Springer Series in Reliability Engineering
For further volumes: http://www.springer.com/series/6917
Toshio Nakagawa
Stochastic Processes with Applications to Reliability Theory
123
Dr. Toshio Nakagawa Department of Business Administration Aichi Institute of Technology 1247 Yachigusa, Yakusa-cho 470-0392 Toyota Japan e-mail:
[email protected]
ISSN 1614-7839 ISBN 978-0-85729-273-5
e-ISBN 978-0-85729-274-2
DOI 10.1007/978-0-85729-274-2 Springer London Dordrecht Heidelberg New York British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library Ó Springer-Verlag London Limited 2011 Apart from any fair dealing for the purposes of research or private study, or criticism or review, as permitted under the Copyright, Designs and Patents Act 1988, this publication may only be reproduced, stored or transmitted, in any form or by any means, with the prior permission in writing of the publishers, or in the case of reprographic reproduction in accordance with the terms of licenses issued by the Copyright Licensing Agency. Enquiries concerning reproduction outside those terms should be sent to the publishers. The use of registered names, trademarks, etc., in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant laws and regulations and therefore free for general use. The publisher makes no representation, express or implied, with regard to the accuracy of the information contained in this book and cannot accept any legal responsibility or liability for any errors or omissions that may be made. Cover design: eStudio Calamar, Berlin/Figueres Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com)
Preface
I have learned first reliability theory from the book Mathematical Theory of Reliability [1] written by Barlow in 1965 that is the most famous and excellent one theoretically up to now. I was a marvelously lucky reader, unfortunately, I could not understand throughly this book at that time because my poor mathematical tools and abysmal ignorance about stochastic processes. I have now already published three monographs Maintenance Theory of Reliability [2], Shock and Damage Models in Reliability Theory [3], and Advanced Reliability Models and Maintenance Policies [4] in which I have surveyed mainly maintenance policies in reliability theory on the research results of the author and my colleagues. Most of the three books have been written based on basic theory of stochastic processes and their mathematical tools. A number of graduate students, researchers, and engineers demand of us some book written in an easy style on stochastic processes to be able to understand readily reliability theory. Recently, plants, satellites, computer and information systems have become more large-scale and complex, and most products are distributed all over the world. If they or some of them would fail and have trouble, it might incur many serious and catastrophic situations and heavy damage. More and better maintenance is required constantly from the economical and environmental points as public infrastructures and operating plants become old in advanced nations. Reliability and maintenance theory is more useful for protecting such severe matters and environmental considerations, and moreover, for making good and safe living. Stochastic processes can be described by probabilistic phenomena in some space at each point of time t. The knowledge of stochastic processes and mathematical tools is indispensable for engineers, managers and researchers in reliability and maintenance. This book introduces basic stochastic processes from the viewpoint of elementary mathematics, mainly based on the books Stochastic Processes [5], Applied Stochastic System Modeling [6], and Introduction to Stochastic Processes [7]. Many reliability examples are cited in this book to explain concretely stochastic processes. They are quoted mainly from the books [1–4]. Furthermore, several interesting reliability examples appeared in two famous and v
vi
Preface
classical books [8, 9]. However, they have become old and their important contributions are going to be forgotten. This book is aimed at quoting possible examples from the two books. This book is composed of eight chapters: Chapter 1 is devoted to the introduction to stochastic processes and reliability theory. Chapters 2 –5 are devoted to standard and fundamental stochastic processes such as Poisson processes, renewal processes, Markov chains, Markov processes, and Markov renewal processes. These four chapters are written mainly based on the books [5–7] and are introduced from the viewpoint of elementary mathematics without using hard proof. Chapter 6 is devoted to cumulative processes that are related greatly to shock and damage models in one field of reliability. Chapter 7 introduces simply Brownian motion and Lévy processes, that might be omitted at the first reading. To understand and explain stochastic processes easily and actually, a lot of examples are quoted from reliability models through each chapter. As final examples of reliability models, Chapter 8 takes up redundant systems and shows systematically how to use well the tools of stochastic processes to analyze them. The reader could learn both stochastic process and reliability theory from this book at the same time. I wish to thank most kindly Professor E. Çinlar, Professor S. M. Ross, and Professor S. Osaki for referring to the good books written by them and the other books in references. I wish to express my special thanks to Professor Fumio Ohi and Professor Mingchih Chen for their careful reviews of this book, and to Dr. Kodo Ito, Dr. Satoshi Mizutani, Dr. Sayori Maeji, and Mr. Xufeng Zhao for their support in writing and typing this book. Finally, I would like to express my sincere appreciation to Professor Hoang Pham, Rutgers University, and Editor Anthony Doyle, Springer-Verlag, London, for providing the opportunity for me to write this book. Toyota, April 2010
Toshio Nakagawa
References 1. 2. 3. 4. 5. 6. 7. 8. 9.
Barlow RE, Proschan F (1965) Mathematical theory of reliability. Wiley, NewYork Nakagawa T (2005) Maintenance theory of reliability. Springer, London Nakagawa T (2007) Shock and damage models in reliability theory. Springer, London Nakagawa T (2008) Advanced reliability models and maintenance policies. Springer, London Ross SM (1983) Stochastic processes. Wiley, New York Osaki S (1992) Applied stochastic system modeling. Springer, Berlin Çinlar E (1975) Introduction to stochastic processes. Prentice-Hall, Englewood Cliffs, NJ Takács L (1960) Stochastic processes. Wiley, New York Cox DR (1962) Renewal theory. Methuen, London
Contents
1
Introduction . . . . . . . . . 1.1 Reliability Models . 1.1.1 Redundancy . 1.1.2 Maintenance . 1.2 Stochastic Processes. 1.3 Further Studies . . . . 1.4 Problems 1 . . . . . . . References . . . . . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
1 2 2 2 2 5 5 6
2
Poisson Processes . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Exponential Distribution. . . . . . . . . . . . . . . . 2.1.1 Properties of Exponential Distribution 2.1.2 Poisson and Gamma Distributions . . . 2.2 Poisson Process . . . . . . . . . . . . . . . . . . . . . 2.3 Nonhomogeneous Poisson Process . . . . . . . . 2.4 Applications to Reliability Models . . . . . . . . 2.4.1 Replacement at the Nth Failure. . . . . . 2.4.2 Software Reliability Model . . . . . . . . 2.5 Compound Poisson Process . . . . . . . . . . . . . 2.6 Problems 2 . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
7 8 9 14 18 27 34 34 36 36 42 45
3
Renewal Processes . . . . . . . . . . . . . . . . . . . . . 3.1 Definition of Renewal Process . . . . . . . . . 3.2 Renewal Function . . . . . . . . . . . . . . . . . . 3.2.1 Age and Residual Life Distributions 3.2.2 Expected Number of Failures . . . . 3.2.3 Computation of Renewal Function . 3.3 Age Replacement Policies . . . . . . . . . . . . 3.3.1 Renewal Equation . . . . . . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
47 48 50 58 61 64 66 66
. . . . . . . .
. . . . . . . .
vii
viii
Contents
3.3.2 Optimum Replacement Policies . . . . . Alternating Renewal Process . . . . . . . . . . . . 3.4.1 Ordinary Alternating Renewal Process 3.4.2 Interval Reliability . . . . . . . . . . . . . . 3.4.3 Off Distribution . . . . . . . . . . . . . . . . 3.4.4 Terminating Renewal Process . . . . . . 3.5 Modified Renewal Processes. . . . . . . . . . . . . 3.5.1 Geometric Renewal Process . . . . . . . 3.5.2 Discrete Renewal Process . . . . . . . . . 3.6 Problems 3 . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
69 73 73 76 79 82 84 84 86 91 93
4
Markov Chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 Discrete-Time Markov Chain . . . . . . . . . . . . . . . . . . . 4.1.1 Transition Probabilities . . . . . . . . . . . . . . . . . . 4.1.2 Classification of States . . . . . . . . . . . . . . . . . . 4.1.3 Limiting Probabilities . . . . . . . . . . . . . . . . . . . 4.1.4 Absorbing Markov Chain . . . . . . . . . . . . . . . . 4.2 Continuous-Time Markov Chain . . . . . . . . . . . . . . . . 4.2.1 Transition Probabilities . . . . . . . . . . . . . . . . . . 4.2.2 Pure Birth Process and Birth and Death Process . 4.2.3 Limiting Probabilities . . . . . . . . . . . . . . . . . . . 4.3 Problems 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
95 97 97 101 103 108 111 111 112 119 120 121
5
Semi-Markov and Markov Renewal Processes . . . . . . . . 5.1 Markov Process . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Embedded Markov Chain . . . . . . . . . . . . . . . . . . . . 5.2.1 Transition Probabilities . . . . . . . . . . . . . . . . . 5.2.2 First-Passage Distributions . . . . . . . . . . . . . . . 5.2.3 Expected Numbers of Visits to States . . . . . . . 5.2.4 Optimization Problems . . . . . . . . . . . . . . . . . 5.3 Markov Renewal Process with Nonregeneration Points 5.3.1 Type 1 Markov Renewal Process . . . . . . . . . . 5.3.2 Type 2 Markov Renewal Process . . . . . . . . . . 5.4 Problems 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
. . . . . . . . . . . .
123 124 132 133 136 138 140 140 141 143 147 148
6
Cumulative Processes . . . . . . . . . 6.1 Standard Cumulative Process . 6.2 Independent Process . . . . . . . 6.3 Modified Damage Models . . . 6.3.1 Imperfect Shock . . . . . 6.3.2 Random Failure Level .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
149 150 155 162 162 164
3.4
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . . . . . . .
. . . . . .
. . . . . . . . . . .
. . . . . .
. . . . . . . . . . .
. . . . . .
. . . . . . . . . . .
. . . . . .
. . . . . . . . . . .
. . . . . .
. . . . . . . . . . .
. . . . . .
Contents
6.3.3 Damage with Annealing. . . . . 6.3.4 Increasing Damage with Time 6.4 Replacement Models . . . . . . . . . . . . 6.4.1 Three Replacement Policies . 6.4.2 Optimum Policies . . . . . . . . 6.5 Problems 6 . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . .
ix
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
166 169 170 171 173 179 180
7
Brownian Motion and Lévy Processes. . . . . . . . . . . . . . 7.1 Brownian Motion and Wiener Processes. . . . . . . . . . 7.2 Three Replacements of Cumulative Damage Models 7.2.1 Three Damage Models . . . . . . . . . . . . . . . . 7.2.2 Numerical Examples . . . . . . . . . . . . . . . . . 7.3 Lévy Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 Problems 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
183 183 187 187 189 191 196 196
8
Redundant Systems . . . . . . . . . . . . . . . . . . . . . . . . . 8.1 One-Unit System . . . . . . . . . . . . . . . . . . . . . . . 8.1.1 Poisson Process . . . . . . . . . . . . . . . . . . . 8.1.2 Nonhomogeneous Poisson Process . . . . . . 8.1.3 Renewal Process . . . . . . . . . . . . . . . . . . . 8.1.4 Alternating Renewal Process . . . . . . . . . . 8.2 Two-Unit Standby System . . . . . . . . . . . . . . . . . 8.2.1 Exponential Failure and Repair Times . . . 8.2.2 General Failure and Repair Times . . . . . . 8.3 Standby System with Spare Units . . . . . . . . . . . . 8.3.1 First-Passage Time to Systems Failure. . . . 8.3.2 Expected Number of Failed Units . . . . . . . 8.3.3 Expected Cost and Optimization Problems . 8.4 Problems 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
199 200 200 200 201 202 203 204 205 211 211 213 215 216 217
Appendix A: Laplace Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . .
219
Appendix B: Answers to Selected Problems . . . . . . . . . . . . . . . . . . . .
225
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
249
. . . . . . . . . . . . . . .
. . . . . . . . . . . . . . .
Chapter 1
Introduction
All things from near oneself to space form stochastic processes whether large or small because their real lives are limited. Of course, each lifetime of human being forms a stochastic process [1], however, such process could not be analyzed exactly because given conditions and constrains would be too many and much complex at the present, as shown in a toy of life games. Most reliability models are much simpler and more clear than human being, and much more, they have to be analyzed practically and be of great use in actual fields. The importance of reliability will be greatly enhanced by environmental considerations, and moreover, for the protection of natural resources and the earth. Reliability techniques have to be developed and expanded as objective models become more complex and large-scale. They also will be applied not only to daily life, but also to a variety of other fields because consumers, workers, and managers must make, buy, sell, and use, and handle articles and products with a sense of safety and security. A wide knowledge of probabilities, statistics, and stochastic processes are needed for learning reliability theory mathematically. Stochastic processes have been applied in many fields such as engineering, physics, biology, electronics, management science, economics, and psychology, and in operations research, they have spread widely over queueing, finance, and insurance. Especially, in reliability theory, stochastic processes are the most powerful mathematical tools for analyzing reliability models. We aim to write this book about stochastic processes from viewpoints of reliability theory and applications: This book introduces basic stochastic processes such as Poisson processes, renewal processes, Markov chains and processes, and furthermore, introduces cumulative processes, Brownian motion, and Lévy processes. A lot of examples cited from reliability models demonstrate how to apply stochastic processes correctly.
T. Nakagawa, Stochastic Processes, Springer Series in Reliability Engineering, DOI: 10.1007/978-0-85729-274-2_1, Ó Springer-Verlag London Limited 2011
1
2
1 Introduction
1.1 Reliability Models 1.1.1 Redundancy High system reliability can be achieved by redundancy and maintenance. A classical problem is to determine how reliability can be improved by using redundant units. Various expressions of reliability quantities of many redundant systems were summarized [2, 3]. A two-unit system is the most fundamental redundant model, and its reliability measures and optimum maintenance policies were surveyed deeply [4]. For systems with redundant or spare units, we have to determine how many number of units should be provided initially. It is well-known as fault tolerance [5] in which systems continue to function correctly in the presence of hardware failures and software errors. Redundancy techniques of systems for improving reliability and achieving fault tolerance are classified commonly in the following forms [5, 6]: hardware redundancy, software redundancy, information redundancy, and time redundancy.
1.1.2 Maintenance Maintenance of units after failure may be costly, and sometimes requires a long time. Thus, the most important problem is to determine when and how to maintain preventively units before failure. However, it is not wise to maintain units with unnecessary frequency. From such viewpoints, the commonly considered maintenance policies are preventive replacement for units without repair, and preventive maintenance for units with repair on a specific schedule. Classifying into three large groups, planned maintenance of units are done at a certain age, a periodic time, or a specified number of occurrences. Consequently, the object of optimization problems is to determine the frequency and timing of several kinds of maintenance and replacement policies in accordance with maintenance costs and effects, and system structures and circumstances.
1.2 Stochastic Processes All reliability models consisting of random time factors form stochastic processes. We briefly explain stochastic processes as examples of reliability systems with maintenance: Consider the simplest one-unit system with repair or replacement whose time is negligible, i.e., the unit is operating and is repaired or replaced at failure, where the time required for repair and replacement is negligible. When the repair or replacement is completed, the unit becomes as good as new and begins to operate.
1.2 Stochastic Processes
3
If failures occur exponentially, i.e., the unit fails constantly during the time interval ðt; t þ dt irrespective of time t, the system forms a Poisson process. Roughly speaking, failures occur randomly in ½t; t þ dt with probability kdt for constant k [ 0; and interarrival times between failures have an exponential distribution ð1 ekt Þ: Then, it is said that failures occur in a Poisson process with rate k: A Poisson process is the simplest and the most fundamental stochastic process. In Chap. 2, we define a Poisson process theoretically and explain various properties of an exponential distribution with memoryless property. As examples of a Poisson process, we consider a parallel system, a spare part problem, an inspection policy, a backward time problem, and so on. Furthermore, we introduce a nonhomogeneous Poisson process in which failures occur in ½t; t þ dt with probability kðtÞdt; and a compound Poisson process in which an amount of some quantities such as customers, damage, costs and profits occurs in a Poisson process. To apply two processes to reliability models, we take up periodic replacements, and shock and damage models. If failures occur generously as an extended process of a Poisson one, the system forms a renewal process, i.e., a renewal theory arises from the study of selfrenewing aggregates. The process plays a major role in the analysis of probability models with sums of independent nonnegative random variables, and is more useful tool in reliability theory because many reliability models have a renewal property by replacement and maintenance. In Chap. 3, we define a renewal process theoretically, and investigate and give a renewal function in theorem forms. The most important matters in a renewal process are how to compute renewal functions and to form renewal equations. Optimum age and periodic replacement policies are shown as practical examples. Furthermore, we introduce an alternating renewal process where the repair or replacement time needs a nonnegligible time for a one-unit system, and a discrete renewal process with only discrete times. As examples of two processes, we give a one-unit system with repair and a discrete age replacement, and obtain their reliability quantities using a renewal theory. We have considered a Poisson process with only one state in which the unit is operating. When we consider a one-unit system whose repair time is nonnegligible, the process forms an alternating renewal process with two states that repeats up and down alternately. In addition, when we consider two types of repair maintenance according to minor or major failures of the unit, the process forms a stochastic process with three states. In such ways, if the process transits among many states and has a Markov property in which the future behavior depends only on the present state and is independent of its past history, it forms a Markov process. If the duration times of states are multiples of time unit such as day, week, month, year, and a certain number of occurrences, then the process is called a discrete-time Markov chain, and if the duration time are distributed exponentially, it is called a continuous-time Markov chain. In Chap. 4, we show how to compute transition probabilities of a Markov chain and to obtain its several useful quantities from them. As examples of reliability models, we take up a one-unit system with repair
4
1 Introduction
and a two-unit parallel system, and derive their reliability measures by using the techniques of a Markov chain. We consider a one-unit system with two types of repair maintenance. If neither up and down times are exponential, the process does not have the Markov property at any time except only transition times which are called regeneration points. The process forms a Markov renewal process or a semi-Markov process that is a marriage of renewal processes and Markov chains. If we are interested mainly in the number of visits to states, the process is called a Markov renewal process, and if we are interested in what state the process is at time t, it is called a semi-Markov process. However, there is not so much difference between two given names. In Chap. 5, we derive first-passage distributions, transition probabilities, and renewal functions by forming renewal equations. As an example of a standby system with repair, we obtain a mean time to failure, transition probabilities, and expected numbers of visits to states. Markov renewal processes give powerful plays for analyzing reliability systems because they include renewal and Markov processes. Learning the techniques of Markov renewal processes is positively necessary for analyzing reliability models. Naturally, such processes are not all-round. For example, we consider a two-unit parallel system with one repair person when both failure and repair times are not exponential that seems to be a simpler system. However, we cannot analyze the system because there is no generation point, i.e., the process has no Markov property at any time. Next, consider a two-unit standby system with one repair person. Then, the process has two regeneration points only at times when one unit fails and another is in standby, and the repair of one unit is completed and another waits for repair [4]. From the above viewpoints, we introduce two types of Markov renewal processes with nonregeneration points and analyze an n-unit parallel and a two-unit standby systems with repair. Suppose that shocks occur in a stochastic process and each amount of damage due to shocks is additive. This process forms a cumulative process that is also called a jump process [7] or doubly stochastic process [8]. The analysis of such processes are more difficult than the already defined stochastic processes because the process is made up of two dependent stochastic processes. In Chap. 6, we define a cumulative process and investigate its properties. There are many examples of shock models, cumulative damage models, and independent damage models, and furthermore, their optimum maintenance policies. State space of the above stochastic processes may be finite because they are usually defined as the number of operating or failed units and the degree of wear for most reliability models. As far as reliability applications are concerned, we consider the processes with only a finite number of states. A continuous-time Markov process with continuous state space forms a Brownian motion, a Wiener process, and a Lévy process. This is now one of the most useful stochastic processes in applied sciences such as physics, economics, communication theory, biology, management science, and mathematical statistics. However, such processes cannot be much applicable to reliability models because of continuous state space. In Chap. 7, we make only the definitions of a Brownian motion and a Lévy
1.2 Stochastic Processes
5
process. As examples, we give continuous wear models in which wear increases continuously with time and consider their replacement policies. Finally, in Chap. 8, as examples of some redundant systems, we show how to analyze them by ready and handy tools of the stochastic processes. This would be greatly helpful for understanding the stochastic processes and learning reliability theory.
1.3 Further Studies This book is written mainly based on the books [9–11] for learning basic stochastic processes. In similar contents of such books, we recommend to read the books [12, 13] for a deeper understanding. There were contained many examples of stochastic processes in various fields of applications [14, 15] and in reliability systems [16, 17]. The names of stochastic processes such as stationary process, Gaussian process, martingales, branching process, and so on, appeared [18]. For more applications of stochastic processes, there were [19–21] for queuing theory, [22] for insurance and finance, and [20] for computer applications. It would be impossible to form stochastic models and to analyze them for more complex systems with too many constrains. Then, simulation would be a useful tool for analyzing such systems. We do not touch on the simulation at all in this book and recommend to master its techniques by [21, 23, 24]. We introduce simply the Brownian motion and Lévy process because these processes are not very applicable to reliability models and are more difficult to understand over our mathematical level. For further studies of such processes, refer to [25–27]. This book would be suited well to learn elementary stochastic processes for students with a major of reliability, and researchers and workers engaged in reliability engineering and management, with reference to books [6, 28–30]. As other text books, refer to recent published books [31–33]. For more further and recent studies of reliability, edited books and hand books [34–36] would be advisable to look through. In Appendix A and B, we present Laplace and LaplaceStieltjes transforms and make answers to selected problems.
1.4 Problems 1 1.1 Search for examples of stochastic processes on your daily life, or read the books [14–17] and give some examples. 1.2 Examine why there is no generation point for a two-unit parallel system with one repair person where the failure and repair times are not exponential [4]. What about the same system with two repair persons?
6
1 Introduction
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24.
25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36.
Klein JP, Moeschberger ML (2003) Survival analysis. Springer, New York Ushakov IA (1994) Handbook of reliability engineering. Wiley, New York Birolini A (1999) Reliability engineering theory and practice. Springer, New York Nakagawa T (2002) Two-unit redundant models. In: Osaki S (eds) Stochastic models in reliability and maintenance. Springer, Berlin, pp 165–185 Lala PK (1985) Fault tolerant and fault testable hardware design. Prentice-Hall, London Nakagawa T (2008) Advanced reliability models and maintenance policies. Springer, London Abdel-Hameed M (1984) Life distribution properties of devices subject to a pure jump damage process. J Appl Probab 21:816–825 Grandell J (1976) Doubly stochastic poisson process. Lecture notes in mathematics 529. Springer, New York Ross SM (1983) Stochastic processes. Wiley, New York Osaki S (1992) Applied stochastic system modeling. Springer, Berlin Çinlar E (1975) Introduction to stochastic processes. Prentice-Hall, Englewood Cliffs Karlin S, Taylor HM (1975) A first course in stochastic processes. Academic Press, New York Tijms HC (2003) A first course in stochastic models. Wiley, Chichester Ross SM (2000) Introduction to probability models. Academic Press, San Diego Bhat UN, Miller GK (2002) Elements of applied stochastic processes. Wiley, Hoboken, NJ Beichelt FE, Fatti LP (2002) Stochastic processes and their applications. CRC Press, Boca Raton Beichelt FE (2006) Stochastic processes in science, engineering and finance. Chapman & Hall, Boca Raton Prabhu NU (2007) Stochastic processes. World Scientific, Singapore Bhat UN, Basawa IV (1992) Queeing and related models. Oxford University Press, Oxford Trivedi KS (2002) Probabilty and statistics with reliability, queueing and computer science applications. Wiley, New York Stewart WJ (2009) Probability, Markov chains, queues, and simulation. Princeton University Press, Princeton Rolski T, Schmidli H, Schmidt V, Teugels J (1999) Stochastic processes for insurance and finance. Wiley, Chichester Ross SM (1997) Simulation. Academic Press, New York Papadopoulos C, Limnious N (2002) Quick Monte Carlo methods in stochastic systems and reliability. In: Osaki S (eds) Stochastic models in reliability and maintenance. Springer, Berlin, pp 307–333 Karatzas I, Shreve S (1991) Brownian motion and stochastic calculus. Springer, New York Bertoin J (1996) Lévy processes. Cambridge University Press, Cambridge Barndorff-Nielsen OE, Mikosch T, Resnick SI (eds) (2001) Lévy processes: theory and applications. Birkhävser, Boston Barlow RE, Proschan F (1965) Mathematical theory of reliability. Wiley, New York Nakagawa T (2005) Maintenance theory of reliability. Springer, London Nakagawa T (2007) Shock and damage models in reliability theory. Springer, London Dhillon BS (2002) Applied reliability and quality. Springer, London O’Connor PDT (2002) Practical reliability engineering. Wiley, Chichester Rausand M, H/yland A (2004) System reliability theory. Wiley, Hoboken Osaki S (eds) (2002) Stochastic models in reliability and maintenance. Springer, Berlin Pham H (eds) (2003) Handbook of reliability engineering. Springer, London Misra KB (2008) Handbook of performability engineering. Springer, London
Chapter 2
Poisson Processes
It is well-known that most units operating in a useful life period, and complex systems that consist of many kinds of components, fail normally due to random causes independently over the time interval. Then, it is said in technical terms of stochastic processes, that failures occur in a Poisson process that counts the number of failures through time. This is a natural modeling tool in reliability problems. Some reliability measures such as MTTF (Mean Time To Failure), availability, and failure rate are estimated statistically from life data and are in practical use under such modelings without much theoretical arguments. Furthermore, because a Poisson process has stationary and independent properties, it is much convenient for formulating stochastic models in mathematical reliability theory. A Poisson process is the simplest stochastic process that arises in many applications for arrival processes. It is of great importance to know the properties of a Poisson process and to learn how to apply the process to reliability models in the beginning: In Sect. 2.1, we investigate properties of an exponential distribution, because most units are assumed practically to have its distribution as their failure times. Then, it is shown that the number of failures has a Poisson distribution when the failure time has an exponential distribution. As applied examples, we discuss about the number of spare units for constructing parallel and standby systems that should be previously prepared. In Sect. 2.2, a Poisson process is defined theoretically, based on such above facts, and its interarrival times are shown to have an exponential distribution. Furthermore, when the failure was detected at time t; its failure occurs uniformly over ½0; t: As such applied examples, we take up an inspection and two backup models, and derive their optimum policies. It has been assumed often in maintenance models that a unit undergoes minimal repair at failures. This means that the failure rate is not changed by any maintenance, and forms a nonhomogeneous Poisson process. Underlying the process, failures occur successively, depending on the operating time t: In Sect. 2.3, we summarize the properties of the process. In Sect. 2.4, we apply these results to the replacement model where the unit is replaced at the Nth failure and the software
T. Nakagawa, Stochastic Processes, Springer Series in Reliability Engineering, DOI: 10.1007/978-0-85729-274-2_2, Ó Springer-Verlag London Limited 2011
7
8
2 Poisson Processes
reliability model, and investigate its properties. Finally, in Sect. 2.5, we consider a cumulative damage model where shocks occur in a Poisson process and the total damage due to shocks are additive. This is called a compound Poisson process. We survey probabilistic properties of the process and apply damage models with a failure level K where the unit fails when the total damage exceeds K: For a more deep study of Poisson processes, refer to the book [1].
2.1 Exponential Distribution Suppose that a unit begins to operate at time 0 and is replaced with a new identical one immediately at each failure, where the time required for replacement is negligible. The successive operating units have failure P times Xk ðk ¼ 1; 2; . . .Þ from the beginning of their operation to failure and Sn nk¼1 Xk in Fig. 2.1, where the horizontal axis represents the time-axis t from t ¼ 0 to the right direction in all figures of this book. It is assumed that Xk are independent and identically distributed random variables, and have an identical exponential distribution FðtÞ PrfXk tg ¼ 1 ekt ;
ð2:1Þ
and its density function f ðtÞ
dFðtÞ ¼ F 0 ðtÞ ¼ kekt dt
for k [ 0 and any k; that decreases strictly from k to 0. The mean and variance are EfXk g ¼
Z1
1 tkekt dt ¼ ; k
0
VfXk g ¼
Z1
ð2:2Þ 1 t2 kekt dt ðEfXk gÞ2 ¼ 2 : k
0
Then, it is said that the unit has a failure time X; omitting subscript k; and fails according to an exponential distribution ð1 ekt Þ with parameter k or MTTF 1=k: Furthermore, from Appendix A, the LS (Laplace-Stieltjes) transform of an exponential distribution is
Fig. 2.1 Process of failures
2.1 Exponential Distribution
9
sX
F ðsÞ Efe
g¼
Z1 e
st
dFðtÞ ¼
0
Z1
est kekt dt ¼
k sþk
ð2:3Þ
0
for any ReðsÞ [ 0; where note that we omit ReðsÞ [ 0 throughout this book. From the assumptions that random variables are independent and identical, n EfX1 þ X2 þ þ Xn g ¼ nEfXg ¼ ; k n VfX1 þ X2 þ þ Xn g ¼ nVfXg ¼ 2 ; k n k sðX1 þX2 þþXn Þ sX n g ¼ Efe g ¼ : Efe sþk
ð2:4Þ
If two random variables X1 and X2 are independent and have the respective exponential distributions with mean 1=k1 and 1=k2 ; then PrfX1 X2 g ¼
Z1
ek2 t k1 ek1 t dt ¼
k1 : k1 þ k2
0
In general, if k random variables Xk are independent and have an exponential distribution with mean 1=kk ðk ¼ 1; 2; . . .Þ; and Xm ¼ minfX1 ; X2 ; . . .; Xn g; then PrfX1 ¼ Xm g ¼
Z1
eðk2 þk3 þþkn Þt k1 ek1 t dt ¼
k1 : k1 þ k2 þ þ kn
0
2.1.1 Properties of Exponential Distribution We investigate and summarize various properties of an exponential distribution ð1 ekt Þ appearing frequently in reliability theory: (1) From the MTTF or the mean life time of the unit 1=k; Fð1=kÞ ¼ 1 e1 0:632; i.e., 63.2% of units have failed at the time of MTTF (Example 2.13). (2) The reliability function RðtÞ FðtÞ ¼ 1 FðtÞ that is the probability that the unit is operating at time t without failure is RðtÞ ¼ ekt : Thus, the failure rate or hazard rate is [2, p. 5, 3, p. 22]
10
2 Poisson Processes
hðtÞ
f ðtÞ kekt ¼ kt ¼ k; e RðtÞ
irrespective of time t: Thus, if the failure rate would be estimated from actual life data, then the MTTF would be obtained by taking the reciprocal of k: For example, when k ¼ 104 (1/h), the MTTF is 104 h and vice versa. Therefore, because hðtÞDt Prft\X t þ DtjX [ tg means the probability that the unit with age t will fail in ðt; t þ Dt for small Dt; the probability that a failure occurs in a small interval ðt; t þ Dt is constant for any time t: That is, the failure rate of an exponential distribution has the property of random failures occurring constantly in a small interval time t; irrespective of ages or elapsed operating times. In general, the failure rate curve, i.e., the bathtub curve [4, p. 188] (Fig. 2.5) is divided into three periods; (1) the initial failure period in which hðtÞ is DFR (Decreasing Failure Rate), (2) the random failure rate period in which hðtÞ is CFR (Constant Failure Rate), and (3) the wearout failure period in which hðtÞ is IFR (Increasing Failure Rate). In this case, an exponential distribution has the properties of DFR and IFR, and shows the failure phenomenon observed in a random failure period. Thus, when units have an exponential failure distribution, there is no necessity for making any maintenance of them theoretically. For such reasons, the exponential distribution might be of limited applications. However, the exponential distribution plays a fitting role for complex systems that consist of many kinds of components, each of which has a variety pattern of malfunction and replacement and makes the failure pattern of the system as a whole [2, p. 18]. It is well-known that the exponential distribution is fitted well to life data of electric and electronic components. (3) Using the Taylor expansion of an exponential function, ðkhÞ2 ðkhÞ3 þ ¼ kh þ oðhÞ; 2! 3! where oðhÞ is an infinite decimal higher than h; i.e., oðhÞ=h ! 0 as h ! 0: Thus, we have the inequalities ðkhÞ2 \1 ekh \kh kh 2 1 ekh ¼ kh
for a small kh [ 0: These inequalities are very useful for computing equations numerically, involving exponential functions because kt is very small in actual reliability models, as shown in Table 2.1. Table 2.1 Comparative table of 1 ekh and approximation kh ðkhÞ2 =2 for kh kh 1 ekh kh ðkhÞ2 =2 1 101 102 103 104
0.6321205 0:951625 101 0:99501 102 0:99949 103 0:9999 104
0.5 0:95 101 0:995 102 0:9995 103 0:9999 104
2.1 Exponential Distribution
11
Example 2.1 (MTTF of exponential distribution) When the failure distribution is an exponential distribution ð1 ekt Þ; we ask the following questions: (a) How much is the failure rate k and the MTTF 1=k when RðtÞ ¼ 0:99 at t ¼ 103 h? (b) What fraction time t of the MTTF are to satisfy that FðtÞ ¼ 101 ; 102 ; 103 ; and 104 ? The solution of (a) is easily given as 3
ek 10 ¼ 0:99; i.e., k ¼ 103 log 0:99 1:005 105 (1/h); 1 MTTF ¼ ¼ 99;499 (h): k We can compute approximately, k 103 ¼ 1 0:99 ¼ 0:01; i.e., e k ¼ 105 (1/h);
f ¼ 105 (h): M TTF
In this case, the difference between the exact and approximate values are very small. The solution of (b) is given by computing kt ¼
t t ¼ : 1=k MTTF
For example, 1 ekt ¼ 0:1; kt ¼ log 0:9 ¼ 0:10536: Approximately, from (3), e ¼ 0:1 ¼ 1=10: The time t that satisfies FðtÞ ¼ 0:1 is also called a B10 life. kt e for FðtÞ ¼ 101 ; 102 ; 103 ; Table 2.2 presents kt and their approximate values kt 4 and 10 : This table indicates that when FðtÞ ¼ 10n ðn ¼ 1; 2; . . .Þ; the fraction time t of the MTTF is approximately 10n ¼ 1=10n ; and its value becomes very good as n becomes large. From such results, we can easily estimate the fraction time of the MTTF for a small FðtÞ by setting t=MTTF ¼ FðtÞ: Example 2.2 (Preventive maintenance) [3, p. 8] Consider operating units such as scales, computers, and production systems where its preventive maintenance is
e for FðtÞ Table 2.2 Comparative table of kt and approximation kt FðtÞ kt
e kt
101 102 103 104
101 102 103 104
1:0536 101 1:0050 102 1:0005 103 1:0000 104
12
2 Poisson Processes
done only at time T ð0\T 1Þ and its failure is detected. It is supposed that the unit has a failure distribution FðtÞ. We get some earning per unit of time and does not have any earning during the interval if it fails before time T: The average time during ½0; T in which we get some earning is lðTÞ ¼ 0 FðTÞ þ TFðTÞ ¼ TFðTÞ;
ð2:5Þ
and lð0Þ ¼ lð1Þ ¼ 0: Differentiating lðTÞ with respect to T and setting it equal to zero, FðTÞ Tf ðTÞ ¼ 0;
i.e.;
hðTÞ ¼
f ðTÞ 1 ¼ : FðTÞ T
ð2:6Þ
Thus, an optimum time T that maximizes lðTÞ in (2.5) is given by a unique solution of (2.6), when FðtÞ is IFR. In particular, when FðtÞ ¼ 1 ekt ; T ¼ 1=k; i.e., we should do the preventive maintenance at the MTTF. Example 2.3 (Discount cost) [5, p. 82] When we consider the total expected cost of an operating unit for an infinite time span, we should evaluate the present values of all replacement costs by using an appropriate discount rate [3, p. 78]. Suppose that a continuous discount with rate a ð0\a\1Þ is used for the cost incurred at replacement time. That is, the present value of replacement cost c at at at time 0. Then, because the total replacement cost is time P1 t is ce c n¼1 exp½aðX1 þ X2 þ þ Xn Þ; the expected cost is Cc
1 X
Efexp½aðX1 þ X2 þ þ Xn Þg ¼ c
n¼1
1 X
n EfeaX g :
n¼1
Thus, when X is distributed exponentially, from (2.4), n n 1 Z 1 1 X X k ck at kt e ke dt ¼ c ¼ : C¼c a þ k a 0 n¼1 n¼1 For example, if the mean failure time is 1,000 h, the replacement cost is 20 dollars, and the interest rate is 5 % per year, i.e., k ¼ 1=1; 000; c ¼ 20; and a ¼ 0:05=ð365 24Þ; then C ¼ 20 365 24=ð1000 0:05Þ ¼ 3; 504 dollars. In general, for any negative function gðtÞ for t 0 (Problem 2.1), ( ) Z1 1 X gðSn Þ ¼ k gðtÞdt: E n¼1
0
In this example, gðtÞ ¼ ceat : (4) From the assumption that X is a random variable with an exponential distribution ð1 ekt Þ; the conditional survival probability that the unit does not fail in a finite interval time t; given that it is operating at time u ðu 0Þ is [4, p. 74]
2.1 Exponential Distribution
13
PrfX [ u þ tjX [ ug ¼
PrfX [ u þ tg ekðuþtÞ ¼ ku e PrfX [ ug
¼ ekt ¼ PrfX [ tg; that is independent of any time u; i.e., for all t; u 0; PrfX [ u þ tg ¼ PrfX [ ug PrfX [ tg:
ð2:7Þ
The condition that the unit is operating at time u; i.e., its age is u; has no effect on the next operating time t; i.e., its residual or remaining time. This is called the memoryless property or lack of memory property. Conversely, the only continuous and positive function that satisfies (2.7) is an exponential function PrfX [ tg ¼ ekt (Problem 2.2). That is, only the exponential distribution has the memoryless property. For such reasons, as one of failure distributions, an exponential distribution comes to be frequently used for forming simpler and analyzing easier stochastic models. Example 2.4 (Parallel system) Consider a parallel system that consists of n identical units and fails when all units have failed, i.e., at least one of n units is operating, the system is also operating [6, p. 8]. Each unit has an independent and identical failure distribution ð1 ekt Þ: Because the system with n units has a failure distribution ð1 ekt Þn ; its MTTF is (Problems 2.3–2.8) Z1 Z1 n 1 1 xn 1X 1 kt n 1 ð1 e Þ dt ¼ dx ¼ : ð2:8Þ 1x k k k¼1 k 0
0
On the other hand, let XðkÞ ðk ¼ 1; 2; . . .; nÞ be the failure time from the ðk 1Þth to kth unit of n units. The failure time of the system is given by Xð1Þ þ Xð2Þ þ XðnÞ : Then, because XðkÞ has an exponential distribution 1 ekðnkþ1Þt with mean 1=½ðn k þ 1Þk from the memoryless property, the MTTF is n n 1X 1 1X 1 ¼ ; EfXð1Þ þ Xð2Þ þ þ XðnÞ g ¼ k k¼1 n k þ 1 k k¼1 k that agrees with (2.8), and it is called a harmonic series (Problem 2.8). We have the inequalities [7] n 1 X 1 log n þ \ \ log n þ 1 ðn ¼ 2; 3; . . .Þ; n k¼1 k and
n X 1 k¼1
k
ð2:9Þ
C þ log n;
where C is Euler’s constant and C 0:577215. . .: From such results, the MTTF of a parallel system increases with n very slowly and logarithmically to infinity, and is given approximately by (2.9) for large n (Problem 2.5).
14
2 Poisson Processes
Next, because the reliability function is n Rn ðtÞ ¼ 1 1 ekt ;
ð2:10Þ
its failure density function is n1 fn ðtÞ ¼ R0n ðtÞ ¼ nkekt 1 ekt : Hence, n n1 k Rn ðtÞ 1 1 1 ekt 1 X ¼ ¼ 1 ekt ; n1 fn ðtÞ nk ekt ð1 ekt Þ nk k¼0 that decreases with t: Thus, the failure rate hn ðtÞ ¼
fn ðtÞ nk ¼ Pn1 kt Þk Rn ðtÞ k¼0 ð1 e
ð2:11Þ
increases strictly with t from 0 to k ðn 2Þ that is the failure rate of each unit. Furthermore, because 1 hnþ1 ðtÞ
n1 h X n k i 1 1 1 ekt 1 ekt ¼ [ 0; hn ðtÞ nðn þ 1Þk k¼0
the failure rate hn ðtÞ decreases strictly with n ðn 1Þ from k to 0 for a given t [ 0:
2.1.2 Poisson and Gamma Distributions We are often interested in the number of failures and replacements of units that have occurred in the interval ½0; t: For this purpose, we introduce the total sum of random variables Sn X1 þ X2 þ þ Xn ðn ¼ 1; 2; . . .Þ that represents the time of the nth failure (Fig. 2.1) when FðtÞ PrfXk tg ¼ 1 ekt for all k 1: Then, we derive the distribution PrfSn tg ¼
Zt 0
1 X kðkuÞn1 ku ðktÞk kt e du ¼ e ðn 1Þ! k! k¼n
¼1
n1 X ðktÞk k¼0
k!
ekt
ðn ¼ 1; 2; . . .Þ;
ð2:12Þ
that is called an Erlang distribution or a gamma distribution with order n: We prove (2.12) by the mathematical induction as follows: Using the total sum of independent random variables with an exponential distribution 1 ekt ;
2.1 Exponential Distribution
15
PrfS2 tg ¼ PrfX1 þ X2 tg ¼
Zt h
i 1 ekðtuÞ keku du
0
¼ 1 ð1 þ ktÞe
kt
:
Assuming that when n ¼ k; PrfSk tg ¼
Zt
kðkuÞk1 ku e du; ðk 1Þ!
0
we easily have Zt h ikðkuÞk1 1 ekðtuÞ PrfSkþ1 tg ¼ PrfSk þ Xkþ1 tg ¼ eku du ðk 1Þ! 0
¼
1 X ðktÞj
j!
j¼k
1 X ðktÞ kt ðktÞj kt e ¼ e ; k! j! j¼kþ1 k
ekt
that concludes that (2.12) holds for all n 1: Equation (2.12) is alternatively proved by the technique of the LS transform in Appendix A: Recalling that the LS transform of a random variable Xk with an exponential distribution 1 ekt is given in (2.3), the LS transform of Sn is n k : EfesSn g ¼ EfesðX1 þX2 þþXn Þ g ¼ sþk On the other hand, the LS transform of (2.12) is Z1
est
kðktÞn1 kt e dt ¼ ðn 1Þ!
k sþk
n :
0
Therefore, because an original function is in one-to-one correspondence with its LS transform, it is shown that Sn has a gamma distribution in (2.12), and its mean and variance are also given in (2.4). A gamma distribution with order n represents the distribution of the nth failure of a unit with an exponential distribution. If a unit fails every at the nth minor fault, each of which occurs exponentially, its failure time has a gamma distribution. In other words, a standby redundant system (Example 2.6) with n units fail according to a gamma distribution with order n: A density function of Sn given in (2.12) is generalized as gðtÞ
kðktÞa1 kt e CðaÞ
k; a [ 0;
ð2:13Þ
16
2 Poisson Processes
R1 where CðaÞ 0 xa1 ex dx for a [ 0; and is called a gamma function with order a: The mean, variance, and LS transform of a random variable X with a gamma density function in (2.13) are, respectively, a a a k sX ; ð2:14Þ EfXg ¼ ; VfXg ¼ 2 ; Efe g ¼ k sþk k and the failure rate is hðtÞ ¼ R 1 t
ðktÞa1 ekt ðkuÞa1 eku du
:
ð2:15Þ
Thus, 1 ¼ hðtÞ
a1 Z1 Z1 a1 u x kðutÞ 1þ e du ¼ ekx dx; t t t
0
that implies that hðtÞ decreases from 1 to k for 0\a\1; is constant k for a ¼ 1; and increases from 0 to k for a [ 1: We define NðtÞ maxn fSn tg that represents the number of failures in ½0; t; where fNðtÞg is called a counting process, that is one of stochastic processes. If S1 ¼ X1 [ t; then NðtÞ ¼ 0; i.e., no failure occurs in ½0; t: Then, we have the relation fSn tg , fNðtÞ ng;
ð2:16Þ
because if the nth failure occurs up to time t; then the number of failures in time t is greater than n; and conversely, if the number of failures in time t is greater than n; then the nth failure occurs up to time t (Fig. 2.1). Thus, from (2.12) and (2.16), PrfNðtÞ ng ¼ PrfSn tg ¼
Zt 0
1 X kðkuÞn1 ku ðktÞk kt e du ¼ e : ðn 1Þ! k! k¼n
Hence, PrfNðtÞ ¼ ng ¼ PrfSn tg PrfSnþ1 tg ¼
ðktÞn kt e n!
ðn ¼ 0; 1; 2; . . .Þ;
ð2:17Þ
and its mean and variance are 1 X ðktÞn kt n e ¼ kt; n! n¼1 1 X ðktÞn kt VfNðtÞg ¼ e ðktÞ2 ¼ kt: n2 n! n¼1
EfNðtÞg ¼
ð2:18Þ
2.1 Exponential Distribution
17
Equation (2.17) is called a Poisson distribution with mean kt that was originally derived as a limiting distribution of a binomial distribution by S.O. Poisson (1873) (Problem 2.9). The number of defects due to random causes normally in some time, area, volume, and so on, has a Poisson distribution. Example 2.5 (Poisson distribution) A unit fails exponentially at a failure rate k ¼ 103 (1/h). Then, calculate the probabilities that k ðk ¼ 0; 1; 2Þ failures occur in 500 h. In this case, because kt ¼ 103 500 ¼ 0:5; PrfNð500Þ ¼ 0g ¼ e0:5 0:607; PrfNð500Þ ¼ 1g ¼ 0:5e0:5 0:303; PrfNð500Þ ¼ 2g ¼
ð0:5Þ2 0:5 0:076: e 2!
Furthermore, the probability that the number of failures is equal to or more than 3 in 500 h is PrfNð500Þ 3g ¼ 1 PrfNð500Þ 2g 0:25 0:5 ¼ 1 1 þ 0:5 þ 0:0144: e 2 This probability is very small because the mean failure time 1=k ¼ 103 h is two times of t ¼ 500: Example 2.6 (Scheduling problem) Consider a standby system where an operating unit is replaced with one of identical spares at each failure and each unit has an exponential failure distribution 1 ekt : We are interested in how many number of spare units should be provided initially in order to assure with probability a ð0\a\1Þ that the system will remain operating in time t [2, p. 49]. An optimum number n for fixed a and t is given by a unique minimum that satisfies PrfNðtÞ ng ¼
n X ðktÞk k¼0
k!
ekt a:
Next, consider the scheduling problem where the system has to work for a job with working time S and its distribution WðtÞ PrfS tg [6, p. 84] and WðtÞ 1 WðtÞ: It is assumed that if the work of a job is accomplished when at least one unit is operating, it needs cost c2 ; if the work is not accomplished when all units have failed, it needs cost c1 with c1 [ c2 ; and the cost of a standby system with n units is c0 n: Then, the expected cost is CðnÞ ¼ c2 þ ðc1 c2 Þ
Z1 0
WðtÞdF ðnÞ ðtÞ þ c0 n ðn ¼ 0; 1; 2; . . .Þ:
ð2:19Þ
18
2 Poisson Processes
Table 2.3 Optimum numbers n of units for a standby system k=w c0 =ðc1 c2 Þ 0.5 0.3 0.1 0.05 0.01
1.0
2.0
5.0
0 1 3 4 6
0 1 3 5 9
0 0 3 7 16
In particular, when WðtÞ ¼ 1 ewt ; CðnÞ ¼ c2 þ ðc1 c2 Þ½F ðwÞn þc0 n:
ð2:20Þ
To find an optimum number n that minimizes CðnÞ; from the inequality Cðn þ 1Þ CðnÞ 0; ½F ðwÞn ½1 F ðwÞ
c0 c 1 c2
ðn ¼ 0; 1; 2; . . .Þ;
ð2:21Þ
whose left-hand decreases strictly from 1 F ðwÞ to 0. Thus, if F ðwÞ\ðc1 c2 c0 Þ=ðc1 c2 Þ; then there exists a finite and unique minimum n ð1 n \1Þ that satisfies (2.21). If F ðwÞ ðc1 c2 c0 Þ=ðc1 c2 Þ; then n ¼ 0; i.e., we should not provide any units for a job. In addition, when FðtÞ ¼ 1 ekt ; (2.21) is wkn ðw þ kÞ
nþ1
c0 c1 c2
ðn ¼ 0; 1; 2; . . .Þ:
ð2:22Þ
If w=ðw þ kÞ [ c0 =ðc1 c2 Þ; then a positive n exists. Table 2.3. presents the optimum n that decreases with c0 =ðc1 c2 Þ (Problem 2.10).
2.2 Poisson Process We consider a stochastic process fXðtÞ; t 2 Tg in which XðtÞ changes randomly with time t 2 T in a space by the probabilistic law, i.e., XðtÞ is a random variable for each t in time set and fXðtÞ; t 2 Tg is its collection. We restrict ourselves to a nonnegative time parameter t for the time set T: If the time set is countable, i.e., T ¼ f0; 1; 2; . . .g; the process fXðkÞ; k ¼ 0; 1; 2; . . .g is called a discrete-time stochastic process, and if the time set is continuous, i.e., T ¼ Set of fa nonnegative real numberg; the process fXðtÞ; t 0g is called a continuous-time stochastic process. If a random variable XðtÞ presents the working condition of a unit at every time and at every day or week, the time set is continuous and discrete, respectively. We omit fk ¼ 0; 1; 2; . . .g and ft 0g for simplicity unless otherwise stated. A set of all possible values of XðtÞ for the process fXðtÞg is called a state space. We restrict ourselves to a discrete state space, where XðtÞ ¼ 0; 1; 2; . . .; except for the cumulative processes in Chap. 6, and Brownian motion and Lévy process in
2.2 Poisson Process
19
Chap. 7. When the state space in such stochastic processes is continuous, we need more mathematical tools for their analysis. If the process fXðtÞg is observed over time t; such realization of XðtÞ is called a sample path or sample function. Furthermore, we introduce the following two properties of a stochastic process: It is said that the difference Xðtk Þ Xðtk1 Þ is called the increment, and XðtÞ has independent increments if Xðt1 Þ Xðt0 Þ; Xðt2 Þ Xðt1 Þ; . . .; Xðtn Þ Xðtn1 Þ are independent for all 0 t0 \t1 \ \tn1 \tn ; where it is called that tk is an arrival time of the kth event and tk tk1 is an interarrival time. In addition, the process fXðtÞg is said to have stationary or homogeneous increments if Xðt þ uÞ XðtÞ has the same distribution for all t 0: If XðtÞ has independent and stationary increments, then XðtÞ and Xðt þ uÞ XðtÞ are independent, and Xðt þ uÞ XðtÞ has the same distribution as XðuÞ: Thus, we have the relation EfXðt þ uÞg ¼ EfXðt þ uÞ XðtÞ þ XðtÞg ¼ EfXðuÞg þ EfXðtÞg: The properties of stationary and independent increments in reliability mean intuitively that any failures occur independently and stationarity in ½t; t þ u; being dependent only on time interval u irrespective of the beginning time t and other intervals. For example, when we consider two intervals ½t1 ; t1 þ u and ½t2 ; t2 þ u for t1 þ u\t2 ; failures occur independently in two intervals and the probability of failures in two intervals is equal to that in ½0; u: In general, the property of independent increment is more important than that of stationary increment for most reliability models, because they might be easy to fail with their age. We consider the stochastic process fNðtÞg; where NðtÞ ¼ 0; 1; 2; . . .; that is called a counting process fNðtÞg; because NðtÞ counts the total number of events occurred in time t: Definition 2.1 If (i) (ii) (iii) (iv)
Nð0Þ ¼ 0; the process fNðtÞg has stationary and independent increments, PrfNðhÞ ¼ 1g ¼ kh þ oðhÞ; PrfNðhÞ 2g ¼ oðhÞ;
then a counting process fNðtÞg is said to be a Poisson process with rate k [ 0: Using this definition, we derive the probability Pk ðtÞ PrfNðtÞ ¼ kjNð0Þ ¼ 0g ¼ PrfNðtÞ ¼ kg ðk ¼ 0; 1; 2; . . .Þ; ð2:23Þ P1 that represents the probability that k events occur in ½0; t; and k¼0 Pk ðtÞ ¼ 1: From (ii), (iii), and (iv), 8 < 1 kh þ oðhÞ for k ¼ j; PrfNðt þ hÞ ¼ k j NðtÞ ¼ jg ¼ kh þ oðhÞ for k ¼ j þ 1; ð2:24Þ : oðhÞ for k [ j þ 1:
20
2 Poisson Processes
Thus, from (ii) and (2.24), P0 ðt þ hÞ ¼ PrfNðt þ hÞ ¼ 0g ¼ PrfNðtÞ ¼ 0; Nðt þ hÞ NðtÞ ¼ 0g ¼ PrfNðtÞ ¼ 0g PrfNðt þ hÞ NðtÞ ¼ 0g ¼ PrfNðtÞ ¼ 0g PrfNðhÞ ¼ 0g ¼ P0 ðtÞ½1 kh þ oðhÞ: Hence, P0 ðt þ hÞ P0 ðtÞ oðhÞ ¼ kP0 ðtÞ þ : h h Letting h ! 0; we have the differential equation dP0 ðtÞ ¼ kP0 ðtÞ; dt
i.e.;
dP0 ðtÞ=dt ¼ k: P0 ðtÞ
By integration and setting P0 ð0Þ ¼ 1 from (i), P0 ðtÞ ¼ ekt :
ð2:25Þ
Similarly, from (2.24), Pk ðt þ hÞ ¼ PrfNðt þ hÞ ¼ kg ¼ PrfNðtÞ ¼ k 1; Nðt þ hÞ NðtÞ ¼ 1g þ PrfNðtÞ ¼ k; Nðt þ hÞ NðtÞ ¼ 0g þ
k X
PrfNðtÞ ¼ k j; Nðt þ hÞ NðtÞ ¼ jg
j¼2
¼ Pk1 ðtÞ½kh þ oðhÞ þ Pk ðtÞ½1 kh þ oðhÞ þ oðhÞ: Hence, Pk ðt þ hÞ Pk ðtÞ oðhÞ ¼ kPk1 ðtÞ kPk ðtÞ þ : h h Letting h ! 0; dPk ðtÞ ¼ kPk1 ðtÞ kPk ðtÞ; dt i.e., e
kt
P0k ðtÞ
d ekt Pk ðtÞ þ kPk ðtÞ ¼ ¼ kekt Pk1 ðtÞ: dt
Recalling that P0 ðtÞ ¼ ekt in (2.25), when k ¼ 1; d½ekt P1 ðtÞ ¼ k: dt
ð2:26Þ
2.2 Poisson Process
21
By integration under the initial condition of P1 ð0Þ ¼ 0; P1 ðtÞ ¼ ktekt : h i Using the mathematical induction by setting Pk1 ðtÞ ¼ ðktÞk1 =ðk 1Þ! ekt in (2.26), d ekt Pk ðtÞ kðktÞk1 : ¼ ðk 1Þ! dt Similarly, by integration under Pk ð0Þ ¼ 0; Pk ðtÞ ¼
ðktÞk kt e k!
ðk ¼ 0; 1; 2; . . .Þ;
ð2:27Þ
that corresponds to a Poisson distribution in (2.17). Conversely, when a counting process NðtÞ has a Poisson distribution in (2.27), and identically distributed the interarrival times Xk ðk ¼ 1; 2; . . .Þ are independent kt [4, p. 69]: It is easily shown according to an exponential distribution 1 e that from (2.16), PrfX1 tg ¼ PrfS1 tg ¼ PrfNðtÞ 1g ¼ 1 PrfNðtÞ ¼ 0g ¼ 1 ekt ; i.e., X1 has an exponential distribution 1 ekt : Similarly, PrfX2 tjX1 ¼ ug ¼ PrfNðt þ uÞ NðuÞ 1jX1 ¼ ug ¼ 1 PrfNðt þ uÞ NðuÞ ¼ 0g ¼ 1 PrfNðtÞ ¼ 0g ¼ 1 ekt : Generally, PrfXnþ1 tjSn ¼ ug ¼ PrfNðt þ uÞ NðuÞ 1jSn ¼ ug ¼ 1 PrfNðt þ uÞ NðuÞ ¼ 0g ¼ 1 ekt : Therefore, each interarrival time Xk is independent and identically distributed exponentially in (2.1). Thus, a Poisson process has memoryless property. In addition, because an exponential density kekt decreases strictly, a Poisson process has shorter interarrival times than longer ones. From (2.16), the arrival time Sn of the nth event has a gamma distribution 1 X ðktÞk kt e : PrfSn tg ¼ k! k¼n Applying the stationary property of (ii) to (2.27), PrfNðt þ uÞ NðuÞ ¼ kjNð0Þ ¼ 0g ¼ PrfNðtÞ ¼ kg ¼ Pk ðtÞ for any t; u 0: Using this result, we can make an alternative definition of a Poisson process [4, p. 67, 8, p. 31]:
22
2 Poisson Processes
Definition 2.2 If (i) Nð0Þ ¼ 0; (ii) the process fNðtÞg has independent increments, (iii) the probability that k events occur in any interval t is PrfNðt þ uÞ NðuÞ ¼ kg ¼
ðktÞk kt e k!
ðk ¼ 0; 1; 2; . . .Þ
for any t; u 0; then a counting process fNðtÞg is a Poisson process. It can be easily seen that two definitions are equivalent: From the above results, Definition 2.1 implies Definition 2.2. It can be clearly verified that the process has stationary increments from (iii) of Definition 2.2. Furthermore, from (iii) and (3) of Sect. 2.1.1, PrfNðhÞ ¼ 1g ¼ khekh ¼ kh þ oðhÞ; 1 X ðkhÞk kh PrfNðhÞ 2g ¼ e ¼ oðhÞ; k! k¼2 that follows that Definition 2.2 implies Definition 2.1. A Poisson process involves all properties shown in Definitions 2.1 and 2.2. Example 2.7 (Poisson distribution) It has been estimated from the past life data that failures of some unit occur in a Poisson process and its expected number is twice the average per year, i.e., 365 24 ¼ 8; 760 h. Then, from kt ¼ 2; the MTTF is 1 t 8; 760 ¼ ¼ ¼ 4; 380 h: k 2 2 Thus, the unit has a failure exponential distribution 1 et=4380 and failures n o occur in a Poisson distribution ½t=4380k =k! et=4380 ðk ¼ 0; 1; 2; . . .Þ: The number of units that should be provided to assure with probability 0:90 to operate continuously during one year without replenishment is given by the minimum number that satisfies, from Example 2.6, n X 2k k¼0
k!
e2 0:9;
i.e., its number is 5, including the first operating unit. Example 2.8 (Superposition of Poisson process) [5, p. 87] Let N1 ðtÞ be the number of failed units from one factory arriving at a repair shop in ½0; t and N2 ðtÞ be the number of failed units from the other. If the arrival times from two factories are independent and have the respective Poisson processes with rates k1 and k2 ; the total number NðtÞ of failed units arriving at a repair shop has the probability
2.2 Poisson Process
23
PrfNðtÞ ¼ ng ¼ PrfN1 ðtÞ þ N2 ðtÞ ¼ ng ¼
n X
PrfN1 ðtÞ ¼ k; N2 ðtÞ ¼ n kg
k¼0
¼
n X ðk1 tÞk k¼0
k!
ek1 t
ðk2 tÞnk k2 t e ðn kÞ!
¼
k nk n ½ðk1 þ k2 Þtn ðk1 þk2 Þt X n! k1 k2 e n! k1 þ k2 k!ðn kÞ! k1 þ k2 k¼0
¼
½ðk1 þ k2 Þtn ðk1 þk2 Þt e : n!
Thus, the process fNðtÞg is also a Poisson process with rate k1 þ k2 (Problem 2.11). Example 2.9 (Decomposition of Poisson process) [5, p. 88] Let NðtÞ be the number of failed units occurring at a factory in ½0; t and be a Poisson process with rate k: Classifying into two large groups of failed units, the number N1 ðtÞ of minor ones occurs with probability p and the number N2 ðtÞ of major ones occurs with probability q 1 p; independent of N1 ðtÞ: Then, because the number N1 ðtÞ has a binomial distribution with p; given that n failures have occurred, the joint probability is PrfN1 ðtÞ ¼ k; N2 ðtÞ ¼ n kg ¼ PrfN1 ðtÞ ¼ k; N2 ðtÞ ¼ n kjNðtÞ ¼ ng PrfNðtÞ ¼ ng n k nk ðktÞn kt ðpktÞk pkt ðqktÞnk qkt e ¼ e e pq ¼ : n! k! ðn kÞ! k Thus, the two Poisson processes fN1 ðtÞg and fN2 ðtÞg are independent and have the Poisson processes with rates pk and qk; respectively (Problem 2.12). In general, when the total number NðtÞ of failed units with a Poisson process with rate k are classified into k groups of Nj ðtÞ ðj ¼ 1; 2; . . .; kÞ with probability pj P P where kj¼1 Nj ðtÞ ¼ NðtÞ and kj¼1 pj ¼ 1; the joint probability is PrfN1 ðtÞ ¼ n1 ; N2 ðtÞ ¼ n2 ; . . .; Nk ðtÞ ¼ nk g ðp1 ktÞn1 p1 kt ðp2 ktÞn2 p2 kt ðpk ktÞnk pk kt ¼ ... ; e e e n1 ! n2 ! nk !
ð2:28Þ
that is called a multi-Poisson process. Next, when events occur in a Poisson process, we obtain the distribution of the interarrival time X; given that there was an event in ½0; t [4, p.71, 8, p.36]. This probability is given by, for u t;
24
2 Poisson Processes
PrfX ujNðtÞ ¼ 1g ¼ ¼
PrfNðuÞ ¼ 1; NðtÞ NðuÞ ¼ 0g PrfNðtÞ ¼ 1g kueku ekðtuÞ u ¼ ; ktekt t
ð2:29Þ
that is a uniform distribution over ½0; t: That is, when the event was detected at time t; it occurs constantly over ½0; t: Example 2.10 (Binomial distribution) [9, p. 76] Let NðtÞ be a Poisson process with rate k: Then, from Definition 2.2, for 0 u t, PrfNðuÞ ¼ k; NðtÞ ¼ ng PrfNðtÞ ¼ ng PrfNðuÞ ¼ kg PrfNðtÞ NðuÞ ¼ n kg ¼ PrfNðtÞ ¼ ng n o k ku ½ðkuÞ =k!e ½kðt uÞnk =ðn kÞ! ekðtuÞ ¼ ½ðktÞn =n!ekt k n u u nk 1 ¼ ðk ¼ 0; 1; 2; . . .; nÞ; t t k
PrfNðuÞ ¼ kjNðtÞ ¼ ng ¼
ð2:30Þ
that is a binomial distribution with parameters n and u=t: A uniform distribution sometimes appears in the cases where events occur only in the finite interval and when an event was detected at time t; without any information of its arrival time, as shown in (2.29). Example 2.11 (Inspection policy) [2, p. 113] Consider an inspection policy where the failure time of a unit is uniformly distributed over ½0; S that is called a working interval [6, p. 96]. To detect a failure, the unit is checked at planned times Tk ðk ¼ 1; 2; . . .; NÞ; where TN ¼ S (Fig. 2.2). Let c1 be the cost of one check and c2 be the cost per unit of time for the time elapsed between a failure and its detection at the next checking time. Then, the expected cost until failure detection is Tkþ1 N 1 Z X dt ½ðk þ 1Þc1 þ c2 ðTkþ1 tÞ : ð2:31Þ C1 ðNÞ ¼ S k¼0 Tk
Differentiating C1 ðNÞ with respect to Tk and setting it equal to zero, c1 Tkþ1 Tk ¼ Tk Tk1 : c2 Fig. 2.2 Process of an inspection policy
2.2 Poisson Process
25
Solving for Tk ; Tk ¼ kT1
kðk 1Þ c1 : c2 2
Setting TN ¼ S; Tk ¼
kS c1 þ kðN kÞ 2c2 N
ðk ¼ 0; 1; 2; . . . ; NÞ:
From Tkþ1 Tk [ 0; S c1 ðN 2k 1Þ [ 0: þ N 2c2 When k ¼ N 1; NðN 1Þ c2 S ; \ 2 c1
NðN þ 1Þ c2 S : ð2:32Þ
2 c1 P Thus, N must be a unique minimum that satisfies Nk¼1 k c2 S=c1 : Note that the summation of integers from 1 to N plays an important rule in obtaining optimum policies for some partition models [6, p. 39] (Problem 2.13). For example, when S ¼ 100 and c1 =c2 ¼ 7; the checking number is N ¼ 5 and checking times are 34, 61, 81, 94, 100. Next, consider the backup policy with checkpoint times in which we execute the rollback operation to the latest checkpoint [6, p. 96] and reconstruct the consistent state. Let c2 be the cost per unit of time for the backup operation to the latest checking time Tk when the failure was detected between Tk and Tkþ1 (Fig. 2.3). The other assumptions are the same as those in the inspection policy. Then, the expected cost until the backup operation is done to the latest checking time when a unit has failed at time t is C2 ðNÞ ¼
i.e.;
Tkþ1 N1 Z X k¼0
dt ½kc1 þ c2 ðt Tk Þ : S
ð2:33Þ
Tk
Differentiating C2 ðNÞ with respect to Tk and setting it equal to zero, c1 Tkþ1 Tk ¼ Tk Tk1 : c2 Therefore, the problem of optimizing checkpoint intervals of the backup operation corresponds to that of the inspection policy when the failure time has a uniform distribution. This is the reason, as failures occur uniformly over any interval. Fig. 2.3 Process of backup policy
26
2 Poisson Processes
Fig. 2.4 Excess cost and shortage cost of backward time T
Example 2.12 (Backward time problem) [6, p. 88] Suppose that when a unit has failed, and its failure time is unknown, we go back to time T ð0 T tÞ from time t to detect its failure and call T a planned backward time. Then, we introduce the following costs: Cost c1 ðxÞ is the excess cost for the time x from a failure to the backward time, c2 ðxÞ is the shortage cost for the time x from the backward time to a failure, and c0 ðTÞ is the cost required for the backward time with c0 ð0Þ 0 (Fig. 2.4). It is assumed that a failure is distributed uniformly over ½0; t based on the result obtained in (2.29). Then, the expected cost for the backward time T is Z T Z tT 1 CðTjtÞ ¼ c1 ðxÞdx þ c2 ðxÞdx þ c0 ðTÞ: ð2:34Þ t 0 0 In particular, when ci ðtÞ ¼ ci t ði ¼ 0; 1; 2Þ; the expected cost is i 1h 2 CðTjtÞ ¼ c1 T þ c2 ðt TÞ2 þ c0 T: 2t
ð2:35Þ
Differentiating CðTjtÞ with respect to T and setting it equal to zero, T¼
c 2 c0 t: c 2 þ c1
ð2:36Þ
Therefore, an optimum T ð0\T \tÞ that minimizes (2.35) is given by (2.36) for c2 [ c0 : If c2 c0 ; then T ¼ 0; i.e., we should not go back at all because the backward cost is high (Problem 2.14). This result in (2.29) is generalized as follows [4, p. 71, 8, p. 37]: The kth arrival times Sk ðk ¼ 1; 2; . . .; nÞ; given that NðtÞ ¼ n; has the probability kh1 ekh1 . . .khn ekhn ekðth1 h2 hn Þ ½ðktÞn =n!ekt n! ¼ n h1 h2 . . .hn t
Prftk \Sk \tk þ hk ; k ¼ 1; 2; . . .; njNðtÞ ¼ ng ¼
2.2 Poisson Process
27
for any 0\t1 \t2 \ \tn \tnþ1 ¼ t and tk þ hk \tkþ1 for small hk ; and hence, its density function is f ðt1 ; t2 ; . . .; tn jNðtÞ ¼ nÞ ¼
n! : tn
ð2:37Þ
That is, n arrival times S1 ; S2 ; . . .; Sn given that NðtÞ ¼ n are independent and identically distributed uniformly over the interval ½0; t: In addition [10, p. 15], from Example 2.10, PrfSk u; NðtÞ ¼ ng PrfNðtÞ ¼ ng PrfNðuÞ k; NðtÞ ¼ ng ¼ PrfNðtÞ ¼ ng Pn j¼k Prf NðuÞ ¼ jg PrfNðtÞ NðuÞ ¼ n jg ¼ PrfNðtÞ ¼ ng nj j n X n u u : ¼ 1 t t j j¼k
PrfSk ujNðtÞ ¼ ng ¼
ð2:38Þ
Clearly (Problem 2.15), EfSk jNðtÞ ¼ ng ¼
kt ; nþ1
EfSk Sk1 jNðtÞ ¼ ng ¼
t : nþ1
ð2:39Þ
2.3 Nonhomogeneous Poisson Process Suppose that a unit begins to operate at time 0: If the unit fails, then it undergoes minimal repair and begins to operate again. It is assumed that the time for repair is negligible. Let us denote 0 S0 S1 Sn1 Sn . . . be the successive failure times of the unit, and Xn Sn Sn1 ðn ¼ 1; 2; . . .Þ be the times between failures (Fig. 2.1) with FðtÞ PrfX1 tg and FðtÞ 1 FðtÞ: We define to make minimal repair at failure as follows [3, p. 96]: Definition 2.3 The unit undergoes minimal repair at failures if and only if PrfXn xjSn1 ¼ tg ¼
Fðt þ xÞ FðtÞ FðtÞ
ðn ¼ 2; 3; . . .Þ
ð2:40Þ
for x [ 0 such that FðtÞ\1: The function ½Fðt þ xÞ FðtÞ=FðtÞ is called the failure rate and represents the probability that a unit with age t fails in the interval ðt; t þ x: The definition in reliability models means that the failure rate remains undisturbed by any minimal repair of failures, i.e., a unit after each minimal repair has the same failure rate as
28
2 Poisson Processes
previous failure. Imperfect PM (Preventive Maintenance) models in which the failure rate after PM reduces were considered and their optimum policies were discussed analytically [3, p. 171]. A variety of failure rate modelling was extensively collected [12]. Assume that FðtÞ has a density function f ðtÞ and hðtÞ f ðtÞ=FðtÞ is continuous. The function hðtÞ is also called the instantaneous failure rate or simply the failure rate and has the same monotone property as ½Fðt þ xÞ FðtÞ=FðtÞ [2, p. 23]. Rt Moreover, HðtÞ 0 hðuÞdu is called the cumulative hazard function and satisfies the relation FðtÞ ¼ eHðtÞ that decreases with t: Because, from the definition of hðtÞ; 0 FðtÞ d log FðtÞ ¼ ¼ hðtÞ: dt FðtÞ By integration under the initial condition Fð0Þ ¼ 1; log FðtÞ ¼
Zt
hðuÞdu ¼ HðtÞ;
i.e.;
FðtÞ ¼ eHðtÞ :
ð2:41Þ
0
We derive the distribution of the nth failure time Sn ; 1 n1 X X ½HðtÞk HðtÞ ½HðtÞk HðtÞ Gn ðtÞ PrfSn tg ¼ e e ¼1 k! k! k¼n k¼0
ðn ¼ 1; 2; . . .Þ: ð2:42Þ
By the mathematical induction, G1 ðtÞ ¼ PrfX1 tg ¼ FðtÞ ¼ 1 eHðtÞ ; Z1 Gnþ1 ðtÞ ¼ PrfXnþ1 t ujSn ¼ ugdGn ðuÞ 0
¼
Zt
FðtÞ FðuÞ ½HðuÞn1 dFðuÞ ðn 1Þ! FðuÞ
0
¼1
n1 X ½HðtÞk k¼0
¼1
k!
e
HðtÞ
Zt
½HðuÞn1 hðuÞdu ðn 1Þ!
0
n X ½HðtÞk k¼0
e
HðtÞ
k!
eHðtÞ
ðn ¼ 1; 2; . . .Þ:
The density function of Gn ðtÞ is gn ðtÞ G0n ðtÞ ¼
½HðtÞn1 HðtÞ ½HðtÞn1 e f ðtÞ; hðtÞ ¼ ðn 1Þ! ðn 1Þ!
2.3 Nonhomogeneous Poisson Process
29
and hence, we have the relation Gn ðtÞ ¼
Zt 0
1 X ½HðuÞn1 ½HðtÞk HðtÞ dFðuÞ ¼ e : ðn 1Þ! k! k¼n
ð2:43Þ
From the above results, EfSn g
Z1 0
n1 Z X ½HðtÞk HðtÞ e Gn ðtÞdt ¼ dt; k! k¼0 1
0
EfXn g ¼ EfSn g EfSn1 g ¼
Z1
ð2:44Þ ½HðtÞn1 HðtÞ e dt: ðn 1Þ!
0
Next, we prove that when hðtÞ increases (decreases), EfXn g decreases (increases) with n and converge to 1=hð1Þ as n ! 1 [3, p. 98]. Using the relation (Problem 2.16) ½HðtÞn ¼ n!
Zt
½HðuÞn1 hðuÞdu; ðn 1Þ!
0
and from the assumption that hðtÞ increases, it follows that 9 8 Z1
¼
0
Z1
½HðtÞn1 HðtÞ dt; e ðn 1Þ!
0
and hence, EfXn g decreases with n: Furthermore, from (2.43), Z1 0
½HðtÞn HðtÞ dt ¼ e n!
Z1
½HðtÞn hðtÞ HðtÞ dt e n! hðtÞ
0
1
hð1Þ
Z1 0
½HðtÞn 1 dFðtÞ ¼ : n! hð1Þ
30
2 Poisson Processes
On the other hand, for any 0\T\1; Z1
½HðtÞn HðtÞ dt ¼ e n!
ZT
0
½HðtÞn HðtÞ dt þ e n!
½HðtÞn HðtÞ dt e n!
T
0
Z1
ZT
½HðtÞn HðtÞ 1 e dt þ : n! hðTÞ
0
Thus, because Z1 lim
n!1
½HðtÞn HðtÞ 1 e dt ; n! hðTÞ
0
and T is arbitrary, Z1 lim
n!1
½HðtÞn HðtÞ 1 : dt ¼ e hð1Þ n!
0
Similarly, when hðtÞ decreases, we can prove that EfXn g increases with n and converge to 1=hð1Þ as n ! 1; where we set that 1=hð1Þ 0 for hð1Þ ¼ 1 and 1=hð1Þ 1 for hð1Þ ¼ 0: Note that when hðtÞ increases strictly to 1; EfXn g decreases strictly to 0: The failure rate hðtÞ and the cumulative hazard Rt function HðtÞ 0 hðuÞdu in reliability theory are called an intensity function and a mean value function in stochastic processes, respectively. Furthermore, let NðtÞ be the number of failures of a unit in ½0; t; i.e., NðtÞ maxn fSn tg: Then, using the relation with NðtÞ and Sn in (2.16), Pn ðtÞ PrfNðtÞ ¼ ng ¼ PrfSn t\Snþ1 g ¼ Gn ðtÞ Gnþ1 ðtÞ ½HðtÞn HðtÞ e ðn ¼ 0; 1; 2; . . .Þ; n! EfNðtÞg ¼ VfNðtÞg ¼ HðtÞ: ¼
ð2:45Þ
In particular, when FðtÞ ¼ 1 ekt ; i.e., HðtÞ ¼ kt; PrfSn tg ¼ Gn ðtÞ ¼ 1
n1 X ðktÞk k¼0
PrfXn tg ¼ 1 ekt ;
k!
ekt ;
1 EfXn g ¼ ; k
n EðSn Þ ¼ ; k
EfNðtÞg ¼ VfNðtÞg ¼ kt;
that agree with those of the Poisson process with rate k: Referring to the above results and expanding Definition 2.1, we can generalize the Poisson process with the parameter of time t [4, p. 77, 8, p. 46]:
2.3 Nonhomogeneous Poisson Process
31
Definition 2.4 If (i) (ii) (iii) (iv)
Nð0Þ ¼ 0; the process fNðtÞg has independent increments, PrfNðt þ hÞ NðtÞ ¼ 1g ¼ kðtÞh þ oðhÞ; PrfNðt þ hÞ NðtÞ 2g ¼ oðhÞ;
then a counting process fNðtÞg is said to be a nonhomogeneous or nonstationary Poisson process that eliminates the stationary from a Poisson process. Using this definition, we derive the probability Pn ðuÞ PrfNðt þ uÞ NðtÞ ¼ ng ¼
½Hðt þ uÞ HðtÞn ½HðtþuÞHðtÞ e n!
ðn ¼ 0; 1; 2; . . .Þ
ð2:46Þ
for a fixed t: From (ii), (iii), and (iv), P0 ðu þ hÞ ¼ PrfNðt þ u þ hÞ NðtÞ ¼ 0g ¼ PrfNðt þ uÞ NðtÞ ¼ 0; Nðt þ u þ hÞ Nðt þ uÞ ¼ 0g ¼ PrfNðt þ uÞ NðtÞ ¼ 0g PrfNðt þ u þ hÞ Nðt þ uÞ ¼ 0g ¼ P0 ðuÞ½1 kðt þ uÞh þ oðhÞ: Hence, P0 ðu þ hÞ P0 ðuÞ oðhÞ ¼ kðt þ uÞP0 ðuÞ þ : h h Letting h ! 0; we have the differential equation P00 ðuÞ ¼ kðt þ uÞP0 ðuÞ;
i.e.;
dP0 ðuÞ=du ¼ kðt þ uÞ: P0 ðuÞ
By integration and setting P0 ð0Þ ¼ 1; log P0 ðuÞ ¼
Zu
kðt þ xÞdx ¼ ½Hðt þ uÞ HðtÞ;
0
that implies P0 ðuÞ ¼ e½HðtþuÞHðtÞ : Similarly, by the same method used in the Poisson process, we can prove (2.46). Clearly, EfNðt þ uÞ NðtÞg ¼
1 X ½Hðt þ uÞ HðtÞn HðtþuÞHðtÞ e n : n! n¼0
¼ Hðt þ uÞ HðtÞ:
32
2 Poisson Processes
Finally, the distribution of the interarrival time X1 ¼ S1 ; given that there was an event in ½0; t; is [4, p. 78], for u t; PrfNðuÞ ¼ 1; NðtÞ NðuÞ ¼ 0g PrfNðtÞ ¼ 1g PrfNðuÞ ¼ 1g PrfNðtÞ NðuÞ ¼ 0g ¼ PrfNðtÞ ¼ 1g
PrfS1 ujNðtÞ ¼ 1g ¼
¼
HðuÞeHðuÞ e½HðtÞHðuÞ HðuÞ : ¼ HðtÞ HðtÞeHðtÞ
ð2:47Þ
Example 2.13 (Expected number of failures) A random variable Y HðXÞ; that represents the expected number of failures until the failure time, has the following distribution
ð2:48Þ PrfY tg ¼ PrfHðXÞ tg ¼ Pr X H 1 ðtÞ ¼ 1 et ; where H 1 is the inverse function of H: Thus, Y has an exponential distribution with mean 1; and EfHðXÞg ¼ V fHðXÞg ¼ 1: In other words, by transforming the failure times S1 ; S2 ; . . . to HðS1 Þ; HðS2 Þ; . . .; a nonhomogeneous Poisson process corresponds to a Poisson process with rate 1: Moreover, x1 that satisfies Hðx1 Þ ¼ 1 is called a characteristic life in the probability paper of a Weibull distribution. This represents the mean lifetime that about 63.2 % of units have failed until time x1 ; as shown in (1) of Sect. 2.1.1. In general, xn that satisfies Hðxn Þ ¼ n ðn ¼ 1; 2; . . .Þ represents the time that the expected number of failures is n; when failures occur in a nonhomogeneous Poisson with a mean value function HðtÞ: From Hðxn Þ ¼ n; Zxn
hðtÞdt ¼ 1
Zxn or
xn1
hðtÞdt ¼ n
ðn ¼ 1; 2; . . .Þ:
0
In particular, when FðtÞ is IFR, xn xn1
1 ; hðxn1 Þ
i.e.;
xn
1 þ xn1 : hðxn1 Þ
Thus, denoting that xn
1 þ xn1 hðxn1 Þ
ðn ¼ 2; 3; . . .Þ;
xn xn : Similarly, xn
1
xn1 : hðxn Þ
ð2:49Þ
2.3 Nonhomogeneous Poisson Process
33
Thus, denoting xn that satisfies xn
1 ¼ xn1 hðxn Þ
ðn ¼ 2; 3; . . .Þ;
xn xn : The function 1=hðtÞ is called an instantaneous MTBF [11, p. 71]. Furthermore, from Theorem 4.4 of [2, p. 27] when F is IFR (DFR) and l EfXg; FðlÞ ðÞe1 : Thus, HðlÞ ð ÞHðx1 Þ; i.e., l ð Þx1 : Example 2.14 (Weibull distribution) Suppose that HðtÞ ¼ kta for a [ 0; i.e., the failure time has a Weibull distribution FðtÞ ¼ 1 expðkta Þ with a shape parameter a: Then, the failure rate is hðtÞ ¼ kata1 ; and decreases for 0\a\1; constant for a ¼ 1; and increases for 1\a\1: So that, the Weibull distribution represents a typical failure distribution that has three patterns in the bath-tub curve (Fig. 2.5). In this case, the process is sometimes called a Weibull process. Furthermore, from (2.49), xn ¼ ðn=kÞ1=a ðn ¼ 1; 2; . . .Þ and l ¼ ð1=kÞ1=a Cð1 þ 1=aÞ: Therefore, (i) x1 \l for 0\a\1; (ii) x1 ¼ l for a ¼ 1; and x1 [ l for a [ 1: Table 2.4. presents xn ; xn and xn when x1 ¼ x1 ¼ x1 ; and EðSn Þ from (2.44) for a ¼ 2 and k ¼ 0:01: This indicates that xn [ EðSn Þ; however, xn are almost the same as EðSn Þ for n 2: Many Weibull models were summarized [13]. Let 0 S0 \S1 \S2 \. . . be the successive times of a nonhomogeneous Poisson process with a mean value function HðtÞ: Then, we have the relation, from Example 2.13, fSnþ1 Sn [ tg , fHðSnþ1 Þ [ HðSn þ tÞg: In addition, because HðSnþ1 Þ HðSn Þ is independent of S1 ; S2 ; . . .; Sn and has an exponential distribution with mean 1, PrfSnþ1 Sn [ tjS1 ; S2 ; . . .; Sn g ¼ PrfHðSnþ1 Þ [ HðSn þ tÞjS1 ; S2 ; . . .; Sn g ¼ PrfHðSnþ1 Þ HðSn Þ [ HðSn þ tÞ HðSn Þg ¼ e½HðSn þtÞHðSn Þ :
Fig. 2.5 Bath-tub curve of the failure rate
34
2 Poisson Processes
Table 2.4 Comparisons of times of expected number when a ¼ 2 and k ¼ 0:01
n
xn
xn ðx1 ¼ x1 Þ
xn
EðSn Þ
1 2 3 4 5 6 7 8 9 10
10.00 14.14 17.32 20.00 22.36 24.49 26.46 28.28 30.00 31.62
10.00 15.00 18.33 21.06 23.43 25.56 27.52 29.34 31.04 32.65
10.00 13.66 17.07 19.84 22.25 24.41 26.39 28.23 29.95 31.59
8.86 13.29 16.62 19.39 21.81 23.99 25.99 27.85 29.59 31.23
The interarrival times for a nonhomogeneous Poisson are not independent, however, still have conditionally exponential distribution [5, p. 97].
2.4 Applications to Reliability Models 2.4.1 Replacement at the Nth Failure A new unit begins to operate at time 0; and when it fails, only minimal repair is made. It is assumed that the repair and replacement times are negligible. The failure times of the R 1unit have a density function f ðtÞ and a distribution FðtÞ with finite mean l 0 FðtÞdt; failure rate hðtÞ f ðtÞ=FðtÞ; and cumulative hazard Rt rate HðtÞ 0 hðuÞdu: Suppose that the unit is replaced only at the Nth failure. Then, the mean time to the Nth failure is, from (2.44), N 1 Z X
1
EfSN g ¼
k¼0
Pk ðtÞdt;
0
where Pk ðtÞ f½HðtÞk =k!geHðtÞ ðk ¼ 0; 1; 2; . . .Þ: Consider one cycle from the beginning of operation to the replacement. Let c1 be the cost of minimal repair and c2 be the replacement cost at the Nth failure. The expected cost rate is [3, p. 106] CðNÞ
Expected cost until replacement c1 ðN 1Þ þ c2 ¼ PN1 R 1 Mean time to replacement k¼0 0 Pk ðtÞdt
ðN ¼ 1; 2; . . .Þ: ð2:50Þ
2.4 Applications to Reliability Models
35
When the failure rate hðtÞ increases strictly to hð1Þ; we derive an optimum number N that minimizes CðNÞ in (2.50). From the inequality CðN þ 1Þ CðNÞ 0; PN1 R 1 c2 k¼0 0 Pk ðtÞdt R1 ðN ¼ 1; 2; . . .Þ: ð2:51Þ ðN 1Þ
c1 P ðtÞdt N 0 Letting LðNÞ be the left-hand side of (2.51), " 1 N Z X LðN þ 1Þ LðNÞ ¼ Pk ðtÞdt R 1 k¼0
0
0
# 1 R1 [ 0; PNþ1 ðtÞdt 0 PN ðtÞdt 1
R1 because 0 Pk ðtÞdt decreases strictly from (2.44) to 1=hð1Þ when hðtÞ increases strictly. Thus, if a solution to (2.51) exists, then it is unique, and the resulting cost rate is R1 0
c1 PN 1 ðtÞdt
\CðN Þ R 1 0
c1 : PN ðtÞdt
ð2:52Þ
Furthermore, we have the inequality (Problem 2.17) LðNÞ [ R 1 0
l PN ðtÞdt
ðN ¼ 2; 3; . . .Þ:
Therefore, if lim R 1
N!1
0
l c2
; PN ðtÞdt c1
i.e., if hð1Þ c2 =ðlc1 Þ; then a solution to (2.51) exists, and it is unique. In addition, if N is a unique minimum that satisfies Z1
PN ðtÞdt
lc1 ; c2
0
then N N: Example 2.15 (Optimum number) Suppose that the failure time of a unit has a Weibull distribution, i.e., FðtÞ ¼ expðkta Þ for a [ 1: Then, hðtÞ increases strictly from 0 to infinity, and (Problem 2.18) Z1 0
1 CðN þ 1=aÞ ; PN ðtÞdt ¼ 1=a a k CðN þ 1Þ
N 1 Z X
1
k¼0
0
Pk ðtÞdt ¼
CðN þ 1=aÞ : k1=a CðNÞ
Thus, there exists a finite and unique minimum N that satisfies (2.51) which is given by
36
2 Poisson Processes
c2 c1 N ¼ ða 1Þc1
1=a þ 1;
ð2:53Þ
where ½x denotes the greatest integer contained in x:
2.4.2 Software Reliability Model In software reliability modeling [14], it is assumed that HðtÞ is the expected number of faults detected by time t and the number of faults detected at any time is proportional to the current number of faults. Let a [ 0 be the total number of faults, and bðtÞ be the error detection rate per fault at time t: Then, because the number of faults at time t is a HðtÞ; we have the differential equation dHðtÞ ¼ bðtÞ½a HðtÞ: dt Solving this equation under the initial condition Hð0Þ 0 (Problem 2.19), 8 2 39 Zt = < ð2:54Þ HðtÞ ¼ a 1 exp4 bðuÞdu5 : ; : 0
This is called a software reliability growth model. In particular, when bðtÞ ¼ b [ 0; HðtÞ ¼ a 1 ebt ; and Hð1Þ ¼ a: A variety of generalized software reliability growth models were summarized extensively [11, 15], and several interesting new topics were cited [16].
2.5 Compound Poisson Process Consider a standard cumulative damage model [17, p. 16, 18]: Shocks occur in a Poisson Process with rate k: A unit is subjected to shocks and suffers some damage due to shocks. Let random variables Xk ðk ¼ 1; 2; . . .Þ denote a sequence of interarrival times between successive shocks, and random variables Wk ðk ¼ 1; 2; . . .Þ denote the damage produced by the kth shock, where W0 0 (Fig. 2.6). It is assumed that the sequence of fWk g is nonnegative, independently, and identically distributed according to PrfWk xg GðxÞ with finite mean, and furthermore, Wk is independent of Xj ðj 6¼ kÞ:
2.5 Compound Poisson Process
37
Fig. 2.6 Process for a standard cumulative damage model
Let NðtÞ denote the total number of shocks up to time t: Then, define a random variable ZðtÞ
NðtÞ X
Wk
ðNðtÞ ¼ 0; 1; 2; . . .Þ;
ð2:55Þ
k¼0
where ZðtÞ represents the total damage at time t: This is called a compound Poisson Process and will be discussed in detail in Chap. 7. The distribution of ZðtÞ defined in (2.55) is ( NðtÞ ) X Wj x PrfZðtÞ xg ¼ Pr j¼0
¼
1 X k¼0
¼ ¼
(
Pr
k X j¼0
) Wj x NðtÞ ¼ k PrfNðtÞ ¼ kg
1 X
ðktÞk kt e PrfW0 þ W1 þ þ Wk xg k! k¼0
1 X
ðktÞk kt GðkÞ ðxÞ e : k! k¼0
ð2:56Þ
Letting G ðsÞ be the LS transform of GðxÞ; the LS transform of (2.56) is Z1 0
esx d PrfZðtÞ xg ¼
1 X
½G ðsÞk
k¼0
ðktÞk kt e k!
¼ expfkt½1 G ðsÞg: Because EfNðtÞg ¼ VfNðtÞg ¼ kt in (2.18),
ð2:57Þ
38
2 Poisson Processes
( ( NðtÞ )) X ¼ EfNðtÞgEfWk g ¼ ktEfWg; EfZðtÞg ¼ E E Wk NðtÞ k¼1
VfZðtÞg ¼ E Z 2 ðtÞ ½EfZðtÞg2 ( ( NðtÞ )) NðtÞ X X ½EfZðtÞg2 ¼E E Wk Wj NðtÞ j¼1 k¼1
¼ VfNðtÞg½EfWg2 þ EfNðtÞgVfWg ¼ ktE W 2 :
ð2:58Þ
When a random variable X has the LS transform function F ðsÞ; i.e., F ðsÞ R 1 sx d PrfX xg; the above results are derived by (Appendix A) 0 e dF ðsÞ d2 F ðsÞ 2 EfXg ¼ ; EfX g ¼ : ds s¼0 ds2 s¼0 Next, it is assumed that the unit fails when the total damage has exceeded a failure level K ð0\K\1Þ: Usually, a failure level K is statistically estimated and is already known. We are interested in a random variable defined by Y mint fZðtÞ [ Kg; i.e., PrfY tg represents the distribution of the failure time and EfYg is the MTTF of the unit. In stochastic processes, PrfY tg is called the firstpassage distribution to some event. Because the events of fY tg and fZðtÞ [ Kg are equivalent, from (2.56), UðtÞ PrfY tg ¼ PrfZðtÞ [ Kg 1 h 1 i X X ðktÞj kt GðkÞ ðKÞ Gðkþ1Þ ðKÞ ¼ e ; j! k¼0 j¼kþ1
ð2:59Þ
and its LS transform is
U ðsÞ
Z1 0
est dUðtÞ ¼
1 h i k kþ1 X GðkÞ ðKÞ Gðkþ1Þ ðKÞ : sþk k¼0
Thus, the MTTF and variance are 1 dU ðsÞ 1X 1 EfYg ¼ ¼ GðkÞ ðKÞ ¼ ½1 þ MG ðKÞ; ds s¼0 k k¼0 k 1 2 d U ðsÞ 2X VfYg ¼ ðEfYgÞ2 ¼ 2 ðk þ 1ÞGðkÞ ðKÞ ðEfYgÞ2 ; 2 ds k k¼0 s¼0
ð2:60Þ
P ðkÞ where MG ðKÞ 1 k¼1 G ðKÞ represents the expected number of shocks before the total damage exceeds a failure level K: Furthermore, the failure rate of the unit with a failure level K is
2.5 Compound Poisson Process
U0 ðtÞ ¼ rðtÞ 1 UðtÞ
39
i h P1 ðkÞ ðkþ1Þ ðKÞ kðktÞk =k! k¼0 G ðKÞ G h i : P1 ðkÞ k k¼0 G ðKÞ ðktÞ =k!
ð2:61Þ
Example 2.16 (MTTF) Suppose that shocks occur in a Poisson process with rate k and each damage due to shocks is exponential with mean 1=h; i.e., FðtÞ ¼ 1 ekt and GðxÞ ¼ 1 ehx : Then, from (2.57), Z1 esx d PrfZðtÞ xg ¼ ekt½s=ðsþhÞ : 0
By inversion [2, p. 80, 17, p. 23], 2
3 Zx p ffiffiffiffiffiffi ffi p ffiffiffiffiffiffiffiffiffi PrfZðtÞ xg ¼ ekt 41 þ kht ehu u1=2 I1 ð2 khtuÞdu5;
ð2:62Þ
0
where Ii ðxÞ is the Bessel function of order i for the imaginary argument defined by 1 2kþi X x 1 : Ii ðxÞ 2 k!ðk þ iÞ! k¼0 Thus, from (2.62), 2
3 ZK pffiffiffiffiffiffiffiffiffi p ffiffiffiffiffiffi ffi PrfY tg ¼ 1 ekt 41 þ kht ehu u1=2 I1 2 khtu du5:
ð2:63Þ
0
Furthermore, from (2.58) and (2.60) (Problem 2.20), kt hK þ 1 ; EfYg ¼ ; h k 2kt 2hK þ 1 ; V fZðtÞg ¼ 2 ; VfYg ¼ h k2 EfZðtÞg ¼
ð2:64Þ
where note that EfZðtÞg increases linearly with time t: In addition, we have the relation EfZðtÞg t ¼ : K þ 1=h EfYg Finally, we give two modified damage models in reliability theory:
40
2 Poisson Processes
(1) Independent Damage Model [17, p. 21, 19]. Suppose that the total damage is not additive, i.e., any shock does no damage unless its amount has not exceeded a failure level K: If the damage due to some shock has exceeded for the first time a failure level K; then the unit fails. The same assumptions as those of the previous model are made except that the total damage is additive. A typical example of this model is the fracture of brittle materials such as glasses, and semiconductor parts that have failed by some over current of fault voltage. In this model, the probability that the unit fails exactly at the ðk þ 1Þth shock ðk ¼ 0; 1; 2; . . .Þ is ½GðKÞk ½GðKÞkþ1 : Thus, the distribution of time to failure is PrfY tg ¼ 1 expfkt½1 GðKÞg;
ð2:65Þ
that is an exponential distribution with parameter k½1 GðKÞ; and hence, the MTTF and the failure rate are EfYg ¼
1 ; k½1 GðKÞ
rðtÞ ¼ k½1 GðKÞ ¼
VfYg ¼
1 k ½1 GðKÞ2 2
;
1 ; EfYg
ð2:66Þ
that is constant for any t 0: (2) Random Failure Level [17, p. 29, 19]. Suppose that a failure level K is not constant and is a random variable with a distribution LðxÞ PrfK xg; where Lð0Þ 0: Then, for the cumulative damage model, the distribution of time to failure is, from (2.59)
PrfY tg ¼
1 1 Z h X k¼0
0
1 i X ðktÞj kt e ; GðkÞ ðxÞ Gðkþ1Þ ðxÞ dLðxÞ j! j¼kþ1
ð2:67Þ
and the MTTF and the failure rate are, respectively, 1 1X EfYg ¼ k k¼0
Z1 0
GðkÞ ðxÞdLðxÞ;
h i P1 R 1 ðkÞ ðkþ1Þ ðxÞ dLðxÞ kðktÞk =k! k¼0 0 G ðxÞ G h i : rðtÞ ¼ P1 R 1 ðkÞ k k¼0 0 G ðxÞdLðxÞ ðktÞ =k!
ð2:68Þ
2.5 Compound Poisson Process
41
For the independent 1 damage model, 1 Z n 1 o X X ðktÞj kt PrfY tg ¼ e ; ½GðxÞk ½GðxÞkþ1 dLðxÞ j! k¼0 j¼kþ1 0
1 Z 1X
ð2:69Þ
1
EfYg ¼
k k¼0
½GðxÞk dLðxÞ:
0
Example 2.17 (Exponential distribution) (Problem 2.21) Suppose that FðtÞ ¼ 1 ekt ; GðxÞ ¼ 1 ehx and LðxÞ ¼ 1 ebx : Then, for the independent damage model, 1 1 PrfY tg ¼ 1 expðktehK Þ; EfYg ¼ ¼ ehK : rðtÞ k For the random failure level, Z1h i GðkÞ ðxÞ Gðkþ1Þ ðxÞ dLðxÞ ¼ 0
Z1
est d PrfY tg ¼
0
By inversion,
bhk ðh þ bÞkþ1
;
kþ1 1 X k bhk kb : ¼ kþ1 s þ k sðh þ bÞ þ kb ðh þ bÞ k¼0
kbt ; PrfY tg ¼ 1 exp hþb 1 hþb ¼ ; EfYg ¼ rðtÞ kb
ð2:70Þ
hþb 2 VfYg ¼ : kb
ð2:71Þ
In addition, for the independent damage model, PrfY [ tg ¼
Z1
expðktehx Þbebx dx
0
Z 1 X ðktÞk
1
¼
k¼0
k!
1 1 ¼ EfYg ¼ rðtÞ k
beðbþkhÞx dx ¼
k¼0
0
Z1
( hx
bx
e be 0
1 X ðktÞk
dx ¼
k!
b ; b þ kh
b kðbhÞ
ðb [ hÞ;
1
ðb hÞ:
(3) Nonhomogeneous Poisson process. Suppose that shocks occur in a nonhomogeneous Poisson process with an intensity function hðtÞ and a mean value function HðtÞ: Then, because the probability of NðtÞ is given in (2.45),
42
2 Poisson Processes
PrfZðtÞ xg ¼
1 X
GðkÞ ðxÞ
k¼0
EfYg ¼
1 X
ðkÞ
½HðtÞk HðtÞ e ; k!
G ðKÞ
k¼0
Z1
EfZðtÞg ¼ EfWgHðtÞ;
½HðtÞk HðtÞ e dt: k!
0
For the independent damage model, PrfY tg ¼ 1 e
½1GðKÞHðtÞ
;
EfYg ¼
Z1
e½1GðKÞHðtÞ dt:
0
The failure rate is rðtÞ ¼ ½1 GðKÞhðtÞ; that has the same property as that of an intensity function hðtÞ:
2.6 Problems 2 2.1 For any negative function gðtÞ in Example 2.3 [5, p. 84], show that ( ) Z1 1 X gðSn Þ ¼ k gðtÞdt: E n¼1
0
2.2 Show that a continuous failure distribution FðtÞ that satisfies the memoryless properly, i.e., Fðt þ uÞ ¼ FðtÞFðuÞ for any t; u 0; is only an exponential distribution [8, p. 24]. 2.3 Let XðkÞ ðk ¼ 1; 2; . . .; nÞ be the kth failure time of n units in Example 2.4. Then, noting that Xð1Þ be the failure time of a series system in which when one of n units fails, the system fails, and XðnÞ be the failure time of a parallel system when all of n units have failed, the system fails. Then, show that the failure distributions of the two systems are, respectively,
Pr Xð1Þ t ¼
Zt
nkenku du ¼ 1 enkt ;
0
Pr XðnÞ t ¼
Zt
nð1 ekt Þn1 kekt dt ¼ ð1 ekt Þn ;
0
EfXð1Þ g ¼
1 ; nk
EfXðnÞ g ¼
n 1X 1 : k k¼1 k
2.6 Problems 2
43
2.4 When FðtÞ is IFR in Example 2.4, show that the MTTF ln of a parallel system with n units has the inequality [2, p. 65, 6, p. 9] l1 ln l1
n X 1 k¼1
k
:
2.5 Check numerically in Example 2.4 that when the failure time has a Weibull a distribution FðtÞ ¼ 1 et ða 1Þ; the MTTF of a parallel system is given by approximately [20], n X 1
k k¼1
!1=a ;
ðC þ log nÞ1=a :
2.6 When the acquisition cost of one unit is c1 and the replacement cost of a failed system is c2 in Example 2.4, show that the expected cost rate is [6, p. 8] CðnÞ ¼
nc1 þ c2 P ð1=kÞ nk¼1 ð1=kÞ
ðn ¼ 1; 2; . . .Þ;
and derive analytically an optimum n that minimizes CðnÞ: 2.7 When the failure time is exponential, obtain the MTTF of a k-out-of-n system. 2.8 A harmonic series in (2.8) appears in a record time [8, p. 47]: Let X1 ; X2 ; . . . be independent and identically distributed random variables. It is said that a record occurs at time n if Xn [ maxfX1 ; X2 ; . . .; Xn1 g: In addition, let Nn denote the total number of records until time n: Then, show [8, p. 286] that n n X X 1 1 1 EfNn g ¼ ; VfNn g ¼ 1 : k k k k¼1 k¼1 This is applied to a breaking test [7]: Suppose that there are n wooden beams. To know the strength of which beams can bear a minimum burden, we make the following experiment: We record a burden B1 on which the first beam breaks. Next, we bear B1 on the second beam. If the beam does not break, then B2 [ B1 : Conversely, if it breaks, we record a burden B2 : Similarly, we bear minfB1 ; B2 g on the third beam. If the beam does not break, then B3 [ minfB1 ; B2 g; and conversely, if it breaks, we record B3 : Then, in the same method of a record time, Pn show that the expected number of broken beams by such experiment is k¼1 ð1=kÞ: Compute the expected number of broken beams when n ðn ¼ 10; 100; 1000Þ experiments are made. 2.9 Show that the limiting distribution of binomial distribution is a Poisson distribution [4, p. 40]. 2.10 For a n units parallel system in Example 2.6, where each unit has an identical failure distribution FðtÞ; the expected cost rate is [6, p. 84]
44
2 Poisson Processes
CðnÞ ¼ c2 þ ðc1 c2 Þ
Z1
WðtÞd½FðtÞn þ c0 n ðn ¼ 0; 1; 2; . . .Þ:
0
An optimum number n is given by a unique minimum that satisfies Z1 c0 ½FðtÞn FðtÞdWðtÞ : c1 c2 0
In particular, when WðtÞ ¼ 1 ext and FðtÞ ¼ 1 ekt ; the above equation is n X x c0 n : ð1Þk k x þ ðk þ 1Þk c1 c2 k¼0 Derive the above equations, compute optimum n that minimize CðnÞ for k=x and c0 =ðc1 c2 Þ; and compare them with Table 2.3. 2.11 Suppose that failed units from two factories arrive at a repair shop according to Poisson processes with rates k1 and k2 in Example 2.8. Show how many number of spare units should be provided in order to assure with probability a that the system will remain operating in time t in Example 2.8. 2.12 Suppose that failures are roughly classified into two kinds of failures; minor failure and major failure in Example 2.9. Minor and major failures occur at the rates of 0.90 and 0.10, respectively. When kt ¼ 1; 2; 3; compute the probabilities that the number of minor failures is k ðk ¼ 0; 1; 2; 3Þ: 2.13 When the objective function in Example 2.11 is [6, p. 41], CðNÞ ¼ AN þ
B N
or
CðNÞ ¼ AN B1=N
ðN ¼ 1; 2; . . .Þ
for parameters A; B [ 0; show that an optimum N is a unique minimum that satisfies N X B k : 2A k¼1 2.14 Consider the multiple backward times Tk ðk ¼ 1; 2; . . .; NÞ; where Sk Pk j¼1 Tj ; S0 0; and SN t in Example 2.12 [21]. 2.15 Prove (2.39). When 10 failures have occurred in 100 h, derive the probabilities that kðk ¼ 0; 1; 2Þ failures have occurred in 10 h in Example 2.10, and the third failure has occurred until 50 h in (2.38). Rt 2.16 Prove in p. 29 that when HðtÞ 0 hðuÞdu; i.e., hðtÞ H 0 ðtÞ; ½HðtÞn ¼ n!
Zt 0
½HðuÞn1 hðuÞdu ðn 1Þ!
ðn ¼ 1; 2; . . .Þ:
2.6 Problems 2
45
2.17 Prove in (2.51) that when hðtÞ increases strictly, PN1 R 1 l k¼0 0 Pk ðtÞdt R1 ðN 1Þ [ R 1 0 PN ðtÞdt 0 PN ðtÞdt where Pk ðtÞ ¼ f½HðtÞk =k!geHðtÞ and l 2.18 Derive in Example 2.15 that Z1 0
1 CðN þ 1=aÞ ; PN ðtÞdt ¼ 1=a a k CðN þ 1Þ
N 1 Z X
R1 0
ðN ¼ 2; 3; . . .Þ;
eHðtÞ dt:
1
k¼0
Pk ðtÞdt ¼
0
CðN þ 1=aÞ : k1=a CðNÞ
2.19 Prove in Sect. 2.4.2 that the solution of the differential equation dHðtÞ ¼ bðtÞ½a HðtÞ; dt is given by (2.54). 2.20 Derive EfZðtÞg; VfZðtÞg from (2.57), and EfYg; VfYg from (2.60) in Example 2.16. 2.21 Derive all equations in Example 2.17.
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18.
Kingman JFC (1993) Poisson processes. Oxford University Press, Oxford Barlow RE, Proschan F (1965) Mathematical theory of reliability. Wiley, New York Nakagawa T (2005) Maintenance theory of reliability. Springer, London Osaki S (1992) Applied stochastic system modeling. Springer, Berlin Çinlar E (1975) Introduction to stochastic processes. Prentice-Hall, Englewood Cliffs Nakagawa T (2008) Advanced reliability models and maintenance policies. Springer, London Havil J (2003) GAMMA: exploring Euler’s constant. Princeton University Press, Princeton Ross SM (1983) Stochastic processes. Wiley, New York Beichelt FE, Fatti LP (2002) Stochastic processes and their applications. CRC Press, Boca Raton Tijms HC (2003) A first course in stochastic models. Wiley, Chichester Yamada S (1994) Software reliability models: fundamentals and applications. JUSE Press, Tokyo Finkelstein M (2008) Failure rate modelling for reliability and risk. Springer, London Murthy DNP, Xie M, Jiang R (2004) Weibull models. Wiley, Hoboken Goel AL, Okumoto K (1979) Time-dependent error-detection rate model for software reliability and other performance measures. IEEE Trans Reliab R-28:206–211 Pham H (2006) System software reliability. Springer, London Pham, H (ed) (2008) Recent advances in reliability and quality in design. Springer, London Nakagawa T (2007) Shock and damage models in reliability theory. Springer, London Cox DR (1962) Renewal theory. Methuen, London
46
2 Poisson Processes
19. Esary JD, Marshall AW, Proschan F (1973) Shock models and wear processes. Ann Probab 1:627–649 20. Yun WY, Nakagawa T (2010) Note on MTTF of a parallel system. In: 16th ISSAT International Conference on Reliability and Quality in Design, pp 235–237 21. Chen M, Mizutani S, Nakagawa T (2010) Optimal backward and backup policies in reliability theory. J Oper Res Soc Jpn 53:101–118
Chapter 3
Renewal Processes
In Chap. 2, we have defined a Poisson process and shown that the interarrival times are distributed exponentially. However, in general, such times have arbitrary distributed random variables, that is called a renewal process. In most reliability systems, they may deteriorate with age and break at last, and become like new by suitable replacement or maintenance. A renewal process arises from the study of self-renewing aggregates, and a renewal theory plays an important role in the analysis of stochastic models with the total sum of independent nonnegative random variables. The results can be applied to a wide variety of both theoretical and practical probability problems in engineering, natural, economical, and social sciences. A renewal process in this chapter is the most fundamental process, and a renewal theory is an important theory in stochastic processes. Many basic stochastic systems form renewal processes essentially. Renewal processes are the largest and most crucial chapter in stochastic processes for studying reliability theory because learning a reliability theory is learning a renewal theory: In Sect. 3.1, we define renewal processes by generalizing Poisson processes. Renewal processes are characterized by renewal functions. In Sect. 3.2, it is shown that a renewal function is obtained by solving a renewal equation, and its properties are given in the name of a renewal theory. Some theorems in a renewal theory are cited as theorems without mathematical proof because it needs more mathematical preparation. But, it would be sufficient on theorem forms to apply them to reliability theory. The most fundamental policy in maintenance models is a replacement policy. In Sect. 3.3, we show how to apply a renewal theory to standard replacement policies. Using the techniques and results of a renewal theory, optimum policies for some replacement models are discussed analytically. In Sect. 3.4, we introduce an alternating renewal process that repeats on and off, up and down, or between operating and failure states. A one-unit system with repair maintenance which is the most basic reliability system is given as a typical example, and its properties are examined and the interval reliability is given by using the results of an alternating renewal process. Finally, in Sect. 3.4, we take up a geometric and a discrete renewal processes as examples of modified renewal processes. A geometric process is the process whose
T. Nakagawa, Stochastic Processes, Springer Series in Reliability Engineering, DOI: 10.1007/978-0-85729-274-2_3, Ó Springer-Verlag London Limited 2011
47
48
3 Renewal Processes
interarrival times decrease or increase geometrically, and a discrete process is defined to be a renewal process with discrete interarrival times.
3.1 Definition of Renewal Process We define renewal processes by generalizing interarrival times Xk in a Poisson process: Consider a sequence of independent and nonnegative random variables fX1 ; X2 ; . . .g in Fig. 2.1, where it is assumed that PrfXk ¼ 0g\1 for all k because of avoiding trivial cases. Suppose that X2 ; X3 ; . . . have an arbitrary identical distribution FðtÞ with finite mean l; however, X1 possibly has a different distribution F1 ðtÞ with finite l1 from FðtÞ; in which both F1 ðtÞ and FðtÞ are not degenerate at time t ¼ 0 and F 1 ðtÞ 1 F1 ðtÞ: P We introduce the total sum of random variables Sn nk¼1 Xk ; where S0 0 for convenience, and define NðtÞ maxn fSn tg that represents the number of renewals in the time interval ½0; t; where the process fNðtÞg is generally called a counting process. Clearly, from (2.16), fSn tg , fNðtÞ ng: It is said that the nth renewal occurs at time Sn : We call epochs at which Sn ends renewal points, renewal times or regeneration points, and the time from the beginning of Sn to its end a renewal cycle with time interval Xn : We use properly two words of renewal point and regeneration point in the right place. Note that we have the memoryless property only at renewal points except F1 ðtÞ and FðtÞ are exponential distributions. In other words, the process starts newly at each renewal point. We will use the terms of events and renewals interchangeably. We define renewal processes according to three types of F1 ðtÞ: Definition 3.1 (i) If FðtÞ ¼ F1 ðtÞ; i.e., all random variables are identically distributed, then the counting process fNðtÞg is called an ordinary renewal process or renewal process for short. (ii) If F1 ðtÞ and FðtÞ are not the same, then the counting process fNðtÞg is called a modified or delayed renewal process. Rt (iii) If F1 ðtÞ 0 F 1 ðuÞdu=l that will be given in (3.24), then the counting process fNðtÞg is called an equilibrium or stationary renewal process. Example 3.1 (Reliability systems) Consider a unit that is replaced with a new one upon failures: A unit begins to operate immediately after the replacement whose time is negligible. Suppose that the failure distribution of each new unit is FðtÞ: If a new unit is installed at time t ¼ 0; then all failure times have the same distribution, and hence, we have an ordinary renewal process. On the other hand, if an old unit is in use at time t ¼ 0; then X1 represents its residual lifetime and could be different from the failure time of a new unit, and hence, we have a modified renewal process. In particular, if the time origin point at which we start observing
3.1 Definition of Renewal Process
49
the process is sufficiently large after the installation of an initial operating unit, it may be considered that the process has already entered the stationary state and Rt the failure distribution of X1 is given by 0 F 1 ðuÞdu=l; then we have an equilibrium renewal process. We may consider an operating unit that is repaired upon failures and whose repair time may be assumed to be negligible because they are small usually compared with failure times. If we observe only the operating states that repeat successively, then it may be said that most reliability systems can form renewal processes. Therefore, renewal processes are the most fundamental stochastic process in reliability theory. In general, if two positive random variables X and Y are independent and have the respective distributions FðtÞ PrfX tg and GðtÞ PrfY tg; then the sum of X and Y has the distribution PrfX þ Y tg ¼
Zt
Fðt uÞdGðuÞ ¼
0
Zt
Gðt uÞdFðuÞ FðtÞ GðtÞ;
0
and its LS transform is, from Appendix A, Z1
est d PrfX þ Y tg ¼ EfesðXþYÞ g ¼ EfesX gEfesY g
0
¼
Z1
est dFðtÞ
0
Z1
est dGðtÞ ¼ F ðsÞG ðsÞ;
0
i.e., Z1
est d½FðtÞ GðtÞ ¼ F ðsÞG ðsÞ;
0
where the asterisk is called the pairwise Stieltjes convolution. Denoting F ð0Þ ðtÞ 1 for t 0 and F ðnÞ ðtÞ PrfX2 þ X3 þ þ Xnþ1 tg Zt ¼ PrfX3 þ X4 þ þ Xnþ1 t ugd PrfX2 ug 0
¼
Zt
F ðn1Þ ðt uÞdFðuÞ
0
¼ FðtÞ FðtÞ
ðn ¼ 1; 2; . . .Þ;
ð3:1Þ
50
3 Renewal Processes
F ðnÞ ðtÞ is called the n-fold Stieltjes convolution of FðtÞ with itself. That is, F ðnÞ ðtÞ represents the distribution of the total sum of independent and identical random P ðnÞ variables nþ1 k¼2 Xk : In addition, the LS transform of F ðtÞ is, from Appendix A, Z1 est dF ðnÞ ðtÞ ¼ ½F ðsÞn ðn ¼ 0; 1; 2; . . .Þ; ð3:2Þ 0
R1
where note that 0 est dF ð0Þ ðtÞ ¼ 1 for any ReðsÞ 0: Furthermore, from the relation of Sn and NðtÞ; PrfNðtÞ ¼ 0g ¼ PrfX1 [ tg ¼ 1 F1 ðtÞ; PrfNðtÞ ¼ ng ¼ PrfSn t and Snþ1 [ tg ¼ PrfSn tg PrfSnþ1 tg ¼ F1 ðtÞ F ðn1Þ ðtÞ F1 ðtÞ F ðnÞ ðtÞ
ðn ¼ 1; 2; . . .Þ:
ð3:3Þ
3.2 Renewal Function We want to know properties of the number of renewals in ½0; t and define its expected number as MðtÞ EfNðtÞg; that is called the renewal function, and the derivative of MðtÞ is mðtÞ dMðtÞ=dt; that is called the renewal density. From (3.3) (Problem 3.1), the renewal function is MðtÞ ¼
1 X n¼1 1 X
n PrfNðtÞ ¼ ng ¼
1 X
PrfNðtÞ ng
n¼1
PrfSn tg ¼
n¼1
1 X
F1 ðtÞ F ðn1Þ ðtÞ:
ð3:4Þ
n¼1
It is fairly easy to show that MðtÞ is finite for all t 0 because PrfXk ¼ 0g\1 (Problem 3.1). Furthermore, from the notations of the convolutions in (3.1) and (3.4), MðtÞ ¼ F1 ðtÞ þ
1 Z X n¼1
mðtÞ ¼ f1 ðtÞ þ
Zt
t
F ðnÞ ðt uÞdF1 ðuÞ ¼
0
Zt 0
½1 þ M0 ðt uÞdF1 ðuÞ; ð3:5Þ
m0 ðt uÞf1 ðuÞdu;
0
where M0 ðtÞ is the renewal function of an ordinary renewal process with distriP P1 ðnÞ ðnÞ bution FðtÞ; i.e., M0 ðtÞ 1 ðtÞ; and n¼1 F ðtÞ; m0 ðtÞ dM0 ðtÞ=dt ¼ n¼1 f f ðtÞ and f1 ðtÞ are the respective density functions of FðtÞ and F1 ðtÞ: Such types of equations as (3.5) are called renewal-type equations and appear frequently in the
3.2 Renewal Function
51
analysis of reliability models, because most systems are renewed after maintenance as shown in Sect. 3.3. It is of great importance in reliability analysis to search for renewal points and set up renewal equations. Generally, solutions of renewal equations can be computed by only numerical methods, however, they are obtained explicitly in the form of LS transforms that will be shown in Sect. 3.2.3. The LS transform of MðtÞ is, from (3.4) or (3.5), Z1 F1 ðsÞ est dMðtÞ ¼ M ðsÞ : ð3:6Þ 1 F ðsÞ 0
Thus, MðtÞ is determined by F1 ðtÞ and FðtÞ: When F1 ðtÞ and FðtÞ are identical, i.e., the process forms an ordinary renewal process, MðtÞ ¼ M0 ðtÞ; and (3.6) implies F ðsÞ ¼ M ðsÞ=½1 þ M ðsÞ: Therefore, FðtÞ and MðtÞ determines uniquely with each other because the LS transform determines its original function uniquely. In particular, when the process forms a stationary renewal process, from (iii) of Definition 3.1, F1 ðsÞ
¼
Z1
est dF1 ðtÞ ¼
1 F ðsÞ ; ls
0
and hence, from (3.6) and Table A.2 in Appendix A, M ðsÞ ¼
1 ; ls
i.e.;
t MðtÞ ¼ : l
After this, it is assumed that F1 ðtÞ ¼ FðtÞ; i.e., we deal with only an ordinary renewal process. Example 3.2 (Gamma and normal distributions) Suppose that FðtÞ is a gamma Pk1 distribution, i.e., FðtÞ ¼ 1 j¼0 ðktÞj =j! ekt ðk ¼ 1; 2; . . .Þ: In particular, when k ¼ 1; from Sect. 2.1.2, k M ðsÞ ¼ ; s
i.e.,
MðtÞ ¼ kt;
that is derived directly, as shown in (2.18). When k ¼ 2; FðtÞ ¼ 1 ð1 þ ktÞekt ; and 2 k F ðsÞ ¼ ; sþk and hence, from (3.6), M ðsÞ ¼
k2 k k : ¼ sðs þ 2kÞ 2s 2ðs þ 2kÞ
52
3 Renewal Processes
Using the LS transform method in Appendix A such that Z1 e
st
Z1
a dðatÞ ¼ ; s
0
est dð1 eat Þ ¼
a ; sþa
0
the inverse LS transform of M ðsÞ is MðtÞ ¼
kt 1 1 e2kt : 2 4
In general, because the probability of n renewals in ½0; t for a renewal process is equal to the probability of either nk; nk þ 1; . . .; or nk þ k 1 events in ½0; t for a Poisson process with rate k (Fig. 3.1) [1, p. 57], PrfNðtÞ ¼ ng ¼
nkþk1 X j¼nk
ðktÞj kt e ; j!
and hence, because mðtÞdt is the probability of a renewal in ðt; t þ dt; " # 1 X ðktÞnk1 kt kdt: e mðtÞdt ¼ ðnk 1Þ! n¼1 Thus, a renewal function is MðtÞ ¼
Zt 0
¼
1 Z X kðkuÞnk1 ku mðuÞdu ¼ e du ðnk 1Þ! n¼1 t
0
1 X 1 X ðktÞj n¼1 j¼nk
j!
ekt :
ð3:7Þ
On the other hand, from Sect. 2.1.2 and (3.6), the LS transform of MðtÞ for a gamma distribution with order k is M ðsÞ ¼
kk ðs þ kÞk kk
;
and inverting its LS transform [1, p. 57, 2, p. 121, 3]
Fig. 3.1 Relationship between Poisson and renewal processes for a gamma distribution with order k
3.2 Renewal Function
53
MðtÞ ¼
k1 i kt 1 X hj h 1 ektð1hj Þ ; þ k k j¼1 1 hj
ð3:8Þ
where hj e2pji=k ðj ¼ 1; 2; . . .; k 1Þ is the distinct root of sk ¼ 1 and i is an imaginary order. In particular, when k ¼ 3 [4, p. 162] (Problem 3.2), pffiffiffi
3 1 2 p kt þ ; MðtÞ ¼ kt 1 þ pffiffiffi e1:5kt sin 2 3 3 3 and when k ¼ 4; MðtÞ ¼
pffiffiffi 1 3 1 p kt þ e2kt þ 2ekt sin kt þ : 4 2 2 4
Figure 3.2 draws the renewal functions MðtÞ for k ¼ 1; 2; 3 when k ¼ 1: Rt Suppose generously that FðtÞ ¼ 0 ½kðkxÞa1 ekx =CðaÞdx for parameters k [ 0 and a [ 0: Because the sum of two random variables with gamma distributions with the respective parameters ða; kÞ and ðb; kÞ has a gamma distribution with parameter ða þ b; kÞ (Problem 3.3), Zt kðkxÞna1 kx e dx; F ðnÞ ðtÞ ¼ ð3:9Þ CðnaÞ 0
and hence, 1 Z X kðkxÞna1 kx e dx; MðtÞ ¼ CðnaÞ n¼1 t
0
that agrees with (3.7) when a ¼ k ðk ¼ 1; 2; . . .Þ: Fig. 3.2 Renewal functions MðtÞ for k ¼ 1; 2; 3
ð3:10Þ
54
3 Renewal Processes
When Xk ðk ¼ 1; 2; . . .Þ is normally distributed with mean l and variance r2 for l r; because Sn is normally distributed with mean nl and variance nr2 ; from (3.4), 1 1 X X t nl ; ð3:11Þ PrfSn tg ¼ U pffiffiffi MðtÞ ¼ nr n¼1 n¼1 where UðxÞ is a standard normal distribution with mean 0 and variance 1 (Problems 3.4 and 3.5). P Next, we derive the variance V fNðtÞg [5, p. 89]: Because n2 ¼ 2 nj¼1 j n; EfNðtÞ2 g ¼
1 X
n2 PrfNðtÞ ¼ ng ¼ 2
n¼1 1 X
¼2
1 X
n PrfNðtÞ ng MðtÞ
n¼1
n PrfSn tg MðtÞ ¼ 2
n¼1
1 X
nF ðnÞ ðtÞ MðtÞ:
n¼1
Taking the LS transform on both sides above, Z1
1 n o X est dE NðtÞ2 ¼ 2 n½F ðsÞn M ðsÞ n¼1
0
F ðsÞ 2 F ðsÞ ¼2 þ 1 F ðsÞ 1 F ðsÞ ¼ 2½M ðsÞ2 þM ðsÞ:
ð3:12Þ
Inverting (3.12), n o E NðtÞ2 ¼ 2MðtÞ MðtÞ þ MðtÞ; and hence, V fNðtÞg ¼ 2MðtÞ MðtÞ þ MðtÞ ½MðtÞ2 :
ð3:13Þ
Because MðtÞ MðtÞ ½MðtÞ2 ; VfNðtÞg MðtÞ þ ½MðtÞ2 : In particular, when FðtÞ ¼ 1 ekt ; V fNðtÞg ¼ kt; that agrees with (2.18). Furthermore, when FðtÞ ¼ 1 ð1 þ ktÞekt ; from Example 3.2 and (3.13) (Problem 3.6), V fNðtÞg ¼
kt kt 1 þ ð1 e2kt Þ þ ð1 e4kt Þ: 4 2 16
3.2 Renewal Function
55
We summarize some important limiting theorems and results of renewal theory [5–7]: R1 Theorem 3.1 If l2 0 t2 dFðtÞ\1 and r2 l2 l2 ; 2 t r 1 MðtÞ ¼ þ þ oð1Þ as t ! 1; ð3:14Þ 2l2 2 l R1 and if l3 0 t3 dFðtÞ\1; 4 r2 t 5r 2r2 3 2l3 þ þ V fNðtÞg ¼ 3 þ þ oð1Þ as t ! 1; ð3:15Þ 4l4 l2 l 4 3l3 where oðhÞ=h ! 0 as h ! 0: Proof Expanding F ðsÞ with respect to s; F ðsÞ ¼ 1 ls þ
1 2 1 r þ l2 s2 l3 s3 þ o s3 : 2 3!
Substituting F ðsÞ for (3.6) and arranging them, 2 1 r l2 þ oð1Þ; M ðsÞ ¼ þ 2l2 sl 4 1 1 r2 l 2 3r r2 3 l3 þ þ oð1Þ: ½M ðsÞ2 ¼ 2 2 þ þ þ l3 4l4 2l2 4 3l3 sl s
ð3:16Þ
Inverting M ðsÞ follows (3.14). Furthermore, substituting (3.16) for (3.12) and inverting it, we have (3.15) from (3.13) and (3.14) (Problem 3.6). h From this theorem, MðtÞ and mðtÞ are approximately given by t r2 1 1 þ 2 ; mðtÞ ; l 2l 2 l r2 t 5r4 2r2 3 2l3 V fNðtÞg 3 þ 4 þ 2 þ 3 4l l l 4 3l MðtÞ
ð3:17Þ
for large t: Furthermore, if r l; then MðtÞ
t 1 : l 2
Theorem 3.2 If l and r2 are finite, MðtÞ 1 ! ; t l
1 mðtÞ ! ; l
V fNðtÞg r2 ! 3 l t
as t ! 1:
ð3:18Þ
This theorem that is easily derived from Theorem 3.1 is called an elementary renewal theorem and is the most important one in a renewal process. This is also
56
3 Renewal Processes
applied not only in the reliability theory but in also many other practical fields. Another proof of Theorem 3.2 that is more direct than that of the above was shown [1, p. 55]. A random variable X with distribution FðtÞ is said to be lattice if there exists P h [ 0 such that 1 n¼0 PrfX ¼ nhg ¼ 1: We give Blackwell’s theorem [8, p. 347]: Theorem 3.3 (Blackwell’s theorem) If X is not lattice, then for h [ 0; h lim ½Mðt þ hÞ MðtÞ ¼ : l
t!1
This theorem shows that the expected number of renewals in ½t; t þ h tends to h=l approximately as t ! 1; i.e., it is proportional to the time duration where the proportional constant is 1=l: Furthermore, we introduce a renewal reward process [6, p. 77]. For instance, if we consider the total reward produced by the successive production of some system, then the process is said to form a renewal reward process, where the successive production can be described by a renewal process and the total rewards caused by production may be additive. Chapter 6 will deal in detail with a cumulative process that is another name of a renewal reward process. Define that some reward Wk is earned at the kth renewal point ðk ¼ 1; 2; . . .Þ: When a sequence of pairs fXk ; Wk g is independent and identically distributed, PNðtÞ ZðtÞ k¼1 Wk is the total reward earned in ½0; t: Theorem 3.4 If EfWg EfWk g\1; EfZðtÞg EfWg ! t l
as t ! 1:
This is proved by intuition because if NðtÞ and W are independent, EfZðtÞg ¼ EfNðtÞgEfWg ¼ MðtÞEfWg: Thus, using Theorem 3.2, EfZðtÞg MðtÞ EfWg ¼ EfWg ! t t l
as t ! 1:
The detailed proof of Theorem 3.4 was shown [6, p. 78]. This theorem shows that if one cycle is denoted by the time interval between renewals, the expected reward per unit of time for an infinite time span is equal to the expected reward per cycle, divided by its mean time, i.e., lim
t!1
EfZðtÞg Expected reward of one renewal cycle ¼ : t Mean time of one renewal cycle
ð3:19Þ
3.2 Renewal Function
57
This function is adopted by appropriate objective functions of maintenance models where the reward of maintenance is given by costs. There will exist many optimization problems with aim to minimize them in reliability theory [1], as shown in Sect. 3.3.2 Applying the asymptotic formulae of (3.17) to the usual central limit theorem, we have a so-called central limit theorem for a renewal process: Theorem 3.5 (
NðtÞ t=l lim Pr pffiffiffiffiffiffiffiffiffiffiffiffiffi x t!1 r2 t=l3
)
1 ¼ pffiffiffiffiffiffi 2p
Zx
eu
2
=2
du;
ð3:20Þ
1
i.e., NðtÞ is asymptotically normally distributed with mean t=l and variance r2 t=l3 for large t: The detailed proof was found in [5, p. 93, 6, p. 62]. Example 3.3 (Asymptotic normal distribution) We compute approximately the number n when the probability that the number of failures is less than n in t ¼ 10; 000 h of operation is given by 0.90. The failure distribution of each unit is unknown, but, it is estimated from the sample data that the mean and variance are l ¼ 1; 000 h and r2 ¼ 200; 000 ðhÞ2 ; respectively. Then, from Theorem 3.5, when t ¼ 10; 000; NðtÞ t=l Nð10; 000Þ 10 pffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 2 r2 t=l3 is approximately normally distributed with mean 0 and variance 1. Thus,
Nð10; 000Þ 10 n 10 pffiffiffi PrfNð10; 000Þ ng ¼ Pr pffiffiffi 2 2 pffiffi ðn10Þ= 2 Z 1 2
pffiffiffiffiffiffi eu =2 du ¼ 0:90: 2p 1
pffiffiffiffiffiffi R u0 u2 =2 e du ¼ 0:90; Because u0 ¼ 1:28 that satisfies ð1= 2pÞ 1 pffiffiffi n 10 þ 1:28 2 ¼ 11:81: In this case, it can be said that with probability more than 0.90, the system with 12 units, i.e., the MTTF is 1:2 104 h, including the first unit, will continue to operate for 104 hours without any supply (Problem 3.7). Theorem 3.6 (Key or Fundamental renewal theorem) If gðtÞ is bounded in ½0; 1Þ and FðtÞ is not lattice, then [2, p. 123, 5, p. 95, 9, p. 295],
58
3 Renewal Processes
Zt lim
t!1
1 gðt uÞdMðuÞ ¼ l
0
Z1
ð3:21Þ
gðtÞdt: 0
This is proved intuitively that from Theorem 3.3 [6, p. 65], Mðt þ hÞ MðtÞ dMðtÞ 1 ¼ lim ¼ ; t!1 dt h l Zt Z1 Zt 1 1 gðt uÞdMðuÞ ¼ lim gðt uÞdu ¼ gðtÞdt: lim t!1 x!1 l l lim lim
h!0 t!1
0
0
0
Recall that mðtÞ ¼ dMðtÞ=dt is called the renewal density and represents the expected number of renewals per unit of time.
3.2.1 Age and Residual Life Distributions Let dðtÞ t SNðtÞ and cðtÞ SNðtÞþ1 t that denote the shortage and excess random variables (Fig. 3.3), respectively. In reliability theory, dðtÞ represents the current age and cðtÞ represents the residual or remaining life of an operating unit. Thus, bðtÞ dðtÞ þ cðtÞ ¼ SNðtÞþ1 SNðtÞ ¼ XNðtÞþ1 represents the total life of the unit. Theorem 3.7 (i) PrfcðtÞ xg ¼ Fðt þ xÞ
Zt
Fðt þ x uÞdMðuÞ:
ð3:22Þ
0
(ii)
PrfdðtÞ xg ¼
FðtÞ
R tx 0
Fðt uÞdMðuÞ 1
for x t; for x [ t:
ð3:23Þ
(iii) The limiting distribution is 1 lim PrfdðtÞ xg ¼ lim PrfcðtÞ xg ¼ t!1 t!1 l
Zx FðuÞdu: 0
Fig. 3.3 Age dðtÞ and residual life cðtÞ in a renewal process
ð3:24Þ
3.2 Renewal Function
59
Proof From Fig. 3.3 [1, p. 58, 5, p. 97], PrfcðtÞ xg ¼ Pr SNðtÞ t\SNðtÞþ1 t þ x 1 X Prft\Skþ1 t þ x; NðtÞ ¼ kg ¼ k¼0
¼
1 X
PrfSk t\Skþ1 t þ xg
k¼0
¼
1 Z X k¼0
¼
Prft u\Xkþ1 t þ x ugd PrfSk ug
0
1 Z X k¼0
t
:
t ðkÞ
½Fðt þ x uÞ Fðt uÞdF ðuÞ
0
¼ Fðt þ xÞ FðtÞ þ
Zt
½Fðt þ x uÞ Fðt uÞdMðuÞ
0
¼ Fðt þ xÞ
Zt
Fðt þ x uÞdMðuÞ:
0
Furthermore, because both events fdðtÞ [ xg and fcðt xÞ [ xg mean that there is no renewal in ½t x; t; we have the equality PrfdðtÞ [ xg ¼ Prfcðt xÞ [ xg; that implies (3.22) from (3.23). Applying Theorem 3.6 to (3.22) and (3.23), 1 lim PrfcðtÞ xg ¼ 1 t!1 l
Z1
1 Fðt þ xÞdt ¼ l
0
Zx FðuÞdu; 0
and lim PrfdðtÞ xg ¼ lim Prfcðt xÞ xg ¼ lim PrfcðtÞ xg;
t!1
t!1
t!1
that follows (3.24). This means that when the process started at t ¼ 1; looking backwards and forwards from the present time t; the age and the residual life have the same distribution in a renewal process. h From (i) of Theorem 3.7 (Problem 3.8), 2 3 Zt Z1 EfcðtÞg ¼ 4FðxÞ þ Fðx uÞdMðuÞ5dx t
0
¼ l½1 þ MðtÞ t;
ð3:25Þ
60
3 Renewal Processes
that is also easily derived because EfcðtÞg ¼ E SNðtÞþ1 t ¼ l½1 þ MðtÞ t: It is also of interest that the mean of the above limiting distribution in (3.24) is (Problem 3.8) 1 l
Z1
tFðtÞdt ¼
l l2 l2 l r2 ¼ þ ; þ 2l 2 2 2l
ð3:26Þ
0
that is greater than half of the mean interval time l: In general [2, p. 128, 6 p. 69] (Problem 3.9),
lim PrfbðtÞ xg ¼
t!1
Zx
1 l
udFðuÞ:
ð3:27Þ
0
Example 3.4 (Exponential distribution) When FðtÞ ¼ 1 ekt ; PrfdðtÞ xg ¼ PrfcðtÞ xg ¼ lim PrfdðtÞ xg t!1
¼ lim PrfcðtÞ xg ¼ 1 ekx ; t!1
and its mean time is 1=k: In addition, PrfbðtÞ xg ¼ PrfdðtÞ þ cðtÞ xg ¼ lim PrfbðtÞ xg t!1
¼ 1 ð1 þ kxÞe
kx
;
i.e., the failure time of the unit that is operating at time t has a gamma distribution with order 2, and its mean life is twice as much as the usual unit. Example 3.5 (Uniform distribution) Suppose that FðtÞ is distributed uniformly on ½0; 10; i.e., FðtÞ ¼ t=10 for 0 t 10 and l ¼ 5: In this case, from (3.24), 1 lim PrfcðtÞ xg ¼ t!1 5
Zx
1
u x x2 du ¼ 10 5 100
0
for 0 x 10; and its mean is 10=3 [ l=2 ¼ 2:5: From (3.27), lim PrfbðtÞ xg ¼
t!1
for 0 x 10 and its mean is 20=3:
x2 100
3.2 Renewal Function
61
3.2.2 Expected Number of Failures The renewal function MðtÞ in a renewal process and the cumulative hazard function HðtÞ in a nonhomogeneous Poisson process play a great role in reliability theory. We summarize and compare the properties of MðtÞ and HðtÞ: Suppose that the unit begins to operate at time 0 and fails according to a failure distribution FðtÞ with finite mean l and a density function f ðtÞ dFðtÞ=dt: 3.2.2.1 Properties of MðtÞ The unit is repaired immediately at failures and operates again like new. (a) Renewal function MðtÞ represents the expected number of failures in ½0; t and is, from (3.5), Zt
MðtÞ ¼ FðtÞ þ
Mðt uÞdFðuÞ
0 1 X
¼
F ðnÞ ðtÞ ¼ FðtÞ þ FðtÞ FðtÞ þ :
n¼1
The LS transform is, from (3.6), M ðsÞ ¼
F ðsÞ : 1 F ðsÞ
Thus, FðtÞ and MðtÞ determine uniquely with each other. However, it is not easy to compute MðtÞ from FðtÞ except FðtÞ is a gamma distribution (Example 3.2). (b) Renewal density is mðtÞ ¼ dMðtÞ=dt; and mðtÞdt represents the probability of a renewal occurring in ½t; t þ dt; that is given by a renewal equation mðtÞ ¼ f ðtÞ þ
Zt
mðt uÞf ðuÞdu;
0
¼
1 X
f ðnÞ ðtÞ ¼ f ðtÞ þ f ðtÞ f ðtÞ þ :
n¼1
The Laplace transform is
m ðsÞ ¼
Z1 0
where f ðsÞ
R1 0
est f ðtÞdt:
est mðtÞdt ¼
f ðsÞ ; 1 f ðsÞ
62
3 Renewal Processes
3.2.2.2 Properties of HðtÞ The unit is repaired at failures and undergoes minimal repair defined in Sect. 2.3. (a) Cumulative hazard function HðtÞ represents the expected number of failures in ½0; t and is, from Sect. 2.3, FðtÞ ¼ 1 eHðtÞ ;
HðtÞ ¼ log FðtÞ:
i.e.,
Hence, ½HðtÞ2 ½HðtÞ3 þ þ ; 2! 3! ½FðtÞ2 ½FðtÞ3 þ þ : HðtÞ ¼ FðtÞ þ 2 3 FðtÞ ¼ HðtÞ
Clearly, 0
hðtÞ H ðtÞ ¼ f ðtÞ½1 þ FðtÞ þ FðtÞ2 þ : (b) Hazard rate or failure rate function is hðtÞ dHðtÞ=dt; and hðtÞdt represents the probability of a failure occurring in ½t; t þ dt; given that the age is t; and the unit undergoes minimal repair at each failure that is given by Rt f ðtÞ hðuÞdu ; i.e., hðtÞ ¼ FðtÞ ¼ 1 e 0 : 1 FðtÞ When FðtÞ ¼ 1 ekt ; MðtÞ ¼ HðtÞ ¼ kt;
mðtÞ ¼ hðtÞ ¼ k:
3.2.2.3 Inequalities of MðtÞ and HðtÞ It is important to investigate MðtÞ and HðtÞ for maintenance models [11, p. 5] when FðtÞ is IFR, i.e., hðtÞ increases with t: (a) When FðtÞ with finite mean l and variance r2 is IFR [1, p. 54] (Problem 3.10), l r; MðxÞ Mðt þ xÞ MðtÞ; x MðxÞ lim ½Mðt þ xÞ MðtÞ ¼ ; t!1 l t V½NðtÞ MðtÞ ; l t t tFðtÞ t 1 MðtÞ R t : 1Rt l l FðuÞdu FðuÞdu 0 0
3.2 Renewal Function
63
Because mðtÞdt is the probability of a renewal in ½t; t þ dt; mðtÞ ¼
Zt
hðxÞd PrfdðtÞ xg:
0
Thus, when FðtÞ is IFR, mðtÞ hðtÞ;
i.e.;
MðtÞ HðtÞ:
In particular, when f ðtÞ ¼ k2 tekt ; for t [ 0; k k2 t 1 e2kt \hðtÞ ¼ ; 2 1 þ kt kt 1 MðtÞ ¼ ð1 e2kt Þ\HðtÞ ¼ kt logð1 þ ktÞ: 2 4 mðtÞ ¼
(b) Because ea 1 þ a; and from (a) of Sect. 3.2.2.2, HðtÞ FðtÞ ; FðtÞ HðtÞ 1 þ HðtÞ FðtÞ and hence, when FðtÞ is IFR, HðtÞ FðtÞ : FðtÞ MðtÞ HðtÞ 1 þ HðtÞ FðtÞ Furthermore, we have the inequality HðtÞ FðtÞ Rt ; t 0 FðuÞdu
ð3:28Þ
that is proved as follows: Let QðtÞ HðtÞ
Zt
FðuÞdu tFðtÞ;
0
we have that Qð0Þ ¼ 0 and 0
Q ðtÞ ¼ hðtÞ
Zt
FðuÞdu FðtÞ FðtÞ½thðtÞ HðtÞ:
0
Furthermore, Q0 ð0Þ ¼ 0 and 00
0
Q ðtÞ ¼ h ðtÞ
Zt 0
½FðtÞ FðuÞdu þ f ðtÞ
Zt 0
½hðtÞ hðuÞdu 0;
64
3 Renewal Processes
that implies (3.28). Therefore (Problem 3.11), MðtÞ FðtÞ HðtÞ FðtÞ : Rt hðtÞ R 1 t t t FðuÞdu 0 FðuÞdu
ð3:29Þ
3.2.2.4 Mean Time between Failures (MTBF) Let Xk be the failure time between the ðk 1Þth and the kth failures ðk ¼ 1; 2; . . .Þ: Then, for a renewal process, Z1 EfXk g ¼ FðtÞdt ¼ l 0
for all k 1; and for a nonhomogeneous Poisson process, from (2.44), EfXk g ¼
Z1
½HðtÞk1 HðtÞ dt e ðk 1Þ!
ðk ¼ 1; 2; . . .Þ:
0
Thus, when FðtÞ is IFR, from Sect. 2.3, l ¼ EfX1 g EfX2 g ; and l EfXk g !
1 hð1Þ
as k ! 1:
Furthermore, x1 that satisfies Hðx1 Þ ¼ 1 is called characteristic life in the probability paper of a Weibull distribution, i.e., it represents the mean lifetime that about 1 e1 63:2% of units have failed until time x1 , as shown in (1) of Sect. 2.1.1. When FðtÞ is IFR, t1 that satisfies Mðt1 Þ ¼ 1 is greater than x1 because MðtÞ HðtÞ: This means that more than 63.2% of units have failed until time t1 at which the expected number of failures is 1 for a renewal process.
3.2.3 Computation of Renewal Function In general, it is not so easy to compute a renewal function except the interarrival distribution FðtÞ is a gamma distribution. The following three methods of computing a renewal function MðtÞ are presented [12, p. 310, 13]: 3.2.3.1 Gamma Approximation Suppose that the distribution FðtÞ of a random variable X is approximately replaced by a gamma distribution as follows: From (a) of Sect. 3.2.2.1,
3.2 Renewal Function
65
MðtÞ ¼
1 X
F ðnÞ ðtÞ
n¼1
k X
F ðnÞ ðtÞ þ
n¼1
1 X
ðnÞ
FG ðtÞ
ð3:30Þ
n¼kþ1
for k ¼ 2; 3; . . .; and FG ðtÞ is a gamma distribution, i.e., 1 FG ðtÞ ¼ CðaÞ
Zt
ka xa1 ekx dx
for k [ 0 and a [ 0;
0 ðnÞ
and FG ðtÞ is given in (3.9). Because the infinite series is truncated, this computation can be established by choosing that EfXg ¼ na=k and VfXg ¼ na=k2 and ðkÞ using the stopping criterion for the first k such that FG ðtÞFG ðtÞ=F G ðtÞ\e for some specified e [ 0: 3.2.3.2 Laplace Inversion The LS transform of MðtÞ is given in Example 3.2 and is obtained by inverting it. The inversion formula for computing the renewal density is Zbþic
1 mðtÞ ¼ lim c!1 2pi
est M ðsÞds;
ð3:31Þ
bic
pffiffiffiffiffiffiffi where i ¼ 1; b [ maxfr; 0g, and r is a radius of convergence. But, (3.31) would be hard to use actually. Several methods of computing (3.31) are summarized [12, p. 462, 13]: The typical method is, applying the Fourier series to (3.31), 2ebt mðtÞ ¼ p
Z1
ReðM ðb þ iuÞÞ cosðxtÞdx
0 bt
2e ¼ p
Z1
ð3:32Þ ImðM ðb þ iuÞÞ sinðxtÞdx:
0
3.2.3.3 Discretization Algorithm From (3.5), the renewal function MðtÞ is computed numerically by solving the renewal equation MðtÞ ¼ FðtÞ þ
Zt 0
Mðt uÞdFðuÞ:
66
3 Renewal Processes
Xie [14] proposed the discretization algorithm for solving the above Volterra integral equation: Divide the time interval ½0; t by n line segments with equal length d; and denote Mi ¼ MðidÞ and Ki ¼ FðidÞ ði ¼ 0; 1; 2; . . .; nÞ at each point i ¼ 0; d; 2d; . . .; nd ¼ t; where Fi Fðði 0:5ÞdÞ: Then, the following recursive formula gives the approximation of the renewal function Mn ; starting from M0 0; " # i1 X 1 Ki þ ðMj Mj1 ÞFijþ1 Mi1 F1 : Mi ¼ ð3:33Þ 1 F1 j¼1 This formula is easy to program and surprisingly shows accurate results. The renewal function of a Weibull distribution was presented in [12, p. 312] (Problem 3.4).
3.3 Age Replacement Policies 3.3.1 Renewal Equation We can generalize the renewal equation in (3.5) to the following equation form [5, p. 95, 6, p. 65] gðtÞ ¼ hðtÞ þ
Zt
gðt uÞdFðuÞ;
ð3:34Þ
0
where hðtÞ and FðtÞ are known, and gðtÞ is unknown. This is called the renewaltype equation. Taking the LS transform of (3.34),
g ðsÞ
Z1
est dgðtÞ ¼
h ðsÞ ¼ h ðsÞ½1 þ M ðsÞ: 1 F ðsÞ
0
Thus, inverting the LS transform g ðsÞ; gðtÞ ¼ hðtÞ þ
Zt
hðt uÞdMðuÞ:
ð3:35Þ
0
Such modified types of equations appear well in reliability theory, as shown in (3.22) and (3.23). We show how to set up renewal-type equations for a variety of reliability models by paying close attention to renewal points and making some suitable modifications: Consider an age replacement model where a unit is replaced at time Tð0\T 1Þ after its installation or at failure, whichever occurs first. The event fT ¼ 1g represents that no planned replacement is done at all. It
3.3 Age Replacement Policies
67
is assumed that failures are instantly detected and each failed unit is replaced with a new one, where its replacement time is negligible, and so, a newly installed unit begins to operate instantly. Furthermore, suppose that the failure time Xk ðk ¼ 1; 2; . . .Þ of each operating unit is independent andR has an identical failure distri1 bution FðtÞ PrfXk tg with finite mean l 0 FðtÞdt\1 throughout this section. A new unit is installed at time t ¼ 0: Then, an age replacement procedure generates a renewal process as follows: Let fXk g1 k¼1 be the failure times of successive operating units. Define a new random variable Zk minfXk ; Tgðk ¼ 1; 2; . . .Þ: Then, fZk g1 k¼1 represents the intervals between replacements caused by either failures or planned replacements (Fig. 3.4). In other words, because PrfZk ¼ Xk g ¼ PrfXk Tg ¼ FðTÞ; PrfZk ¼ Tg ¼ PrfXk [ Tg ¼ FðTÞ; such trials corresponds to a Bernoulli trial [5, p. 34] whose outcome can be classified as either success or failure with the respective probabilities p FðTÞ and q FðTÞ: A sequence of random variables fZk g1 k¼1 ; where Zk is independently and identically distributed, forms a renewal process with an interarrival distribution
FðtÞ for t T; PrfZk tg ¼ 1 for t [ T: Thus, the mean time between replacements is EfZk g ¼
ZT
tdFðtÞ þ TFðTÞ ¼
0
ZT FðtÞdt: 0
First, let Y1 be the first time at which the unit is replaced before time T by failure, and G1 ðtÞ PrfY1 tg: Then, by paying attention to the renewal point of replacement times, G1 ðtÞ ¼
Zt 0
DðuÞdFðuÞ þ
Zt
G1 ðt uÞFðuÞdDðuÞ;
0
where DðtÞ is a degenerate distribution at time T; i.e., DðtÞ 0 for t\T and 1 for t T: Taking the LS transform and solving it for G1 ðsÞ; Fig. 3.4 Process of age replacement with planned time T
68
3 Renewal Processes
G1 ðsÞ
Z1 e
st
0
R 1 st R T st DðtÞdFðtÞ e dFðtÞ 0R e dG1 ðtÞ ¼ ¼ 0 sT : 1 st 1 0 e FðtÞdDðtÞ 1 e FðTÞ
ð3:36Þ
Similarly, letting Y2 be the first time at which the unit is replaced before failure by the planned replacement, G2 ðtÞ PrfY2 tg ¼
Zt
FðuÞdDðuÞ þ
Zt
0
G2 ðt uÞDðuÞdFðuÞ;
0
and its LS transform is
G2 ðsÞ
Z1
est dG2 ðtÞ ¼
0
esT FðTÞ : RT 1 0 est dFðtÞ
ð3:37Þ
Therefore, the mean times of Y1 and Y2 are, respectively, ZT 1 G1 ðsÞ 1 l1 EfY1 g ¼ lim FðtÞdt; ¼ s!0 s FðTÞ 0
1 G2 ðsÞ 1 l2 EfY2 g ¼ lim ¼ s!0 s FðTÞ
ð3:38Þ
ZT FðtÞdt: 0
It can be easily seen that l1 decreases from 1=hð0Þ to l when FðtÞ is IFR, l2 increases from 0 to 1; and when FðTÞ ¼ 1=2; l1 ¼ l2 (Problem 3.12). The above results are easily derived by solving the respective renewal equations ZT l1 ¼ ðT þ l1 ÞFðTÞ þ tdFðtÞ; 0
l2 ¼ TFðTÞ þ
ZT
ðt þ l2 ÞdFðtÞ;
0
and
EfZk g ¼ 1=
1 1 : þ l1 l2
It is of interest that 2EfZk g is given by the harmonic mean of l1 and l2 : Next, let M1 ðtÞ be the expected number of replacements due to failures in ½0; t: We have a renewal equation for M1 ðtÞ M1 ðtÞ ¼
Zt 0
½1 þ M1 ðt uÞDðuÞdFðuÞ þ
Zt 0
M1 ðt uÞFðuÞdDðuÞ:
3.3 Age Replacement Policies
69
Forming the LS transform and solving for M1 ðtÞ;
M1 ðsÞ ¼
Z1
st
e
RT dM1 ðtÞ ¼ R T0 0
0
est dFðtÞ
sest FðtÞdt
:
ð3:39Þ
Similarly, let M2 ðtÞ be the expected number of replacements due to the planned replacement in ½0; t: We have a renewal equation for M2 ðtÞ M2 ðtÞ ¼
Zt
M2 ðt uÞDðuÞdFðuÞ þ
0
Zt
½1 þ M2 ðt uÞFðuÞdDðuÞ:
0
Thus,
M2 ðsÞ ¼
Z1
est dM2 ðtÞ ¼ R T 0
0
est FðTÞ sest FðtÞdt
:
ð3:40Þ
Then, by using a Tauberian Theorem in Appendix A.3, lim
t!1
M1 ðtÞ FðTÞ 1 ¼ ; ¼ lim sM1 ðsÞ ¼ R T s!0 t FðtÞdt l1 0
M2 ðtÞ FðTÞ 1 ¼ : lim ¼ lim sM2 ðsÞ ¼ R T t!1 s!0 t FðtÞdt l2
ð3:41Þ
0
Therefore, letting MðtÞ be MðtÞ M1 ðtÞ þ M2 ðtÞ that is the expected number of replacements in ½0; t; lim
t!1
MðtÞ 1 1 1 1 ¼ þ : ¼ RT ¼ t FðtÞdt EfZk g l1 l2 0
Three limiting results of M1 ðtÞ; M2 ðtÞ and MðtÞ satisfy the elementary renewal theorem.
3.3.2 Optimum Replacement Policies We give optimum policies for age and periodic replacements as applied examples of renewal processes [1, p. 84, 11, p. 70]. 3.3.2.1 Age Replacement We take up the age replacement model where the unit is replaced at time T or at failure, whichever occurs first, as shown in Sect. 3.3.1. The model forms a renewal process RT with an interarrival distribution PrfZk tg and its mean time EfZk g ¼ 0 FðtÞdt:
70
3 Renewal Processes
We consider the problem of minimizing the expected cost per unit of time for an infinite time span. Introduce the following replacement costs: Cost c1 is incurred for each failed unit that is replaced; this includes all costs resulting from the failure and its replacement. Cost c2 ð\c1 Þ is incurred for each exchange of nonfailures at time T: Let M1 ðtÞ and M2 ðtÞ denote the expected replacement numbers of failures and b nonfailures and the expected cost CðtÞ denote c1 M1 ðtÞ þ c2 M2 ðtÞ in ½0; t: Then, from (3.41), the expected cost per unit of time for an infinite time span is
b M1 ðtÞ M2 ðtÞ CðtÞ CðTÞ lim þ c2 lim c1 t!1 t t!1 t t ð3:42Þ c1 c2 c1 FðTÞ þ c2 FðTÞ ¼ þ ¼ : RT l 1 l2 0 FðtÞdt We consider the renewal cycle from one replacement to the next replacement. Then, the pairs of time and cost on each cycle are independently and identically distributed, and both have finite means. The expected cost of one cycle is c1 PrfXk Tg þ c2 PrfXk [ Tg ¼ c1 FðTÞ þ c2 FðTÞ; RT and the mean time of one cycle is EfZk g ¼ 0 FðtÞdt: Thus, the expected cost per unit of time is easily given in (3.42) from (3.19). We call CðTÞ the expected cost rate and generally adopt it as the objective function of the optimization problem for age replacement with an infinite time span [11, p. 71]. We derive an optimum planned replacement time T that minimizes the expected cost rate CðTÞ in (3.42). It is assumed that the failure rate hðtÞ f ðtÞ=FðtÞ is differentiable and increases strictly, i.e., h0 ðtÞ dhðtÞ=dt [ 0; where f ðtÞ is a density function of FðtÞ: It is clearly seen that Cð0Þ lim CðTÞ ¼ 1; T!0
Cð1Þ lim CðTÞ ¼ T!1
c1 : l
Thus, there exists a positive T ð0\T 1Þ: In addition, differentiating CðTÞ with respect to T and setting it equal to zero, ZT hðTÞ
FðtÞdt FðTÞ ¼
c2 : c 1 c2
0
Letting Q1 ðTÞ be the left-hand side of (3.43), Qð0Þ ¼ 0 and Q1 ð1Þ ¼ lhð1Þ 1: Differentiating Q1 ðTÞ with respect to T; dQ1 ðTÞ ¼ h0 ðTÞ dT
ZT FðtÞdt [ 0: 0
ð3:43Þ
3.3 Age Replacement Policies
71
Therefore, we have the optimum policy when K c1 =½lðc1 c2 Þ: (i) If hð1Þ [ K; then there exists a finite and unique T ð0\T \1Þ that satisfies (3.43), and the resulting cost rate is CðT Þ ¼ ðc1 c2 ÞhðT Þ: (ii) If hð1Þ K; then T ¼ 1; i.e., the unit is replaced only at failure, and the expected cost rate is Cð1Þ ¼ c1 =l:
3.3.2.2 Three Periodic Replacement We take up three periodic replacements where the unit is replaced at periodic times kT ðk ¼ 1; 2; . . .Þ: These replacement models form the simplest renewal process with a constant time T ð0\T 1Þ: First, we consider the block replacement where each failed unit between planned replacements is instantly detected and is replaced with a new one. That is, the model also forms a renewal process with a renewal function MðtÞ between planned replacements. Then, because the expected number of failures in one planned replacement cycle is MðTÞ; the expected cost rate is, from (3.19), C1 ðTÞ ¼
c1 MðTÞ þ c2 ; T
ð3:44Þ
where c1 ¼ cost of replacement at each failure and c2 ¼ cost of planned replacement at time T: Clearly, from (3.18), limT!1 C1 ðTÞ ¼ c1 =l: Compare the expected costs with age and block replacements: Letting AðTÞ c1 FðTÞ þ c2 FðTÞ and BðTÞ c1 MðTÞ þ c2 ; we have the renewal equation (Problem 3.13) BðTÞ ¼ AðTÞ þ
ZT
BðT tÞdFðtÞ;
or
BðTÞ ¼ AðTÞ þ
0
ZT
AðT tÞdMðtÞ;
0
i.e., AðTÞ and BðTÞ determine each other. We seek an optimum planned replacement time T that minimizes C1 ðTÞ in (3.44). It is assumed that MðtÞ is differentiable and define mðtÞ dMðtÞ=dt: Then, differentiating C1 ðTÞ with respect to T and setting it equal to zero, TmðTÞ MðTÞ ¼
c2 : c1
ð3:45Þ
This is a necessary condition that there exists a finite T ; and the resulting cost rate is C1 ðT Þ ¼ c1 mðT Þ:
72
3 Renewal Processes
Next, we consider the periodic replacement where each failed unit between replacements undergo minimal repair, as shown in Sect. 2.4.1. It is assumed that the repair times are negligible and the failure rate remains undisturbed by any minimal repair. That is, the replacement model also forms a nonhomogeneous Poisson process with a cumulative hazard function HðtÞ between planned replacements, where FðtÞ ¼ 1 eHðtÞ and H 0 ðtÞ ¼ hðtÞ: Then, because the expected number of failures in one planned replacement cycle is HðTÞ from (2.45), the expected cost rate is, from (3.19), C2 ðTÞ ¼
c1 HðTÞ þ c2 ; T
ð3:46Þ
where c1 ¼ cost of minimal at each failure, and c2 is given in (3.44). When FðtÞ is IFR, C2 ðTÞ C1 ðTÞ for the same replacement costs from HðTÞ MðTÞ: We seek an optimum planned time T that minimizes C2 ðTÞ in (3.46). Differentiating C2 ðTÞ with respect to T and setting it equal to zero, c2 ThðTÞ HðTÞ ¼ ; ð3:47Þ c1 and the resulting cost rate is C2 ðT Þ ¼ c1 hðT Þ: It is shown that if the failure rate hðtÞ increases strictly, then the left-hand side of (3.47) also increases strictly [11, p. 102]. So that, if a solution T to (3.47) exists, then it is unique. Finally, we consider the periodic replacement where each failed unit remains in a failed state for the time interval from failure to its planned replacement. Then, the mean time from a failure to the replacement is ZT
ðT tÞdFðtÞ ¼
0
ZT FðtÞdt: 0
Thus, the expected cost rate is [11, p. 120] RT c1 0 FðtÞdt þ c2 C3 ðTÞ ¼ ; T
ð3:48Þ
where c1 ¼ downtime cost per unit of time elapsed between a failure and its replacement, and c2 is given in (3.44). Differentiating C3 ðTÞ with respect to T and setting it equal to zero, TFðTÞ
ZT 0
c2 FðtÞdt ¼ ; c1
ZT or
tdFðtÞ ¼
c2 : c1
ð3:49Þ
0
Thus, if l [ c2 =c1 ; then there exists an optimum time T3 that satisfies uniquely (3.49), and the resulting cost rate is
3.3 Age Replacement Policies
73
C3 ðT3 Þ ¼ c1 FðT Þ: In general, the results of three periodic replacements are summarized as follows: The expected cost rate is [11, p. 125] RT c1 0 uðtÞdt þ c2 ; ð3:50Þ CðTÞ ¼ T where uðtÞ is hðtÞ; mðtÞ; and FðtÞ; respectively. Differentiating CðTÞ with respect to T and setting it equal to zero, TuðTÞ
ZT
c2 uðtÞdt ¼ ; c1
0
ZT or
tduðtÞ ¼
c2 : c1
ð3:51Þ
0
If there exists T that satisfies (3.51), then the expected cost rate is CðT Þ ¼ c1 uðT Þ: Conversely, if a finite T does not exists, i.e., T ¼ 1; and the unit does not undergo any planned replacement, then Cð1Þ ¼ c1 uð1Þ:
3.4 Alternating Renewal Process Alternating renewal processes are the process that combines two renewal processes fXk g and fYk g generated alternately (Fig. 3.5) [7]. That is, each sequence of fXk g and fYk g forms a renewal process, and the sequence of the sum of fXk þ Yk g also forms a renewal process. In this case, Xk and Yk are allowed to be dependent. Most redundant systems with repair maintenance that repeat up and down states alternately, generate alternating renewal processes. Furthermore, the process is a special case of Markov renewal processes in Chap. 5.
3.4.1 Ordinary Alternating Renewal Process We consider a simple system with one unit that repeats on and off states. Suppose that fXk g and fYk g ðk ¼ 1; 2; . . .Þ are the respective on and off times of the kth renewal cycle and have the distribution FðtÞ PrfXk tg with finite mean l and Fig. 3.5 Realization of alternating renewal process
74
3 Renewal Processes
the distribution GðtÞ PrfYk tg with finite mean b: To analyze the system, we define the following states: State 0: Unit is on. State 1: Unit is off. When the unit starts to State 0 at time 0; the unit repeats States 0 and 1 alternately (Fig. 3.5). Let Mij ðtÞ ði; j ¼ 0; 1Þ denote the expected number of State j in ½0; t when the unit goes into State i at time 0; where the first visit to State j is not counted when i ¼ j: Then, we have the following renewal equations: M01 ðtÞ ¼ FðtÞ þ
Zt
M11 ðt uÞdFðuÞ;
0
M10 ðtÞ ¼ GðtÞ þ
Zt
M00 ðt uÞdGðuÞ;
ð3:52Þ
0
M11 ðtÞ ¼
Zt
M01 ðt uÞdGðuÞ;
M00 ðtÞ ¼
0
Zt
M10 ðt uÞdFðuÞ:
0
Thus, taking the LS transforms of both sides of the above equations and solving them, F ðsÞ G ðsÞ ; M ; ðsÞ ¼ 10 1 F ðsÞG ðsÞ 1 F ðsÞG ðsÞ ðsÞ ¼ G ðsÞM01 ðsÞ; M00 ðsÞ ¼ F ðsÞM10 ðsÞ; M11 M01 ðsÞ ¼
ð3:53Þ
where the R 1 asterisk of the function denotes the LS transform with itself, i.e., U ðsÞ 0 est dUðtÞ for any function UðtÞ: Next, let ZðtÞ denote the state of the process at time t; and define Pij ðtÞ PrfZðtÞ ¼ jjZð0Þ ¼ ig ði; j ¼ 0; 1Þ: The probability Pij ðtÞ is called the transition probability from State i to State j and represents the probability that the unit is in State j at time t if it starts from State i at time 0: By the similar method for obtaining (3.52), P00 ðtÞ ¼ FðtÞ þ
Zt
P10 ðt uÞdFðuÞ;
0
P11 ðtÞ ¼ GðtÞ þ
Zt
P01 ðt uÞdGðuÞ;
ð3:54Þ
0
P10 ðtÞ ¼
Zt 0
P00 ðt uÞdGðuÞ;
P01 ðtÞ ¼
Zt 0
P11 ðt uÞdFðuÞ:
3.4 Alternating Renewal Process
75
Thus, again taking the LS transforms, P00 ðsÞ ¼ P10 ðsÞ
1 F ðsÞ ; 1 F ðsÞG ðsÞ
¼G
ðsÞP00 ðsÞ;
P11 ðsÞ ¼ P01 ðsÞ
1 G ðsÞ ; 1 F ðsÞG ðsÞ
¼F
ð3:55Þ
ðsÞP11 ðsÞ:
Thus, comparing (3.53) and (3.55), we have the following relations: Zt
P01 ðtÞ ¼
Gðt uÞdM01 ðuÞ ¼ M01 ðtÞ M00 ðtÞ;
0
Zt
P10 ðtÞ ¼
Fðt uÞdM10 ðuÞ ¼ M10 ðtÞ M11 ðtÞ:
0
These relations between renewal functions and transition probabilities would be useful for the analysis of more complex systems. Next, it is assumed that distributions FðtÞ and GðtÞ have finite variances rF 2 and rG 2 ; respectively. Then, 1 F ðsÞ ¼ 1 ls þ l2 þ r2F s2 þ oðs2 Þ; 2 1 G ðsÞ ¼ 1 bs þ b2 þ r2G s2 þ oðs2 Þ: 2 Substituting these equations for (3.53), M01 ðsÞ ¼
1 1 l 1 r2 þ r2G þ oð1Þ: þ þ F l þ b s l þ b 2 2ðl þ bÞ2
ðsÞ gives, for large t; The LS inversion of M01
M01 ðtÞ ¼
t l 1 r2 þ r2G þ oð1Þ: þ þ F l þ b l þ b 2 2ðl þ bÞ2
M00 ðtÞ ¼
t 1 r2 þ r2G þ F þ oð1Þ; l þ b 2 2ðl þ bÞ2
P00 ðtÞ ¼
l þ oð1Þ; lþb
ð3:56Þ
Similarly,
P01 ðtÞ ¼
b þ oð1Þ: lþb
ð3:57Þ
Clearly, lim
t!1
Mij ðtÞ 1 ¼ t lþb
ðj ¼ 0; 1Þ;
l EfXk g lim Pi0 ðtÞ ¼ 1 lim Pi1 ðtÞ ¼ ¼ ; t!1 t!1 l þ b EfXk g þ EfYk g
ð3:58Þ
76
3 Renewal Processes
that are independent of an initial State i in the limit. Note that (3.58) represents the steady-state availability of a one-unit system [11, p. 9]. Example 3.6 (One-unit system with repair) Suppose that an operating unit is repaired when it fails, and it begins to operate again. In this case, States 0 and 1 represent that the unit is operating and under repair, respectively. It is assumed that the failure time has an exponential distribution FðtÞ ¼ 1 ekt and the repair time has an exponential distribution GðtÞ ¼ 1 eht : Then, it is easy to invert the LS transforms as follows: i k h P01 ðtÞ ¼ 1 eðkþhÞt ; kþh 2 h i kht k 1 eðkþhÞt : þ M01 ðtÞ ¼ kþh kþh Thus, for large t; P01 ðtÞ
k ; kþh
M01 ðtÞ
2 kh k : tþ kþh kþh
Next, when FðtÞ ¼ 1 ð1 þ ktÞekt and the time for repair is constant b; P01 ðsÞ ¼
k2 ð1 ebs Þ
;
M01 ðsÞ ¼
b ; 2=k þ b
M01 ðtÞ
2
ðs þ kÞ
k2 ebs
k2 2
ðs þ kÞ k2 ebs
;
and for large t; P01 ðtÞ
t : 2=k þ b
3.4.2 Interval Reliability Interval reliability is defined as the probability that at a specified time, the unit is operating and will continue to operate for an interval duration [1, p. 7, 11, p. 11]. Then, the interval reliability Rðx; T0 Þ for an interval of duration x starting at time T0 is Rðx; T0 Þ PrfZðuÞ ¼ 0; T0 u T0 þ xg;
ð3:59Þ
and its limit of Rðx; T0 Þ as T0 ! 1 is called the limiting interval reliability. When the unit begins to operate at time 0 in Example 3.6, the interval reliability is
h k ðkþhÞT0 kx kx Rðx; T0 Þ P00 ðT0 Þe e ; þ e ¼ ð3:60Þ kþh kþh
3.4 Alternating Renewal Process
77
and its limiting interval reliability is RðxÞ lim Rðx; T0 Þ ¼ T0 !1
h kx e : kþh
ð3:61Þ
Suppose that the unit has a failure distribution FðtÞ with mean l and a repair distribution GðtÞ with mean b in Sect. 3.4.1. When the unit begins to operate at time 0; the interval reliability is [11, p. 48] Rðx; T0 Þ ¼ FðT0 þ xÞ þ
ZT0
FðT0 þ x uÞdM00 ðuÞ;
ð3:62Þ
0
where M00 ðtÞ is given in (3.53). Hence, its Laplace transform is R1 Z1 esx x est FðtÞdt ; esT0 Rðx; T0 ÞdT0 ¼ R ðx; sÞ 1 F ðsÞG ðsÞ
ð3:63Þ
0
and its limiting interval reliability is
RðxÞ lim Rðx; T0 Þ ¼ lim sR ðx; sÞ ¼ T0 !1
s!0
R1 x
FðtÞdt : lþb
ð3:64Þ
Next, suppose that we set the preventive maintenance (PM) at time T ð0\T 1Þ for the operating unit [11, p. 141]. However, the PM is not done in the interval ½T0 ; T0 þ x even if it is time for PM. It is assumed that the distribution of time to the PM completion is the same as the repair distribution GðtÞ: Let DðtÞ be a degenerate distribution placing unit mass at T; i.e., DðtÞ 0 for t\T and 1 for t T: Then, in a similar way of obtaining (3.52), M00 ðtÞ ¼
Zt
M10 ðt uÞ½DðuÞdFðuÞ þ FðuÞdDðuÞ;
0
M10 ðtÞ ¼
Zt
½1 þ M00 ðt uÞdGðuÞ:
0
Taking the LS transforms,
h i RT G ðsÞ 1 0 sest FðtÞdt h i: ðsÞ ¼ M00 RT 1 G ðsÞ 1 0 sest FðtÞdt
ð3:65Þ
By the similar method of obtaining (3.62), the interval reliability is RðT; x; T0 Þ ¼ FðT0 þ xÞDðT0 Þ þ
ZT0 0
FðT0 þ x uÞDðT0 uÞdM00 ðuÞ:
ð3:66Þ
78
3 Renewal Processes
Taking the Laplace transform of (3.66) and substituting it for (3.65),
R ðT; x; sÞ
Z1
esT0 RðT; x; T0 ÞdT0
0
ð3:67Þ
R Tþx
est FðtÞdt x ; ¼ RT 1 G ðsÞ þ G ðsÞ 0 sest FðtÞdt esx
that agrees with (3.63) when T ¼ 1 (Problem 3.14). Thus, the limiting interval reliability is RðT; xÞ lim RðT; x; T0 Þ ¼ lim sR ðT; x; sÞ T0 !1 s!0 R Tþx FðtÞdt ¼ R Tx ; 0 FðtÞdt þ b
ð3:68Þ
that also agrees with (3.64) when T ¼ 1 (Problem 3.15). We seek an optimum PM time T that maximizes RðT; xÞ for a fixed x [ 0: Let kðt; xÞ ½Fðt þ xÞ FðtÞ=FðtÞ for t 0 and kð1; xÞ limt!1 kðt; xÞ 1: Then, note that both kðt; xÞ and the failure rate hðtÞ f ðtÞ=FðtÞ have the same property [1, p. 23, 11, p. 6]. It is assumed that kðt; xÞ increases strictly with t: Then, differentiating RðT; xÞ in (3.68) with respect to T and setting it equal to zero, 2 T 3 Z ZT kðT; xÞ4 FðtÞdt þ b5 ½Fðt þ xÞ FðtÞdt ¼ b; ð3:69Þ 0
0
whose left-hand side QðT; xÞ increases strictly from Qð0Þ ¼ bFðxÞ to Qð1; xÞ ¼ kð1; xÞðl þ bÞ
Zx FðtÞdt: 0
Therefore, if Qð1; xÞ [ b; then there exists a finite and unique T ð0\T \1Þ that satisfies (3.69), and the resulting interval reliability is RðT ; xÞ ¼
FðT þ xÞ ; FðT Þ
that represents the probability that the unit with age T does not fail in ½T ; T þ x: Example 3.7 When FðtÞ ¼ 1 ð1 þ ktÞekt ; kx ekx ; kðt; xÞ ¼ 1 1 þ 1 þ kt
3.4 Alternating Renewal Process
79
that increases strictly with t to ð1 ekx Þ: Thus, if Q ð1; xÞ [ b; i.e., x [ b; an optimum T is a unique solution of the equation kTðx bÞ xð1 ekT Þ ¼ ð1 þ kxÞb; and the interval reliability is RðT ; xÞ ¼
1þ
kx ekx : 1 þ kT
In this case, an optimum T decreases with x to a unique solution of T
1 ekT ¼ b: k
Conversely, if x b; then T ¼ 1 and the interval reliability in (3.68) is given in (3.64).
3.4.3 Off Distribution Let NðtÞ and DðtÞ be the number of on states and the total amount of time spent in off states in ½0; t; respectively. It is said in reliability theory that DðtÞ is called a downtime [1, p. 79, 11, p. 45]. Then, from the formula of the total sum of independent random variables, PrfY1 þ Y2 þ þ Yn xjNðtÞ ¼ ng PrfNðtÞ ¼ ng ¼ PrfY1 þ Y2 þ þ Yn xg PrfX1 þ þ Xn t x\X1 þ þ Xnþ1 g h i ¼ GðnÞ ðxÞ F ðnÞ ðt xÞ F ðnþ1Þ ðt xÞ : Thus, the downtime distribution is
P1 ðnÞ ðnÞ ðnþ1Þ ðt xÞ n¼0 G ðxÞ½F ðt xÞ F PrfDðtÞ xg ¼ 1
for t [ x; for t x: ð3:70Þ
The above equation can be written as PrfDðt þ xÞ xg ¼
1 X n¼0
(P PrfDðtÞ [ xg ¼
0
h i GðnÞ ðxÞ F ðnÞ ðtÞ F ðnþ1Þ ðtÞ ;
1 ðnÞ n¼0 G ðxÞ
Gðnþ1Þ ðxÞ F ðnþ1Þ ðt xÞ
for t [ x; for t x:
Next, we investigate the properties of DðtÞ: Denote Ld mint fDðtÞ [ dg as the first time that the total time spent in off states, i.e., the total downtime, exceeds a threshold level d: We have the following relation:
80
3 Renewal Processes
PrfLd tg ¼ PrfDðtÞ [ dg 1 h i X ¼ GðnÞ ðdÞ Gðnþ1Þ ðdÞ F ðnþ1Þ ðt dÞ:
ð3:71Þ
n¼0
Thus, the mean time that the total downtime first exceeds d is EfLd g ¼ d þ l
1 X
GðnÞ ðdÞ:
ð3:72Þ
n¼0
The mean downtime in ½0; t is EfDðtÞg ¼
Zt
xd PrfDðtÞ xg:
ð3:73Þ
0
In general, it is very difficult to compute (3.73). It is noted easily from Sect. 3.4.1 that EfDðtÞg ¼
Zt
P01 ðuÞdu;
ð3:74Þ
0
and its LS transform is Z1
est P01 ðtÞdt ¼
F ðsÞ½1 G ðsÞ : s½1 F ðsÞG ðsÞ
ð3:75Þ
0
Thus, from (3.57) EfDðtÞg lim ¼ lim P01 ðtÞ ¼ lim s t!1 t!1 s!0 t
Z1
est P01 ðtÞdt ¼
b : lþb
ð3:76Þ
0
Example 3.8 (One-unit system with repair) We consider the same one-unit system in Example 3.6. The downtime distribution is [1, p. 80] PrfDðtÞ [ xg ¼
1 X ðkxÞn n¼0
n!
ekx
1 X ½hðt xÞk hðtxÞ e ; k! k¼nþ1
or 2
3 Zx pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ffi PrfDðtÞ xg ¼ ehðtxÞ41þ khðtxÞ eky y1=2 I1 2 khðtxÞy dy5; 0
3.4 Alternating Renewal Process
81
where I1 ðxÞ is the Bessel function of order 1 for the imaginary argument defined by I1 ðxÞ
1 X ðx=2Þ2jþ1 j¼0
j!ðj þ 1Þ!
:
In addition, from (3.72) and (3.74), 1 EfLd g ¼ d þ ð1 þ hdÞ; k i k k h t 1 eðkþhÞt : EfDðtÞg ¼ 2 kþh ðk þ hÞ Takács [10] also proved the following important theorem: Theorem 3.8 If r2F \1 and r2G \1; then 9 8 > > > > Zx = < DðtÞ bt=ðl þ bÞ 1 2 r ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi p ffiffiffiffiffi ffi ¼ x eu =2 du: lim Pr h i t!1 > > 2p > > ; : ðbrF Þ2 þðlrG Þ2 t=ðl þ bÞ3 1
ð3:77Þ
That is, if the means and variances of FðtÞ and GðxÞ are statistically estimated, then the probability of the total time spent in off states, i.e., downtime DðtÞ is approximately computed for large t; by using a standard normal distribution. Example 3.9 (Asymptotic downtime distribution) [11, p. 50] We wish to compute the time d when the probability that the system is down more than d in t ¼ 10; 000 hours of operation is given by 0.90. The failure and repair distributions are unknown, but from the sample data, the estimates of means and variances are l ¼ 1; 000;
rF 2 ¼ 100; 000;
b ¼ 100;
rG 2 ¼ 400:
Then, from Theorem 3.8, when t ¼ 10; 000; DðtÞ bt=ðl þ bÞ Dð10; 000Þ 909:09 rhffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ i 102:56 ðbrF Þ2 þðlrG Þ2 t=ðl þ bÞ3 is approximately normally distributed with mean 0 and variance 1: Hence,
Dð10; 000Þ 909:09 d 909:09 [ PrfDðtÞ [ dg ¼ Pr 102:56 102:56 Z1 1 2
pffiffiffiffiffiffi eu =2 du ¼ 0:90: 2p ðd909:09Þ=102:56
82
3 Renewal Processes
pffiffiffiffiffiffi R 1 2 Because u0 ¼ 1:28 such that ð1= 2pÞ u eu =2 du0 ¼ 0:90; d 909:09 102:56 1:28 ¼ 777:81; i.e., the system is down more than 778 h in t ¼ 10; 000 h with probability 0:90 (Problem 3.16).
3.4.4 Terminating Renewal Process Consider the first time that an amount of the time spent in off states exceeds a ek Yk ; threshold level U with distribution WðtÞ: We define that if Yk U then Y e and conversely, if U\Yk then U k U ðk ¼ 1; 2; . . .Þ: The process ends with the first event of fU\YN g that is called the terminating process e 1 ; X2 ; Y e2 ; . . .; XN1 ; Y eN1 ; XN ; U e g; and its interest quantities are the total fX1 ; Y PN1 N ek Þ þ XN þ U e N ðN ¼ 1; 2; . . .Þ and its sum of random variables ZN k¼1 ðXk þ Y P1 distribution LðtÞ N¼1 PrfZN tg: The probability that Yk is not greater than U and Yk t is BðtÞ PrfYk U; Yk tg ¼
Zt WðuÞdGðuÞ; 0
and the probability that Yk is greater than U and U t is e PrfU\Yk ; U tg ¼ BðtÞ
Zt GðuÞdWðuÞ: 0
Thus, from the formula of the total sum of independent random variables, LðtÞ ¼
1 X
PrfZN tg ¼
N¼1
1 X
e F ðkÞ ðtÞ BðkÞ ðtÞ FðtÞ BðtÞ:
ð3:78Þ
k¼0
Therefore, the LS transform of LðtÞ is
L ðsÞ
Z1
est dLðtÞ ¼
e ðsÞ F ðsÞ B ; 1 F ðsÞB ðsÞ
ð3:79Þ
0
and its mean time is l
Z1 0
R1 1 L ðsÞ l þ 0 GðtÞWðtÞdt ¼ R1 : tdLðtÞ ¼ lim s!0 s 0 WðtÞdGðtÞ
ð3:80Þ
3.4 Alternating Renewal Process
83
In particular, when a random variable U is constant cðc [ 0Þ; i.e., WðtÞ 0 for t\c; and 1 for t c; Rc l þ 0 GðtÞdt F ðsÞesc GðcÞ Rc L ðsÞ ¼ ; l¼ ; ð3:81Þ 1 F ðsÞ 0 est dGðtÞ GðcÞ where c is called a critical allowed time for maintenance [15]. Example 3.10 (Repair limit policy) [11, p. 51] Consider a one-unit system that is repaired or replaced if it fails. Let l denote the mean failure time of the unit and GðtÞ denote the repair distribution of the failed unit with finite mean b: When the unit fails, its repair is started immediately, and when the repair is not completed within time T ð0 T\1Þ; that is called the repair limit time, it is replaced with a new one. If Zk minfYk ; Tg, then the sequence of fXk ; Zk g ðk ¼ 1; 2; . . .Þ forms an alternating renewal process. Let c1 be the replacement cost of a failed unit, caused by failure and replacement, and cr ðtÞ be the expected repair cost during ½0; t; that includes all costs incurred due to repair and downtime during ½0; t: Consider one cycle from the beginning of an operating unit to the repair or replacement completion. Then, the expected cost of one cycle is ½c1 þ cr ðTÞGðTÞ þ
ZT
cr ðtÞdGðtÞ ¼ c1 GðTÞ þ
ZT
0
GðtÞdcr ðtÞ;
0
and the mean time of one cycle is l þ TGðTÞ þ
ZT
tdGðtÞ ¼ l þ
0
ZT GðtÞdt: 0
Thus, from (3.19), the expected cost rate is RT c1 GðTÞ þ 0 GðtÞdcr ðtÞ CðTÞ ¼ : RT l þ 0 GðtÞdt
ð3:82Þ
Clearly, c1 Cð0Þ lim CðTÞ ¼ ; T!0 l
R1 Cð1Þ lim CðTÞ ¼ T!1
0
GðtÞdcr ðtÞ ; lþb
that represent the expected cost rates with only replacement and only repair maintenance, respectively. Suppose that cr ðtÞ ¼ at for a [ 0: Then, the expected cost rate is, from (3.82), RT c1 GðTÞ þ a 0 GðtÞdt CðTÞ ¼ : ð3:83Þ RT l þ 0 GðtÞdt
84
3 Renewal Processes
We find an optimum repair limit time T that minimizes CðTÞ: It is assumed that rðtÞ gðtÞ=GðtÞ; where gðtÞ is a density function of GðtÞ: Then, differentiating CðTÞ with respect to T and setting it equal to zero, 2 3 ZT al ð3:84Þ rðTÞ4l þ GðtÞdt5 þ GðTÞ ¼ : c1 0
Letting QðTÞ be the left-hand side of (3.84), Qð0Þ lrð0Þ þ 1;
Qð1Þ ¼ ðl þ bÞrð1Þ:
Thus, if rðtÞ decreases strictly and Qð0Þ [ al=c1 [ Qð1Þ; there exists uniquely a finite and positive T ð0\T \1Þ that satisfies (3.84), and the resulting cost rate is C ðT Þ ¼ a c1 r ðT Þ: If Qð0Þ al=c1 then T ¼ 0; i.e., no repair should be done. If Qð1Þ al=c1 then T ¼ 1; i.e., no replacement should be done (Problem 3.17).
3.5 Modified Renewal Processes This section takes up two modified renewal processes.
3.5.1 Geometric Renewal Process Suppose that a renewal process is defined by a sequence of random variables fX1 ; X2 =a; . . .; Xk =ak1 ; . . .g for a [ 0 with distribution Fðak1 tÞ PrfXk =ak1 tg; where FðtÞ has finite mean l and variance r2 [16]. It is shown in a one-unit system that each operating time decreases geometrically with the number of repairs for a [ 1: Then, the mean time and variance are
Xk l Xk r2 E k1 ¼ k1 ; V k1 ¼ 2ðk1Þ : a a a a Pn When Sn k¼1 Xk =ak1 ; r2 a2 a2ðn1Þ lða anþ1 Þ EfSn g ¼ ; ; V fSn g ¼ a2 1 a1 and for a [ 1; EfSn g !
la ; a1
V fSn g !
r2 a 2 a2 1
as n ! 1:
3.5 Modified Renewal Processes
85
Furthermore, a renewal function Mðt; aÞ Zt
Mðt; aÞ ¼ FðtÞ þ
P1
n¼1
PrfSn tg is, from (3.5),
M ðaðt uÞ; aÞdFðuÞ;
ð3:85Þ
0
because PrfSn tg ¼
Zt
PrfSn1 aðt uÞgdFðuÞ
ðn ¼ 1; 2; . . .Þ:
0
The distributions of the age dðtÞ and the residual life cðtÞ at time t in (3.22) and (3.23) are rewritten as, respectively, 1 Z t X PrfcðtÞ xg ¼ Fðt þ xÞ F ðan ðt þ x uÞÞdF ðnÞ ðuÞ;
PrfdðtÞ xg ¼
8 <
n¼1
FðtÞ
:
0
X1 Z n¼1
tx
Fðan ðt uÞÞdF ðnÞ ðuÞ
for x t;
1
for x [ t:
:
0
ð3:86Þ In general, it is very difficult to derive Mðt; aÞ explicitly even if FðtÞ is exponential, and so, there is nothing but to compute it numerically. For example, when FðtÞ ¼ 1 ekt [17, p. 156] (Problem 3.18), " # n n X Y 1 F ðnÞ ðtÞ ¼ 1 ð3:87Þ exp ak1 kt ðn ¼ 1; 2; . . .Þ: kj 1 a k¼1 j¼1;j6¼k The renewal process with interarrival times fX1 ; aX1 ; a2 X1 ; . . .g is also called the quasi-renewal process [18, p. 53]. Example 3.11 (Replacement policy) [15, p. 201] It is assumed in Sect. 2.4.1 that the time between the ðk 1Þth and kth failures is denoted by Xk =ak1 and its mean time is l=ak1 ðk ¼ 1; 2; . . .Þ for a [ 1: Then, the expected cost rate is (Problem 3.19) C1 ðNÞ ¼
c1 ðN 1Þ þ c2 ðN ¼ 1; 2; . . .Þ; lða aNþ1 Þ=ða 1Þ
where costs c1 and c2 C1 ðN þ 1Þ C1 ðNÞ 0;
are
given
aNþ1 a c2 ðN 1Þ c1 a1
in
(3.44).
From
ðN ¼ 1; 2; . . .Þ:
ð3:88Þ the
inequality
ð3:89Þ
86
3 Renewal Processes
It is easily seen that the left-hand side of (3.89) increases strictly from a to 1: Thus, there exists a finite and unique N ð1 N \1Þ that satisfies (3.89). If a c2 =c1 ; then N ¼ 1: In addition, because the left-hand side of (3.89) also increases with a; N decreases with a. Next, consider an alternating renewal process defined by two sequences of fXk =ak1 g and fYk =bk1 g for a; b [ 0: Example 3.12 (One-unit system) Consider a one-unit system in Example 3.6 where the failure time for the kth operation is Xk =ak1 with mean time l=ak1 for a 1 and the repair time for the kth repair is Yk =bk1 with mean time b=bk1 for 0\b 1: That is, the mean failure times decrease with their number, and conversely, the mean repair times increase with their number. Suppose that the unit is replaced with a new one at the Nth failure. Then, the expected cost rate is (Problem 3.20) c1 ðN 1Þ þ c2 ðN ¼ 1; 2; . . .Þ; PN1 k1 Þ þ k1 Þ k¼1 ðl=a k¼1 ðb=b
C2 ðNÞ ¼ PN
ð3:90Þ
where c1 ¼ cost of repair and c2 ¼ replacement cost at the Nth failure. Then, from C2 ðN þ 1Þ C2 ðNÞ 0; PN1 PN k1 þ k¼1 b=bk1 c2 k¼1 l=a ðN 1Þ ðN ¼ 1; 2; . . .Þ; ð3:91Þ N N1 c1 l=a þ b=b P where 01 0: Thus, if l b l b þ k1 [ kþ1 þ k ; k a b a b i.e., the mean times between failures decrease, then there exists a finite and unique N ð1 N \1Þ that satisfies (3.91). If l=ðl=a þ bÞ c2 =c1 ; then N ¼ 1:
3.5.2 Discrete Renewal Process Until now, it has been assumed in a renewal process that the sequence of random variables fXk g has an identical continuous distribution FðtÞ: However, the times to some event in actual fields are often measured by the number of cycles, might not be recorded at its exact instant time and be observed statistically per day, per month, per year, and so on. In failure studies, the failure times of units such as tires of jet fighters, switching devices, railroad tracks, and ball bearings might be measured discretely. In general, the conversion from continuous models to discrete ones is possible. However, the results are often too complicated to be useful, unless skillful treatment is used. This section simply shows how to make the conversion.
3.5 Modified Renewal Processes
87
Consider the time over an infinitely long cycle j ðj ¼ 1; 2; . . .Þ and renewal cycle k ðk ¼ 1; 2; . . .Þ: Suppose that Xk has an identical probability function pj ðj ¼ 1; 2; . . .Þ with finite mean l and a probability distribution Pj ; i.e., pj P PrfXk ¼ jg; Pj ji¼1 pi ðj ¼ 1; 2; . . .Þ; and 1 X
l
jpj ¼
j¼1
1 X ð1 Pj Þ\1; j¼1
where p0 P0 0: The generating function of a probability function is defined as P j p ðzÞ 1 j¼1 z pj for jzj 1 from Appendix A.4, and hence, P ðzÞ
1 X
zj Pj ¼
j¼1
Let the total sum of random variables Sn ðnÞ pj
PrfSn ¼ jg; where ð2Þ
ð1Þ pj
¼ pj and
ð0Þ pj
p ðzÞ : 1z
Pn
k¼1
ð3:92Þ
Xk and its probability function
1: Then,
¼ PrfS2 ¼ jg ¼ PrfX1 þ X2 ¼ jg
pj
¼
j X
PrfX2 ¼ j ijX1 ¼ ig PrfX1 ¼ ig ¼
i¼1
j X
pji pi ;
i¼1
and its generating function is 1 X
ð2Þ
z j pj
¼
j¼1
j 1 X X zj pji pi ¼ ½p ðzÞ2 : j¼1
i¼1
Similarly, by using the mathematical induction, 1 X
ðnÞ
z j pj
¼ ½p ðzÞn
ðn ¼ 0; 1; 2; . . .Þ:
ð3:93Þ
j¼1 ðnÞ
The generating function of Pj 1 X
ðnÞ
zj Pj
¼
j¼1
ð0Þ
Prf Sn jg; where Pj
j 1 X X ½p ðzÞn ðnÞ zj pi ¼ 1z j¼1 i¼1
1; is
ðn ¼ 0; 1; 2; . . .Þ:
Thus, the probability function of Nj maxn f Sn jg is Pr Nj ¼ 0 ¼ Prf X1 [ jg ¼ 1 Pj ; Pr Nj ¼ n ¼ Prf Sn j and Snþ1 [ jg ¼ Prf Sn jg Prf Snþ1 jg ¼
ðnÞ Pj
ðnþ1Þ Pj
ð3:94Þ
ð3:95Þ
ðn ¼ 0; 1; 2; . . .Þ: The renewal function Mj E Nj ðj ¼ 1; 2; . . .Þ that represents the expected number of renewals in cycle j is
88
3 Renewal Processes
Mj ¼
1 X
1 X ðnÞ n Pr Nj ¼ n ¼ Pj
n¼1
¼ Pj þ
n¼1 j X
ð3:96Þ
ðj ¼ 1; 2; . . .Þ;
pi Mji
i¼1
and its generating function is, from (3.92), M ðzÞ
1 X
zj Mj ¼
j¼1
P ðzÞ 1 p ðzÞ ¼ : 1 p ðzÞ 1 z 1 p ðzÞ
ð3:97Þ
In particular, when pj ¼ pqj1 ðj ¼ 1; 2; . . .Þ; where q 1 p; p ðzÞ ¼
pz ; 1 qz
P ðzÞ ¼
1 pz ; 1 z 1 qz
and hence, from (3.97), M ðzÞ ¼
pz ð1 zÞ
2
¼p
1 X
jzj ;
i.e.,
Mj ¼ jp ¼
j¼1
j ; E fX k g
that is the same result as the exponential case of the continuous renewal process shown in Example 3.2. Example 3.13 (Age replacement policy) [11, p. 80] Consider the unit with a failure probability function pj that should be operating for the time over an infinitely long cycle j ðj ¼ 1; 2; . . .Þ: The unit is replaced at a planned cycle N ðN ¼ 1; 2; . . .Þ after its installation or at failure, whichever occurs first. Letting Zk minfXk ; N g ðk ¼ 1; 2; . . .Þ; a sequence of fZk g1 k¼1 forms a discrete renewal process with an independent and identical distribution
Pj for j N; i¼1 pi Prf Zk jg ¼ 1 for j [ N; and its mean time is E fZk g ¼
N X j¼1
jpj þ N
1 X j¼Nþ1
pj ¼
N X 1 X
pj :
i¼1 j¼i
Thus, by the similar method of obtaining (3.42), the expected cost rate is c1 PrfXk N g þ c2 PrfXk [ N g EfZk g P P c1 Nj¼1 pj þ c2 1 j¼Nþ1 pj ðN ¼ 1; 2; . . .Þ; ¼ PN P1 i¼1 j¼i pj
CðNÞ ¼
where replacement costs are given in (3.42).
ð3:98Þ
3.5 Modified Renewal Processes
89
We seek an optimum replacement cycle N that minimizes CðNÞ: Let hj P pj = 1 i¼j pi ðj ¼ 1; 2; . . .Þ be the discrete failure rate of the probability function pj with finite mean l: Suppose that hj increases strictly with j: From the inequality CðN þ 1Þ CðNÞ 0;
hNþ1
N X 1 X i¼1 j¼i
pj
N X
pj
j¼1
c2 c1 c 2
ðN ¼ 1; 2; . . .Þ:
ð3:99Þ
Letting LðNÞ be the right-hand side of (3.99),
LðN þ 1Þ LðNÞ ¼ ðhNþ2 hNþ1 Þ
N þ1 X 1 X
pj [ 0;
i¼1 j¼i
Lð1Þ lim LðNÞ ¼ lh1 1: N!1
Therefore, if Lð1Þ [ c2 =ðc1 c2 Þ; i.e., h1 [ K; where K c1 =½lðc1 c2 Þ; then there exists a finite unique minimum N ð1 N \1Þ that satisfies (3.99), and the resulting cost rate is ðc1 c2 ÞhN CðN Þ\ðc1 c2 ÞhN þ1 : Conversely, if h1 K; then N ¼ 1; i.e., no planned replacement should be done. Note that 0\hj 1 from the definition of the failure rate hj : Thus, if K 1; then N ¼ 1: In particular, assume that the failure distribution is a negative binomial distribution pj ¼ jp2 qj1 ðj ¼ 1; 2; . . .Þ; where q 1 p ð0\p\1Þ: Then, l ¼ ð1 þ qÞ=p and hj ¼ jp2 =ðjp þ qÞ that increases strictly from p2 to p: Therefore, from the above results, if c1 q [ c2 ð1 þ qÞ; an optimum N ð1 N \1Þ is given by a finite and unique minimum that satisfies ðN þ 1Þpð1 þ qÞ þ qNþ2 c1 : Np þ 1 c1 c 2 For example, when c1 ¼ 10; c2 ¼ 1; and l ¼ 9; i.e., p ¼ 1=5; N ¼ 4 and Cð4Þ ¼ 0:92: In this case, Cð1Þ ¼ c1 =l ¼ 1:11: Consider the discrete alternating renewal process with fXk g and fYk g that have the respective probability functions pj PrfXk ¼ jg and qj PrfYk ¼ jg; Pj Pj Pj i¼1 pi ; and Qj i¼1 qi : Other assumptions are the same as those of Sect. 3.4. Let Mij ðnÞ ði; j ¼ 0; 1Þ denote the expected number of visits to State j in n cycles, starting from State i at cycle 1; and its generating function is Mij ðzÞ ¼ P1 n n¼1 z Mij ðnÞ for jzj\1: Then, the results in Sect. 3.4 are rewritten as
90
3 Renewal Processes
M01 ðnÞ ¼ Pn þ
n X
pk M11 ðn kÞ;
M10 ðnÞ ¼ Qn þ
k¼1
M11 ðnÞ ¼
n X
n X
qk M00 ðn kÞ;
k¼1
qk M01 ðn kÞ;
M00 ðnÞ ¼
k¼1
n X
pk M10 ðn kÞ:
k¼1
Forming the generating functions as the above equations and solving them for Mij ðzÞ; ðzÞ ¼ p ðzÞM10 ðzÞ; M00 M11 ðzÞ ¼ q ðzÞM01 ðzÞ;
1 p ðzÞ ; 1 z 1 p ðzÞq ðzÞ 1 q ðzÞ : M10 ðzÞ ¼ 1 z 1 p ðzÞq ðzÞ M01 ðzÞ ¼
ð3:100Þ
Next, let ZðnÞ denote the state of the process at cycle n; and define Pij ðnÞ ¼ PrfZðnÞ ¼ jjZð1Þ ¼ ig: Then, from (3.54), the results in continuous time are rewritten as n X P00 ðnÞ ¼ 1 Pn þ pk P10 ðn kÞ; k¼1
P11 ðnÞ ¼ 1 Qn þ
n X
qk P01 ðn kÞ;
k¼1
P10 ðnÞ ¼
n X
qk P00 ðn kÞ;
P01 ðnÞ ¼
k¼1
P10 ðzÞ ¼ q ðzÞP00 ðzÞ;
pk P11 ðn kÞ;
k¼1
and their generating functions Pij ðzÞ P01 ðzÞ ¼ p ðzÞP11 ðzÞ;
n X
P1
n¼1
zn Pij ðnÞ are
1 1 p ðzÞ ; 1 z 1 p ðzÞq ðzÞ 1 1 q ðzÞ : P11 ðzÞ ¼ 1 z 1 p ðzÞq ðzÞ P00 ðzÞ ¼
ð3:101Þ
Clearly, P01 ðnÞ ¼ M01 ðnÞ M00 ðnÞ;
P10 ðnÞ ¼ M10 ðnÞ M11 ðnÞ:
Example 3.14 (One-unit system with repair) Consider a one-unit system with repair maintenance whose failure and repair distributions are geometric, i.e., pk ¼ pqk1 and qk ¼ abk1 ðk ¼ 1; 2; . . .Þ; where p þ q ¼ 1 and a þ b ¼ 1: Then, the results in Example 3.6 are rewritten as pz ; ð1 zÞ½1 ðb pÞz pz 1 bz M01 ðzÞ ¼ : 1 z ð1 zÞ½1 ðb pÞz P01 ðzÞ ¼
3.5 Modified Renewal Processes
91
Inverting the generating functions for 1 [ b p [ 0 (Problem 3.21), p ½1 ðb pÞn ; pþa
p p½1 ðb pÞn ; na þ M01 ðnÞ ¼ pþa pþa P01 ðnÞ ¼
and as n ! 1; P01 ðnÞ !
p ; pþa
M01 ðnÞ pa ! : n pþa
3.6 Problems 3 3.1 Prove that MðtÞ\1 for all t\1 [6, p. 57, 9, p. 287]. 3.2 In Example 3.2, show that when k ¼ 3 and k ¼ 4 [4, p. 162], respectively, pffiffiffi
3 1 2 1:5kt p p ffiffi ffi sin kt þ ; e MðtÞ ¼ kt 1 þ 2 3 3 3
pffiffiffi 1 3 1 p : MðtÞ ¼ kt þ e2kt þ 2ekt sin kt þ 4 2 2 4 3.3 Prove that when X1 and X2 have a gamma distribution with the respective parameters ða; kÞ and ðb; kÞ; X X1 þ X2 has a gamma distribution ða þ b; kÞ: a 3.4 When FðtÞ ¼ 1 eðktÞ ; show that ( 2 ) 1 1 1 2 1 EfXg ¼ C 1 þ ; VfXg ¼ 2 C 1 þ C 1þ ; k a a a k and using the discretization algorithm, compute MðtÞ for a = 1, 1.5, 2.0, 2.5, 3.0 and t = 1, 2, ......, 10 when k = 1, and compare them with t/E{X} given by the approximation of an elementary renewal theorem in Theorm 2.3 3.5 For a two-unit parallel system with an exponential failure distribution ð1 ekt Þ; show that the failure distribution of the system is FðtÞ ¼ ð1 ekt Þ2 and a renewal function is [24, p. 161]
2 1 3kt kt ð1 e Þ : MðtÞ ¼ 3 3 3.6 When FðtÞ ¼ 1 ð1 þ ktÞekt ; derive VfNðtÞg and (3.15) in Theorem 3.1. 3.7 In Example 3.3, compute the number n of units when the probability that the number of failures is less than n in t ¼ 104 hours is given by a ¼ 0:95; 0:99:
92
3 Renewal Processes
3.8 Show (3.25) and (3.26) [19, p. 259]. 3.9 Prove (3.27). 3.10 In inequalities of MðtÞ and HðtÞ; prove that t tFðtÞ 1 MðtÞ R t : 0 FðuÞdu 0 FðuÞdu
Rt
3.11 In (3.29), prove when F(t) is IFR, FðtÞ Rt
FðuÞdu
hðtÞ 1 R
FðtÞ
:
FðuÞdu
t
0
3.12 In (3.38), prove that when FðtÞ is IFR, l1 decreases from 1=hð0Þ to l; and l2 increases from 0 to 1: 3.13 In Sect. 3.3.2.2, prove that when AðTÞ ¼ c1 FðTÞ þ c2 FðTÞ and BðTÞ c1 MðTÞ þ c2 ; BðTÞ ¼ AðTÞ þ
ZT
BðT tÞdFðtÞ; BðTÞ ¼ AðTÞ þ
Z
T
AðT tÞdMðtÞ: 0
0
3.14 Show that the interval reliability in (3.67), when T0 is a random variable with an exponential distribution ð1 eat Þð0\a\1Þ; is RðT; x; aÞ ¼ aR ðT; x; aÞ; and derive an optimum policy that maximizes RðT; x; aÞ [11, p. 143] when the failure rate increases strictly. 3.15 If the time for PM has a distribution G2 ðtÞ with mean b2 and the time for repair has G1 ðtÞ with mean b1 ; then show that the limiting interval reliability in (3.68) is [11, p. 143] R Tþx FðtÞdt x RðT; xÞ ¼ R T : 0 FðtÞdt þ b1 FðTÞ þ b2 FðTÞ 3.16 In Example 3.9, compute the time d when the probability that the system is down more than d in t ¼ 10; 000 h of operation is given by a ¼ 0:95; 0:99: 3.17 In Example 3.10, derive optimum policies when (i) cr ðtÞ ¼ at; GðtÞ ¼ pffi 1 eh t ; and (ii) cr ðtÞ ¼ at2 ; GðtÞ ¼ 1 eht : 3.18 Derive (3.87). 3.19 Derive (3.88) and (3.89). 3.20 Derive (3.90) and (3.91). 3.21 In Example 3.14, derive P01 ðnÞ and M01 ðnÞ: 3.22 A generalized renewal process in which Snþ1 ¼ Sn þ Anþ1 Xnþ1
ðn ¼ 0; 1; 2; . . .Þ
is applied to a imperfect repair model [20]. When An ¼ a; derive PrfXnþ1 xjSn ¼ tg:
3.6 Problems 3
93
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20.
Barlow RE, Proschan F (1965) Mathematical theory of reliability. Wiley, New York Prabhu NU (2007) Stochastic processes. World Scientific, Singapore Parzen E (1962) Stochastic process. Holden-Day, San Francisco Beichelt FE (2006) Stochastic processes in science, engineering and finance. Chapman & Hall, Boca Raton Osaki S (1992) Applied stochastic system modeling. Springer, Berlin Ross SM (1983) Stochastic processes. Wiley, New York Cox DR (1962) Renewal theory. Methuen, London Feller W (1968) An introduction to probability theory and its application, vol I, 3rd edn. Wiley, New York Çinlar E (1975) Introduction to stochastic processes. Prentice-Hall, Englewood Cliffs Takács L (1960) Stochastic processes. Wiley, New York Nakagawa T (2005) Maintenance theory of reliability. Springer, London Tijms HC (2003) A first course in stochastic models. Wiley, Chichester Dohi T, Kaio N, Osaki S (2002) Renewal processes and their computational aspects. In: Osaki S (ed) Stochastic models in reliability and maintenance. Springer, Berlin Xie M (1989) On the solution of renewal-type integral equations. Commun Statist B 18:281– 293 Calabro SR (1962) Reliability principles and practices. McGraw-Hill, New York Lam Yeh (2007) Geometric process and its applications. World Scientific, Singapore Nakagawa T (2007) Shock and damage models in reliability theory. Springer, London Wang H, Pham H (2006) Reliability and optimal maintenance. Springer, London Rausand M, Høyland A (2004) System reliability theory. Wiley, Hoboken Kijima M (2002) Generalized renewal processes and general repair models. In: Osaki S (ed) Stochastic models in reliability and maintenance. Springer, Berlin
Chapter 4
Markov Chains
In Sect. 3.4, we consider the system that repeats up and down such as operating and failed states alternately. Next, as one example of extended models, we take up the system with repair maintenance: The system begins to operate at time 0 and undergoes repair according to two types of failures such as minor and major failures. After the repair completion, the system becomes like new and begins to operate again. To form the stochastic modeling of the above system, we define the following three states: State 0: System is operating. State 1: System is under repair according to minor failure. State 2: System is under repair according to major failure. In general, the repair time and cost required for major failure might be greater than those for minor failure. The system repeats transitions between the states of 0 ! 1 ! 0 or 0 ! 2 ! 0 (Fig. 4.1). We explain simply and clearly Markov processes on the subject of the system: Suppose that the failure occurs according to a general distribution FðtÞ, and when it has occurred, the probability of minor failure is a ð0 a 1) and the probability of major failure is 1 a, where both probabilities are independent of failure times. The distribution of time to the repair completion of minor and major failures are G1 ðtÞ and G2 ðtÞ, respectively. If we pay no regard to transition times and are concerned only with transition probabilities among three states, we express the process in the following matrix form: 2 1 00 1 0 0 a 1a P ¼ 1 @1 0 0 A: 2 1 0 0
T. Nakagawa, Stochastic Processes, Springer Series in Reliability Engineering, DOI: 10.1007/978-0-85729-274-2_4, Ó Springer-Verlag London Limited 2011
ð4:1Þ
95
96
4 Markov Chains
Fig. 4.1 Process of the system with two types of failures
The stochastic process with discrete-state space such as (4.1) is called a discretetime Markov chain. Using the transition matrix in (4.1), we can obtain stochastic behaviors of the system in Sect. 4.1. Next, assume that FðtÞ ¼ 1 ekt and Gj ðtÞ ¼ 1 ehj t (j ¼ 1; 2). We investigate the states of the system at time t. Let ZðtÞ denote the state of the system at time t when it starts from State 0, and define that Pj ðtÞ PrfZðtÞ ¼ jg (j ¼ 0; 1; 2), that is called a transition probability. Note that the process has the memoryless property irrespective of elapsed operating and repair times because of exponential distribution in Sect. 2.1.1. Thus, we have the following equations: For small h [ 0, P0 ðt þ hÞ ¼ P0 ðtÞekh þ P1 ðtÞð1 eh1 h Þ þ P2 ðtÞð1 eh2 h Þ þ oðhÞ; P1 ðt þ hÞ ¼ P1 ðtÞeh1 h þ P0 ðtÞað1 ekh Þ þ oðhÞ; P2 ðt þ hÞ ¼ P2 ðtÞeh2 h þ P0 ðtÞð1 aÞð1 ekh Þ þ oðhÞ; where oðhÞ=h ! 0 as h ! 0. From these equations, P0 ðt þ hÞ P0 ðtÞ 1 ekh 1 eh1 h 1 eh2 h oðhÞ þ P1 ðtÞ þ P2 ðtÞ þ ¼ P0 ðtÞ ; h h h h h P1 ðt þ hÞ P1 ðtÞ 1 eh1 h 1 ekh oðhÞ þ P0 ðtÞa þ ¼ P1 ðtÞ ; h h h h P2 ðt þ hÞ P2 ðtÞ 1 eh2 h 1 ekh oðhÞ þ P0 ðtÞð1 aÞ þ ¼ P2 ðtÞ : h h h h As h ! 0, we have the differential equations P00 ðtÞ ¼ kP0 ðtÞ þ h1 P1 ðtÞ þ h2 P2 ðtÞ; P01 ðtÞ ¼ h1 P1 ðtÞ þ akP0 ðtÞ; P02 ðtÞ ¼ h2 P2 ðtÞ þ ð1 aÞkP0 ðtÞ: In a matrix form, 0
1 0 P00 ðtÞ k h1 P0 ðtÞ @ P01 ðtÞ A ¼ @ ak h1 P02 ðtÞ ð1 aÞk 0
1 h2 0 APðtÞ: h2
ð4:2Þ
4 Markov Chains
97
Therefore, by solving the above differential equations mathematically and numerically, we can obtain the probabilities Pj ðtÞ (j ¼ 0; 1; 2). The stochastic process, whose conditional probability of transition to the future states depends only on the present state and is independent of the past history, is called a Markov property. The stochastic process with Markov property at any time is a Markov process. Many real-world phenomena can be modeling stochastically by Markov processes because we may observe the phenomena only at Markov property points without keeping track of the time elapsed since the last point. In other words, if we know the state for any specific value of time, we can predict the stochastic behavior beyond the point. Stochastic processes with independent increments defined in Sect. 2.2 always have the Markov property. Markov processes can be analyzed by forming differential equations such as (4.2) and are also called a continuous-time Markov chain. More detailed discussions for Markov processes are done in Sect. 4.2.
4.1 Discrete-Time Markov Chain 4.1.1 Transition Probabilities Consider a discrete-time stochastic process fXn ; n ¼ 0; 1; 2; . . .g with a finite discrete-state set f0; 1; 2; . . .; mg. Note that the event fXn ¼ jg represents that the process is in State j (j ¼ 0; 1; . . .; m) at time n (n ¼ 0; 1; 2; . . .). That is, the process moves among (m þ 1) states at some unit of time, time unit, or jump according to a given probability law. Then, we want to know mainly in which state the process is at time n and converges as n becomes larger. If we suppose that PrfXnþ1 ¼ jjX0 ¼ i0 ; X1 ¼ i1 ; . . .; Xn ¼ in g ¼ PrfXnþ1 ¼ jjXn ¼ in g
ð4:3Þ
for all i0 ; i1 ; . . .; in and all n 0, then the process fXn ; n ¼ 0; 1; 2; . . .g is said to be a Markov chain (Markov 1806). This property indicates that, given the value of Xn , the future value of Xnþ1 does not depend on the value of Xk for 0 k n 1. If the probability of Xnþ1 being in State j, given that Xn is in State i, is independent of n, i.e., PrfXnþ1 ¼ jjXn ¼ ig Pij ; then the process has a stationary or homogeneous (one-step) transition probability, i.e., the process transits from State i to State j with probability Pij and remains in State i with probability Pii regardless of n. In other words, the transition probability from State i to State j does not change even if time varies. We restrict ourselves only to discrete-time Markov chains with stationary transition probabilities and a finite state space, because most reliability systems form such processes. A matrix
98
4 Markov Chains
P fPij g is called a transition probability matrix, and manifestly, Pij satisfies P Pij 0 and m j¼0 Pij ¼ 1 for any i, j. Let P denote the transition matrix given by 0
P00 B P10 B . B . B . P ðPij Þ ¼ B B Pi0 B . @ . . Pm0
P01 P10 .. . Pi1 .. . Pm1
1 P0m P1m C .. C C . C C: . . . Pim C .. C A . . . . Pmm ... ...
Example 4.1 (Alternating renewal process) Consider an alternating renewal process in Sect. 3.4, where a unit begins to operate at time 0 and undergoes repair at failures. After the repair completion, a unit becomes like new and begins to operate again. Then, we define State 0: unit is operating. State 1: unit is under repair. The process has the simplest Markov chain with two states f0; 1g, and a transition probability matrix is P00 P01 0 1 P : ¼ P10 P10 1 0 The process repeats alternately between operating and repair. When the repair is made, the unit operates like new with probability 0:8, however, it undergoes repair again because of its imperfectness with probability 0:2. Then, a transition probability matrix is 0 1 P¼ : 0:8 0:2 It is helpful to understand the process whose behavior is illustrated by using state transition diagram as shown in Figs. 4.1 and 4.2, where the number circled denotes a state and the number appearing on an arc is a transition probability or a transition rate (Fig. 4.3). It is indispensable to draw a state transition diagram and to make clear of the behavior for modeling reliability systems stochastically. Example 4.2 (One-unit system) Consider a one-unit system with repair in Example 3.6, where the failure and repair times have exponential distributions
Fig. 4.2 Diagram of oneunit system
4.1 Discrete-Time Markov Chain
99
Fig. 4.3 Diagram of oneunit system with exponential failure and repair times
ð1 ekt Þ and ð1 eht Þ, respectively. To know the behavior of the system, we check the unit at periodic times kT (k ¼ 1; 2; . . .). If the unit has failed at some check, it undergoes repair immediately, and otherwise, leaves alone. Similarly, if the repair is completed at some check, it begins to operate, and otherwise, continues to make the repair. Then, because of the memoryless property of exponential distributions, the transition matrix is kT P00 P01 1 ekT e P¼ ¼ ; P10 P10 1 ehT ehT and the state transition diagram is shown in Fig. 4.3. When T ! 1, and ekT ¼ 0 and ehT ¼ 0:2, the transition matrices correspond to those in Example 4.1, respectively. A Markov chain is completely specified by the transition probabilities Pij and an initial probability PrfX0 ¼ ig. Let Pnij denote the probability that the process goes from State i to State j in n transitions that is called the n-step transition probability, i.e., Pnij PrfXnþk ¼ jjXk ¼ ig P0ij P0ii
0
ðn ¼ 1; 2; . . .Þ;
for i 6¼ j;
1;
P n 1 n Pnij 0; m j¼0 Pij ¼ 1 for all n 0, and Pij Pij for any i, j. Note that Pij is independent of the present time k and depends only on the time duration n. Then, Pnij ¼ PrfXnþk ¼ jjXk ¼ ig ¼
m X
PrfXnþk ¼ j; Xrþk ¼ ljXk ¼ ig
l¼0
¼ ¼
m X l¼0 m X
PrfXnþk ¼ jjXrþk ¼ lgPrfXrþk ¼ ljXk ¼ ig Pril Pnr lj
ðr ¼ 0; 1; 2; . . .; nÞ;
ð4:4Þ
l¼0
that is called the Chapman–Kolmogorov equation. In a matrix form, from (4.4), PðnÞ Pnij ¼ Pð1Þ Pðn1Þ ¼ P Pðn1Þ ¼ ¼ Pn ;
ð4:5Þ
100
4 Markov Chains
i.e., PðnÞ can be computed by the nth power of matrix P. Thus, (4.4) is written in the following matrix form Pn ¼ Pr Pnr
ðr ¼ 0; 1; 2; . . .; nÞ;
ð4:6Þ
where note that Pð0Þ ¼ P0 ¼ I is the identity matrix where all diagonal elements are 1. The Chapman–Kolmogorov equation in (4.4) plays a great role in analyzing Markov chains that will be shown in (4.25). Therefore, the probability that the process is in State j at time n, irrespective of an initial state, is pj ðnÞ PrfXðnÞ ¼ jg ¼
m X
PrfXðnÞ ¼ j;
Xð0Þ ¼ ig
i¼0
¼ ¼
m X i¼0 m X
PrfXðnÞ ¼ jjXð0Þ ¼ igPrfXð0Þ ¼ ig pi ð0ÞPnij
ðj ¼ 0; 1; 2; . . .; mÞ;
ð4:7Þ
i¼0
P where note that m j¼0 pj ðnÞ ¼ 1 for all n 0. Letting pðnÞ fp0 ðnÞ; p1 ðnÞ; . . .; pm ðnÞg, (4.7) is written in a matrix form pðnÞ ¼ pð0ÞPðnÞ
ðn ¼ 0; 1; 2; . . .Þ:
ð4:8Þ
Example 4.3 In Example 4.2, setting a 1 ekT and b 1 ehT [1, p. 101], 1a a ; P¼ b 1b ! ð1 aÞa þ ð1 bÞa ð1 aÞ2 þ ab 2 : P ¼ ð1 aÞb þ ð1 bÞb ab þ ð1 bÞ2 In general (Problem 4.1), 1 P ¼ aþb n
b b
a a
ð1 a bÞn þ aþb
a b
a : b
If the initial state is pð0Þ ¼ ð1; 0Þ, i.e., the unit starts to operate at time 0, 1 ð1 a bÞn ða; aÞ ðb; aÞ þ aþb aþb b þ að1 a bÞn a½1 ð1 a bÞn ; : ¼ aþb aþb
pðnÞ ¼
For example, when kT ¼ 0:1 and hT ¼ 0:2, i.e., a ¼ 0:095 and b ¼ 0:181, pðnÞ ¼ ð0:656 þ 0:344ð0:724Þn ; 0:344½1 ð0:724Þn Þ
4.1 Discrete-Time Markov Chain
101
that goes to pð1Þ limn!1 pðnÞ ¼ ð0:656; 0:344Þ, and the values of 0.656 and 0.344 are approximately equal to the availability h=ðk þ hÞ and the unavailability k=ðk þ hÞ in Example 4.2.
4.1.2 Classification of States Two States i and j communicate if and only if there exist integers n 0 and k 0 such that Pkij [ 0 and Pnji [ 0. By using a communication relation, all states are classified into several classes. Furthermore, we define the first-passage distribution as Fijn PrfXn ¼ j; Xk 6¼ j; k ¼ 1; 2; . . .; n 1jX0 ¼ ig ðn ¼ 1; 2; . . .Þ; that is the probability that starting in State i, the first transition into State j occurs at P n the nth transition, where Fij0 0 for all i, j. In addition, we define Fij 1 n¼1 Fij that represents the eventual transition probability from State i to State j. All states in a finite Markov chain are classified into two kinds of states: (i) If Fii ¼ 1, then State i is said to be recurrent, (ii) If Fii \1, then State i is said to be transient. A recurrent set of states is one in which all states communicate and which cannot be left, once it is entered. A transient set of states is one in which all states communicate and which might be left even if it is entered. We have the following relation between a transition probability Pnij and a firstpassage distribution Fijn Pnij ¼
n X
Fijk Pnk jj
ðn ¼ 1; 2; . . .Þ;
ð4:9Þ
k¼0
because Pnij is denoted by the probability that the process transits from State i to State j for the first time at the kth step (k ¼ 1; 2; . . .; n) and returns to State j at the ðn kÞth step, where note that P0ij ¼ 0 for i 6¼ j, and P0jj ¼ 1. Let P ðzÞ and Fij ðzÞ be the generating functions of Pnij and Fijn , i.e., Pij ðzÞ P1 n n P1 n ijn n¼0 z Pij and Fij ðzÞ n¼0 z Fij for jzj\1 from Appendix A.4. Then, (4.9) is rewritten by the generating function, for i 6¼ j, Pij ðzÞ ¼ ¼
1 X n¼0 1 X k¼0
zn Pnij ¼
1 X n¼0
zk Fijk
1 X n¼0
zn
n X
Fijk Pnk ij
k¼0
zn Pnjj ¼ Fij ðzÞPjj ðzÞ;
ð4:10Þ
102
4 Markov Chains
and for i ¼ j, Pii ðzÞ ¼ 1 þ Fii ðzÞPii ðzÞ;
Pii ðzÞ ¼
i.e:;
1 : 1 Fii ðzÞ
ð4:11Þ
Next, let Mijn denote the expected number of visits to State j in the nth transition if the process starts from State i, not including the first at time 0. Then, clearly Pn k ðn ¼ 1; 2; . . .Þ; k¼1 Pij Mijn ¼ ð4:12Þ 0 ðn ¼ 0Þ: Forming the generating function of (4.12), for i 6¼ j, Mij ðzÞ
1 X
zn Mijn ¼
Pij ðzÞ ; 1z
zn Miin ¼
Pii ðzÞ 1 : 1z
n¼0
Mii ðzÞ
1 X n¼0
ð4:13Þ
Using Theorem A.5 of Appendix A, when State i is recurrent, lim z!1
1 X n¼0
zn Fiin ¼ lim Fii ðzÞ ¼ z!1
1 X
Fiin ¼ 1;
n¼0
and hence, from (4.11), lim Pii ðzÞ ¼ lim z!1
z!1
1 X
zn Pnii ¼ 1:
n¼0
Thus, using Theorem A.5, 1 X
Pnii ¼ 1:
n¼0
P1
n n¼0 Pii
¼ 1, we assume that Fii \1. Then, from (4.11), Next, when limz!1 Pii ðzÞ\1 using Theorem A.5. In addition, using Theorem A.5 again, P1 n P1 n n¼0 Pii \1 that contradicts the assumption. Thus, if n¼0 Pii ¼ 1, then Fii ¼ 1. When State i is transient, we can make similar discussions. Therefore, we have the following results: Theorem 4.1 P1 n (i) State i is recurrent if and only if P ¼ 1. P1n¼0 nii (ii) State i is transient if and only if P n¼0 ii \1. The above results show that for any recurrent State i, the process can revisit State i infinitely. For any transient State i, the process transits to states of another class after it revisits to State i finitely, and cannot return to State i.
4.1 Discrete-Time Markov Chain
103
Example 4.4 In Example 4.3, it is assumed that 0\a\1 and b ¼ 0, i.e., any failed units remain a failed state without making repair. Then, a transition probability matrix is 1a a P¼ ; 0 1 and 1 X
1 Pn00 ¼ 1 þ ð1 aÞ þ ð1 aÞ2 þ ¼ ; a n¼0
1 X
Pn11 ¼ 1 þ 1 þ ¼ 1;
n¼0
that implies that State 0 is transient and State 1 is recurrent. In general, the transition probability of a finite Markov chain can be expressed in a matrix form [1, p. 124, 2, p. 36]
(4.14)
where C1 ; C2 ; . . .; Ck are all the sets of recurrent classes and T is a set of transient states.
4.1.3 Limiting Probabilities P n Recall that if State i is recurrent, then Fii ¼ 1 and 1 conversely, n¼0 Pii ¼ 1, andP P1 n n if State i is transient, then Fii \1 and n¼0 Pii \1. Define that lj 1 n¼1 nFjj is the mean recurrence time for State j (j ¼ 0; 1; . . .; m). Theorem 4.2 (i) If State jis recurrent, then Pnjj !
1 [0 lj
as n ! 1:
ð4:15Þ
(ii) If State jis transient, then Pnjj ! 0 as n ! 1:
ð4:16Þ
104
4 Markov Chains
(iii) If State jis recurrent, then Pnij !
Fij : lj
ð4:17Þ
Proof From (4.12), Mjjn Mjjn1 ¼ Pnjj !
1 lj
as n ! 1:
P n Furthermore, because m j¼0 Pjj ¼ 1 for any n 0, 1=lj [ 0. If State j is transient, P1 n n n¼1 Pjj \1, and hence, Pjj ! 0 as n ! 1. If a Markov chain starts from State i, visits State j in a recurrent class with probability Fijn and results infinitely State j. Thus, noting that Fij represents the eventual transition probability from State i to State j, we get easily (iii). h Theorem 4.3 If States i and j (i; j ¼ 0; 1; . . .; m) are in the same recurrent class, pj limn!1 Pnij [ 0, that is a unique and positive solution to satisfy pj ¼
m X
pi Pij
ðj ¼ 0; 1; . . .; mÞ;
i¼0
m X
pj ¼ 1:
ð4:18Þ
j¼0
Proof From the assumption that both States i and j are in the same recurrent class, Fij ¼ 1, and hence, limn!1 Pnij ¼ 1=lj [ 0 from (i) and (iii) of Theorem 4.2. Furthermore, from (4.7), pj ðnÞ ¼
m X
pi ðn 1ÞPij ;
i¼0
m X
pj ðnÞ ¼ 1 for all n 1:
j¼0
Letting n ! 1, we easily have (4.18).
h
Two methods of solving (4.18) were proposed [3, p. 107]; direct method and iterative method. However, because the number of states is not so large in most reliability models, it would be sufficient to solve (4.18) by the direct method. Note that (4.18) is expressed in a matrix form p ðp0 ; p1 ; . . .; pm Þ ¼ pP:
ð4:19Þ
n Therefore, we summarize the liming probabilities P1 ij limn!1 Pij of a finite Markov chain in (4.14)
4.1 Discrete-Time Markov Chain
105
1 ði; j 2 Cl l ¼ 1; 2; . . .; kÞ; lj Fij ¼ ði 2 T; j 2 Cl l ¼ 1; 2; . . .; kÞ; lj
P1 ij ¼ P1 ij
P1 ij ¼ 0 ði 2 Cl ; j 2 Cr
ð4:20Þ
l 6¼ rÞ;
P1 ij ¼ 0 ði; j 2 TÞ; P1 ij ¼ 0 ði 2 Cl
l ¼ 1; 2; . . .; k; j 2 TÞ:
Example 4.5 When kT ¼ 0:1 and hT ¼ 0:2, i.e., a ¼ 0:095 and b ¼ 0:181 in Example 4.3, a transition probability matrix is 0 P¼ 1
0 0:905 0:181
1 0:095 ; 0:819
that consists of one recurrent class. Then, the limiting probabilities p0 and p1 are, from Theorem 4.3, 0:905 0:095 : ðp0 ; p1 Þ ¼ ðp0 ; p1 Þ 0:181 0:819 Thus, solving two equations p0 ¼ 0:905p0 þ 0:181p1 ;
p0 þ p1 ¼ 1;
p0 0:656 and p1 0:344 that are equal to pð1Þ in Example 4.3. In a transition probability matrix in (4.1), the limiting transition probabilities are given by solving simultaneous equations p1 ¼ p0 a;
p2 ¼ p0 ð1 aÞ;
p0 þ p1 þ p2 ¼ 1:
Thus, p0 ¼ 1=2, p1 ¼ a=2, and p2 ¼ ð1 aÞ=2 (Problem 4.2). Example 4.6 Suppose that a transition probability matrix is given by
106
4 Markov Chains
Then, the recurrent classes are C1 ¼ f0g, C2 ¼ f1; 2g, C3 ¼ f3; 4g, and a transient class is T ¼ f5; 6g. The transition diagram among seven states is shown in Fig. 4.4 For a recurrent class C2 ¼ f1; 2g, the transition probability matrix is 1 0 1 1 C B P2 ¼ @ 2 2 A; 1 3 4 4 and the limiting probabilities are, from (4.18), 1 p1 ¼ ; 3 Similarly, C3 ¼ f3; 4g,
0
2 P ¼ @3 1
1 1 3 A; 0
2 p2 ¼ : 3
3 p3 ¼ ; 4
1 p4 ¼ : 4
We derive the first-passage distributions Fij (i 2 T; j 2 Ck (k ¼ 2; 3)) defined in Sect. 4.1.2 [1, p. 127]. Theorem 4.4 The first-passage distribution Fij from State i (i 2 T) to State j (j 2 Ck ) ðk ¼ 1; 2; . . .Þ satisfies X X Pil þ Pil Flj ði 2 T; j 2 Ck Þ; ð4:21Þ Fij ¼ l2Ck
l2T
and its matrix form is ðFij Þ ¼ ðI QÞ1 Rk 1;
ð4:22Þ
where 1 is a column vector where all elements are 1. Proof The first-passage distribution Fij has two possibilities: one is to transit from State i to State l (l 2 Ck ) directly, and the other is to transit from State i to State l ðl 2 T) once including itself and then to transit to State j (j 2 Ck ). Thus, we have (4.21). Rewriting (4.21) in a matrix form, Fig. 4.4 Transition diagram of Example 4.6
4.1 Discrete-Time Markov Chain
107
Fij ¼ Rk 1 þ QðFij Þ; i.e., ðFij Þ ¼ ðI QÞ1 Rk 1: Example 4.7 In Example 4.6, 0 35 ; Q¼ 1 0 3
h ðI QÞ ¼
i.e:;
Hence, the inverse matrix is ðI QÞ1 ¼
5 4 5 12
3 4 5 4
1 13
35 : 1
:
Thus, using Theorem 4.4, 0 5 B ðI QÞ1 R1 1 ¼ @ 4 5 12 0 5 B ðI QÞ1 R2 1 ¼ @ 4 5 12 0 5 B ðI QÞ1 R3 1 ¼ @ 4 5 12
10 1 1
0 5 1 CB 5 C B 8 C A@ Að1Þ ¼ @ A; 1 5 17 2 4 24 0 1 1 1 3 10 1 1 0A B C 4 C@ ¼ @ 8 A; A 10 1 5 1 0 0 24 4 1 011 3 10 1 0 1 C B C 4 CB 10 ¼ @ 4 A: A@ A 1 1 5 1 0 6 4 4 3 4
Consequently, 5 F50 ¼ ; 8 17 F60 ¼ ; 24
1 ðj ¼ 1; 2Þ; 8 1 F6j ¼ ðj ¼ 1; 2Þ; 24 F5j ¼
where note that 58 þ 18 þ 14 ¼ 1 and
17 24
1 4 1 F6j ¼ 4
F5j ¼
ðj ¼ 3; 4Þ; ðj ¼ 3; 4Þ;
1 þ 24 þ 14 ¼ 1. Therefore, from (4.20),
5 1 1 1 P1 ; P1 50 ¼ ; 51 ¼ ¼ 8 8 3 23 1 3 3 1 1 1 ; P1 ; P1 53 ¼ ¼ 54 ¼ ¼ 4 4 16 4 4 16 17 1 1 1 P1 ; P1
¼ ; 60 ¼ 61 ¼ 24 24 3 72 1 3 3 1 1 1 1 1 P63 ¼ ¼ ; P64 ¼ ¼ ; 4 4 16 4 4 16 P4 P 4 1 where note that j¼0 P1 j¼0 P6j ¼ 1. 5j ¼
1 2 1 P1 ; 52 ¼ ¼ 8 3 11
P1 62 ¼
1 2 1
¼ ; 24 3 36
108
4 Markov Chains
Collecting these results, a matrix of limiting probabilities is
4.1.4 Absorbing Markov Chain We are sometimes interested in stochastic behaviors before system failure. An absorbing Markov chain consists of one in which all recurrent states are absorbing. The transition probability matrix of an absorbing Markov chain with k absorbing states and transient states can be expressed, from (4.14),
(4.23)
where 0 is zero matrix in which all elements are zero. We obtain the expected number of visits to State j, starting from State i (i; j 2 T), i.e., ði; j 2 TÞ: Mij ¼ EfNj ð1ÞjXð0Þ ¼ ig Theorem 4.5 The expected number of visits from State i to State j is [2, p. 46] Mij ¼ ðI QÞ1 : ð4:24Þ Proof Letting Ujn be the function that takes 1 when the process is State j at time n, P n and otherwise, takes 0, it can be easily seen that Nj ð1Þ ¼ 1 n¼0 Uj . Thus,
4.1 Discrete-Time Markov Chain
Mij ¼
( E
1 X n¼0
¼ ¼
109
Ujn Xð0Þ
)! ¼i
¼
1 n o X E Ujn jXð0Þ ¼ i n¼0
1 h i X ðnÞ ðnÞ Pij 1 þ 1 Pij 0
! ¼
n¼0 1 X
1 X
!
! ðnÞ Pij
n¼0
Qn ¼ ðI QÞ1 :
n¼0
Note that at least one of sums of each row in Q is \1 because Q is a matrix of transient states, and hence, the inverse of a matrix ðI QÞ exists. h Example 4.8 In Example 4.7, ðI QÞ1
0
5 5B 4 ¼ @ 5 6 12
1 3 4C 5 A: 4
Thus, the expected numbers of visits to transient States 5 and 6 before absorbing are 5=4 þ 3=4 ¼ 2 from State 5 and 5=12 þ 5=4 ¼ 5=3 from State 6, where note that 2 and 5=3 are the mean times from States 5 and 6 to an absorbing state, respectively, because each transition time is a unit of time in a Markov chain. Example 4.9 (Two-unit parallel system with spare units) Consider a two-unit parallel system with no repair, however, spare units in inventory: Two units begin to operate at time 0. If one of operating unit fails before time T, the spare unit is ordered immediately and is placed into service as soon as it is delivered after a lead time L. If two operating units do not fail up to time T, the spare unit is ordered at time T and is delivered after time L. In this case, when two units have failed, the spare unit is put into service and the spare unit is ordered immediately. When only one unit has failed, the spare unit is put into service, and when two units are operating, the spare unit is put into inventory. It is assumed that each unit fails according to an identical exponential distribution ð1 ekt Þ. We define the states: State State State State State State
0: 1: 2: 3: 4: 5:
Two units are operating and there is no spare. Two units are operating and spare is ordered. One unit is operating and spare is ordered. Two units are operating and spare is delivered. One unit is operating and spare is delivered. Two units have failed and spare is delivered.
We draw a state transition diagram of the above system in Fig. 4.5. Referring to the diagram, transition probabilities are P02 ¼ 1 e2kT ; P13 ¼ e2kL ; P01 ¼ e2k T ; kL kL kL 2 P14 ¼ 2e ð1 e Þ; P15 ¼ ð1 e Þ ; P24 ¼ ekL ; P25 ¼ 1 ekL ; P30 ¼ P40 ¼ P52 ¼ 1;
110
4 Markov Chains
Fig. 4.5 Diagram of twounit parallel system with spare units in inventory
and its matrix form is 1 00 0 0 e2kT 1B 0 B0 2B 0 0 P¼ B 3B 1 0 B 4 @1 0 5 0 0
2 1 e2kT 0 0 0 0 1
3 0 e2kL 0 0 0 0
4 5 1 0 0 2ekL ð1 ekL Þ ð1 ekL Þ2 C C ekL 1 ekL C C: C 0 0 C A 0 0 0 0
When kT ¼ 0:1 and kL ¼ 0:05, the transition probability matrix is 0 1 2 P¼ 3 4 5
1 00 0 0:8187 B0 0 B B0 0 B B1 0 B @1 0 0 0
2 0:1813 0 0 0 0 1
3 0 0:9048 0 0 0 0
4 5 1 0 0 0:0928 0:0024 C C 0:9512 0:0488 C C: 0 0 C C 0 0 A 0 0
Suppose that State 5 is in a failed state of the system and is absorbing. Striking out row 5 and column 5 of I P and inverting it, we can compute the expected number of visits to State j before system failure 0 1 2 3 4
0 1 2 3 4 1 488:944 400:303 88:360 392:387 95:557 B 488:800 401:185 88:604 393:241 95:549 C C B B 484:079 396:319 88:749 388:483 95:597 C : C B @ 488:944 400:303 88:630 393:387 95:557 A 488:944 400:303 88:630 392:387 96:557 0
Thus, the mean time from Sate 0 to State 5 is 1; 465:551 ð1=kÞ by summing the number of column 0. For example, when 1=k ¼ 100, the mean time is 146, 555:1 (Problem 4.3).
4.2 Continuous-Time Markov Chain
111
4.2 Continuous-Time Markov Chain 4.2.1 Transition Probabilities Consider a continuous-time stochastic process fXðtÞ; t 0g with a finite state set f0; 1; 2; . . .; mg. Note that the event fXðtÞ ¼ jg represents that the process is in State j (j ¼ 0; 1; 2; . . .; m) at time t. Then, we want to know mainly in which state the process is at time t and the process converges as t becomes larger. If we suppose that PrfXðtÞ ¼ jjXðt0 Þ ¼ i0 ; Xðt1 Þ ¼ i1 ; . . .; Xðtn Þ ¼ in g ¼ PrfXðtÞ ¼ jjXðtn Þ ¼ in g for all i0 , i1 , . . ., in and 0 t0 \t1 \ \tn \t, then the process fXðtÞg is said to be a continuous-time Markov chain. This property indicates that, given the value of Xðtn Þ, the future value of XðtÞ for t [ tn does not depend on the value of Xðtn1 Þ for any tn1 \tn and depends only on the present state. If the probability of Xðt þ uÞ being in State j, given that XðuÞ is in State i, is independent of u 0, i.e., PrfXðt þ uÞ ¼ jjXðuÞ ¼ ig Pij ðtÞ; then the process has a stationary or homogeneous transition probability that depends only on the time difference t. We restrict ourselves to continuous-time Markov chains with stationary transition probabilities and a finite state space f0; 1; 2; . . .; mg like discrete-time Markov chains. A continuous-time Markov chain is a stochastic process in which the process moves among states in accordance with a discrete-time Markov chain, however, its sojourn time spent in each state to visit the next state is independent and distributed exponentially. A matrix PðtÞ ¼ fPij ðtÞg is called a transition probability matrix at time t, and P satisfies Pij ðtÞ 0 and m j¼0 Pij ðtÞ ¼ 1 for any t 0 and i, j. In a similar way of obtaining (4.4), Pij ðt þ uÞ ¼ PrfXðt þ uÞ ¼ jjXð0Þ ¼ ig m X ¼ PrfXðt þ uÞ ¼ j; XðuÞ ¼ kjXð0Þ ¼ ig k¼0
¼ ¼
m X k¼0 m X
PrfXðt þ uÞ ¼ jjXðuÞ ¼ kgPrfXðuÞ ¼ kjXð0Þ ¼ ig Pik ðuÞPkj ðtÞ
ð4:25Þ
k¼0
for t, u 0, that is called the Chapman–Kolmogorov equation. In a matrix form, (4.25) is rewritten as Pðt þ uÞ ¼ PðtÞPðuÞ:
ð4:26Þ
112
4 Markov Chains
However, we cannot calculate the transition probability by the matrix operation as shown in Sect. 4.1.1.
4.2.2 Pure Birth Process and Birth and Death Process Recall that a Poisson process is a counting process in which the process has stationary and independent increments, and the interarrival times have an identical exponential distribution with mean 1=k. Expanding the Poisson process, the law of probability of arrival events depends on the total number of events that has already arrived, i.e., the interarrival times are independent and distributed exponentially with mean 1=kk that depends on the present State k. In this case, states are denoted by the total number of events. Definition 4.1 If fXðtÞg is a continuous-time Markov chain with (i) Xð0Þ ¼ 0, (ii) PrfXðt þ hÞ XðtÞ ¼ 1jXðtÞ ¼ kg ¼ kk h þ oðhÞ, (iii) PrfXðt þ hÞ XðtÞ 2jXðtÞ ¼ kg ¼ oðhÞ, then the process fXðtÞg is said to be a pure birth process with parameters kk and transits only from k to k þ 1. Using this definition, we derive the stationary transition probability Pk ðtÞ PrfXðtÞ ¼ kjXð0Þ ¼ 0g ¼ PrfXðtÞ ¼ kg
ðk ¼ 0; 1; . . .; mÞ;
that represents that k events occur in the time interval ½0; t. By the similar method of obtaining (2.25), from Definition 4.1, P0 ðt þ hÞ ¼ PrfXðt þ hÞ ¼ 0g ¼ PrfXðtÞ ¼ 0gPrfXðt þ hÞ XðtÞ ¼ 0jXðtÞ ¼ 0g ¼ P0 ðtÞ½1 k0 h þ oðhÞ: Hence, as h ! 0, dP0 ðtÞ ¼ k0 P0 ðtÞ: dt By integration and setting P0 ð0Þ ¼ 1 from (i), P0 ðtÞ ¼ ek0 t :
ð4:27Þ
4.2 Continuous-Time Markov Chain
113
Similarly, Pk ðt þ hÞ ¼ PrfXðt þ hÞ ¼ kg ¼ PrfXðtÞ ¼ k 1; Xðt þ hÞ XðtÞ ¼ 1g þ PrfXðtÞ ¼ k; Xðt þ hÞ XðtÞ ¼ 0g þ
k X
PrfXðtÞ ¼ k j; Xðt þ hÞ XðtÞ ¼ jg
j¼2
¼ Pk1 ðtÞ½kk1 h þ oðhÞ þ Pk ðtÞ½1 kk h þ oðhÞ þ oðhÞ: Hence, as h ! 0, dPk ðtÞ ¼ kk Pk ðtÞ þ kk1 Pk1 ðtÞ; dt i.e., ekk t ½P0k ðtÞ þ kk Pk ðtÞ ¼
d½ekk t Pk ðtÞ ¼ kk1 ekk t Pk1 ðtÞ: dt
By integration under Pk ð0Þ ¼ 0 for k 1, Pk ðtÞ ¼ kk1 ekk t
Zt
ekk u Pk1 ðuÞdu ðk ¼ 1; 2; . . .Þ:
ð4:28Þ
0
Therefore, from (4.27) and (4.28), we can compute Pk ðtÞ successively. Furthermore, interarrival times are independent and have an exponential distribution with mean 1=kk (k ¼ 0; 1; 2; . . .; m), as shown in Sect. 2.2. Example 4.10 (Parallel system) Consider a parallel redundant system with n identical units each of which has an independent exponential distribution ð1 ekt Þ. Then, setting that kk ðn kÞk (k ¼ 0; 1; . . .; n 1), the system forms a pure birth process. Solving (4.28) under P0 ðtÞ ¼ enkt , n ð1 ekt Þk eðnkÞkt ðk ¼ 0; 1; . . .; nÞ; Pk ðtÞ ¼ k that is easily derived by using a binomial distribution (Problem 4.4). In addition, it is shown that the mean time to system failure is given in (2.8). Next, consider the stochastic process that states increase or decrease according to arrivals or departures of events. Definition 4.2 If the process fXðtÞg is a continuous-time Markov chain with (i) Xð0Þ ¼ i, (ii) PrfXðt þ hÞ XðtÞ ¼ 1jXðtÞ ¼ kg ¼ kk h þ oðhÞ,
114
4 Markov Chains
(iii) PrfXðt þ hÞ XðtÞ ¼ 1jXðtÞ ¼ kg ¼ lk h þ oðhÞ, (iv) PrfjXðt þ hÞ XðtÞj [ 1jXðtÞ ¼ kg ¼ oðhÞ, then the process fXðtÞg is said to be a birth and death process with parameters fkk ; lkþ1 g and transits only from k to k 1 or k þ 1, where kk and lkþ1 (k ¼ 0; 1; 2; . . .) are called the birth and death rates, respectively. It follows easily from (ii–iv) that PrfXðt þ hÞ XðtÞ ¼ 0jXðtÞ ¼ 0g ¼ 1 k0 h þ oðhÞ; PrfXðt þ hÞ XðtÞ ¼ 0jXðtÞ ¼ kg ¼ 1 ðkk þ lk Þh þ oðhÞ
ðk ¼ 1; 2; . . .Þ:
Using this definition, we derive the stationary transition probability Pij ðtÞ PrfXðtÞ ¼ jjXð0Þ ¼ ig
ði; j ¼ 0; 1; . . .; mÞ:
From the Chapman–Kolmogorov equation in (4.25), Pij ðt þ hÞ ¼
m X
Pik ðtÞPkj ðhÞ:
k¼0
Thus, from (ii–iv), Pij ðt þ hÞ ¼ Pij1 ðtÞPrfXðt þ hÞ XðtÞ ¼ 1jXðtÞ ¼ j 1g þ Pij ðtÞPrfXðt þ hÞ XðtÞ ¼ 0jXðtÞ ¼ jg þ Pijþ1 ðtÞPrfXðt þ hÞ XðtÞ ¼ 1jXðtÞ ¼ j þ 1g m X Pik ðtÞPrfXðt þ hÞ XðtÞ ¼ j kjXðtÞ ¼ kg þ k¼0;k6¼j1; j; jþ1
¼ Pij1 ðtÞ½kj1 h þ oðhÞ þ Pij ðtÞ½1 ðkj þ lj Þh þ oðhÞ þ Pijþ1 ðtÞ½ljþ1 h þ oðhÞ þ oðhÞ: Hence, as h ! 0, dPij ¼ kj1 Pij1 ðtÞ ðkj þ lj ÞPij ðtÞ þ ljþ1 Pijþ1 ðtÞ dt
ðj ¼ 1; 2; . . .; m 1Þ: ð4:29Þ
In particular, for the respective cases of j ¼ 0 and j ¼ m, dPi0 ðtÞ ¼ k0 Pi0 ðtÞ þ l1 Pi1 ðtÞ; dt dPim ðtÞ ¼ km1 Pim1 ðtÞ lm Pim ðtÞ: dt
ð4:30Þ
Therefore, from (4.29) and (4.30), we can compute Pij ðtÞ under the condition of Pii ð0Þ ¼ PrfXð0Þ ¼ ig ¼ 1;
Pij ð0Þ ¼ 0
for i 6¼ j:
4.2 Continuous-Time Markov Chain
115
2λ
Fig. 4.6 Transition diagram among three states
0
λ 1
θ
2 θ
Example 4.11 (two-unit parallel system) Consider a two-unit parallel system with one repair person, i.e., only one failed unit is repaired by the repair person. It is assumed that two units begin to operate at time 0 and fail according to an identical exponential distribution ð1 ekt Þ. Failed units are repaired by one repair person according to an exponential distribution ð1 eht Þ, become like new ones, and start to operate again immediately upon the repair completion. Then, we define the states: State 0: Two units are operating. State 1: One unit is operating and the other is under repair. State 2: One unit is under repair and the other waits for repair. Note that states represent the number of failed units. Setting that k0 ¼ 2k, k1 ¼ k, and l1 ¼ l2 ¼ h, the transition diagram among three states is shown in Fig. 4.6. Setting that P00 ð0Þ 1, (4.29) and (4.30) for the system become P000 ðtÞ ¼ 2kP00 ðtÞ þ hP01 ðtÞ; P001 ðtÞ ¼ 2kP00 ðtÞ ðk þ hÞP01 ðtÞ þ hP02 ðtÞ; P002 ðtÞ ¼ kP01 ðtÞ hP02 ðtÞ: From Appendix A, using the Laplace–Stieltjes (LS) transformation of Z1 est dP0ij ðtÞ ¼ s½Pij ðsÞ Pij ð0Þ; 0
the above differential equations are sP00 ðsÞ s ¼ 2kP00 ðsÞ þ hP01 ðsÞ; sP01 ðsÞ ¼ 2kP00 ðsÞ ðk þ hÞP01 ðsÞ þ hP02 ðsÞ; sP02 ðsÞ ¼ kP01 ðsÞ hP02 ðsÞ: Solving these equations, s2 þ ðk þ 2hÞs þ h2 ; s2 þ ð3k þ 2hÞs þ 2k2 þ 2kh þ h2 2ks þ 2kh ; P01 ðsÞ ¼ s2 þ ð3k þ 2hÞs þ 2k2 þ 2kh þ h2 P00 ðsÞ ¼
116
4 Markov Chains
P02 ðsÞ ¼
2k2 ; s2 þ ð3k þ 2hÞs þ 2k2 þ 2kh þ h2
where note that P00 ðsÞ þ P01 ðsÞ þ P02 ðsÞ ¼ 1. Expanding the right-hand side of P00 ðsÞ in partial fractions, P00 ðsÞ ¼ C þ
As Bs þ ; s þ w1 s þ w2
where h2 ; 2k2 þ 2kh þ h2 ðs þ w1 Þðs þ w2 Þ ¼ s2 þ ð3k þ 2hÞs þ 2k2 þ 2kh þ h2 ; C
i.e., w1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 3k þ 2h þ k2 þ 4kh ; 2
w2
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 3k þ 2h k2 þ 4kh : 2
From the following identity equation Cðs þ w1 Þðs þ w2 Þ þ Asðs þ w2 Þ þ Bsðs þ w1 Þ ¼ s2 þ ðk þ 2hÞs þ h2 ; A
2kðw1 k hÞ ; w1 ðw1 w2 Þ
B
2kðw2 k hÞ : w2 ðw2 w1 Þ
Thus, using the inverse LS transform of Z1
est dewt ¼
s ; sþw
0
the transition probability is h2 2k w1 k h w1 t w2 k h w2 t þ e e : P00 ðtÞ ¼ 2 w1 w2 2k þ 2kh þ h2 w1 w2 Similarly, 2kh 2k w1 þ h w1 t w2 þ h w2 t þ e e ; w1 w2 2k2 þ 2kh þ h2 w1 w2 2k2 2k2 1 w1 t 1 w2 t þ e e ; P02 ðtÞ ¼ 2 w2 2k þ 2kh þ h2 w1 w2 w1 P01 ðtÞ ¼
where note that P00 ðtÞ þ P01 ðtÞ þ P02 ðtÞ ¼ 1. Thus, the limiting probabilities Pj limt!1 Pij ðtÞ are
4.2 Continuous-Time Markov Chain
P0
117
h2 ; 2k2 þ 2kh þ h2
P1
2kh ; 2k þ 2kh þ h2 2
and P2 ¼ 1 P0 P1 that represents the unavailability for a two-unit parallel system with one repair person. Next, suppose that there exists two repair persons. In this case, any units are repaired immediately upon failures. Then, setting k0 ¼ 2k, k1 ¼ k, l1 ¼ h, and l2 ¼ 2h, P000 ðtÞ ¼ 2kP00 ðtÞ þ hP01 ðtÞ; P001 ðtÞ ¼ 2kP00 ðtÞ ðk þ hÞP01 ðtÞ þ 2hP02 ðtÞ; P002 ðtÞ ¼ kP01 ðtÞ 2hP02 ðtÞ: By the same method of the previous model, s2 þ ðk þ 3hÞs þ 2h2 ; ðs þ 2k þ 2hÞðs þ k þ hÞ 2kðs þ 2hÞ ; P01 ðsÞ ¼ ðs þ 2k þ 2hÞðs þ k þ hÞ
P00 ðsÞ ¼
P02 ðsÞ ¼
2k2 : ðs þ 2k þ 2hÞðs þ k þ hÞ
Forming the inverse LS transforms of P0j ðsÞ, 2 h þ keðkþhÞt ; kþh i 1 h i 2k h P01 ðtÞ ¼ 1 eðkþhÞt h þ keðkþhÞt ; kþh kþh i 2 k h ðkþhÞt 1e P02 ðtÞ ¼ kþh P00 ðtÞ ¼
(Problem 4.5). Example 4.12 We compute explicitly Pj ðtÞ (j ¼ 0; 1; 2) given in (4.2). Forming the LS transforms of (4.2), sP0 ðsÞ s ¼ kP0 ðsÞ þ h1 P1 ðsÞ þ h2 P2 ðsÞ; sP1 ðsÞ ¼ h1 P1 ðsÞ þ akP0 ðsÞ; sP2 ðsÞ ¼ h2 P2 ðsÞ þ ð1 aÞkP0 ðsÞ:
118
4 Markov Chains
Solving these equations, ðs þ h1 Þðs þ h2 Þ ; þ sðk þ h1 þ h2 Þ þ ð1 aÞkh1 þ akh2 þ h1 h2 ak P1 ðsÞ ¼ P ðsÞ; s þ h1 0 ð1 aÞk P2 ðsÞ ¼ P ðsÞ: s þ h2 0 P0 ðsÞ ¼
s2
Forming the inverse LS transforms of Pj ðsÞ, h1 h2 ðw1 h1 Þðw1 h2 Þ w1 t e þ ð1 aÞkh1 þ akh2 þ h1 h2 w1 ðw1 w2 Þ ðw2 h1 Þðw2 h2 Þ w2 t e þ ; w2 ðw2 w1 Þ akh2 akðw1 h2 Þ w1 t P1 ðtÞ ¼ e ð1 aÞkh1 þ akh2 þ h1 h2 w1 ðw1 w2 Þ akðw2 h2 Þ w2 t e ; w2 ðw2 w1 Þ ð1 aÞkh1 ð1 aÞkðw1 h1 Þ w1 t P2 ðtÞ ¼ e ð1 aÞkh1 þ akh2 þ h1 h2 w1 ðw1 w2 Þ ð1 aÞkðw2 h1 Þ w2 t e ; w2 ðw2 w1 Þ P0 ðtÞ ¼
where
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 w1 k þ h1 þ h2 þ ðk þ h1 þ h2 Þ2 4½ð1 aÞkh1 þ akh2 þ h1 h2 ; 2 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 k þ h1 þ h2 ðk þ h1 þ h2 Þ2 4½ð1 aÞkh1 þ akh2 þ h1 h2 : w2 2
From (4.29) and (4.30), the transition probability matrix satisfies the differential equation P0 ðtÞ ¼ PðtÞA; where the initial condition of Pð0Þ ¼ I, where 0 k0 k0 0 0 B l1 ðk1 þ l1 Þ k 0 1 B B 0 ðk þ l Þ k l 2 2 2 2 B AB . .. .. .. B .. . . . B @ 0 0 0 0 0 0 0 0
ð4:31Þ
I is an identity matrix and ... ... ...
0 0 0 .. .
0 0 0 .. .
1
C C C C C: C ... C . . . ðkm1 þ lm1 Þ km1 A ... lm lm
4.2 Continuous-Time Markov Chain
119
Solving (4.31), PðtÞ ¼ eAt ¼ I þ
1 X An t n
n!
n¼1
:
ð4:32Þ
In general, we can compute the transition probabilities from (4.32), however, it is difficult to do so except in simple cases.
4.2.3 Limiting Probabilities If all states communicate with each other, i.e., the process consists of one recurrent class, then there exists the limiting probabilities Pj lim Pij ðtÞ [ 0
ðj ¼ 0; 1; 2; . . .; mÞ;
t!1
Pm
Pj ¼ 1, as shown in Sect. 4.1.3.
lim P0ij ðtÞ ¼ 0:
ð4:33Þ
that are independent of the initial State i and Furthermore,
j¼0
t!1
It can be clearly seen that there exits limt!1 P0ij ðtÞ from (4.29) because the limiting probabilities exists. Thus, if P0ij ðtÞ 6¼ 0, then there exists T [ 0 such that P0ij ðtÞ c;
or
P0ij ðtÞ c
for t T and c [ 0. If P0ij ðtÞ c, then ZT0
P0ij ðtÞdt cðT0 TÞ
T
for T0 [ T, i.e., Pij ðT0 Þ Pij ðTÞ cðT0 TÞ; that leads to the mathematical contradiction because the right-hand side goes to infinity as T0 ! 1. Therefore, (4.33) is proved. It is also proved in the case of P0ij ðtÞ c. Using the above results, from (4.29) and (4.30), the limiting probabilities satisfy the simultaneous equations, k0 P0 þ l1 P1 ¼ 0; kj1 Pj1 ðkj þ lj ÞPj þ ljþ1 Pjþ1 ¼ 0 km1 Pm1 lm Pm ¼ 0:
ðj ¼ 1; 2; . . .; m 1Þ;
ð4:34Þ
120
4 Markov Chains
Summing for i ¼ 0; 1; 2; . . .; j 1 in (4.34), kj1 kj1 kj2 Pj1 ¼ Pj2 ¼ ¼ Pj ¼ lj lj lj1 Recalling that
Pm
j¼0
j Y kk1 k¼1
lk
! P0 :
Pj ¼ 1, Qj
Pj ¼ where
Q0
k¼1
ðkk1 =lk Þ Pmk¼1Qj 1 þ j¼1 k¼1 ðkk1 =lk Þ
ðj ¼ 0; 1; . . .; mÞ;
ð4:35Þ
1.
Example 4.13 (n-unit parallel system) Consider an n-unit parallel system with one repair person. Then, setting that kk ¼ ðn kÞk and lkþ1 ¼ h (k ¼ 0; 1; . . .; n) in (4.35), the limiting probabilities are ðk=hÞj =ðn jÞ! Pj ¼ Pn k k¼0 ðk=hÞ =ðn kÞ!
ðj ¼ 0; 1; 2; . . .; nÞ:
In particular, when n ¼ 2, Pj agree with the limiting probabilities Pj of a two-unit parallel system with one repair person in Example 4.11.
4.3 Problems 4 4.1 Prove in Example 4.3 that by the mathematical induction, a a b a 1 ð1 a bÞn Pn ¼ ; þ aþb aþb b a b b b þ að1 a bÞn a½1 ð1 a bÞn ; : pðnÞ ¼ aþb aþb When a þ b ¼ 1, show that for n 1, 1a a n ; pðnÞ ¼ ð1 a; aÞ: P ¼ 1a a 4.2 The unit begins to operate at time 0, and when it fails, it is replaced with a new one [4, p. 113]. In addition, if the unit has not failed until Nth period, it is also replaced with a new one. It is assumed that the unit has a failure probability pn at time n. Then, the probability that the unit with age n fails at time n is pn rn P1
k¼n
pk
;
4.3 Problems 4
121
that is called a discrete failure rate [5, p. 18, 6, p. 9]. Show that the transition probabilities are given by Pi1 ¼ 1 Piiþ1 ¼ ri
ði ¼ 1; 2; . . .; N 1Þ;
PN1 ¼ 1;
and derive the limiting probabilities pj (j ¼ 1; 2; . . .; N). 4.3 In Example 4.9, compute P and ðI PÞ1 by striking out row 5 and column 5 when kt0 ¼ 0:9 and kL ¼ 0:1. 4.4 In Example 4.10, prove that by using the binomial distribution with n and p 1 ekt , n ð1 ekt Þk eðnkÞkt : Pk ðtÞ ¼ k 4.5 Derive the transition probabilities of two-unit standby system with one and two repair persons, respectively.
References 1. 2. 3. 4. 5. 6.
Osaki S (1992) Applied stochastic system modeling. Springer, Berlin Kemeny JG, Snell JL (1960) Finite Markov chains. Nostrand Co, Princeton Tijms HC (2003) A first course in stochastic models. Wiley, Chichester Ross SM (1983) Stochastic processes. Wiley, New York Barlow RE, Proschan F (1965) Mathematical theory of reliability. Wiley, New York Nakagawa T (2005) Maintenance theory of reliability. Springer, London
Chapter 5
Semi-Markov and Markov Renewal Processes
We consider the same one-unit system in Chap. 4. If FðtÞ and Gj ðtÞ (j ¼ 1; 2) are not exponential, the process does not have the Markov property at any time t and has its property at only epochal times of transition to states. Let ZðtÞ denote the state of the process at time t; Nj ðtÞ (j ¼ 0; 1; 2) denote the number of times that the process visits State j in ½0; t when the system begins to operate at time 0, and define that Pj ðtÞ PrfZðtÞ ¼ jg;
Fj ðtÞ PrfNj ðtÞ [ 0g;
where Fj ðtÞ is called a first-passage distribution to State j: Then, using the techniques of forming renewal equations in Sect. 3.3, F0 ðtÞ ¼ aFðtÞ G1 ðtÞ þ ð1 aÞFðtÞ G2 ðtÞ; F1 ðtÞ ¼ aFðtÞ þ ð1 aÞFðtÞ G2 ðtÞ F1 ðtÞ; F2 ðtÞ ¼ ð1 aÞFðtÞ þ aFðtÞ G1 ðtÞ F2 ðtÞ;
ð5:1Þ
where the asterisk denotes the pairwise Stieltjes convolution, i.e., aðtÞ bðtÞ Rt 0 bðt uÞdaðuÞ: Furthermore, using Fj ðtÞ; the transition probabilities are P0 ðtÞ ¼ 1 FðtÞ þ F0 ðtÞ P0 ðtÞ; P1 ðtÞ ¼ F1 ðtÞ ½1 G1 ðtÞ þ G1 ðtÞ P1 ðtÞ;
ð5:2Þ
P2 ðtÞ ¼ F2 ðtÞ ½1 G2 ðtÞ þ G2 ðtÞ P2 ðtÞ: Thus, using Fj ðtÞ in (5.1), we can obtain the transition probabilities Pj ðtÞ: The process fZðtÞg is called a semi-Markov process because it has the Markov property only at regeneration points at which the state transitions take place. We investigate mainly the properties of the transition probabilities Pj ðtÞ of a semi-Markov process in Sect. 5.2 as an example of a standby redundant system.
T. Nakagawa, Stochastic Processes, Springer Series in Reliability Engineering, DOI: 10.1007/978-0-85729-274-2_5, Ó Springer-Verlag London Limited 2011
123
124
5
Semi-Markov and Markov Renewal Processes
Finally, define that renewal functions Mj ðtÞ EfNj ðtÞg (j ¼ 0; 1; 2) that represent the expected number of visits to State j in ½0; t: In a similar way of obtaining (5.2), M0 ðtÞ ¼ F0 ðtÞ ½1 þ M0 ðtÞ; M1 ðtÞ ¼ F1 ðtÞ ½1 þ G1 ðtÞ M1 ðtÞ; M2 ðtÞ ¼ F2 ðtÞ ½1 þ G2 ðtÞ M2 ðtÞ:
ð5:3Þ
Thus, using Fj ðtÞ in (5.1), we can obtain Mj ðtÞ from (5.3) (Problem 5.1). The process fNj ðtÞg is called a Markov renewal process that combines a renewal process in Chap. 3 and a Markov chain in Chap. 4. We investigate mainly the properties of Mj ðtÞ for a Markov renewal process in Sect. 5.3. However, there is no clear difference between semi-Markov and Markov renewal processes. Markov chain, Markov process, Semi-Markov process, and Markov renewal process are called Markov process in one word, i.e., Markov processes are a general term of the stochastic process with the Markov property. State space is usually defined by the number of units that are working satisfactorily. As far as the applications to reliability theory is concerned, we consider only a finite number of states, contrast with a queueing theory. We mention only the theory of stationary Markov processes with a finite-state space. It is shown that transition probabilities, first-passage distributions, and renewal functions are given by forming renewal equations. Furthermore, some limiting properties are summarized when all states communicate.
5.1 Markov Process Consider a stochastic process with a finite state space f0; 1; 2; . . .; mg that makes transitions from state to state in accordance with a Markov chain with stationary transition probabilities. However, in the process, the amount of time spent in each state until the next transition is not always constant and exponential, but has a random variable shown in Definition 5.1. Consider a stochastic process fXðnÞ; Sn ; n ¼ 0; 1; 2; . . .g where Sn denotes the random variable of the nth arrival time at which the process just moves from one state to another, and XðnÞ denotes the state of the process at time Sn : Definition 5.1 The stochastic process fXðnÞ; Sn g is called a Markov renewal process [1, p. 167, 2, p. 313] if PrfXðn þ 1Þ ¼ j; Snþ1 Sn tjXð0Þ; Xð1Þ; . . .; XðnÞ ¼ i; S0 ; S1 ; . . .; Sn g ¼ PrfXðn þ 1Þ ¼ j; Snþ1 Sn tjXðnÞ ¼ i; Sn g Qij ðtÞ for all i; j ¼ 0; 1; 2; . . .; m and t 0:
ð5:4Þ
5.1 Markov Process
125
The probability Qij ðtÞ denotes that after making a transition into State i; the next process makes a transition into State P j; in an amount of time less than or equal to t: Clearly, we have Qij ðtÞ 0 and m j¼0 Qij ð1Þ ¼ 1; where Qij ð1Þ represents the probability that the next process makes a transition into State j; given that the process goes into State i; i.e., Qij ð1Þ forms a discrete-time Markov chain in Sect. 4.1. We call the probability Qij ðtÞ a mass function or a one-step transition probability [1, p. 167]. Letting Gij ðtÞ ¼
Qij ðtÞ Qij ð1Þ
for Qij ð1Þ [ 0;
then Gij ðtÞ represents the conditional probability that the process makes a transition in an amount of time less than or equal to t; given that the process goes from State i to State j at the next transition. In other words, Gij ðtÞ ¼ PrfSnþ1 Sn tjXðnÞ ¼ i; Xðn þ 1Þ ¼ jg is the distribution of the sojourn time that the process takes in State i given that the next visiting state is j: Because XðnÞ denotes the state of the process immediately after the nth transition has occurred for n 1; the stochastic process fXðnÞ; n ¼ 0; 1; 2; . . .g is called an embedded Markov chain with transition probabilities Qij ð1Þ: If the process makes a transition from one state to another with one unit of time, i.e., Gij ðtÞ ¼ 0 for t\1; and 1 for t 1; then an embedded Markov chain becomes a discrete-time Markov chain in Sect. 4.1. Furthermore, if an amount of time spent in State i depends only on State i and is distributed exponentially independent of the next state; Gij ðtÞ ¼ 1 eki t for constant ki [ 0; the process is said to be a continuous-time Markov chain in Sect. 4.2. In addition, the process becomes a renewal process in Chap. 3 if it has only one state and becomes an alternating renewal process in Sect. 3.4 if it has two states. If we let ZðtÞ denote the state of the process at time t; then the stochastic process fZðtÞ; t 0g is called a semi-Markov process. Let Nj ðtÞ denote the number of times that the process visits State j in ½0; t: It follows from the renewal theory that with probability 1, Nj ðtÞ\1 for t 0: The stochastic process fN0 ðtÞ; N1 ðtÞ; N2 ðtÞ; . . .; Nm ðtÞg is called a Markov renewal process, and EfNj ðtÞg are called Markov renewal functions. The event fNj ðtÞ ¼ kg represents the number of k visits to State j in ½0; t: An embedded Markov chain records the state of the process at each transition point, a semi-Markov process records the state of the process at each time point t; and a Markov renewal process records the total number of times that each state has been visited in ½0; t: Letting Hi ðtÞ denote the distribution of an amount of time spent in State i until the process makes a transition to the next state, Hi ðtÞ
m X
Qij ðtÞ;
j¼0
that is called the unconditional distribution for State i: We suppose that Hi ð0Þ\1 for all i: Denoting
126
5
gi
Z1
Semi-Markov and Markov Renewal Processes
tdHi ðtÞ and
lij
0
Z1
tdGij ðtÞ;
0
it is easily seen that gi ¼
m X
Qij ð1Þlij ;
j¼0
which represents the mean time spent in State i: We denote transition probabilities, first-passage distributions, and Markov renewal functions as, respectively, Pij ðtÞ PrfZðtÞ ¼ jjZð0Þ ¼ ig; Fij ðtÞ PrfNj ðtÞ [ 0jZð0Þ ¼ ig; Mij ðtÞ EfNj ðtÞjZð0Þ ¼ ig: We have the following renewal equations that show the relationships for Pij ðtÞ; Fij ðtÞ and Mij ðtÞ in terms of the mass functions Qij ðtÞ: Pii ðtÞ ¼ 1 Hi ðtÞ þ
m Z X k¼0
Pij ðtÞ ¼
t m Z X k¼0
t
Pki ðt uÞdQik ðuÞ;
0
ð5:5Þ
Pkj ðt uÞdQik ðuÞ for i 6¼ j;
0
Zt m X
Fij ðtÞ ¼ Qij ðtÞ þ
k¼0;k6¼j
Mij ðtÞ ¼ Qij ðtÞ þ
m Z X k¼0
Fkj ðt uÞdQik ðuÞ;
ð5:6Þ
0 t
Mkj ðt uÞdQik ðuÞ:
ð5:7Þ
0
Therefore, the mass functions Qij ðtÞ determine Pij ðtÞ; Fij ðtÞ and Mij ðtÞ uniquely. Furthermore, Pii ðtÞ ¼ 1 Hi ðtÞ þ
Zt
Pii ðt uÞdFii ðuÞ;
0
Pij ðtÞ ¼
Zt 0
Pjj ðt uÞdFij ðuÞ
ð5:8Þ for i 6¼ j;
5.1 Markov Process
127
Mij ðtÞ ¼ Fij ðtÞ þ
Zt
Mjj ðt uÞdFij ðuÞ:
ð5:9Þ
0
Taking the LS transforms of the above equations, Pii ðsÞ ¼
1 Hi ðsÞ ; 1 Fii ðsÞ
Pij ðsÞ ¼ Fij ðsÞPjj ðsÞ
for i 6¼ j;
Mij ðsÞ ¼ Fij ðsÞ½1 þ Mjj ðsÞ;
ð5:10Þ ð5:11Þ
where the asterisk denotes the LS transform of the function with itself. Thus, if the first-passage distributions Fij ðtÞ are obtained, then the transition probabilities Pij ðtÞ and renewal functions Mij ðtÞ are given by (5.10) and (5.11), respectively. Consider the process in which all states communicate, i.e., Gii ð1Þ ¼ 1; lii \1 for all i; and each Gii ðtÞ is a nonlattice distribution. It is said that the process consists of one positive recurrent class. Then, Gij ð1Þ ¼ 1;
lij \1; n X Qik ð1Þlkj : lij ¼ gi þ k¼0;k6¼j
Furthermore, lim Mij ðtÞ ¼
Z1
t!1
Pij ðtÞdt;
0
Mij ðtÞ 1 ¼ \1; t!1 t ljj lim
lim Pij ðtÞ ¼
t!1
gj \1: ljj
In this case, there exist limt!1 Mij ðtÞ=t and limt!1 Pij ðtÞ: From a Tauberian theorem in Appendix A.3 that if for some nonnegative integer n; lims!0 sn U ðsÞ ¼ C; then limt!1 UðtÞ=tn ¼ C=n!; Mij ðtÞ ¼ lim sMij ðsÞ; s!0 t Zt 1 Pij ðuÞdu ¼ lim Pij ðsÞ: lim Pij ðtÞ ¼ lim t!1 t!1 t s!0
lim
t!1
0
Example 5.1 (Alternating renewal process) An alternating renewal process in Sect. 3.4.1 has the mass functions Q01 ðtÞ ¼ FðtÞ and Q10 ðtÞ ¼ GðtÞ: Then, from (5.5),
128
5
P00 ðtÞ ¼ 1 Q01 ðtÞ þ
Zt
Semi-Markov and Markov Renewal Processes
P10 ðt uÞdQ01 ðuÞ
0
¼ FðtÞ þ
Zt
P10 ðt uÞdFðuÞ;
0
P11 ðtÞ ¼ 1 Q10 ðtÞ þ
Zt
P01 ðt uÞdQ10 ðuÞ
0
¼ GðtÞ þ
Zt
P01 ðt uÞdGðuÞ;
0
P01 ðtÞ ¼
P10 ðtÞ ¼
Zt
P11 ðt uÞdQ01 ðuÞ ¼
Zt
0
0
Zt
Zt
P00 ðt uÞdQ10 ðuÞ ¼
0
P11 ðt uÞdFðuÞ;
P00 ðt uÞdGðuÞ;
0
that agree with (3.54). Furthermore, from (5.7), M00 ðtÞ ¼
M11 ðtÞ ¼
Zt
M10 ðt uÞdQ01 ðuÞ ¼
Zt
0
0
Zt
Zt
M01 ðt uÞdQ10 ðuÞ ¼
0
M01 ðtÞ ¼ Q01 ðtÞ þ
M10 ðtÞ ¼ Q10 ðtÞ þ
M10 ðt uÞdFðuÞ;
M01 ðt uÞdGðuÞ;
0
Zt
M11 ðt uÞdQ01 ðuÞ ¼ FðtÞ þ
Zt
0
0
Zt
Zt
0
M00 ðt uÞdQ10 ðuÞ ¼ GðtÞ þ
M11 ðt uÞdFðuÞ;
M00 ðt uÞdGðuÞ;
0
that agrees with (3.52). Example 5.2 (Intermittent fault) [3, p. 113] Faults occur in a communication system according to an exponential distribution ð1 ekt Þ and are hidden. When the duration time X in hidden fault exceeds an upper limit time Y; faults become permanent failures eternally, and otherwise, they get out of a hidden state. That is, if the event fX Yg occurs, then hidden faults disappear, and conversely, if the event fX [ Yg occurs, then they become permanent failure. It is assumed that both
5.1 Markov Process
129
random variables X and Y are independent and have exponential distributions PrfX tg ¼ 1 elt and PrfY tg ¼ 1 eht ; respectively. We define the following states of intermittent faults: State 0: No fault occurs and the system is in a normal condition. State 1: Hidden fault occurs. State 2: Permanent failure occurs. We have the following mass functions Qij ðtÞ (i ¼ 0; 1; j ¼ 0; 1; 2) from State i to State j in time t from Fig. 5.1: Q01 ðtÞ ¼ 1 ekt ; Q10 ðtÞ ¼ PrfX Y; X tg ¼
Zt
ehu l elu du ¼
i l h 1 eðlþhÞt ; lþh
0
Q12 ðtÞ ¼ PrfX [ Y; Y tg ¼
Zt
elu h ehu du ¼
ð5:12Þ
i h h 1 eðlþhÞt : lþh
0
Let Pij ðtÞ denote the transition probabilities from State i at time 0 to State j at time t: Then, we have the following renewal equations: P00 ðtÞ ¼ 1 Q01 ðtÞ þ Q01 ðtÞ P10 ðtÞ; P01 ðtÞ ¼ Q01 ðtÞ P11 ðtÞ; P10 ðtÞ ¼ Q10 ðtÞ P00 ðtÞ; P11 ðtÞ ¼ 1 Q10 ðtÞ Q12 ðtÞ þ Q10 ðtÞ P01 ðtÞ;
ð5:13Þ
P02 ðtÞ ¼ Q01 ðtÞ P12 ðtÞ; P12 ðtÞ ¼ Q12 ðtÞ þ Q10 ðtÞ P02 ðtÞ: Taking the LS transforms of (5.13) and rearranging them, 1 Q01 ðsÞ ; 1 Q01 ðsÞQ10 ðsÞ Q01 ðsÞ 1 Q10 ðsÞ Q12 ðsÞ P01 ðsÞ ¼ ; 1 Q01 ðsÞQ10 ðsÞ Q ðsÞ 1 Q01 ðsÞ ; P10 ðsÞ ¼ 10 1 Q01 ðsÞQ10 ðsÞ 1 Q10 ðsÞ Q12 ðsÞ : P11 ðsÞ ¼ 1 Q01 ðsÞQ10 ðsÞ P00 ðsÞ ¼
Fig. 5.1 Transition diagram of intermittent fault
ð5:14Þ
130
5
Semi-Markov and Markov Renewal Processes
Thus, substituting (5.12) for (5.14) and taking the inverse LS transforms, the transition probabilities from State i to State j (i; j ¼ 0; 1; 2) are 1 ½ðl þ h x2 Þex2 t ðl þ h x1 Þex1 t ; x1 x2 k l ðex2 t ex1 t Þ; P10 ðtÞ ¼ ðex2 t ex1 t Þ; P01 ðtÞ ¼ x1 x2 x 1 x2 1 P11 ðtÞ ¼ ½ðk x2 Þex2 t ðk x1 Þex1 t ; x1 x2 P02 ðtÞ ¼ 1 P00 ðtÞ P01 ðtÞ; P12 ðtÞ ¼ 1 P10 ðtÞ P11 ðtÞ; P00 ðtÞ ¼
ð5:15Þ
where qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 x1 k þ l þ h þ ðk þ l þ hÞ2 4kh ; 2 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 x2 k þ l þ h ðk þ l þ hÞ2 4kh : 2 Next, let Mij ðtÞ (i; j ¼ 0; 1) be the expected number of visits to State j from State i before State 2 in time t: Then, we have the following renewal equations: M00 ðtÞ ¼ Q01 ðtÞ M10 ðtÞ; M11 ðtÞ ¼ Q10 ðtÞ M01 ðtÞ;
M01 ðtÞ ¼ Q01 ðtÞ ½1 þ M11 ðtÞ; M10 ðtÞ ¼ Q10 ðtÞ ½1 þ M00 ðtÞ:
ð5:16Þ
Taking the LS transforms of (5.16) and rearranging them, M00 ðsÞ ¼
Q01 ðsÞQ10 ðsÞ ; 1 Q01 ðsÞQ10 ðsÞ
ðsÞ ¼ M01
Q01 ðsÞ : 1 Q01 ðsÞQ10 ðsÞ
ð5:17Þ
Thus, substituting (5.12) for (5.17) and taking the inverse LS transforms, kl 1 1 x2 t x1 t ð1 e Þ ð1 e Þ ; M00 ðtÞ ¼ x 1 x 2 x2 x1 ð5:18Þ k x1 l h x2 l h x1 t x2 t ð1 e Þ ð1 e Þ : M01 ðtÞ ¼ x1 x2 x1 x2 Example 5.3 (ARQ model) ARQ (Automatic-Repeat-Request) models with intermittent faults that are mainly employed in data transmission to achieve high reliability of communication [3, p. 110]: ARQ strategies are adopted when data transmission due to faults, fail. However, the data throughput decreases significantly if retransmissions are repeated without limitation. To keep the level of data throughput, when all numbers N of transmissions have failed, the system is inspected and maintained. We repeat the above procedure until the data transmission succeeds.
5.1 Markov Process
131
It is assumed in Example 5.2 that when the system is in State 0, the data transmission succeeds with probability 1, when the system is in State 1, it succeeds with probability q and fails with probability p 1 q; and when the system is in State 2, it fails with probability 1. When the data transmission fails, the retransmission incurs in time T1 : When N retransmissions have failed consecutively including the first one, the system is inspected and preventively maintained, and the same data transmission begins new after time T2 : Suppose that the system is in State 0 at time 0. Then, the probability that the data transmission succeeds until the Nth number is PðNÞ ¼ 1 pP01 ðT1 Þ P02 ðT1 Þ þ pP01 ðT1 Þ½1 pP11 ðT1 Þ P12 ðT1 Þ
N 2 X
½ pP11 ðT1 Þj ;
j¼0
and the probability that all N retransmissions have failed is ( ) N2 X j N1 ; PðNÞ ¼ P02 ðT1 Þ þ pP01 ðT1 Þ P12 ðT1 Þ ½ pP11 ðT1 Þ þ½ pP11 ðT1 Þ j¼0
where PðNÞ þ PðNÞ ¼ 1: We call the time from the beginning of transmission to N failures or success as one period. Then, the mean time of one period is ( l1 ¼ T1 1 pP01 ðT1 Þ P02 ðT1 Þ þ pP01 ðT1 Þ½1 pP11 ðT1 Þ P12 ðT1 Þ (
N2 X
) j
ðj þ 2Þ½ pP11 ðT1 Þ þNPðNÞ
j¼0
pP01 ðT1 ÞP12 ðT1 Þ ¼ T1 1 þ ðN 1Þ P02 ðT1 Þ þ 1 pP11 ðT1 Þ
) 1 pP11 ðT1 Þ P12 ðT1 Þ 1 ½ pP11 ðT1 ÞN1 þ pP01 ðT1 Þ : 1 pP11 ðT1 Þ 1 pP11 ðT1 Þ
Thus, the mean time in which the data transmission succeeds is l2 ðNÞ ¼ l1 þ PðNÞ½T2 þ l2 ðNÞ; i.e., l1 þ T2 PðNÞ PðNÞ T1 A þ ðN 1ÞB n o T2 þ ¼ 1 pP11 ðT1 Þ C þ D 1 ½ pP11 ðT1 ÞN1
l2 ðNÞ ¼
ðN ¼ 1; 2; . . .Þ; ð5:19Þ
132
5
Semi-Markov and Markov Renewal Processes
where T1 ½P00 ðT1 Þ þ qP01 ðT1 Þ ; A T1 þ T 2 1 pP11 ðT1 Þ pP01 ðT1 ÞP12 ðT1 Þ ; B T1 P02 ðT1 Þ þ 1 pP11 ðT1 Þ C P00 ðT1 Þ þ qP01 ðT1 Þ;
D pP01 ðT1 Þ
P10 ðT1 Þ þ qP11 ðT1 Þ : 1 pP11 ðT1 Þ
From the inequality l2 ðN þ 1Þ l2 ðNÞ 0; 1 þ C=D ½1 pP11 ðT1 Þ½ pP11 ðT1 Þ
N1
ðN 1Þ
A 1 þ B 1 pP11 ðT1 Þ
ðN ¼ 1; 2; . . .Þ: ð5:20Þ
Letting LðNÞ be the left-hand side of (5.20), Lð1Þ lim LðNÞ ¼ 1; N!1
LðN þ 1Þ LðNÞ ¼
Lð1Þ ¼
1 þ C=D ; 1 pP11 ðT1 Þ
1 þ C=D ½pP11 ðT1 ÞN [ 0: ½P11 ðT1 ÞN
Therefore, an optimum N (1 N \1) that minimizes l2 ðNÞ in (5.19) is given by a finite and unique minimum that satisfies (5.20). If C=D ðA=BÞ½1 pP11 ðT1 Þ; then N ¼ 1 (Problem 5.2).
5.2 Embedded Markov Chain Consider a standby redundant system with n þ 1 repair persons and one operating unit supported by n identical spare units. Each unit fails according to a general distribution FðtÞ with finite mean l and undergoes repair immediately. When failed units are repaired, they become as good as new units and rejoin spare ones. It is assumed that the repair time of each failed unit is an independent random variable with an identical exponential distribution ð1 eht Þ with finite mean 1=h: Let denote by ZðtÞ the number of units undergoing repair at time t: We say that the system is in State j at time t if ZðtÞ ¼ j (j ¼ 0; 1; 2; . . .; n þ 1). We also say that system failure occurs when the system is in Sate n þ 1: Denote by s0 ; s1 ; . . .; sk ; . . . the failure times of operating units. For simplicity of computations, we restrict ourselves to the case of s0 0: If we denote Zk Zðsk 0Þ for k ¼ 1; 2; . . .; then Zk represents the number of units undergoing repair immediately before the kth failure occurrence, where Z0 0: It is noted that the process fZk ; k ¼ 0; 1; 2; . . .g
5.2 Embedded Markov Chain
133
forms an embedded Markov chain and fZðtÞ; t 0g forms a semi-Markov process with stationary transition probabilities and a finite state space.
5.2.1 Transition Probabilities We derive the transition probabilities by using binomial moments [4] and the expected number of failures before system failure: Let denote the transition probabilities as ðkÞ
Pij ðtÞ ¼ PrfZk ¼ j; sk tjZ0 ¼ ig
ði; j ¼ 0; 1; 2; . . .; nÞ
for k ¼ 1; 2; . . .: By similar arguments of obtaining the Chapman–Kolmogorov equations of (4.4) and (4.25), ðkÞ Pij ðtÞ
¼
n Z X l¼0
t ð1Þ
ðk1Þ
Plj ðt uÞdPil
ðuÞ:
ð5:21Þ
0
Furthermore, because Z1 has a binomial distribution with parameters i þ 1 and ehu under the given conditions, PrfZ1 ¼ jjs1 s0 ¼ u; Z0 ¼ ig 8 < i þ 1 ð1 ehu Þiþ1j ejhu ¼ j : 0
ði ¼ 0; 1; . . .; n 1; j i þ 1Þ; ðj [ i þ 1Þ:
Dropping the condition s1 s0 ¼ u; i.e., the operating unit fails in ðu; u þ du; 8 Z t iþ1 < ð1 ehu Þiþ1j ejhu dFðuÞ ði ¼ 0; 1; . . .; n 1; j i þ 1Þ; ð1Þ Pij ðtÞ ¼ j 0 : 0 ðj [ i þ 1Þ: ð5:22Þ If Z0 ¼ n; then ð1Þ Pnj ðtÞ
¼
Zt
ð1Þ
ðn þ 1Þh eðnþ1Þhu Pn1j ðt uÞdu:
ð5:23Þ
0 ðkÞ
Next, introduce the rth binomial moment of the LS transform of Pij ðtÞ as follows:
134
5
Semi-Markov and Markov Renewal Processes
1
ðkÞ Uir ðsÞ
n Z X j j¼r
r
ðkÞ
est dPij ðtÞ
ðr ¼ 0; 1; 2; . . .; nÞ
0
for k ¼ 1; 2; . . .: Then, taking the LS transforms of (5.21)–(5.23) and arranging them, iþ1 n þ 1 ðs þ rhÞdin ð1Þ ; Uir ðsÞ ¼ F ðs þ rhÞ s þ ðn þ 1Þh r r ðkÞ ðk1Þ ðk1Þ Uir ðsÞ ¼ F ðs þ rhÞ Uir ðsÞ þ Uir1 ðsÞ nþ1 s þ rh ðk1Þ ðsÞ ðr ¼ 0; 1; 2; . . .; nÞ; U s þ ðn þ 1Þh in r
ð5:24Þ
where din is the Kronecker delta, i.e., din 0 for i 6¼ n and din ¼ 1 for i ¼ n: From ðkÞ (5.24), Uir ðsÞ can be determined recursively for every k and r: It is more convenient to introduce the generating function in Appendix A.4 Uir ðz; sÞ
1 X
ðkÞ
zk Uir ðsÞ
for jzj 1:
k¼1
Taking the generating function of (5.24), iþ1 Uir ðz; sÞ ¼ zF ðs þ rhÞ Uir ðz; sÞ þ Uir1 ðz; sÞ þ r
nþ1 s þ rh ½Uin ðz; sÞ þ din : s þ ðn þ 1Þh r Solving for Uir ðz; sÞ; iþ1 zF ðs þ rhÞ Uir ðz; sÞ ¼ Uir1 ðz; sÞ þ 1 zF ðs þ rhÞ r
nþ1 s þ rh ½Uin ðz; sÞ þ din : s þ ðn þ 1Þh r Denote Cr ðz; sÞ
r Y
zF ðs þ lhÞ 1 zF ðs þ lhÞ l¼0
ðr ¼ 0; 1; . . .; nÞ;
C1 ðz; sÞ 1:
ð5:25Þ
5.2 Embedded Markov Chain
135
Then, dividing both side of (5.25) by Cr ðz; sÞ; iþ1 Uir ðz; sÞ Uir1 1 ¼ þ Cr ðz; sÞ Cr1 ðz; sÞ Cr1 ðz; sÞ r nþ1 s þ rh Uin ðz; sÞ þ din C r r1 ðz; sÞ s þ ðn þ 1Þh
ðr ¼ 1; 2; . . .; nÞ;
Ui0 ðz; sÞ Uin ðz; sÞ þ din ¼1s : s þ ðn þ 1Þh C0 ðz; sÞ
ð5:26Þ
ð5:27Þ
Writing down (5.26) for r 1; r 2; . . .; 1; adding them, and from (5.27), r r X iþ1 nþ1 Uir ðz; sÞ X 1 s þ jh Uin ðz; sÞ þ din ¼ ðz; sÞ ðz; sÞ s þ ðn þ 1Þh Cr ðz; sÞ C C l l r1 l1 l¼0 l¼0 ðr ¼ 0; 1; . . .; nÞ:
ð5:28Þ
Setting r ¼ n in (5.28) and rearranging it, Piþ1 iþ1 =Cl1 ðz; sÞ Uin ðz; sÞ þ din ¼ Pnþ1 l¼0 : l nþ1 s þ ðn þ 1Þh ðs þ lhÞ=Cl1 ðz; sÞ l¼0
ð5:29Þ
l
Thus, from (5.28), " Uir ðz; sÞ ¼ Cr ðz; sÞ
r X iþ1
1 ðz; sÞ C l l1 l¼0 # Pr nþ1 iþ1 X iþ1 ðz; sÞ ðs þ lhÞ=C 1 l1 l¼0 l ; P nþ1 Cl1 ðz; sÞ nþ1 l ðs þ lhÞ=Cl1 ðz; sÞ l¼0 l¼0 l ð5:30Þ
that is given in terms of the binomial moment. By inverting the above binomial moment,
Pij ðz; sÞ
1 X k¼1
k
Z1
z
e 0
st
ðkÞ dPij ðtÞ
¼
n X r r¼j
j
ð1Þrj Uir ðz; sÞ:
ð5:31Þ
136
5
Semi-Markov and Markov Renewal Processes ðkÞ
Therefore, the transition probabilities Pij PrfZk ¼ jjZ0 ¼ ig for the process fZk ; k ¼ 0; 1; . . .g are given by the generating function Pij ðzÞ
1 X
ðkÞ zk Pij
¼
k¼1
lim Pij ðz; sÞ s!0
n X r
¼
r¼j
j
ð1Þrj Uir ðz; 0Þ:
ð5:32Þ
The total expected number of visits to State j in ½0; t immediately before the P ðkÞ operating unit fails, initially starting from State i; is 1 k¼1 Pij ðtÞ: In particular, the P ðkÞ total expected number of system failures in ½0; t is 1 k¼1 Pin ðtÞ:
5.2.2 First-Passage Distributions We derive the first-passage distributions using the results of transition probabilities in Sect. 5.2.1: Let denote the first-passage distribution ðkÞ
Fij ðtÞ PrfZk ¼ j; Zl 6¼ j for l ¼ 1; 2; . . .; k 1; sk tjZ0 ¼ ig ði; j ¼ 0; 1; . . .; nÞ for k ¼ 1; 2; . . . : Then, by similar methods of obtaining (4.9), ð1Þ
ð1Þ
Pij ðtÞ ¼ Fij ðtÞ; ðkÞ Pij ðtÞ
¼
ðkÞ Fij ðtÞ
þ
k1 Z X l¼1
t ðklÞ
Pjj
ðlÞ
ðt uÞdFij ðuÞ
ð5:33Þ
0 ðkÞ
for k ¼ 2; 3; . . . : The first-passage distribution Fij ðtÞ can be calculated recursively from (5.33) and are uniquely determined. Introduce the following notation
Fij ðz; sÞ
1 X k¼1
k
Z1
z
ðkÞ
est dFij ðtÞ for jzj 1:
0
Then, taking the generating functions and the LS transforms of (5.33), Fij ðz; sÞ ¼
Pij ðz; sÞ : 1 þ Pjj ðz; sÞ
ð5:34Þ
5.2 Embedded Markov Chain
137
Substituting (5.31) for (5.34), Pn r rj Cr ðz; sÞ r¼j j ð1Þ hP P nþ1 r iþ1 =Cl1 ðz; sÞ nþ1 ðs þ lhÞ=Cl1 ðz; sÞ l¼0 l¼0 l l i Piþ1 iþ1 Pr nþ1 ð5:35Þ l¼0 l =Cl1 ðz; sÞ l¼0 l ðs þ lhÞ=Cl1 ðz; sÞ : Fij ðz; sÞ ¼ P jþ1 jþ1 =Cl1 ðz; sÞ l¼0 hP l i Pnþ1 nþ1 r n rj C ðz; sÞ ðz; sÞ ð1Þ ðs þ lhÞ=C r l1 r¼j j l¼rþ1 l Therefore, the first-passage distribution ðkÞ
Fij PrfZk ¼ j; Zl 6¼ k for l ¼ 1; 2; . . .; k 1; jZ0 ¼ ig is given by the generating function Fij ðzÞ
1 P
ðkÞ
zk Fij
k¼1
hP P n rj r iþ1 C ðzÞ ð1Þ =Cl1 ðzÞ nl¼0 j =Cl ðzÞ r r¼j l¼0 l i Pr1 n Piþ1 iþ1 ðzÞ ðzÞ =C =C l¼0 l1 l l¼0 l l hP i ; ¼P Pn n r jþ1 jþ1 n rj ðzÞ C ðzÞ ðzÞ =C ð1Þ =C l1 r l l¼0 r¼j j l¼r l l Pn r j
ð5:36Þ
where Cr ðzÞ Cr ðz; 0Þðr ¼ 1; 0; 1; . . .Þ: In particular, for i\j; Fij ðzÞ
Piþ1 iþ1
l ¼ Pl¼0 jþ1 iþ1 l¼0
l
=Cl1 ðzÞ =Cl1 ðzÞ
:
Next, let lij be the mean first-passage time from State i to State j: Then, from (5.35), for i\j; jþ1 X 1 Fij ð1; sÞ ¼l s!0 s l¼1
lij lim
jþ1 l
iþ1 l
1 ; Cl1
ð5:37Þ
Q where Cr rl¼1 fF ðlhÞ=½1 F ðlhÞg (r ¼ 1; 2; . . .) and C0 1: In particular, when i ¼ 1 and j ¼ n; l1n agrees with [5]. Note that State -1 means the initial condition that one unit begins to operate and n units are in standby at time 0.
138
5
Semi-Markov and Markov Renewal Processes
5.2.3 Expected Numbers of Visits to States We obtain two reliability quantities from the first-passage distributions; (1) the expected number of visits to State j before system failure and (2) the expected number of failures and repairs before system failure. First, we derive the probabilities of visits to State j at least once before returning ðkÞ to State i: Let denote fm;ij (k ¼ 1; 2; . . .) for an embedded Markov chain fZk ; k ¼ 0; 1; 2; . . .g the probability of first-passage from State i to State k without ðkÞ passing through State m: In general, fm;ij is known as taboo probability for a Markov chain. We evidently have, for j 6¼ m; ð1Þ
ð1Þ
ðkÞ
ðkÞ
Fij ¼ fm;ij ; Fij ¼ fm;ij þ
k1 X
ðklÞ ðlÞ
ðk ¼ 2; 3; . . .Þ;
Fmj fj;im
ð5:38Þ
l¼1
because, the first-passage from State i to State j for a Markov chain fZk ; k ¼ 1; 2; . . .g occurs without passing through State m (6¼ j) or by passing through State m: By exchanging j and m in (5.38), (5.38) is hold and uniquely deterðkÞ mine fm;ij : Introduce the following generating function Fm;ij ðzÞ
1 X
ðkÞ
zk fm;ij
for jzj 1:
k¼1
Then, from (5.38), for j 6¼ m; ðzÞ ¼ Fm;ij
Fij ðzÞ Fim ðzÞFmj ðzÞ : 1 Fjm ðzÞFmj ðzÞ
ð5:39Þ
ðzÞ and the probability Substituting (5.36) for (5.39), we can obtain Fm;ij
fm;ij
1 X k¼1
ðkÞ
fm;ij ¼ lim Fm;ij ðzÞ z!1
ð5:40Þ
of the first-passage from State i to State j without passage through State m: Furthermore, because fi;ij ¼ 1 fj;ii (Problem 5.3), from (5.37) and (5.39), ," # X j jþ1 X 1 jþ1 ri r ð1Þ fi;ij ¼ 1 Cr for i\j n; ð5:41Þ i l Cl1 r¼i l¼rþ1 that is the probability of visits to State j at least once before returning to State i:
5.2 Embedded Markov Chain
139
Next, we give two reliability quantities from the probability fj;jn : Let denote by ðkÞ Mj
(j ¼ 0; 1; . . .; n 1) the number of visits to State j for a Markov chain fZk ; k ¼ 1; 2; . . .g; including the first before system failure. Clearly, ðkÞ
PrfMj
X
¼ lg ¼
ðk Þ ðk Þ
ðk
Þ ðk Þ
fn;jj1 fn;jj2 fn;jjj1 fj;jnj
ðk; l ¼ 1; 2; . . .Þ:
k1 þk2 þ þkj ¼k;k1 ;k2 ;...;kj [ 0
ð5:42Þ By operating the convolution in (5.42), 1 X
ðkÞ
¼ lg ¼ ½Fn;jj ðzÞl1 Fj;jn ðzÞ;
ðkÞ
¼ lg ¼ ð1 fj;jn Þl1 fj;jn :
zk PrfMj
ð5:43Þ
k¼1
and from (5.40) and (5.43), 1 X
PrfMj
k¼1
Thus, the expected number Mj of visits to State j before system failure is Mj ¼
1 X 1 X 1 ðkÞ l PrfMj ¼ lg ¼ ; fj;jn l¼1 k¼1
and from (5.41), Mj ¼
n X r¼j
rj
ð1Þ
X nþ1 1 r nþ1 Cr : j l C l1 l¼rþ1
ð5:44Þ
Let M denote the total expected number of failures before system failures, when one unit starts to operate and n units are in standby as spare ones at time 0: Then,
M ¼1þ
n1 X
Mj ;
j¼0
because each failure causes one visit before system failure and the last failure causes system failure. From (5.44), nþ1 X 1 nþ1 M¼ : j Cj1 j¼1 The expected number of repairs before system failure is M ðn þ 1Þ: Note that lM is also the mean time to system failure that agrees with l1n in (5.37).
140
5
Semi-Markov and Markov Renewal Processes
5.2.4 Optimization Problems Consider the total cost of the operation for the system before system failure. It is assumed that the cost of each unit is c0 ; the repair cost incurred for each failed unit is c1 ; and the cost incurred for the system failure is c2 : Then, the total expected cost including the cost of system failure is ðn þ 1Þc0 þ ½M ðn þ 1Þc1 þ c2 ; and hence, the expected cost rate is CðnÞ ¼
ðn þ 1Þðc0 c1 Þ þ Mc1 þ c2 : lM
ð5:45Þ
We can easily determine an optimum number of spare units that minimizes the total expected cost before system failure. If the total cost C before system failure is previously given, we can find an upper bound of the number of spare units such that ðn þ 1Þc0 þ ½M ðn þ 1Þc1 þ c2 C:
ð5:46Þ
We can determine an optimum number of spare units that minimizes (5.45) within this number. In this case, because CðnÞ in (5.45) decreases to c0 =l (Problem 5.4), an optimum n is given by a unique maximum that satisfies (5.46).
5.3 Markov Renewal Process with Nonregeneration Points This section explains unique modifications of Markov renewal processes and applies them to redundant repairable systems including some nonregeneration points [6]. It has already been shown that such modifications give powerful plays for analyzing two-unit redundant systems [7] and communication systems [8]. It is assumed that the Markov renewal process under consideration has only one positive recurrent class, because we restrict ourselves to applications of reliability models. Consider the case where epochs at which the process enters some states are not regeneration points. Then, we partition a state space S into S ¼ S [ Sy (S \ Sy ¼ ;), where S is the portion of the state space such that the epoch entering State i (i 2 S ) is not a regeneration point, and Sy is such that the epoch entering State i (i 2 Sy ) is a regeneration point, where S and Sy are assumed not to be empty. Define the mass function Qij ðtÞ from State i (i 2 Sy ) to State j (j 2 S) by the probability that after entering State i; the process makes a transition into State j; in an amount of time less than or equal to t: However, it is impossible to define mass functions Qij ðtÞ for i (i 2 S ), because the epoch entering State i is not a regenðk ;k2 ;...;kl Þ
eration point. We define the new mass function Qij 1
ðtÞ that is the probability
5.3 Markov Renewal Process with Nonregeneration Points
141
that after entering State i (i 2 Sy ), the process next makes transition into State k1 ; k2 ; . . .; kl (k1 ; k2 ; . . .; kl 2 S ), and finally, enters State j (j 2 S), in an amount of time less than or equal to t:
5.3.1 Type 1 Markov Renewal Process Consider a Markov renewal process with m þ 1 states that consists of Sy ¼ f0g and S ¼ f1; 2; . . .; mg in Fig. 5.2. The process starts in State 0, i.e., Zð0Þ ¼ 0; and makes transitions into State 1; 2; . . .; m; and comes back to State 0. Then, from straightforward renewal arguments, the first-passage distributions are F01 ðtÞ ¼ Q01 ðtÞ; ð1;2;...;j1Þ
F0j ðtÞ ¼ Q0j
ð1;2;...;mÞ
F00 ðtÞ ¼ Q00
ðtÞ ðj ¼ 2; 3; . . .; mÞ;
ð5:47Þ
ðtÞ;
the renewal functions are ð1;2;...;mÞ
ðtÞ M01 ðtÞ M01 ðtÞ ¼ Q01 ðtÞ þ Q00 ¼ Q01 ðtÞ þ F00 ðtÞ M01 ðtÞ; ð1;2;...;j1Þ
M0j ðtÞ ¼ Q0j
ðtÞ þ F00 ðtÞ M0j ðtÞ
M00 ðtÞ ¼ F00 ðtÞ þ F00 ðtÞ M00 ðtÞ; and the transition probabilities are Fig. 5.2 State transition diagram for Type 1 Markov renewal process
ðj ¼ 2; 3; . . .; mÞ;
ð5:48Þ
142
5 ð1Þ
ð1;2;...;mÞ
P01 ðtÞ ¼ Q01 ðtÞ Q02 ðtÞ þ Q00 ¼ P0j ðtÞ ¼
Semi-Markov and Markov Renewal Processes
ðtÞ P01 ðtÞ
ð1Þ Q01 ðtÞ Q02 ðtÞ þ F00 ðtÞ P01 ðtÞ; ð1;2;...;j1Þ ð1;2;...;jÞ Q0j ðtÞ Q0jþ1 ðtÞ þ F00 ðtÞ
P0j ðtÞ
ðj ¼ 2; 3; . . .; mÞ;
P00 ðtÞ ¼ 1 Q01 ðtÞ þ F00 ðtÞ P00 ðtÞ; ð1;2;...;mÞ
ð5:49Þ
ð1;2;...;mÞ
where Q0mþ1 ðtÞ ¼ Q00 ðtÞ ¼ F00 ðtÞ: Taking the LS transforms on both sides of (5.48) and (5.49), M01 ðsÞ ¼
Q01 ðsÞ ðsÞ ; 1 F00 ð1;2;...;j1Þ
Q0j ðsÞ 1 F00 ðsÞ F00 ðsÞ M00 ðsÞ ¼ ðsÞ ; 1 F00 ðsÞ ¼ M0j
ðj ¼ 2; 3; . . .; mÞ;
ð5:50Þ
and ð1Þ
P01 ðsÞ ¼
Q01 ðsÞ Q02 ðsÞ ; ðsÞ 1 F00 ð1;2;...;j1Þ
ð1;2;...;jÞ
ðsÞ Q0jþ1 ðsÞ 1 F00 1 Q01 ðsÞ P00 ðsÞ ¼ ðsÞ ; 1 F00 P0j ðsÞ ¼
Q0j
ðsÞ
ðj ¼ 2; 3; . . .; mÞ;
ð5:51Þ
P where note that m j¼0 P0j ðsÞ ¼ 1: From (5.50) and (5.51), the renewal functions and the transition probabilities are computed explicitly upon inversion, however, this might not be easy except in simple cases. Example 5.4 (Parallel system) Consider an n-unit parallel redundant system: When at least one of n units is operating, the system is operating. When all units are down simultaneously, the system fails and will begin to operate again immediately by replacing all failed units with new ones. Each unit operates independently and has an identical failure distribution FðtÞ: The states are denoted by the total number of failed units. When all units begin to operate at time 0, the mass functions are Q01 ðtÞ ¼ 1 ½FðtÞn ; n X n ð1;2;...;j1Þ Q0j ½FðtÞi ½FðtÞni ðtÞ ¼ i i¼j
ðj ¼ 2; 3; . . .; nÞ:
Because the system begins to operate immediately at system failure,
5.3 Markov Renewal Process with Nonregeneration Points
ð1;2;...;nÞ
Q00
ð1;2;...;n1Þ
ðtÞ ¼ Q0n
143
ðtÞ ¼ FðtÞn :
If the system operates again after a maintenance time Y with distribution GðtÞ PrfY tg at system failure, then States 0 and n are regeneration points and ð1;2;...;nÞ ðtÞ ¼ FðtÞn GðtÞ ¼ F00 ðtÞ: Thus, substituting the above Qn0 ðtÞ ¼ GðtÞ; Q00 equations for (5.50) and (5.51), we can obtain the renewal functions and the transition probabilities.
5.3.2 Type 2 Markov Renewal Process Consider a Markov renewal process with Sy ¼ f0g and S ¼ f1; 2; . . .; mg in Fig. 5.3. The process starts in State 0 and is only permitted to make a transition into one State j 2 S ; and then return to State 0. The LS transforms of the first-passage distributions, the renewal functions, and the transition probabilities are ðsÞ ¼ F00
m X
ðiÞ
Q00 ðsÞ;
i¼1 ðsÞ F0j
¼
,"
Q0j ðsÞ
1
m X i¼1;i6¼j
Fig. 5.3 State transition diagram for Type 2 Markov renewal process
# ðiÞ Q00 ðsÞ
ð5:52Þ ðj ¼ 1; 2; . . .; mÞ;
144
5
Semi-Markov and Markov Renewal Processes
F00 ðsÞ ðsÞ ; 1 F00 Q0j ðsÞ ðj ¼ 1; 2; . . .; mÞ; M0j ðsÞ ¼ ðsÞ 1 F00 P 1 m j¼1 Q0j ðsÞ P00 ðsÞ ¼ ; 1 F00 ðsÞ M00 ðsÞ ¼
ðjÞ
P0j ðsÞ
Q0j ðsÞ Q00 ðsÞ ¼ ðsÞ 1 F00
ð5:53Þ
ð5:54Þ
ðj ¼ 1; 2; . . .; mÞ:
The process corresponds to a special one of Type 1 when m ¼ 1: That is, it is the simplest state space with one nonregeneration point. The process takes two alternate States 0 and 1. When the epoch entering State 1 is also a regeneration point, the process becomes an alternating renewal process described in Sect. 3.4. Example 5.5 (Inventory model) We give the inventory model of spare units [6] as one example of Markov renewal processes with nonregeneration points: A unit begins to operate at time 0. If the operating unit fails before time T; the spare unit is ordered immediately and is placed into service as soon as it is delivered after a lead time L: If the operating unit does not fail up to time T; the spare one is ordered at time T and is delivered after the lead time L: In this case, the spare unit is put into service when the original unit has failed or is put into inventory when it does not fail. Define the following states: State 0: Unit is operating and spare one is not delivered. State 1: Unit is operating and spare one is ordered. State 2: Unit is operating and spare one is in inventory. State 3: Operating unit has failed during the order of spare unit. State 4: Spare unit is ordered because of the failure of operating unit before time T: The system with the above five states forms a Markov renewal process with regeneration point f0g and nonregeneration points f1; 2; 3; 4g (Fig. 5.4).
Fig. 5.4 State transition diagram of spare part inventory
5.3 Markov Renewal Process with Nonregeneration Points
145
Assume that the failure times of each operating unit are identical and have a general distribution FðtÞ with mean time l: In addition, assume that PrfT tg AðtÞ and PrfL tg ¼ BðtÞ are the degenerate distributions placing unit masses at T and L; respectively. Then, the mass functions of the above process are Q01 ðtÞ ¼
Zt FðuÞdAðuÞ; 0
ð1Þ
Q02 ðtÞ ¼
Zt
FðuÞd½AðuÞ BðuÞ;
0 ð1;2Þ
Q00 ðtÞ ¼
Zt
½AðuÞ BðuÞdFðuÞ;
0 ð1Þ
Q03 ðtÞ ¼
Zt
½AðuÞ BðuÞdFðuÞ;
0
ð1;3Þ
Q00 ðtÞ ¼
8 Z t
Q04 ðtÞ ¼
:
0
9 =
½FðuÞ FðvÞdAðvÞ dBðu vÞ; ;
Zt AðuÞdFðuÞ;
Q40 ðtÞ ¼ BðtÞ:
0
Thus, the LS transforms of the above mass functions are Q01 ðsÞ ¼ esT FðTÞ; ð1Þ
Q02 ðsÞ ¼ esðTþLÞ FðT þ LÞ; Z1 ð1;2Þ est dFðtÞ; Q00 ðsÞ ¼ TþL ð1Þ Q03 ðsÞ
¼
Z1
st
e
dFðtÞ
T ð1;3Þ Q00 ðsÞ
est dFðtÞ;
TþL sðTþLÞ
¼e
Q04 ðsÞ ¼
Z1
ZT
½FðT þ LÞ FðTÞ;
est dFðtÞ;
0
Q40 ðsÞ ¼ esL :
146
5
Semi-Markov and Markov Renewal Processes
The LS transform of the recurrence time distribution to State 0 is ð1;2Þ
H00 ðsÞ ¼ Q00
¼
Z1 e
ð1;3Þ
ðsÞ þ Q00
ðsÞ þ Q04 ðsÞQ40 ðsÞ
st
sðTþLÞ
dFðtÞ þ e
sL
½FðT þ LÞ FðTÞ þ e
ZT
TþL
est dFðtÞ;
0
where note that H00 ð0Þ ¼ 1: Thus, the mean recurrence time to State 0 is
l00
Z1
ZTþL
tdH00 ðtÞ ¼ l þ
FðtÞdt: T
0
Using the above results, the LS transforms of the transition probabilities are P00 ðsÞ ¼
1 Q01 ðsÞ Q04 ðsÞ ; ðsÞ 1 H00
P01 ðsÞ ¼
Q01 ðsÞ Q02 ðsÞ Q03 ðsÞ ; ðsÞ 1 H00
P02 ðsÞ ¼
Q02 ðsÞ Q00 ðsÞ ; ðsÞ 1 H00
ð1Þ
ð1Þ
ð1Þ
ð1;2Þ
ð1Þ
ð1;3Þ
Q03 ðsÞ Q00 ðsÞ ; ðsÞ 1 H00 Q ðsÞ½1 Q40 ðsÞ ; P04 ðsÞ ¼ 04 ðsÞ 1 H00 P03 ðsÞ ¼
where note that P00 ðsÞ þ P01 ðsÞ þ P02 ðsÞ þ P03 ðsÞ þ P04 ðsÞ ¼ 1: Thus, the limiting probabilities Pj limt!1 P0j ðtÞ ¼ lims!0 P0j ðsÞ are 1 P0 ¼ l00
ZT FðtÞdt;
1 l00
Z1 FðtÞdt; TþL
ZTþL FðtÞdt; T
0
P2 ¼
1 P1 ¼ l00
2 TþL 3 Z 1 P3 ¼ 4 ½FðTÞ FðtÞdt5; l00 T
LFðTÞ ; P4 ¼ l00 where also note that P0 þ P1 þ P2 þ P3 þ P4 ¼ 1 (Problem 5.5).
5.3 Markov Renewal Process with Nonregeneration Points
147
Furthermore, the LS transforms of the expected number of visits to State j are ðsÞ ¼ M00
H00 ðsÞ ; ðsÞ 1 H00
ðsÞ ¼ M01
Q01 ðsÞ ; ðsÞ 1 H00
ðsÞ ¼ M03
Q03 ðsÞ ; 1 H00 ðsÞ
ð1Þ
Q02 ðsÞ ; 1 H00 ðsÞ Q04 ðsÞ M04 ðsÞ ¼ ðsÞ ; 1 H00 ðsÞ ¼ M02
ð1Þ
and the limiting expected number of visits to State j per unit of time, i.e., Mj ðsÞ are limt!1 M0j ðtÞ=t ¼ lims!0 sM00 1 FðTÞ ; M1 ¼ ; l00 l00 FðT þ LÞ FðTÞ M3 ¼ ; l00
M0 ¼
FðT þ LÞ ; l00 FðTÞ M4 ¼ ; l00 M2 ¼
where note that M0 ¼ M2 þ M3 þ M4 and M1 = M2 + M3 (Problem 5.6).
5.4 Problems 5 5.1 Derive Fj ðsÞ; Pj ðsÞ and Mj ðsÞðj ¼ 0; 1; 2Þ by taking the LS transforms of (5.1)–(5.3). 5.2 It is assumed in Example 5.3 that when the system is in State i (i ¼ 0; 1; 2), the data transmission succeeds with probability qi and fails with probability pi 1 qi ; where 0 p0 p1 p2 1 [3, p. 116]. Then, obtain the mean time l3 ðNÞ in which the data transmission succeeds and derive analytically an optimum number N that minimizes l3 ðNÞ: 5.3 Prove that fi;ij ¼ 1 fj;ii . 5.4 Prove in (5.45) that CðnÞ decrease to c0 =l for c0 c1 : 5.5 Derive Pj and Mj (j ¼ 0; 1; 2; 3; 4) in Example 5.4. 5.6 It is assumed in Example 5.5 that c1 is the expedited order cost of a spare unit when the original unit fails before time T; c2 (\c1 ) is the ordinary order cost at time T; c4 is the inventory cost per unit of time in State 2, and c3 ([c4 ) is the system failure cost per unit of time in States 3 and 4. Then, when FðtÞ ¼ 1 ekt ; derive an optimum time T that minimizes the expected cost rate CðTÞ ¼ c1 M4 þ c2 M1 þ c3 ðP3 þ P4 Þ þ c4 P2 :
148
5
Semi-Markov and Markov Renewal Processes
References 1. 2. 3. 4. 5.
Osaki S (1992) Applied stochastic system modeling. Springer, Berlin Çinlar E (1975) Introduction to stochastic processes. Prentice-Hall, Englewood Cliffs Nakagawa T (2008) Advanced reliability models and maintenance policies. Springer, London Takács L (1962) Introduction to the theory of queues. Oxford University Press, New York Srinivasan VS (1968) First emptiness in the spare parts problem for repairable components. Oper Res 16:407–415 6. Nakagawa T, Osaki S (1976) Markov renewal process with some nonregeneration points and their applications to reliability theory. Microelectron Reliab 15:633–636 7. Nakagawa T (2002) Two-unit redundant models. In: Osaki S (ed) Stochastic models in reliability and maintenance. Springer, Berlin, pp 165–185 8. Yasui K, Nakagawa T, Sandoh H (2002) Reliability models in data communication systems. In: Osaki S (ed) Stochastic models in reliability and maintenance. Springer, New York, pp 281–306
Chapter 6
Cumulative Processes
Failures of units or systems are generally classified into two failure modes: Catastrophic failure in which units fail by some sudden shock, and degradation failure in which units fail by physical deterioration suffered from some damage. In the latter case, units fail when the total damage due to shocks has exceeded a critical failure level. This is called a cumulative damage model or shock model with additive damage and can be described theoretically by a cumulative process [1] in stochastic processes. Damage models can be applied to actual units that are working in architecture, industry, service, information, and computers, and were summarized [2]. Let fNðtÞg be a counting process that represents the total numbers of events up to time t defined in Chaps. 2 and 3. Next, let Wk be some amount of increments due to the kth event. Then, define ZðtÞ
NðtÞ X
Wk
ðNðtÞ ¼ 0; 1; 2; . . .Þ;
ð6:1Þ
k¼0
that is the total amount of increments at time t: The stochastic process fZðtÞg is a cumulative process [1] or is called a jump process [3] or doubly stochastic process [4] (Problem 6.1). In Sects. 6.1 and 6.2, we investigate the properties of the process fZðtÞg when increments due to each event are additive or not, respectively. We apply such results to damage models and obtain their reliabilities, mean times to failure, and failure rates. Furthermore, Sect. 6.3 proposes four modified damage models. In Sect. 6.4, we summarize compactly replacement policies for damage models where a unit is replaced at a planned time, a shock number or a damage level. The continuous damage model where the total damage increases with time will be dealt in Chap. 7 by applying a Brownian motion and a Lévy process. In this way, cumulative processes can be applied to not only reliability models but also stochastic models such as queueing, insurance, finance, shot noise, and stochastic duel [2, p. 170].
T. Nakagawa, Stochastic Processes, Springer Series in Reliability Engineering, DOI: 10.1007/978-0-85729-274-2_6, Ó Springer-Verlag London Limited 2011
149
150
6 Cumulative Processes
6.1 Standard Cumulative Process Consider a cumulative process in the context of Cox [1]: Let random variables Xk ðk ¼ 1; 2; . . .Þ denote a sequence of interarrival times between successive events that form Poisson or renewal processes introduced in Chaps. 2 and 3. Next, suppose that an increment Wk ðk ¼ 1; 2; . . .Þ such as damage, wear, fatigue, cost, and risk is associated with the kth event, where W0 0: Two random variables Xk and Wk might be possibly dependent. It is assumed that the sequence of fWk g is nonnegative, independently, and identically distributed, and is independent of Xj ðk 6¼ jÞ: Then, consider the cumulative process fZðtÞg defined in (6.1). Example 6.1 (Replacement cost) Suppose that each failed unit is replaced immediately at failures with a new one, i.e., Xk is the failure time between replacements. If Wk is the replacement cost for the kth failure, then ZðtÞ is the total replacement cost up to time t: Furthermore, if Xk is less than a planned replacement time T; Wk is the replacement cost c1 after failure, and conversely, if Xk is greater than T; Wk is the replacement cost c2 before failure, then ZðtÞ is the total b replacement cost for an age replacement policy in Sect. 3.3 and ZðtÞ ¼ CðtÞ in (3.42). On the other hand, suppose that an operating unit is replaced only at periodic times kT and each failed unit undergoes some repair, i.e., Xk is always equal to T: If Wk is the sum of repair and replacement costs in the kth replacement interval, ZðtÞ is the total cost up to time t for periodic replacement policies in Sect. 3.3.2. Example 6.2 (Cumulative damage model) A unit is subjected to shocks and suffers some damage, and each damage is additive. It is assumed that the unit fails when the total damage has exceeded a failure level K ð0\K\1Þ for the first time (Fig. 6.1).
Fig. 6.1 Process for a standard cumulative damage model
6.1 Standard Cumulative Process
151
We are interested in the following quantities of the cumulative process fZðtÞg : (i) PrfZðtÞ xg; the distribution of the total increment at time t: (ii) EfZðtÞg; the total expected increment at time t: In addition, let K be the threshold level of the total increment. Defining Y mint fZðtÞ [ Kg; (iii) PrfY tg; the first-passage distribution to a critical point. (iv) EfYg; the mean time to a critical point. Suppose that the sequences of fXk g and fWk g are non-negative, independently, and identically distributed, and both of Xk and Wk are independent of each other. It is assumed that FðtÞ PrfXk tg with finite mean EfXk g l and GðxÞ PrfWk xg with finite mean EfWk g x for any t; x 0: Then, from (3.3), the number NðtÞ of events in ½0; t has the following probability PrfNðtÞ ¼ ng ¼ F ðnÞ ðtÞ F ðnþ1Þ ðtÞ
ðn ¼ 0; 1; 2; . . .Þ;
where uðnÞ ðtÞ denotes the n-fold Stieltjes convolution of any function uðtÞ with itself, and uð0Þ ðtÞ 1 for t 0: Thus, ( NðtÞ ) ( ) n X X Wk x; NðtÞ ¼ n ¼ Pr Wk xjNðtÞ ¼ n PrfNðtÞ ¼ ng Pr k¼0
k¼0 ðnÞ
¼ G ðxÞ½F ðnÞ ðtÞ F ðnþ1Þ ðtÞ: Therefore, the distribution of ZðtÞ defined in (6.1) is ( NðtÞ ) X Wk x PrfZðtÞ xg ¼ Pr k¼0
¼ ¼
1 X n¼0 1 X
(
Pr
n X
) Wk xjNðtÞ ¼ n PrfNðtÞ ¼ ng
k¼0
GðnÞ ðxÞ½F ðnÞ ðtÞ F ðnþ1Þ ðtÞ;
ð6:2Þ
n¼0
and the survival probability is PrfZðtÞ [ xg ¼
1 X ½GðnÞ ðxÞ Gðnþ1Þ ðxÞF ðnþ1Þ ðtÞ:
ð6:3Þ
n¼0
The total expected increment at time t is EfZðtÞg ¼
Z1
xd PrfZðtÞ xg
0
¼x
1 X n¼1
F ðnÞ ðtÞ ¼ xMF ðtÞ;
ð6:4Þ
152
6 Cumulative Processes
P ðnÞ where MF ðtÞ EfNðtÞg ¼ 1 n¼1 F ðtÞ is a renewal function with distribution FðtÞ and represents the expected number of events occurred in ½0; t; and its properties have been investigated in Sect. 3.2. It can be intuitively known that EfZðtÞg is given by the product of the average amount of increment suffered from each event and the expected number of events in time t: This is useful for estimating the total expected increment at time t: R1 Furthermore, from Theorem 3.1, if li 0 ti dFðtÞ\1ði ¼ 2; 3Þ and r2F l2 l2 ; then as t ! 1; 2 t rF 1 þ oð1Þ; MF ðtÞ ¼ þ 2l2 2 l 4 r2 t 5rF 2r2F 3 2l3 þ þ VfNðtÞg ¼ F3 þ þ oð1Þ; 4l4 l2 l 4 3l3 where oðhÞ=h ! 0 as h ! 0: Thus, when FðGÞ has finite mean lðwÞ and variance r2F ðr2G Þ; approximately, for large t (Problem 6.2), ( ( NðtÞ )) X Wk jNðtÞ ¼ EfNðtÞgEfWk g EfZðtÞg ¼ E E
k¼1
t r2F l2 x þ ; 2l2 l VfZðtÞg ¼ EfZ 2 ðtÞg ðEfZðtÞgÞ2 ( ( NðtÞ )) NðtÞ X X ¼E E Wk Wj jNðtÞ ðEfZðtÞgÞ2 k¼1
ð6:5Þ
j¼1
¼ VfNðtÞgðEfWk gÞ2 þEfNðtÞgVfWk g 2 t rF r2G 5r4F 2r2F 3 2l3 r2G r2F þ þ þ þ þ 1 : x2 4l4 l2 2 l2 l l2 x2 4 3l3 Moreover, because EfZðtÞg x ¼ ; lim t!1 t l
VfZðtÞg x2 r2F r2G ¼ ; þ t!1 l l2 x 2 t lim
ð6:6Þ
by applying Theorem 3.5 to the cumulative process fZðtÞg; we have the following asymptotic distribution of ZðtÞ given by a normal distribution: ( ) Zx ZðtÞ xt=l 1 2 2 x ¼ pffiffiffiffiffiffi lim Pr eu =2 du: ð6:7Þ 2 2 2 2 t!1 ðx t=lÞ rF =l þ rG =x 2p 1
Example 6.3 (Cumulative damage model) In the cumulative damage model given in Example 6.2, we wish to estimate the total damage when the probability that is
6.1 Standard Cumulative Process
153
more than z in t ¼ 30 days of operation is given by 0.90. The distributions FðtÞ and GðxÞ of shock times and the amount of damage are unknown, but from sample data, the following estimations of means and variances are made: l ¼ 2 days;
r2F ¼ 5 ðdaysÞ2 ;
x ¼ 1;
r2G ¼ 0:5:
In this case, from (6.5), EfZð30Þg 15:125: Then, from (6.7), when t ¼ 30; ZðtÞ xt=l Zð30Þ 15 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 2 5:12 ðx2 t=lÞ rF =l2 þ r2G =x2 is approximately normally distributed with mean 0 and variance 1. Hence,
Zð30Þ 15 z 15 PrfZð30Þ [ zg ¼ Pr [ 5:12 5:12 Z1 1 2 eu =2 du ¼ 0:90: pffiffiffiffiffiffi 2p ðz15Þ=5:12
pffiffiffiffiffiffi R 1 2 Because u0 ¼ 1:28 satisfies ð1= 2pÞ u0 eu =2 du ¼ 0:90; z ¼ 15 5:12 1:28 8:45: Thus, the total damage is more than 8.45 in 30 days with probability 0.90. Next, when a failure level is known as K ¼ 10;
Zð30Þ 15 10 15 [ PrfZðtÞ [ 10g ¼ Pr 5:12 5:12 Z1 1 2 pffiffiffiffiffiffi eu =2 du 0:836: 2p 0:98
Thus, the probability that the unit with a failure level K ¼ 10 fails in 30 days is about 0:836 (Problem 6.3). Next, suppose that the process ends when the total increment exceeds a threshold level K that is called a critical point. In this case, the probability that the total increment exceeds a threshold level K at the ðn þ 1Þth event is pnþ1 ¼ GðnÞ ðKÞ Gðnþ1Þ ðKÞ ðn ¼ 0; 1; 2; . . .Þ: Define Y mint fZðtÞ [ Kg: Because the event of fY tg and fZðtÞ [ Kg are equivalent, the first-passage distribution to a critical point is, from (6.3), UðtÞ PrfY tg ¼ PrfZðtÞ [ Kg 1 X ½GðnÞ ðKÞ Gðnþ1Þ ðKÞF ðnþ1Þ ðtÞ; ¼ n¼0
ð6:8Þ
154
6 Cumulative Processes
and its Laplace-Stieltjes (LS) transform from Appendix A is
U ðsÞ
Z1
est dUðtÞ ¼
1 X
½GðnÞ ðKÞ Gðnþ1Þ ðKÞ½F ðsÞnþ1 ;
ð6:9Þ
n¼0
0
where u ðsÞ denotes the LS transform of any function uðtÞ; i.e., u ðsÞ R 1 st duðtÞ: Thus, the mean time to a critical point is 0 e Z1 dU ðsÞ EfYg ¼ tdPrfY tg ¼ ds 0 ð6:10Þ s¼0 1 X ¼l GðnÞ ðKÞ ¼ l½1 þ MG ðKÞ; n¼0
P1
where MG ðKÞ n¼1 GðnÞ ðKÞ represents the expected number of events before a critical point. Similarly, when GðxÞ has finite mean x and variance r2G ; from (3.17), approximately, K r2G þ x2 þ EfYg l : ð6:11Þ 2x2 x In addition, when GðxÞ has an IFR property defined in Sect. 3.2.2 (Example 6.4), it has been shown that x x 1\MG ðxÞ ; x x from (a) of Sect. 3.2.2.3. Thus, lK K \EfYg l þ1 : x x In Example 6.3, EfYg is approximately 21.5 days and 20\EfYg 22: Next, suppose that events occur in a nonhomogeneous Poisson process with an Rt intensity function hðtÞ and a mean value function HðtÞ 0 hðuÞdu in Sect. 2.3. Then, from (2.43), ½HðtÞn HðtÞ ðn ¼ 0; 1; 2; . . .Þ: e n! P j HðtÞ Thus, by replacing F ðnÞ ðtÞ with 1 formally, we can obtain the j¼n f½HðtÞ =j!ge following quantities: PrfNðtÞ ¼ ng ¼
6.1 Standard Cumulative Process
PrfZðtÞ xg ¼
EfYg ¼
155
1 X
½HðtÞn HðtÞ GðnÞ ðxÞ ; e n! n¼0
1 X
ðnÞ
G ðKÞ
n¼0
Z1
EfZðtÞg ¼ xHðtÞ;
½HðtÞn HðtÞ e dt: n!
ð6:12Þ
0
If each event occurs at a constant time T ð0\T\1Þ; i.e., FðtÞ is the degenerate distribution placing unit mass at time T; and FðtÞ 0 for t\T and 1 for t T; then Z1 ð½t=TÞ ðKÞ; EfYg ¼ Gð½t=TÞ ðKÞdt; PrfY tg ¼ 1 G 0
where ½t=T denotes the greatest integer less than or equal to t=T: Finally, when GðxÞ 0 for x\1 and 1 for x 1; and K ¼ n; PrfY tg ¼ F ðnþ1Þ ðtÞ;
EfYg ¼ ðn þ 1Þl:
6.2 Independent Process Consider the independent process where the total increment is not additive, and the process ends when one amount of increment exceeds a threshold level K: In the damage model, when the damage due to some shock has exceeded for the first time a failure level K; then the unit fails (Fig. 6.2). The practical examples of such models are the fracture of brittle materials and semiconductor parts that have failed by some over-current or fault voltage (Problem 6.4). The generalized model with three types of shocks where shocks with a small level of damage cause no damage to the unit, shocks with a large level of damage result in failure, and shocks with an intermediate level result in failure only with some probability are considered [5]. In this case, the probability that the increment exceeds firstly a threshold level K at the ðn þ 1Þth event is Pnþ1 ¼ ½GðKÞn ½GðKÞnþ1 : Thus, the first-passage distribution to a critical point is PrfY tg ¼
1 X f½GðKÞn ½GðKÞnþ1 gF ðnþ1Þ ðtÞ; n¼0
Fig. 6.2 Process for a standard cumulative damage model
ð6:13Þ
156
6 Cumulative Processes
and its LS transform is Z1
est dPrfY tg ¼
½1 GðKÞF ðsÞ : 1 GðKÞF ðsÞ
ð6:14Þ
0
Hence, the mean time to a critical point is l : 1 GðKÞ
EfYg ¼
ð6:15Þ
If shocks occur in a nonhomogeneous Poisson process with a mean value function HðtÞ; then PrfY tg ¼
1 X
f1 ½GðKÞn g
n¼0
½HðtÞn HðtÞ e n!
¼ 1 e½1GðKÞHðtÞ ; Z1 EfYg ¼ e½1GðKÞHðtÞ dt:
ð6:16Þ
0
If events occur at a constant time T; PrfY tg ¼ 1 ½GðKÞ½t=T ; Z1 EfYg ¼ ½GðKÞ½t=T dt: 0
Example 6.4 (Failure rate) It is important to investigate the properties of failure rates practically and theoretically. Let Pn denote the probability of surviving the first n shocks ðn ¼ 0; 1; 2; . . .Þ; where P0 1; and Fn ðtÞ be the probability that n shocks occur in time t; where F0 ðtÞ 1: Then, the survival function that the unit does not fail in time t is written in the following general form: From (6.2), UðtÞ ¼ PrfY [ tg ¼
1 X n¼0
Pn PrfNðtÞ ¼ ng ¼
1 X
Pn ½Fn ðtÞ Fnþ1 ðtÞ:
n¼0
Let fn ðtÞ be a density function of Fn ðtÞ: The failure rate is rðtÞ
P1 0 U ðtÞ ðPn Pnþ1 Þfnþ1 ðtÞ : ¼ P1n¼0 P UðtÞ n¼0 n ½Fn ðtÞ Fnþ1 ðtÞ
ð6:17Þ
6.2 Independent Process
157
In particular, when shocks occur in a Poisson process with rate k; i.e., P j kt Fn ðtÞ ¼ 1 ; j¼n ½ðktÞ =j!e 1 X ðktÞn kt UðtÞ ¼ Pn ð6:18Þ e : n! n¼0 The probabilistic properties of UðtÞ were extensively investigated [6, 7]. The failure rate is, from (6.17), P1
Pnþ1 ½ðktÞn =n! k: rðtÞ ¼ k 1 Pn¼0 1 n n¼0 Pn ½ðktÞ =n! When Pn ¼ qn ; i.e., q ¼ GðKÞ and the total damage is not additive, from (6.18), UðtÞ ¼ ekð1qÞt ;
rðtÞ ¼ kð1 qÞ;
that is constant. Any distribution FðtÞ is said to have the property of IFR (Increasing Failure Rate) if and only if ½Fðt þ xÞ FðtÞ=FðtÞ increases with t for x [ 0 and FðtÞ\1 [8, p. 23], where FðtÞ 1 FðtÞ: Furthermore, it has been proved that FðtÞ is IFR 0 if and only if hðtÞ F ðtÞ=FðtÞ increases with t: The failure rate has been discussed in Sect. 2.3. In this model, the following properties (i) and (ii) were proved [9]: (i) The failure rate rðtÞ in (6.17) increases with t if ðPn Pnþ1 Þ=Pn increases with n: When Pnþ1 =Pn decreases with n and fnþ1 ðtÞ ¼ ½ðktÞn =n!ekt ; we prove that rðtÞ increases with t: Differentiating rðtÞ in (6.17) with respect to t; " # 1 1 X k2 ðktÞn X ðPnþ1 Pkþ1 Pnþ2 Pk Þ : ½UðtÞ2 n¼0 n! k¼0 The bracket is 1 1 X X ðktÞn ðktÞk Pkþ1 Pnþ1 Pnþ1 Pk n! k! Pk Pnþ1 n¼0 k¼0 " # n k 1 nþ1 1 X ðktÞ X X ðktÞ Pkþ1 Pnþ2 ðktÞk Pkþ1 Pnþ2 ¼ þ Pnþ1 Pk Pk n! k! k! Pk Pnþ1 Pk nþ1 n¼0 k¼0 k¼nþ1 1 nþ1 X X ðktÞn ðktÞk Pkþ1 Pnþ2 Pnþ1 Pk ¼ n! k! Pk Pnþ1 n¼0 k¼0 1 nþ1 X X ðktÞnþ1 ðktÞk1 Pnþ2 Pkþ1 Pnþ1 k Pk þ ðn þ 1Þ! k! Pnþ1 Pk n¼0 k¼0 n k 1 nþ1 X ðktÞ X ðktÞ Pkþ1 Pnþ2 ðn þ 1 kÞ 0; Pnþ1 Pk ¼ ðn þ 1Þ! k! Pk Pnþ1 n¼0 k¼0 that completes the proof.
158
6 Cumulative Processes
In addition, when the total damage is additive, from (6.8), UðtÞ ¼
1 X
GðnÞ ðKÞ
n¼0
ðktÞn kt e : n!
ð6:19Þ
Rt (ii) The failure rate average 0 rðuÞdu=t increases with t because ½GðnÞ ðxÞ1=n decRt reases with n: Note that if rðtÞ increases, then 0 rðuÞdu=t also increases. In P j hK particular, when Pn ¼ GðnÞ ðKÞ ¼ 1 ; Pnþ1 =Pn decreases j¼n ððhKÞ =j!Þe strictly with n from ð1 ehK Þ to 0. Thus, the failure rate rðtÞ increases strictly from kehK to k (Example 6.5). In general, for the cumulative damage model with a failure level K; the failure rate is, setting that Pn ¼ GðnÞ ðKÞ and Fn ðtÞ ¼ F ðnÞ ðtÞ in (6.17), P1 ½GðnÞ ðKÞ Gðnþ1Þ ðKÞf ðnþ1Þ ðtÞ : ð6:20Þ rðtÞ Pn¼0 1 ðnÞ ðnÞ ðnþ1Þ ðtÞ n¼0 G ðKÞ½F ðtÞ F When shocks occur in a nonhomogeneous Poisson process with an intensity function hðtÞ and a mean value function HðtÞ [7], from (6.12), UðtÞ ¼
1 X n¼0
Pn
½HðtÞn HðtÞ e : n!
ð6:21Þ
(iii) The failure rate rðtÞ increases if both hðtÞ and ðPn Pnþ1 Þ=Pn increase. Rt (iv) The failure rate average 0 rðuÞdu=t increases if both rðtÞ=t and ðPn Pnþ1 Þ=Pn increase. When the total damage is additive, (6.19) is UðtÞ ¼
1 X n¼0
GðnÞ ðKÞ
½HðtÞn HðtÞ : e n!
ð6:22Þ
Then, properties (iii) and (iv) are rewritten as (Problem 6.5): (v) The failure rate rðtÞ increases if both hðtÞ and ½GðnÞ ðKÞ Gðnþ1Þ ðKÞ=GðnÞ ðKÞ increase. Rt (vi) The failure rate average 0 rðuÞdu=t increases if both HðtÞ=t and ½GðnÞ ðKÞ Gðnþ1Þ ðKÞ=GðnÞ ðKÞ increase. Such results were compactly summarized [10]. Moreover, when shocks occur in the birth process [11], in the counting process [12], and in the Lévy process [13, 14], similar results were obtained. Example 6.5 (Exponential damage and shock times) Suppose that FðtÞ ¼ 1 ekt and GðxÞ ¼ 1 ehx ; i.e., shocks occur in a Poisson process with rate k and each
6.2 Independent Process
159
damage due to shocks is exponential with mean 1=h: In this case, both nonhomogeneous Poisson and renewal processes form the same Poisson process, i.e., 1 X ½HðtÞk
Fn ðtÞ ¼ F ðnÞ ðtÞ ¼
k!
k¼n
eHðtÞ ¼
1 X ðktÞk k¼n
k!
ekt
ðn ¼ 0; 1; 2; . . .Þ:
In the cumulative damage model, from (6.2), Z1
esx d PrfZðtÞ xg ¼ ekts=ðsþhÞ :
0
By inversion [8, p. 80], 2
3 Zx p ffiffiffiffiffiffi ffi p ffiffiffiffiffiffiffiffiffi PrfZðtÞ xg ¼ ekt 41 þ kht ehu u1=2 I1 ð2 khtuÞdu5; 0
where Ii ðxÞ is the Bessel function of order i for the imaginary argument defined by Ii ðxÞ
1 2jþi X x j¼0
2
1 : j!ðj þ iÞ!
Thus, from (6.8), 2
3 ZK p ffiffiffiffiffiffi ffi p ffiffiffiffiffiffiffiffiffi UðtÞ ¼ PrfY tg ¼ 1 ekt 41 þ kht ehu u1=2 I1 ð2 khtuÞdu5: 0
Furthermore, from (6.4), (6.6), and (6.10), EfZðtÞg ¼
kt ; h
VfZðtÞg ¼
2kt ; h2
EfYg ¼
hK þ 1 : k
The failure rate is, from (6.17), pffiffiffiffiffiffiffiffiffiffi kekthK I0 ð2 khtK Þ rðtÞ ¼ pffiffiffiffiffiffiffi R K pffiffiffiffiffiffiffiffiffi : 1 þ kht 0 ehu u1=2 I1 ð2 khtuÞdu Because GðnÞ ðKÞ Gðnþ1Þ ðKÞ ðhKÞn =n! ¼ P 1 j GðnÞ ðKÞ j¼n ðhKÞ =j! increases with n from ehK to 1 (Problem 6.6), rðtÞ increases from kehK to k: In the independent damage model, from (6.13), PrfY tg ¼ 1 exp ktehK ;
160
6 Cumulative Processes
and from (6.15), EfYg ¼
1 1 ¼ ehK : rðtÞ k
Example 6.6 (Replacement policy) Consider a replacement policy with two kinds of damage: Shocks occur in a renewal process according to FðtÞ with finite mean l, and the amount Wk of damage 1 due to the kth shock has an identical distribution GðxÞ PrfWk xg with finite mean x: Each damage 1 is additive, and the unit fails when the total damage exceeds a failure level K: Suppose that the unit is replaced at time T ð0\T 1Þ or at failure, whichever occurs first. Then, the probability that the unit is replaced before failure at time T is, from (6.8), UðTÞ ¼
1 X
GðnÞ ðKÞ½F ðnÞ ðTÞ F ðnþ1Þ ðTÞ;
n¼0
and the probability that it is replaced at failure is UðTÞ ¼
1 X ½GðnÞ ðKÞ Gðnþ1Þ ðKÞF ðnþ1Þ ðTÞ: n¼0
Thus, the mean time to replacement is
TUðTÞ þ
ZT 0
tdUðtÞ ¼
1 X n¼0
ðnÞ
G ðKÞ
ZT
½F ðnÞ ðtÞ F ðnþ1Þ ðtÞdt:
0
Introduce the replacement costs: cK is the replacement cost at failure and cT is the replacement cost at time T; where cK [ cT : Then, from (3.19), the expected cost rate is P ðnÞ ðnÞ ðnþ1Þ cK ðcK cT Þ 1 ðTÞ n¼0 G ðKÞ½F ðTÞ F e CðTÞ ¼ : RT P1 ðnÞ ðnÞ ðnþ1Þ ðtÞdt n¼0 G ðKÞ 0 ½F ðtÞ F
ð6:23Þ
Next, that another damage 2 occurs in a non homogeneous Poisson process with Rt an intensity function hðtÞ and a mean value function HðtÞ 0 hðuÞdu; i.e., the probability of n occurrences of damage 2 in ½0; t is given in (2.43). It is assumed that damage 2 occurs independently for damage 1 and its damage is not additive. When damage 2 occurs, the unit undergoes only minimal repair and its amount of damage 2 becomes zero. Thus, the expected number of occurrences of damage 2, i.e., minimal repair, before the replacement is, from (2.45),
6.2 Independent Process
HðTÞ
1 X
161
Z 1 X ðTÞþ ½GðnÞ ðKÞGðnþ1Þ ðKÞ HðtÞdF ðnþ1Þ ðtÞ T
ðnÞ
ðnÞ
G ðKÞ½F ðTÞF
ðnþ1Þ
n¼0
¼
1 X
n¼0
GðnÞ ðKÞ
n¼0
ZT
0
½F ðnÞ ðtÞF ðnþ1Þ ðtÞhðtÞdt:
0
Therefore, letting cM be the minimal repair cost and adding the total minimal e repair cost before replacement to CðTÞ in (6.23), P ðnÞ ðnÞ ðnþ1Þ cK ðcK cT Þ 1 ðTÞ n¼0 G ðKÞ½F ðTÞ F R T ðnÞ P1 ðnÞ ðnþ1Þ þcM n¼0 G ðKÞ 0 ½F ðtÞ F ðtÞhðtÞdt CðTÞ ¼ : RT P1 ðnÞ ðnÞ ðnþ1Þ ðtÞdt n¼0 G ðKÞ 0 ½F ðtÞ F
ð6:24Þ
Clearly, Cð0Þ lim CðTÞ ¼ 1; T!0
Cð1Þ lim CðTÞ ¼
cK þ cM
P1
n¼0
T!1
R1 GðnÞ ðKÞ 0 ½F ðnÞ ðtÞF ðnþ1Þ ðtÞhðtÞdt ; l½1 þ MG ðKÞ
whose denominator represents the mean time to failure in (6.10). Thus, there exists a positive T ð0\T 1Þ that minimizes CðTÞ in (6.24). We find an optimum time T that minimizes CðTÞ: Differentiating CðTÞ with respect to T and setting it equal to zero, ( ðcK cT Þ rðTÞ
1 X
ðnÞ
G ðKÞ
n¼0
ZT
½F ðnÞ ðtÞ F ðnþ1Þ ðtÞdt
0
1 X ½GðnÞ ðKÞ Gðnþ1Þ ðKÞF ðnþ1Þ ðTÞ
)
n¼0
þ cM
1 X n¼0
ðnÞ
G ðKÞ
ZT
½F ðnÞ ðtÞ F ðnþ1Þ ðtÞ½hðTÞ hðtÞdt ¼ cT ;
ð6:25Þ
0
where rðTÞ is given in (6.20). It can be clearly seen that if rðtÞ increases strictly and hðtÞ increases, or rðtÞ increases and hðtÞ increases strictly (Problem 6.7), then the left-hand side of (6.25) increases strictly from 0 to
162
6 Cumulative Processes
ðcK cT8Þflrð1Þ½1 þ MG ðKÞ 1g 9 Z1 1 = < X þ cM lhð1Þ½1 þ MG ðKÞ GðnÞ ðKÞ ½F ðnÞ ðtÞ F ðnþ1Þ ðtÞhðtÞdt : ; : n¼0 0
ð6:26Þ Thus, if (6.26) is greater than cT ; then there exists a finite and unique T ð0\T \1Þ that satisfies (6.25). When HðtÞ ¼ at; i.e., hðtÞ ¼ a ða [ 0Þ; from (6.26), if lrð1Þ½1 þ MG ðKÞ [ cK =ðcK cT Þ then there exists a finite T that satisfies (6.25). In P j hK ; addition, when FðtÞ ¼ 1 ekt and GðnÞ ðKÞ ¼ 1 j¼n ½ðhKÞ =j!e rðTÞ ¼
k
P1
n¼0 ½G
ðKÞ Gðnþ1Þ ðKÞ½ðkTÞn =n! n ðnÞ n¼0 G ðKÞ½ðkTÞ =n!
ðnÞ
P1
increases strictly from kehK to k; as shown in Example 6.5. Thus, if hK [ cT =ðcK cT Þ; then there exists a finite and unique T ð0\T \1Þ that satisfies (6.25).
6.3 Modified Damage Models We give the following four modified damage models [2, p. 28]: Damage model with imperfect shock where some shock may produce no damage, a failure level K is a random variable with a general distribution LðxÞ; the total damage decreases exponentially with time, and the total damage increases with time. A variety of shock models were widely summarized [14].
6.3.1 Imperfect Shock It has been assumed that shocks occur in a renewal process with FðtÞ and each damage due to shocks is distributed generally with GðxÞ: However, it may be considered the case where some shocks do not produce any damage. Suppose that each shock does some damage with probability p (0\p 1Þ and no damage with probability q 1 p: Then, the distribution of time where some shocks does the damage in time t is F1 ðtÞ ½1 þ qFðtÞ þ qFðtÞ qFðtÞ þ pFðtÞ; and its LS transform is F1 ðsÞ ¼
pF ðsÞ : 1 qF ðsÞ
ð6:27Þ
6.3 Modified Damage Models
163
Hence, the mean time to the first damage due to shocks is EfY1 g
Z1
l tdF1 ðtÞ ¼ : p
0
Thus, by replacing FðtÞ with F1 ðtÞ; we can get the results in the cases where shocks are imperfect. In particular, when FðtÞ ¼ 1 ekt ; F1 ðtÞ ¼ 1 epkt and EfY1 g ¼ 1=ðpkÞ; and PrfNðtÞ ¼ ng ¼
ðpktÞn pkt e n!
ðn ¼ 0; 1; 2; . . .Þ:
Similarly, when shocks occur in a non homogeneous Poisson process with a mean value function HðtÞ; ½pHðtÞn pHðtÞ e ðn ¼ 0; 1; 2; . . .Þ; n! EfNðtÞg ¼ VfNðtÞg ¼ pHðtÞ:
PrfNðtÞ ¼ ng ¼
Therefore, substituting F1 ðtÞ in (6.27) for FðtÞ in (6.2), (6.4), (6.8), (6.10) and (6.20), PrfZðtÞ xg; EfZðtÞg; PrfY tg; EfYg; and rðtÞ are given. In particular, from (6.9) and (6.10), Z1 e 0
st
1 X pF ðsÞ nþ1 ðnÞ ðnþ1Þ dPrfY tg ¼ ½G ðKÞ G ðKÞ ; 1 qF ðsÞ n¼0 l EfYg ¼ ½1 þ MG ðKÞ: p
The corresponding results for the independent damage model are, from (6.14) and (6.15), Z1
est dPrfY tg ¼
p½1 GðKÞF ðsÞ ; 1 ½q þ pGðKÞF ðsÞ
0
EfYg ¼
l : p½1 GðKÞ
Example 6.7 (Parallel system) Consider a parallel redundant system with n identical units, each of which fails at shocks with probability p; where q 1 p; and shocks occur in a renewal process with mean interval l (Problem 6.8). Let Wk be the total number of units that fail at the kth ðk ¼ 1; 2; . . .Þ shock. Then, because P the probability that one unit fails until the kth shock is kj¼1 pqj1 ¼ 1 qk ; the mean time to system failure is [2, p. 62]
164
6 Cumulative Processes 1 X
kl PrfW1 þ W2 þ þ Wk1 n 1 and W1 þ þ Wk ¼ ng
k¼1
¼
1 X
kl½ð1 qk Þn ð1 qk1 Þn
k¼1 1 X
n X n 1 ð1Þkþ1 ¼l ½1 ð1 q Þ ¼ l : 1 qk k k¼0 k¼1 k n
6.3.2 Random Failure Level Most units have individual variations in their ability to withstand shocks and are operating in different environments. In such cases, a failure level K is not constant and would be random. Consider the case where a failure level K is a random variable with a general distribution LðxÞ such that Lð0Þ ¼ 0: Then, for the cumulative damage model, the distribution of time to failure is, from (6.8), PrfY tg ¼
1 X
F
ðnþ1Þ
ðtÞ
Z1
n¼0
½GðnÞ ðxÞ Gðnþ1Þ ðxÞdLðxÞ;
ð6:28Þ
0
and its mean time is, from (6.10),
EfYg ¼ l
Z1
½1 þ MG ðxÞdLðxÞ:
ð6:29Þ
0
The failure rate is, from (6.20), P1 ðnþ1Þ R 1 ðnÞ f ðtÞ 0 ½G ðxÞ Gðnþ1Þ ðxÞdLðxÞ R : rðtÞ ¼ Pn¼0 1 ðnÞ ðnþ1Þ ðtÞ 1 GðnÞ ðxÞdLðxÞ n¼0 ½F ðtÞ F 0
ð6:30Þ
For the independent damage model,
PrfY tg ¼
1 X
F
ðnþ1Þ
n¼0
EfYg ¼ l
0
1 1 Z X n¼0
Z1 ðtÞ f½GðxÞn ½GðxÞnþ1 gdLðxÞ;
0
ð6:31Þ n
½GðxÞ dLðxÞ:
6.3 Modified Damage Models
165
For the cumulative model with imperfect shock, Z1 e
st
0
Z1 1 X pF ðsÞ nþ1 dPrfY tg ¼ ½GðnÞ ðxÞ Gðnþ1Þ ðxÞdLðxÞ: ðsÞ 1 qF n¼0
ð6:32Þ
0
Example 6.8 (Random failure level) Suppose that all random variables are exponential. When FðtÞ ¼ 1 ekt ; F1 ðsÞ ¼ pk=ðs þ pkÞ; i.e., F1 ðtÞ ¼ 1 epkt by inversion. Thus, substituting k for pk; we can obtain the corresponding results. When GðxÞ ¼ 1 ehx and a failure level K also has an exponential distribution LðxÞ ¼ 1 ebx ; Z1
½GðnÞ ðxÞ Gðnþ1Þ ðxÞdLðxÞ ¼
0
bhn ðb þ hÞnþ1
:
Thus, from (6.28), Z1
st
e 0
nþ1 1 X k bhn kb : dPrfY tg ¼ ¼ nþ1 sþk sðb þ hÞ þ kb ðb þ hÞ n¼0
By inversion,
kbt PrfY tg ¼ 1 exp ; bþh 1 1 h ¼ þ1 : EfYg ¼ rðtÞ k b For the independent damage model,
PrfY [ tg ¼
Z1
exp ktehx bebx dx
0
Z 1 X ðktÞn
1
¼
n¼0
EfYg ¼
n!
1 1 ¼ rðtÞ k
beðbþnhÞx dx ¼
0
Z1 0
ehx bebx dx ¼
1 X ðktÞn n¼0
(
n!
b ; b þ nh
b kðbhÞ
ðb [ hÞ;
1
ðb hÞ:
166
6 Cumulative Processes
6.3.3 Damage with Annealing The total damage in the usual reliability models is additive and does not decrease. In some materials, annealing, i.e., lessening the damage, can take place in materials such as rubber, fiber reinforced plastics, and polyurethane. Takac´s [15] considered the following damage model: If a unit suffers some damage W due to shock, then its damage after time duration t is reduced to Weat ð0\a\1Þ: Define ZðtÞ ¼
NðtÞ X
Wn exp½aðt Sn Þ;
ð6:33Þ
n¼1
P where Sn nk¼1 Xk (Fig. 6.3). Suppose that shocks occur in a Poisson process with rate k: Then, Uðt; xÞ PrfZðtÞ xg forms the following renewal equation [15, p. 105] 8 9 Zx < = dUðt; xÞ ¼ k Uðt; xÞ G½ðx yÞeat dy Uðt; yÞ ; : ; dt 0
and its LS transform is dU ðt; sÞ ¼ k½1 G ðseat ÞU ðt; sÞ; dt R1 R1 where U ðt; sÞ 0 esx dUðt; xÞ and G ðsÞ 0 esx dGðxÞ: Solving the differential equation, 8 9 < Zt = U ðt; sÞ ¼ exp k ½1 G ðseau Þdu ; : ; ð6:34Þ 0 dU ðt; sÞ kxð1 eat Þ EfZðtÞg ¼ ¼ : ds s¼0 a Fig. 6.3 Process for a cumulative damage model with annealing
6.3 Modified Damage Models
167
In addition, when x\1; limt!1 PrfZðtÞ xg exits and its LS transform is 9 8 < k Z1 1 G ðsuÞ = ð6:35Þ du : U ð1; sÞ ¼ exp ; : a u 0
We derive (6.34) by another method: When shocks occur in a Poisson process, from (2.29) and (2.37), unordered random variables of n shock times S1 ; S2 ; . . .; Sn ; given that NðtÞ ¼ n; are independent and identically distributed uniformly over the interval ½0; t; i.e., t Sk ðk ¼ 1; 2; . . .Þ are also distributed uniformly over ½0; t: Thus, when NðtÞ ¼ 1; ZðtÞ ¼ W1 exp½aðt S1 Þ and its LS transform is Z1 e
sx
1 dPrfZðtÞ xg ¼ EfG ½s expðaS1 Þg ¼ t
0
Zt
G ðseau Þdu:
0
Generally, when NðtÞ ¼ n; Z1
2 t 3n Z 1 esx dPrfZðtÞ xg ¼ ðEfG ½s expðaS1 ÞgÞn ¼ 4 G ðseau Þdu5 : t
0
0
So that, using the law of total probability, Z1
2
1 X ðktÞn
3n
Zt
1 ekt 4 G ðseau Þdu5 n! t n¼0 0 8 9 t Z < = ¼ exp k ½1 G ðseau Þdu : : ;
est dPrfZðtÞ xg ¼
0
0
Furthermore, (6.34) is also derived directly as follows [16, p. 40]:
EfZðtÞjNðtÞ ¼ ng ¼ E
( NðtÞ X
) Wk exp½aðt Sk ÞjNðtÞ ¼ n
k¼1
¼
n X E Wk exp½aðt Sk ÞNðtÞ ¼ n k¼1
n X ¼ E Wk NðtÞ ¼ n E exp½aðt Sk ÞNðtÞ ¼ n k¼1
¼ EfWk geat
n X E expðaSk ÞNðtÞ ¼ n : k¼1
168
6 Cumulative Processes
Recalling that Sk is distributed uniformly over ½0; t; ( ) n X expðaSk Þ E ZðtÞNðtÞ ¼ n ¼ xeat E k¼1
¼ xeat
n t
Zt
eax dx ¼
xn ð1 eat Þ: at
0
So that, because EfNðtÞg ¼ kt;
xNðtÞ kx at ð1 e Þ ¼ ð1 eat Þ: EfZðtÞg ¼ E at a Example 6.9 (1) When GðxÞ ¼ 1 ehx ,
U ðt; sÞ ¼
s þ heat sþh
m
ekt ;
where m k=a: Thus, by inversion [2, p. 32], 1 1 X X ðhxeat Þi mþj1 kt PrfZðtÞ xg ¼ e expðhxeat Þ: ð1 eat Þj j i! i¼j j¼0 In a similar way, m h U ð1; sÞ ¼ ; sþh
lim PrfZðtÞ xg ¼
Zx
t!1
hðhuÞm1 hu e du; CðmÞ
0
that is a gamma distribution with mean m=h: (2) When GðxÞ 0 for x\x and 1 for x x; i.e., the damage due to each shock is constant and its amount is x: From the results [15, p. 129], 0 1 m Z1 su 1 e duA; exp@m U ð1; sÞ ¼ u sxc x
where c eC ¼ 1:781072. . . and C 0:577215. . . that is Euler’s constant. By inversion, Rx P j j m ðjÞ xm þ 1 j¼1 ð1Þ ðmÞ =j! jx ðx uÞ I ðuÞdu lim PrfZðtÞ xg ¼ ; t!1 ðxcÞm Cð1 þ mÞ where IðuÞ is uniform over ½0; x:
6.3 Modified Damage Models
169
6.3.4 Increasing Damage with Time Consider the cumulative damage model with two kinds of damage (Fig. 6.4). One of them is caused by shock and is additive, and the other increases proportionately with time. That is, the total damage is accumulated subject to shocks and time at the rate of constant a ða [ 0Þ; independent of shocks. A unit fails whether the total damage exceeds with time or has exceeded a failure level K at some shocks, and its failure is detected only at the time of shocks. Such models would be the life of dry and storage batteries. The replacement policy where the unit is replaced before failure will be discussed of Chap. 7. P in Example 7.4 P Suppose that Sn nk¼1 Xk and Zn nk¼1 Wk ; where S0 ¼ W0 ¼ 0: Because PrfSn tg ¼ F ðnÞ ðtÞ and PrfZn xg ¼ GðnÞ ðxÞ; the distribution of time to detect a failure at some shock is [2, p. 35], PrfY tg ¼
1 X
PrfZn þ aSn \K Znþ1 þ aSnþ1 ; Snþ1 tg
n¼0
¼
8 tu t 1 Z
0
:
0
9 =
½GðnÞ ðK auÞ Gðnþ1Þ ðK aðu þ xÞÞdFðxÞ dF ðnÞ ðuÞ; ; ð6:36Þ
where note that GðnÞ ðxÞ ¼ 0 for x\0: Thus, the mean time to detect a failure at some shock is
Fig. 6.4 Process for a cumulative damage model with two kinds of damages
170
6 Cumulative Processes
EfYg ¼
8 1 1 1 Z
¼l
0
:
0
K=a 1 Z X n¼0
9 = h i ðt þ xÞ GðnÞ ðK atÞ Gðnþ1Þ ðK aðt þ xÞÞ dFðxÞ dF ðnÞ ðtÞ ;
GðnÞ ðK atÞdF ðnÞ ðtÞ:
ð6:37Þ
0
Finally, if the total damage increases exponentially, i.e., ZðtÞ ¼
NðtÞ X
Wn exp½aðt Sn Þ;
ð6:38Þ
n¼1
then by arguments similar to those of Sect. 6.3.3, when FðtÞ ¼ 1 ekt ; 9 8 = < Zt U ðt; sÞ ¼ exp k ½1 G ðseau Þdu ; ; : 0
kxðeat 1Þ ; EfZðtÞg ¼ 2a 1 3 Z k 1 G ðsuÞ 5 U ð1; sÞ ¼ exp4 du : a u
ð6:39Þ
1
6.4 Replacement Models Suppose that shocks occur in a renewal process with distribution FðtÞ and each causes a random amount of damage with distribution GðxÞ to a unit. These damages are additive, and the unit fails when the total damage has exceeded a failure level K: When the failure during actual operation may be costly or dangerous, it is of great importance to avoid such terrible situation. It would be wise to exchange the unit before failure at a lower cost. If we have no information on the condition of the unit, its maintenance should be done at a planned time. On the other hand, if we could get the number of shocks up to now and the amount of damage at shock times, its maintenance should be done at a pre-specified number of shocks or at a damage level. A variety of replacement models subjected to shocks were summarized [2]. This section summarizes the replacement policies for cumulative damage models using the theories of cumulative processes: In Sect. 6.4.1, the unit is replaced before failure at a planned time T; at a shock number N; or a damage level Z; whichever occurs first. Introducing the respective replacement costs for T; N and Z; we obtain the expected cost rate CðT; N; ZÞ: In Sect. 6.4.2, we derive analytically optimum T ; N and Z that minimize the expected cost rates for three policies. Furthermore, we discuss optimum N and Z that minimize the expected
6.4 Replacement Models
171
cost rate CðN; ZÞ and give a numerical example. Finally, we consider the periodic replacement with minimal repair at failures for a cumulative damage model and obtain the expected cost rate.
6.4.1 Three Replacement Policies As per the preventive replacement policies, the unit is replaced before failure at time T ð0\T 1Þ; at shock number N ðN ¼ 1; 2; . . .Þ; or at damage level Z ð0 Z KÞ; whichever occurs first. In addition, it is assumed that the unit is replaced at K or Z without considering to replace it at N when the total damage has exceeded K or Z at shock N: The probability that the unit is replaced at time T is PT ¼
N 1 h i X F ðnÞ ðTÞ F ðnþ1Þ ðTÞ GðnÞ ðZÞ; n¼0
the probability that it is replaced at shock N is PN ¼ F ðNÞ ðTÞGðNÞ ðZÞ; the probability that it is replaced at damage Z is PZ ¼
N 1 X
F
ðnþ1Þ
ðTÞ
n¼0
ZZ
½GðK xÞ GðZ xÞdGðnÞ ðxÞ;
0
and the probability that it is replaced at a failure level K is ZZ N 1 X ðnþ1Þ F ðTÞ GðK xÞdGðnÞ ðxÞ; PK ¼ n¼0
0
where note that PT þ PN þ PZ þ PK ¼ 1 (Problem 6.9). Thus, the mean time to replacement is ZT N1 h i X ðnÞ ðnþ1Þ ðnÞ ðNÞ F ðTÞ F ðTÞ G ðZÞ þ G ðZÞ t dF ðNÞ ðtÞ EfYg ¼ T n¼0
þ
þ
T N 1 Z X n¼0 0 T N1 XZ n¼0
¼
N 1 X n¼0
t dF ðnþ1Þ ðtÞ
ZZ
0
½GðK xÞ GðZ xÞdGðnÞ ðxÞ
0
t dF ðnþ1Þ ðtÞ
0
GðnÞ ðZÞ
ZT h 0
ZZ
GðK xÞdGðnÞ ðxÞ
0
i F ðnÞ ðtÞ F ðnþ1Þ ðtÞ dt:
172
6 Cumulative Processes
For the above replacement model, we introduce the following replacement costs: Costs cT ; cN ; cZ ; and cK are the respective replacement costs at time T; at shock N; at damage Z; and at level K; where cK is higher than the three costs cT ; cN and cZ : Then, the expected cost rate is, from (3.19) in Sect. 3.2, CðT; N; ZÞ ¼
cT PT þ cN PN þ cZ PZ þ cK PK EfYg i XN1 h ðnÞ ðnþ1Þ cK ðcK cT Þ F ðTÞ F ðTÞ GðnÞ ðZÞ n¼0
ðcK cN ÞF ðnÞ ðTÞGðNÞ ðZÞ RZ PN1 ðnþ1Þ ðcK cZ Þ n¼0 F ðTÞ 0 ½GðK xÞ GðZ xÞdGðnÞ ðxÞ ¼ : RT PN1 ðnÞ ðnÞ ðnþ1Þ ðtÞdt n¼0 G ðZÞ 0 ½F ðtÞ F ð6:40Þ When the unit is replaced only after failure, the expected cost rate is, from (6.10), CðT; N; ZÞ ¼ C lim T!1 N!1 Z!K
cK : l½1 þ MG ðKÞ
ð6:41Þ
The unit might be working as long as possible in the case where the replacement cost after failure is not so high. In this case, the unit should be replaced before failure at time T; shock number N; or damage level Z; whichever occurs last. Then, the probability that the unit is replaced at time T is PT ¼
1 h X
ih i F ðnÞ ðTÞ F ðnþ1Þ ðTÞ GðnÞ ðKÞ GðnÞ ðZÞ ;
n¼N
the probability that it is replaced at shock N is h i ðNÞ PN ¼ F ðTÞ GðNÞ ðKÞ GðNÞ ðZÞ ; the probability that it is replaced at damage Z is PZ ¼
1 X
F
ðnþ1Þ
ðTÞ
n¼N
ZZ
½GðK xÞ GðZ xÞdGðnÞ ðxÞ;
0
and the probability that it is replaced at a failure level K is PK ¼
1 h i X ðnÞ ðNÞ ðNÞ F ðnÞ ðTÞ F ðnþ1Þ ðTÞ G ðKÞ þ F ðTÞG ðKÞ n¼N
þ
1 X n¼N
F
ðnþ1Þ
ðTÞ
ZZ 0
GðK xÞdGðnÞ ðxÞ;
6.4 Replacement Models
173
where note that PT þ PN þ PZ þ PK ¼ 1 (Problem 6.9). Thus, the mean time to replacement is T
1 h ih i X F ðnÞ ðTÞ F ðnþ1Þ ðTÞ GðnÞ ðKÞ GðnÞ ðZÞ n¼N
þ
Z1
h i t dF ðNÞ ðtÞ GðNÞ ðKÞ GðNÞ ðZÞ
T 1 Z X
1
þ
t dF
ðnþ1Þ
ðtÞ
n¼N
þ
T T Z 1 X n¼0
þ
n¼0
þ
0 1
N 1 Z X
½GðK xÞ GðZ xÞdGðnÞ ðxÞ
0
h
t dF ðnþ1Þ ðtÞ GðnÞ ðKÞ Gðnþ1Þ ðKÞ
i
h i t dF ðnþ1Þ ðtÞ GðnÞ ðKÞ Gðnþ1Þ ðKÞ
T
1 1 Z X n¼N
ZZ
t dF ðnþ1Þ ðtÞ
T
¼ l½1 þ MG ðKÞ
ZZ
GðK xÞdGðnÞ ðxÞ
0 1 h X
ðnÞ
ðnÞ
G ðKÞ G ðZÞ
n¼N
i Z1h
i F ðnÞ ðtÞ F ðnþ1Þ ðtÞ dt:
T
Therefore, by the similar method of obtaining (6.40), the expected cost rate is P ðnÞ ðnþ1Þ ðTÞ GðnÞ ðKÞ GðnÞ ðZÞ cK ðcK cT Þ 1 n¼N F ðTÞ F h i ðNÞ ðcK cN ÞF ðTÞ GðNÞ ðKÞ GðNÞ ðZÞ RZ P ðnþ1Þ ðcK cZ Þ 1 ðTÞ 0 ½GðK xÞ GðZ xÞdGðnÞ ðxÞ n¼N F R 1 ðnÞ : CðT;N;ZÞ ¼ P ðnÞ ðnÞ ðnþ1Þ ðtÞ dt l½1þMG ðKÞ 1 n¼N G ðKÞ G ðZÞ T F ðtÞ F ð6:42Þ Clearly, when T ! 1; N ! 1, or Z ! K; the expected cost rate agrees with (6.41).
6.4.2 Optimum Policies We derive analytically the optimum planned time T ; shock number N ; and damage level Z that minimize the expected cost rates.
174
6 Cumulative Processes
6.4.2.1 Optimum T* Suppose that the unit is replaced only at time T or at failure, whichever occurs first. Then, the expected cost rate is, from (6.40), C1 ðTÞ N!1 lim CðT; N; ZÞ Z!K
ðnÞ P ðnÞ ðnþ1Þ cK ðcK cT Þ 1 ðTÞ n¼0 G ðKÞ F ðTÞ F ¼ : RT P1 ðnÞ ðnÞ ðnþ1Þ ðtÞdt n¼0 G ðKÞ 0 ½F ðtÞ F
ð6:43Þ
Note that limT!0 C1 ðTÞ ¼ 1 and limT!1 C1 ðTÞ is given in (6.41). We seek an optimum time T that minimizes C1 ðTÞ in (6.43) for cK [ cT : Then, differentiating C1 ðTÞ with respect to T and setting it equal to zero, rðTÞ
1 X
ðnÞ
G ðKÞ
n¼0
¼
ZT h
1 h i i X F ðnÞ ðtÞ F ðnþ1Þ ðtÞ dt GðnÞ ðKÞ Gðnþ1Þ ðKÞ F ðnþ1Þ ðTÞ n¼0
0
cT ; cK c T
ð6:44Þ
where rðTÞ is given in (6.20). It can be clearly seen that if rðTÞ increases strictly with T; then the left-hand side of (6.44) also increases strictly from 0 to lrð1Þ½1 þ MG ðKÞ 1 (Problem 6.10). Thus, if rð1Þ½1 þ MG ðKÞ [
cK ; lðcK cT Þ
then there exists a finite and unique T that satisfies (6.44), and the resulting cost rate is C1 ðT Þ ¼ ðcK cT ÞrðT Þ: Conversely, if rð1Þ½1 þ MG ðKÞ
cK ; lðcK cT Þ
then T ¼ 1; i.e., the unit is replaced only at failure, and the expected cost rate is given in (6.41). If a failure level K is distributed according to LðxÞ, as shown in Sect. 6.3.2, the expected cost rate becomes R1 P ðnÞ ðnþ1Þ cK ðcK cT Þ 1 ðTÞ 0 GðnÞ ðxÞdLðxÞ n¼0 F ðTÞ F : ð6:45Þ C1 ðTÞ ¼ R1 P1 R T ðnÞ ðnþ1Þ ðtÞdt ðnÞ ðxÞdLðxÞ ½ F ðtÞ F G n¼0 0 0
6.4 Replacement Models
175
In particular, suppose that shocks occur in a nonhomogeneous Poisson process and K is distributed exponentially, i.e., LðxÞ ¼ 1 ebx : Then, replacing F ðnÞ ðtÞ R1 F ðnþ1Þ ðtÞ with f½HðtÞn =n!geHðtÞ and setting that 0 GðnÞ ðxÞdLðxÞ ¼ ½G ðbÞn ;
C1 ðTÞ ¼
cK ðcK cT Þe½1G ðbÞHðTÞ ; RT ½1G ðbÞHðtÞ dt 0 e
ð6:46Þ
that corresponds to an age replacement policy by setting that FðtÞ ¼ e½1G ðbÞHðtÞ in (3.42). 6.4.2.2 Optimum N* Suppose that the unit is replaced only at shock N ðN ¼ 1; 2; . . .Þ or at failure, whichever occurs first. Then, the expected cost rate is, from (6.40), C2 ðNÞ lim CðT; N; ZÞ T!1 Z! K
¼
cK ðcK cN ÞGðNÞ ðKÞ PN1 ðnÞ l n¼0 G ðKÞ
ðN ¼ 1; 2; . . .Þ:
ð6:47Þ
Forming the inequality C2 ðN þ 1Þ C2 ðNÞ 0 to seek an optimum N that minimizes C2 ðNÞ for cK [ cN ; rNþ1
N1 X
GðnÞ ðKÞ ½1 GðNÞ ðKÞ
n¼0
cN cK cN
ðN ¼ 1; 2; . . .Þ;
ð6:48Þ
where rNþ1
GðNÞ ðKÞ GðNþ1Þ ðKÞ : GðNÞ ðKÞ
If rN increases strictly with N; i.e., GðNþ1Þ ðxÞ=GðNÞ ðxÞ decreases strictly with N; then the left-hand side of (6.48) also increases strictly with N to r1 ½1 þ MG ðKÞ 1; where r1 limN!1 rN 1 (Problem 6.10). Thus, if r1 ½1 þ MG ðKÞ [ cK =ðcK cN Þ; then there exists a finite and unique minimum N (1 N \1) that satisfies (6.48). In particular, when GðxÞ ¼ 1 ehx ; rNþ1 increases to 1: In this case, if hK [ cN =ðcK cN Þ then there exists a finite and unique minimum N ð1 N \1Þ that satisfies (6.48). 6.4.2.3 Optimum Z* Suppose that the unit is replaced only at damage Z ð0 Z KÞ or at failure, whichever occurs first. Then, the expected cost rate is, from (6.40),
176
6 Cumulative Processes
C3 ðZÞ lim CðT; N; ZÞ T!1
h i RZ cK ðcK cZ Þ GðKÞ 0 GðK xÞdMG ðxÞ ¼ : l½1 þ MG ðZÞ N!1
ð6:49Þ
We seek an optimum level Z that minimizes C3 ðZÞ for cK [ cZ : Differentiating C3 ðZÞ with respect to Z and setting it equal to zero, ZK cZ ½1 þ MG ðK xÞdGðxÞ ¼ ; ð6:50Þ cK cZ KZ
whose left-hand side increases strictly from 0 to MG ðKÞ: Thus, if MG ðKÞ [ cZ =ðcK cZ Þ; then there exists a finite and unique Z ð0\ Z \KÞ that satisfies (6.50), and its resulting cost rate is C3 ðZ Þ ¼
ðcK cZ ÞGðK Z Þ : l
In particular, when GðxÞ ¼ 1 ehx ; (6.50) is simplified as hZ ehðKZÞ ¼
cZ : c K cZ
6.4.2.4 Optimum N* and Z* We can consider optimum policies ðN ; Z Þ and ðT ; Z Þ that minimize the expected cost rates C4 ðN; ZÞ and C5 ðT; ZÞ (Problem 6.11) from CðT; N; ZÞ in (6.40), respectively. In particular, we derive optimum policies N and Z that minimize C4 ðN; ZÞ : The expected cost rate is, from (6.40), C4 ðN; ZÞ lim CðT; N; ZÞ T!1
cZ þ ðcN cZ ÞGðNÞ ðZÞ PN1 R Z ðnÞ þ ðcK cZ Þ n¼0 0 ½1 GðK xÞdG ðxÞ ¼ : PN1 l n¼0 GðnÞ ðZÞ Differentiating C4 ðN; ZÞ with respect to Z and setting it equal to zero, " # N1 gðNÞ ðZÞ X ðnÞ ðNÞ ðcN cZ Þ PN1 G ðZÞ G ðZÞ ðnÞ n¼1 g ðZÞ n¼0 Z N1 Z X ½GðK xÞ GðK ZÞdGðnÞ ðxÞ ¼ cZ ; þ ðcK cZ Þ n¼0
0
ð6:51Þ
ð6:52Þ
6.4 Replacement Models
177
where gðnÞ ðxÞ dGðnÞ ðxÞ=dx: From the inequality C4 ðN þ 1; ZÞ C4 ðN; ZÞ 0; " # N1 GðNÞ ðZÞ GðNþ1Þ ðZÞ X ðnÞ ðNÞ ðcN cZ Þ G ðZÞ þ G ðZÞ GðNÞ ðZÞ n¼0 2 PN1 ðnÞ ZZ G ðZÞ ½1 GðK xÞdGðNÞ ðxÞ þ ðcK cZ Þ4 n¼0ðNÞ G ðZÞ 0 3 Z Z N1 X ½1 GðK xÞdGðnÞ ðxÞ5 cZ : ð6:53Þ n¼0
0
Substituting (6.52) for (6.53), " # GðNÞ ðZÞ GðNþ1Þ ðZÞ gðNÞ ðZÞ þ PN1 ðcN cZ Þ ðnÞ GðNÞ ðZÞ n¼1 g ðZÞ 1 ðcK cZ Þ ðNÞ G ðZÞ
ZZ
½GðK xÞ GðK ZÞdGðNÞ ðxÞ 0:
ð6:54Þ
0
Thus, if cN cZ then there exists no finite N that satisfies (6.54) for any Z: It is assumed that cK [ cZ [ cN and GðxÞ ¼ 1 ehx : Then, (6.52) and (6.54) are simplified as, respectively, ( ) N1 X 1 1 ðhZÞN1 =ðN 1Þ! X ðhZÞj hZ X ðhZÞj hZ e e ðcZ cN Þ PN2 j j! j! j¼N n¼0 j¼n j¼0 ðhZÞ =j! þ ðcK cZ Þ
N1 X 1 X ðhZÞj hK e ¼ cZ ; j! n¼0 j¼nþ1
ð6:55Þ
and (
ðhZÞN =N! ðhZÞN1 =ðN 1Þ! ðcZ cN Þ P1 þ PN2 j j j¼N ðhZÞ =j! j¼0 ðhZÞ =j! P1 j hK j¼Nþ1 ðhZÞ =j!e ðcK cZ Þ P1 0: j hZ j¼N ðhZÞ =j!e
)
ð6:56Þ
P j hZ to 1, the left-hand side of Because ½ðhZÞn =n!= 1 j¼n ½ðhZÞ =j! increases from e (6.56) goes to cZ cN as N ! 1: Thus, there exists a finite N ð1 N \1Þ that satisfies (6.56). Table 6.1 presents the optimum shock number N and damage level Z ; and expected cost rate C4 ðN ; Z Þ for K ¼ 10; 20; 30; 40; 50 when h ¼ 0:5; l ¼ 1; cN ¼ 1; cZ ¼ 10 and cK ¼ 100: In case of N ¼ 1; C4 ð1; ZÞ decreases with Z and
178
6 Cumulative Processes
Table 6.1 Optimum N and Z ; and expected cost rate C4 ðN ; Z Þ when h ¼ 0:5; l ¼ 1; cN ¼ 1; cZ ¼ 10 and cK ¼ 100 K 10 20 30 40 50 N Z C4 ðN ; Z Þ
1 1 1.606
3 18.2 0.421
6 26.7 0.204
9 36.2 0.128
12 45.8 0.091
C4 ð1; 1Þ 1:606: Both optimum N and Z increase with K for K 20: Note that the ratio of Z to K are approximately 90%; i.e., Z =K 0:90; and N ð1=hÞ=K increase slowly from 0.20 to 0.48 with K for K 10: Furthermore, by the regression analysis, we have the regression line N ¼ 0:28K 2:2 for K 10 (Problem 6.12).
6.4.2.5 Periodic Replacement It is assumed in Sect. 6.4.1 that the total damage due to shocks is additive when it has not exceeded a failure level K; however, it is not additive at any shock after it has exceeded K (Fig. 6.5). In this case, the minimal maintenance is done at each shock and the damage level remains in K: Suppose that the unit is replaced at periodic times kT ðk ¼ 1; 2; . . .Þ for a planned time T ð0\T 1Þ (Problem 6.14). Then, the expected number of minimal maintenances, i.e., the expected number of shocks in the case where the total damage remains in K when it has reached K; is 1 h X
ðnÞ
G ðKÞ G
ðnþ1Þ
n¼0
¼
ZT h 1 iX i ðKÞ ðk þ 1Þ F ðkÞ ðT tÞ F ðkþ1Þ ðT tÞ dF ðnþ1Þ ðtÞ k¼0
1 X
0
½1 GðnÞ ðKÞF ðnÞ ðTÞ:
n¼1
Let cM be the cost of minimal maintenance. Then, by the similar method of obtaining (6.40), the expected cost rate is ( ) 1 X 1 ðnÞ ðnÞ F ðTÞ½1 G ðKÞ ; ð6:57Þ CðTÞ ¼ c T þ cM T n¼1 where cT is the replacement cost at time T: We seek an optimum time T that minimizes CðTÞ: Differentiating CðTÞ with respect to T and setting it equal to zero, 1 X n¼1
½1 GðnÞ ðKÞ½Tf ðnÞ ðTÞ F ðnÞ ðTÞ ¼
cT ; cM
6.4 Replacement Models
179
Fig. 6.5 Process for PM at time T
i.e., 1 h X
1 GðnÞ ðKÞ
i ZT
n¼1
td f ðnÞ ðtÞ ¼
cT ; cM
ð6:58Þ
0
where f ðnÞ ðtÞ dF ðnÞ ðtÞ=dt: Thus, if 1 X ½1 GðnÞ ðKÞT½f ðnÞ ðTÞ0 [ 0; n¼1
and 1 X
ðnÞ
½1 G ðKÞ
n¼1
Z1
td f ðnÞ ðtÞ [
cT ; cM
0
then there exists a finite and unique T (0\T \1) that satisfies (6.58). In particular, when FðtÞ ¼ 1 ekt ; (6.58) is 1 X ðkTÞn n¼1
n!
ekT
n X cT ½GðkÞ ðKÞ GðnÞ ðKÞ ¼ ; c M k¼1
ð6:59Þ
whose left-hand side increases strictly from 0 to MG ðKÞ: Therefore, if MG ðKÞ [ cT =cM then there exists a finite and unique T ð0\T \1Þ that satisfies (6.59).
6.5 Problems 6 6.1 Search for examples of cumulative processes on your daily life and investigate some examples [2]. 6.2 Derive (6.5).
180
6 Cumulative Processes
6.3 In Example 6.3, compute the total damage when a ¼ 0:95; 0.99 and the probability that the unit with K ¼ 15; 20 fails in 30 days. 6.4 Give some examples of independent damage models [2, p. 21]. 6.5 Prove (v) and (vi) in Example 6.4. P k 6.6 Prove in Example 6.5 that ½ðhKÞn =n!= 1 k¼n ½ðhKÞ =k! increases with n from ehK to 1. 6.7 Prove in (6.25) that if rðtÞ increases strictly and hðtÞ increases, or rðtÞ increases and hðtÞ increases strictly, the left-hand side of (6.25) increases strictly from 0 to (6.26). 6.8 In Example 6.7, consider the optimization problem in which the system is replaced when N ðN ¼ 1; 2; . . .; nÞ units have failed as the preventive replacement policy [2, p. 62]. 6.9 Prove that PT þ PN þ PZ þ PK ¼ 1 in Sect. 6.4.1 6.10 Prove in Sect. 6.4.2 that if rðTÞ and rN increase strictly with T and N; respectively, then (6.44) and (6.48) increase strictly. 6.11 Derive in Sect. 6.4.2.4 optimum T and Z that minimize C5 ðT; ZÞ limN!1 CðT; N; ZÞ when cT ¼ cZ and shocks occur in a Poisson process [2, p. 53]. 6.12 In Table 6.1, derive the regression line N ¼ 0:28K 2:2: 6.13 In Sect. 6.4.2.5, obtain the expected cost CðT; NÞ where the unit is replaced at time T or shock number N; whichever occurs first.
References 1. Cox DR (1962) Renewal theory. Methuen, London 2. Nakagawa T (2007) Shock and damage models in reliability theory. Springer, London 3. Abdel-Hameed M (1984) Life distribution properties of devices subject to a pure jump damage process. J Appl Probab 21:816–825 4. Grandell J (1976) Doubly stochastic poisson process. Lecture notes in mathematics 529. Springer, New York 5. Finkelstein MS, Zarudnij VI (2001) A shock process with a non-cumulative damage. Relib Eng Syst Saf 71:103–107 6. Gaver DP Jr (1963) Random hazard in reliability problems. Technometrics 5:211–226 7. Barlow RE, Proschan F (1975) Statistical theory of reliability and life testing. Halt, Rinehart & Winston, New York 8. Barlow RE, Proschan F (1965) Mathematical theory of reliability. Wiley, New York 9. Esary JD, Marshall AW, Proschan F (1973) Shock models and wear processes. Ann Probab 1:627–649 10. Shaked M (1984) Wear and damage processes from shock models in reliability theory. In: Abdel-Hameed M, Çinlar E, Quinn J (eds) Reliability theory and models. Academic, Orlando 11. Abdel-Hameed M, Proschan F (1975) Shock models with underlying birth process. J Appl Probab 12:18–28 12. Block HW, Savits TH (1978) Shock models with NBUE survival. J Appl Probab 15:621–628
References
181
13. Abdel-Hameed M (1984) Life distribution properties of devices subject to a Lévy wear process. Math Oper Res 9:606–614 14. Nikulim MS, Limnios N, Balakrishnan N, Kahle W, Huber-Carol C (2010) Advances in degradation modeling. Birhäuser, Boston 15. Takács L (1960) Stochastic processes. Wiley, New York 16. Ross SM (1983) Stochastic processes. Wiley, New York
Chapter 7
Brownian Motion and Lévy Processes
Wiener (1923) and Lévy (1939) mathematically gave the theoretical foundation and construction of the Brownian Motion based on the phenomenon of small particles observed by Brown (1827) and Einstein (1905). This is called the Brownian Motion or Wiener process, and is an example of a Markov process with continuous time and state space. This is now one of the most useful stochastic processes in applied sciences such as physics, economics, communication theory, biology, management science, and mathematical statistics, and recently, it is also used to make several valuable models in finance. Section 7.1 describes a brief outline of a Brownian Motion process and considers its application to a continuous damage level. The total damage might be estimated only at periodic times or with time roughly. Section 7.2 takes up three damage models and compares numerically their optimum replacement policies. Section 7.3 introduces a Lévy process and gives two damage models to show how to apply it to reliability models.
7.1 Brownian Motion and Wiener Processes We define a Brownian Motion process or a Wiener process [1, p. 184, 2, p. 340]: Definition 7.1 A stochastic process fXðtÞg is said to be a Brownian Motion process if (i) Xð0Þ ¼ 0; (ii) the process fXðtÞg has stationary and independent increment, (iii) the process fXðtÞg is normally distributed with mean 0 and variance r2 t for any t [ 0: For any t; u [ 0; the increments Xðt þ uÞ XðuÞ and XðtÞ XðuÞ for t [ u have a normal density with mean 0, and variance r2 t and r2 ðt uÞ; respectively, i.e., the density f ðx; tÞ of PrfXðt þ uÞ XðuÞ xg ¼ Prf XðtÞ xg is
T. Nakagawa, Stochastic Processes, Springer Series in Reliability Engineering, DOI: 10.1007/978-0-85729-274-2_7, Ó Springer-Verlag London Limited 2011
183
184
7 Brownian Motion and Lévy Processes
1 2 2 f ðx; tÞ ¼ pffiffiffiffiffiffiffi ex =ð2r tÞ ; 2ptr
ð7:1Þ
and its Laplace transform is (Problem 7.1) Z1
esx f ðx; tÞdx ¼ es
2 2
r t=2
ð7:2Þ
:
1
When r ¼ 1; BðtÞ XðtÞ=r is called a standard Brownian Motion because VfXðtÞ=rg ¼ t: Thus, any Brownian Motion can convert to the standard process. For any 0 t0 \t1 \ \tn \t; from the properties of independent and stationary increments, PrfXðtÞ xjXðt0 Þ ¼ x0 ; . . .; Xðtn Þ ¼ xn g ¼ PrfXðtÞ xjXðtn Þ ¼ xn g ¼ PrfXðtÞ Xðtn Þ x xn g; that states that the process has a Markov property, and its distribution function is, from (7.1), 1 PrfXðtÞ Xðtn Þ x xn g ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2pðt tn Þr
xxn Z
eu
2
=½2r2 ðttn Þ
du;
ð7:3Þ
1
i.e., XðtÞ Xðtn Þ has a normal density with mean 0 and variance r2 ðt tn Þ for any t tn ; that does not depend on tn : When ZðtÞ ¼ lt þ rBðtÞ; where BðtÞ is a standard Brownian Motion, ZðtÞ is called a Brownian Motion with drift parameter l and variance r2 [1, 2]. Example 7.1 (Continuous damage model) The total damage or wear ZðtÞ usually increases swayingly with time t from Zð0Þ ¼ 0; and the unit fails when ZðtÞ has exceeded a failure level K (Fig. 7.1). It is assumed that ZðtÞ ¼ AðtÞt þ rBðtÞ for AðtÞ 0 and BðtÞ is a standard Brownian Motion. Then, the reliability RðtÞ at time t; in other words, the probability that the total damage has not exceeded K in time t; is RðtÞ ¼ PrfZðtÞ Kg ¼ PrfAðtÞt þ rBðtÞ Kg; R1 and the MTTF is given by 0 RðtÞdt:
ð7:4Þ
(1) When AðtÞ l (constant), K ¼ k (constant), and BðtÞ is normally distributed with mean 0 and variance t; k lt pffi ; RðtÞ ¼ PrfrBðtÞ k ltg ¼ U ð7:5Þ r t where UðxÞ is the standard normal distribution with mean 0 and variance 1, pffiffiffiffiffiffi R x 2 i.e., UðxÞ ð1= 2pÞ 1 eu =2 du (Problem 7.2).
7.1 Brownian Motion and Wiener Processes
185
Fig. 7.1 Process for a continuous damage model
(2) When BðtÞ 0; K ¼ k and AðtÞ is normally distributed with mean l and variance r2 =t; k lt pffi ; ð7:6Þ RðtÞ ¼ PrfAðtÞ k=tg ¼ U r t that becomes equal to (7.5). (3) When AðtÞ l; BðtÞ 0 and K is normally distributed with mean k and variance r2 ; k lt RðtÞ ¼ Prflt Kg ¼ U : ð7:7Þ r When K is normally distributed with mean k and variance r2 t; RðtÞ is equal to (7.5) and (7.6). pffiffiffiffiffi Replacing a r= lk and b k=l in (7.5) or (7.6), " rffiffiffi rffiffiffi!# 1 b t ; RðtÞ ¼ U a t b which is called the Birnbaum–Saunders distribution [3]. This is widely applied to fatigue failure for material strength subject to stresses [4, 5]. When ZðtÞ ¼ AðtÞt þ rBðtÞ; if AðtÞ; BðtÞ; and K are deterministic, i.e., AðtÞ l; BðtÞ ¼ 1; and K ¼ k; then the unit fails at time ðk rÞ=l: By fitting appropriate probability distributions to AðtÞ; BðtÞ; and K, and estimating their parameters for actual units, the function ZðtÞ can be used as a continuous damage or wear function in cumulative damage models [6, p. 26]. Furthermore, the process defined by XðtÞ eZðtÞ for t 0 is called a geometric Brownian Motion, and its state space is the interval ½0; 1Þ [1, p. 193]. The process
186
7 Brownian Motion and Lévy Processes
is applied to models of share prices, interest rates, and stock options in finance. When ZðtÞ ¼ lt þ BðtÞ defined in (7.1), from (7.2), n o 1 E esZðtÞ ¼ exp st l þ sr2 ; 2 and hence, setting that s ¼ 1; n o 2 EfXðtÞg ¼ E eZðtÞ ¼ eðlþr =2Þt :
ð7:8Þ
Example 7.2 [7, p. 379]. Suppose that a unit begins to operate at time 0 and a failed unit undergoes repair, where the repair time is negligible. It is assumed that YðtÞ c1 ½XðtÞ 1 is the total maintenance cost due to failures over the time interval ½0; t and is given by n o h i 2 cðtÞ EfYðtÞg ¼ c1 E eZðtÞ 1 ¼ c1 eðlþr =2Þt 1 : Suppose that the unit is replaced at time T: The expected cost rate is, from Sect. 3.3.2.2,
CðTÞ
cðTÞ þ c2 ¼ T
h i 2 c1 eðlþr =2ÞT 1 þ c2 T
;
ð7:9Þ
where c2 is the replacement cost at time T: Clearly, lim CðTÞ ¼ lim C1 ðTÞ ¼ 1:
T!0
T!1
Thus, there exists an optimum T that minimizes (7.9). Differentiating CðTÞ with respect to T and setting it equal to zero,
h i c r2 2 2 2 lþ Teðlþr =2ÞT eðlþr =2ÞT 1 ¼ ; 2 c1
ð7:10Þ
whose left-hand side increases from 0 to 1: Therefore, there exists a unique optimum T (0\T \1) that satisfies (7.10), and the resulting cost rate is (Problems 7.3 and 7.4) r2 ðlþr2 =2ÞT : ð7:11Þ e CðT Þ ¼ c1 l þ 2
7.2 Three Replacements of Cumulative Damage Models
187
7.2 Three Replacements of Cumulative Damage Models We consider three cumulative damage models: A unit is subjected to shocks and suffers some damage due to it (see Sect. 7.2.1.1); the amount of damage due to shocks is measured only at periodic times (see Sect. 7.2.1.2), and the total damage increases linearly with time (see Sect. 7.2.1.3). The total damage is additive and the unit fails when the total damage has exceeded a failure level K (0\K\1) for three models. Models 2 and 3 would be practically used as the approximation ones of Model 1 and Models 1 and 2, respectively. As per the replacement policy, the unit is replaced at a planned time T: The expected cost rates of each model are obtained, and optimum policies that minimize them are derived analytically. Furthermore, optimum replacement times of each model are compared numerically when each model has the same mean amount of damage at any periodic times.
7.2.1 Three Damage Models We consider the following three damage models for an operating unit [8].
7.2.1.1 Standard Model The unit is subjected to shocks and suffers some damage at each shock. Let random variables Xk (k ¼ 1; 2; . . .) denote a sequence of interarrival times between successive shocks, and random variables Wk (k ¼ 1; 2; . . .) denote the damage produced by the kth shock, where W0 ¼ 0 (Fig. 6.1). In addition, let NðtÞ denote the random variable that is the total number of shocks up to time t: Then, define a random variable ZðtÞ ¼
NðtÞ X
Wk
ðNðtÞ ¼ 0; 1; 2; . . .Þ;
ð7:12Þ
k¼0
that represents the total damage at time t: It is assumed that FðtÞ PrfXk tg with finite mean l and variance r2F and GðxÞ PrfWk xg with finite mean x and variance r2G for t; x 0: Then, from (6.3), (6.5) and (6.8), PrfZðtÞ xg ¼
1 X
GðkÞ ðxÞ½F ðkÞ ðtÞ F ðkþ1Þ ðtÞ;
k¼0
EfZðtÞg ¼ xMF ðtÞ;
x2 t r2F r2G VfZðtÞg þ ; l l2 x 2
ð7:13Þ
188
7 Brownian Motion and Lévy Processes
where /ðkÞ ðtÞ (k ¼ 1; 2; . . .) denotes the k-fold Stieltjes convolution of any funcP ðkÞ tion /ðtÞ with itself and /ð0Þ ðtÞ 1 for t 0; and M/ ðtÞ 1 k¼1 / ðtÞ: Suppose that the unit fails when the total damage has exceeded a failure level K: As per the replacement policy, the unit is replaced at time T or at failure, whichever occurs first. Letting cK be the replacement cost at failure and cT (cT \cK ) be the replacement cost at time T; the expected cost rate is, from (6.23) and (6.43), P ðkÞ ðkÞ ðkþ1Þ cK ðcK cT Þ 1 ðTÞ k¼0 G ðKÞ½F ðTÞ F : ð7:14Þ C1 ðTÞ ¼ RT P1 ðkÞ ðkÞ ðkþ1Þ ðtÞdt k¼0 G ðKÞ 0 ½F ðtÞ F We find an optimum time T1 that minimizes C1 ðTÞ: Let f ðtÞ be a density function of FðtÞ and P1 ðkþ1Þ f ðTÞ½GðkÞ ðKÞ Gðkþ1Þ ðKÞ : Q1 ðTÞ Pk¼0 1 ðkÞ ðkÞ ðkþ1Þ ðTÞ k¼0 G ðKÞ½F ðTÞ F By the method similar to Example 6.6, if Q1 ðTÞ increases strictly with T and Q1 ð1Þ½1 þ MG ðKÞ [ cK =½lðcK cT Þ; then there exists a finite and unique T1 (0\T1 \1) that satisfies (Problem 7.5) Q1 ðTÞ
1 X k¼0
1 X
ðkÞ
G ðKÞ
ZT
½F ðkÞ ðtÞ F ðkþ1Þ ðtÞdt
0
F ðkþ1Þ ðTÞ½GðkÞ ðKÞ Gðkþ1Þ ðKÞ ¼
k¼0
cT : cK cT
ð7:15Þ
7.2.1.2 Periodic Model Each amount Wn (n ¼ 1; 2; . . .) of damage is measured only at periodic times nT0 (n ¼ 1; 2; . . .) for given T0 (0\T0 \1) and has an identical distribution GT ðxÞ PrfWn xg with mean xT and variance r2T : Then, ðnÞ
PrfZðnT0 Þ xg ¼ GT ðxÞ; EfZðnT0 Þg ¼ nxT ;
VfZðnT0 Þg ¼ nr2T :
ð7:16Þ
Suppose that the unit is replaced at time NT0 or at failure, whichever occurs first. Then, the expected cost rate is, from (6.47), ðNÞ
C2 ðNÞ ¼
cK ðcK cT ÞGT ðKÞ PN1 ðnÞ GT ðKÞ T0 n¼0
ðN ¼ 1; 2; . . .Þ:
ð7:17Þ
7.2 Three Replacements of Cumulative Damage Models
189
We find an optimum number N that minimizes C2 ðNÞ: From the inequality C2 ðN þ 1Þ C2 ðNÞ 0; Q2 ðN þ 1Þ
N1 X
ðnÞ
ðNÞ
GT ðKÞ ½1 GT ðKÞ
n¼0
cT ; c K cT
ð7:18Þ
where ðN1Þ
Q2 ðNÞ
GT
ðNÞ
ðKÞ GT ðKÞ ðN1Þ
GT
ðKÞ
:
Thus, if Q2 ðNÞ increases strictly with N and Q2 ð1Þ½1 þ MGT ðKÞ [ cK =ðcK cT Þ; then there exists a finite and unique minimum N (1 N \1) that satisfies (7.18) (Problem 7.6).
7.2.1.3 Continuous Model The total damage ZðtÞ usually increases with time t: Suppose that ZðtÞ ¼ AðtÞt where AðtÞ has a distribution LA ðtÞ with mean a and variance r2A =t: Then, PrfZðtÞ xg ¼ PrfAðtÞ x=tg ¼ LA ðx=tÞ; EfZðtÞg ¼ at;
VfZðtÞg ¼ r2A t:
ð7:19Þ
Suppose that the unit is replaced at time T or at failure, whichever occurs first. Then, the expected cost rate is C3 ðTÞ ¼
cK ðcK cT ÞLA ðK=TÞ : RT 0 LA ðK=tÞdt
ð7:20Þ
We find an optimum time T3 that minimizes C3 ðTÞ: Let lA ðtÞ be a density function of LA ðtÞ and rA ðtÞ be the failure rate of LA ðtÞ; i.e., rA ðtÞ lA ðtÞ=LA ðtÞ: Then, if rA ðtÞ increases strictly, then there exists a finite and unique T3 ð0\T3 \1Þ that satisfies (Problem 7.7) rA ðK=TÞ
ZT
LA ðK=tÞdt ½1 LA ðK=TÞ ¼
cT : cK c T
ð7:21Þ
0
7.2.2 Numerical Examples We compute optimum policies numerically for the three models: It is assumed that FðtÞ ¼ 1 ekt and GðxÞ ¼ 1 ex=x for Model 1, GT ðxÞ ¼ 1 ex=xT for Model 2, and AðtÞ has a normal distribution NðxA ; r2A =tÞ for Model 3.
190
7 Brownian Motion and Lévy Processes
In addition, we set the mean and variance of the total damage at any time nT0 as equal for all models, i.e., EfZðnT0 Þg ¼ kxnT0 ¼ nxT ¼ xA nT0 ;
VfZðnT0 Þg ¼ 2kx2 nT0 ¼ nx2T ¼ r2A nT0 :
Hence, k¼
2 ; T0
x¼
xT ; 2
xA ¼
xT ; T0
r2A ¼
x2T : T0
When T0 ¼ 1 and xT ¼ 1; k ¼ 2; x ¼ 1=2; xA ¼ 1 and r2A ¼ 1: Optimum policies are, for Model 1, from (7.15), P1
T j j j Z 1 X ð2KÞj 2K X j¼0 ½ð2TÞ ð2KÞ =j!j! e P1 Pj j k j! k¼0 j¼0 ½ð2KÞ =j! k¼0 ½ð2TÞ =k! j¼0 0
2
1 X ð2TÞj j¼1
j!
e2T
j1 X ð2KÞk
k!
k¼0
e2K ¼
ð2tÞk 2t e dt k!
1 ; cK =cT 1
ð7:22Þ
for Model 2, from (7.18), N 1 X 1 N1 k X K N =N! K j K X N K 1 P1 e e ; k =k!Þ ðK j! k! =c c K T 1 k¼N k¼0 j¼k k¼0
ð7:23Þ
and for Model 3, from (7.21), pffiffiffiffi pffiffiffiffi ZT pffiffiffiffi ðK=T þ 1Þ/ðK= T T Þ K pffi K pffiffiffiffi pffiffiffiffi pffiffiffiffi U pffi t dt 1 U pffiffiffiffi T t T 2 T UðK= T T Þ 0
1 ¼ ; cK =cT 1
ð7:24Þ
pffiffiffiffiffiffi R x pffiffiffiffiffiffi 2 2 where /ðxÞ ð1= 2pÞex =2 and UðxÞ ð1= 2pÞ 1 eu =2 du: Table 7.1 presents optimum T1 and C1 ðT1 Þ for cK =cT ¼ 5; 10; 20; 50 and K ¼ 5; 10; 15; 20: This indicates that T1 decreases when cK =cT and 1=K increase. It is of interest that C1 ðT1 Þ have their minimum values at K ¼ 15 for cK =cT 10: Table 7.2 presents optimum N T0 and C2 ðN T0 Þ for the same parameters cK =cT and K as Table 7.1. This shows the similar tendencies to Table 7.1, however, C2 ðN Þ decreases with K and cT =cK : Table 7.3 presents optimum T3 and C3 ðT3 Þ: This shows the same tendencies as Table 7.2 for both T3 and C3 ðT3 Þ: These tables indicate that T1 is equal to N T0 and N T0 [ T3 for large cK =cT and K; and T1 =T3 changes from 1:0 to 2:2 in this case. It is natural that T1 ; N T0 and T3 decrease when cK =cT and 1=K increase. The values of T3 are too small compared with N T0 : If T0 and 1=k would be small, T3 would approach N T0 :
7.2 Three Replacements of Cumulative Damage Models
191
Table 7.1 Optimum T1 and C1 ðT1 Þ when T0 ¼ 1 and lT0 ¼ 1 cK =cT
K 5
5 10 20 50
10
15
20
T1
CðT1 Þ
T1
CðT1 Þ
T1
CðT1 Þ
T1
CðT1 Þ
2.9 2.2 1.7 1.3
0.520 0.668 0.840 1.127
5.9 5.0 4.3 3.6
0.212 0.247 0.284 0.336
9.3 8.2 7.3 6.4
0.126 0.141 0.157 0.178
13.3 12.2 11.5 11.0
0.110 0.145 0.207 0.385
Table 7.2 Optimum N T0 and C2 ðN Þ when T0 ¼ 1 and lT0 ¼ 1 cK =cT
K 5
10
5 10 20 50
15
20
N T0
C2 ðN Þ
N T0
C2 ðN Þ
N T0
C2 ðN Þ
N T0
C2 ðN Þ
3.0 2.0 2.0 1.0
0.508 0.684 0.887 1.330
6.0 5.0 4.0 4.0
0.213 0.253 0.299 0.377
9.0 8.0 7.0 6.0
0.128 0.146 0.164 0.190
13.0 12.0 10.0 9.0
0.089 0.100 0.110 0.122
Table 7.3 Optimum T3 and C3 ðT3 Þ when T0 ¼ 1 and lT0 ¼ 1 cK =cT
K 5
5 10 20 50
10
15
20
T3
CðT3 Þ
T3
CðT3 Þ
T3
CðT3 Þ
T3
CðT3 Þ
2.5 2.0 1.8 1.6
0.483 0.559 0.626 0.704
6.0 5.1 4.6 4.1
0.198 0.219 0.237 0.260
9.2 8.7 7.8 7.1
0.119 0.130 0.139 0.149
13.1 12.0 11.3 10.6
0.084 0.090 0.095 0.102
7.3 Lévy Process We briefly explain a Lévy process appearing in stochastic processes [9]. Definition 7.2 A stochastic process fXðtÞg is said to be a Lévy process if (i) Xð0Þ ¼ 0; (ii) the increments Xðt1 Þ Xðt0 Þ; Xðt2 Þ Xðt1 Þ; . . .; Xðtn Þ Xðtn1 Þ are independent for all 0 t0 \t1 \ \tn ; (iii) the distribution of increment Xðtk Þ Xðtk1 Þ depends only the difference tk tk1 : A Lévy process is a stochastic process with independent and stationary increments. For the strict definition of a Lévy process, we have to add two regularity conditions, however, we omit them to exceed mathematically the limit of the book
192
7 Brownian Motion and Lévy Processes
[10, 11]. From this definition, the increment XðtÞ XðuÞ is independent of the process fXðxÞ; 0 x ug and has the same distribution of Xðt uÞ for t u 0: Special cases of Lévy processes are the Brownian Motion with continuous increment and the Poisson process with jump increment. In other words, the Poisson process is the jump Lévy process and the Brownian Motion is the continuous Lévy process. The Lévy process has been applied to various fields such as physics, finance, and statistics [11]. However, there have been few papers treated with reliability models using Lévy processes: When damage and wear models form a Lévy process, the damage models have been discussed analytically [12–14]. Example 7.3 (Continuous damage model) It is assumed in Sect. 7.2.1.3 that pffi ZðtÞ ¼ xt þ Bt ; where Bt has an exponential distribution 1 ex=r t : That is, the total damage usually increases linearly with time t; however, it undergoes positively some change according to an exponential distribution. Then, pffi PrfZðtÞ xg ¼ PrfBt x xtg ¼ 1 exp½ðx xtÞ=r t; pffi EfZðtÞg ¼ xt þ r t;
ð7:25Þ
2
VfZðtÞg ¼ r t:
Thus, when the unit is replaced at time T; the expected cost rate is, from (7.20), pffiffiffiffi cK ðcK cT Þf1 exp½ðK xTÞ=r T g : C1 ðTÞ ¼ RT pffi 0 f1 exp½ðK xtÞ=r tgdt
ð7:26Þ
Differentiating C1 ðTÞ with respect to T and setting it equal to zero, pffiffiffiffi ZT pffi ðK=T þ xÞ exp½ðK xTÞ=r T pffiffiffiffi pffiffiffiffi f1 exp½ðK xtÞ=r tgdt 2r T f1 exp½ðK xTÞ=r T g 0
pffiffiffiffi exp½ðK xTÞ=r T ¼
ð7:27Þ
1 : cK =cT 1
Table 7.4 presents optimum T and expected cost rate C1 ðT Þ=cT for cK =cT ¼ 5; 10; 20; 50 and K ¼ 5; 10; 15; 20 when r2 ¼ 1 and x ¼ 1: Compared to Table 7.3, optimum T are less than T3 because the average amount of damage is greater than that in the continuous model (Sect. 7.2.1.3). However, if the damage level would be estimated at a lowest level and additional damage suffered from environment change and frequent disasters would be caused randomly, this would be applied to such models (Problem 7.8). On the other hand, if ZðtÞ ¼ xt Bt ; i.e., the total damage changes negatively according to an exponential distribution, then
7.3 Lévy Process
193
Table 7.4 Optimum T and C1 ðT Þ=cT when x ¼ r2 ¼ 1 K cK =cT 5 5 10 20 50
10
15
20
T
C1 ðT Þ=cT
T
C1 ðT Þ=cT
T
C1 ðT Þ=cT
2.2 1.6 1.1 0.8
0.7619 1.1029 1.3361 1.8131
4.8 3.8 3.1 2.3
0.2920 0.3648 0.4461 0.5683
7.8 6.4 5.3 4.3
0.1696 0.2045 0.2424 0.2983
T 10.9 9.2 7.8 6.5
pffi PrfZðtÞ xg ¼ PrfBt xt xg ¼ exp½ðxt xÞ=r t; pffi EfZðtÞg ¼ xt r t; VfZðtÞg ¼ r2 t:
C1 ðT Þ=cT 0.1163 0.1371 0.1594 0.1918
ð7:28Þ
Thus, the expected cost rate is (Problem 7.9) pffiffiffiffi cK ðcK cT Þ exp½ðxT KÞ=r T C2 ðTÞ ¼ : RT pffi 0 exp½ðxt KÞ=r tdt
ð7:29Þ
Example 7.4 [6, p. 35] Consider the damage model with two kinds of damage in Sect. 6.3.4. One of them is caused by shock and is additive in Sect. 6.3.1, and the other increases proportionately with time t; i.e., the total damage is YðtÞ at þ ZðtÞ; where ZðtÞ is given in (7.12). The unit fails whether the total damage exceeds a failure level with time or has exceeded in both cases, at some shock, and the failure is detected only at time of shocks. Denote that Sn X1 þ X2 þ þ Xn ; Zn ¼ W1 þ W2 þ þ Wn ðn ¼ 1; 2; . . .), where l EfXn g: The distribution of time to detect a failure at some shock is, from (6.36), 1 X
PrfZn þ aSn \K Znþ1 þ aSnþ1 ; Snþ1 tg
n¼0
¼
8 tu t 1 Z
0
:
0
9 = i GðnÞ ðK auÞ Gðnþ1Þ ðK aðu þ xÞÞ dFðxÞ dF ðnÞ ðuÞ: ; ð7:30Þ
The mean time to detect a failure at time shock is, from (6.37), 1 Z X
K=a
a
n¼0
GðnÞ ðK atÞdF ðnÞ ðtÞ:
ð7:31Þ
0
Next, suppose that the unit is replaced at time T or at failure, i.e., when the total damage has exceeded K at some shock, whichever occurs first [6, p. 35, 15]. Then, the probability that the unit is replaced because failure is detected at some shock is, from (7.30),
194
7 Brownian Motion and Lévy Processes 1 X
PrfZn þ aSn K; Znþ1 þ aSnþ1 [ K; Snþ1 Tg
n¼0 T 1 Z h i X GðnÞ ðK atÞ Gðnþ1Þ ðK atÞ dF ðnþ1Þ ðtÞ ¼ n¼0
þ
0
8 Tt T 1 Z
0
:
0
9 = i GðnÞ ðK atÞ Gðnþ1Þ ðK aðt þ xÞÞ dFðxÞ dF ðnÞ ðtÞ; ; ð7:32Þ
and the probability that it is replaced at time T before failure is 1 X
PrfZn þ aSn K; Sn T Snþ1 g
n¼0 1 Z X
ð7:33Þ
T
¼
n¼0
ðnÞ
ðnÞ
FðT tÞG ðK atÞdF ðtÞ;
0
where (7.32) ? (7.33) = 1. The mean time to replacement is, from (7.32) and (7.33), 1 Z X
T
T
n¼0
FðT tÞGðnÞ ðK atÞdF ðnÞ ðtÞ
0
T 1 Z h i X þ t GðnÞ ðK atÞ Gðnþ1Þ ðK atÞ dF ðnþ1Þ ðtÞ n¼0
0
8
9 = i ðt þ xÞ GðnÞ ðK atÞ GðnÞ ðK aðt þ xÞÞ dFðxÞ dF ðnÞ ðtÞ þ : ; n¼0 0 0 2 3 T Z Tt 1 Z X 4 ¼ FðxÞdx5GðnÞ ðK atÞdF ðnÞ ðtÞ: T Tt 1 Z
n¼0
0
h
0
ð7:34Þ Thus, the expected cost rate is, from (7.33) and (7.34), P RT ðnÞ ðnÞ cK ðcK cT Þ 1 n¼0 0 FðT tÞG ðK atÞdF ðtÞ h i ; CðTÞ ¼ P1 R T R Tt FðxÞdx GðnÞ ðK atÞdF ðnÞ ðtÞ n¼0 0 0
ð7:35Þ
where cK is the replacement cost at failure, and cT is the replacement cost at time T; where cK [ cT :
7.3 Lévy Process
195
Table 7.5 Optimum T when 1=k ¼ 1=h ¼ 1:0 a K ¼ 100 cK =cT 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.5 3.0 4.0 5.0 6.0
K ¼ 200 cK =cT
2
5
10
100
2
5
10
100
81 69 59 52 47 42 39 36 33 31 29 25 22 18 15 13
72 62 54 48 43 39 36 33 31 29 27 23 21 17 14 12
68 58 51 45 41 37 34 32 29 28 26 22 20 16 14 12
59 51 45 40 37 33 31 29 27 25 24 21 18 15 13 11
169 142 123 108 97 87 80 74 68 63 59 51 45 36 30 26
157 133 115 102 92 83 76 70 65 61 57 49 43 35 29 25
151 128 112 99 89 81 74 68 64 59 56 48 43 34 29 25
138 118 103 92 83 76 69 64 60 56 53 46 40 33 28 24
We find an optimum T that minimizes CðTÞ when shocks occur in a Poisson P k kt process, i.e., F ðnÞ ðtÞ ¼ 1 ðn ¼ 0; 1; 2; . . .Þ: Then, differentiating k¼n ½ðktÞ =k!e CðTÞ with respect to T and setting it equal to zero, Z T Z T BðTÞ kcT kt kt 1þ e dBðtÞ e BðtÞ dt ¼ ; ð7:36Þ RT c K cT 1 þ 0 B ðtÞ dt 0 0 where BðtÞ k
1 X ðktÞn n¼0
n!
Gðnþ1Þ ðK atÞ:
Next, it is assumed that GðxÞ ¼ 1 ehx : Then, the average damage per unit of time is a þ k=h: So that, the mean time until the total damage exceeds a failure level K is K=ða þ k=hÞ: Table 7.5 presents the optimum T that satisfies (7.36) for K ¼ 100; 200; cK =cT ¼ 2; 5; 10; 100 and a changes from 0 to 6 when 1=k ¼ 1=h ¼ 1:0: For example, when K ¼ 200; cK =cT ¼ 5; a ¼ 0:2; K=ða þ k=hÞ ¼ 167 and T ¼ 133; i.e., the unit with the meantime 167 to failure should be replaced at time 133. In other words, if the operating time is ð133=167Þ 100 80% of the mean time to failure, the unit should be replaced before failure. This indicates that optimum T decreases when both cK and a increase. This means that if a is large, the total damage increases mainly with time, and hence, we should adopt the scheduled replacement time T (Problem 7.10). Therefore, the rate a is an important factor for deciding the optimum policy (Problem 7.11).
196
7 Brownian Motion and Lévy Processes
7.4 Problems 7 7.1 When the density is 1 2 2 f ðx; tÞ ¼ pffiffiffiffiffiffiffi ex =ð2r tÞ ; 2ptr show that Z1
esx f ðx; tÞdx ¼ es
2 2
r t=2
:
1
7.2 In (1) of Example 7.1, compute the reliability RðtÞ for t ¼ 5; 10 when l ¼ r ¼ 1 and K ¼ 10; and the MTTF. 7.3 Prove in Example 7.2 that an optimum T ð0\T \1Þ is a finite and unique solution of (7.10), and the resulting cost rate is given in (7.11). 7.4 In Example 7.2, CðTÞ in (7.9) is approximately given by # " 2 ðl þ r2 ÞT r2 c2 e þ : CðTÞ ¼ c1 l þ 1þ 2 2 T
7.5
7.6
7.7 7.8 7.9 7.10 7.11
e e e that minimizes CðTÞ: Derive CðTÞ and an optimum T Furthermore, show e numerically that T becomes a good approximation to T : Prove that if Q1 ðTÞ increases strictly with T and Q1 ð1Þ½1 þ MG ðKÞ [ cK = ½lðcK cT Þ; then there exists a finite and unique T1 ð0\T1 \1Þ that satisfies (7.15). Prove that if Q2 ðNÞ increases strictly with N and Q2 ð1Þ½1 þ MGT ðKÞ [ cK = ðcK cT Þ; then there exists a finite and unique minimum N ð1 N \1Þ that satisfies (7.18). Prove that if rA ðtÞ increases strictly to infinity, then there exists a finite and unique T3 that satisfies (7.21). In Example 7.3, rewrite all results when Bt has a Weibull distribution p ffi
PrfBt xg ¼ 1 exp xa =r t for a 1: Derive an optimum T that minimizes C2 ðTÞ in (7.29). Explain the physical meaning of why optimum T decreases with a: In Example 7.4, consider the damage model where the unit is replaced at shock number N or at failure, and discuss an optimum N that minimizes the expected cost [15].
References 1. Ross SM (1983) Stochastic processes. Wiley, New York 2. Karlin S, Taylor HM (1975) A first course in stochastic processes. Academic Press, New York 3. Birnbaum ZW, Saunders SC (1969) A new family of life distributions. J Appl Probab 6:319– 327
References
197
4. Owen WJ, Padgett WJ (2003) Accelerated test models with the Birnbaum–Saunders distribution. In: Pham H (ed) Handbook of reliability engineering. Springer, London, pp 429– 439 5. Park C, Padgett WJ (2005) New cumulative damage models for failure using stochastic processes as initial damage. IEEE Trans Reliab 54:530–540 6. Nakagawa T (2007) Shock and damage models in reliability theory. Springer, London 7. Beichelt FE (2006) Stochastic processes in science, engineering and finance. Chapman & Hall, Boca Raton 8. Ito K, Nakagawa T (2008) Comparison of three cumulative damage models. In: Sheu SH, Dohi T (eds) Advanced reliability modeling III. McGraw-Hill, Taipei, pp 332–338 9. Prabhu NU (2007) Stochastic processes. World Scientific, Singapore 10. Bertoin J (1996) Lévy processes. Cambridge University Press, Cambridge 11. Sato K (2001) Basic results on Lévy processes. In: Bandorff-Nielsen O, Mikosch T, Resnick S (eds) Lévy processes: theory and applications. Birkhäuser, Boston 12. Zuckerman D (1978) Optimal replacement policy for the case where the damage process is a one-sided Lévy process. Stoch Process Appl 7:141–151 13. Abdel-Hameed M (2010) Degradation process: an overview. In: Nikulin MS, Limnios N, Balakrishnan N, Kahle W, Huber-Carol C (eds) Advances in degradation modeling. Birkhäuser, Boston, pp 17–25 14. Lehmann A (2010) Failure time models based on degradation processes. In: Nikulin MS, Limnios N, Balakrishnan N, Kahle W, Huber-Carol C (eds) Advances in degradation modeling. Birkhäuser, Boston, pp 209–233 15. Satow T, Nakagawa T (1997) Replacement policies for a shock model with two kinds of damages. In: Osaki S (ed) Stochastic modeling in innovative manufacturing. Springer Lecture notes in economics and mathematical systems, vol 445. Springer, Berlin, pp 188–195
Chapter 8
Redundant Systems
High system reliability can be achieved by providing redundancy and maintenance. It is not too much to say that learning reliability is analyzing redundant systems and deriving optimum maintenance policies. In the preceding chapters, we have already given many useful examples of reliability models to understand naturally stochastic processes, and conversely, to apply the theory of stochastic processes to actual reliability models. As final examples of reliability models, we take up standard redundant systems and show systematically how to use well the techniques of stochastic processes to analyze them and to obtain their reliability properties theoretically. This would be greatly helpful for understanding stochastic processes and learning reliability theory [1, p. 160]. In Sect. 8.1, we analyze one-unit systems with no repair, that is, the simplest and fundamental reliability system, using Poisson, nonhomogeneous Poisson, and renewal processes. This section aims not only to make differences among three processes, but also to apply renewal processes to standard replacement policies. Furthermore, we obtain several reliability measures of a one-unit system with repair, using an alternating renewal process. In Sect. 8.2, we take up two-unit systems with repair. When both failure and repair times have exponential distributions, we obtain transition probabilities, using a continuous-time Markov chain. Further, when failure and repair times have general distributions, we analyze the system, using a Markov renewal process. This indicates that Markov renewal processes are the most powerful tool for analyzing reliability systems. It has been presented until now that states are well defined in stochastic processes and their behaviors are derived analytically. In Sect. 8.3, it can be shown as an example of a standby system with spare units that the system is analyzed and its several reliability measures are derived by forming renewal equations directly, and by taking their Laplace-Stieltjes (LS) transforms and forming generating functions. This would be useful for analyzing simple noncomplicated reliability models.
T. Nakagawa, Stochastic Processes, Springer Series in Reliability Engineering, DOI: 10.1007/978-0-85729-274-2_8, Ó Springer-Verlag London Limited 2011
199
200
8 Redundant Systems
8.1 One-Unit System 8.1.1 Poisson Process As the simplest redundant system, we consider a one-unit system where a unit begins to operate at time 0 and has a failure distribution ð1 ekt Þ. When the unit fails, it is repaired, becomes like new, and begins to operate again. The repair time is very small compared to its operating time and is assumed to be negligible. In other words, the unit is assumed to be replaced immediately with a new one at each failure and its replacement time is negligible. Let Xk (k ¼ 1; 2; . . .) be the failure times of the kth operating unit that are independent and identically distributed random variables with an exponential distribution ð1 ekt Þ, and Sn X1 þ X2 þ þ Xn (n ¼ 1; 2; . . .). From Chap. 2, the distribution of Sn is a gamma distribution with order n. Thus, the number NðtÞ of failures in ½0; t is, from (2.12) and (2.16), PrfNðtÞ ¼ ng ¼ PrfSn t\Snþ1 g ¼ PrfSn tg PrfSnþ1 tg 1 1 X X ðktÞk kt ðktÞk kt e e ¼ k! k! k¼n k¼nþ1 ¼
ðktÞn kt e n!
ðk ¼ 0; 1; 2; . . .Þ;
ð8:1Þ
that is a Poisson distribution with mean kt. It is said that failures occur in a Poisson process with rate k, i.e., the failures occur randomly independent of its age and operating time, and its residual life at any time t also has the same exponential distribution (1 ekt ) because of its memoryless property.
8.1.2 Nonhomogeneous Poisson Process Suppose that the unit undergoes minimal repair at failures and its failure rate hðtÞ remains unchanged by minimal repair given in Definition 2.3 of Sect. 2.3. The unit begins to operate again after the completion of minimal repair, where the time for minimal repair is negligible. Then, the expected number EfNðtÞg of failures in Rt ½0; t is HðtÞ 0 hðuÞdu, and the number NðtÞ of failures in [0, t] is, from (2.42), PrfNðtÞ ¼ ng ¼
½HðtÞn HðtÞ e n!
ðn ¼ 0; 1; 2; . . .Þ;
ð8:2Þ
that is a Poisson distribution with mean HðtÞ. It is said that failures occur in a nonhomogeneous Poisson process with an intensity function hðtÞ and a mean value function HðtÞ. If FðtÞ is IFR, i.e., hðtÞ increases with time t, failures are apt to fail with age. Some variations for minimal repair and preventive maintenance are
8.1 One-Unit System
201
well-known as the imperfect repair [2] and preventive maintenance, and were summarized [3, p. 96].
8.1.3 Renewal Process Suppose that the unit has a general failure distribution FðtÞ with finite mean l. Then, letting F ðnÞ ðtÞ be the n-fold Stieltjes convolution of FðtÞ with itself, from (3.3), (3.4), and (3.18), PrfSn tg ¼ F ðnÞ ðtÞ; MðtÞ 1 ¼ ; t!1 t l lim
PrfNðtÞ ¼ ng ¼ F ðnÞ ðtÞ F ðnþ1Þ ðtÞ; 1 X F ðnÞ ðtÞ: EfNðtÞg ¼ MðtÞ ¼
ð8:3Þ
n¼1
It is said that failures occur in a renewal process with an interarrival distribution FðtÞ. Most reliability systems fail in accordance with renewal processes. The renewal function MðtÞ plays an important role in analyzing reliability systems. However, in general, it is practically impossible to explicitly express MðtÞ for a general distribution FðtÞ because it is very difficult to compute the n-fold Stieltjes convolution F ðnÞ ðtÞ of FðtÞ explicitly except exponential and gamma distributions given in Example 3.2. Thus, we derive the Laplace-Stieltjes (LS) transform of MðtÞ from Z 1 F ðsÞ M ðsÞ ¼ est dMðtÞ ¼ ; ð8:4Þ 1 F ðsÞ 0 and compute MðtÞ explicitly only by inverting M ðsÞ numerically, as shown in Sect. 3.2.3. A unit is planned to be replaced at periodic times kT (k ¼ 1; 2; . . .) and failures occur in a renewal process between replacements, i.e., each failed unit is replaced with a new one. Then, from (3.44), the expected cost rate is C1 ðTÞ ¼
c1 MðTÞ þ c2 ; T
ð8:5Þ
where c1 ¼ cost of replacement at each failure and c2 ¼ cost of planned replacement at time T. Similarly, suppose that failures occur in a nonhomogeneous Poisson process with a mean value function HðtÞ, i.e., each failed unit undergoes minimal repair. Then, from (3.46), the expected cost rate is C2 ðTÞ ¼
c1 HðTÞ þ c2 ; T
ð8:6Þ
where c1 ¼ cost of minimal repair at each failure and c2 is given in (8.5). Next, suppose that a unit is replaced at time T or at failure, whichever occurs first. Then, defining Zk min fXk ; Tg, the process fZk g has a renewal process with interarrival distribution
202
8 Redundant Systems
PrfZk tg ¼
FðtÞ t T; 1 t [ T;
and hence, EfZk g ¼
Z
T
t dFðtÞ þ TFðTÞ ¼
0
Z
T
FðtÞdt; 0
where uðtÞ 1 uðtÞ for any function uðtÞ. Then, from (3.42), the expected cost rate is C3 ðTÞ ¼
c1 FðTÞ þ c2 FðTÞ ; RT 0 FðtÞdt
ð8:7Þ
where c1 ¼ replacement cost at failure and c2 ¼ replacement cost at time T:
8.1.4 Alternating Renewal Process It is assumed in Sect. 8.1.3 that each repair time is not negligible and has an identical distribution GðtÞ with finite mean b. Let Yk (k ¼ 1; 2; . . .) be the repair times of the kth failed unit that are independent and identically distributed random variables with GðtÞ. We define the following states: State 0: Unit is operating. State 1: Unit is under repair. The unit repeats operating and repair alternately in Figs. 3.5 and 4.2. It is said that the unit forms an alternating renewal process with States 0 and 1. Clearly, the process fXk þ Yk g has a renewal process with interarrival distribution Rt FðtÞ GðtÞ ¼ 0 Gðt uÞdFðuÞ and mean ðl þ bÞ: Suppose that the unit begins to operate at time 0. Let Nj ðtÞ be the number of visits to State j in ½0; t, where the first visit to State 0 at time 0 is not counted, and Mj ðtÞ EfNj ðtÞg that is called the renewal function. Then, from (3.53), Z 1 F ðsÞG ðsÞ ; est dM0 ðtÞ ¼ M0 ðsÞ ¼ 1 F ðsÞG ðsÞ 0 ð8:8Þ Z 1 F ðsÞ st M1 ðsÞ ¼ e dM1 ðtÞ ¼ : 1 F ðsÞG ðsÞ 0 Let ZðtÞ be the process at time t, and Pj ðtÞ PrfZðtÞ ¼ jg (j ¼ 0; 1). Then, from (3.55), P0 ðsÞ ¼
1 F ðsÞ ; 1 F ðsÞG ðsÞ
P1 ðsÞ ¼
F ðsÞ½1 G ðsÞ : 1 F ðsÞG ðsÞ
ð8:9Þ
8.1 One-Unit System
203
Thus, the limiting probabilities Pj limt!1 Pj ðtÞ are P0 ¼ lim P0 ðsÞ ¼ s!0
l ¼ 1 P1 ; lþb
ð8:10Þ
that represents the steady-state availability of the unit. Furthermore, the interval reliability is the probability that at a specified time, the unit is operating and will continue to operate for an interval of duration [3, pp. 48, 140]. Then, the interval reliability Rðx; tÞ for an interval of duration x starting at time t is, from (3.62) to (3.64), Z t Rðx; tÞ ¼ Fðt þ xÞ þ Fðt þ x uÞdM0 ðuÞ; 0
and its Laplace transform in Appendix A is R1 Z 1 esx x est FðtÞdt st e Rðx; tÞdt ¼ : R ðx; sÞ ¼ 1 F ðsÞG ðsÞ 0 Thus, the limiting interval reliability is
RðxÞ lim Rðx; tÞ ¼ lim sR ðx; sÞ ¼ t!1
s!0
R1 x
FðtÞdt : lþb
ð8:11Þ
Next, suppose that an operating unit undergoes repair at time T or at failure, whichever occurs first, from the beginning of its operation, and define Zk min fXk ; Tg. Then, the process fZk ; Yk g forms an alternating renewal process. On the other hand, when the repair is not completed until time T, it is replaced with a new one, that is called a repair limit policy in Example 3.7. Then, defining Zk min fYk ; Tg, the process fXk ; Zk g forms an alternating renewal process. In this RT case, because EfZk g ¼ 0 GðtÞdt, the expected cost rate is, from (3.82), RT c1 GðTÞ þ 0 GðtÞdcr ðtÞ CðTÞ ¼ ; ð8:12Þ RT l þ 0 GðtÞdt where c1 ¼ replacement cost of a failed unit when its repair is not completed until time T and cr ðtÞ ¼ repair cost during ½0; t (Problem 8.1).
8.2 Two-Unit Standby System A two-unit standby redundant system with a single repair facility is one of the most fundamental and important redundant systems in reliability theory [4]. The system consists of two units, where one unit is used for operation and the other is in standby as an initial condition. If an operating unit fails, then it undergoes repair immediately and the other standby unit takes over its operation. Either of the two
204
8 Redundant Systems
units is alternately operating. It is said that system failure occurs when both units are down simultaneously. The transition probabilities, the first-passage distributions, and the renewal functions are derived by using the techniques of Markov processes.
8.2.1 Exponential Failure and Repair Times An operating unit has an exponential failure distribution (1 ekt ), and a failed unit also has an exponential repair distribution (1 eht ). If an operating unit fails and the other unit is in standby, the failed unit undergoes repair immediately, and the standby unit takes over its operation immediately. However, if an operating unit fails while the other unit is under repair, the failed unit has to wait until the repair facility become empty. This means that system failure has occurred. It is also assumed that a failed unit recovers its functioning upon repair completion and becomes as good as new. The unit in standby neither deteriorates nor fails in the standby interval. Two units are used alternately for its operation, as described above. Even if system failure occurs, the system can operate again upon repair completion. That is, the system repeats up and down alternately. All random variables considered here are independent of each other. To analyze the above system, we define the following three system states that represent the states of the process: State 0: One unit is operating and the other unit is in standby. State 1: One unit is operating and the other unit is under repair. State 2: One unit is under repair and the other unit is waiting for repair. The system states defined above are regeneration points because the failure and repair times are exponential. Let fXðtÞ ¼ jg denote that the process is in State j (j ¼ 0; 1; 2) at time t. Then, the process fXðtÞg forms a continuous-time Markov chain with finite state set f0; 1; 2g. Let Pij ðtÞ PrfXðtÞ ¼ jjXð0Þ ¼ ig be the transition probabilities from State i to State j at time t. By the similar method in Example 4.11, P000 ðtÞ ¼ kP00 ðtÞ þ hP01 ðtÞ; P001 ðtÞ ¼ kP00 ðtÞ ðk þ hÞP01 ðtÞ þ hP02 ðtÞ; P002 ðtÞ ¼ kP01 ðtÞ hP02 ðtÞ: Taking the LS transformations of the above equations, sP00 ðsÞ s ¼ kP00 ðsÞ þ hP01 ðsÞ; sP01 ðsÞ ¼ kP00 ðsÞ ðk þ hÞP01 ðsÞ þ hP02 ðsÞ; sP02 ðsÞ ¼ kP01 ðsÞ hP02 ðsÞ:
8.2 Two-Unit Standby System
205
Solving these equations, s2 þ ðk þ 2hÞs þ h2 ; s2 þ 2ðk þ hÞs þ k2 þ kh þ h2 kðs þ hÞ ; P01 ðsÞ ¼ s2 þ 2ðk þ hÞs þ k2 þ kh þ h2
P00 ðsÞ ¼
P02 ðsÞ ¼
k2 : s2 þ 2ðk þ hÞs þ k2 þ kh þ h2
Taking the inverse LS transforms, h2 k x1 k h x1 t x2 k h x2 t ; P00 ðtÞ ¼ 2 þ e e x1 x2 k þ kh þ h2 x1 x2 kh k x1 þ h x1 t x2 þ h x2 t ; P01 ðtÞ ¼ 2 þ e e 2 x1 x2 x1 x2 k þ kh þ h k2 k2 1 x1 t 1 x2 t P02 ðtÞ ¼ 2 þ e e ; ð8:13Þ x2 k þ kh þ h2 x1 x2 x1 where P00 ðtÞ þ P01 ðtÞ þ P02 ðtÞ ¼ 1 (Problem 8.2), and x1 k þ h þ
pffiffiffiffiffi kh;
x2 k þ h
pffiffiffiffiffi kh:
8.2.2 General Failure and Repair Times An operating unit has a general failure distribution FðtÞ with finite mean l, and a failed unit has a general repair distribution GðtÞ with finite mean b. The other assumptions are the same ones in Sect. 8.2.1. However, State 0 is not a regeneration point except when an operating unit fails for the first time from the beginning of its operation at time 0, and the failure time is exponential. From this reason, we define an initial time instant: State -1: One unit begins to operate and the other unit is in standby. The other states are the same as those in Sect. 8.2.1. An epoch at which the process makes a transition into State 1 is a regeneration point. However, the epochs at which the system makes a transition into State j (j ¼ 0; 2) are not regeneration points, except the failure and repair times are exponential (Fig. 5.3). We analyze the system by using the techniques of a Markov renewal process in Sect. 5.3. Define a mass function Qij ðtÞ from State i (i ¼ 1; 1) to State j (j ¼ 0; 1; 2) by the probability that after making a transition into State i, the process next makes a transition into State j, in a smaller amount of time than
206
8 Redundant Systems
time t in a Markov renewal process. Then, from Sect. 5.3.2, we have the following mass functions: Q11 ðtÞ ¼ FðtÞ; Z t FðuÞdGðuÞ; Q10 ðtÞ ¼
Q12 ðtÞ ¼
0
Z
t
GðuÞdFðuÞ:
ð8:14Þ
0
It is impossible to define the mass function Q01 ðtÞ and Q21 ðtÞ because the epochs for States 0 and 2 are not regeneration points. Thus, define a new mass ð0Þ ð2Þ function Q11 ðtÞ ðQ11 ðtÞÞ that is the recurrence time distribution for State 1 via ð0Þ
ð2Þ
State 0 (State 2). That is, Q11 ðtÞ ðQ11 ðtÞÞ is the probability that after making a transition into State 1, the process next makes a transition into State 0 (State 2) and returns to State 1 in time t. Then, from Sect. 5.3.2, Z t Z t ð0Þ ð2Þ Q11 ðtÞ ¼ GðuÞdFðuÞ; Q11 ðtÞ ¼ FðuÞdGðuÞ: ð8:15Þ 0
0
Therefore, the LS transforms of the mass functions are Q11 ðsÞ ¼ F ðsÞ; Z 1 est FðtÞdGðtÞ; Q10 ðsÞ ¼ Z0 1 ð0Þ Q11 ðsÞ ¼ est GðtÞdFðtÞ; 0
Q12 ðsÞ ð2Þ
¼
Q11 ðsÞ ¼
Z
1
Z0 1
est GðtÞdFðtÞ;
ð8:16Þ
est FðtÞdGðtÞ:
0
8.2.2.1 First-Passage Distribution We obtain the first-passage distributions and their mean times by using the mass functions. Let Hij ðtÞ denote the first-passage distributions from State i (i ¼ 1; 1) to State j (j ¼ 0; 1; 2). First, suppose that the process starts in the epoch for State 1. Then, from (8.15), the recurrence time distribution H11 ðtÞ for State 1 is ð0Þ
ð2Þ
H11 ðtÞ ¼ Q11 ðtÞ þ Q11 ðtÞ ¼ FðtÞGðtÞ; that is the probability that the process comes back to State 1 via State 0 or State 2, whichever occurs first, and its LS transform is, from (8.16), Z 1 Z 1 Z 1 st st H11 ðsÞ ¼ e dH11 ðtÞ ¼ e FðtÞdGðtÞ þ est GðtÞdFðtÞ: ð8:17Þ 0
0
0
Thus, the mean recurrence time l11 for State 1 is Z 1 t dH11 ðtÞ ¼ l þ b c; l11 ¼ 0
ð8:18Þ
8.2 Two-Unit Standby System
207
R1 where c 0 FðtÞGðtÞdt. The first-passage distributions H1j ðtÞ (j ¼ 0; 2) are given by solving the following renewal-type equation: ð2Þ
H10 ðtÞ ¼ Q10 ðtÞ þ Q11 ðtÞ H10 ðtÞ;
ð8:19Þ
where the asterisk mark represents the Stieltjes convolution, i.e., aðtÞ bðtÞ Rt 0 bðt uÞdaðuÞ for any functions aðtÞ and bðtÞ. The first term at the right-hand side in (8.19) represents the probability that the process goes to State 0 directly, and the second term is the probability that the process goes to State 0 after returning to State 1 via State 2. Similarly, ð0Þ
H12 ðtÞ ¼ Q12 ðtÞ þ Q11 ðtÞ H12 ðtÞ:
ð8:20Þ
If the process starts from State -1, then, H1j ðtÞ ¼ Q11 ðtÞ H1j ðtÞ
ðj ¼ 0; 2Þ:
Therefore, taking the LS transforms of (8.20)–(8.22) and solving them, R1 Q ðsÞQ ðsÞ F ðsÞ 0 est FðtÞdGðtÞ R1 H10 ðsÞ ¼ 11 ð2Þ10 ¼ ; 1 0 est FðtÞdGðtÞ 1 Q11 ðsÞ R1 Q11 ðsÞQ12 ðsÞ F ðsÞ 0 est GðtÞdFðtÞ R : ¼ H12 ðsÞ ¼ 1 ð0Þ 1 0 est GðtÞdFðtÞ 1 Q11 ðsÞ
ð8:21Þ
ð8:22Þ
The mean first-passage times l1j (j ¼ 0; 2), starting from State -1, are 1 H10 ðsÞ b ; ¼ l þ R1 s!0 s FðtÞdGðtÞ 0 " # 1 H12 ðsÞ 1 ¼ l 1 þ R1 ; ¼ lim s!0 s GðtÞdFðtÞ 0
l10 ¼ lim l12
ð8:23Þ
that represent the mean times to system recovery and system failure, respectively. Furthermore, the mean time of the busy period of a repair facility is 1 H10 ðsÞ b : ¼ l10 l ¼ R 1 s!0 s 0 FðtÞdGðtÞ
l10 ¼ lim
The mean downtime ld after system failure is Z 1 Z 1 ð2Þ ld ¼ t dQ11 ðtÞ t dQ12 ðtÞ 0 Z0 1 FðtÞGðtÞdt ¼ b c: ¼ 0
208
8 Redundant Systems
8.2.2.2 Expected Number of Visits to States Consider the expected number of visits to State j (j ¼ 0; 1; 2) during the interval ½0; t. Let the renewal function Mij ðtÞ be the expected number of occurrences of State j in ½0; t, starting in the epoch for State i (i ¼ 1; 1) at time 0, where the first visit to j is not counted if i ¼ j. First, when the process starts from State 1 at time 0, we have the following renewal-type equation: ð0Þ
ð0Þ
ð2Þ
M10 ðtÞ ¼ Q10 ðtÞ Q11 ðtÞ þ Q11 ðtÞ ½1 þ M10 ðtÞ þ Q11 ðtÞ M10 ðtÞ ¼ Q10 ðtÞ þ H11 ðtÞ M10 ðtÞ;
ð8:24Þ
where the first term is the probability that the process goes to State 0 directly and then remains in State 0, the second term is the expected number that the process returns to State 1 via State 0 and goes to State 0, and finally, the third term is the expected number that the process returns to State 1 via State 2 and then goes to State 0. Similarly, ð0Þ
ð2Þ
M11 ðtÞ ¼ Q11 ðtÞ ½1 þ M11 ðtÞ þ Q11 ðtÞ ½1 þ M11 ðtÞ ¼ H11 ðtÞ ½1 þ M11 ðtÞ; ð2Þ
ð0Þ
ð2Þ
M12 ðtÞ ¼ Q12 ðtÞ Q11 ðtÞ þ Q11 ðtÞ M12 ðtÞ þ Q11 ðtÞ ½1 þ M12 ðtÞ
ð8:25Þ
¼ Q12 ðtÞ þ H11 ðtÞ M12 ðtÞ: The expected numbers of visits to State j (j ¼ 0; 1; 2) in ½0; t, starting from State -1, are M11 ðtÞ ¼ Q11 ðtÞ ½1 þ M11 ðtÞ; ð8:26Þ M1j ðtÞ ¼ Q11 ðtÞ M1j ðtÞ ðj ¼ 0; 2Þ: Taking the LS transforms of (8.24)–(8.26) and solving them, R1 Q11 ðsÞQ10 ðsÞ F ðsÞ 0 est FðtÞdGðtÞ ; M10 ðsÞ ¼ ðsÞ ¼ ðsÞ 1 H11 1 H11 Q11 ðsÞ F ðsÞ ðsÞ ¼ M11 ¼ ; ðsÞ 1 H11 ðsÞ 1 H11 R1 Q11 ðsÞQ12 ðsÞ F ðsÞ 0 est GðtÞdFðtÞ : M12 ðsÞ ¼ ðsÞ ¼ ðsÞ 1 H11 1 H11
ð8:27Þ
In addition, if FðtÞ and GðtÞ are non-lattice and their mean times are finite, from Theorem 3.2, there exist the limits Mj limt!1 Mij ðtÞ=t (j ¼ 0; 1; 2) of expected numbers of visits to State j per unit of time in the steady-state that are independent of an initial State i. Thus, by applying a Tauberian theorem in Appendix A.3 to (8.27), respectively,
8.2 Two-Unit Standby System
209
Z 1 1 M0 ¼ ¼ FðtÞdGðtÞ; l11 0 Z 1 1 ðsÞ ¼ GðtÞdFðtÞ; M2 ¼ lim sM12 s!0 l11 0 1 ðsÞ ¼ ¼ M0 þ M2 : M1 ¼ lim sM11 s!0 l11 lim sM10 ðsÞ s!0
A quantity of some interest is the total number of units that have failed and have been repaired in ½0; t, respectively. Letting Mif ðtÞ and Mir ðtÞ be the respected number of failed and repaired units in ½0; t, when the process starts from State i (i ¼ 1; 1), ð2Þ
M1f ðtÞ ¼ Q12 ðtÞ Q11 ðtÞ þ H11 ðtÞ ½1 þ M1f ðtÞ ¼ FðtÞGðtÞ þ H11 ðtÞ ½1 þ M1f ðtÞ;
ð8:28Þ
ð0Þ
M1r ðtÞ ¼ Q10 ðtÞ Q11 ðtÞ þ H11 ðtÞ ½1 þ M1r ðtÞ ¼ FðtÞGðtÞ þ H11 ðtÞ ½1 þ M1r ðtÞ: If the process starts from State -1, then M1f ðtÞ ¼ Q11 ðtÞ ½1 þ M1f ðtÞ;
ð8:29Þ
M1r ðtÞ ¼ Q11 ðtÞ M1r ðtÞ: Thus, combining (8.28) and (8.29), M1f ðsÞ ¼
M1r ðsÞ ¼
h i ð2Þ Q11 ðsÞ 1 þ Q12 ðsÞ Q11 ðsÞ ðsÞ 1 H11
h i ð2Þ Q11 ðsÞ Q10 ðsÞ þ Q11 ðsÞ ðsÞ 1 H11
; ð8:30Þ
:
Clearly, Mf lim
t!1
M1f ðtÞ M1r ðtÞ ¼ Mr ¼ lim ¼ M1 : t!1 t t
8.2.2.3 Transition Probabilities Denote that Pij ðtÞ is the transition probability that the process is in State j (j ¼ 0; 1; 2) at time t, starting from State i (i ¼ 1; 1) at time 0. Then, we have the following renewal equations:
210
8 Redundant Systems
ð0Þ
P10 ðtÞ ¼ Q10 ðtÞ Q11 ðtÞ þ H11 ðtÞ P10 ðtÞ ¼ FðtÞGðtÞ þ H11 ðtÞ P10 ðtÞ; P11 ðtÞ ¼ 1 Q10 ðtÞ Q12 ðtÞ þ H11 ðtÞ P11 ðtÞ
ð8:31Þ
¼ FðtÞGðtÞ þ H11 ðtÞ P11 ðtÞ; ð2Þ
P12 ðtÞ ¼ Q12 ðtÞ Q11 ðtÞ þ H11 ðtÞ P12 ðtÞ ¼ FðtÞGðtÞ þ H11 ðtÞ P12 ðtÞ: Furthermore, if the process starts from State -1, then P10 ðtÞ ¼ 1 Q11 ðtÞ þ Q11 ðtÞ P10 ðtÞ ¼ FðtÞ þ Q11 ðtÞ P10 ðtÞ; ðj ¼ 0; 2Þ: P1j ðtÞ ¼ Q11 ðtÞ P1j ðtÞ
ð8:32Þ
Thus, the LS transforms of P1j ðtÞ (j ¼ 0; 1; 2) are h i ð0Þ Q11 ðsÞ Q10 ðsÞ Q11 ðsÞ
P10 ðsÞ ¼ 1 Q11 ðsÞ þ ðsÞ 1 H11 Q ðsÞ 1 Q10 ðsÞ Q12 ðsÞ P11 ðsÞ ¼ 11 ; 1 H11 ðsÞ h i ð2Þ Q11 ðsÞ Q12 ðsÞ Q11 ðsÞ : P12 ðsÞ ¼ ðsÞ 1 H11
; ð8:33Þ
It is evident that because P10 ðsÞ þ P11 ðsÞ þ P12 ðsÞ ¼ 1, P10 ðtÞ þ P11 ðtÞ þ P12 ðtÞ ¼ 1. Note that P11 ðtÞ ¼ M11 ðtÞ M10 ðtÞ M12 ðtÞ. If FðtÞ and GðtÞ are non-lattice and their means are finite, there exist the limiting probabilities Pj limt!1 Pij ðtÞ (j ¼ 0; 1; 2) that are independent of an initial State i. Thus, by applying a Tauberian theorem to (8.33), b ; l11 lþb ; P1 ¼ lim P11 ðsÞ ¼ 1 þ s!0 l11 l P2 ¼ lim P12 ðsÞ ¼ 1 : s!0 l11 P0 ¼ lim P10 ðsÞ ¼ 1 s!0
It is noted that AðtÞ ¼ P10 ðtÞ þ P11 ðtÞ represents the pointwise availability of the system at time t, and AðtÞ ¼ P12 ðtÞ represents the pointwise unavailability at time t, given that two units are initial good. It is also noted that A ¼ P0 þ P1 ¼ l=l11 represents the steady-state availability and A ¼ P2 represents the steady-state
8.2 Two-Unit Standby System
211
unavailability (Problem 8.3). In other words, the so-called availability represents the ratio of the mean operating time to the total of mean operating and repair times [3, p. 9].
8.3 Standby System with Spare Units Consider the system with a main unit and n spare units [5]. It is assumed that the spare units are statistically not identical to the main unit, but have the same functions as the main unit, if they take over the system operation. The system operates in such a way that if the main unit fails then it undergoes repair immediately and one of the spare units replaces it. As soon as the repair of the main unit is completed, it begins to operate and the operating spare unit is available for further use. A failed spare unit is scrapped. The system can operate until the nth spare unit fails. In other words, the system fails when the last spare unit fails while the main unit is under repair. The operating behaviors of this model require the following knowledge: (1) Distribution of time and mean time to first system failure, given that n spare units are provided at time 0. (2) Probability that the number of failed spare units is exactly equal to n and its expected number in ½0; t. These quantities are derived from renewal equations and used for two optimization problems and to determine an initial number of spare units to stock.
8.3.1 First-Passage Time to System Failure Assume that the failure time of the main unit has a general distribution FðtÞ with mean l and its repair time has a general distribution GðtÞ with mean b. That is, the main unit forms an alternating renewal process that repeats operating and failed states. It is also assumed that the failure time of each operating spare unit has a general distribution FS ðtÞ with mean lS , even if it has been used before, i.e., the life of spare units that have already been in service is not affected by the past. All random variables considered here are independent. Unless otherwise stated, all units are good at time 0, and any failures are instantaneously detected and repaired or scrapped. Each switchover is perfect and each switchover time is instantaneous. Using the modified methods of an alternating renewal process in Sect. 3.4, we obtain the reliability properties of the system. First, suppose that n spare units are provided at time 0. Let Hk ðtÞ (k ¼ 1; 2; . . .; nÞ denote the first-passage distribution
212
8 Redundant Systems
to system failure when k spare units are provided at time 0. Then, we have the following renewal equation: Z t ðnÞ GðuÞdFS ðuÞ Hn ðtÞ ¼ FðtÞ 0 ) Z th n1 i X ðkÞ ðkþ1Þ þ Hnk ðtÞ FS ðuÞ FS ðuÞ dGðuÞ ðn ¼ 1; 2; . . .Þ; 0
k¼0
ð8:34Þ ðkÞ
where the asterisk mark represents the convolution, and FS ðtÞ (k ¼ 1; 2; . . .) ð0Þ
represents the k-fold Stieltjes convolution of FS ðtÞ with itself and FS ðtÞ 1 for t 0. The first term of the bracket on the right-hand side in (8.34) is the distribution of the time that all of n spare units fail before the first repair completion of the failed main unit. The second term is the distribution of the time that k (k ¼ 0; 1; . . .; n 1) spare units fail exactly until the first repair completion of the main unit, and then, the system with n k spare units operates again. The first-passage distribution Hn ðtÞ can be calculated recursively and uniquely determined from (8.34). Introduce the notation of the generating function of the LS transform: Z 1 1 X n z est dHn ðtÞ H ðz; sÞ n¼1
0
for jzj 1. Thus, taking the LS transform of (8.34) and forming the generating function H ðz; sÞ, R P ðnÞ n 1 st F ðsÞ 1 GðtÞdFS ðtÞ n¼1 z 0 e h i : ð8:35Þ H ðz; sÞ ¼ R P n 1 st F ðnÞ ðtÞ F ðnþ1Þ ðtÞ dGðtÞ 1 F ðsÞ 1 S S n¼0 z 0 e Moreover, let ln denote the mean first-passage time to system failure. Then, the renewal equation for ln is easily given, from (8.34), ln ¼ l þ
Z
1h 0
ðnÞ 1FS ðtÞ
i
GðtÞdt þ
n1 X k¼0
Z lnk 0
1h
ðkÞ
ðkþ1Þ
FS ðtÞFS
i ðtÞ dGðtÞ
ðn ¼ 1; 2; . . .Þ;
ð8:36Þ
and hence, its generating function is i R 1h P ðnÞ lz=ð1 zÞ þ 1 zn 0 1 FS ðtÞ GðtÞdt n¼1 i l ðzÞ z n ln ¼ : R h P n 1 F ðnÞ ðtÞ F ðnþ1Þ ðtÞ dGðtÞ 1 1 z n¼1 S S n¼0 0 1 X
ð8:37Þ
8.3 Standby System with Spare Units
213
8.3.2 Expected Number of Failed Units We derive the expected number of failed spare units in ½0; t. Let Mn ðtÞ (n ¼ 0; 1; . . .) be the that the total number of failed spare units in ½0; t Pprobability 1 are n, and MðtÞ n¼1 nMn ðtÞ. Then, in a similar way of obtaining (8.34), M0 ðtÞ ¼ FðtÞ þ FðtÞ F S ðtÞGðtÞ þ M0 ðtÞ ( h i ðnÞ ðnþ1Þ ðtÞ GðtÞ Mn ðtÞ ¼ FðtÞ FS ðtÞ FS þ
n X
Mnk ðtÞ
k¼0
Z th 0
ðkÞ FS ðuÞ
Z
t
F S ðuÞdGðuÞ ;
0
ðkþ1Þ FS ðuÞ
i
) dGðuÞ
ðn ¼ 1; 2; . . .Þ: ð8:38Þ
Introduce the notation M ðz; sÞ
1 X
zn
n¼0
Z
1
est dMn ðtÞ:
0
Then, from (8.38),
nh i o R1 P ðnÞ ðnþ1Þ 1 F ðsÞ 1 1 zn 0 est d FS ðtÞ FS ðtÞ GðtÞ n¼0 h i : M ðz; sÞ ¼ R P n 1 st F ðnÞ ðtÞ F ðnþ1Þ ðtÞ dGðtÞ 1 F ðsÞ 1 S S n¼0 z 0 e ð8:39Þ It can be easily shown that limz!1 M ðz; sÞ ¼ 1. The LS transform of the expected number of failed spare units in ½0; t is Z 1 1 X dM ðz; sÞ M ðsÞ n est dMn ðtÞ ¼ lim z!1 dz 0 n¼1 R 1 F ðsÞ 0 est GðtÞdMS ðtÞ ; ð8:40Þ ¼ 1 F ðsÞG ðsÞ P ðnÞ where note that MS ðtÞ 1 n¼1 FS ðtÞ is the renewal function for spare units. Furthermore, the expected number of failed spare units per unit of time in the steady-state is MðtÞ M lim ¼ t!1 t
R1 0
MS ðtÞdGðtÞ : lþb
ð8:41Þ
214
8 Redundant Systems
Equation (8.41) can be intuitively explained because the numerator represents the total expected number of failed spare units during the repair time of the main unit, and the denominator represents the mean time from the beginning of the operating main unit to its repair completion. Further, we obtain the expected number mn of failed main units until system failure for the system with n spare units. Because Z
1
0
n1 Z X
ðnÞ
GðtÞdFS ðtÞ þ
k¼0
1h
ðkÞ
ðkþ1Þ
i ðtÞ dGðtÞ ¼ 1;
ðkþ1Þ
i ðtÞ dGðtÞ:
FS ðtÞ FS
0
and by the similar method of obtaining (8.36), mn ¼ 1 þ
n1 X
Z
1h
mnk
ðkÞ
FS ðtÞ FS
0
k¼0
Thus, the generating function of mn is m ðzÞ
1 X n¼1
z n mn ¼
1
z=ð1 zÞ i : R 1 h ðnÞ ðnþ1Þ n ðtÞ dGðtÞ n¼0 z 0 FS ðtÞ FS
P1
ð8:42Þ
Example 8.1 (Exponential repair time) Suppose that GðtÞ ¼ 1 eht . Then, from (8.37), the mean time to system failure, given that n pare units are provided at time 0, is (Problem 8.4) 1 1 FS ðhÞ ln ¼ l þ n l þ : h FS ðhÞ Note that, by adding one spare unit to the system, the mean time to system increases by constant A, independent of the number of the former spare units, where A ðl þ 1=hÞ½1 FS ðhÞ=FS ðhÞ. Furthermore, the LS transform of the expected number of failed spare units in ½0; t is, from (8.40), F ðsÞFS ðs þ hÞ ; M ðsÞ ¼ 1 FS ðs þ hÞ f1 ½h=ðs þ hÞF ðsÞg and its limit per unit of time is, from (8.41) (Problem 8.4), M¼
1 FS ðhÞ 1 ¼ : l þ 1=h 1 FS ðhÞ A
In a similar way, the expected number of failed main units until system failure is, from (8.42) (Problem 8.4), mn ¼ 1 þ n
1 FS ðhÞ ; FS ðhÞ
8.3 Standby System with Spare Units
215
where note that mn ¼
ln þ 1=h : l þ 1=h
8.3.3 Expected Cost and Optimization Problems We derive the expected cost rate by introducing costs incurred for each failed main unit and failed spare units when infinite spare units are provided at time 0. This expected cost is easily deduced from the expected number of failed units. Then, we compare the expected cost rates for the system with both main unit and spare units, and for the system with only spare units, and determine which of them is more economical. Cost c1 is incurred for each failed main unit that is repaired, and cost cS is incurred for each failed spare unit that is replaced and scrapped. Letting CðtÞ be total expected cost incurred in ½0; t, CðtÞ c1 MR ðtÞ þ cS MðtÞ; where MR ðtÞ represents the expected number of repairs for the failed main unit, and recall that MðtÞ is the expected number of failed spare units in ½0; t. It is easily seen in Sect. 3.4 that MR lim
t!1
MR ðtÞ 1 ¼ : t lþb
Thus, the expected cost rate is, from (3.19) and (8.41), CðtÞ ¼ c1 MR þ cS M tR 1 c1 þ cS 0 MS ðtÞdGðtÞ ; ¼ lþb
C lim
t!1
ð8:43Þ
that is also equal to the total expected cost of one cycle from the beginning of the operating main unit to its repair completion. On the other hand, if only a spare unit is allowed to operate, then the expected cost rate is CS ¼
cS : ls
Therefore, comparing (8.43) and (8.44), C ðÞCS if and only if Z 1 lþb c1 ðÞcS MS ðtÞdGðtÞ : ls 0
ð8:44Þ
ð8:45Þ
216
8 Redundant Systems
In particular, when GðtÞ ¼ 1 eht , (8.45) becomes l þ 1=h FS ðhÞ : c1 ðÞcS ls 1 FS ðhÞ
ð8:46Þ
In this case, if the mean time ln to system failure is required to be greater than DT , then we can provide a minimum number of spare units that satisfies n
DT l 1 FS ðhÞ : l þ 1=h FS ðhÞ
If the expected cost until the mean time to system failure that is given by 1 FS ðhÞ cS n þ c1 mn ¼ c1 þ n cS þ c1 FS ðhÞ is less than DC , we can provide a maximum number of spare units that satisfies n
cS þ c1
DC c1
1 FS ðhÞ =FS ðhÞ
(Problem 8.5).
8.4 Problems 8 8.1 In (8.12), derive an optimum repair limit time T that minimizes CðTÞ when cr ðtÞ ¼ at for a [ 0 [3, p. 52]. 8.2 Derive the transition probabilities of a two-unit standby system with standby failure, i.e., when the unit in standby has the failure rate kS , and the other assumptions are the same ones in Sect. 8.2.1 [4]. 8.3 Consider the same two-unit standby system in Sect. 8.2.2 where each unit has a failure distribution FðtÞ with mean l and a failed unit has a repair distribution G1 ðtÞ with mean b1 . We adopt the following PM policy for an operating unit. When an operating unit works for a specified time T without failure, we stop its operation for PM. It is assumed that the time to PM completion has a general distribution G2 ðtÞ with mean b2 , and any unit becomes as good as new upon repair or PM completion. Furthermore, we make the following two assumptions. (a) PM of an operating unit is done only if the other unit is in standby. (b) An operating unit that has forfeited PM because of (1) undergoes PM just upon repair or PM completion of the other unit. Under the above assumptions, obtain the availability and derive an optimum PM time that maximizes it [3, p. 144].
8.4 Problems 8
217
8.4 In Example 8.1, derive ln , mn and M. 8.5 Consider the PM policy in Sect. 8.3 where the operating main unit undergoes PM at time T (0\T 1) after its installation or is repaired at failure, whichever occurs first. Spare units are infinite and work temporarily during the interval of the repair or PM time of the main unit. It is assumed that the PM time has the same distribution GðtÞ as the repair time. The main unit is as good as new upon repair or PM and begins to operate immediately. The costs incurred for each failed main unit and each spare unit are c1 and cs , respectively. The PM cost c2 with c2 \c1 is incurred for each non-failed main unit. Then, obtain the expected cost rate and derive an optimum PM time T that minimizes it [5].
References 1. 2. 3. 4.
Birolini A (1999) Reliability engineering theory and practice. Springer, New York Brown M, Proschan F (1983) Imperfect repair. J Appl Prob 20:851–859 Nakagawa T (2005) Maintenance theory of reliability. Springer, London Nakagawa T (2002) Two-unit redundant models. In: Osaki S (ed) Stochastic models in reliability and maintenance. Springer, Berlin, pp 165–185 5. Nakagawa T, Osaki S (1976) Reliability analysis of a one-unit system with unrepairable spare units and it optimization applications. Oper Res Q 27:101–110
Appendix A: Laplace Transform
This appendix presents some results of Laplace and Laplace–Stieltjes (LS) transforms and give useful tools used in reliability theory [1–4].
A.1 Laplace and Laplace–Stieltjes Transforms Let f ðtÞ be a function of t specified for t 0 and s be a complex number. Then, the Laplace transform of f ðtÞ is defined by
Lff ðtÞg f ðsÞ ¼
Z1
est f ðtÞdt;
ðA:1Þ
0
that is a function of the complex number s. Although s is defined as the complex number, we do not need the knowledge of complex analysis and may use the Laplace transform as if it is a real number in reliability theory. If the above integral (A.1) converges for ReðsÞ [ s0 , then it always exists for ReðsÞ [ s0 . For example, suppose that f ðtÞ is a function of t 0 such that jf ðtÞj Meat for some M and a. Then, f ðsÞ exists for all ReðsÞ [ a. The function f ðtÞ is called the inverse Laplace transform of f ðsÞ and is written f ðtÞ ¼ L1 ff ðsÞg:
ðA:2Þ
When f ðtÞ is a probability density of a nonnegative random variable X, sX
Efe
g¼
Z1
est f ðtÞdt ¼ f ðsÞ:
ðA:3Þ
0
Similarly, let FðtÞ be a function of t 0: The Laplace–Stieltjes (LS) transform of FðtÞ is defined by
219
220
Appendix A: Laplace Transform
F ðsÞ
Z1
est dFðtÞ
ðA:4Þ
0
for ReðsÞ [ s0 . When FðtÞ is a probability distribution of a random variable X, the LS transform of FðtÞ is EfesX g ¼
Z1
est dFðtÞ ¼ F ðsÞ;
ðA:5Þ
0
that agrees with the Laplace transform f ðsÞ in (A.3). From (A.5), the nth moment of a random variable X is dF ðsÞ 1 F ðsÞ EfXg ¼ ¼ lim ; ds s¼0 s!0 s dn F ðsÞ ðn ¼ 1; 2; . . .Þ; EfX n g ¼ dsn s¼0 VfXg ¼ EfX 2 g ðEfXgÞ2 : There is no theoretical difference greatly between Laplace and LS transforms, because
F ðsÞ ¼
Z1 e 0
st
dFðtÞ ¼ s
Z1
est FðtÞdt:
ðA:6Þ
0
In other words, the LS transform can be obtained by s times the Laplace transform. It would be more convenient to use LS transforms because we deal mainly with distribution forms more than density ones in reliability theory. However, we sometimes use the Laplace transform for some functions in p. 61, p. 77, p. 78, p.184 and p. 196.
A.2 Properties We always use LS transforms, except for some special functions through this book and summarize their brief properties. LS transforms play significant, important roles in making integral and differential calculuses, and in calculating distributions of the sum of random variables. Let F ðsÞ be the LS transform of FðtÞ in (A.5).
Appendix A: Laplace Transform
1. The LS transform of
221
Rt
FðuÞdu is 3 Zt Z1 Z1 1 st 4 5 e d FðuÞdu ¼ est FðtÞdt ¼ F ðsÞ: s 0
2
0
0
ðA:7Þ
0
2. The LS transform of F 0 ðtÞ is Z1
est dF 0 ðtÞ ¼ s
0
Z1
est F 0 ðtÞdt ¼ s½F ðsÞ Fð0Þ:
ðA:8Þ
0
Rt 3. If the function HðtÞ is given by the convolution HðtÞ ¼ 0 Gðt uÞdFðuÞ, and the LS transforms of GðtÞ and FðtÞ are G ðsÞ and F ðsÞ, respectively, the LS transform of HðtÞ is 2 t 3 Z1 Z est d4 Gðt uÞdFðuÞ5 H ðsÞ ¼ 0
¼
Z1
0 st
e
Z1 dGðtÞ
0
est dFðtÞ
ðA:9Þ
0
¼ G ðsÞF ðsÞ: This is easily obtained by another method: If two random variables X and Y are independent and have distributions GðtÞ and FðtÞ, respectively, the LS transforms of distribution of Z X þ Y is EfesZ g ¼ EfesðXþYÞ g ¼ EfesX gEfesY g ¼ G ðsÞF ðsÞ: Table A.1 presents some properties of LS transforms and Table A.2 presents LS transforms of some functions [1, 4]
A.3 Abelian and Tauberian Theorems We show two theorems that are very useful for performing limiting operations with transforms: Theorem A.1 (An Abelian Theorem) If lim
t!1
FðtÞ C for a 0; ¼ ta Cða þ 1Þ
ðA:10Þ
222
Appendix A: Laplace Transform
Table A.1 Properties of LS transforms
Table A.2 LS transforms of functions
FðtÞ
F ðsÞ
F1 ðtÞ þ F2 ðtÞ aFðtÞ Fðt aÞða [ 0Þ FðatÞ eat FðtÞ ða [ 0Þ F 0 ðtÞ ¼ dFðtÞ=dt tF 0 ðtÞ Rt FðuÞdu R0t 0 F2 ðt uÞd F1 ðuÞ limt!0 FðtÞ limt!1 FðtÞ
F1 ðsÞ þ F2 ðsÞ aF ðsÞ esa F ðsÞ F ðs=aÞ ½s=ðs þ aÞF ðs þ aÞ s½F ðsÞ Fð0Þ sdF ðsÞ=ds F ðsÞ=s F1 ðsÞF2 ðsÞ lims!1 F ðsÞ lims!0 F ðsÞ
FðtÞ
F ðsÞ
1 t t n ðn ¼ 1; 2; . . .Þ ta ða [ 1Þ eat ða [ 0Þ teat ða [ 0Þ tn eat ða [ 0Þ
1 1=s n!=sn Cða þ 1Þ=sa s=ðs þ aÞ s=ðs þ aÞ2 sn!=ðs þ aÞnþ1
tb eat ða [ 0; b [ 1Þ cos at sin at log t
sCðb þ 1Þ=ðs þ aÞbþ1 s2 =ðs2 þ a2 Þ sa=ðs2 þ a2 Þ C log sy
y C Euler’s constant and C = 0.577215…
then lim sa F ðsÞ ¼ C; s!0
where CðaÞ
R1 0
ðA:11Þ
ta1 et dt is a gamma function defined in (2.13).
TheoremR A.2 (A Tauberian Theorem) If FðtÞ increases and its LS transform 1 F ðsÞ ¼ 0 est dFðtÞ exists for ReðsÞ [ 0, and if lim sa F ðsÞ ¼ C s!0
for a 0;
ðA:12Þ
then lim
t!1
FðtÞ C ¼ : ta Cða þ 1Þ
ðA:13Þ
The readers refer to Example 4.11, Problems 4.5, and Sect. 8.2.1 as applications to differential equations of LS transforms.
Appendix A: Laplace Transform
223
In general, it is very difficult to making the inversion of LS transforms. However, with rapid progress of personal computers and a significant development of Laplace inversion methods, it is not so difficult to compute numerically some probabilities and other quantities in reliability models. Tijms [2] summarized some useful methods of Laplace inversion (see Sect. 3.2).
A.4 Generating Function LS transforms are used for continuous functions, while generating functions [2, p. 449, 5, p. 446] are used for discrete probability distributions and functions with variables n (n ¼ 0; 1; 2; . . .; or n ¼ 1; 2; . . .Þ as shown in Sects. 3.5.2, 4.1.2 and 8.3. Let an (n ¼ 0; 1; 2; . . .) be a sequence of real number. If a ðzÞ
1 X
an z n
ðA:14Þ
n¼0
exists for jzj 1, then a ðzÞ is called the generating function of the sequence fan g1 n¼0 . The discrete function can be recovered from a0 ¼ a ð0Þ and 1 dn a ðzÞ an ¼ n! dzn z¼0
ðn ¼ 1; 2; . . .Þ:
ðA:15Þ
When pn (n ¼ 0; 1; 2) is a probability function of random variable X, 1 X dp ðzÞ EfXg ¼ npn ¼ ; dz z¼1 n¼1 1 X d2 p ðzÞ nðn 1Þpn ¼ ; EfXðX 1Þg ¼ dz2 z¼1 n¼1 VfXg ¼ EfX 2 g ðEfXgÞ2 : For example, when a probability distribution is geometric distribution, i.e., pn ¼ pqn1 (n ¼ 1; 2; . . .), the generating function is p ðzÞ ¼
1 X
zn pqn1 ¼
n¼1
pz ; 1 qz
and dp ðzÞ p ¼ 2 dz z¼1 ð1 qzÞ
z¼1
1 ¼ ; p
d2 p ðzÞ 2pq ¼ dz2 z¼1 ð1 qzÞ3
z¼1
¼
2q : p2
224
Appendix A: Laplace Transform
Thus, the mean and variance of a geometric distribution are 1 2q 1 1 q EfXg ¼ ; VfXg ¼ 2 þ 2 ¼ 2 : p p p p p P The sequence cn ¼ nk¼0 ak bnk (n ¼ 0; 1; 2; . . .) where the generating functions of fan g and fbn g are a ðzÞ and b ðzÞ for jzj 1, respectively, the generating function of fcn g is c ðzÞ
1 X n¼0
zn
n X
ak bnk ¼ a ðzÞb ðzÞ
k¼0
for jzj 1, that corresponds to the property of (A.9) in continuous case. For example, when both probability functions are Poisson distributions, i.e., an ¼ ½ðk1 tÞn =n!ek1 t and bn ¼ ½ðk2 tÞn =n!ek2 t ðn ¼ 0; 1; 2; . . .Þ, their generating functions are a ðzÞ ek1 tð1zÞ ;
b ðzÞ ek2 tð1zÞ ;
and hence, c ðzÞ ¼ exp½ðk1 þ k2 Þtð1 zÞ; that shows the properties of superposition and decomposition of a Poisson process as shown in Examples 2.8 and 2.9, respectively. Abelian and Tauberian theorems are rewritten in discrete case [5 p. 447, 6, p. 418]: Theorem A.3 (An Abelian Theorem) If limn!1 an ¼ a, then limð1 zÞ z!1
1 X
an zn ¼ a:
n¼0
Theorem A.4 (A Tauberian Theorem) If limz!1 ð1 zÞ limn!1 nðan an1 Þ ¼ 0, then
P1
n¼0
an zn ¼ a and
lim an ¼ a:
n!1
Theorem A.5 Let fan g1 n¼0 be a nonnegative sequence of real numbers with generating function in (A.14). For constant a and b, 1 X
an ¼ a
n¼0
lim
m!1
m 1X an m n¼0
,
lim a ðzÞ ¼ a;
,
limð1 zÞa ðzÞ ¼ c:
z!1
! ¼c
z!1
Extending such a generating function, the universal generating function is introduced and is applied to reliability analysis of multi-state systems [7–9].
Appendix B: Answers to Selected Problems
Chapter 2 2.1 From Sn has a gamma distribution with order n in (2.12) [10, p. 84,], EfgðSn Þg ¼
Z1
gðtÞd PrfSn tg ¼
0
Z1 gðtÞ
kðktÞn1 kt e dt: ðn 1Þ!
0
Hence, E
( 1 X
) gðSn Þ
1 Z X
1
¼
n¼1
n¼1
kðktÞn1 kt gðtÞ e dt ¼ k ðn 1Þ!
0
Z1 gðtÞdt: 0
2.2 See [11, p. 24]. 2.3 EfXð1Þ g ¼
Z1
enkt dt ¼
1 ; nk
0
EfXðnÞ g ¼
Z1
0
n 1X 1 : 1 ð1 enkt Þ dt ¼ k k¼1 k a
2.5 Compute numerically when FðtÞ ¼ 1 et [12], MTTF ¼
Z1
a 1 ð1 et Þn dt;
0
and compare it with
Pn
k¼1
1=k
1=a
and ðC þ log nÞ1=a .
225
226
Appendix B: Answers to Selected Problems
2.6
From [13, p. 9], an optimum number n (1 n \1) is given by a finite and unique minimum that satisfies nþ1 X nþ1
kþ1 k¼1 2.7
From [13, p. 12], MTTF ¼
n Z X n j¼k
j
c2 c1
1
ðn ¼ 1; 2; . . .Þ:
ejkt ð1 ekt Þnj dt ¼
0
n 1X 1 : k j¼k j
2.8 From [11, p. 286], let Ik ¼ 1 if there is a record at time t, Ik ¼ 0 otherwise. Then, because PrfIk ¼ 1g ¼ ðk 1Þ!=k! ¼ 1=k; n n n X X X 1 Ik ; EfNn g ¼ EfIk g ¼ Nn ¼ ; k k¼1 k¼1 k¼1 n X 1
1 VarfIk g ¼ 1 : VarfNn g ¼ k k k¼1 k¼1 n X
When n ¼ 10; 100; 1000, respectively, n X 1 k¼1
k
¼ 2:92897;
5:18738;
7:48547:
2.9 See [1, p. 40]. 2.11 The number of spare units is given a minimum number that satisfies n X ðk1 þ k2 Þk k¼0
k!
eðk1 þk2 Þ a:
2.12 Compute the probabilities, for p ¼ 0:90; 0:10 and kt ¼ 1; 2; 3; ðpktÞk pkt e k!
ðk ¼ 0; 1; 2; 3Þ:
2.13 From the inequality CðN þ 1Þ CðNÞ 0; B NðN þ 1Þ ; A
i.e.,
N X k¼1
k
B : 2A
2.14 Consider the multiple backward times Tk (k ¼ 1; 2; . . .; N), where Sk Pk j¼1 Tj , S0 0 and SN t [14]. If we cannot find the actual failure time until time Sk1 from time t by the backward operation, we execute again the next backward time with Tk . If we can find the failure time, i.e., the system fails in ðt Sk ; t Sk1 , then the process ends. Finally, if we cannot find the failure time until SN1 , we make the final backward with TN ¼ t SN1 , and can find it certainly.
Appendix B: Answers to Selected Problems
227
It is assumed that c1 þ c2 Tk is the cost for the kth backward. When the failure time is detected at the kth backward, the expected cost is 1 ðc1 k þ c2 Sk Þ½Fðt Sk1 Þ Fðt Sk Þ; FðtÞ and hence, the total expected cost is CðT1 ; T2 ; . . .; TN Þ ¼
N1 1 X ðc1 k þ c2 Sk Þ½Fðt Sk1 Þ Fðt Sk Þ: FðtÞ k¼1
We find optimum times Tk ðk ¼ 1; 2; . . .; NÞ that minimize CðT1 ; T2 ; . . .; TN Þ for a given t [ 0: Differentiating CðT1 ; T2 ; . . .; TN Þ with respect to Sk and setting it equal to zero, Skþ1 Sk ¼
Fðt Sk1 Þ Fðt Sk Þ c1 c2 f ðt Sk Þ
ðk ¼ 1; 2; . . .; N 1Þ:
Setting that xk t Sk ; xk xkþ1 ¼
Fðxk1 Þ Fðxk Þ c1 c2 f ðxk Þ
ðk ¼ 1; 2; . . .; N 1Þ;
ðB:1Þ
where xN 0 and x0 t. Then, the total expected cost is Cðx1 ; x2 ; . . .; xN Þ ¼
N 1 X ½c1 k þ c2 ðt xk Þ½Fðxk1 Þ Fðxk Þ: FðtÞ k¼1
ðB:2Þ
Therefore, we solve the simultaneous equations in (B.1) for given N 1 and compute the total expected cost in (B.2). Finally, comparing the total expected cost for N 1, we can get an optimum number N and backward times Sk ¼ t xk (k ¼ 1; 2; . . .; N ). 2.15 t k1 Z j k1 X X u u nj kt n n j!ðnjÞ! EfSk jNðtÞ ¼ ng ¼ 1 du ¼ t ¼ ; j j ðnþ1Þ! nþ1 t t j¼0 j¼0 0
EfSk Sk1 jNðtÞ ¼ ng ¼
kt ðk 1Þt t ¼ : nþ1 nþ1 nþ1
Compute when k ¼ 0; 1; 2; PrfNð10Þ ¼ kjNð100Þ ¼ 10g ¼ PrfS3 50jNð100Þ ¼ 10g ¼
10
k 10 X 10 k¼3
k
10 100
k
50 100
1
10 100
10k ;
k 50 10k 1 ¼ 0:945: 100
228
Appendix B: Answers to Selected Problems
2.16 Differentiate ½HðtÞn =n! with t and integrate it from 0 to t, where note that Hð0Þ 0. R1 2.17 Noting that l ¼ 0 P0 ðtÞdt; 1 PN1 R 1 N 1 Z X 1 k¼1 0 Pk ðtÞdt R1 ½Pk ðtÞ PN ðtÞdt [ 0; ðN 1Þ ¼ R 1 0 PN ðtÞdt 0 PN ðtÞdt k¼1 0
R1
because when hðtÞ increases strictly, 0 Pk ðtÞdt decreases strictly from (2.44) for N 2. 2.19 Setting that KðtÞ 1 HðtÞ=a, solve the differential equation by using the separation of variables method. 2.20 From (2.57), d exp½kts=ðs þ hÞ kt EfZðtÞg ¼ ¼ h; ds s¼0 2 d exp½kts=ðs þ hÞ kt 2kt VfZðtÞg ¼ ¼ 2; 2 ds h h s¼0
and from (2.60), 1 EfYg ¼ ð1 þ hKÞ; k 2 1 1 X 2X ðhKÞj hK 1 2hK þ 1 ðk þ 1Þ ð1 þ hKÞ ¼ : VfYg ¼ 2 e j! k k k¼0 k2 j¼k
Chapter 3 3.2
Considering that the inverse LS transform of 1=½ðs þ aÞ2 þ b2 is given by " pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi # a2 þ b2 at 1 1 b e sin bt þ tan 1 ; b a2 þ b2 a derive when k ¼ 3; 4; respectively, M ðsÞ ¼ M ðsÞ ¼
k3 2
2
s½ðs þ 3k=2Þ þ 3k =4
¼
k4 sðs þ 2kÞ½ðs þ kÞ2 þ k2
k ks þ 3k2 ; 3s 3½ðs þ 3k=2Þ2 þ 3k2 =4
¼
k k k2 : 4s 4ðs þ 2kÞ 2½ðs þ kÞ2 þ k2
3.3 Use the LS transform of a gamma distribution in (2.14).
Appendix B: Answers to Selected Problems
229 a
Table B.1 MðtÞ and t=EfXg when FðtÞ ¼ 1 eðktÞ and k ¼ 1 t a
1 2 3 4 5 6 7 8 9 10 10/E{X}
3.4 3.5
1.0
1.5
2.0
2.5
3.0
1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 9.00 10.00 10.00
0.84 1.95 3.05 4.16 5.27 6.38 7.49 8.59 9.70 10.81 11.08
0.75 1.89 3.02 4.15 5.28 6.41 7.54 8.86 9.79 10.92 11.28
0.70 1.84 2.97 4.10 5.23 6.35 7.48 8.61 9.74 10.86 11.27
0.67 1.80 2.93 4.05 5.17 6.29 7.41 8.53 9.65 10.77 11.20
Table B.1 is computed. From Example 2.4, FðtÞ ¼ ð1 ekt Þ2 , and hence, Z1 2k2 ; est dFðtÞ ¼ F ðsÞ ¼ ðs þ kÞðs þ 2kÞ 0
M ðsÞ ¼ 3.7
F ðsÞ 2k2 ¼ : 1 F ðsÞ s2 þ 3ks
Use the inverse LS transform in Table A.2. pffiffiffi n 10 þ 1:64 2 ¼ 12:32 for a ¼ 0:95; pffiffiffi n 10 þ 2:33 2 ¼ 13:29 for a ¼ 0:99:
3.8
From (3.25), 2 3 Z1 Zt 4FðxÞ þ Fðx uÞdMðuÞ5dx t
0
¼
Z1
FðxÞdx
0
¼ltþ
Zt
2 3 Z1 Z t FðxÞdx þ 4 Fðx uÞdMðuÞ5dx
2 t 0 3 Z t Z1 Ztu FðxÞdx þ 4 FðxÞdx FðxÞdx5dMðuÞ 0
Zt 0
¼ l t þ lMðtÞ þ
0
Zt 0
¼ l½1 þ MðtÞ t:
0
FðxÞdx
0
Zt 0
Mðt xÞFðxÞdx
230
Appendix B: Answers to Selected Problems
From (3.26), 1 l
Z1
1 tFðtÞdt ¼ 2l
0
Z1
1 FðtÞdt ¼ 2l 2
0
Z1
t2 dFðtÞ ¼
l2 þ r2 : 2l
0
3.9 From [3, p. 128] and Fig. B.1, PrfbðtÞ xg ¼ PrfXNðtÞþ1 xg ¼
Zt
½FðxÞ Fðt uÞdMðuÞ:
tx
Thus, by Theorem 3.6, 1 lim PrfbðtÞ xg ¼ t!1 l
Zx
1 ½FðxÞ FðuÞdu ¼ l
0
Zx udFðuÞ: 0
3.10 Let NA ðtÞ and NB ðtÞ be the number of replacements of units for the age and block replacements with planned replacement time T in ½0; t of Sect. 3.3.2. Then, from [16, p. 67], EfNB ðtÞg EfNA ðtÞg; lim
t!1
EfNB ðtÞg MðTÞ þ 1 EfNA ðtÞg 1 ¼ lim ¼ RT : t!1 t T t FðtÞdt 0
Thus, t 1: 0 FðuÞdu
MðtÞ R t
e A ðtÞ and N e B ðtÞ be the number of failures of units for the age and Next, let N block replacements in ½0; t. Then, from [16, p. 69], when FðtÞ is IFR, e A ðtÞg; e B ðtÞg Ef N Ef N e B ðtÞg MðTÞ e A ðtÞg Ef N Ef N FðTÞ lim ¼ lim ¼ RT : t!1 t!1 t T t FðtÞdt 0
Fig. B.1 Process of the life in a renewal process
Appendix B: Answers to Selected Problems
231
Thus, tFðtÞ : 0 FðuÞdu
MðtÞ R t 3.12 1 lim T!0 FðTÞ
ZT
1 ; FðtÞdt ¼ hð0Þ
1 lim T!1 FðTÞ
0
ZT
FðtÞdt ¼ l;
0
where if hðtÞ ! 0 as t ! 0 then 1=hð0Þ is assumed to be infinity. In addition, when FðtÞ is IFR, 0 @ 1 FðTÞ
ZT
10 FðtÞdtA ¼
0
1 FðTÞ½FðTÞ2
" RT 0
FðtÞdt
FðTÞ
RT 0
FðtÞdt
# hðTÞ 0:
Similarly, 1 lim T!0 FðTÞ
ZT
FðtÞdt ¼ 0;
1 lim ¼ T!1 FðTÞ
0
Clearly,
RT 0
ZT
FðtÞdt ¼ 1:
0
FðtÞdt=FðTÞ increases even when FðtÞ is not IFR.
3.13 AðTÞ þ
ZT
BðT tÞdFðtÞ
0
¼ c1 FðTÞ þ c2 FðTÞ þ
ZT
½c1 MðT tÞ þ c2 dFðtÞ
0
¼ c1 MðTÞ þ c2 ; AðTÞ þ
ZT
AðT tÞdMðtÞ
0
¼ c1 FðTÞ þ c2 FðTÞ þ
ZT
½c1 FðT tÞ þ c2 FðT tÞdMðtÞ
0
¼ c1 MðTÞ þ c2 : 3.14 When T0 has an exponential distribution ð1 eat Þ, from (3.67),
232
Appendix B: Answers to Selected Problems
RðT; x; aÞ ¼
R Tþx
eat FðtÞdt : RT 1 G ðaÞ þ aG ðaÞ 0 eat FðtÞdt aeax
x
Clearly, Rð0; x; aÞ ¼ 0;
R1 aeax x eat FðtÞdt Rð1; x; aÞ ¼ : 1 G ðaÞF ðaÞ
Differentiating RðT; x; aÞ with respect to T and setting it equal to zero, 2 3 ZT kðT; xÞ41 G ðaÞ þ aG ðaÞ eat FðtÞdt5 0
aG ðaÞ
ZT
eat ½FðtÞ Fðt þ xÞdt ¼ 1 G ðaÞ;
ðB:3Þ
0
where kðt; xÞ
Fðt þ xÞ FðtÞ : FðtÞ
Note that kðt; xÞ and hðtÞ have the same property. Letting LðTÞ be the lefthand side of (B.3), Lð0Þ ¼ FðxÞ½1 G ðaÞ 1 G ðaÞ; Lð1Þ ¼ ½kð1; xÞ 1½1 G ðaÞF ðaÞ Z1 ax þ 1 G ðaÞ þ aG ðaÞe eat FðtÞdt: x
If the failure rate increase strictly, then 2 L0 ðTÞ ¼ k0 ðT; xÞ41 G ðaÞ þ aG ðaÞ
ZT
3 eat FðtÞdt5 [ 0:
0
Thus, we have the following optimum policy: (i) If Lð1Þ [ 1 G ðaÞ then there exists a finite and unique T ð0\T \1Þ that satisfies (B.3), and the interval reliability is RðT ; x; aÞ ¼
FðT þ xÞ : G ðaÞFðT Þ
(ii) If Lð1Þ 1 G ðaÞ then T ¼ 1.
Appendix B: Answers to Selected Problems
233
3.16 From Example 3.5, when a ¼ 0:95 and 0:99, respectively, d 909:09 102:56 1:64 ¼ 740:89; d 909:09 102:56 2:33 ¼ 670:13: pffi 3.17 When GðtÞ ¼ 1 eh t , (3.84) is pffiffiffi
alpffiffiffiffi 1 hl T 1 eh T ¼ ; c1 h 2
ðB:4Þ
and when cr ðtÞ ¼ at2 and GðtÞ ¼ 1 eht , from (3.82), T
1 ehT c1 h ¼ : hðlh þ 1Þ 2a
ðB:5Þ
In both cases, there exist optimum T that satisfies (B.4) and (B.5). 3.18 From [17, p. 156], the LS transform of an exponential distribution Fk ðtÞ ¼ ð1 ekk t Þ is kk =ðs þ kk Þ, and hence, the LS transform of the nth convolution F ðnÞ ðtÞ of Fk ðtÞ is F
ðnÞ
ðsÞ ¼
Z1
est dF ðnÞ ðtÞ ¼
0
n Y kk : s þ kk k¼1
Inverting F ðnÞ ðsÞ, ðnÞ
F ðtÞ ¼ 1
n X k¼1
! kj ekk t : k k j k j¼1;j6¼k n Y
In particular, when kk ¼ ak1 k, ðnÞ
F ðtÞ ¼ 1
n X k¼1
! 1 expðak1 ktÞ: kj 1 a j¼1;j6¼k n Y
3.19 The mean time to replacement is N X l lða aNþ1 Þ ¼ : k1 a a1 k¼1
3.20 The mean operating and repair times until the unit is replaced at the Nth failure are, respectively, N X l ; k1 a k¼1
N 1 X b : k1 b k¼1
234
Appendix B: Answers to Selected Problems
3.21 pz p 1 1 ¼ ð1 zÞ½1 ðb pÞz p þ a 1 z 1 ðb pÞz 1 p X ¼ ½1 ðb pÞn zn ; p þ a n¼0 pz 1 bz p a p z ¼ þ 1 zð1 zÞ½1 ðb pÞz p þ a 1 z 1 ðb pÞz 1 z 1 1 X p X ½a þ pðb pÞn zn zk ¼ p þ a n¼0 k¼1 1 p X p½1 ðb pÞn n ¼ na þ z : p þ a n¼0 pþa 3.22 PrfXnþ1 xjSn ¼ tg ¼
Fðx þ atÞ FðatÞ : 1 FðatÞ
Chapter 4 4.1 P Pn ¼
1a b
a 1b
b a 1 ð1 a bÞn a þ aþb aþb b a b
¼ Pnþ1 : 1a b þ að1 a bÞn a½1 ð1 a bÞn pðnÞ P ¼ ; aþb aþb b
a b
a 1b
¼ pðn þ 1Þ: 4.2
P¼
1 2 .. . N1 N
(
1 r1 r2 .. .
2 1 r1 0 .. .
3 0 1 r2 .. .
... ... ...
N1 0 0 .. .
N 0 0 .. .
rN1 1
0 0
0 0
... ...
0 0
1 rN1 0
)
:
Appendix B: Answers to Selected Problems
From p ¼ pP, p1 ¼
N X
235
pj r j ;
j¼1
pjþ1 ¼ pj ð1 rj Þ
ðj ¼ 1; 2; . . .; N 1Þ:
Thus, Pj1 ð1 ri Þ i; pj ¼ P hi¼1 N j1 j¼1 Pi¼1 ð1 ri Þ where rN 1 and P0i¼1 1; 4.4 Let FðtÞ be the identical failure probability of each unit at time t. Then, the probability that k units have failed and n k units have not failed independently at time t is n ½FðtÞk ½1 FðtÞnk ðk ¼ 0; 1; 2; . . .; nÞ: k 4.5 For one repair person, the LS transforms of the differential equations are ðs þ kÞP00 ðsÞ ¼ s þ hP01 ðsÞ; ðs þ k þ hÞP01 ðsÞ ¼ kP00 ðsÞ þ hP02 ðsÞ; ðs þ hÞP02 ðsÞ ¼ kP01 ðsÞ: Solving these equations for P0i ðsÞ ði ¼ 0; 1; 2Þ and inverting them, h2 k x1 k h x1 t x2 k h x2 t ; þ e e P00 ðtÞ ¼ 2 x1 x2 k þ kh þ h2 x1 x2 kh k x1 þ h x1 t x2 þ h x2 t P01 ðtÞ ¼ 2 þ e e ; x1 x2 k þ kh þ h2 x1 x2 k2 k2 1 x1 t 1 x2 t þ e e P02 ðtÞ ¼ 2 ; x2 k þ kh þ h2 x1 x2 x1 where x1 k þ h þ
pffiffiffiffiffi kh;
x2 k þ h
pffiffiffiffiffi kh:
For two repair persons, 2h2 k x1 k 2h x1 t x2 k 2h x2 t ; þ e e x1 x2 k2 þ 2kh þ 2h2 x1 x2 2kh k x1 þ 2h x1 t x2 þ 2h x2 t ; þ e e P01 ðtÞ ¼ 2 2 x1 x2 x1 x2 k þ 2kh þ h k2 k2 1 x1 t 1 x2 t P02 ðtÞ ¼ 2 ; þ e e x2 k þ 2kh þ 2h2 x1 x2 x1 P00 ðtÞ ¼
236
Appendix B: Answers to Selected Problems
where
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 2k þ 3h þ 4kh þ h2 ; 2 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 x2 2k þ 3h 4kh þ h2 : 2
x1
Chapter 5 5.1 F0 ðsÞ ¼ F ðsÞ½aG1 ðsÞ þ ð1 aÞG2 ðsÞ; aF ðsÞ ð1 aÞF ðsÞ ; F ; ðsÞ ¼ 2 1 ð1 aÞF ðsÞG2 ðsÞ 1 aF ðsÞG1 ðsÞ Fj ðsÞ½1 Gj ðsÞ 1 F ðsÞ P0 ðsÞ ¼ ðsÞ ¼ ; P ðj ¼ 1; 2Þ; j 1 F0 ðsÞ 1 Fj ðsÞGj ðsÞ
F1 ðsÞ ¼
M0 ðsÞ ¼
F0 ðsÞ ; 1 F0 ðsÞ
Mj ðsÞ ¼
Fj ðsÞ 1 Fj ðsÞGj ðsÞ
ðj ¼ 1; 2Þ:
5.2 Let Qij be the probability that when the system is in State i ði ¼ 0; 1Þ; the data transmission has failed at j times (j ¼ 1; 2; . . .; N) consecutively. Then, using (5.15), we have the following equations related to Qij [13, p. 116]: Q0j ¼ P00 ðT1 Þp0 Q0j1 þ P01 ðT1 Þp1 Q1j1 þ P02 ðT1 Þpj2 ; Q1j ¼ P10 ðT1 Þp0 Q0j1 þ P11 ðT1 Þp1 Q1j1 þ P12 ðT1 Þpj2 : Forming the generating functions of Q0j and Q1j , solving them, and inverting their generating functions, 1 fx1 p1 ½P11 ðT1 Þ P01 ðT1 Þgx1j x1 x2 fx2 p1 ½P11 ðT1 Þ P01 ðT1 Þgx2j
Q0j ¼
þ
p2 x1 x2
fx1 P02 ðT1 Þ p1 ½P11 ðT1 ÞP02 ðT1 Þ P01 ðT1 ÞP12 ðT1 Þg
x1j p2j fx2 P02 ðT1 Þ p1 ½P11 ðT1 ÞP02 ðT1 Þ P01 ðT1 ÞP12 ðT1 Þg x1 p2 ! x2j p2j ðj ¼ 0; 1; 2; . . .; NÞ; x2 p2
Appendix B: Answers to Selected Problems
237
where
1 p0 P00 ðT1 Þ þ p1 P11 ðT1 Þ 2 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi þ ½p0 P00 ðT1 Þ p1 P11 ðT1 Þ2 þ 4p0 p1 P01 ðT1 ÞP10 ðT1 Þ ; 1 x2 p0 P00 ðT1 Þ þ p1 P11 ðT1 Þ 2 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ½p0 P00 ðT1 Þ p1 P11 ðT1 Þ2 þ 4p0 p1 P01 ðT1 ÞP10 ðT1 Þ :
x1
Thus, the mean time l3 ðNÞ is PN1 T1 j¼0 Q0j þ T2 Q0N l3 ðNÞ ¼ 1 Q0N
ðN ¼ 1; 2; . . .Þ:
From l3 ðN þ 1Þ l3 ðNÞ 0,
Q0N
N 1 X Q0N T2 ð1 Q0N Þ Q0j : Q0Nþ1 T1 j¼0
ðB:6Þ
Let LðNÞ be the left-hand side of (B.6). If Q0jþ1 =Q0j increases strictly, then LðNÞ increases strictly. Therefore, if Lð1Þ [ T2 =T1 , then there exists a finite and unique minimum N ð1 N \1Þ that satisfies (B.6). 5.3 fi;ij is the probability of first-passage from State i to State j without returning to State i, and fj;ii is the probability of returning from State i to State i without passing through State j, and hence, fi;ij þ fj;ii ¼ 1. 5.4 From (5.44) and (5.45), ðn þ 1Þðc0 c1 Þ þ c2 lCðnÞ ¼ þ c1 : Pnþ1 n þ 1 C j1 j¼1 j Thus, because nþ1 n X n nþ1 n 1 X 1 1X 1 0; n þ 1 j¼1 Cj1 n j¼1 j¼1 j Cj1 j c0 lim CðnÞ ¼ ; n!1 l CðnÞ decreases to c0 =l. 5.6 CðTÞ ¼ c1 M4 þ c2 M1 þ c3 ðP3 þ P4 Þ þ c4 P2 R TþL R1 c1 FðTÞ þ c2 FðTÞ þ c3 T FðtÞdt þ c4 TþL FðtÞdt : ¼ R TþL l þ T FðtÞdt
238
Appendix B: Answers to Selected Problems
Clearly, Cð0Þ ¼
c 2 þ c3
RL
R1 FðtÞdt þ c4 L FðtÞdt ; RL l þ 0 FðtÞdt
0
Cð1Þ ¼
c1 þ c3 L : lþL
Differentiating CðTÞ with respect to T and setting it equal to zero, 8 9 2 3 ZT < = LðTÞ kðT; LÞ c3 l þ c4 4L þ FðtÞdt5 c1 FðTÞ c2 FðTÞ : ; 0 2 3 ZTþL þ ½ðc1 c2 ÞhðTÞ c4 4l þ FðtÞdt5 ¼ 0;
ðB:7Þ
T
where hðtÞ f ðtÞ=FðtÞ and kðt; LÞ ½Fðt þ LÞ FðtÞ=FðtÞ. If hðtÞ increases strictly, then kðt; LÞ increases strictly with t, and hence, LðTÞ increases strictly with T from Lð0Þ to Lð1Þ, where 2 Lð0Þ ¼ FðLÞðc3 l þ c4 L c2 Þ þ 4l þ
ZL
3 FðtÞdt5½ðc1 c2 Þhð0Þ c4 ;
0
Lð1Þ ¼ kð1; LÞðc3 l c1 Þ þ ðl þ LÞfðc1 c2 Þhð1Þ c4 ½1 kð1; LÞg: Therefore, we have the following optimum policy: (i) If Lð0Þ 0 then T ¼ 0, i.e., the order is made at the same time as the beginning of operation. (ii) If Lð0Þ\0 and Lð1Þ [ 0 then there exists a finite and unique T ð0\T \1Þ that satisfies (B.7). (iii) If Lð1Þ 0 then T ¼ 1.
Chapter 6 6.3 From Problem 3.7 and 3.16, when a ¼ 0:95, 0:99, respectively, z ¼ 15 5:12 1:64 ¼ 6:60; z ¼ 15 5:12 2:33 ¼ 3:07; PrfZð30Þ [ 15g ¼ 0:500;
PrfZð30Þ [ 20g ¼ 0:164:
Appendix B: Answers to Selected Problems
239
6.6 1 1 ðhKÞnþ1 X ðhKÞk ðhKÞn X ðhKÞk ðn þ 1Þ! k¼n k! n! k¼nþ1 k! " # n X 1 1 X ðhKÞ ðhKÞkþ1 ðhKÞk ¼ ðn þ 1Þ [ 0: ðn þ 1Þ! k¼n k! k! k¼nþ1
6.7
Differentiating (6.25) with respect to T, 0
0
½ðcK cT Þr ðTÞ þ cM h ðTÞ
1 X n¼0
6.8
ðnÞ
G ðKÞ
ZT
½F ðnÞ ðtÞ F ðnþ1Þ ðtÞdt:
0
From [17, p. 62], the probability that the system is replaced after failure is N k X X 1 n k Pn ¼ ; ð1Þk pnk ð1Þi k i 1 qni i¼0 k¼0 the probability that it is replaced before failure, i.e., the total number of failed units is N þ 1; N þ 2; . . .; n 1, is n1 k X X 1 n k k nk ; PN ¼ ð1Þ p ð1Þi k i 1 qni i¼0 k¼Nþ1 and the mean time to replacement is N N k X X 1 n nk lNþ1 ¼ l ð1Þk k i 1 qni i¼0 k¼0
ðN ¼ 0; 1; 2; . . .; n 1Þ:
Thus, the expected cost rate is CðNÞ
cn Pn þ cN PN lNþ1
P P cN þ ðcn cN Þ Nk¼0 nk ð1Þk pnk ki¼0 ki ð1Þi ½1=ð1 qni Þ ¼ P PNk nk l Nk¼0 nk ð1Þk i¼0 ½1=ð1 qni Þ i ðN ¼ 0; 1; 2; . . .; n 1Þ; where cn ¼ replacement cost of a failed system and cN ¼ replacement cost before system failure with cN \cn . When the number n of units is given, we can obtain an optimum number n that minimizes CðNÞ.
240
Appendix B: Answers to Selected Problems
6.10 Differentiating the left-hand side of (6.44) with respect to T, r0 ðTÞ
1 X
GðnÞ ðKÞ
n¼0
ZT
½F ðnÞ ðtÞ F ðnþ1Þ ðtÞdt:
0
From (6.48), rNþ2
N N1 X X GðnÞ ðKÞ ½1 GðNþ1Þ ðKÞ rNþ1 GðnÞ ðKÞ þ ½1 GðNÞ ðKÞ n¼0
n¼0
¼ ðrNþ2 rNþ1 Þ
N X
GðnÞ ðKÞ:
n¼0
6.11 When cT ¼ cZ and shocks occur in a Poisson process, i.e., F ðnÞ ðtÞ ¼ P1 j kt , from (6.40), j¼n ½ðktÞ =j!e C5 ðT; ZÞ lim CðT; N; ZÞ N!1 RZ P ðnþ1Þ cT þ ðcK cT Þ 1 ðTÞ 0 GðK xÞdGðnÞ ðxÞ n¼0 F P ¼ : ðnþ1Þ ðTÞGðnÞ ðZÞ ð1=kÞ 1 n¼0 F Differentiating C5 ðT; ZÞ with respect to T and setting it equal to zero, QðT; ZÞ
1 X
F ðnþ1Þ ðTÞGðnÞ ðZÞ
n¼0
1 X
F
ðnþ1Þ
ðTÞ
n¼0
ZZ
GðK xÞdGðnÞ ðxÞ ¼
cT ; cK c T
ðB:8Þ
0
and differentiating C5 ðT; ZÞ with respect to Z and setting it equal to zero, GðK ZÞ
1 X
F ðnþ1Þ ðTÞGðnÞ ðZÞ
n¼0
1 X
F
ðnþ1Þ
ðTÞ
n¼0
ZZ
GðK xÞdGðnÞ ðxÞ ¼
cT ; c K cT
ðB:9Þ
0
where P1 QðT; ZÞ
RZ ðTÞ F ðnþ1Þ ðTÞ 0 GðK xÞdGðnÞ ðxÞ P1 : ðnÞ ðnþ1Þ ðTÞGðnÞ ðZÞ n¼0 ½F ðTÞ F
n¼0 ½F
ðnÞ
Thus, there does not exist both T ð0\T \1Þ and Z ð0\Z \KÞ that satisfy (B.8) and (B.9) simultaneously, because QðT; ZÞ\GðK ZÞ. As further studies, when cZ [ cT , obtain both optimum T and Z that minimize
Appendix B: Answers to Selected Problems
241
P ðnÞ ðnþ1Þ cK ðcK cT Þ 1 ðTÞGðnÞ ðZÞ n¼0 ½F ðTÞ F P1 ðnþ1Þ ðcK cZ Þ n¼0 F ðTÞ RZ 0 ½GðK xÞ GðZ xÞdGðnÞ ðxÞ C5 ðT; ZÞ ¼ RT P1 ðnÞ ðnÞ ðnþ1Þ ðtÞdt n¼0 G ðZÞ 0 ½F ðtÞ F
:
6.13 Suppose that the unit is replaced at time T or at shock number N, whichever occurs first [17, p. 90]. The probability that the unit is replaced at time T is PN1 ðnÞ F ðTÞ, the probability that it is replaced at shock number N is Pn¼0 1 ðnÞ n¼N F ðTÞ, and the expected number of minimal maintenance is N1 X
½F ðnÞ ðTÞ F ðnþ1Þ ðTÞ
n¼1
n X
½1 GðiÞ ðKÞ
i¼1
1 N 1 X X ½F ðnÞ ðTÞ F ðnþ1Þ ðTÞ ½1 GðiÞ ðKÞ þ n¼N
¼
N1 X
i¼1
½1 GðnÞ ðKÞF ðnÞ ðTÞ:
n¼1
The mean time to replacement is T½1 F
ðNÞ
ðTÞ þ
ZT tdF 0
ðNÞ
ðtÞ ¼
ZT
½1 F ðNÞ ðtÞdt:
0
Thus, the expected cost rate is PN1 PN1 ðnÞ cN þ ðcT cN Þ n¼1 F ðTÞ þ cM n¼1 ½1 GðnÞ ðKÞF ðnÞ ðTÞ : CðT; NÞ ¼ RT ðNÞ ðtÞdt 0 ½1 F
Chapter 7 7.1 Z1 1
f ðx; tÞdx ¼ e
s2 r2 t=2
Z1 1
" # 1 ðx þ sr2 tÞ2 2 2 pffiffiffiffiffiffiffi exp dx ¼ es r t=2 : 2r2 t 2ptr
242
Appendix B: Answers to Selected Problems
pffiffiffi 7.2 When t ¼ 5, Rð5Þ ¼ Uðð10 5Þ= 5Þ ¼ 0:987, and 2 3 Z1 Z1 Zt pffiffi 2 10 t 1 MTTF ¼ dt ¼ 4 pffiffiffiffiffiffiffi eð10xÞ =2 x dx5dt 10:5: U pffi t 2pt 0
0
0
When t ¼ 10, Rð10Þ ¼ Uð0Þ ¼ 0:500. 2 7.3 Letting LðTÞ be the left-hand side of (7.10), where A elþr =2 ; Lð0Þ ¼ 0; Lð1Þ ¼ 1 and
L0 ðTÞ ¼ A2 TeAT [ 0;
CðT Þ ¼ c1 AeAT :
7.4 Using ea 1 þ a þ a2 =2, from (7.9), r2 ðl þ r2 =2ÞT c2 e 1þ þ : CðTÞ ¼ c1 l þ 2 T 2 e e e Clearly, Cð0Þ ¼ Cð1Þ ¼ 1. Differentiating CðTÞ with respect to T and setting it equal to zero, rffiffiffiffiffiffiffi 1 2c2 e T ¼ l þ r2 =2 c1 e and T from (7.10), and compare them, and assure that T \ T e, Compute T 2 because for T [ 0 and l þ r =2 [ 1; ATeAT ðeAT 1Þ [
ðATÞ2 : 2
7.5 Letting LðTÞ be the left-hand side of (7.15), Lð0Þ ¼ 0; 0
L ðTÞ ¼
Lð1Þ ¼ lQ1 ð1Þ½1 þ MG ðKÞ 1;
Q01 ðTÞ
1 X
ðkÞ
G ðKÞ
k¼0
ZT
½F ðkÞ ðtÞ F ðkþ1Þ ðtÞdt:
0
7.6 Letting LðNÞ be the left-hand side of (7.18), Lð1Þ ¼ Q2 ð1Þ½1 þ MGT ðKÞ 1; LðN þ 1Þ LðNÞ ¼ ½Q2 ðN þ 2Þ Q2 ðN þ 1Þ
N X n¼0
7.7 Letting LðTÞ be the left-hand side of (7.21), Lð0Þ ¼ 0;
Lð1Þ ¼ 1; ZT 0 0 L ðTÞ ¼ rA ðK=TÞ LA ðK=tÞdt: 0
ðnÞ
GT ðKÞ [ 0:
Appendix B: Answers to Selected Problems
243
7.8 (7.26) and (7.27) are rewritten as, respectively, pffiffiffiffi cK ðcK cT Þf1 exp½ðK xTÞa =r T g CðTÞ ¼ ; RT pffi a 0 f1 exp½ðK xtÞ =r tgdt pffiffiffiffi ðK xTÞa1 ½K þ ð2a 1ÞxT exp½ðK xTÞa =r T pffiffiffiffi pffiffiffiffi 2rT T 1 exp½ðK xTÞa =r T ZT pffiffiffiffi pffi f1 exp½ðK xtÞa =r tgdt exp½ðK xTÞa =r T 0
¼
1 : cK =cT 1
7.9 From (7.29), C2 ð0Þ ¼ 1 and C2 ð1Þ ¼ R 1 0
cK pffi : exp½ðxt KÞ=r tdt
Differentiating C2 ðTÞ with respect to T and setting it equal to zero, K þ xT pffiffiffiffi 2rT T
ZT
pffiffiffiffi pffi exp½ðxt KÞ=r tdt þ exp½ðxT KÞ=r T ¼
cK ; c K cT
0
whose left-hand decreases strictly to 0. Hence, T ¼ 1.
Chapter 8 8.1 When cr ðtÞ ¼ at, (8.12) is [18, p. 53] RT c1 GðTÞ þ a 0 GðtÞdt CðTÞ ¼ : RT l þ 0 GðtÞdt R1 If 0 GðtÞdt b then Cð1Þ ¼ ab=ðl þ bÞ and C(0)=c1/l. Let rðtÞ gðtÞ=GðtÞ, where gðtÞ is a density function of GðtÞ. Differentiating CðTÞ with respect to T and setting it equal to zero, 2 3 ZT al rðTÞ4l þ GðtÞdt5 þ GðTÞ ¼ : ðB:10Þ c1 0
244
Appendix B: Answers to Selected Problems
If rðtÞ decreases strictly then the left-hand side of (B.10) decreases from lrð0Þ þ 1 to ðl þ bÞrð1Þ. Therefore, we have the following optimum policy where k
al c1 ; c1 l
K
al : c1 ðl þ bÞ
(i) If rð0Þ k then T ¼ 0. (ii) If rð0Þ [ k and rð1Þ\K then there exists a finite and unique T ð0\T \1Þ that satisfies (B.10). (iii) If rð1Þ K then T ¼ 1. 8.2 P000 ðtÞ ¼ ðk þ ksÞP00 ðtÞ þ hP01 ðtÞ; P001 ðtÞ ¼ ðk þ ksÞP00 ðtÞ ðk þ hÞP01 ðtÞ þ hP02 ðtÞ; P002 ðtÞ ¼ kP01 ðtÞ hP02 ðtÞ: Taking the LS transforms of the above equations, solving and inverting them, h2 k þ ks x1 k h x1 t x2 k h x2 t ; þ e e P00 ðtÞ ¼ x1 x2 ðk þ hÞðk þ ksÞ þ h2 x1 x2 hðk þ ksÞ k þ ks x1 þ h x1 t x2 þ h x2 t þ e e P01 ðtÞ ¼ ; x1 x2 ðk þ hÞðk þ ksÞ þ h2 x1 x2 kðk þ ksÞ kðk þ ksÞ 1 x1 t 1 x2 t ; þ e e P02 ðtÞ ¼ x2 ðk þ hÞðk þ ksÞ þ h2 x1 x2 x1 where qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 x1 ½2k þ 2h þ ks þ 4kh þ k2s ; 2 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 x2 ½2k þ 2h þ ks 4kh þ k2s : 2 8.3 State State State State
0: 1: 2: 3:
One One One One
unit unit unit unit
is is is is
operating and the other is on standby. operating and the other is under repair. operating and the other is under PM. under repair or PM and the other waits for repair.
Appendix B: Answers to Selected Problems
245
From [18, p. 149], the limiting probabilities Pj (j ¼ 0; 1; 2; 3) of transition probabilities are 1 P0 ¼ c11 Pj ¼
1 cjj
ZT 0
ZT
1 P3 ¼ c11
1 G1 ðtÞFðtÞdt þ c22
Gj ðtÞFðtÞdt
ZT
G2 ðtÞFðtÞdt;
0
ðj ¼ 1; 2Þ;
0
ZT 0
1 G1 ðtÞFðtÞdt þ c22
ZT
G2 ðtÞFðtÞdt;
0
where 2
3 R 1 G1 ðtÞdFðtÞ b1 þ G1 ðtÞFðtÞdt þ 4b2 þ G2 ðtÞFðtÞdt5 TR 1 ; 1 T G2 ðtÞdFðtÞ 0 0 2 3 R1 ZT ZT 1 G2 ðtÞdFðtÞ b2 þ G2 ðtÞFðtÞdt þ 4b1 þ G1 ðtÞFðtÞdt5 R 1T : G 1 ðtÞdFðtÞ T ZT
c11
c22
ZT
0
0
The availability is AðTÞ P0 þ P1 þ P2 h i iR h RT R1 RT 1 c1 þ 0 G1 ðtÞFðtÞdt 1 T G2 ðtÞdFðtÞ þ c2 þ 0 G2 ðtÞFðtÞdt T G1 ðtÞdFðtÞ i iR ¼h ; h RT R1 RT 1 b1 þ 0 G1 ðtÞFðtÞdt 1 T G2 ðtÞdFðtÞ þ b2 þ 0 G2 ðtÞFðtÞdt T G1 ðtÞdFðtÞ
where ci
Z1
Gi ðtÞFðtÞdt;
qi
0
Z1
Gi ðtÞdFðtÞ
ði ¼ 1; 2Þ:
0
Let QðtÞ
q1 G2 ðtÞ q2 G1 ðtÞ ; q1 q2
k
q2 ; q1 c2 q2 c1
K
q1 : lðq1 q2 Þ
Suppose that G1 ðtÞ\G2 ðtÞ for 0\t\1 and the failure rate increases strictly.
246
Appendix B: Answers to Selected Problems
(i) If hð1Þ [ K; q1 c2 [ q2 c1 , and hð0Þ\k or hð1Þ [ K and q1 c2 q2 c1 , then there exists a finite and unique T ð0\T \1Þ that satisfies 2 T 3 Z1 Z ZT QðtÞFðtÞdt5 QðtÞdFðtÞ hðTÞ4 QðtÞFðtÞdt þ 0
¼
q1
0
R1
QðtÞdFðtÞ þ q2 q1 q2
0
RT 0
0
QðtÞdFðtÞ :
(ii) If hð1Þ\K then T ¼ 1. (iii) If q1 c2 [ q2 c1 and hð0Þ [ k, then T ¼ 0. 8.4 When GðtÞ ¼ 1 eht , P j n lz=ð1 zÞ þ ð1=hÞ 1 n¼1 z ½1 FS ðhÞ l ðzÞ ¼ 1 ½1 FS ðhÞ=½1 zFS ðhÞ 1 1 X 1 1 FS ðhÞ X n ¼ lþ nz þ l zn ; h FS ðhÞ n¼1 n¼1 z=ð1 zÞ 1 ½1 FS ðhÞ=½1 zFS ðhÞ 1 1 X 1 FS ðhÞ X n nz þ zn ; ¼ FS ðhÞ n¼1 n¼1
m ðzÞ ¼
M¼
1 1 FS ðhÞ : MS ðhÞ ¼ l þ 1=h l þ 1=h 1 FS ðhÞ
8.5 From [19], the total expected cost of one cycle from the beginning of the operation of the main unit to repair or PM completion is 2 FðTÞ4c2 þ cS
Z1
3
2
MS ðtÞdGðtÞ5 þ FðTÞ4c1 þ cS
Z1
0
3 MS ðtÞdGðtÞ5;
0
and the mean time of one cycle is
FðTÞðl þ TÞ þ FðTÞl þ
ZT
tdFðtÞ ¼ l þ
0
where MS ðtÞ
P1
n¼1
ðnÞ
ZT FðtÞdt; 0
FS ðtÞ. The expected cost rate is CðTÞ ¼
e c 1 FðTÞ þ ec 2 FðTÞ ; RT l þ 0 FðtÞdt
Appendix B: Answers to Selected Problems
247
where e c 1 c1 þ cS
Z1 0
MS ðtÞdGðtÞ;
e c 2 c 2 þ cS
Z1
MS ðtÞdGðtÞ:
0
Using the results in Sect. 3.3.2.1, we can obtain an optimum policy.
References 1. Osaki S (1992) Applied stochastic system modeling. Springer, Berlin 2. Tijms HC (2003) A first course in stochastic models. Wiley, Chichester 3. Beichelt FE, Fatti LP (2002) Stochastic process and their applications. CRC Press, Boca Raton 4. Rausand M, Høyland A (2004) System reliability theory. Wiley, Hoboken 5. Bhat UN, Miller GK (2002) Elements of applied stochastic processes. Wiley, Hoboken 6. Feller W (1968) An introduction to probability theory and its application, vol I, 3rd edn. Wiley, New York 7. Gnedenko B, Ushakov I (1995) Probabilistic reliability engineering. Wiley, New York 8. Levitin L, Lisnianski A (2003) Multi-state systems reliability analysis and optimization. In: Pham H (ed) Handbook of reliability engineering. Springer, London 9. Levitin G (2005) The universal generating function in reliability analysis and optimization, Springer, London 10. Çinlar E (1975) Introduction to stochastic processes. Prentice-Hall, Englewood Cliffs 11. Ross SM (1983) Stochastic processes. Wiley, New York 12. Yun WY, Nakagawa T (2010) Note on MTTF of a parallel system. In: 16th ISSAT International conference on reliability and quality in design, pp 235–237 13. Nakagawa T (2008) Advanced reliability models and maintenance policies. Springer, London 14. Chen M, Mizutani S, Nakagawa T (2010) Optimal backward and backup policies in reliability theory. J Oper Res Soc Jpn 53:101–118 15. Prabhu NU (2007) Stochastic processes. World Scientific, Singapore 16. Barlow RE, Proschan F (1965) Mathematical theory of reliability. Wiley, New York 17. Nakagawa T (2007) Shock and damage models in reliability theory. Springer, London 18. Nakagawa T (2005) Maintenance theory of reliability. Springer, London 19. Nakagawa T, Osaki S (1976) Reliability analysis of a one-unit system with unrepairable spare units and it optimization applications. Oper Res Q 27:1019–110
Index
A Abelian theorem, 221–222, 224 Age, 58–78 Age replacement, 3, 66–70, 88, 175 Allowed time, 83 Alternating renewal process, 3, 47, 73–83, 86, 98, 125, 127–128, 199, 202–203, 211 Automatic repeat request (ARQ), 130–132 Availability, 7, 76, 101, 203, 210
Cumulative hazard, 28–30, 34, 44, 61–63, 72 Cumulative process, 1, 4, 18, 149–179 Cumulative damage, model, 4, 8, 36–41, 149–179, 185, 187–189
B Backup policy, model, 7, 25–26 Backward time, 3, 26–27, 44 Bathtub curve, 10, 33 Bernoulli trial, 67 Binomial distribution, 17, 23–24, 43, 113, 121, 133 Birth process, 112–114, 158 Birth and death process, 112, 114–119 Blackwell’s theorem, 56 Block replacement, 71 Branching process, 5 Brownian motion, 1, 4–5, 18, 149, 183–186, 192
E Elementary renewal theorem, 55, 69 Embedded Markov chain, 125, 132–139 Expected cost, rate, 12, 24–26, 34, 43, 70–73, 83, 85, 88, 140, 147, 160, 170, 172–178, 180, 186–189, 192–194, 196, 201–203, 215–217 Expected number of visits, failures, 102, 108, 110, 124, 130, 136, 138, 147, 208
C Central limit theorem, 57, 152–153 Chapman-Kolmogorov equation, 99, 111, 114, 133 Checkpoint, 25 Communication system, 128–130, 140 Compound Poisson process, 3, 8, 36–41 Computer, system, 5, 11 Counting process, 16, 19, 31, 48, 112, 149, 158
D Degenerate distribution, 67, 77, 145, 155 Doubly stochastic process, 4, 149 Downtime, 4, 79–83
F Failure rate, 7, 9–11, 14, 16–17, 27–28, 30, 33–35, 38, 40, 42, 62, 70, 72, 78, 89, 93, 121, 149, 156–159, 164, 189, 200, 216 Fault tolerance, 2 Finance, 1, 5, 149 First-passage time, distribution, 137, 206–207, 211–212
G Gamma distribution, 14–15, 21, 51–53, 60, 64–65, 91, 168, 200–201
249
250
G (cont.) Gamma function, 16 Gaussian process, 5 Generating function, 87–90, 101, 134–137, 212–214 Geometric distribution, 86 Geometric process, 47, 84–86
H hazard rate, function, 9, 10 harmonic series, 13, 43
I Imperfect maintenance, 28 Imperfect repair, 93, 201 Increasing failure rate (IFR), 10, 12, 33, 43, 62–63, 68, 72, 92, 154, 157, 200 Independent increment, 19, 112, 183, 191 Inspection model, policy, 3, 7, 24–25 Insurance, 1, 5, 149 Intensity function, 30, 154, 158, 200 Intermittent fault, 128–130 Interval reliability, 47, 76–79, 203 Inventory model, cost, 144–147
Index Markov property, 3, 97, 123, 184 Markov renewal process, 4, 73, 123–147, 199, 205 Martingale, 5 Mass function, 125 Mean time to failure (MTTF), 4, 7–13, 38–43, 57, 113, 149, 156, 164, 184, 207, 211, 214 Mean time between failures (MTBF), 64, 86 Mean value function, 30, 33, 154, 158, 160, 163, 200 Memoryless property, 13, 42, 48, 96, 99, 200 Minimal repair, maintenance, 7, 27, 34, 62, 72, 160, 171, 178–180, 200 Minor failure, 3, 44, 95–96
N Negative binomial, 89 Normal distribution, 51, 54, 57, 81, 152, 184, 189 Nonhomogeneous Poisson process, 3, 7, 27, 34, 41, 61, 64, 175
O One-unit system, 2–4, 47, 76, 80, 83–84, 86, 90, 98, 123, 199, 200
J Jump process, 4, 149
K k-out-of-n system, 43
L Laplace transform, 3–74, 184, 196, 219–222 Laplace–Stieltjes (LS) transform, 5, 8, 15, 38, 49–53, 61, 66–67, 74–78, 80, 82, 115–117, 127, 129, 130, 142–148, 154–156, 167, 201, 204–209, 212–214, 219–222 Laplace inversion, 65, 76, 205 L´evy process, 1, 4–5, 18, 149, 158, 183–186 Lifetime, 32
M Maintenance, policy, 2, 4, 7, 47, 51 Major failure, 3, 44, 95 Markov chain, 1, 3–4, 95–119, 124–125, 138–139, 199, 204 Markov process, 1, 3, 119, 124–132
P Parallel system, 3–5, 13, 43, 92, 109–110, 113, 115–118, 120, 142, 163 Periodic replacement, 3, 69, 71–73, 150, 178–179 Poisson distribution, 7, 17, 21, 22, 43, 200, 224 Poisson process, 1–44, 47, 52, 61, 64, 112, 154, 156–159, 163, 166–167, 175, 180, 192, 200–201, 224 Preventive maintenance, 2, 11–12, 28, 77, 200–201
Q Queueing, 1, 5, 124, 149
R Regeneration point, 4, 48, 123, 140, 147 Renewal density, 50, 58, 61, 65 Renewal equation, 3–4, 47, 51, 61, 65–66, 68–69, 71, 74, 123–124, 126, 129–130, 166, 209–212
Index Renewal function, 3–4, 47, 50–53, 61, 64–66, 71, 75, 85, 87, 92, 124–127, 141–143, 152, 201–202, 204, 208, 213 Renewal point, 48, 51, 56, 66–67 Renewal process, 1, 3–4, 47–52, 55–89, 92–93, 98, 124–127, 140–144, 150, 159–160, 163, 170, 199, 201–203, 205–206, 211 Renewal theory, 3, 47, 125 Repair, limit, 2–5, 47, 73, 76–77, 83, 86, 95, 98–99, 117, 132, 186, 199, 202–204, 211 Replacement model, policy, 2, 7, 34, 42, 47, 84, 86, 149–150, 160, 170–178, 186–195, 199, 201–203 Residual life, time, 13, 48, 58–59, 85, 200 Redundant unit, system, 2, 5, 15, 73, 113, 123, 132, 140, 142, 163, 199–216 Reward process, 56
S Sample path, 19 Scheduling, time, 17 Semi-Markov process, 4, 123–125, 133 Series system, 42 Shock, model, 3–4, 8, 36–41, 149–180, 187, 193–196 Software reliability, 7, 36
251 Spare unit, part, 2, 7, 17, 44, 109, 110, 132, 140, 144, 199, 211–217, 226 Standby system, 4, 7, 15, 17–18, 199, 203–205, 211–216 Stationary, increment, 7, 19, 22, 97, 111, 124, 133, 183, 191 Stationary process, 5
T Tauberian theorem, 69, 127, 208, 210, 221–222, 224 Transition probability, 3–4, 74, 75, 95–130, 133–136, 142–146, 199, 204, 210, 216 Two types of failures, 95–96 Two-unit system, 2, 4–5, 92, 109, 115–117, 120–121, 140, 199, 203, 211, 216
U Uniform distribution, 24–25, 60 Up time, 4
W Wear model, 5, 192 Weibull distribution, 32–33, 35, 43, 64, 66, 196 Wiener process, 4, 183