value to the cur-
rent organization
• Conclusion of a contract or time-limited project
• Move to another position and different set of responsibilities
within the organization
• Plans to retire
• Serious illness or sudden death

The goal of succession planning is to continuously identify and devel-


op high-performing leaders capable of meeting the future needs of the
organization. Goals should include a formal identification process sup-
ported by leadership development that builds leadership capacity within
the organization. Many organizations now use succession planning to
address the succession of workers at many different levels as they leave
the workforce. A workforce-planning strategy should include ongoing
Workforce Management 345

workforce training and leadership development to meet the looming chal-


lenge of an aging workforce.
Samuel Greengard, a noted management author, has pointed out sev-
eral key steps for successful succession planning in “Five Keys to
Successful Succession Planning”:

1. Identify key leadership criteria.


2. Find future leaders and motivate them.
3. Create a sense of responsibility within the organization.
4. Align succession planning with the corporate culture.
5. Measure results and reinforce desired behavior so that employ-
ees are prepared and trained for the jobs of tomorrow.

Succession planning ensures that there are highly qualified people in


all positions, not just today, but tomorrow, next year, and five years from
now. Succession planning establishes a process that recruits employees,
develops their skills and abilities, and prepares them for advancement, all
while retaining them to guarantee a return on the organization’s training
investment. Succession planning involves the following steps:

• Determine critical positions.


• Identify current and future competencies for positions.
• Create assessment and selection tools.
• Identify gaps in current employee and candidate competency
levels.
• Develop individual development plans for employees.
• Develop and implement coaching and mentoring programs.
• Assist with leadership transition and development.
• Develop an evaluation plan for succession management.

Developing a Leadership Culture


Managers and supervisors should be encouraged to become leaders.
Directive, autocratic management styles will soon become a thing of the
past. The leaders of tomorrow will need to be facilitative. Establishing
agreed-upon results with individual performers, the management role will
provide needed resources, support, and coaching. Part of the leader’s role
will be to help people reach their full potential, and then raise the bar to
keep them growing.
Leaders inspire rather than direct. They coach, encourage, and guide.
Effective leaders earn agreement with their people about what has to be
346 Chapter 10

done. They determine—with their team members—what resources are


needed to get the job done. They arrange those resources and then get out
of the way so their people can perform.
Leaders appreciate and recognize their people. People are hungry for
appreciation. Thanking people every day, being sincere and specific, seek-
ing creative ways to show care and appreciation are some valuable com-
ponents of good leadership styles.
Leaders also endeavor to educate and train all managers to coach their
team members on an ongoing basis. They focus on helping each worker
learn how to improve performance. Continuous improvement is much
more effective than annual appraisal alone. Supervisors/ managers need to
learn new coaching techniques to deal with the new diversified workforce.

Establishing and Maintaining Learning Environment


As mentioned in the previous section, an environment conducive to
learning is maintained by being supportive to personal and professional
growth through education, training, and people development to help every
employee become more competent and confident. Personal and corporate
growth objectives are linked to build an environment that bonds people to
the organization. When people can meet their personal needs through their
employment, they’ll be more likely to stay.

“If you think training/education is expensive;


try to count the cost of ignorance.”
Anonymous

Creating a Flexible Work Environment


Rigidity in the employment environment is waning. On the contrary,
there is a trend toward more flexible work hours. Work policies that allow
working from home are being demanded. The organization’s policies need
to be reviewed and changes made if necessary to meet the needs of this
new work environment.
The move away from formality at work can also be seen in dress
attire. It started with “casual Fridays” and has now moved to the accept-
ance of “corporate casual dress” every day. Relationships have also
become more informal. Research shows that the informality removes sta-
tus barriers and improves measurable productivity. This change to infor-
mality does not come without consequences. Policies and procedures
must be changed to allow this informality, but with guidance to protect
Workforce Management 347

personnel from harassment and inappropriate behavior toward and


between employees.

10.8 Measures of Performance

The processes involved in workforce management, like other process-


es, need to be measured and evaluated to make improvements. A few
examples of performance measures and benchmark data include:

1. Percent of training budget ($ and man-hours per employee).


This measure is related to how much we are budgeting for people
development. The benchmark is about 5%.
2. Number (or %) of certified professionals as of total work-
force. This measure is related to skill sets of M&R workforce,
including technicians.
3. Number of papers being presented or published at mainte-
nance and reliability-related conferences. This measure is
related to people development and benchmarking and sharing
information with industry peers.
4. Number of people among M&R staff supporting/involved
with industry-sponsored standardization teams or profession-
al society’s teams. This measure is related to benchmarking and
sharing information with industry peers

10.9 Summary

Big changes are happening in today’s workforce. These changes have


nothing to do with downsizing, global competition, or stress; they result
from a distinct generation gap. Young people entering the workforce each
bring a more diversified background and have much different attitudes
about work. They want a life-work balance. They want to be led, not man-
aged—and certainly not micro-managed. The new mode is flexibility and
informality. A large proportion of managers of the veteran era have been
trained in relatively autocratic and directive methods that don’t sit well
with today’s employees.
There’s a very serious labor shortage right around the corner.
According to data from the Bureau of Labor Statistics, there will be more
jobs than people to fill them. In addition, people entering the workforce,
specifically in the M&R field, lack basic skill sets as well as reading and
348 Chapter 10

math capabilities. Additionally, there is a need to have people develop-


ment programs in every organization to ensure a supply of skilled and tal-
ented workers.
The labor shortage will intensify. Finding qualified employees will be
difficult. Workers will move easily from job to job in a flow we used to
call “job hopping.” The continual shifting will be commonplace.
Organizations will be caught unaware until they suddenly lose their
best people. It is important to be prepared and have your program in place
to retain them if possible, or to have back-up workers trained and ready to
take new responsibility.
Organizations don’t have to wait until they have an opening to recruit
qualified applicants. They should have a list of screened potential employ-
ees, ready to call when we have the right opportunity for them.
Internships, co-operative programs, and similar tools can be used to
recruit and evaluate future employees. Succession planning for all posi-
tions and training of identified employees need to be developed. These
steps will guarantee the availability of the right people when we need
them. Succession planning is not just a simple replacement strategy. It
involves forecasting workforce needs and developing ongoing strategies
to meet those needs. Every person hired must meet minimum qualification
standards. This practice sends a very clear message that we won’t allow
anyone who may drag us down. By adhering to this rule, organizations
reinforce the high performance of current employees.
The organizational structure needs to be examined critically. Is the
organization structured in an optimal way for smooth operation? Does it
support productivity, accountability, and profitability? If not, changes
have to be instituted to concentrate energies for results. People need to be
cross-trained to enable them to work productively together across depart-
mental or functional lines.
Despite all these efforts to attract, train, retain, and minimize the loss
of critical skills and talent, the fact remains that we are under increasing
pressure to adopt a more flexible workforce model in order to turn fixed
cost into variable cost. Shared services and outsourcing need to be evalu-
ated to obtain a cost-effective solution. Let us not forget that today’s
workforce—which consists primarily of Baby Boomers, Gen Xers, and
Gen Yers with diversified, ethnic backgrounds—requires different sets of
benefits and management styles to retain them within the organization.
Finally, getting a workforce with the required knowledge and skill
sets requires a long-term commitment. Aligning the workforce with over-
Workforce Management 349

all business strategy is just a starting point. Organizations need to devel-


op systematic approaches in attracting, developing, and retaining key
young talent, while at the same time minimizing the loss of those critical
skills that older workers possess in abundance. Additionally, only by con-
tinually challenging the workforce can we build a more flexible workforce
model. It is people or, more aptly, the right people that make things hap-
pen.

10.10 Self Assessment Questions

Q10.1 Who is Dr. W. Edwards Deming? What was his message?

Q10.2 How are Dr. Deming’s principles related to the workforce?

Q10.3 Who were the quality gurus who revolutionized Japanese


industry in the 1960s? What did they do?

Q10.4 Explain the employee life cycle. What is its importance?

Q10.5 What is the largest expense for an organization?

Q10.6 What does “Generation Gap” mean? How does it impact an


organization?

Q10.7 What can we do to leverage different generations’ employees


to our advantage?

Q10.8 Why is the communication process important to us?

Q10.9 What happens in the communication process?

Q10.10 How should we determine training needs?

10.11 References and Suggested Reading

Archer, Ron. On Teams. Irwin Professional Publication, 1995.


Deming, W. Edwards. Out of the Crisis. MIT Press, Paperback issue,
2000.
350 Chapter 10

DDI, Inc. Engaging the Four Generations. www.ddiworld.com


Frost, Peter et al. HRM Reality, 2nd edition. Prentice Hall, 2002.
Greengard, Samuel. Finding the Work You Love. AARP
Books/Sterling, 2008.
The Herman Group. Red Alert 2006: Workforce Trends.
www.hermangroup.com
Strauss, William and Howe, Neil. Generations. William Morrow
Publication, 1991.
U.S. Department of Labor, Bureau of Labor Statistics. Outlook
Handbook 2008/9 edition.
Workforce Management. www.workforce.com
Chapter 11
Maintenance Analysis and
Improvement Tools

Every problem is an opportunity.


—Kilchiro Toyoda, founder of Toyota

11.1 Introduction
11.2 Key Terms and Definitions
11.3 Maintenance Root Cause Analysis Tools
11.4 Six Sigma and Quality Maintenance Tools
11.5 Lean Maintenance Tools
11.6 Other Analysis and Improvement Tools
11.7 Summary
11.8 Self Assessment Questions
11.9 References and Suggested Reading

After reading this chapter, you will be able to understand:


• Why are analysis tools necessary?
• Types of analysis tools available
• What analysis tools should be used? When and where should
they be used?
• Application of 6 sigma in a non-production environment
• What is meant by lean and VSM, and how they apply to
maintenance

351
352 Chapter 11

11.1 Introduction

Organizations must continually improve processes, reduce costs, and


cut waste to remain competitive. Data (O&M data) should be analyzed
using various techniques and tools in order to develop and implement
effective plans that can lead to improvements in assets and processes.
Recent industry surveys have indicated that although many organiza-
tions have started investing time and effort to improve their processes, it
isn’t unusual to see the same problems popping up over and over. The
impact of these problems on customers (internal and external), employ-
ees, profitability, and competitiveness have been well documented. One
factor making such problems highly visible is the formalized management
systems guided by documents such as ISO 9000. A new requirement of
“continuous improvement” in ISO 9001 requires organizations to collect
and analyze data on process performance using audits, internal perform-
ance indicators, and customer feedback. Any problems that are identified
are to have corrective action taken to prevent recurrence.
Unfortunately, insufficient effort has been placed on providing guid-
ance on how to carry out an effective diagnosis to identify the causes of
problems. Organizations generally try to fix the symptoms of the prob-
lems instead of fixing the root causes. They try often to implement what
we may call a “duct tape solution,” hoping it will address the problem.
Meanwhile, the risks associated with repeat problems have significantly
increased.
In addition, our assets /systems are getting very complex. Although
the identification of problems is more rigorous, the ability to solve them
has not necessarily improved at the same rate. Much of the training that is
generally provided is too high level and philosophical, or is not focused
on analytical problem solving. People are not being taught how to think
logically and deductively. They lack the knowledge of what tools to use
and how to apply them appropriately.
This chapter provides a tool box with many problem solving tools in
it which are focused more heavily on the analytical process involved in
finding the actual causes of problems. This chapter will discuss various
tools and techniques that are available—ranging from simple checklists
and spreadsheets to sophisticated modeling software that is helpful for
solving problems. We will focus our discussion on a few of these tools,
briefly describing situations in which they can appropriately be used.
Maintenance Analysis and Improvement Tools 353

There are many tools available to us. For the sake of streamlining our
discussion; they have been classified in four major categories. These are:

• Maintenance Root Cause Analysis Tools


• Six Sigma and Quality Maintenance Tools
• Lean Maintenance Tools
• Other Analysis and Improvement Tools

Each of these will be discussed in the next four sections.

11.2 Key Terms and Definitions

5 Whys
A problem-solving technique for discovering the root cause of
a problem. This technique helps users to get to the root of the
problem quickly by simply asking “why” a number of times
until the root cause become evident.

Barrier Analysis
A technique often used, particularly in process industries,
based on tracing energy flows. It has a focus on barriers to
those flows, and helps to identify how and why the barriers did
not prevent the energy flows from causing damage.

Cause-and-Effects Analysis
Also called Ishikawa or fishbone chart. It identifies many pos-
sible causes for an effect or problem, and then sorts ideas into
useful categories to help in developing appropriate corrective
actions.

Cause Mapping
A simple, but effective method of analyzing, documenting,
communicating, and solving a problem to show how individual
cause-and-effect relationships are inter-connected.
354 Chapter 11

Checklists
A generic tool that can be developed for a wide variety of pur-
poses. It is a structured, pre-prepared form for collecting,
recording, and analyzing data as the work progresses. Some
examples are operator’s start-up checklist, PM checklist, and
maintainability checklist used by designers.

Control Charts
A graph used to display how a process changes over time.
Comparing current data to historical control limits indicates
process variations whether the process is in control or out of
control.

Design for Six Sigma (DFSS)


A systematic methodology using tools and training to enable
the design of products, processes, and services that meet cus-
tomer expectations at 6 Sigma quality levels. DFSS optimizes
the design process to achieve a very high quality and repeat-
able 6 Sigma performances. It follows a five-phase process
called DMADV (Define – Measure – Analyze – Design –
Verify), which is sometimes synonymously referred to as
DFSS.

Design of Experiments (DOE)


A method for carrying out carefully planned experiments on a
process. Usually, design of experiments involves a series of
experiments that start by looking broadly at a large numbers of
variables and then focusing on the few critical ones.

Failure Modes and Effects Analysis (FMEA)


A technique to examine an asset, process, or design to deter-
mine potential ways it can fail and the potential effects (conse-
quences); subsequently to identify appropriate mitigation tasks
for highest priority risks.

Fault Tree
This analysis tool is constructed starting with the final failure
(or event) and progressively tracing each cause that led to the
previous cause. This continues till the trail can be traced back
Maintenance Analysis and Improvement Tools 355

no further. Once the fault tree is completed and checked for


logical flow, it is determined what changes would prevent the
sequence of causes (or events) with marked consequences
from occurring again.

Flow Chart
A graphical summary of the process steps (such as production,
storage, transportation) and flows (movement of information
and materials) that make up a procedure or process from
beginning to end. This information is used in defining, docu-
menting, studying, and improving the system. Also called flow
diagram, flow process chart, or network diagram.

Mistake Proofing
Mistake proofing, also known as Poka-Yoke (Japanese equiva-
lent), is the use of any automatic device or method that either
makes it impossible for an error to occur or makes the error
immediately obvious once it has occurred.

Muda
Japanese lean word for waste; non-value-added work.

Mura
Japanese lean word for unevenness; inconsistency.

Muri:
Japanese lean word for overburden; unreasonable work.

Pareto Analysis
This bar graph displays variances by the number of their
occurrences. Variances are shown in their descending order to
identify the largest opportunities for improvement, and to sepa-
rate the critical few from the trivial many. The concept is also
known as 80/20 Principle.

PDCA — Deming’s Improvement Cycle


Plan – Do – Check – Act (PDCA) is known as Deming’s
methodology to make improvements.
356 Chapter 11

Root Cause Analysis


Identification and evaluation of the reason for an undesirable
condition or non-conformance. A methodology that leads to
the discovery of the cause of a problem or root cause.

Scatter Diagram
A diagram that graphs pairs of numerical data, one variable on
each axis, to look for a trend or a relationship .

Six Sigma
This methodology systemically analyzes processes to reduce
process variations and also to eliminate wastes. Six Sigma is
also used to further drive productivity and quality improve-
ments in any type of organization. DMAIC (Define – Measure
– Analyze – Improve – Control) represents the steps used to
guide implementation of the Six Sigma process.

Standard Deviation
Standard deviation measures variations of values from the
mean. It is denoted by the Greek letter (σ) and is calculated
using the following formula:

where ∑ = sum of, X = observed values, X bar (X with a line


i
over the top) = arithmetic mean, and n = number of observa-
tions.

Stratification
A technique that separates data gathered from a variety of
sources so that a pattern can be seen.

Theory of Constraints (TOC)


Concepts and methodology aimed mainly at achieving the
most efficient flow of material in a plant. Basically it is a
scheduling and inventory control philosophy that proposes that
any organization has a chain of interdependent links (depart-
Maintenance Analysis and Improvement Tools 357

ments, functions, resources); some may have potential for


greater performance, but cannot realize it because of a weak
link — bottleneck (constraint). TOC supports identification
and removal of bottlenecks. It is also sometime called bottle-
neck analysis.

Value Stream Mapping (VSM)


VSM is a tool that helps to visualize and understand the flow
of information and material as it makes its way through the
process value stream. It identifies steps which are not adding
any value — they are waste and needed to be removed from
the process or improved.

11.3 Maintenance Root Cause Analysis Tools


Root Cause Analysis (RCA)
Root Cause Analysis (RCA), or Root Cause Failure Analysis (RCFA)
as it is sometime called, is a step-by-step methodology that leads to the
discovery of the prime cause (or the root cause) of the failure. If the root
cause of a failure is not addressed in a timely fashion, the failure will
repeat itself, usually causing unnecessary loss of production and increas-
ing the cost of maintenance. RCA is a structured way to arrive at the root
cause, thus facilitating elimination of the cause and not just symptoms
associated with it.
Assets, components, and processes can fail for a number of reasons.
But usually there is a definite progression of actions and consequences
that lead to a failure. An RCA investigation traces the cause and effect trail
from the failure back to the root cause. RCA is more like a detective at
work trying to solve a crime or, in a somewhat similar way, the National
Transportation Safety Board (NTSB) trying to piece together evidence
following a plane crash to determine the cause of the failure.
Several studies by many organizations have repeatedly proven that
90% of the time unwanted situations caused by failures are related to
process problems; only about 10% are related to personnel problems. Yet,
most organizations spend far more time looking for culprits, rather than
focusing on finding root causes. Because of this misdirected effort, we
often miss the opportunity to learn and benefit from understanding the
root cause of the unwanted failures and eliminating those causes.
Consider the following two scenarios.
358 Chapter 11

Scenario #1 The Plant Manager walked into the plant and found a
puddle of oil on the floor near a tube assembly machine. The manager
instructed the area supervisor to have the oil cleaned up immediately. The
next day, while in the same area of the plant, the Plant Manager again
found oil on the floor and asked the area supervisor to get the oil cleaned
up from the floor. In fact, the manager was a little upset with the supervi-
sor for not following directions given the day before. His parting words
were either to get the oil cleaned up or he would find someone who would.
Scenario #2 The Plant Manager walked into the plant and found a
puddle of oil on the floor near a tube assembly machine. The manager
asked the area supervisor why there was oil on the floor. The supervisor
indicated that it was due to a leaky oil seal in the hydraulic pipe joint
above. The Plant Manager then asked when the oil seal had been replaced.
The supervisor responded that Maintenance had installed 5 or 6 oil seals
over the past few weeks and each one seemed to leak. The supervisor also
indicated that Maintenance had been talking to Purchasing about the seals
because it seemed they were all failing — leaking prematurely.
The Plant Manager then went to talk to Purchasing about the situation
with the seals. The Purchasing Manager indicated that they had in fact
received a bad batch of oil seals from the supplier as reported by the main-
tenance department. The Purchasing Manager also indicated that they had
been trying for the past month or so to get the supplier to make good on
the last order of 50 seals that all seemed to be bad.
The Plant Manager then asked the Purchasing Manager why they had
purchased from this supplier if their quality was poor. The Purchasing
Manager replied that the supplier was the lowest bidder when quotes were
received from various suppliers. When the Plant Manager asked the
Purchasing Manager why they went with the lowest bidder without con-
sidering bidder’s quality issues, the Purchasing Manager explained that he
was directed by the Finance Manager to reduce cost.
Next, the Plant Manager went to talk to the Finance Manager about
the situation. The Finance Manager noted that his direction to Purchasing
to always take the lowest bid was in response to the Plant Manager’s
memo telling them to be as cost conscious as possible and only purchase
from the lowest bidder, thus saving money. The Plant Manager was horri-
fied to realize that he was the reason there was oil on the plant floor. What
a discovery!!
We may find Scenario # 2 somewhat funny, and even laugh when the
problem comes full circle. We have found that most of the time everyone
Maintenance Analysis and Improvement Tools 359

in the organization tries to do their best and to do the right things. But,
sometimes things don’t work out the way we envision. The root cause of
this whole situation is sub-optimization with no overall vision. Scenario
#2 also provides a good example of how one should proceed to do the root
cause analysis. We need to continue asking “Why?” until a pattern
emerges and the cause of the difficult situation becomes rather obvious.
When we have a problem, how do we approach it for a solution? Do
we jump in and start treating the symptoms, much like continually clean-
ing oil as in Scenario #1? If we only fix the symptoms, based on what we
see on the surface, the problem will almost certainly happen again. Then
we will keep fixing the problem, again and again, but never solving it.
The practice of RCA is predicated on the belief that problems are best
solved by attempting to correct or eliminate root causes, as opposed to
merely addressing the immediately obvious symptoms. By directing cor-
rective measures at root causes, the likelihood of problem recurrence will
be minimized. In many cases, complete prevention of recurrence through
a single intervention is unlikely. Therefore, RCA is often considered to be
an iterative process; it is frequently viewed as a part of a continuous
improvement tool box.
Root cause analysis is not a single, defined methodology; there are
several types or philosophies of RCA in existence. Most of these can be
classified into four, very broadly defined categories based on their field of
application: safety-based, production-based, process-based, and asset fail-
ure-based.

1. Safety-based RCA is performed to find causes of accidents


related to occupational safety, health, and environment.
2. Product- or Production-based RCA is performed to identify
causes of poor quality, production and other problems in
manufacturing related to the product.
3. Process-based RCA is performed to identify causes of prob-
lems related to processes, including business systems.
4. Asset-based RCA is performed for failure analysis of assets
or systems in engineering and the maintenance area.

Despite the seeming disparity in purpose and definition among the


various types of root cause analysis, there are some general principles that
can be considered as universal.
360 Chapter 11

General Principles for RCA

• Aiming corrective measures at root causes is more effective than


merely treating the symptoms of a problem.
• To be effective, RCA must be performed systematically, and
conclusions must be backed up by evidence.
• There is usually more than one root cause for any given
problem.

The Six Steps in Performing an RCA

• Define the problem — the failure.


• Collect data / evidence about issues that contributed to the
problem.
• Identify possible causal factors.
• Develop solutions and recommendations.
• Implement the recommendations.
• Track the recommended solutions to ensure effectiveness.

Step One: Define the Problem

• What happened?
• What were the specific symptoms?

What happened? What failed? How was the problem discovered?


What was the sequence of events that led to the failure or breakdown?
There should be a physical examination of the area or asset involved and
a detailed description of the actual event. There is a high probability of
missing the right cause if the entire problem from every aspect is not
checked.

Step Two: Collect Data / Evidence

• What proof do we have that the problem exists?


• What sequence of events leads to the problem?
• How long has the problem existed?
• What is the impact of the problem?

Analyze the situation to evaluate factors that contributed to the prob-


lem. To maximize the effectiveness of RCA, get all involved together,
Maintenance Analysis and Improvement Tools 361

e.g., operators, maintainers, and others who are familiar with the situation.
People most familiar with the problem can help lead to a better under-
standing of the issues. Identify issues that contributed to the problem col-
lectively. The details of the problem–failure can be organized by using the
‘3W2H’ (what, when, where, how, how much) tool.
Do not make any assumptions when examining a problem. No two
problems are exactly the same in nature and cause. Actually, it is rare for
the exact same failure to occur twice. Each problem should be reviewed
as if you are looking at the situation for the first time. It may be that the
two failure phenomena appear to be same, but the causes can be different.

Step Three: Identify Possible Causal Factors

• What are the causal factors?


• Why does the causal factor exist?
• What is the real reason the problem occurred?

Identify every possible cause of failure. Looking at problems in the


past may be helpful in determining cause of failure; however, you should
not limit your search to past causes. Every possible cause must be consid-
ered and examined.
In performing cause analysis, it will become clear in the process that
some causes will be illogical. Remove all illogical causes after careful
examination. If we determine that the cause of the failure was human
error, separate that cause from the physical causes.
Use these tools to help identify causal factors:

• 5 Whys — Ask Why? Until the root of the problem is found.


• Cause and Effect Diagrams — Create a chart of all of the possible
causal factors, to see where the trouble may have begun.
• Drill Down — Break down a problem into small, detailed parts to
better understand the big picture.
• Ask So what? Determine all the possible consequences of a fact.

These tools are designed to encourage analyzing deeper at each level


of cause and effect. During this stage, identify as many causal factors as
possible. Too often, people identify one or two factors and then stop, but
that’s not sufficient. With RCA, we don’t want to simply treat the most
obvious causes; we may need to dig deeper.
362 Chapter 11

Step Four: Develop Solutions and Recommendations.


Based on the factors that may cause or have already caused the fail-
ure, you need to develop mitigating corrective actions. To develop appro-
priate recommendations, group the cause factors into three basic cause
types:

a) Physical causes — Tangible, material items failed in some way


(for example, Crane’s brakes stopped working; the hydraulic
cylinder rod didn’t stop at the right location).
b) Human causes — The operator or mechanic did something
wrong or did not do something that was needed. Human causes
typically lead to physical causes (for example, low brake fluid
which led to brake failure; the limit switch was not relocated at
the right location after the last set-up change).
c) Organizational / Process causes — A system, process, or policy
that people use to make decisions or do their work is faulty (for
example, no one person was responsible for vehicle maintenance,
and everyone assumed someone else had filled the brake fluid; no
written procedure was available to ensure relocation of the limit
switch after the set-up change).

Root Cause Analysis should investigate all three types of causes. It


involves investigating the patterns of negative effects, finding hidden
flaws in the system, and discovering specific actions that contributed to
the problem. This often means that RCA reveals more than one root cause.

Step Five: Implement Recommendations

a) What can we do to prevent the problem from happening again?


b) How will the solution be implemented?
c) Who will be responsible for it?
d) What are the risks of implementing the solution?
e) How will implementation success be measured?

Develop a plan with a schedule for implementing the suggested solu-


tion or recommendations. The plan should be presented to all stakehold-
ers including management for their approval. The plan should also identi-
fy how implementation progress will be tracked and what metrics should
be used to measure the effectiveness of the recommended solution.
Maintenance Analysis and Improvement Tools 363

Step Six: Track the recommended solutions to ensure effectiveness.


After recommended solutions have been implemented, tracking the
appropriate metrics to measure the effectiveness of the recommended solu-
tion is an essential component of RCA. If the solution has not been effec-
tive, the team should revisit the RCA and modify the solution, then imple-
ment the revised recommendations and start re-measuring the effectiveness

Examples of tools and techniques to perform root cause analysis

• 5 Whys
• Cause and effect diagram or fishbone diagram
• Failure mode and effects analysis (FMEA)
• Pareto analysis — 80/20 rule
• Fault tree analysis
• Barrier analysis
• Cause mapping

Examples of some basic elements of root causes


Materials
• Defective raw material
• Wrong type of material for this job

Asset / Machine
• Incorrect asset /machine used
• Incorrect tool selected
• Poorly maintained or inadequate maintenance
• Poor design
• Poor machine installation
• Defective machine or tool

Environment
• Poorly maintained workplace
• Inadequate job design or layout of work
• Surfaces poorly maintained
• Physical demands of the task
• Forces of nature or Act of God

Safety and Management


• No or poor management involvement
• Inattention to task
364 Chapter 11

• Task hazards not guarded properly


• Poor recognition of hazard
• Previously identified hazards were not eliminated
• High stress demands

Methods
• No or poor procedures
• Not following the procedures
• Poor communication

People System / Involvement


• Lack of or poor training
• Lack of process / machine operating procedures
• Lack of or poor employee involvement

Root Cause Analysis Template


Recommended template for final RCA report:
1. Undesirable Event
2. Undesirable Event Summary
3. Data Summary from FMEA / Pareto Analysis
4. Identified Root Causes
a. Physical
b. Human
c. Organization / Process and Procedures
5. Recommended Corrective Action
6. Implementation Plan
7. Metrics to measure effectiveness
8. Team Members
9. Special / Additional comments

Several computer software programs are available that can be very


useful in supporting and documenting the root analysis. It is suggested
you use a commercially-available program or a spreadsheet to document
the process findings.

5 Whys Analysis
The 5 Whys is a simple problem-solving technique that helps users get
to the root of the problem quickly. Made popular in the 1970s by the
Toyota Production System, the 5 Whys analysis involves looking at any
problem and asking: “Why?” and “What caused this problem? Quite
Maintenance Analysis and Improvement Tools 365

often, the answer to the first why will prompt another why and the answer
to the second why will prompt another, and so on — hence the name the
5 Whys analysis.

Benefits of the 5 Whys include:

• It helps to quickly determine the root cause of a problem.


• It is easy to learn and apply.

How to Use the Tool


When looking to solve a problem, start at the end result and work
backward (toward the root cause), continually asking, “Why?” This
process will need to be repeated over and over until the root cause of
the problem becomes apparent. If it doesn’t quickly give an answer
that’s obviously right, then you may need more sophisticated problem-
solving techniques.

Example
The following example shows the effectiveness of the 5 Whys
analysis as a problem-solving technique:
1. Why is our customer (operations department xyz) unhappy?
• Because we did not deliver our services (fixing the
asset) when we said we would.
2. Why were we unable to meet the agreed-upon schedule for
delivery?
• The job took much longer than we thought it would.
3. Why did it take so much longer?
• Because we underestimated the work requirements.
4. Why did we underestimate the work?
• Because we made a quick estimate of the time needed to
complete it, and did not list the individual steps needed to
complete the total job. In short, we did not do work plan-
ning for this job.
5. Why didn’t we do planning — detailed analysis — for this job?
• Because we were running behind on other projects and were
short on planner resources. Our customer (operations depart-
ment xyz) was forcing us to do the job quickly. Clearly, we
needed better planning, including improved time estimations
and steps needed to complete the job efficiently.
366 Chapter 11

The 5 Whys analysis is an effective tool for uncovering the root causes
of a problem. Because it is so elementary in nature, it can be adapted quick-
ly and applied to almost any problem. Remember, if it doesn’t prompt an
intuitive answer, other problem-solving techniques may need to be applied.

Cause-and-Effects Analysis (or Fishbone Diagram)

What Is a Fishbone Diagram?


Dr. Kaoru Ishikawa, a Japanese quality control statistician, invented
the fishbone diagram, which often looks like the skeleton of a fish, hence
its name. The fishbone diagram is an analytical tool that provides a sys-
tematic way of looking at effects and the causes that create or contribute
to those effects. Because of this function of the fishbone diagram, it is also
called a Cause-and-Effect Diagram.
Whatever name we choose to call this analysis, it helps us in a sys-
tematic and simple way to categorize the many potential causes of prob-
lems and to identify root causes. Usually this analysis is performed as a
team. After listing the possible causes for a problem, the team can analyze
each cause carefully, giving due importance to statements made by each
team member during the brainstorming session, accepting or ruling out
certain causes, and eventually arriving at the root cause of the problem. In
general, fishbone diagrams give us increased understanding of complex
problems by visual means of analyses.

When Should We Use a Fishbone Diagram?


It is helpful to use the fishbone diagram in the following cases:
• To stimulate thinking during a brainstorming session
• When there are many possible causes for a problem
• To evaluate all the possible reasons when a process is beginning
to have difficulties, problems, or breakdowns
• To investigate why an asset or process is not performing proper-
ly or producing the desired results
• To analyze and find the root causes of a complicated problem
and to understand relationships between potential causes
• To dissect problems into smaller pieces

When Should We Not Use a Fishbone Diagram?


Of course, the Fishbone diagram isn’t applicable to every situation.
Listed below are just a few examples in which we should not use the
Fishbone diagram because the diagram either is not relevant or does not
produce the expected results:
Maintenance Analysis and Improvement Tools 367

• The problem is simple or is already known.


• The team size is too small for brainstorming.
• There is a communication problem among the team members.
• The team has experts who can fix any problem without much
difficulty.
How to Construct a Fishbone Diagram
The following five steps are essential when constructing a fish-bone
diagram:
1. Define the problem.
2. Brainstorm.
3. Identify all causes.
4. Select any causes that may be at the root of the problem.
5. Develop corrective action plan to eliminate or reduce the impact
of the causes selected in Step 4.

The first step is fairly simple and straightforward. Define the problem
for which the root cause needs to be identified. Usually the maintenance /
reliability engineer or technical leader chooses the problem that needs a
permanent fix, and that is worth brainstorming with the team.
After the problem is identified, the team leader can start constructing
the Fishbone diagram. The leader defines the problem in a square box to
the right side of a page or worksheet. A straight line is drawn from the left
to the problem box with an arrow pointing towards the box. The problem

Figure 11.1 Fishbone Diagram: Failure of a Hydraulic Pump


368 Chapter 11

box now becomes the fish head and its bones are to be filled in during the
steps that follow. Figure 11.1 provides an example of a Hydraulic Pump
analysis. In this example, a hydraulic pump that is not pumping the desired
output (oil — specified pressure and volume) has become a problem.
The next step is to start identifying major components and suspected
causes of this failure, e.g., bearing failure, motor failure, seal failure, or
shaft failure. All major causes are identified and connected as parts (the
bones) of the Fishbone diagram. Causes of bearing failures are also listed
in this example. The next step is to refine the major causes to find the sec-
ondary causes and other causes occurring under each of the major cate-
gories.
In general, the following steps are taken to draw the fishbone dia-
gram:
1. List the problem/issue to be investigated in the “head of the
fish”.
2. Label each “bone” of the “fish”. Major categories typically
include:
• The 4 Ms:
Methods, Machines, Materials, and Manpower
• The 4 Ps:
Place, Procedure, People, and Policies
• The 4 Ss:
Surroundings, Suppliers, Systems, and Skills
• The 6 Ms
Machine, Method, Materials, Measurement, Man, and
Mother Nature (Environment)
• The 6 EPMs
Equipment/Asset, Process, People, Materials,
Environment, and Management.

a. The team may use one of the categories suggested above, com-
bine them in any manner, or make up others as needed. The cat-
egories are to help organize the ideas.
b. Use an idea-generating technique (e.g., brainstorming) to identi-
fy the factors within each category that may be affecting the
problem or effect being studied. The team should ask, “What are
the issue and its cause and effect?”
c. Repeat this procedure with each factor under the category to
produce sub-factors. Continue asking, “Why is this happening?”
and put additional segments under each factor and subsequently
Maintenance Analysis and Improvement Tools 369

under each sub-factor.


d. Continue until you no longer get useful information when you
ask, “Why is that happening?”
e. Analyze the results of the fishbone after team members agree
that an adequate amount of detail has been provided under each
major category. For example, look for those items that appear in
more than one category. These become the most likely causes.
f. For those items identified as the most likely causes, the team
should reach consensus on their priority. The first item should
be listed the most probable cause.

An example of another fishbone diagram is shown in Figure 11.2. In


this example, an analysis team is trying to understand poor humidity con-
trol problem in a drier application. The team used five specific headings
to prompt the ideas.
Figure 11.3 shows yet another example of a fishbone diagram. Here,
“Maintenance Excellence” is the problem statement (result). The diagram
lists causes — in this case actions needed to be taken in order to achieve

Figure 11.2 Fishbone Diagram: Poor Humidity Control Problem in Drier


370 Chapter 11

Figure 11.3 Elements of Excellence

excellence in maintenance. It should be noted that problem to be solved


could be positive too. We always think a problems to be in a negative
sense. These tools could be used in either way.
Sometimes the fishbone diagram can become very large because the
team may identify many possible causes. This makes the diagram very
complex; comprehending the relationship of the causes can be difficult. A
good fishbone diagram is one which has explored all the possibilities for
a problem but is still easy to understand when developing corrective
action plans.

Failure Modes and Effects Analysis (FMEA)


This analysis tool is also called failure modes, effects and criticality
analysis (FMECA), and potential failure modes and effects analysis.
FMEA is a step-by-step methodology for identifying all possible failures
during the design of an asset (product) — in a manufacturing or assembly
process, in the operations and maintenance phase, or in providing servic-
es.
Failure modes are the ways, or modes, in which something might fail.
Failures are any potential or actual errors or defects that affect the cus-
tomer, user, or asset itself. Effects refer to the consequences of those fail-
ures.
Maintenance Analysis and Improvement Tools 371

Failures are prioritized according to how serious their consequences


are, how frequently they occur, and how easily they can be detected. The
purpose of FMEA is to take actions to eliminate or reduce failures, start-
ing with the highest-priority ones.
FMEA is an economical and effective tool for finding potential fail-
ures early in the design–development phase where it is easier to take
actions to overcome these issues, thereby enhancing reliability through
design. FMEA is used to identify potential failure modes and their effect
on the operation of the assets; it also is helpful when developing effective
PM actions to mitigate consequences of failure. It is an important step in
anticipating what might go wrong with assets. Although anticipating
every failure mode may not possible, the analysis team should formulate
as extensive a list of potential failure modes as possible. Early and consis-
tent use of FMEAs in the design process allows us to design out failures,
in turn making assets more reliable and safe.
It has been observed that designers and engineers often use safety fac-
tor as a way of making sure that the design will work and protect the asset
(product) and user upon failure. In the past, asset and systems designers
have not done a good job designing in reliability and quality into the asset.
The use of a large safety factor does not necessarily translate into a reli-
able asset. In fact, it often leads to an overdesigned product with reliabil-
ity problems.

Types and Usage of FMEAs


There are several types of FMEAs, based on how and where it is used,
e.g., in the design–development, operations, or maintenance phase of the
assets. FMEA should always be performed whenever failures could create
potential harm or injury to the user (operator), environmental challenges,
or breakdown of the asset, in turn causing loss of production. FMEAs can
be classified into the following categories:

a. Design: focuses on components and subsystems


b. Process: focuses on manufacturing and assembly processes
c. Maintenance: focuses on asset functions
d. Service: focuses on service functions
e. Software: focuses on software functions

Although the purpose, terminology, and other details can vary accord-
ing to type (e.g., Process FMEA, Design FMEA), the basic methodology
is similar for all.
372 Chapter 11
Figure 11.4
FMEA Steps
Maintenance Analysis and Improvement Tools 373

Figure 11.4 depicts the sequence in which a FMEA is performed. The


typical sequence of steps answers the following set of questions:

1. What are the components and the functions they provide?


2. What can go wrong?
3. What are the effects?
4 How bad are the effects?
5. What are the causes?
6. How often can they fail?
7. How can this be prevented?
8. Can this be detected?
9. What can be done; what design, process, or procedural changes
can be made?

Published Standards and Guidelines


There are a number of published guidelines and standards for the
requirements and recommended reporting format of FMEAs. Some of the
key published standards for this analysis include

• SAE J1739,
• AIAG FMEA-3
• MIL-STD-1629A (out of print / cancelled)

Automotive Industry Action Group (AIAG) guidelines and reference


manual are very similar to SAE standard – J1739. Figure 11.5 shows the
SAE/AIAG recommended format for performing and reporting FMEA
In addition, many industries and organizations have developed their
own FMEA procedures and formats to meet the specific requirements of
their products and processes. Figure 11.6 illustrates another organization’s
approach to documenting FMEA using a simple spreadsheet. Part a shows
elements of Failure Modes Identification and Effects whereas Part b
shows Prevention Impact and Mitigation Assessment.
In general, FMEA requires the identification of the following basic
information:

i. Items — components
ii. Functions
iii.Failure modes
iv.Effects of Failure
v. Causes of Failure
374 Chapter 11
Figure 11.5
SAE/AIAG
FMEA
Guidelines
Maintenance Analysis and Improvement Tools 375
FMEA with a
Documenting
Figure 11.6a

Spreadsheet
376 Chapter 11

Figure 11.6b Documenting FMEA continued

Figure 11.7 Failure Modes and Causes


Maintenance Analysis and Improvement Tools 377

vi. Probability (Frequency) of Failure


vii. Severity of Effects
viii. Likelihood of Detection
ix. Current Mitigating Plan
x. Recommended Actions

FMEA Procedure
The basic steps for performing a FMEA are outlined below:
1. Establish the objective of FMEA and identify analysis team
members who should consist of representatives from key stake-
holders, e.g., design, operations, maintenance, and materials. If
the FMEA is design-based, the designer should lead this effort.
2. Describe the asset or process and its functions. A good under-
standing of the asset or process under consideration is important
as this understanding simplifies the analytical process.
3. Create a block diagram of the asset or process. In this diagram,
major components or process steps are represented by blocks
connected together by lines that indicate how the components or
steps are related. The diagram shows the logical relationships of
components and establishes a structure around which the FMEA
can be developed. Establish a coding system to identify system
elements.
4. Create a spreadsheet or a form and list items or components and
the functions they provide. The list should be organized in a log-
ical manner according to the subsystem and sub-assemblies,
based on the block diagram.
5. Identify which component could fail, how it will fail (failure
modes), and for each component why it will fail (possible
cause). A failure mode is defined as the manner in which that
item / component could potentially fail while meeting the design
intent. Examples of potential failure modes include:

• Broken / Fractured
• Corrosion
• Deformation
• Clogged / Contamination
• Excess Vibration
• Electrical Short or Open
• Eroded / Worn
378 Chapter 11

A list of suggested failure modes and causes is shown in Figure 11.7.

6. A failure mode in one component can be the cause of a failure


mode in another component. Failure modes should be listed for
functions of each component or process step. At this point, the
failure mode should be identified whether or not the failure is
likely to occur.
7. Identify the causes for each failure mode. A failure cause can be
a design weakness, or any operator or maintenance workman-
ship that may result in a failure. The potential causes for each
failure mode should be identified and documented. The causes
should be listed in technical terms and not in terms of symp-
toms. Examples of potential causes include:
• Improper torque applied
• Improper operating conditions
• Corrosion
• Coordination
• Design induced — wrong part used or wrong control circuit
• Improper alignment
• Excessive loading
• Excessive current /voltage
8. Identify failure mechanisms that will create a random or wear-
out type of failure event.
9. List the effects of each of the failure modes. A failure effect is
defined as the result of a failure mode on the function of the
component or process. Examples of failure effects include:
• Injury to the user–operator
• Creation of hazards material leak — environmental spill
• Loss of all or partial functions
• Creation of unbearable, bad odors
• Degraded performance
• High noise levels
10. Establish a numerical ranking for the severity of the effect. A
common industry standard scale uses 1 to represent no effect
and 10 to indicate very severe with failure affecting system
operation and safety without warning. The intent of the ranking
is to help the analyst determine whether a failure would be a
minor nuisance or a catastrophic occurrence. This enables the
designers and engineers to prioritize the effects of failures and
Maintenance Analysis and Improvement Tools 379

address the critical issues first.


11. Estimate the probability factor for each cause. A numerical
weight should be assigned to each cause that indicates how like-
ly that cause is (probability of the cause occurring). A common
industry standard scale uses 1 to represent not likely and 10 to
indicate inevitable.
12. What is the likelihood that the current control measures will
detect at the onset the cause of the failure in time and thus pre-
vent it from happening?
13. Calculate and review Risk Priority Numbers (RPN). The Risk
Priority Number is a mathematical product of the numerical
Severity, Probability, and Detection ratings: RPN = (Severity) x
(Probability) x (Detection). To use the Risk Priority Number
(RPN) method to assess risk, the analysis team must:
• Rate the severity of each effect of failure.
• Rate the likelihood of occurrence for each cause of failure.
• Rate the likelihood of prior detection for each cause of fail-
ure (e.g., the likelihood of detecting the problem by the
operator or during PM inspection before it fails).
The RPN is calculated by obtaining the product of the three rat-
ings. The RPN can then be used to compare risks within the
analysis and to prioritize corrective action items.
14. Identify current practices including PM plan or controls in place.
The controls strive to prevent the failures or detect them for cor-
rective actions before they impact the operations. The designer
or engineer uses analysis, monitoring, testing, and other tech-
niques to detect failures. These techniques should be assessed to
determine how well they are expected to identify or detect fail-
ure modes. FMEA should then be updated; plans should be
made to address those failures to eliminate them. In some
FMEA applications, the dollar cost of business loss, which con-
sist of downtime cost, failure fixing cost, and frequency of fail-
ure, is also calculated to prioritize the corrective action plan
15. Determine recommended actions to address potential failures
that have a high RPN. These actions may include specific
inspections, PdM /CBM tasks, or improved operations; selection
of better components or materials; limiting environmental stress-
es or the operating range; redesign of the item to avoid the fail-
ure mode; monitoring mechanisms; changing frequency of pre-
380 Chapter 11

ventive maintenance; and inclusion of back-up systems or


redundancy.
16. Assign responsibility and a target completion date for these
actions. This makes responsibility clear-cut and facilitates track-
ing.
17. Track actions taken and re-evaluate risk. After actions have been
taken, re-assess the severity, probability, and detection; then
review the revised RPNs. Determine if any further actions are
required. Update the FMEA as the design, process, or the assess-
ment changes or new information becomes known.

Benefits of FMEA
FMEA helps designers and engineers to improve the reliability of
assets and systems to produce quality products. FMEA analysis helps to
incorporate reliability and maintainability features into the asset design to
eliminate or reduce failures, thereby reducing overall life cycle cost.
Properly performed, FMEA provides several benefits. These include:

• Early identification and elimination of potential asset/process


failure modes
• Prioritization of asset /process deficiencies
• Documentation of risk and actions taken to reduce risk
• Minimization of late changes and associated cost
• Improved asset (product) / process reliability and quality
• Reduction of Life Cycle Costs
• Catalyst for teamwork among design, operations, and
maintenance

Fault Tree
A fault tree is constructed starting with the final failure. It progres-
sively traces each cause that led to the previous cause. This continues until
the trail can be traced back no further. Each result of a cause must clearly
flow from its predecessor. If it becomes evident that a step is missing
between causes, it is added in and need is explained.
Once the fault tree is completed and checked for logical flow, the
investigating team determines what changes to make to prevent or break
the sequence of causes and consequences from occurring again. It is not
necessary to prevent the first, or root cause, from happening. It is merely
necessary to break the chain of events at any point so that the final failure
cannot occur. Often the fault tree leads to an initial design problem. In
Maintenance Analysis and Improvement Tools 381

such a case, redesign becomes necessary. Where the fault tree leads to a
failure of procedures, it is necessary either to address the procedural
weakness or to install a method to protect against the damage caused by
the procedural failure.

11.4 Six Sigma and Quality Maintenance Tools

Six Sigma and Quality


Six Sigma is a quality improvement initiative, developed directly
from Total Quality Management (TQM). It uses much the same toolset
and concepts. The major emphasis of Six Sigma is its focus on reducing
process variation to very low levels.
Jack Welch, retired chairman of General Electric, has been Six
Sigma’s most influential advocate. Companies such as Motorola and
Allied Signal have been incubators and proponents of the movement.
Mikel Harry, the principal architect of Six Sigma methodology at
Motorola, is one of its most enthusiastic champions.
The name Six Sigma warrants some explanation. Imagine that we
make potato chips and weigh bags of potato chips as they come out of a
bagging process. The bags are supposed to weigh 200 grams (about 7
ounces), but the actual weight will vary. If bags are overweight, we are
giving away chips. If they are underweight, we are taking advantage of
our customers. To analyze the process variation, we record the weights,
and use a software program to construct a histogram of the distribution.
We hope that the distribution will be centered on 200 grams and that there
wouldn’t be long tails on either side. If our specification calls for all bags
to be more than 190 grams and less than 210 grams, we can draw the spec-
ification limits on the histogram.
In plotting the histogram, we found a bag that weighed 188 grams.
Obviously, it is out of specification. It may cause trouble with our cus-
tomers if we ship it. What do we do? How many out of specification bags
do we expect to find if we measure 1000 bags? We need to use statistical
tools to predict this. We can find the average weight — or mean weight —
of a bag of potato chips from all the bags we have weighed and calculate
a standard deviation, which provides us an idea of how much variation
there is around the mean. A high standard deviation means that we have a
lot of variation in the process. This is where Six Sigma comes in. (The
Greek letter sigma (σ) is typically used to symbolize the standard devia-
tion in statistical equations.)
With enough data, we can try to fit a curve to the data — drawing a
382 Chapter 11

line that best approximates the mathematical function that really describes
what is going on in the process. Let’s take a normal curve as an example
of one that we might decide to use. If our histogram can be closely
described by a normal distribution, then 68% of the bags we measure will
weigh within one standard deviation, or one sigma, of the mean value. If
the mean is 200 grams, and the standard deviation is 10 grams, then 68%
of the bags would weigh between 190 and 210 grams. Again, if this is a
normal distribution, we would find that 95.5% of the bags weighed with-
in 2 sigma, or 20 grams of the mean. If we consider 3 sigma, or 30 grams,
we would find that 99.7% of all bags would weigh between 170 grams,
and 330 grams. A very few, just 0.3% of all bags, would weigh less than
170 grams or more than 230 grams.
If this is the case, our process is wider than our specification limits.
Some of the bags weigh less than 180 grams, and some weigh more than
220 grams. Does that matter? The answer depends on product or process.
If we were measuring something requiring fine tolerances, or something
expensive, or where precise temperatures or mixture compositions were
critical, we wouldn’t want to find instances where our process was pro-
ducing outputs out of the specification limits. In that case, we need to
reduce the process variation. In the example above, if we can control the
process variation and reduce sigma (σ) to say 3 grams, then we can be
sure to meet and exceed the specification requirements, (200 grams +/- 10
grams), 99.7 % of the time.
The Six Sigma quality movement takes process variation very much
to heart. In fact, Six Sigma advocates believe that for many processes,
there should be six sigma control between the mean and the specification
limits, so that the process is making only a few bad (3.4) parts per million.
Of course, by relaxing the specifications, we can meet Six Sigma require-
ments, but that isn’t usually the way to please customers. Instead, the vari-
ation in the process needs to be driven towards zero, so that the histogram
gets narrower, and fits more comfortably inside the specification limits.
Six Sigma methodologies are not new. They combine elements of sta-
tistical quality control, breakthrough thinking, and management science
— all valuable, powerful disciplines. The application of quality tools and
process improvement can help in achieving excellent results.

Core of Six Sigma and Implementation


As we have discussed, Six Sigma is a statistical concept that measures
a process in terms of defects. Achieving Six Sigma means that the process
is delivering only 3.4 defects per million opportunities (DPMO). Thus, it
Maintenance Analysis and Improvement Tools 383

Figure 11.8 Six Sigma Chart

means the process is working nearly perfectly. One sigma represents


691,462 defects per million opportunities, which translates to a percent-
age of good (non-defective) output of only 30.85 %. This is really very
poor process performance. Figure 11.8 lists sigma level, DPMO, and per-
centage quality output.
To have a good understanding of Six Sigma and the DPMO concept,
let us consider lost luggage at the airport. Many of us have experienced
the frustration of watching the baggage carousel slowly revolve while
waiting for luggage that never arrives. The baggage handling capabilities
of many airlines is around the three or four sigma level. Let us assume it
is three. It means that there are about 66,000 defects for every one million
baggage transactions. This amount equates to approximately 93% proba-
bility that we will get our bags with our scheduled flight. Seven percent
of us are not going to receive our bags. If quality level is four, we will still
have at least one person who does not get his/her baggage. These defects
increases costs for the airlines, because airline employees must deal with
misplaced luggage and very unhappy passengers. Those defects — miss-
ing bags can result in lost business in the future. Similarly, in our manu-
facturing / maintenance process, these defects prohibit us from meeting
customers’ needs on time; they increase costs.
Deviations, variations, defects, or waste — whatever we want to call
it, the end result is the same. It costs money. No matter what business we
are in — distribution, manufacturing, process industry, or services — any
defects or hidden waste in any of our processes will impact our bottom
line.
To get an accurate view of critical processes, we need to understand
the limits on variations. Root causes of variation are explored, and the
384 Chapter 11

classic Deming’s PDCA (Plan, Do, Control and Act) cycle is used to plan
improvements, implement and test them, evaluate if they worked, and
then standardize if they did. However, for problem solving in Six Sigma,
the PDCA cycle has been modified slightly to have a five-phase method-
ology called DMAIC.
DMAIC represents the five phases in the Six Sigma methodology:

• D — Define
• M — Measure
• A — Analyze
• I — Improve
• C — Control

The DMAIC method involves completing the necessary steps in a


sequence. Skipping a phase or jumping around will not produce the
desired results. It is a structured process to solve problems with proper
implementation and follow-up. Each phase requires the following:

Define Phase (D)

• Identify the problems.


• Create a project to combat one or more problems.
• Define the parameters (boundaries) of the project.
• Determine the vital few factors to be measured, analyzed,
improved, and controlled.

Measure Phase (M)

• Select critical to quality (CTQ) characteristics in the process /


product, e.g., Y CTQ, where Y might be corrective maintenance
cost of an asset or system, downtime, etc.
• Define performance standard for Y — what is the desired goal
to achieve?
• Establish and validate process to measure Y.

Analyze phase (A)

• Define improvement objectives for Y.


• Define and identify sources (Xs) creating variations in Y.
• Review/sort potential sources for change in Y.
Maintenance Analysis and Improvement Tools 385

• Identify the vital few sources (Xs) creating variations or


defects
Improve Phase (I)
• Determine variable relationships among vital few factors – Xs.
• Establish operating range for vital few Xs.
• Validate measurement plan for Xs.

Control Phase (C)


• Establish plan to control vital few Xs.
• Implement the plan — process control vital few Xs.
• Review the data per plan.
• Make changes as necessary until the process is in control and
stable.

Pareto Analysis — 80/20 Principles


A Pareto chart is a bar graph that arranges information in such a way
that priorities for process improvement can be established easily. It is a
tool for visualizing the Pareto principle, which states that a small set of
problems, the “vital few,” affecting a common outcome tend to occur
more frequently than the remainder. In other words, 80% of the effects
(failures) are created by 20% of the causes (assets / components). A Pareto
chart can be used to determine which subset of problems should be solved
first, or which problems deserve the most attention. Pareto charts are often
constructed to provide a before-and-after comparison of the effect of
improvement measures. Basically, it is a resource optimization tool that
helps to focus on vital few issues to maximize gains with limited
resources.
The Pareto chart is used to illustrate occurrences of problems or
defects in a descending order. It can be used both during the development
process as well as when components or products are in use. It can depict
which assets are failing more in a specific area or what type of compo-
nents are having more failures than others. Figure 11.9 shows an example
of a Pareto chart for a compressor system with its components and failure
history.
Pareto charts are a key improvement tool because they help us identi-
fy patterns and potential causes of a problem. One trick, often overlooked,
is to create several Pareto charts out of the same set of data. Doing so can
help you quickly scan a number of factors that might contribute to a prob-
lem and focus on those with the greatest potential payback for your efforts.
386 Chapter 11

Figure 11.9 A Pareto Chart for a Compressor System

History of Pareto — Where It Came From


The Pareto Principle, or 80/20 Rule, was first developed in 1906 by
Italian economist Vilfredo Pareto who observed an unequal distribution of
wealth and power in a relatively small proportion of the total population.
His discovery has since been called many names, including the Pareto
Principles, Pareto Laws, the 80/20 rule or principle, the principle of least
effort, and the principle of imbalance. The 80/20 principle states that there
is a built-in imbalance between causes and results, inputs and outputs, and
efforts and rewards. This imbalance is shown by the 80/20 relationship. In
business, many applications of the 80/20 Principle have been validated.
For example, 20% of products usually account for about 80% of dollar
sales value. The same is true for 20% of customers accounting for 80% of
dollar sales value. In turn, 20% of products or customers usually also
account for about 80% of an organization’s profits.
Dr. Joseph M. Juran, the world-renowned quality expert, is credited
with adapting Pareto’s economic observations to business applications.
Dr. Juran observed that 20% of the defects caused 80% of the problems.
Project managers know that 20% of the work (the first 10% and the last
10%) consume 80% of time and resources. We can apply the 80/20 Rule
almost universally, from the science of management to the physical world.
Pareto Analysis is also used in inventory management through an
approach called ABC Classification, as discussed in Chapter 5.
Maintenance Analysis and Improvement Tools 387

Example of Pareto Principle


Generally, the Pareto Principle is the observation — not the law that
most things in life are not distributed evenly. It can mean all of the follow-
ing things:

• 20% of the process issues result in 80% of defects


• 20% of the assets cause 80% of the failures
• 20% of the failures cost 80% of the corrective maintenance
budget
• 20% of the input creates 80% of the result
• 20 % of criminals accounts for 80% of the value of all crimes
• 20% of motorists cause 80% of accidents
• 20% of your clothes will be worn 80% of the time
• And on and on…

Recognize that the numbers don’t have to be exactly 20% and 80%.
The key point is that most things in life, effort, reward, output, etc., are
not evenly distributed; some contribute more than others. The number
20% could be anything from 10–30%; similarly 80% could be 60–90%.
What we need to remember: 20% is the vital few; the remaining are the
others. Thus, 20% of the assets could create 60, 80, or 90% of the failures.
Pareto analysis helps us to prioritize and focus resources to gain maxi-
mum benefit.
Figure 11.10 shows an example of Pareto analysis of plant assets.
The value of the Pareto Principle is that it reminds us to focus on the
20% that are important. Of the things we do during our daily routine, only
20% of our activities really matter — they produce 80% of our benefits.
These are the activities we must identify and emphasize.

11.5 Lean Maintenance Tools

“Lean” is a new buzzword. Words such as lean production, lean man-


ufacturing, lean maintenance, lean management, lean enterprise, and lean
thinking have been abundantly discussed in literature in the last few years.
But what does “Lean” really mean?
As the word says, Lean means literally — LEAN. We all need to be
lean to become or stay healthy. We need to get rid of fat — waste which
we carry around with us. Similarly, in our work environment, we need to
be efficient and effective (in others words, lean) to stay healthy and sur-
vive in today’s competitive environment. Appropriate use of tools is nec-
388 Chapter 11

Figure 11.10 Pareto Analysis of Plant Assets

essary to be efficient and effective, and to ultimately create value for our
customers. In fact, many subject matter experts and authors are saying the
same mantra — get rid of the waste.
For example, Kevin S. Smith, President of TPG Productivity, Inc.,
states, “Lean is a concept, a methodology, a way of working; it’s any
activity that reduces the waste inherent in any business process.”
In their famous book Lean Thinking, James P. Womack and Daniel T.
Jones write that the critical starting point for lean thinking is value. Value
can only be defined by the ultimate customer. It’s only meaningful when
expressed in terms of a specific product (a good or a service, and often
both at once) that meets the customer’s needs at a specific price at a spe-
cific time.

Lean Background
Lean philosophy or thinking is not new. At the turn of the century,
Henry Ford, founder of the Ford Motor Company, was implementing lean
philosophy. Of course he didn’t use the word lean at that time.
John Krafcik, a Massachusetts Institute of Technology (MIT)
researcher in the late 1980s, coined the term Lean Manufacturing while
involved in a study of best practices in automobile manufacturing. The
Maintenance Analysis and Improvement Tools 389

MIT study had examined the methodology developed at Japanese auto


giant Toyota under the direction of production engineer Taiichi Ohno,
who later became known as the father of TPS, the Toyota Production
System, a model of the lean system. At the end of World War II, with
Toyota needing to improve brand image and market share, Ohno reputed-
ly turned to Henry Ford’s classic book, Today and Tomorrow for inspira-
tion. One of Ford’s guiding principles had been the elimination of waste.
Ohno is credited with developing the principles of lean production.
His philosophy, which focused on eliminating waste and empowering
workers, reduced inventory and improved productivity. Instead of main-
taining resources in anticipation of what might be required for future man-
ufacturing, as Henry Ford did with his production line, the management
team at Toyota built partnerships with suppliers. In effect, under the direc-
tion of Ohno, Toyota automobiles became made-to-order. By maximizing
the use of multi-skilled employees, the company was able to flatten their
management structure and focus resources in a flexible manner. Because
of this, Toyota was able to make changes quickly; they were often able to
respond more quickly to market demands than their competitors.
To illustrate the lean thinking, Shigeo Shingo, another Japanese Lean
and Quality expert, observed that only the last turn of a bolt actually tight-
ens it — the rest is just movement. This ever finer clarification of waste
is a key to establishing distinctions between value-added activity, waste,
and non-value-added work. Non-value adding work is a waste that must
be removed. Ohno defined three broad types of waste: Muri, Mura, and
Muda. These are three key Japanese words in lean terminology.

1. Muri: Overburden
2. Mura: Unevenness
3. Muda: Waste, non-value-added work

Muri is all the unreasonable work that an organization imposes on


workers and machines because of poor organizational design, such as car-
rying heavy weights, unnecessary moving, dangerous tasks, even working
significantly faster than usual. It is pushing a person or a machine beyond
its natural limits. This may simply be asking a greater level of perform-
ance from a process than it can handle without taking shortcuts and infor-
mally modifying decision criteria. Unreasonable work is almost always a
cause of multiple variations.
Mura is a traditional Japanese term for unevenness, inconsistency in
physical matter or human spiritual condition. Mura is avoided through JIT
systems which are based on little or no inventory, by supplying the pro-
390 Chapter 11

duction process with the right part, at the right time, in the right amount,
and first-in, first-out component flow. Just in Time systems create a “pull
system” in which each sub-process withdraws its needs from the preced-
ing sub-processes, and ultimately from an outside supplier. When a pre-
ceding process does not receive a request or withdrawal, it does not make
more parts. This type of system is designed to maximize productivity by
minimizing storage overhead.
Muda is a traditional Japanese term for activity that is wasteful and
doesn’t add value or is unproductive.
The original seven muda are:

1. Transportation. Moving material and parts that are not actually


required for the process.
2. Inventory. No extra inventory should be in the system. All com-
ponents, work-in-process, and finished product not being
processed are waste.
3. Motion. People or equipment moving or walking more than
required to perform the work are waste.
4. Waiting. Waiting for the next step or activity. Time not being
used effectively is a waste.
5. Overproduction. Production ahead of demand or need.
6. Inappropriate Processing. Produce only what is needed and
when needed with well-designed processes and assets.
7. Defects. The simplest form of waste involves components or
products that do not meet the specification. They lead to addi-
tional inspections and defects that must be fixed.

First, Muri focuses on the preparation and planning of the process, or


what work can be avoided proactively by design. Next, Mura focuses on
how the work design is implemented and the elimination of fluctuation at
the scheduling or operations level, such as quality and volume. Muda is
then discovered after the process is in place and is dealt with. It is seen
through variation in output. It is the role of management to examine the
Muda in the processes and eliminate the deeper causes by considering the
connections to the Muri and Mura of the system. The Muda and Mura
inconsistencies must be fed back to the Muri, or planning, stage for the
next project.
Lean requires the use of a set of tools that assist in the identification and
steady elimination of waste. Examples of such tools are Brainstorming,
Cause and Effects Analysis, Five S, Kanban (pull systems), Poke-yoke
(error-proofing), Pareto Analysis, and Value Stream Mapping.
Maintenance Analysis and Improvement Tools 391

Lean Maintenance
Much has been written and talked about lean concepts in manufactur-
ing, but what about lean maintenance? Is it merely a subset of lean man-
ufacturing? Is it a natural spinoff from adopting lean manufacturing prac-
tices? Lean maintenance is neither a subset nor a spinoff. Instead, it is a
prerequisite for success as a lean organization. Can we imagine lean JIT
concepts to work without having reliable assets or good maintenance
practices? Of course, we want maintenance to be lean — efficient and
effective — without waste. Lean maintenance has nothing to do with thin-
ning out warm bodies, or more directly, reducing maintenance resources.
Rather, it has to do with enhancing the value-added nature of our mainte-
nance and reliability efforts.
In maintenance, our customers are inherently internal to our organiza-
tion — they are our operations / production departments. One of the pri-
mary responsibilities of maintenance is to provide plant capacity to its
customers. Let’s face a fundamental truth: We can’t be successful with
Lean manufacturing if we don’t have reliable assets, reliable machines.
Lean maintenance is not performing lean (less) corrective or preventive
actions. It is not about facilitating a poor maintenance program.
Maintenance customers expect maintenance and reliability programs to be
optimized — effective and efficient, and fully supporting the need to oper-
ate at designed or required capacity reliably.
The majority of maintenance activities revolve around systems and
the processes that move people, material, and machine together such as
preventive maintenance programs, predictive maintenance programs,
planning and scheduling, computerized maintenance management sys-
tems, and store room and work order systems. We need to apply the prin-
ciples of Lean to these maintenance programs and processes to drive out
the non-value-added activities.
Value stream mapping for key maintenance processes to identify non-
value-added activities should be performed. It will be a good practice to
create current and future states of the maintenance processes in order to
develop a plan to reduce and eliminate wasteful activities. In developing
the current- and future-state maps of maintenance process, we must also
assess the skills and knowledge of our maintenance personnel. A poorly-
skilled person operating within a great system will produce poor results.
Likewise, if we have a good preventive maintenance program, yet our
PMs are poorly structured and designed, our PMs will achieve poor
results. Therefore we need to optimize PM using tools such FMEA/RCM
to give new life to our efforts.
392 Chapter 11

The endless pursuit of waste elimination is the essence of lean main-


tenance. Eliminate waste by understanding the seven wastes discussed
earlier in relation to maintenance. Identify where they exist and eliminate
them. For example:

1. Transportation. Plan and provide materials and tools to reduce


the number of extra trips to the store room to hunt for the right
parts.
2. Inventory. Eliminate or minimize extra inventory in the system.
Keep only the right material / parts / tools in the store room.
3. Motion. Minimize people movement by improved planning.
4. Waiting. Minimize waiting for the next step — another skilled
person or part — by improved planning and scheduling.
5. Overproduction. Develop optimized PMs, FMEA/RCM-based,
perform root cause analysis to reduce failures, etc.
6. Inappropriate Processing. Use the right tools and fixtures to
improve maintenance processes.
7. Defects. Eliminate rework and poor workmanship. Educate /
train maintenance personnel appropriately.

Most organizations, even without proclaiming a Lean maintenance


effort, might actually be engaged in the very activity that will get them
there. For example, TPM and Lean share many traits. Standardization, 5S,
and mistake-proofing are just a few other examples. More importantly,
TPM recognizes that the operator is just as responsible for asset reliabili-
ty as the maintenance person. One of the key objectives of TPM is to elim-
inate the six major losses: breakdown, set up and adjustments, idling and
minor stoppages, operating at reduced speeds, defects, and reduced yield.
It is to these losses that we employ a CMMS, work orders, planning and
scheduling, and other system tools in order to mitigate them.
Our challenge going forward is to identify activities which don’t add
value to maintenance and reliability. Use tools to analyze problems (waste
in the system), develop value-added solutions, and implement practices
that have been discussed earlier in the book to become a lean organization.

Value Stream Mapping (VSM)


Processes, both manufacturing and service, can be compared to a
river. They flow in a natural direction and carry material (products and
information) with them from one point to another. Now, visualize all that
a river carries with it. In addition to value-added elements such as water
Maintenance Analysis and Improvement Tools 393

and fish, it also carries other elements which may not be adding any val-
ues, or instead may be harming it. Similarly, our processes have many
non-value-added and wasteful activities which need to be identified and
eliminated. Value Stream Mapping identifies waste and helps to stream-
line the process for higher productivity.
VSM is a tool commonly used in continuous improvement programs
to help understand and improve the material and information flow within
processes and organizations. As part of the lean methodology, VSM cap-
tures the current issues and presents a realistic picture of the whole
process from end to end in a way that is easy to understand. To some,
VSM is a paper and pencil tool that helps us to see and understand the
flow of information and material as it makes its way through the value
stream. It helps to identify steps that are not adding any value — they are
waste and need to be removed from the process or improved.
VSM is a structured process and is usually carried out in eight steps.

Step 1 — Identify the problem and set expectations.


Step 2 — Select the team.
Step 3 — Select the process to be mapped.
Step 4 — Collect data and produce the current state map.
Step 5 — Critique / evaluate the current state. identify waste and
non-value added items. -
Step 6 — Map the future state;
Step 7 — Create an action plan and implement.
Step 8 — Measure and evaluate results; readjust the plan, if
needed.

Figure 11.11 is an example of the mapping process. In creating the


maps, sometimes standard symbols are used to identify wait, flow, opera-
tion, storage, etc. Also, the following color coding of activities is used to
display the value of activities:

• Green — Value-added activities


• Yellow — Non-value-added activities, but may be required to
meet regulatory or organizational requirements
• Red — Non-value-added activities (waste)

Value stream mapping is not a one-time event. Successful organiza-


tions apply value stream mapping continuously to their processes to get
better results.
394 Chapter 11

Figure 11.11 The Mapping Process

11.6 Other Analysis and Improvement Tools

The Theory of Constraints

The Theory of Constraints (TOC) is an improvement tool originally


developed by Eliyahu M. Goldratt and introduced in his book The Goal.
It is based on the fact that a system is like a chain with its weakest link.
At any point in time, there is most often only one aspect of that system
that is limiting its ability to achieve most of its goal. For that system to
attain any significant improvement, its constraint must be identified and
the whole system must be managed with this constraint in mind.
Therefore, if we want to increase throughput, we must identify and elim-
inate the constraint (or bottleneck).

Process Flow Types There are four primary process flow types in the
TOC lexicon. The four types can be combined in many ways in larger
facilities.

• A-Flow — Material flows in a sequence, such as in an assembly


line. The primary work is done in a straight sequence of events
(one-to-one). The constraint is the slowest operation. The general
flow of material is many-to-one, such as in a plant where many
sub-assemblies converge for a final assembly. The primary prob-
Maintenance Analysis and Improvement Tools 395

lem in A-Flow process is synchronizing the converging lines so


that each supplies the final assembly point at the right time.
• I-Flow — Material flows in a sequence, such as in an assembly
line. The primary work is done in a straight sequence of events
(one-to-one). The constraint is the slowest operation.
• T-Flow — The general flow is that of an I-Flow (or has multiple
lines), which then splits into many assemblies (many-to-many).
Most manufactured parts are used in multiple assemblies and
nearly all assemblies use multiple parts.
• V-Flow — The general flow of material is one-to-many, such as
a process that takes one raw material and makes many final prod-
ucts. Classic examples are meat rendering plants and steel manu-
facturers. The primary problem in V-Flow is “robbing” where one
operation (A) immediately after a diverging point “steals” mate-
rials meant for the other operation (B). Once the material has been
processed by operation A, it cannot come back and be run through
operation B without significant rework.

Once the key constraint has been identified, we need to do everything


possible to maximize the rate of flow through that constraint, and not let
any other portion of the system run at a faster rate. In a well-optimized
system, only one asset is running at full capacity and the remaining run at
only their partial capacity.

The key steps in implementing an effective TOC approach are:

1. Communicate the goal of the process or organization. For exam-


ple, “Make 100 units or more of X products per hour.”
2. Identify the constraint. This is the factor within a process or sys-
tem that prevents the organization from meeting that goal.
3. Exploit the constraint. Make sure that the constraint is dedicated
to what it uniquely adds to the process, and not doing things that
it should not do.
4. Elevate the constraint. If required, increase the capacity of the
constraint by adding additional equipment or reducing the set-up
time.
5. If, as a result of these steps, the constraint has moved to another
system, return to Step 1 and repeat the process for the new
system.
396 Chapter 11

Affinity Analysis or Diagram


The primary purpose of an Affinity analysis, also known as the KJ
method after its author Jiro Kawakita, is to organize ideas, data, facts,
opinions, and issues in naturally-related groups, specifically when a prob-
lem is complex.
An affinity diagram is the result of a creative process focused on find-
ing the major themes affecting a problem by generating a number of ideas,
issues, and opinions. The process identifies these ideas, groups naturally
related items, and identifies the one concept that ties each grouping
together. The team working on a problem reaches consensus by the cumu-
lative effect of individual sorting decisions rather than through discussion.
In many problem solving situations, brainstorming is a common tool
used to gather issues and ideas to solve problems. As a mechanism for
allowing a group of individuals to get ideas and issues on the table, brain-
storming is one of the best methodologies. However, too often, such ses-
sions generate large quantities of issues. These issues can become com-
plex to review and difficult to interpret. It can also be challenging to high-
light particular trends or themes from the gathered issues following a
brainstorming session.
A variety of methods are available to utilize gathered ideas efficient-
ly. Of these methods, Affinity diagrams represent an excellent tool both to
group ideas in a logical way and to capture themes that have developed
during the brainstorming. The following steps are used to develop Affinity
diagram (Figure 11.12):

Identify the problem or issue to solve.

a. Conduct a brainstorming meeting.


b. Record ideas and issues on “post-it notes” or cards.
c. Gather post-it notes and cards into a single place (e.g., a desk or
wall). Sort the ideas into similar groups, patterns, or themes
based on the team’s thoughts. Continue until all notes and cards
have been sorted and the team is satisfied with their groupings.
d. Label each group with a description of what the group repre-
sents and place the name at the top of each group.
e. Capture and discuss the themes or groups and how they may
relate.

Although affinity diagrams are not complicated, getting the most from
them takes a little practice. For example:
Maintenance Analysis and Improvement Tools 397

Figure 11.12 Preparing an Affinity Diagram

• Make sure that ideas and issues that have been captured are
understood. Usually brainstorming sessions have a habit of sim-
plifying issues or agreeing without understanding the concepts
being discussed.
• Do not place the notes in any order. Furthermore, do not deter-
mine categories or headings in advance.
• Allow plenty of time for grouping the ideas. Maybe post the
randomly-arranged notes in a public place and allow grouping to
happen over a few days, if needed.
• Use an appropriate number of groups within the diagram. Too
many can become confusing and unmanageable.

Affinity diagrams are great tools for assimilating and understanding


large amounts of information. The next time you are confronting a large
amount of information or number of ideas and feel overwhelmed at first
glance, use the affinity diagram approach to discover all the hidden link-
ages.

Barrier Analysis (BA)


Barrier analysis is a technique often used particularly in process
industries. It is based on tracing energy flows, with a focus on barriers to
those flows, to identify how and why the barriers did not prevent the ener-
gy flows from causing damage. Assets and systems generally have barri-
ers, defenses, or controls in place to increase their safety. Barrier analysis
398 Chapter 11

can be used to establish the type of barriers that should have been in place
to prevent the incident, or could be installed to increase system safety.
Therefore, Barrier Analysis can be used either proactively, to help to
design effective barriers and control measures, or reactively, to clarify
which barriers failed and why.
Barrier analysis offers a structured way to visualize the events related
to system failure. Barriers can be physical, human action, or system con-
trolled. Barriers can be classified as:
• Physical barriers
• Natural barriers
• Human action barriers
• Administrative barriers

To perform Barrier Analysis, identify the issue to be analyzed. For


example, consider the hydraulic supply line — 1 1/2 inch, 3000 psi
hydraulic fluid supply line — ruptured by overpressure.
List all the barriers that were in place, but failed to stop this rupture.
Consider the circumstances of the incident and assess the performance of
each barrier on this situation. For barriers that are assessed as having
failed, consider why they failed. Evaluate the impact of each failure on the
incident being analyzed. If the failure was causal, then efforts should be
focused on aspects of the system in terms of quality and safety improve-
ment. Record findings and establish improvements needed. Occasionally
detailed analysis has to be performed using the Fishbone, Five Whys, or
other techniques before recommendations should be made.

11.7 Summary

Organizations must continually improve processes, reduce costs, and


cut waste to remain competitive. To make improvements in any process,
data needs to be analyzed, utilizing tools and techniques to develop and
implement corrective actions. A variety of methods, techniques, and tools
are available to us, ranging from a simple checklist to sophisticated mod-
eling software. They can be used effectively to lead us to appropriate cor-
rective actions. We have discussed a few of these tools such as 5 whys,
cause–and-effects diagrams (aka fishbone diagram), Pareto analysis, root
cause analysis, FMEA, Six Sigma, Lean, and Value Stream Mapping.
Applying continuous improvement tools discussed can optimize work
processes and help any organization to improve results, regardless of the
size or type of business environment.
Maintenance Analysis and Improvement Tools 399

11.8 Self Assessment Questions

Q11.1 Into what categories can root cause analysis be classified?

Q11.2 What steps are needed to perform an RCA?

Q11.3 When should we use the fishbone tool?

Q11.4 How is a fishbone diagram constructed? Explain the steps that


are needed.

Q11.5 What is the purpose of the FMEA?

Q11.6 What key steps / elements are needed to perform an FMEA?

Q11.7 What does 6 Sigma mean?

Q11.8 If a process is performing at 5 Sigma level, what percent of


defective parts can be expected?

Q11.9 What’s is the difference between Deming’s improvement cycle


and DMAIC?

Q11.10 What is the primary objective of Pareto analysis?

11.9 References and Suggested Reading


FMEA, 3rd Edition. Reference manual AIAG.
Harry, Mikel and Richard Schroeder. Six Sigma: The Breakthrough
Management Strategy. Currency Publications, 2000.
Koch, Richard. The 80/20 Principle: The Secret to Success by Achieving
More with Les. Currency Books, 1998.
Latino, Robert and Kenneth Latino. Root Cause Analysis: Improving
Performance for Bottom Line Results. CRC Press, 2002.
Tague, Nancy R. The Quality Tool Box. The Quality Press, 2005.

399
Chapter 12
Current Trends and Practices

“Progress is impossible without change;


and those who cannot change their minds, cannot change anything”
—George Bernard Shaw

12.1 Introduction
12.2 Key Terms and Definitions
12.3 Energy Management, Sustainability, and the Green Initiative
12.4 Personnel, Facility, and Arc Flash Safety
12.5 Risk Management
12.6 Corrosion Control
12.7 Systems Engineering and Configuration Management
12.8 Standards and Standardization
12.9 Summary
12.10 Self Assessment Questions
12.11 References and Suggested Reading

After reading this chapter, you will be able to understand:


• Why M&R leaders should worry about energy and sustainability
or “being green”
• What M&R leaders should do about safety and, more
specifically, arc flash safety
• How we go about managing the risks around us
• What is meant by corrosion control
• The effects of standards and standardization on your
organization
401
402 Chapter 12

• What exactly are systems engineering and configuration


management

12.1 Introduction

As stated in the previous chapter, businesses must continually


improve processes to remain competitive. One key means to accomplish-
ing this is through awareness of current trends and innovative best prac-
tices as they enter the maintenance and reliability industry. Chapter 12
captures a few of these trends and practices and briefly discusses how they
might apply in the areas of energy and sustainability, safety, risk manage-
ment, corrosion control, systems engineering/configuration management,
and standards. Of course, as with any trend or practice, further investiga-
tion is necessary to understand not only the context in which a particular
industry or business is applying these trends and practices, but also how
these trends and practices could be applied and implemented to your own
organization.

12.2 Key Terms and Definitions

American National Standards Institute (ANSI)


A not-for-profit, non-government organization which develops
and publishes standards for almost the entire U.S. industry; its
standards-setting procedure provides a forum for discussion
among academics, special interest groups, users, and vendors;
an ANSI standard, though termed voluntary, carries a lot of
clout in the United States and elsewhere.

Configuration
The arrangement and contour of the physical and functional
characteristics of a system, equipment, and related hardware or
software. It also includes controlling and documenting changes
made to the functional characteristics and layout

Configuration Management
A discipline applying technical and administrative direction and
surveillance to identify and document the functional and physi-
cal characteristics of an asset / system called a configuration
item; control changes to those characteristics; and record and
report changes.
Current Trends and Practices 403

Corrosion
Gradual alteration, degradation, or eating away of a metal due to
a chemical or electrochemical reaction between it and its envi-
ronment.

Green Energy
A type of energy that is considered to be environmentally friend-
ly and non-polluting, such as hydro, geothermal, wind, and solar
power.

Hazard
A condition that is prerequisite to a mishap and a contributor to
the effects of the mishap.

Hazardous Material
Any substance that is listed as corrosive, harmful, irritant, reac-
tive, toxic, or highly toxic.

LEED
Leadership in Energy and Environmental Design.

Mishap
An unplanned event or series of events resulting in death, injury,
occupational illness, or damage to or loss of equipment or prop-
erty, or damage to the environment.

Mitigation
A method that eliminates or reduces the consequences, likeli-
hood, or effects of a hazard or failure mode; a hazard control.

Personnel Protection Equipment (PPE)


Safety equipment issued to help employees in protecting them-
selves from the hazards of their work environments. PPE
includes fire retardant or chemical-proof clothing, gloves, hard
hats, respirators, safety glasses, etc.

Risk
A future event that has some uncertainty of occurrence and neg-
ative consequence if it were to occur.
404 Chapter 12

Risk Assessment
The determination of quantitative or qualitative value of risk
related to a concrete situation and a recognized threat (also
called hazard). Quantitative risk assessment requires calcula-
tions of two components of risk: consequence, the magnitude of
the potential loss, and likelihood, the probability that the loss
will occur.

Risk Disposition
One of several different ways to address identified risk.

Risk Index (Magnitude)


Consequence (impact) of risk event X’s probability of occur-
rence.

Risk Management
A continuous process that is accomplished throughout the life
cycle of a system to:
• Identify and measure the unknowns.
• Develop mitigation options.
• Select, plan, and implement appropriate risk mitigations.
• Track the implementation of risk mitigations to ensure
successful risk reduction.

Risk Type
One of several risk attributes: residual, transferred, assumed,
avoided.

Sustainability
The ability to maintain a certain status or process in existing
systems; in general refers to the property of being sustainable;
capacity to endure.

Validation
The act of determining that a product or process, as constituted,
will fulfill its desired purpose.

Verification
The process of assuring that a product or process, as constitut-
ed, complies with the requirements specified for it.
Current Trends and Practices 405

12.3 Sustainability, Energy Management,


and the Green Initiative

What is Sustainability?
Sustainability in general refers to the property of being sustainable.
The widely accepted definition of sustainability or sustainable develop-
ment was given by the World Commission on Environment and
Development in 1987. It defined sustainable development as “forms of
progress that meet the needs of the present without compromising the
ability of future generations to meet their needs.”
Over the past 25 years, the concept of sustainability has evolved to
reflect perspectives of both the public and private sectors. A public policy
perspective would define sustainability as the satisfaction of basic eco-
nomic, social, and security needs now and in the future without undermin-
ing the natural resource base and environmental quality on which life
depends. From a business perspective, the goal of sustainability is to
increase long-term shareholder and social value, while decreasing indus-
try’s use of materials and reducing negative impacts on the environment.
Common to both the public policy and business perspectives is the
recognition of the need to support a growing economy while reducing the
social and economic costs of economic growth. Sustainable development
can foster policies that integrate environmental, economic, and social val-
ues in decision making. From a business perspective, sustainable develop-
ment favors an approach based on capturing system dynamics, building
resilient and adaptive systems, anticipating and managing variability and
risk, and earning a profit. Sustainable development reflects not the trade-
off between business and the environment but the synergy between them.
Practically, sustainability refers to three broad themes, also called pil-
lars, of sustainability: economic, social and environmental. They must all
be coordinated and addressed to ensure the long-term viability of our
community and the planet. The sustainability issue has emerged as a result
of significant concerns about the unintended social, environmental, and
economic consequences of rapid population growth, economic growth,
and the increasing consumption of our natural resources. When consider-
ing existing or new individual, business, industrial and community prac-
tices or projects, one must ensure that economic, social, and environmen-
tal benefits are achieved. Each person, business, and industry has a role
and a responsibility to ensure their individual and collective actions sup-
port the sustainability of our community.
406 Chapter 12

This also means that we must preserve our resources in such a way
that human beings in the future can enjoy them as well. To achieve this,
we must regenerate our resources at a rate that is equal to or faster than
our consumption.
Social sustainability stems from the fact that multiple cultures and
societies all share and inhabit the same planet. These cultures may be dif-
ferent in their histories, backgrounds, and beliefs, but each brings a differ-
ent perspective to the world around them. Therefore, considering the
social side of resources (whether they be land or other physical resources)
must play a part into the overall sustainability equation.
Environmental sustainability is important because it involves natural
resources that human beings need for economic or manufactured capital.
Materials taken from nature are used for solutions that address human
needs. If nature is depleted faster than it can regenerate, human beings
will be left without raw materials. Furthermore, environmental sustain-
ability also involves ensuring that waste emissions are at volumes that
nature can handle. If not, many humans and other living things on Earth
may be harmed to the point of extinction.
Sustainability is really based on a simple principle: Everything that
we need for our survival and well-being depends, either directly or indi-
rectly, on our natural environment. Sustainability creates and maintains
the conditions under which humans and nature can exist in productive har-
mony, that permit fulfilling the social, economic, and other requirements
of present and future generations.

6 Easy Ways to Make a Big Difference in Sustainability


To ensure that we have and will continue to have, the water, materi-
als, and resources to protect human health and our environment, we need
to be aware of sustainability challenges. With these concerns in mind we,
as individuals and as M&R professionals, can act on our own even if a
large scale effort is needed. Our own individual actions can make a real
difference. Below are a few actions we can turn into habits so that we can
start to adopt a more sustainable lifestyle.

1. Reduce what we buy and what we use. Manufacturing uses up


precious natural resources. Everything that we own comes from
raw materials that were once part of the earth. Every object in our
household made an impact on the environment when it was
stripped from the earth and produced. Choosing to buy less is
Current Trends and Practices 407

obvious as a sustainable strategy, but is not always practical.


Sometimes the best choice is to buy well-made and durable items
made with environmentally-friendly material.
2. Save on electricity-energy. Most electricity is generated using
fossil fuels. This makes electrical generation one of the biggest
contributors of greenhouse gasses to the ecosystem. By using less
electricity we can reduce these emissions. There are lots of ways
to reduce electrical consumption such as using energy efficient
appliances, devices, and processes in the plants. (Reducing and
managing energy consumption will be discussed in more detail
later in this section.)
3. Generate less waste. Everything we throw away goes some-
where. Reducing the amount of waste that goes into landfills is an
excellent way of protecting the environment. Recycling and
reusing also reduce what we need to throw away.
4. Reduce water resources. Water is one of the most precious
resources on our planet. Take care of leaking taps, pipes, and toi-
lets. Turn off faucets that are running unnecessarily. Find ways to
reduce the use of water in our processes. Water that has been used
in bathing or washing can also be reused as gray water.
5. Minimize or eliminate need of hazardous material. In our
homes and plants, we use a lot of hazardous – environmentally
bad material such Freon, TCE – trichloroethylene, paints-sol-
vents, etc. We need to replace that material with environmentally
friendly material.
6. Choose greener transportation anytime it is feasible. When we
can reduce or eliminate the greenhouse emissions produced by an
internal combustion engine, we can help in reducing greenhouse
emissions. Taking advantage of public transportation, carpooling,
and riding a bike to work are all geared toward postponing the day
when fossil fuels and life as we know it may be gone.

The Environmental Protection Agency (EPA) has acted primarily as


the United States’ environmental watchdog, striving to ensure that indus-
tries met legal requirements to control pollution. In subsequent years, the
EPA began to develop theory, tools, and practices that enabled it to move
from controlling pollution to preventing it. Today the EPA aims to make
sustainability the next level of environmental protection by drawing on
advances in science and technology to protect human health and the envi-
ronment, and promoting innovative green business practices.
408 Chapter 12

Key Government Regulations and Practices

Executive Order 13423: “Strengthening Federal Environmental, Energy,


and Transportation Management” of 2007 set policy and specific goals for
federal agencies to “conduct their environmental, transportation, and
energy-related activities under the law in support of their respective mis-
sions in an environmentally, economically and fiscally sound, integrated,
continuously improving, efficient, and sustainable manner.”

Executive Order 13514: “Federal Leadership in Environmental, Energy,


and Economic Performance” of 2009 enhances EO 13423 “to establish an
integrated strategy towards sustainability in the Federal Government and
to make reduction of greenhouse gas emissions (GHG) a priority for
Federal agencies.”

The Federal Government Sustainability website includes the latest infor-


mation from federal agencies relevant to developing and maintaining sus-
tainable facilities and to developing and promoting sustainable practices
within their environmental programs.

Greening EPA. EPA implements a wide range of programs to reduce the


environmental impact of its facilities and operations, from building new,
environmentally sustainable structures to improving the energy efficiency
of older buildings.

The Cost of Energy


Energy costs can have a significant impact on the financial perform-
ance of businesses. A recent poll taken by the National Association of
Manufacturers (NAM) revealed that over 90% of small and medium-sized
manufacturing companies believe that higher energy prices are having a
negative impact on their bottom line. In fact, the energy crisis of 2008
caused many businesses to go under. As a result, organizations are review-
ing and updating energy plans to reduce their energy usage.
Throughout the manufacturing process, energy is lost due to equip-
ment inefficiency and mechanical and thermal limitations. Optimizing the
efficiency of these systems can result in significant energy and cost sav-
ings and reduced carbon dioxide emissions. Understanding how energy is
used and wasted—or energy use and loss footprints—can help plants pin-
point areas of energy intensity and ways to improve efficiency. Substantial
opportunities exist to reduce energy wasted in the industrial and service
sectors. Organizations are affected directly by the energy cost of manufac-
Current Trends and Practices 409

turing products, maintaining operations – including offices and receiving


raw materials — and delivering finished goods to the customers.
Energy use can also have significant environmental cost. Onsite com-
bustion of fuels in boilers, ovens, vehicles, and equipment can emit a vari-
ety of regulated pollutants, including carbon dioxide (CO2), carbon
monoxide (CO), sulfur dioxide (SO2), nitrogen oxide (NOx), particulate
matter, volatile organic compounds (VOCs), and a variety of airborne tox-
ins. Combustion pollutant emissions can affect worker health, and trigger
the need for costly permitting, monitoring, and emission controls. More
broadly, reducing air emissions from combustion activities as well as the
safe storage and handling of fuels, spent fuel, and other fluids can help
protect not only workers but also neighboring communities and public
health.
Energy is a vital and often costly input to most production processes
and value streams. Think unnecessary energy usage as another “deadly
waste,” and develop plans to eliminate or reduce it to achieve energy and
environmental excellence. Benefits of Energy Management are:

• Reduced operating and maintenance costs


• Reduced vulnerability to energy and fuel price increases
• Enhanced productivity
• Improved safety
• Improved employee morale and commitment
• Improved environmental quality
• Reduced greenhouse gas emissions
• Remain below air permitting emission thresholds
• Increased overall profit

Energy Sources and End Usage


The energy supply chain begins with electricity, steam, natural gas,
coal, and other fuels supplied to a manufacturing plant from off-site power
plants, gas companies, and fuel distributors. Energy then flows to either a
central energy generation utility system or is distributed immediately for
direct use. Industrial energy systems account for roughly 80% of all ener-
gy used by the industry. On average, 35% of that energy is lost every year
due to inefficient processes and waste. As much as 50 % of this could be
saved by improving the efficiency and reducing energy losses in these sys-
tems. Energy is then processed using a variety of highly energy-intensive
systems, including steam, process heating, and motor-driven equipment
such as compressed air, pumps, and fans. Industrial energy systems are cat-
410 Chapter 12

egorized by the Department of the Energy (DOE) in much the same way:

1. Steam
2. Process Heat
3. Motors, pumps and fans
4. Compressed air

Steam
Over 45% of all the fuel burned is consumed to raise steam. Steam is
used to heat raw materials and treat semi-finished products. It is also a
power source for equipment, as well as for building heat and electricity
generation. Many manufacturing facilities can recapture energy through
the installation of more efficient steam equipment and processes. The
whole system must be considered to optimize energy and cost savings.

Process Heating
Process heating is vital to nearly all manufacturing processes, supply-
ing heat needed to produce basic materials and commodities. Heating
processes consume nearly 20% of all industrial energy use. Advanced
technologies and operating practices offer significant savings opportuni-
ties to reduce process heating costs.

Motors, Pumps and Fans


Motor-driven equipment accounts for about 65% of the electricity
consumed in the industrial sector. Within the nation’s most energy-inten-
sive industries—improvements to motor systems could yield dramatic
energy and cost savings. The key to these savings is applying energy-effi-
ciency equipment or implementing sound energy management practices.

Compressed Air
The compressed air systems account for an estimated $5 billion per
year in energy costs in the U.S. industrial sector. Many industries use
compressed air systems as power sources for tools and equipment used for
pressurizing, atomizing, agitating, and mixing applications. The major
source of waste for this type of energy is leakage. Many users at the shop
floor believe the myth that air is free or costs very little. Optimization of
compressed air systems can provide energy efficiency improvements of
20%–50%.
Although the predominant energy sources used in industry are natural
Current Trends and Practices 411

gas and electricity, industry also uses other energy sources, such as fuel
oil, for producing heat. Some facilities have on-site co-generation, where
they combust a fuel (e.g., natural gas, waste oil, or scraps) to produce heat
and electricity. Understanding the energy end usage —what work we use
the energy to do—reveals more useful information to identify opportuni-
ties for improving efficiency and reducing costs. In an office setting, end-
uses primarily include heating, ventilating, and air conditioning (HVAC),
lighting, and operation of appliances and computers. In an industrial plant,
end uses primarily include process equipment operation, process heating
and cooling, transportation, HVAC, and lighting.
Understanding the costs of energy use can raise awareness of the
potential value of identifying and eliminating energy waste. The costs of
energy use are not always visible to production/operations managers
because they are rolled up into facility overhead costs, rather than
assigned to production areas. Explicitly tracking costs associated with
individual processes or equipment can encourage energy conservation.

Walkthrough Practice to Observe Energy Usage


Walkthrough assessments and observing processes as they actually
run at a facility can be a simple, but effective way to identify waste and
find improvement opportunities. During the walkthrough, look for signs
of unnecessary or inefficient energy use. An IR camera and an Ultrasonic
leak detector gun are good to identify hot/cold spots and leaks. Any num-
ber of questions could be asked about energy usage (with a few listed
below according to categories):

Motors and Machines

1. Are machines left running when not in operation? If so, why?


2. Are energy efficient motors, pumps, and equipment used?
3. Are motors, pumps, and equipment sized according to their
loads? Do motor systems use variable speed drive controls?

Compressed Air

1. If compressed air is used, do we notice any leaks in the com-


pressed air system? When was the last air leak audit performed?
2. Do compressed air systems use the minimum pressure needed to
operate equipment?
412 Chapter 12

Process and Facility Heating and Cooling

1. Are oven and process heating temperatures maintained at higher


levels than necessary?
2. Are work areas heated or cooled more than necessary?
3. Do employees have control over heating and cooling in their work
areas?
4. Are exterior windows or doors opened to adjust heating and
cooling?

Lighting

1. Is lighting focused where we need it?


2. Is lighting controlled by motion sensors in warehouses, storage
areas, and other areas that are intermittently used?
3. Are energy-efficient fluorescent light bulbs used?

Energy Audits and Measuring Energy Usage

While a walkthrough is an excellent way to identify and fix energy


waste that are readily apparent, we may be leaving energy savings on the
table unless we examine energy use more closely. Two strategies for
learning more about its use include:

1. Conduct an energy audit to understand how energy is used — and


possibly wasted — across facility.
2. Measure the energy use of individual production and support
processes.

An energy audit, sometimes referred to as an energy assessment, is a


study of the energy end-uses and performance of a facility. Energy audits
can range in complexity and level of detail, from a simple audit involving
a facility walkthrough and review of utility bills, to a comprehensive
analysis of historical energy use and energy-efficiency investment
options. Energy audits allow managers to compare a plant’s energy use to
industry benchmarks and to identify specific energy savings opportuni-
ties. In many locations, local utilities provide energy audit services for
free or at reduced cost.
Current Trends and Practices 413

Energy Reduction and Process Improvement Strategies


Many energy efficiency best practices can be implemented without
extensive analysis or planning. In plant operations, several strategies can
be employed to reduce energy usage such as:

1. Total Productive Maintenance (TPM). Incorporate energy


reduction best practices into day-to-day autonomous mainte-
nance activities to ensure that equipment and processes run
smoothly and efficiently.
2. Right-Sized Equipment. Replace oversized and inefficient
equipment with smaller equipment tailored to the specific needs
of manufacturing.
3. Plant Layout and Flow. Design or rearrange plant layout to
improve product flow while also reducing energy usage and
associated impacts.
4. Standard Work, Visual Controls, and Mistake-Proofing.
Sustain and support and energy performance gains through stan-
dardized work procedures and visual signals that encourage
energy conservation.

More details are discussed below on a couple of these strategies.

Replace Over-Sized and Inefficient Equipment with Right-Sized


Equipment
Process Improvement often results in the use of right-sized equipment
to meet production needs. Right-sized equipment is designed to meet the
specific needs of manufacturing or an individual process step, rather than
the processing needs for an entire facility. For example, rather than rely-
ing on one large paint booth or parts cleaning tank station to service all
painting and degreasing needs for a facility, Lean principles typically lead
organizations to shift to right-sized paint and degreasing stations that are
embedded in manufacturing cells.
In conventional manufacturing, equipment/systems often are over-
sized to accommodate the maximum anticipated demand. Because pur-
chasing a new, large piece of equipment is often costly and time-consum-
ing, engineers often design in additional “buffer capacity” to be sure that
the equipment does not constrain the production. For example, a fan sys-
tem is usually oversized. Ways in which it could be correctly sized to
reduce energy are:
414 Chapter 12

a. Use smaller, energy-efficient motors. Right-sizing a 75-horse-


power (hp) standard efficiency motor with a 60-hp energy-effi-
cient motor will reduce motor energy consumption by about 25
percent or more.
b. Reduce fan speed by larger pulleys. Replacing an existing belt-
driven pulley with a larger one will reduce its speed, thereby sav-
ing energy costs. Reducing a fan’s speed by 20 percent reduces its
energy consumption by 50 percent.
c. Use of static pressure adjustment for variable air volume
(VAV) systems. Reducing static pressure in VAV system reduces
the fan power consumption. By gradually reducing the static pres-
sure set point to a level low enough to keep occupants comfort-
able, energy consumption can be reduced.

Design Plant Layout to Improve Flow and Reduce Energy Usage


Process improvement focuses on improving the flow of product
through the production process. Arrange equipment and workstations in a
sequence that supports a smooth flow of materials and components
through the process, with minimal transport or delay. The desired outcome
is to have the product move through production in the smallest, quickest
possible increment (one piece). Improving the flow of product and
process inputs can significantly reduce the amount of energy required to
support a production process.
An example of a good design is to use large pipes and small pumps
rather than small pipes and big pumps. Optimizing the whole system
together will lead to dramatically decreased operating costs. The objective
is to minimize friction losses.
In addition to explicitly using process methods to target energy
wastes, facilities can take advantage of other opportunities for energy sav-
ings to install energy-efficient equipment, switch to less polluting fuel
sources, and design products to use less energy. To be most effective,
energy saving efforts should be proactive, strategic, and systematic to
establish an energy management system that aligns with and supports the
organization’s initiatives to achieve the greatest improvements in opera-
tional, energy, and environmental performance.
Identifying and eliminating energy waste through process improve-
ment including Lean and Green initiatives can improve an organization’s
ability to compete in several ways. For example, reducing the energy
intensity of production activities and support processes directly lowers
recurring operating costs with direct bottom line and competitiveness
impacts.
Current Trends and Practices 415

There are three steps involved in developing an energy planning and


management roadmap appropriate to any organization:

1. Initial Assessment. Consider the opportunities, risks, and costs


associated with strategic energy management.
2. Design Process. Understand the organization’s energy needs and
identify the best way to establish an energy management plan.
3. Evaluate Opportunities. Identify and prioritize energy-related
improvement opportunities, such as energy-efficiency actions,
energy-supply options, and energy-related products and services.

Finally, a new energy management standard has been issued by the


International Organization for Standardization (ISO) recently. The stan-
dard known as ISO 50001:2011 establishes a framework to manage ener-
gy for industrial plants; commercial, institutional, or governmental facili-
ties; or entire organizations. Targeting broad applicability across national
economic sectors, it is estimated that the standard could influence up to
60 % of the world’s energy use.
Energy usage is often viewed as a necessary support cost of doing
business, and energy-efficiency efforts can sometimes have difficulty
competing for organizational attention with other operational needs. By
linking energy management to Environmental / Green and Lean activities,
energy-reduction efforts can be tied more directly to process improvement
efforts that are regarded by senior managers as being vital to business suc-
cess.

The Green Initiative


Green energy is a term used to describe sources of energy that are
considered to be environmentally friendly and non-polluting, such as
geothermal, wind, and solar power. These sources of energy may provide
a remedy to the alleged effects of global warming and certain forms of
pollution. They are generally more expensive than traditional energy
sources, but can be purchased with the help of government subsidies.
Several working definitions used for green energy include:

• An alternate term for renewable energy


• Energy generated from sources which do not produce pollutants
(e.g., solar, wind, and wave energies)
• Energy generated from sources that are considered environmen-
tally-friendly (e.g., hydro (water), solar (sun), biomass (landfill),
or wind)
416 Chapter 12

• Energy generated from sources that produce low amounts of


pollution
• Energy that is produced and used in ways that produce relatively
less environmental impact

The Green Building Initiative (GBI) is another initiative to reduce


energy usage and environmental impact. GBI challenges state govern-
ments to demonstrate leadership in energy efficiency and environmental
responsibility in state buildings, while also reducing the environmental
impact state facilities have on our world.
GBI requires the state to reduce grid-based energy usage in its build-
ings 20% by 2015, and, in so doing, reduce greenhouse gas emissions
associated with the production of fossil fuel-based power required to oper-
ate those same buildings.

Leadership in Energy and Environmental Design (LEED) is an


internationally recognized green building certification system, providing
third-party verification that a building or community was designed and
built using strategies intended to improve performance in metrics such as
energy savings, water efficiency, CO2 emissions reduction, improved
indoor environmental quality, and stewardship of resources and sensitivi-
ty to their impacts.
Developed by the U.S. Green Building Council (USGBC), and spear-
headed by LEED founding chairman, Robert K. Watson, LEED is intend-
ed to provide building owners and operators a concise framework for
identifying and implementing practical and measurable green building
design, construction, operations and maintenance solutions.
A LEED-certified building uses significantly less energy and water,
and produces fewer greenhouse gas emissions, than conventional con-
struction. Many state and federal government agencies are mandated to
meet a minimum of LEED Silver certification for new construction and
major renovations of facilities larger than 10,000 square feet. In addition,
smaller buildings are being designed to meet LEED standards. LEED cer-
tification is obtained after submitting an application documenting compli-
ance with the requirements of the rating system as well as paying registra-
tion and certification fees. Certification is granted solely by the Green
Building Certification Institute (GBCI), which is responsible for the third
party verification of project compliance with LEED requirements. See the
details at www.GBCI.org.
Current Trends and Practices 417

The Green Building Certification Institute (GBCI) is a third-party


organization that provides independent oversight of professional creden-
tialing and project certification programs related to green building. GBCI
is committed to ensuring precision in the design, development, and imple-
mentation of measurement processes for green building performance
(through project certification) and green building practice (through pro-
fessional credentials and certificates). Established in 2008 to administer
certifications and professional designations within the framework of the
U.S. Green Building Council’s LEED® Green Building Rating
Systems™, GBCI continues to develop new programs and offer the mar-
ketplace validation that building certifications and professional designa-
tions have met specific, rigorous criteria.

12.4 Personnel, Facility, and Arc Flash Safety

The Relationship of Safety and Reliability


Several reliability and safety experts have observed that reliable
plants are safe plants and safe plants are reliable plants. Furthermore, safe
and reliable plants are usually profitable plants. Safety and reliability have
historically been considered two separate elements of the production oper-
ations system. Recently, both have been proven to be increasingly interre-
lated. In fact, safety is treated as the most important attribute in reliabili-
ty analysis.
Ron Moore, a leading M&R expert and noted author, writes that there
is a strong correlation — an inverse correlation coefficient of 0.87 —
between OEE/Uptime and accident (injury) rate per 100 employees.
Overall Equipment Effectiveness (OEE), a product of equipment avail-
ability, quality, and performance, is a key indicator of reliability and oper-
ational performance. This conclusion is based on his observation and data
from many plants that he had visited for his consulting tours.
A study by Batson, Ray, and Quan reported in the October 2000 issue
of Reliability Magazine made similar observations. This study indicated
that in the organizations where maintenance performance ratings have
increased tenfold, injury frequency and severity have been reduced ten-
fold, in a nearly inverse linear fashion.
Another observation reported in PIMA’s June 2003 conference in
New York referred to a study by a major Pulp and Paper company that
found the company was 28% more likely to have an incident when main-
tenance work was reactive versus work that was planned and scheduled
before execution. The author’s own observations conclude a strong corre-
418 Chapter 12

lation between safety incidents/injuries and reactive maintenance. In a


reactive situation, we might not take the time we should to plan and think
before we take action. The urgent nature of reactive work also pushes
maintenance personnel to take risks they shouldn’t be taking. These obser-
vations strongly imply that organizations that are reliable with excellent
maintenance practices will have lower injury rates. The same behavior
and practices that improve plant operations and reliability also reduce
injuries. Therefore, those organizations with high reliability and safe oper-
ations will be more productive and profitable.

Creating a Safety Culture and a Culture That Cares


When Rosanne Danner, Vice President of Development for DuPont
Safety Resources, was asked to describe her definition of a safety culture,
she noted that she had once asked the same question to one of her col-
leagues. He replied, “What people do when no one is watching.” There is
certainly some truth to that. But if we step back and think about the word
culture, it’s about what people do, how they interact, and how they live
day to day. When we apply it to safety, it moves beyond simply being a
program and becomes part of one’s being. A Safety culture doesn’t just
stay at the organization or workplace. It goes home with us. It’s part of the
fabric of who we are. For example, when I drive home, I automatically put
on my seatbelt. I also make sure that my passengers wear their seatbelts.
I do not use my cell phone when I am driving. These actions are all a nat-
ural extension of following a safety culture at work.
As we have discussed, reliability is not just the responsibility of the
maintenance department, but is for everyone in the organization including
operators, planners, supervisors, designers, material/store handlers, , and
purchasing managers — and includes the organization’s leadership team.
Similarly, safety is not just the responsibility of the safety department. It
is for every one of us to be responsible for our own and our coworker’s
safety. Leadership plays a key role in ensuring that we understand our role
in keeping the workplace safe. Safety and reliability — a good steward-
ship of our resources — should be part of an organization’s core values.
The leadership should become a role model by doing, not just talking.
Consider a scenario where we are walking to attend a meeting in the
plant area. On the way, we find a puddle of water or oil spill on the shop
floor near an assembly area. Should we stop and take care of this spill
before somebody else slips and gets hurt? We are already late to the meet-
ing. We could just keep going and hope that somebody will take care of
this spill. What should we do?
Current Trends and Practices 419

It’s simple. STOP! Get somebody to take care of this potential hazard
before proceeding to the meeting. Yes, we will be late. It’s OK. Apologize
to the meeting group and tell them the truth — the reason for being late.
Also, on the way back, ensure that the hazard has been taken care of and
somebody is finding out its root cause. That’s safety culture.
Instituting a safety culture must begin at the top of the organization.
However, all employees have a responsibility to follow procedures and
think about how they do their work. Usually, we start with an organization
in the reactive stage, where employees are reacting to incidents instead of
thinking about how to prevent or eliminate them. Once employees begin
to view safety as something important to them and something which they
value, they move to the independent stage. This is where they are practic-
ing safety because they want to do it, not because they are being told to
do it. The ultimate goal is the interdependent stage when every employee
is looking out for the other. It’s a “brother’s keeper” mentality. At this
stage, any employee should be comfortable to call out a safety issue to the
point where they will stop a production line if they see a problem, or chal-
lenge a manager who, for example, isn’t wearing a hard hat.
The State of Montana has done a unique thing to create a safety cul-
ture. A Safety Culture Act was enacted in 1993 by the Montana state leg-
islature to encourage workers and employers to come together to create
and implement a workplace safety philosophy. The intent of this act is to
raise workplace safety awareness to a preeminent position in the minds of
all of Montana’s workers and employers. It becomes the responsibility
and duty of the employers to participate in the development and imple-
mentation of safety programs that will meet the specific needs of their
workplace — thereby establishing a safety culture that will help to create
a safe work environment for all future generations of Montanans.

A Safety Process Model


Adhering to a simple process model is another highly effective com-
ponent of an overall strategy for improving the safety in an organization.
The model below focuses on four aspects of safety:
1. Leadership. As stated earlier, leadership involvement is impor-
tant. Leaders must lead and support the safety process whole-
heartedly. They must communicate the importance of safety as
well as the value and respect they have for the people who work
for them.
2. Personnel. Investing in people is paramount to success. The best
organizations will first seek to hire the right people and then
420 Chapter 12

develop their capabilities and skill sets. Be sure to include ques-


tions about safety as part of the hiring process, to gain an under-
standing of a prospective employee’s knowledge of safety, and to
communicate your organization’s commitment to safety.
3. Environment. It is essential to ensure that the overall environ-
ment is safe, assets and systems are properly cared for, operating
practices are adhered to, and engineering standards are followed.
Conduct a design safety review of all equipment from inception
and a full ergonomic review before installation and continue
annually after that. Establish extensive inspection programs to
ensure compliance and be on the lookout for new technologies to
reduce risk.
4. Behavior. Changing organizational behavior is what transforms
an organization from good to world class. When passion for safe-
ty is driven by the leadership team, it filters down to the floor and
will encourage workers to actively care about one another while
fostering interdependence within the organization.

Turn Employees into Safety Leaders


To be successful, organizations should create career paths that turn
employees into safety leaders by making sure that everyone is highly
trained and motivated — not just to succeed, but to exceed expectations.
Workers should be mentored, to help them contribute to the safety
process. The organization should also develop an environment and culture
that supports the belief that every employee can create and maintain a
workplace free of illness and injury. The result of this investment will be
establishing within workers a sense of ownership of the safety process and
a shift within the organization from an independent to an interdependent
culture. This can help drive employees to eliminate unsafe behaviors and
conditions and to focus on eliminating injuries entirely, rather than just
meeting regulatory requirements.
According to OSHA, when a company’s safety culture is strong,
“everyone feels responsible for safety and pursues it on a daily basis;
employees go beyond ‘the call of duty’ to identify unsafe conditions and
behaviors, and intervene to correct them.”
Consider posting the following safety principles throughout the plant /
facility to remind employees of the importance organization places on
safety:
1. Any person can and must confront unsafe behaviors and condi-
tions. No one is authorized to disregard such a warning.
Current Trends and Practices 421

2. No one is expected to perform any function or accept any direc-


tion that they believe is unsafe to themselves or others, or cre-
ates an unsafe situation, regardless of who directs such an
action.
3. Anyone who feels that a process is unsafe will shut down that
process and work with appropriate team members to create a
safe situation.

An organization’s greatest asset is its employees, and protecting them


from illnesses or workplace injuries is critical to success. Operating an
injury-free facility is no longer a dream. In many workplaces, it has
become a reality — and not just for a year, but for several years running.
Creating a workplace that is free of illness and injury begins with one cru-
cial decision: making safety a core value.
Many organizations such as DuPont, Kimberly-Clark, Harley-
Davidson, General Mills, Milliken, and Jacobs Engineering have created
a safety culture in their organizations. They have been able to reduce their
injury and incident rate below 1 per 100 employees. In fact, their goal is
zero injury. Jacobs calls this new initiative “Beyond Zero” to create a
“Culture of Caring.” These results can be attributed to a culture that
embraces safety and empowers employees to maintain a commitment to
safety in everything they do. The key to this success is establishing a safe-
ty-based culture that starts at the top.
In a recent talk at the IMC (International Maintenance Conference),
Bart Jones, Director of Facilities O&M with ATA-Jacobs at Arnold
Engineering and Development (Test) Center, explained his view of the
culture of caring this way
“...when my work family members (employees) go home
each day, I want them to be in better shape, not just physical-
ly but mentally too, than when they came to work. Work is
very important, but it’s not the most important thing in our
lives....
“This can only be achieved when we think of our work
family just like we think of our spouse, parents, children or
grandparents. When we start asking ourselves if we would
want our daughter in the current work environment or if we’d
send our grandmother on the next task we’re about to per-
form, then we’ve really achieved what the Culture of Caring
is all about — respect, treating each other as family, and truly
caring about all aspects of the folk’s lives that we work with
422 Chapter 12

and the ramifications of our decisions on those lives. That


translates to a whole new outlook on how we approach relia-
bility, safety, and maintainability among many other areas.
We’ll begin to look at designing equipment with safety and
maintainability (ergonomics) in mind. We’ll evaluate our
operations and maintenance processes and procedures to
ensure we don’t take un-necessary risks. And we’ll ensure our
folks understand that we all take responsibility for each
other.”

Bart takes this subject very passionately. It’s evident when he talks
about it and in his actions on the programs he runs for ATA-Jacobs. He
speaks from the heart because of personal experience. He lost his 23-year-
old brother in an industrial accident and he knows first-hand how an
injury affects a widespread net of friends and family for the rest of their
lives.
According to DuPont’s Rosanne Danner, of the most common reasons
organizations fail to develop a safety culture are:

1. Lack of commitment from leadership and management. A


safety culture has to start with the CEO setting the right vision of
where we want to be. That person needs to say, “This is how we
do work.” Safety must be part of measuring performance. It’s not
profitability or safety — it’s both. That commitment must extend
down to line management. If line managers see something that is
unsafe, and they don’t call it out, and it happens a second time,
then it becomes an acceptable way to do work. It becomes the
new standard. If something is viewed as not being important to
the manager, employees won’t pay attention to it.
2. Inconsistency in how and where safety is applied. Management
must put in place the right procedures and consistently follow
them. They may start all internal and external meetings with a
safety message or contact. This speaks to being constantly aware
of a person’s surroundings and thinking through actions that a
person would take in a variety of possible scenarios.
3. Loss of focus. Instituting a safety culture is not an overnight
proposition. If it is done correctly, we may see a change in injury
rate sooner, but it will take time to make it ingrained in an organ-
ization. We can’t let the quick results trick us into losing our focus
for the long term.
Current Trends and Practices 423

Implementing a safety policy for any organization should be a top pri-


ority. Employees should be encouraged to report any unsafe conditions
right away and should be trained how to react in an emergency. The pri-
mary goal of a workplace safety policy is to establish the expectation that
it is the responsibility of all employees to create and maintain a safe work
environment.

Arc Flash Hazards


One specific area of safety that should be considered by almost any
organization that maintains their own electrical equipment is Arc Flash
Safety. Bureau of Labor Statistics data reveal that between 1992 and 2002,
electrical accidents in the workplace caused 3,378 deaths and an addition-
al 46,598 non-fatal injuries. About 5% of all workplace deaths were relat-
ed to electrical equipment. These statistics were validated in a second
study involving more than 120,000 employees; this study determined that
arc flash injuries accounted for the largest category of all recorded electri-
cal injuries. Arc flash is responsible for a significant fraction of total elec-
trical deaths and injuries.
As defined by IEEE and the National Fire Protection Association
(NFPA), arc flash is a strong electric current — and sometimes a full-
blown explosion — that passes through air when insulation between ener-
gized conductors or between an energized conductor and ground is no
longer sufficient to contain the voltage between them. This creates a
“short cut” that allows electricity to race from conductor-to-conductor, to
the extreme detriment of any worker standing nearby. Arc flash resembles
a lightning bolt-like charge, emitting heat to reach temperatures of
o
35,000 F, which is hotter than the surface temperature of the sun, in
1/1000 of a second. Anyone exposed to the blast or heat without sufficient
personal protective equipment (PPE) would be severely, and oftentimes
fatally, injured.
An arc flash can cause substantial damage, fire, or injury. The massive
energy released in the fault instantly vaporizes the metal conductors
involved, blasting molten metal and expanding plasma outward with
extreme force. The result of the violent event can cause destruction of the
equipment involved, fire, and injury not only to the worker, but also to
people and equipment nearby.
Usually a fire produces roughly 50% convective heat (flame) and
50% radiant heat. An arc can be up to 90% radiant heat. This level can
produce severe burns when there is little or no flame present. In addition
to the explosive blast of such a fault, destruction also arises from the
424 Chapter 12

intense radiant heat produced by the arc. The metal plasma arc produces
tremendous amounts of light energy from far infrared to ultraviolet.
Surfaces of nearby people and objects absorb this energy and are instant-
ly heated to vaporizing temperatures. The effects of this can be seen on
adjacent walls and equipment, which are often ablated and eroded from
the radiant effects.

Causes and Damage of Arc Flash


Arcs can be initiated by a variety of causes, such as when

• Workers incorrectly think the equipment is de-energized and


begin to work on it energized.
• Workers drop or improperly use tools or components on an
energized system.
• Dust, water, or other contamination accumulate and cause
insulation breakdown.
• Connections loosen, overheat, and reach thermal runaway and
fail.

A hazardous arc flash can occur in any electrical device, regardless of


voltage, in which the energy is high enough to sustain an arc. Potential
places where this can happen include:

• Panel boards and switchboards


• Motor control centers
• Metal clad switch gear
• Transformers
• Motor starters and drive cabinets
• Fused disconnects
• Any place that can have an electrical equipment failure

Exposure to an arc flash frequently results in a variety of serious


injuries and, in some cases, death. Workers have been injured even though
they were ten feet or more away from the arc center. Worker injuries can
include damaged hearing, eyesight, and severe burns requiring years of
skin grafting and rehabilitation. The pressure waves can also propel loose
material like molten metal, pieces of damaged equipment, tools and other
objects, through the air.
Some of the employees at risk from arc flash hazards include mechan-
ics, electricians, and HVAC personnel. The most dangerous tasks include:
Current Trends and Practices 425

1. Removing or installing circuit breakers or fuses


2. Working on control circuits with energized parts exposed
3. Opening or closing circuit breakers or disconnects
4. Applying safety grounds
5. Removing panel covers

Because of the dangers of electrical explosions, OSHA now legally


requires employers to follow the NFPA recommended practices to protect
workers from arc flash exposure. OSHA’s 1910.132(d) and 1926.28(a)
state that the employer is responsible to assess the hazards in the work
place; select, have, and use the correct PPE; and document the assess-
ment. Although OSHA does not, per se, enforce the NFPA 70E standard,
OSHA considers the NFPA standard a recognized industry practice.
Electrical inspectors also are now enforcing the new labeling require-
ments set forth in the 2008 National Electric Code (NEC).
Compliance with OSHA involves adherence to a six-point plan:

1. A facility must provide, and be able to demonstrate, a safety pro-


gram with defined responsibilities.
2. Establish shock and flash protection boundaries.
3. Provide protective clothing (PC) and personal protective equip-
ment (PPE) that meet ANSI standards.
4. Training for workers on the hazards of arc flash.
5. Appropriate tools for safe working.
6. Warning labels on equipment. Note that the labels are provided by
the equipment owners, not the manufacturers. 110.16 NFPA 70E
2009 requires labels with the available incident energy or required
level of PPE.

Arc Flash Prevention


Preventive maintenance, worker training, and an effective safety pro-
gram can significantly reduce arc flash exposure. Preventive maintenance
should be conducted on a routine basis to ensure safe operation. As part
of a preventive maintenance program, equipment should be thoroughly
cleaned and routine inspections should be conducted by qualified person-
nel who understand how to uncover loose connections, overheated termi-
nals, discoloration of nearby insulation, and pitted contacts. A comprehen-
sive preventive maintenance plan should also include:
426 Chapter 12

1. Using corrosion-resistant terminals and insulating exposed metal


parts, if possible
2. Sealing all open areas of equipment to ensure rodents and birds
cannot enter
3. Verifying that all relays and breakers are set and operate
properly
4. Use of CBM technologies such as infrared and ultrasound where
possible

In order to select the proper PPE, incident energy must be known at


every point where workers may be required to perform work on energized
equipment. These calculations need to be performed by a qualified person
such as an electrical engineer. All parts of the body that may be exposed to
the arc flash need to be covered by the appropriate type and quality of PPE.
The best prevention method is to reduce arc flash hazards by design-
ing them out. Three key factors determine the intensity of an Arc Flash
event on personnel. These factors are the quantity of fault current avail-
able in a system, the fault until an arc flash is cleared, and the distance an
individual is from an arc. Various design and equipment configuration
choices can be made to affect these factors and, in turn, reduce the Arc
Flash hazard.

1. Fault current can be limited by using current limiting devices


such as grounding resistors or fuses. If the fault current is limited
to 5 amps or less, then many ground faults self-extinguish and do
not propagate into phase-to-phase faults.
2. Arcing time can be reduced by temporarily setting upstream pro-
tective devices to lower set points during maintenance periods or
by employing zone interlocking (ZSIP).
3. Distance can be mitigated through the use of remote operators or
robots to perform activities that are high risk for Arc Flash inci-
dents like racking breakers on a live electrical bus. The distance
from an arc flash source within which an unprotected person has
a 0% chance of receiving a second degree burn is referred to as
the “flash protection boundary.” Those conducting flash hazard
analyses must determine this boundary, and then must determine
what PPE should be worn within the flash protection boundary.
Current Trends and Practices 427

12.5 Risk Management

What Is Risk Management?

Risk is the potential that a chosen action or activity will lead to a loss,
an undesirable event or outcome. We all take risks in our everyday life.
When we do any work or activity at work, home or in our personal life,
such as driving to work, repairing a machine, engaging in a new venture
or assignment or project, we accept a certain level of risk. Unconsciously
in our mind, we evaluate the risk and its benefits and based on that infor-
mation, we do things because we believe the level of benefits outweigh
the risks. For example, we know that although driving can be dangerous,
it gets us to work or places we want to go in less time. Also, we know from
historical data that the probability of having an automobile accident is rea-
sonably low if we follow the road rules/regulations, such as wearing safe-
ty belts and staying within speed limits. However, sometimes we don’t
follow the rules, underestimate or misunderstand the risk involved, or
simply ignore it for many reasons, therefore resulting in undesirable out-
comes. Risks can also come from uncertainty in project failures (at any
phase in asset/system development, production or sustainment life-
cycles), legal liabilities, operational accidents, natural causes and disasters
as well as deliberate attack from an adversary or events of uncertain root
cause. Risk is officially defined as the combination of the probability of
an event and its consequences (ISO/IEC Guide 73). In all types of tasks
we undertake, there is the potential for events and consequences that con-
stitute opportunities for benefits or threats to success.
Questions arise about how we manage these uncertainties (risks).
Several risk management guidelines and standards have been developed
including those from the Project Management Institute (PMI), actuarial
societies, and ISO standards. Methods, definitions, and goals vary widely
according to whether the risk management method is in the context of
project management, security, engineering, industrial processes, financial
portfolios, actuarial assessments, or public health and safety. Risk
Management is increasingly recognized as a technique that considers
both positive and negative aspects of risk. In the safety field, risk is known
as a hazard and is generally recognized that consequences are only nega-
tive and therefore the management of safety risk is focused on prevention
and mitigation of these hazards
428 Chapter 12

ISO 31000 Standards


An organization may use strategies such as risk acceptance, risk
avoidance, risk retention, risk transfer, or any other strategy (or combina-
tion of strategies) in proper management of future events.
ISO 31000 is intended to be a family of standards relating to risk man-
agement issued by the International Organization for Standardization. The
ISO 31000 family includes the following:

• ISO 31000:2009 Principles and Guidelines on Implementation


• ISO 31010:2009 Risk Management — Risk Assessment
Techniques
• ISO / IEC Guide 73:2009 Risk Management — Vocabulary

ISO 31000:2009 is a new international standard which defines risk as


the effect of uncertainty on objectives, whether positive or negative. This
standard provides a generic framework for establishing the context of
identifying, analyzing, evaluating, treating, monitoring and communicat-
ing risk. ISO 31000:2009 offers guidelines for the design, implementation
and maintenance of risk management processes throughout an organiza-
tion. This approach to formalizing risk management practices will facili-
tate broader adoption by organizations who require an enterprise risk
management standard that accommodates multiple ‘silo-centric’ manage-
ment systems.

Purpose
The purpose of risk management is to prevent, reduce, or control
future impacts of unfavorable events as opposed to reacting to unwanted
events after they have already occurred. The mitigation of every plausible
risk may not be possible and is rather impractical due to resource limita-
tions. Hence, effective risk management requires a process to determine
which risks are actionable and can be mitigated, and which risks are non-
actionable or residual and cannot be mitigated. These risks must be con-
trolled instead (if identified early enough), watched, or transferred while
being accepted by the appropriate authority.

The Risk Management Process


Risk management is (should be) a central part of any organization’s
strategic management. It is the process whereby organizations methodi-
cally address the risks associated with their activities to enable the goal of
achieving sustained benefit within each activity and across all activities of
Current Trends and Practices 429

the organization. The focus of good risk management is the identification


and treatment of these risks. Its objective is to add maximum sustainable
value to all the activities of the organization. It creates the understanding
of the potential upside and downside of all those factors which can affect
the organization. It increases the probability of success, and reduces both
the probability of failure and the uncertainty of achieving the organiza-
tion’s overall objectives.
An effective risk management process should start from the ground
up with participation being encouraged at all levels. It is especially impor-
tant to encourage proactive participation by the subject matter experts and
stakeholders. Proper risk identification and mitigation should also have
the full support of the management chain for success. To ensure this, the
asset owner, manager or project manager should be held accountable for
proper risk handling and be held responsible for residual risks deemed
acceptable. Strong leadership across all relevant stakeholders is needed to
establish an environment for the free and open disclosure and discussion
of risk.
Risk management should be a continuously developing process which
runs throughout the organization’s strategy and the implementation of that
strategy. It should address methodically all the risks surrounding the orga-
nization’s activities past, present and in particular, future that could
endanger achievement of critical objectives. It must be integrated into the
culture of the organization with an effective policy.
Each risk can be classified as Low, Moderate, Serious, or High. The
risk classification is based on the likelihood of occurring and the conse-
quence it may have. Each risk should be assigned to an appropriate per-
son for adjudication. The designated risk owner is responsible for coordi-
nating the initial risk assessment to determine if the risk is actionable and
can be mitigated. A practical mitigation plan should be developed for each
actionable risk. Non-actionable risks cannot be mitigated and are classi-
fied as residual risks that must be accepted by the appropriate authority.
The goal is to identify all significant residual risks early enough to control
their likelihood of occurrence, if practical. A practical control plan should
be developed for each significant residual risk, with the appropriate risk
acceptance authority.

Risk Assessment
The fundamental difficulty in risk assessment is determining the rate
of occurrence because statistical information may not be available on all
categories of past incidents. Furthermore, evaluating the severity of the
430 Chapter 12

consequences (impact) is often quite difficult for intangible assets. Asset


valuation is another question that needs to be addressed. Thus, best edu-
cated opinions and available statistics are the primary sources of informa-
tion. Nevertheless, risk assessment should produce such information for
the management of the organization that the primary risks are easy to
understand and that the risk management decisions may be prioritized.
Thus, there have been several theories and attempts to quantify risks.
Numerous different risk formulae exist, but perhaps the most widely
accepted formula for risk quantification is:

Risk Index (Magnitude) = Consequence (Impact) of Risk event x


Likelihood (Probability) of Occurrence

Figure 12.1 is an example of a risk matrix that allows us to rate poten-


tial risks on two dimensions – likelihood (probability) and consequence
(impact).
The impact of the risk event is commonly assessed on a scale of 1 to
5, where 1 and 5 represent the minimum and maximum possible impact







  
  



 
  



%$#"! !#%!"

Figure 12.1 Example of a Risk Matrix


Current Trends and Practices 431

of an occurrence of a risk (usually in terms of financial losses). However,


the 1 to 5 scale can be arbitrary and need not be on a linear scale.
The probability of occurrence is likewise commonly assessed on a
scale from 1 to 5, where 1 represents a very low probability of the risk
event actually occurring whereas 5 represents a very high probability of
occurrence. This axis may be expressed in either mathematical terms
(event occurs once a year, once in ten years, once in 100 years etc.) or may
be expressed in “plain English” (event occurs here very often, event has
been known to occur here, event has been known to occur in the industry,
etc.). Again, the 1-to-5 scale can be arbitrary or non-linear depending on
decisions by subject-matter experts
There are many different types of risk matrixes in use across industry,
ranging from 3x3 to 6x6 and larger. The 5x5 matrix is generally consid-
ered the standard risk matrix, but other risk matrixes are widely used, such
as, the MIL-STD-882 compliant 4x5 system safety risk matrix.
The risk index can thus take values ranging (typically) from 1 through
25, and this range is usually arbitrarily divided into three sub-ranges. The
overall risk assessment is then Low, Medium or High, depending on the
sub-range containing the calculated value of the risk index. For instance,
the three sub-ranges could be defined as 1 to 8, 9 to 16 and 17 to 25.
Follow these steps to develop a risk impact and probability
table/chart:

1. List all of the likely risks that the asset/system or project faces.
Make the list as comprehensive as possible.
2. Assess the probability of each risk occurring, and assign it a rat-
ing. For example, use a scale of 1 to 5. Assign a score of 1 when
a risk is extremely unlikely to occur, and use a score of 5 when
the risk is extremely likely to occur.
3. Estimate the impact on the asset/system or project if the risk
occurs. Again, do this for each and every risk on the list. Using
a 1–5 scale, assign it a 1 for little impact and a 5 for a huge, cat-
astrophic impact.
4. Map out the ratings on the Risk Impact/Probability Chart.
5. Develop a response to each risk, according to its position in the
chart. Remember, risks in the bottom left corner can often be
ignored whereas those in the top right corner need a great deal
of time and attention.
432 Chapter 12

Risk Mitigation
Once risks have been identified and assessed, all techniques to man-
age the risk fall into one or more of these four major categories known
as ACAT

• Avoidance (eliminate, or not do that activity)


• Control (optimize, mitigate, or reduce risk)
• Accept (accept and budget-plan)
• Transfer (risk share or outsource)

Risk avoidance includes not performing an activity that could carry


risk. An example is not to drive the car in order not to take the risk that
the car may be involved in accident. Avoidance may seem the answer to
all risks, but avoiding risks also means losing out on the potential gain that
accepting the risk may have allowed.
Risk control, reduction, or optimization involves reducing the severi-
ty of the loss or the likelihood of the loss occurring. For example, sprin-
klers are designed to put out a fire to reduce the risk of loss by fire. This
method may cause a greater loss by water damage and, therefore, may not
be suitable. Halon fire suppression systems may mitigate that risk, but the
cost may be prohibitive.
Risk acceptance means accepting the risk and developing a plan or
procedure for what to do if it happens. An example of this may be a plant,
organization, or city’s emergency response plan for an earthquake.
Risk transfer is often used in place of risk sharing in the mistaken
belief that we can transfer risk to a third party through insurance or out-
sourcing. In practice if the insurance company or contractor goes bankrupt
or ends up in court, the original risk is likely to still revert to the first party.
As such in the terminology of practitioners and scholars alike, the pur-
chase of an insurance contract is often described as a “transfer of risk.”
Acknowledging that risks can be positive or negative, optimizing
risks means finding a balance between negative risk and the benefit of the
operation or activity and between risk reduction and effort applied.
Operational Risk, another frequently used term, is defined as the
probability of loss occurring from the internal inadequacies of an organi-
zation’s breakdown in its controls, operations, or procedures. The four
principles of Operational Risk Management (ORM) are:

• Accept risk when benefits outweigh the cost.


• Accept no unnecessary risk.
Current Trends and Practices 433

• Anticipate and manage risk by planning.


• Make risk decisions at the right level.

Creating a Risk Management Plan


Select appropriate controls or countermeasures to measure each risk.
Risk mitigation needs to be approved by the appropriate level of manage-
ment. For instance, a risk concerning the image of the organization should
have top management decision behind it whereas IT management would
have the authority to decide on computer virus risks.
The risk management plan should propose applicable and effective
security controls for managing the risks. For example, an observed high
risk of computer viruses could be mitigated by acquiring and implement-
ing antivirus software. A good risk management plan should contain a
schedule for control implementation and responsible persons for those
actions.
A risk management plan should answer the following to ensure there
is an established process to:

• Determine risk sources and categories.


• Define the parameters used to analyze and categorize risks.
• Establish and maintain the strategy to be used for risk
management.
• Identify and document the risks.
• Determine its relative priority of each identified risk.
• Develop a risk mitigation plan for the most important risks.
• Monitor the status of each risk periodically and update the risk
mitigation plan as appropriate.
• Ensure adequate resources and training plan are in place for risk
management.

Risks associated with projects can be classified into six spheres or


attributes of the processes and end products of a project.
1. Safety risk
An expression of the possibility/impact of a mishap that can cause
death, injury, occupational illness, or damage to or loss of equipment
or property, or damage to the environment, in terms of hazard severi-
ty categories and hazard probability levels

2. Performance risk
The degree to which the proposed system or process design is capa-
434 Chapter 12

ble of meeting the operational requirements, which include reliability,


maintainability, dependability, availability, and testability require-
ments

3. Cost risk
The ability of the system to achieve the program’s life-cycle cost
objectives, including the effects of budget and affordability deci-
sions and the effects of inherent errors in the cost estimating tech-
nique(s) used, given that the system requirements were properly
defined

4. Schedule risk
The adequacy of the time allocated for performing the defined tasks,
including development, production, and testing as well as the effects
of programmatic schedule decisions, the inherent errors in the
schedule estimating technique used, and external physical con-
straints
5. Technology risk
Degree to which the technology proposed for the system has been
demonstrated as capable of meeting all of the project’s objectives

6. Product data access and protection risk


A new area of risk, especially with the industrial espionage and
reverse engineering practices of some companies and even govern-
ments, that provides for protection of proprietary data regarding
processes and products

All of these risks need to be addressed when dealing with specific


projects.

12.6 Corrosion Control

What Is Corrosion?
One specific risk that has become better known is in the area of
Corrosion Control.
Corrosion is a naturally occurring phenomenon commonly defined as
the deterioration of a substance, usually a metal, or its properties because
of a reaction with its environment. In other words, corrosion is the wear-
ing-away of metals due to a chemical reaction.
A better scientific definition of Corrosion is the disintegration of an
Current Trends and Practices 435

engineered material into its constituent atoms due to chemical reactions


with its surroundings. It means electrochemical oxidation of metals in
reaction with an oxidant such as oxygen. Formation of an oxide of iron
due to oxidation of the iron atoms in solid solution is a well-known exam-
ple of electrochemical corrosion, commonly known as rusting. This type
of damage typically produces oxide(s) and/or salt(s) of the original metal.
Corrosion can also refer to other materials than metals, such as ceramics
or polymers, although in this context, the term degradation is more com-
mon.
Corrosion can cause dangerous and expensive damage to everything
from automobiles, home appliances, water supply system pipelines,
bridges, buildings, and industrial plant infrastructure. A two-year long
study, entitled “Corrosion Costs and Prevention Strategies in the United
States” released by the U.S. Federal Highway Administration (FHWA) in
2002 showed that the total annual estimated direct cost of corrosion in the
United States is about $276 billion (approximately 3.1 % of the nation’s
Gross Domestic Product or GDP). It revealed that, although corrosion
management has improved over the past several decades, the United
States and other countries must find more and better ways to implement
optimal corrosion control practices. Other studies performed recently in
China, Japan, the United Kingdom, and Venezuela showed similar to even
more costly results, leading to an estimated worldwide direct cost exceed-
ing $1.8 trillion, which translates to 3 to 4% of the GDP of industrialized
countries.
Corrosion is so prevalent and takes so many forms that its occurrence
and associated costs never will be completely eliminated. However, all
studies estimate that 25 to 30% of annual corrosion costs could be saved
if optimum corrosion management practices were employed.
Corrosion prevention and control can be achieved by incorporation of
the latest state-of-the-art corrosion control technology in the original
equipment design, in the manufacturing, in all levels of maintenance, sup-
ply, and storage processes. The objective is to minimize corrosion by
using design and manufacturing practices that address selection of mate-
rials; coatings and surface treatments; production processes; process spec-
ifications; system geometry; material limitations; environmental
extremes; storage and ready conditions; preservation and packaging
requirements; and repair, overhaul, and spare parts requirements.

Corrosion Control and Protection


The following are four basic methods for Corrosion Control and
436 Chapter 12

Protection:

1. Materials resistant to Corrosion


2. Protective coatings
3. Cathodic protection
4. Corrosion Inhibitors: modify the operating environment

In most cases, effective corrosion control is obtained by combining


two or more of these methods. Corrosion control should be considered at
the design stage of a given facility or system. The methods selected must
be appropriate for the materials used, for the configurations, and for the
types and forms of corrosion which must be controlled.
There are no materials that are immune to corrosion in all environ-
ments. Materials must be matched to the environment that they will
encounter in service. Corrosion Resistance Data are used to assess the
suitability of a material in an environment.
Protective coatings are the most widely used corrosion control tech-
nique. Essentially, protective coatings are a means for separating the sur-
faces that are susceptible to corrosion from the factors in the environment
which cause corrosion to occur. Remember, however, that protective coat-
ings can never provide 100 percent protection of 100 percent of the sur-
face. If localized corrosion at a coating defect is likely to cause rapid cat-
astrophic failure, additional corrosion control measures must be taken.
Coatings are particularly useful when used in combination with other
methods of corrosion control such as cathodic protection or galvanic cor-
rosion
Cathodic protection interferes with the natural action of the electro-
chemical cells that are responsible for corrosion. Cathodic protection can
be effectively applied to control corrosion of surfaces that are immersed
in water or exposed to soil. Cathodic protection in its classical form can-
not be used to protect surfaces exposed to the atmosphere. The use of
anodic metallic coatings such as zinc on steel (galvanizing) is, however, a
form of cathodic protection, which is effective in the atmosphere. There
are two basic methods of supplying the electrical currents required to
interfere with the electrochemical cell action.
The first method, cathodic protection with galvanic anodes, uses the
corrosion of an active metal, such as magnesium or zinc, to provide the
required electrical current. In this method, called sacrificial or galvanic
anode cathodic protection, the active metal is consumed in the process of
protecting the surfaces where corrosion is controlled and the anodes must
be periodically replaced. In the second method, impressed current cathod-
Current Trends and Practices 437

ic protection, an alternative source of direct electrical current, usually a


rectifier that converts alternating current to direct current is used to pro-
vide the required electrical current. In this system, the electrical circuit is
completed through an inert anode material that is not consumed in the
process.
The best application of Cathodic Protection is usually to protect stor-
age tanks, particularly underground and piping systems. Based on the type
of fluid contained in the tanks to be protected, it is possible to use galvan-
ic anodes or impressed current. Usually the surfaces to be protected with
cathodic protection are also coated to reduce the current requirement and
increase the life of the galvanic anodes. The anodes used in cathodic pro-
tection for tanks must be periodically inspected and replaced when con-
sumed.
Another method of corrosion control often neglected is modifying the
operating environment. Using a selective backfill around a buried struc-
ture, using corrosion inhibitors in power plant or in engine cooling sys-
tems, and modifying structures to provide adequate drainage are all exam-
ples of the use of this method of corrosion control. Although best
employed during the design stage, in some cases, actions taken to correct
corrosion problems through modifying the environment can be taken after
a system is built. Careful identification and characterization of corrosion
problems will often reveal opportunities for changing the environment to
control corrosion.

12.7 Systems Engineering and Configuration Management

Imagine that a craft person we have sent to repair an asset finds out
that the new spare won’t fit or the new motor has a different footprint
(frame size) from what’s documented in the CMMS system. Suppose we
have ordered a special purpose machine and, upon installation, it does not
do what it is supposed to do. In both cases, the system requirements or
configurations were not documented properly or they were misinterpret-
ed. Has this happened in your plant? If we had followed systems engineer-
ing and configuration management practices appropriately, we would
have minimized such issues.
Systems Engineering (SE) is an interdisciplinary engineering man-
agement process that evolves and verifies an integrated, life-cycle bal-
anced set of system solutions that satisfy customer needs. Systems engi-
neering management is accomplished by integrating three major cate-
gories:
438 Chapter 12

• A product development phase that controls the design process and


provides baselines that coordinate design effort
• A systems engineering process that provides a structure for solv-
ing design problems and tracking requirements flow through
design process
• Life cycle integration that involves customers in the design,
building, and installation (including commissioning) process, and
ensures that the product developed is viable throughout its life.

Configuration management (CM), a component of SE, is a critical


discipline in delivering products that meet customer requirements and that
are built according to approved design documentation. In addition, it
tracks and keeps updated system documentation which includes drawings,
manuals, operations/maintenance procedures, training, etc.
CM is the methodology of effectively managing the life cycle of
assets and products in the plant. It prohibits any change of the asset’s
form, fit, and function without a thorough, logical process that considers
the impact proposed changes have on life cycle cost.

Systems Engineering (SE)


The term systems engineering can be traced back to Bell Telephone
Laboratories in the 1940s. The need to identify and manipulate the prop-
erties of a system as a whole — which in complex engineering projects
may greatly differ from the sum of the parts’ properties — motivated the
Department of Defense, NASA, and other industries to apply this disci-
pline.
The purpose of SE is to provide a structured but flexible process that
transforms requirements into specifications, architecture, and configura-
tion baselines. The discipline of this process provides the control and
traceability to develop solutions that meet customer needs. Life cycle inte-
gration, a key component of SE process, is achieved through integrated
development — that is, concurrent consideration of all life cycle needs
during the development process.
The key primary functions of systems engineering support the system
(product):
1. Design/Development. This function includes the activities
required to evolve the system from customer needs to product
/process solutions.
2. Build/Construction/Manufacture. This function includes the
fabrication and construction of unique systems and subsystems.
Current Trends and Practices 439

3. Deployment/Fielding/Commissioning. These activities are nec-


essary to deliver, install, check out, train, operate, and field the
system to achieve full operational capabilities.
4. Operations. The user (system owner) operates systems safely as
designed (not to be abused).
5. Support/Maintenance. This area includes activities necessary
to provide operational support including maintenance, logistics,
and materials management.
6. Disposal. It considers the activities necessary to ensure that
once the asset or system has completed its useful life, it is
removed in a way that meets all applicable regulations.
7. Training. These are the activities necessary to achieve and
maintain the knowledge and skill levels for efficiently and effec-
tively performing operations, maintenance, and other support
functions.
8. Verifications. These activities are necessary for evaluating the
progress and effectiveness of the system’s processes, and to
measure specification (requirements) compliance.

Systems engineering is a standardized, disciplined management


process for developing system solutions. It provides a methodical
approach to system development in an environment of change and uncer-
tainty. It ensures that the correct technical tasks get done during the devel-
opment through planning, tracking, and coordinating. Systems engineer-
ing covers the “cradle to grave” life cycle process.

Configuration Management (CM)


Configuration management is a field of management that focuses on
establishing and maintaining the consistency of product performance, and
its functional and physical attributes with its requirements, design, and
operational information throughout its life.
Configuration management was first developed by the U.S.
Department of Defense in the 1950s as a technical management disci-
pline. The concepts have been widely adopted by numerous technical
management models, including systems engineering, integrated logistics
support, Capability Maturity Model Integration (CMMI), ISO 9000, proj-
ect management methodology, and product lifecycle management.
Configuration management is used to maintain an understanding of
the status of complex assets with a view to maintaining the highest level
of serviceability for the lowest cost. Complex assets such as automobiles,
440 Chapter 12

aircraft, and major capital equipment sometimes consist of hundreds to


thousands of parts. In addition, there are related tooling, fixtures, gauges,
templates, test equipment, and control software. It is estimated that a part
may undergo ten engineering changes or more over its life. This suggests
that an organization may evaluate and process many hundreds to thou-
sands of engineering changes for a complex system. It takes a significant
amount of effort to keep baseline and documentation current.
Over the life cycle of systems, the manufacturer, supplier, and owner
must assure that the as-designed configuration at any point in time will
satisfy functional requirements and that the hardware and software actu-
ally delivered (as-built configuration) corresponds to the approved as-
designed configuration. The configuration management emphasis should
be continued during O&M phase to ensure all documentation is current.
As a result, the configuration management effort required for a complex
system is significant. Usually, computerized systems such as
CMMS/EAM/ERP may be required to support configuration management
if an organization is to avoid being drowned in a sea of paper and non-
value-added administration.
An organization’s configuration management program includes an
evaluation process which identities, examines, and selects assets/systems,
computer software, and documents that will be part of the CM program.
The evaluation process also provides for periodic assessment of the pro-
gram elements throughout the lifetime of the program.
Typical examples of documents included in a configuration manage-
ment program include:

1. System descriptions
2. Drawings
3. Special studies and reports including safety inspections or
investigations
4. Operations and maintenance procedures, guidelines, and
acceptance criteria
5. Instrument and control set points
6. Quality assurance and quality control documents
7. Vendor/suppliers manuals
8. Regulatory requirements, codes, and standards
9. Modification including capital project packages
10. Component and part lists
11. Specifications and purchase orders information for major and
critical assets
Current Trends and Practices 441

12. Asset/system performance and maintenance records


13. Welding qualification records
14. Pressure vessels/systems integrity and inspection records
15. Design criteria/requirements
16. Operations/maintenance training records

Documents and records should be continuously updated to include all


approved changes and should be accurately reflected in output documents
such as drawings, system descriptions, specifications, and procedures.
A cautionary note: The CM program will cost significant resources to
implement. Therefore, organizations must evaluate what documents and
records should be part of the CM program based on the asset’s complexi-
ty, criticality, and a value-added assessment of CM.
The CM program also helps to be in compliance with ISO 9000, ISO
31000, and 55000, etc., requirements. This will ensure that the organiza-
tion remains in the appropriate configuration with the right and current
documentation during its operating life cycle. Configuration management
affects the entire organization. Every plant and facility must have an
effective configuration management program to eliminate or mitigate the
negative impact of uncontrolled, undocumented changes in the configura-
tion of its assets. Enforcing a logical, disciplined process to evaluate,
design, procure, implement, operate, and maintain modifications to major
and critical assets eliminates most of the excessive maintenance cost
caused by poor records.

12.8 Standards and Standardization

• The inch or millimeter is a standard of measurement.


• Words are standards of communication.
• Traffic lights are safety standards.
• Octane numbers of gasoline are quality standards.
• “No more than 1% shrinkage” is a performance standard.

These are just few examples of standardization – application of stan-


dards. Standardization has a major impact on our lives, yet most of us
know little about the process or about the standards themselves. Do you
remember using camera film marked ISO 100, 200 etc.? We know that
camera film marked as ISO 200 is likely to give good results in a camera
with the film speed set at 200. But few understand that the ISO 200 mark-
ing on the package means that the film conforms to a standard established
442 Chapter 12

by the International Organization for Standardization (ISO), an interna-


tional organization that writes standards.

What Is a Standard?
A standard is defined by the National Standards Policy Advisory
Committee as:

“A prescribed set of rules, conditions, or requirements concerning


definitions of terms; classification of components; specification of
materials, performance, or operations; delineation of procedures; or
measurement of quantity and quality in describing materials,
products, systems, services, or practices.”

In layman’s terms, a standard is a rule or requirement that is deter-


mined by a consensus opinion of users and that prescribes the accepted
and (theoretically) the best criteria for a product, process, test, or proce-
dure. The general benefits of a standard are safety, quality, and reliability,
interchangeability of parts or systems, and consistency across internation-
al borders.

History of Standards
Standards are known to have existed as early as 7000 B.C. when
cylindrical stones were used as units of weight in Egypt. One of the first
known attempts at standardization in the Western world occurred in 1120.
King Henry I of England ordered that the ell, the ancient yard, should be
the exact length of his forearm, and that it should be used as the standard
unit of length in his kingdom.
History also notes that, in 1689, the Boston city fathers recognized the
need for standardization when they passed a law making it a civic crime
to manufacture bricks in any size other than 9x4x4. The city had just been
destroyed by fire, and the city fathers decided that standards would assure
rebuilding in the most economic and fastest way possible.
With the advent of the Industrial Revolution in the 19th century, the
increased demand to transport goods from place to place led to advanced
modes of transportation. The invention of the Railroad was a fast, eco-
nomical and effective means of sending products cross-country. This feat
was made possible by the standardization of the railroad gauge, which
established the uniform distance between two rails on a track. Imagine the
chaos and wasted time if a train starting out in New York had to be
unloaded in St. Louis because the railroad tracks did not line up with the
Current Trends and Practices 443

train’s wheels. Early train travel in America was hampered by this phe-
nomenon. The government worked with the railroads to promote use of
the most common railroad gauge in the United States at the time, which
measured 4 feet, 8 1/2 inches, a track size that originated in England. This
gauge was mandated for use in the Transcontinental Railroad in 1864 and
by 1886 had become the U.S. standard.
In 1904, a fire broke out in the basement of the John E. Hurst &
Company Building in Baltimore. After taking hold of the entire structure,
it leaped from building to building until it engulfed an 80-block area of
the city. To help combat the flames, reinforcements from New York,
Philadelphia, and Washington, D.C. immediately responded—but to no
avail. Their fire hoses could not connect to the fire hydrants in Baltimore
because they did not fit the hydrants in Baltimore. Forced to watch help-
lessly as the flames spread, the fire destroyed approximately 2,500 build-
ings and burned for more than 30 hours.
It was evident that a new national standard had to be developed to pre-
vent a similar occurrence in the future. Up until that time, each municipal-
ity had its own unique set of standards for firefighting equipment. As a
result, research was conducted regarding over 600 fire hose couplings
from around the country and one year later a national standard was creat-
ed to ensure uniform fire safety equipment and the safety of Americans
nationwide.
This was the beginning of standardization and standards development
in the 20th century to support interchangeability of parts, components, and
safety. In 1918, the American National Institute of Standards (ANSI), a
not-for-profit organization, was founded by support of several profession-
al societies, such as ASME, IEEE, ASCE, ASTM, etc., to support devel-
opment of standards.

How are standards developed today?


Most standards are developed by committees of volunteers, which can
include members of industry, government, and the public. In the United
States, the American National Standards Institute (ANSI) acts as a “par-
ent” organization, helping to coordinate volunteers and ensure that the
development process emphasizes four main issues: requirements for due
process, appeals procedures, the mandatory consideration of negative
votes or comments, and support for “committee balance.” Balance is
achieved when all parties having an interest in the outcome of a standard
have an opportunity to participate and where no single interest can domi-
nate the outcome.
444 Chapter 12

In the United States alone, approximately 30,000 current voluntary


standards have been developed by more than 400 organizations. Many of
these organizations, known as Standards Development Organizations
(SDO), are professional societies or not-for-profit organizations such as
ANSI, ASME, ASTM, IEEE, UL, etc. These do not include a much
greater number of procurement specifications (developed and used by
federal, state, and local procurement authorities), as well as mandatory
codes, rules, and regulations containing standards developed and adopted
at federal, state, and local levels. In addition, numerous foreign national,
regional, and international organizations produce standards of interest and
importance to U.S. manufacturers and exporters.

Benefits and Types of Standards


We use standards to achieve a level of safety, quality, and consisten-
cy in the products and processes that affect our lives. In short, standards
make our lives safer, easier, and better. Standards are also vital tools of
industry and commerce. They often provide the basis for buyer-seller
transactions; hence, they have tremendous impact on organizations and
nations, and even on the economic fabric of the world market.
Standards are a powerful tool for organizations of all sizes, support-
ing innovation and increasing productivity. Effective standardization pro-
motes forceful competition and enhances profitability, enabling a business
to take a leading role in shaping the industry itself. Standards allow organ-
izations to:

• Implement and maintain best practices


• Support safety of people and environment
• Improve productivity — reduce cost
• Attract and assure customers
• Demonstrate market leadership
• Create competitive advantage

Standards can be classified in two categories:


• Specifications — codes
• Process improvement — management

Specifications/codes help to standardize parts — components for


interchangeability and safety of products. Process improvement is related
to management of processes. We all are familiar with ISO 9000, which is
Current Trends and Practices 445

a quality management standard. It could be used for managing any process.


It was developed back in the 1970s; since then, it has gone through many
iterations and has been widely accepted as the management standard world-
wide. Currently, the 9000 series of management standards consist of:

• ISO 9000:2005 Quality management systems — Fundamentals


and vocabulary
• ISO 9001:2008 Quality management systems — Requirements
• ISO 9004:2009 Managing for the sustained success of an
organization — A quality management approach,

The 9001:2008 is the key standard which contains the requirements. This
standard contains the following key sections:

• Section 1: Scope
• Section 2: Normative Reference
• Section 3: Terms and definitions (specific to ISO 9001, not
specified in ISO 9000)
• Section 4: Quality Management System
• Section 5: Management Responsibility
• Section 6: Resource Management
• Section 7: Product Realization
• Section 8: Measurement, analysis and improvement

In effect, users need to address all Sections 1 through 8, but only Sections
4 through 8 need implementing within a quality management system.
Although ISO 9001 is known as the Quality Management System stan-
dard, but it could be applied, with some tailoring, to any process such as
logistics–supply chain, design, and asset management. Some organizations
such as Aerospace Testing Alliance (ATA) at Arnold Engineering
Development (Test) Center (AEDC) and Jacobs have implemented ISO
9001 to all of their work processes successfully including asset manage-
ment.
However, many experts in maintenance and asset management area
globally believe there is a gap in the area of asset management standards.
An international effort is underway to support and develop an international
standard for asset management known as ISO 55000. This family of stan-
dards has the following three standards:
446 Chapter 12

ISO 55000: Asset management — Overview, principles and terminology


ISO 55001: Asset management — Management systems — Requirements
ISO 55002: Asset management — Management systems — Guidelines
for the application of ISO 55001

The overall purpose of these three International Standards is to provide


a cohesive set of information in the field of Asset Management Systems
that will:

1. Enable users of the standards to understand the benefits, key con-


cepts, and principles of asset, asset management, and asset manage-
ment systems.
2. Harmonize the terminology being used in this field.
3. Enable users to know and understand the minimum requirements of
an effective management system to manage their assets.
4. Provide a means for such management systems to be assessed
(either by the users themselves, or by external parties).
5. Provide guidance on how to implement the minimum
requirements.

The specific scopes for each standard are:

• ISO 55000 is to provide an overview of the field of asset manage-


ment, including a description and explanation of the importance of
key concepts and principles relating to asset management and asset
management systems, as well as defining the terminology needed in
this discipline.
• ISO 55001 is to define “requirements” for the development, mainte-
nance, and improvement of a management system for the manage-
ment of an organization’s assets to achieve its stated strategic objec-
tives
• ISO 55002 is to provide guidelines for the application of the require-
ments specified in ISO 55001. It will discuss and describe contextu-
al situations and differences that can affect the use of asset manage-
ment principles and requirements in general, and provide guidance
on the establishment, implementation, maintenance, and improve-
ment of a management system for asset management, and its coordi-
nation with other management systems.
Current Trends and Practices 447

Currently, the following standards related to asset management and


plant/facilities are available; they can be implemented to improve the
overall performance:

• ISO 9000:2008 Quality Management


• PAS 55 Asset Management (Guidelines /
specifications)
• ISO 55000x Asset Management (Standards
Under Development)
• ISO 50001:2011 Energy Management
• ISO 31000:2009 Environment Management
• ISO 18000:2009 Risk Management

An asset management specific standard will aid maintenance and reli-


ability organizations in establishing standards for their processes and
become a leader in the maintenance and reliability industry as well.

12.9 Summary

Being aware of current trends and innovative best practices allow


businesses to continually improve processes to remain competitive. A few
of these trends and innovative practices have been discussed: energy and
sustainability, safety (including arc flash safety), risk management, corro-
sion control, systems engineering/configuration management, and stan-
dards.
Sustainable development is defined as “forms of progress that meet
the needs of the present without compromising the ability of future gen-
erations to meet their needs” and refers to three broad themes, also called
“pillars”: economic, social, and environmental. Sustainability also car-
ries with it a reduction in energy usage by a company. Because energy
costs can have a significant impact on the financial performance of busi-
nesses, energy reduction measures should be taken by companies as the
prices of energy continue to rise; you should also investigate green initia-
tives and their applications to your company or industry.
Several reliability and safety experts have observed that reliable
plants are safe plants and safe plants are reliable plants, and that the com-
bination makes for profitable plants. Therefore, any company trying to
achieve maintenance and reliability excellence should consider the impor-
tance of creating a safety-minded culture throughout their organization,
beginning with a fostering of a safety attitude amongst its entire leader-
448 Chapter 12

ship team. One specific area of safety worth mentioning is that of arc
flash, which has the ability to negatively impact a facility and endanger its
personnel, to which prevention measures should be evaluated and imple-
mented as necessary.
Risk is defined as the potential that a chosen action or activity will
lead to a loss, an undesirable event or outcome, of which we all take risks
in our everyday life, whether at work or in our personal lives. Risk
Management is increasingly recognized as a technique that considers both
positive and negative aspects of risk and should be a continuously devel-
oping process which runs throughout the organization’s strategy and the
implementation of that strategy, integrated into the culture of the organi-
zation with an effective policy. Risk management approaches can be
divided into four major categories: Avoidance (eliminate, or not do that
activity), Control (optimize, mitigate, or reduce risk), Accept (accept and
budget/plan), and Transfer (risk share or outsource). A risk management
plan should be developed to propose applicable and effective security
controls for managing the risks.
Corrosion is a naturally occurring phenomenon commonly defined as
the deterioration of a substance, usually a metal, or its properties because
of a reaction with its environment. Corrosion can cause dangerous and
expensive damage to everything and is so prevalent that it takes many
forms will never be completely eliminated. However, studies estimate that
25 to 30% of annual corrosion costs could be saved if optimum corrosion
management practices were employed. There are four basic methods for
corrosion control and protection: materials resistant to corrosion, protec-
tive coatings, cathodic protection, and corrosion inhibitors to modify the
operating environment. In most cases, effective corrosion control is
obtained by combining two or more of these methods and should also be
considered at the design stage of a given facility or system. The methods
selected must be appropriate for the materials used, for the configurations,
and for the types and forms of corrosion which must be controlled.
Systems Engineering and Configuration Management are techniques
that should be considered not only for the products that are manufactured
but also for the assets/systems maintained by an organization. Systems
Engineering (SE) is an interdisciplinary engineering management process
that evolves and verifies an integrated, life-cycle balanced set of system
solutions that satisfy customer needs. Configuration management (CM), a
component of SE, is a critical discipline in delivering products that meet
customer requirements and that are built according to approved design
documentation. It addition, it tracks and keeps updated all appropriate sys-
Current Trends and Practices 449

tem documentation.
Standards are rules or requirements that are determined by a consen-
sus opinion of users and that prescribe the accepted and (theoretically) the
best criteria for a product, process, test, or procedure. The general bene-
fits of a standard are safety, quality, reliability, interchangeability of parts
or systems, and consistency across international borders. Most standards
are developed by committees of volunteers, which can include members
of industry, government, and the public. Effective standardization pro-
motes forceful competition and enhances profitability, enabling a business
to take a leading role in shaping the industry itself. An asset management
specific standard will aid maintenance and reliability organizations in
establishing standards for their processes and become a leader in this
industry as well.
Further analysis of how each of these applies to your specific business
or industry is necessary to understand how to apply these trends and prac-
tices, if your company is engaged in activities to which each specifically
applies.

12.10 Self Assessment Questions

Q12.1 Define sustainability. Why is it important to organizations?

Q12.2 What process improvement strategies can be used to reduce


plant energy consumption?

Q12.3 What four major categories of equipment/systems use the


majority of energy in the industry, as defined by DOE?

Q12.4 Generally, the electricity bill is broken down by what types of


charges? What can be done to minimize the total electric ener-
gy cost?

Q12.5 Define the major categories of risk to which a product (asset)


or project may be exposed.

Q12.6 Why is configuration management important? Discuss its


application in the maintenance – asset management area.
Q12.7 What strategies are used to reduce the impact of arc flash
hazards?
450 Chapter 12

Q12.8 Why do we use standards? How they can be classified?

Q12.9 What is the intent of the ISO 55000 family of standards?

12.11 References and Suggested Reading

U.S. Department of Energy. 20 Ways to Save Energy Now.


www.eere.energy.gov/consumer/industry/20ways.html
www.energy.gov
Seminar / Paper by Ron Moore at MARCON 2011 and Reliability
Maintenance Center /UTK meetings 2011
Blanchard, Benjamin S. Systems Engineering Management. Prentice-
Hall, 1997.
Bureau of Labor Statistics data at www.bls.gov
NFPA website www.nfpa.org
OSHA website www.osha.gov
MIL-STD-882
ANSI website www.ansi.org
ISO website www.iso.org
ISO/US TAG -PC 251 committee website: www.uspc251tag.org
www.iso55000.info
www.wikipedia.com
Appendix for Chapter 1

Best Practices Q–A


Answer Key and Explanations

Q. 1 a Q. 31 b
Q. 2 b Q. 32 a
Q. 3 a Q. 33 b
Q. 4 a Q. 34 a
Q. 5 a Q. 35 b
Q. 6 a Q. 36 a
Q. 7 a Q. 37 a
Q. 8 a Q. 38 b
Q. 9 a Q. 39 b
Q. 10 a Q. 40 a
Q. 11 a Q. 41 b
Q. 12 a Q. 42 b
Q. 13 a Q. 43 a
Q. 14 a Q. 44 a
Q. 15 a Q. 45 a
Q. 16 a Q. 46 b
Q. 17 a Q. 47 b
Q. 18 b Q. 48 c
Q. 19 a Q. 49 c
Q. 20 b Q. 50 b
Q. 21 b Q. 51 b
Q. 22 a Q. 52 d
Q. 23 a Q. 53 b
Q. 24 b Q. 54 c
Q. 25 b Q. 55 a
Q. 26 c
Q. 27 b
Q. 28 a
Q. 29 a
Q. 30 a

451
452 Appendix

1. Best Practices are practices that are defined and applied by an


organization to improve their operation. These practices may or
may not be proven, but results are found to be acceptable.
a. True
b. False

Answer: a — True
A best practice is a business function, a practice, or a process, that is
considered superior to all other known methods. It’s a documented strate-
gy and approach used by the most respected, competitive, and profitable
organizations. A best practice when implemented appropriately should
improve performance and efficiency in a specific area. See more details in
Chapter 1.

2. Maintainability is measured by PM schedule compliance.


a. True
b. False

Answer: b — False
Maintainability is defined as ease of maintenance; it’s primarily
measured by Mean Time to Repair (MTTR). See more details in Chapter
6.

3. All maintenance personnel’s time should be covered by


work orders.
a. True
b. False

Answer: a — True
All maintenance personnel’s time should be counted and document-
ed in CMMS to ensure all repair and maintenance costs are accurate.
See more details in Chapters 3 and 4.

4. Operations and Maintenance must work as a team to


achieve improved OEE.
a. True
b. False

Answer: a — True
OEE is calculated as Availability X Performance X Quality.
Operations and Maintenance both impact this metric and need to work
together as a team to achieve higher OEE. See more details in Chapter 7.
Appendix 453

5. Best practices would indicate that 90% or more of all


maintenance work is planned.
a. True
b. False

Answer: a — True
It’s good practice to plan 90% or more work. Planned work costs
2–3 times less than reactive work. See more details in Chapter 4.

6. 100% of PM and PdM tasks should be developed using


FMEA / RCM methodology.
a. True
b. False

Answer: a — True
All PM / PdM tasks should be developed using FMEA / RCM
methodology. This ensures cost effective and correct tasks to mitigate
certain risks and to find failures before they fail. See more details in
Chapter 3 and 8.

7. Utilization of assets in a world-class facility should be


about 85% or better.
a. True
b. False

Answer: a — True
Assets cost money to procure and maintain. They should be utilized
98% or better to get high ROI. Of course, our M&R task is to ensure
their availability; we need some time to perform maintenance too. See
more details in Chapter 4 and 9.

8. 100% of maintenance personnel’s (craft) time should be


scheduled.
a. True
b. False

Answer: a — True
100 % of maintenance personnel, specifically craft available hours,
should be scheduled. Scheduling compliance analysis should provide
opportunity to reduce/eliminate waste and improve productivity. See
more details in Chapter 4.
454 Appendix

9. Time-based PMs should be less than 20% of all PMs.


a. True
b. False

Answer: a – True
It’s a good and cost effective practice to do more run/cycle-based
and condition-based PM. It’s good practice to have calendar-based PMs
20% or less. If assets are operating 24/7, calendar-based PM could be a
higher percentage. See more details in Chapters 3, 4, and 8.

10. The 10% rule of PM is applied on critical assets.


a. True
b. False

Answer: a — True
This rule implies that time-based PM must be accomplished in 10%
of the time frequency or it is out of compliance. Many organizations use
this metric “PM Compliance” as a measurement of their maintenance
department’s performance, which is a good metric. But, we need to
ensure that critical assets are being maintained properly at the right time,
within 10% of time frequency. See more details in Chapters 3 and 4.

11. Most emergency work orders should be written by


production/operations.
a. True
b. False

Answer: a — True
Most emergency work orders should be written by the
production–operators. Operators are on the shop floor all the time and
they should know what needs to be fixed to meet the production sched-
ule. However, maintenance should also write WOs if emergencies arise.
Emergency and unscheduled work cost many times more than routine
scheduled work. See more details in Chapters 4 and 7.

12. It should be a common practice for Operators to perform PMs.


a. True
b. False

Answer: a — True
Appendix 455

This is true assuming the organization is deploying TPM as one of


the best practices. Under TPM philosophy, operators do perform PMs
and support maintenance. See more details in Chapter 7.

13. The P–F interval can be applied to visual inspections.


a. True
b. False

Answer: a — True
Yes, the primary objective is to detect a fault, or find the start of
one, and correct it before it fails. Detection can be visual or by using
predictive technologies. See more details in Chapter 8.

14. Understanding of a P-F curve should help in optimizing PM


frequency.
a. True
b. False

Answer: a — True
Yes, the primary objective is to detect a fault, or find the start of
one, and correct it before it fails. The PM frequency should be less than
the P-F interval. See more details in Chapter 8.

15. The best method of measuring the Reliability of an asset is by


counting downtime events.
a. True
b. False

Answer: a — True
Yes, Reliability is measured by MTBF, which is operating time
divided by the number of failures, or downtime events. See more details
in Chapter 6.

16. The primary purpose of scheduling is to coordinate maintenance


jobs for greatest utilization of the maintenance resources.
a. True
b. False

Answer: a — True
It’s true. The goal is to get all work completed as scheduled. See
more details in Chapter 4.
456 Appendix

17. What percentage of your assets should be ranked critical based


on the risk to business?
a. Less than 20%
b. 20–50%
c. over 50%

Answer: a — Less than 20%


We need to take care of all the assets cost effectively. However, we
don’t have unlimited resources to do that. We need to do everything pos-
sible to ensure all critical assets are being maintained properly. It has
been found that on average 20% +/-5% should be considered critical
assets as a good practice. See more details in Chapters 3 and 4.

18. Vibration monitoring can detect uniform impeller wear.


a. True
b. False

Answer: b — False
Uniform wear will not create any unbalance, hence no vibration. See
more details in Chapter 8.

19. Understanding the known and likely causes of failures can help
design a maintenance strategy for an asset to prevent or predict
failure.
a. True
b. False

Answer: a — True
If we understand the failure mechanism — how a part or component
can fail, we could develop a mitigating maintenance strategy to prevent
failure. See more details in Chapters 8 and 11.

20. Reliability can be improved easily after a maintenance plan has


been put into operation.
a. True
b. False

Answer: b — False
Reliability is a design attribute. It means that reliability is based on
how an asset is designed — with what type of components and their
Appendix 457

configurations. A maintenance plan can’t change the basic (inherent) relia-


bility unless the components are changed or redesigned with more reli-
able, higher MTBF parts. See more details in Chapters 6 and 8.

21. What percentage of maintenance work should be proactive?


a. 100%
b. 85% or more
c. 50%

Answer: b — 85%
The proactive work is defined as all work minus
unscheduled/unplanned work. We know that the planned and scheduled
work cost less than unscheduled, reactive work. We have found that it’s
good or best practice to have proactive work 85% or more. See more
details in Chapters 3 and 4.

22. MTBF is measured by operating time divided by the number


of failures of an asset.
a. True
b. False

Answer: a — True
MTBF is mean time between failures. It is calculated by dividing
operating time by the number of failures, or downtime events. See more
details in Chapter 6.

23. Maintenance cost will decrease as reliability increases.


a. True
b. False

Answer: a — True
Initially, when we are starting an M&R improvement plan, mainte-
nance cost may go up, but eventually it should come down as reliability
increases. With an increase in reliability or reduction in the number of
failures, assets are more available to perform their function See more
details in Chapter 3 and 6.
458 Appendix

24. The “F” on the P–F Interval indicates that equipment is still
functioning.
a. True
b. False

Answer: b — False
In the P–F interval curve, F stands for failure and P stand for poten-
tial failure. See more details in Chapters 6 and 8.

25. A rule of thumb is that, on average, an experienced planner can


plan work for how many craft people?
a. 5
b. 15
c. 25 or more

Answer: b — 15
On average, an experience planner should able to plan work for
about 15 +/5 craft people, depending upon type of work. Usually 15 is a
good number. See more details in Chapter 4.

26. Which of the following is not a primary objective for implement-


ing a Planning process?
a. Reduce reactive work
b. Prevent delays during the maintenance process
c. Mesh the production schedule and the maintenance
schedule

Answer: c — Mesh the production schedule and the maintenance sched-


ule
It’s not a P&S job to mesh maintenance and production schedules.
However, they may use production schedules to improve their schedule
or to identify conflicts. See more details in Chapter 4.

27. The best method of measuring the reliability of an asset is by:


a. MTTR
b. MTBF
c. Both
Appendix 459

Answer: b — MTBF
Reliability is measured by MTBF, which is operating time divided
by the number of failures. MTTR is a measure of maintainability. See
more details in Chapters 6 and 9.

28. With the exception of emergency work orders, Planning and


Scheduling will benefit all maintenance work.
a. True
b. False

Answer: a — True
The P&S should benefit all maintenance work. Planned and sched-
uled work cost much less and work get accomplished in a timely man-
ner. See more details in Chapter 4.

29. Leading KPIs predict results.


a. True
b. False

Answer: a — True
The leading indicators are process indicators and they lead to the
results. For example, PM compliances and backlog are leading indica-
tors. See more details in Chapter 9.

30. The 6th S in the 6 S (also called 5 S plus) process stands for
safety.
a. True
b. False

Answer: a — True
The original Five S (5 S — sort, set, shine, standardize, and sustain)
is a basic, systematic process for improving productivity, quality, and
housekeeping. Lately, a 6th S has been added to focus on safety too. 5 S
was originated in Japan. See more details in Chapter 7.

31. RCM stands for:


a. Regimented Centers of Maintenance
b. Reliability Centered Maintenance
c. Reliable Centers of Maintenance (uses best
practices)
460 Appendix

Answer: b — Reliability Centered Maintenance


See more details in Chapter 8.

32. The objective of RCM is to preserve functions.


a. True
b. False

Answer: a — True
The objective of RCM is to preserve function. See more details in
Chapter 8.

33. An MRO storeroom shouldn’t be stocking parts for


emergencies.
a. True
b. False

Answer: b — False
The FMEA/RCM analysis should provide us with details of failure
modes and what part we need to stock. However, in some cases, it may
be cost effective not to stock if parts can be procured or made available
locally in couple of hours. See more details in Chapter 5.

34. The inventory turnover ratio for MRO store should be:
a. Less than 2
b. Between 4–6
c. Over 6

Answer: a — Less than 2


Yes, it should be less than 2 for maintenance-related spares. Mostly
these are long lead and costly “A” type inventory. See more details in
Chapter 5.

35. PM compliance is a __________ KPI.


a. Lagging
b. Lagging or leading
c. Leading

Answer: b — Lagging/Leading
PM compliance is a leading indicator for availability and lagging
indicator for work execution. See more details in chapter 9.
Appendix 461

36. Quality is one key component of OEE.


a. True
b. False

Answer: a — True
OEE = Availability X Performance X Quality. See more details in
Chapter 7.

37. Reliability and Maintainability can only be designed in.


a. True
b. False

Answer: a — True
Reliability and Maintainability are design attributes. It means that
reliability and maintainability depend upon how the asset is designed
and with what type of components and configurations. A maintenance
strategy can’t change the basic (inherent) reliability. However, training
the workforce in repair techniques and providing the right tools will
improve the asset availability. See more details in Chapter 6.

38. Creating a reliability culture from a reactive mode can be


accomplished in a short period of time if enough resources are
made available.
a. True
b. False
Answer: b — False
Changing culture takes a long time. We can’t change a sustainable
culture overnight. It’s a journey that takes many years and is not depend-
ent on recourses; it’s time. See more details in Chapter 2.

39. Karl Fischer’s Coulometric Titration Method is an effective


technique to determine the metallic content (in PPM) in an oil
sample.
a. True
b. False

Answer: b — False
Karl Fisher’s method is used for determining water content — Parts
per Million (PPM) in an oil sample. See more details in Chapter 8.
462 Appendix

40. An IR thermography window is an effective method to satisfy


NFPA 70E arc flash requirements.
a. True
b. False

Answer: a — True
IR thermography windows are being used effectively to detect any
hot spots or potential problems in electrical cabinets, switchgears, etc.,
and help in meeting NFPA -70 E arc flash requirements. See more
details in Chapters 8 and 12.

41. FMEA is applicable only to assets currently in use.


a. True
b. False

Answer: b — False
FMEA can apply to any asset whether in use or not. In fact, it is a
good application in new systems being designed/ developed to identify
potential failure modes. See more details in Chapters 8 and 11.

42. RCM methodology can’t be used effectively on new


systems being designed.
a. True
b. False

Answer: b — False
RCM can be used on new or “in use” systems. In fact, it’s a good
application for new systems under development to use RCM methodolo-
gy to identify potential failure modes and to develop an effective PM
plan. See more details in Chapter 8.

43. Properly training the M&R workforce can increase asset and
plant availability.
a. True
b. False

Answer: a — True
Training the M&R workforce in application of new tools/techniques
will reduce repair time, resulting in higher availability. See more details
in Chapters 10 and 11.
Appendix 463

44. TPM is a type of maintenance performed by the operators.


a. True
b. False

Answer: a — True
Total Productive Maintenance (TPM) is another maintenance strate-
gy where an operator does some maintenance, sometimes called first
level, e.g., changing filters, minor adjustments, etc., and becomes part of
the maintenance crew in supporting major repairs. See more details in
Chapter 7.

45. Lagging KPIs are the results of a process.


a. True
b. False

Answer: a — True
Lagging indicators are the results. For example, Maintenance cost
and availability are lagging indicators. See more details in Chapter 9.

46. EOQ improves the inventory turn ratio.


a. True
b. False

Answer: b — False
EOQ (Economical Order Quantity) calculates the optimum order
quantity to optimize inventory cost. It does not impact the inventory
turnover ratio. See more details in Chapter 5.

47. New incoming oil from the supplier is always clean and
ready to be used.
a. True
b. False

Answer: b — False
It has been found that the incoming oil is not usually clean and does
not meet cleanliness requirements. The best organizations are establish-
ing oil cleaning systems to clean all incoming oil to ensure that new oil
meets cleanliness requirements. See more details in Chapter 8.
464 Appendix

48. Which phase of asset life cycle has the highest cost?
a. Design
b. Acquisition
c. O & M

Answer: c — O&M
O&M phase usually has the highest cost in an asset’s life cycle. See
more details in Chapter 6.

49. Most of the maintenance costs become fixed


a. After installation
b. During operations
c. During design

Answer: c — Design
Most maintenance costs get fixed during the design phase. See more
details in Chapter 6.

50. RCM provides best results when used


a. During Operation /Production
b. During Design / Development
c. After an asset has failed or keeps failing

Answer: b — Design
RCM should be used in the design phase to get maximum benefits.
See more details in Chapters 6 and 8.

51. How soon we can restore an asset is measured by:


a. MTBF
b. MTTR
c. MTBMA
d. None of the above

Answer: b — MTTR
MTTR is a measure of how soon can we bring the asset back to
operations. See more details in Chapter 6.
Appendix 465

52. Availability is a function of:


a. MTBF
b. MTTR
c. Uptime
d. Uptime and downtime

Answer: d — Uptime and downtime


Availability is defined by uptime divided by uptime plus downtime.
See more details in Chapter 6.

53. The failure rate of a component / asset can be calculated by


knowing:
a. Number of failures
b. MTBF
c. MTTR
d. Uptime

Answer: b — MTBF
Failure rate is the inverse of MTBF. See more details in Chapter 5.

54. The biggest benefit of a Failure Modes and Effects Analysis


occurs during:
a. Operations phase
b. Maintenance phase
c. Design phase
d. None of the above

Answer: c — design
FMEA should be performed during the design phase to identify fail-
ure modes and those that could be eliminated or their impact reduced or
mitigated cost effectively. See more details in Chapters 6 and 11.

55. PM schedule compliance should be equal or greater than


95%.
a. True
b. False

Answer: a — True
High (95% or better) PM schedule compliance will catch potential
failures before they happen, thereby reducing the unexpected break-
downs. See more details in Chapters 3 and 4.
INDEX

Index Terms Links

5S 190 203–208 214


390
5 S plus 3 459
5 whys 353 364–366
6S 13

ABC 127
AC high potential testing (HiPot) 268
acceleration 248
accountability 285
accuracy 135–137
acoustics 55
acquisition 15 176
action plans 29
active inventory 123
active redundancy 168–170
admire 316
affinity analysis 396–397
affordability 70
age exploration 221
agent 38
alarms 62

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

alignment 253
American National Standards Institute 402
analysis 351–399
analytical ferrography 263–264
ANSI 402
arc flash hazards 423–426
AS/RS 133–134
ASME 206
assessment 43 76–78 404
asset 3 5 10
12 38 39
51 211–212 221
363 453 464
465
critical 11 97
life cycle 15 176–178
audits 412–413
Austin, Nancy 26
automated ID 143
automated storage 133
autonomous maintenance 60 197–198
availability 16 28 156
157–161 166 190
209 212 462
465
awareness 37

baby boomers 313 318–319


backlog 52 98 113

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

balanced scorecard 292–296


bar code 143–144
barricades 111
barrier analysis 353 397–398
Barringer, Paul 177
bathtub curve 161–162
behavior 37 420
benchmark 3–4 6–9 76
286–287 298–305 334–335
Bennis, Warren 27 30
best practices 1–10 75 286
451–465
best-in-class 286
bill of material 85 120
bins 131 136
brainstorming 390
breakdowns 8–9 95 198
broadband trending 252
budget 118
budget 285
Burns, James McGregor 24
business process perspective 295
Byham, Tacy 43

calendar-based PM 93
capital project 52 108
capital project maintenance, see CPM
carousels 133
carrying costs 139–140

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

category codes 91
causal factors 361
cause mapping 353
cause-and-effects analysis 353 366-370 390
CBM 39 52 55–57
58 66 77
84 87 94
220 221 245–274
426
centralized 341–342
certification 313 335–338
certified maintenance & reliability pro
fessional, see CMRP
chains 205
change 23 28 321–322
change agent 38
change management 21 36–38 43
charisma 28
checklist 103 354
classification 91–98 122–124 145–146
cleaning 61
CM 52 59 87
93 94–95 153
CMMS / EAM 41 51 63–74
77–78 84 85
96 106 120
134–135 136 437
CMRP 30 336–338
collaboration 43
collect data 360–361

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

color codes 206–207


commonly used parts 120
communication 28 37 110–111
286 313 323–329
competence 28 43–44
compliance 14 86 113
460 465
schedule 16
component 51 170–172 465
compressed air 410–411
compressor X1 system 173
computerized maintenance management
system, see CMMS / EAM
condition based maintenance, see CBM
condition monitoring 246–248
condition-directed tasks 221
configuration 402 438–441
consensus 43
continuous improvement 202 243
control charts 354
cooling 412
coordinators 85 88
corona 221
corporate strategy 35
corrective maintenance, see CM
correlation analysis 247
corrosion 403 434–437

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

cost 6 12 38
69 137–138 176–178
230 244 408–409
434 457 464
Coulometric Titration Method 14 461
Covey, Steven 20 33
CPM 52
craft supervisor 88 91
creating, culture 39–42
critical assets 11 221 454
456
criticality 97
cultural change 21
culture 14 19–47 345–346
461
customer perspective 295–296
customers 154

data analysis 68–69


data collection 246 297–298
de-centralized 341–342
decision tree 238–240
define problem 360
Deming, Dr. W. Edwards 310–312 355
design 15 67–68 181–184
199 200 228
reliability 151–186
six sigma 354
design of experiments (DOE) 354

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

detection interval 252


development 200 329-338
discipline 37
displacement 248
diversity 313 340–341
division of labor 189
documentation 244
downtime 16 158 211
213
dust control 111
duties 332

EAM, see CMMS / EAM


ease of use 72
economic order quantity, see EOQ
eddy current testing 273
effectiveness 363
efficiency 213 244
electrical condition monitoring 266–268
electrical measurements 56 247
emergency showers 111
emergency spares 120
emergency work orders 11 13 95
454
emissivity 221
employee involvement 313
employee life cycle 313 314-316
encoding 324
energizing 28

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

energy costs 408–409


energy usage 409-412
enterprise asset management, see
CMMS / EAM
environment 201 332 363
420
EOQ 15 135 138
139 463
equipment 244 332 413–414
equipment design 199 200
evaluation 43
experience 44
external benchmarking 299
eye baths 111

failure 12 39 73–74
156 165 222
224–225 274–275 277
456
failure mode 222 227–228
Failure Mode and Effect Analysis, see
FMEA
failure rate 16 156 164
166 465
fans 410
fault tree 354–355 380–381
feedback 325
ferrography 222 263–264
financial perspective 296

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

Fischer, Karl 461


fishbone diagram 366–370
five S, see 5 S
flexibility 321 346–347
flow chart 355
flow rates 56 247
FMEA 10 14 16
51 69 119
182 222 236–238
277 354 370–380
453 462 465
foaming 264
focused improvement 198–199
forcing frequencies 251
Ford, Henry 389
Fosbury, Dick 2
frequencies 251
functional block diagram 232–234
functional failure 222 234–235

generation gap 317–323


generation Xers 313 319–321
generation Yers 314 321–323
goals 28–29 36 43
287
green energy 403 415–416
guidelines 373–377

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

Harari, Oren 24 26
harmonic distortion 270
hazard 403
hazardous material 403
health 201
Heap, Harold 226
heating 412
hidden failure 222
hire 315
Howe, Neil 317
human error 237
hybrid organization 341–342
hydrostatic testing 272–273

impact 97
impeller wear 12
implementation 241–242 362
improvement 76–78 202
improvement cycle 355
inactive inventory 145
infrared 426
infrared spectroscopy 263
infrared thermography 55 248 254–256
462
infrequently used inventory 123
inherent reliability 5
initiatives 43

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

inspire 315–316
internal benchmarking 299
inventory 14 117–149
costs 126
shrinkage 146
stratification 124–128
types 122–128
inventory turnover ratio 120 460

Japanese words 192


jishu hozen 197–198
JIT 121 190
job plans 107
job priority 96
job task analysis 314 330–333
job template 103
Jones, Daniel T. 388
just-in-time inventory, see JIT

kaizen 198–199
kanban 390
Kaplan, Robert 292
key performance indicators, see KPIs
kitting 87 113 136–137
knowledge 3 9–16 44
kobetsu 198
kouzes, James M. 42

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

KPIs 13–15 43 68
76–77 288 292
459 460 463
Krafcik, John 388

labeling 205
labor 189
lagging indicators 291–292
layout 128–134 413–415
leaders 27
leadership 42 19–47 345–346
419
attributes 27
secrets 24–26
Leadership Practices Inventory 42
leading indicators 291–292
leak detection 258–260
lean maintenance 387–393
lean manufacturing 190
learning 319
learning and growth perspective 294–295
LEED 403 416–417
life cycle 69 176–178 244–245
464
asset 15
lighting 412
limits 247
Lincoln, Abraham 26 83
liquid handling 111–112

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

listening 325–327
logic tree analysis (LTA) 238–240
logistics 182
loss 8–9 198–199 212–213
lubricant analysis 248 260–266
lubricating 61

M&R 6-16 31 39
M&R analysis 351–399
magnetic particle testing (MPT) 272
maintainability 10 14 151–186
452 461
maintenance 4 31 49–80
453
approaches 53–59
assessment 76–78
autonomous 60
backlog 52
capital project 52
condition based 52 55–57
corrective 52 59
cost 6–7 12 15
38
improvement 76–78
metrics 290
operator based 52 60–63
predictive 52
preventive 53 57–58 200
proactive 53 58–59

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

maintenance (Cont.)
productive 195
quality 74–76
reactive 53
run-to-failure 53
maintenance & reliability, see M&R
maintenance department 34
maintenance plan 5
man to part 130–132
management by wandering around 6
materials 117–149 363
Matteson, Tom 226
MBWA 26
mean time between failures, see
MTBF
mean time to repair, see MTTR
measures 42–44 151–186
meetings 328–329
megohmmeter testing 269
Mentzer, Bill 226
metrics 287 289 290
mishap 403
mission 21 28
mission statement 21 32–35
mistake proofing 355 413
mitigation 403 432–433
mobile technology 70
modular drawers 131
moisture 262
morale 321

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

motivation 3
motor current readings 268
motor current signature analysis
(MCSA) 268
motors 410 411
m-out-of-n reliability 170–172
MRO 14 460
MTBF 12 13 16
68 156–158 163
166 173 179
457 459 464
465
MTBMA 16 464
MTTR 13 16 156
158–159 180 459
464 465
muda 355 390
multiple component system 171
mura 355 389–390
muri 355 389

Nakajima, Seiichi 195


Nanus, Burt 27
narrowband trending 252
networking 287 319
new hires 331
new technologies 143
Nippondenso 194
noise control 112

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

non-destructive testing 271-273


Norton, David 293
Nowland, Stan 226

O&M 15 177
objectives 28 287
OBM 52 60–63 94
OEE 10 14 191
208–214 417 452
461
OEM 277
office improvement 201
Ohno, Taiichi 389
oil 463
oil analysis 56
oil contamination 265-266
oil levels 205
operating context 223
operating environment 5 180
operations 187–217
operator based maintenance, see OBM
operator driven reliability 191 193
operators 61 454
optimization 77 134–145 219–282
455
oragnizational initiatives 43
ordering 139
organization 203–208
organizational culture 21–22

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

organizational structure 314 338–347


original equipment manufacturers (OEM) 193
outsourcing 342–344
overall equipment effectiveness, see OEE
ownership 285

pallet rack 131


Palmer, Doc 99 107
parallel systems 168–170
pareto analysis 355 385–387 390
part to man 130–132
partial discharge detection 56 221
parts 9 101–102 117–149
244
parts kitting 87 137
pattern recognition 247
PDCA (Deming’s improvement cycle) 355
PdM 10 52 55–57
66 221 245–274
453
people 43
people development 329–338 364
performance 4 42–44 56
113 147–147 209
212–214 283–307 347
433–434
performance review 331
personal attributes 44
personal protection equipment 106 111

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

personnel 10 11 107
329–338 419–420 452
453
Peters, Tom 26
P-F interval 11 12 223
224 455 458
Phillips, Donald 26
physical plant 34
pillars of TPM 197-201
planned maintenance 199
planned work 8 10 85
113
planner 85 88 90
458
planning 13 83–86 458
459
capabilities 102–105
checklist 104
ineffective 102
process 98–105
plant 38 42 230
413
PM 10 11 14
16 53 57–58
66 75 77
84 86 87
93 153 195
219–282 453 465
calendar-based 93
condition based 94

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

PM(Cont.)
operator based 94
run-based 94
time-based 454
poke-yoke 390
Posner, Barry Z. 42
potential failure 223
Powell, Gen. Colin 24–26
power factor 270 319
PPE 106 111 403
423
predictive maintenance, see PdM
pressure 56 61 247
prestige 319
preventive maintenance, see PM
prioritization 91-98
proactive maintenance 12 36 53
58-59 99–100 457
procedures 61–62 101
process heating 410
process improvements 28
product data 434
production/operations 8–9
productive maintenance 195
productivity 67–68 244
professional success 43
prognosis 223
pumps 410

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

qualified people 314


quality 14 74–76 209
213 381–385 461
quality control 155
quality defects 199
quality maintenance 199

radio frequency identification (RFID) 144–145


radio frequency monitoring 269–270
radiography 271
ranges 247
rarely used inventory 123
RAV 6–7 9 296
RBD 167–169 171–176
RCA 10 69 453
RCM 13 15 39
53 69 152
157 183 220–221
223 225–245 459-460
462 464
analysis 229–245
elements 228–229
history 226–227
other processes 243–245
principles 227–228
reactive CM 95
reactive maintenance, see RM

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

reactive thinking 36
reactive work 99
redundancy 168–170
regulations 277
relationships 333
reliability 4 11 12
53 63 151–186
455–457 461
culture 14
design 181–184
failure distribution 162–163
inherent 5
metrics 290
models 170–172
requirements 178–179
specifications 178–181
reliability block diagram, see RBD
reliability culture 38–42 461
repairs 41 62 95
165 166
replacement asset value, see RAV
reporting 68–69
reputation 154
resistance 37
resources 28 100–101 285
338–347 455
retire 316
retrieval system 130 133–134
rewards 189
rework 113

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

risk 12 75 403–404
risk management 427–434
RM 53
roles 37 87–91
root cause analysis 356 357–381
rotating bomb oxidation test (RBOT) 264
RTF 53 220 223
238 241 278–279
run-based PM 94
run-to-failure maintenance, see RTF
rust prevention 264

safety 13 182 201


228 363–364 417–422
433
sampling 264–265
scaffolding 112
scatter diagram 356
schedule 8 321 434
453 455 459
465
compliance 8 10 16
86
scheduled CM 95
scheduled work 86
schedulers 86 88 90
scheduling 11 13 66
84 86 105–107
scope management 109

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

seiketsu 207–208
seiri 203
seiso 207
seiton 204–207
service stock 121
set-in-order 204–207
setup 198
shelf life 135
shelves 131
shine 207
shock pulse method 56 253
shutdowns 108–112
signature analysis 252
silent generation 317–318
six sigma 354 356 381–385
459
skill levels 100–101
skills development 333–334
SMART test 289
Smith, Audrey 43
SMRP 30 34–35 83
336–337
SMRP Initiative 301–303
Society of Maintenance & Reliability
Professionals, see SMRP
solid waste 111–112
solutions 362
sort 203
spare parts 121
spectrometric metals analysis 263

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

spectrum analysis 253


speed 199
standard deviation 356
standardize 207–208
standards 373–377 441–447
statistical process analysis 247
status codes 90
steam 410
steps 101
stock keeping unit (SKU) 121
stock supplies 130
stocking levels 139
stock-out 9 146
stoppage 198–199
store 121
storeroom 14 119 128–134
460
strategic framework 29
strategy 21 29 35
338–339
stratification 121 356
Strauss, William 317
structured maintenance 54–55
success 29 43
success flow 23
succession planning 314 344–345
supervisor 107
suppliers 15
surge comparison testing 268–269
sustain 208

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

sustainability 404–408
sustaining, culture 39–42
system affordability 70
system boundary 231–232
system selection 230–232
systems engineer 88 437–439

tasks 77 228 240-242


332
Taylor, Frederick 2
teams 37
teamwork 285 328
technology 70 143 270–271
337 434
TEEP 191 209 210
214
temperature 56 61 247
theory of constraints (TOC) 356-357 394–396
thermography 14 55
time domain reflectometry 269
time-based PMs 454
time-directed tasks 223
tools 101–102 134–145 205
332
total acid number (TAN) 263
total base number (TBN) 263
total effective equipment performance,
see TEEEP

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

total productive maintenance, see


TPM
total quality management (TQM) 195
TPM 15 60 190
191 194–202 413
463
training 28 37 75
200 320 331–332
333–335 462
trend analysis 247
turnaround 86 108–112
turnover 43 120–121 147
460
turnover ratio 14

ultrasonic measurements 55 248 256-260


271-272 426
uptime 16 157
utilization 10
utilization rate 191 211–212 453
455

validation 404
value stream mapping (VSM) 357 390 392–393
values 29
variance 146
velocity 248

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

vendors 72 146
verification 404
vibration 247–253 456
vibration analysis 55
vibration monitoring 12
viscosity 223 262
vision 21 26 28–32
287
vision statement 21 30
visual and odor tests 261–262
visual controls 413
visual workplace 192

water test 262


Waterman, Robert 26
waveform analysis 253
wear 12
wear particle analysis 260–266
Womack, James P. 388
work, impact 97
work classification 91–98
work flow 87–92
work instruction 242
work management 81–115
work orders 10 11 13
65 87 90
454
work performer 88
work plan 87

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

workforce 15 62 309–350
world-class 288 310
wrench time 107

This page has been reformatted by Knovel to provide easier navigation.

You might also like