Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

International Journal of Production Research

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/tprs20

Selective maintenance scheduling under


stochastic maintenance quality with multiple
maintenance actions

Chaoqun Duan, Chao Deng, Abolfazl Gharaei, Jun Wu & Bingran Wang

To cite this article: Chaoqun Duan, Chao Deng, Abolfazl Gharaei, Jun Wu & Bingran Wang
(2018) Selective maintenance scheduling under stochastic maintenance quality with multiple
maintenance actions, International Journal of Production Research, 56:23, 7160-7178, DOI:
10.1080/00207543.2018.1436789

To link to this article: https://doi.org/10.1080/00207543.2018.1436789

Published online: 15 Feb 2018.

Submit your article to this journal

Article views: 1856

View related articles

View Crossmark data

Citing articles: 145 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tprs20
International Journal of Production Research, 2018
Vol. 56, No. 23, 7160–7178, https://doi.org/10.1080/00207543.2018.1436789

Selective maintenance scheduling under stochastic maintenance quality with multiple


maintenance actions
Chaoqun Duana,b∗, Chao Denga∗, Abolfazl Gharaeib, Jun Wuc and Bingran Wangd
a
School of Mechanical Science and Engineering, Huazhong University of Science and Technology, Wuhan, P.R. China; bDepartment
of Mechanical and Industrial Engineering, University of Toronto, Toronto, Canada; cSchool of Naval Architecture and Ocean
Engineering, Huazhong University of Science and Technology, Wuhan, P.R. China; dDivision of Engineering Science, University of
Toronto, Toronto, Canada
(Received 3 August 2017; accepted 17 January 2018)

Many systems are required to perform a series of missions with finite breaks between any two consecutive missions. To
improve the probability of system successfully completing the next mission, maintenance action is carried out on compo-
nents during the breaks. In this work, a selective maintenance model with stochastic maintenance quality for multi-com-
ponent systems is investigated. At each scheduled break, a set of maintenance actions with different degrees of impact
are available for each component. The impact of a maintenance action is assumed to be random and follow an identified
probability distribution. The corresponding maintenance cost and time are modelled based on the expected impact of the
maintenance action. The objective of selective maintenance scheduling is to find the cost-optimal maintenance action for
each component at every scheduled break subject to reliability and duration constraints. A simulated annealing algorithm
is used to solve the complicated optimisation problem where both multiple maintenance actions and stochastic quality
model are taken into account. Two illustrative numerical examples and a real case study have been solved to demonstrate
the performance of the proposed approach. A comparison with deterministic maintenance shows the importance of con-
sidering the proposed stochastic quality in selective maintenance scheduling.
Keywords: selective maintenance scheduling; stochastic maintenance quality; multiple maintenance actions; conditional
reliability; simulated annealing algorithm

1. Introduction
Many systems in the industry are required to perform several missions with finite breaks between missions. In order to
maintain good conditions of the systems, we have to decide which components should be given maintenance activities
in the limited time allotted between missions (Dao, Zuo, and Pandey 2014). This kind of maintenance problem is called
selective maintenance, and it exists commonly in many industrial applications, such as energy systems, production
equipment, and military fields (Wong, Chan, and Chung 2012; Khatab et al. 2017).
Different from opportunistic maintenance, which utilises the finite breaks and shares set-up cost/work (Koochaki
et al. 2012; Xia et al. 2013, 2015, 2016, 2017; Xia, Tao, and Xi 2017), selective maintenance focuses on selecting
appropriate maintenance actions under limited maintenance resources (e.g. system reliability, availability, budget) during
mission beaks. The purpose of selective maintenance is to find trade-offs between maintenance expenditure and system
risks or/and plant profits. Rice, Cassady, and Nachlas (1998) first introduced the selective maintenance model for a ser-
ies-parallel system consisting of identical components under limited maintenance time. In their paper, the authors con-
sidered constant component failure rates and only one type of maintenance action, i.e. replacement of failed component.
Cassady, Pohl, and Paul Murdock (2001b) further proposed a generalised modelling framework for optimising selective
maintenance decisions. Three selective maintenance optimisation problems were presented with deterministic model
parameters, which are constant maintenance time and maintenance cost. To make the selective maintenance problem
more practical, Cassady, Murdock, and Pohl (2001a) considered component lifetime follows Weibull distribution, and
multiple maintenance options, i.e. minimal repair, preventive maintenance, and replacement, are available between mis-
sions. In all these mentioned selective maintenance models, the maintenance quality is considered to be perfect and the
system is assumed to be either in as-good-as-new or as-bad-as-old condition after maintenance. To describe actual influ-

∗Corresponding authors. Email: duancq@mie.utoronto.ca (C. Duan), dengchao@mail.hust.edu.cn (C. Deng)


This article has been corrected with minor changes. These changes do not impact the academic content of the article.

© 2018 Informa UK Limited, trading as Taylor & Francis Group


International Journal of Production Research 7161

ence of maintenance, imperfect maintenance models such as age reduction coefficient, hazard rate adjustment factor, and
hybrid imperfect maintenance have been utilised in various maintenance applications (Xia et al. 2012, 2013, 2015,
2016; Xia, Tao, and Xi 2017; Xia et al. 2017; Duan, Deng, and Wang 2017a, 2017b). Liu and Huang (2010) considered
imperfect quality of maintenance in selective maintenance framework for multi-state systems. An age reduction coeffi-
cient is employed to describe the imperfect maintenance quality, and a comparative analysis between the strategies with
and without considering imperfect maintenance indicated that incorporating imperfect maintenance quality into selective
maintenance modelling achieves better outcomes. Afterwards, Zhu et al. (2011) used the age reduction coefficient to
model the quality of imperfect preventive maintenance in selective maintenance for machining line systems. Pandey
et al. (2013) also proposed a generalised hybrid imperfect maintenance model including both age reduction coefficient
and hazard rate adjustment factor in scheduling selective maintenance for binary systems. Selective maintenance prob-
lem for multi-state systems under imperfect maintenance is later addressed by Pandey, Zuo, and Moghaddass (2013). In
Doostparast, Kolahan, and Doostparast (2014), different levels of maintenance are provided to improve the reliability of
a component under system reliability requirement and maintenance resources constraint. The effects of imperfect mainte-
nance actions are distinct and defined on component’s reliability.
Most of the existing research on selective maintenance problems assumed the maintenance action can affect compo-
nent’s state with a certain degree, which means the maintenance improvement is fixed with corresponding maintenance level
(Tsai, Liu, and Lio 2011; Dijoux, Fouladirad, and Nguyen 2016; Liu et al. 2016; Zhao and Xie 2017). However, this
assumption may be violated in real-world situations, where the working environment is uncertain and the repair is not
always perfectly performed. In fact, the maintenance improvement is closely related to the repairman’s expertise,
maintenance methods, tools and environments. It’s more practical and reasonable to assume that the maintenance
improvement is stochastic and its quality can be modelled as a random variable represented by an appropriate probabil-ity
distribution. In non-selective maintenance framework, Wu and Clements-Croome (2005) assumed the quality of a
preventive maintenance action is a random variable following the uniform distribution. Two frequently studied maintenance
models, a hazard rate adjustment factor and an age reduction coefficient, were investigated. A paper presented by Liu et al.
(2011) assumed that the possible values for the age reduction coefficient fall within fixed support at time t and fol-low a
normal distribution. Khatab and Aghezzaf (2016) presented a stochastic age reduction coefficient to describe the
maintenance improvement, and assumed the improvement following a Beta distribution. Even though, the assumptions of
random improvement factors are more realistic than the assumptions of constant improvement factors (Zhang, Gaudoin, and
Xie 2015), no paper has been found with regard to considering random improvement for solving selective mainte-nance
scheduling problem.
In this paper, we consider a selective maintenance scheduling problem under stochastic maintenance quality with
multiple maintenance options. The maintenance quality is assumed to obey a continuous probability distribution defined on
interval [0, 1]. The maintenance cost and time invested are constructed according to maintenance quality to reflect the
maintenance restoration effect. At each scheduled break, various maintenance levels which have different effects on
component’s age are available to ensure the reliability of completing next mission remaining at desirable level. Due to the
limitations of maintenance resources and break duration, not all the components are likely to be maintained and maintenance
action needs to be chosen properly to achieve maximum benefits. For such a maintenance problem, tradi-tional enumeration-
based techniques, such as branch and bound method (Shih 1979; Batun and Azizoğlu 2009; Shim 2009), are inefficient as
they suffer from the excessive computational times. In this regard, a simulated anneal-ing algorithm is employed to solve the
resulting optimisation problem. The simulated annealing algorithm is usually not restricted to the problem size and structure,
and it has been widely used in constrained problem optimisations with providing good solutions in reasonable computational
time, e.g. Pedamallu and Ozdamar (2008); Liu et al. (2010); Şahin, Ertoğral, and Türkbey (2010); Wang et al. (2015). Other
interesting applications of the simulated annealing algorithm for maintenance scheduling can be found in Nakagawa (2005);
Rasmekomen and Parlikad (2016).
The present paper extends the age reduction model presented by Wu and Clements-Croome (2005) to a multi-level
setting with multi-component systems in a selective maintenance framework. The maintenance decision is made to first
select the best subset of components, and then to choose an appropriate level of maintenance for each selected component.
The optimality criterion is the minimisation of the long-run maintenance cost under reliability requirement and break con-
straint. To our knowledge, this is the first paper providing an optimisation tool to solve the selective maintenance scheduling
under stochastic quality of maintenance with multiple levels. The rest of this paper is organised as follows. Section 2
describes the characteristics of the system. Section 3 presents the stochastic maintenance model and corresponding main-
tenance cost and time structures. The conditional reliability model of completing the next mission is also derived.
Section 4 reviews the non-homogenous Poisson process (NHPP) and applies for modelling the failures process of the sys-
tem. In Section 5 ~, the selective maintenance optimisation problem is formulated A. simulated annealing algorithm is emplo-
yed to obtain the optimum solution, which is presented in Section 6. Section 7 provides two illustrative numerical
7162 C. Duan et al.

examples and a real case study to justify the implementation of the computational algorithm and experimental results.
Comparisons with the previous approach as well as with the case of deterministic maintenance quality are also given.
Section 8 contains concluding remarks.

2. System description and analysis


Consider a system composed of N independent components C1, …, CN. The system operates to execute a sequence of
missions running one after another one. At the end of each mission, a scheduled mission break of a limited length starts,
and inspection with following possible maintenance activities can be performed to keep the conditional reliability of
completing next mission at a desirable level. When some proper maintenance activities are done, the new mission starts
immediately.
For a given component Ci, a list of corresponding Li + 1 options (levels) {0, 1, ⋯, li, ⋯, Li} are available at sched-
uled breaks, where the actions can be categorised into three types:
• Inspection only: this type of action usually involves simple services, such as lubricating, adjusting/calibrating,
tightening the loose parts, cleaning dust, and adding supplements (oil, waters, etc.). It usually requires negligible
cost compared to maintenance activity. These services do not improve the condition of the system, it is actually
the special case of maintenance with level li = 0, which is denoted as maintenance 0 in this paper.
• Intermediate maintenance level: denoted by li where 0 < li < Li. These actions are mainly employed for repairing
some parts from different levels. Examples for this type of maintenance are disassembly and reassembly of
machinery, surface treatments of the moving parts and calibrations. The actions usually bring the system’s health
condition to somewhere between the as good as new and as bad as old.
• Replacement or overhaul: denoted by Li, this type of maintenance action is to overhaul or replace a component
with a new one. It is frequently adopted for the key components to avoid potentially serious damages. In addition,
the components which have undergone several times intermediate maintenance and are not worthy to go on
using, may also take this type of action.
We assume that the system requires finishing a certain quota in each mission with constant production rate, thus the
duration of each mission time is constant and denoted by Δ. At the end of each mission, a scheduled break is initiated,
and all the components are regularly inspected. At the inspection, depending on the component deterioration state and
the required system reliability level to execute the next mission, a component Ci may be selected for one of maintenance
actions in the set {1, …, li, …, Li} such that required level of the system reliability is maintained under constrained
break time with the minimal long-run expected average cost.
The maintenance activity presented here has indistinct impact on components condition. Different maintenance level
has different degrees of impact. The impact of each maintenance level is modelled as a random variable. The
mainte-nance cost and the time assigned to each maintenance activity are related to its maintenance level. For a given
compo-nent, a maintenance level li has its own cost and time requirements.
To assess the maintenance cost and time consumed in maintenance activity, we first need to evaluate the impact and the
improvement of each maintenance level. In the next section, the quality of each maintenance level is characterised, the
corresponding maintenance cost and time models are constructed based on the quality of maintenance level, then the
effect of maintenance on the system conditional reliability is derived.

3. Maintenance model formulation


3.1 Modelling the random quality of maintenance
In the literature related to maintenance policy, a system is typically assumed to be restored to a certain condition by a
maintenance action with deterministic quality (Nakagawa 1979; Pham and Wang 1996; Tsai, Wang, and Tsai 2004;
Nodem, Gharbi, and Kenné 2011; Ben Mabrouk, Chelbi, and Radhoui 2016; Liu et al. 2016; Alaswad and Xiang
2017). After a maintenance action, the age t of the system is reduced to t × (1 − η) where η for 0 ≤ η ≤ 1 is the age
reduction coefficient. Accordingly, the system becomes as good as new if its age is reset to zero with age reduction coef-
ficient g ¼ 1, while it becomes as bad as old if the age has no change with η = 0. The former case corresponds to the
replacement or overhaul which restores the system to completely new state, while the latter case corresponds to inspec-
tion which only some clean-up services and adjustments are applied.
However, due to the differences in improvement between individual components as well as in repairman’s expertise
degrees, it is usually difficult to specify precisely the quality of a maintenance action in real-world environments. Even
International Journal of Production Research 7163

in response to the same maintenance action, the maintenance improvement may vary for a same component, and the
restoration of a maintenance action is not fixed in reality. Throughout this paper, the quality of maintenance is assumed
to be stochastic, and a random age reduction coefficient is utilised to model the maintenance impact. For a given com-
ponent Ci, each maintenance level li is assigned an age reduction coefficient represented by a random variable gli . Since
gli 2 ½0; 1, the probability density function (PDF) and cumulative distribution function (CDF) of the random variable gli
are only defined on interval [0,1], and denoted by f ðgli Þ and Fðgli Þ, respectively.
For each maintenance level li, the corresponding random variable gli follows a Beta distribution for which the f ðgli Þ
is defined on the support [0,1] as:

1
f ðgli Þ ¼ gai 1 ðgli Þbi 1 (1)
Bðai ; bi Þ li
R1
where αi and βi, for αi > 0, βi > 0, are shape parameters; Bðai ; bi Þ ¼ 0 t ai 1 ð1  tÞbi 1 dt is the Beta function. According
to the property of Beta distribution, the average value lgl and the standard deviation rgli of gli are given by
i

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ai ai bi
lgl ¼ ; rg l i ¼ (2)
i ai þ b i ðai þ bi Þ2 ðai þ bi þ 1Þ
In the equation above, both lgl and rgli specify the quality of each maintenance level li. The lgl evaluates the average
i i
maintenance quality while the rgli characterises the maintenance uncertainty. Specifically, a low value of rgli indicates a
low influence induced by external factors, i.e. maintenance techniques, environment factors, etc., while a high value of
rgli indicates a high dispersion of maintenance quality resulting from complicated external factors.
The random age reduction coefficient is depicted for some cases of maintenance actions in Figure 1. In the figure,
the curves correspond to the PDF of three intermediate maintenance actions on the support [0,1] with average age
reduction coefficients 0.25, 0.5, 0.75 under standard deviation 0.01. The average value of the age reduction coefficient
reflects the expected improvement of the maintenance action. For intermediate maintenance, the average value of Beta
distribution is between 0 and 1 depending on different levels. In the case of inspection or replacement, the average value
of Beta distribution is 0 or 1. After defining quality of the maintenance, we next evaluate the related maintenance cost
and time, which are presented in the following subsection.

Figure 1. Random age reduction coefficient (PDF).


7164 C. Duan et al.

3.2 Maintenance cost and time evaluation


In real-world environments, the maintenance quality improves with the amount of the budget invested during mainte-
nance activities (Duan, Deng, and Wang 2017b). Liu and Huang (2010) showed that the relationship of maintenance
improvement and corresponding incurred cost, which is given by

MCli b
gli ¼ ða  Þ (3)
MCPi

where MCPi is the preventive replacement cost of component Ci; MCli is the cost of intermediate maintenance li; a and
b are the related adjustment parameters. Considering stochastic maintenance quality, the relationship of age reduction
coefficient being a function of maintenance cost must be reviewed and adapted to the present paper’s context.
In the present work, age reduction coefficient gli is determined by the expected value lgl and standard derivation
i
rgli . Depending on the system performance, a component may or may not be selected for the maintenance. When it is
not selected, the corresponding maintenance cost is nil. However, if a component Ci is selected for maintenance with
intermediate level of li, 0 < li < Li, it consumes some maintenance budget. The expression for maintenance cost of the
component Ci can be expressed as:
dc rgl
MCli ¼ rc  MCPi  ðlgl Þ i (4)
i

where rc and dc (rc, dc > 0) are characteristic constants that determine the exact relationship between age reduction coef-
ficient and corresponding intermediate maintenance cost through (4). It is related to the inherent characteristics of the
component, and can be estimated via the collected repair cost and reliability/failure data of the component. The rgli
defined in equation reveals the stability level of the maintenance li, which means the smaller rgli , the higher maintenance
cost. Also, if li = Li, then lgl ¼ 1, which corresponds to preventive replacement with cost MCPi . The equation above
i
states that maintenance budget consumed by Ci, when it receives a maintenance action of level li, is equal to MCli
depending on expected value and standard derivation of the age reduction coefficient as well as cost of preventive
replacement.
In addition to cost model, maintenance time can be used as a measure of maintenance quality, since high quality of
maintenance action usually incurs relative long time. The maintenance time assigned to intermediate maintenance level
is related to the improvement of that maintenance level. In the same manner, the time to perform an intermediate main-
tenance of level li for 0 < li < Li on component Ci is evaluated as

dT rgl
MTli ¼ rT  TPi  ðlgl Þ i (5)
i

where TPi is the time required to perform preventive replacement of component Ci; rT and dT are the corresponding char-
acteristic constants that determine the exact relationship between age reduction coefficient and corresponding intermedi-
ate maintenance time. If li = Li, it corresponds to preventive replacement with time TPi . Therefore, a high expected
value with a low standard derivation of the improvement level induces a high maintenance cost with long maintenance
time, and the health condition of maintained component will have a great recovery.

3.3 Conditional reliability evaluation


According to the age reduction model, component’s effective age can be improved through a maintenance action. If
a maintenance action with an eligible level li is performed on component Ci at the end of the mth mission, then the
effec-tive age Ai(m + 1 ) o f Ci at the beginning of the next mission is evaluated as:
Ai ðm þ 1Þ ¼ ð1  gli Þ  Bi ðmÞ (6)
where gli is the age reduction coefficient of maintenance level li. Bi(m) is the effective age of Ci at the end of the mth
mission. If at the scheduled break only inspection action is given, the age of component Ci has a change with expected
value of zero. The inspection action corresponds to maintenance level 0 whose average value of random age reduction
coefficient is equal to 0. If a maintenance li for 0 < li < Li is set up, the effective age will restore to ð1  gli Þ  Bi ðmÞ. In
the case where a replacement of component Ci is chosen, the component becomes in average as good as new after
replacement with a random age reduction coefficient gLi whose average value is equal to 1.
International Journal of Production Research 7165

Suppose that the system has just executed the mth mission and become available for maintenance, the conditional
reliability that component Ci successfully completes the next mission given that it received a maintenance action of level
li at current scheduled break can be expressed as

Ri ðm þ 1jm; li Þ ¼ PrðTi [ Aðm þ 1Þ þ DjTi [ Aðm þ 1Þ; li Þ (7)


where Ti denotes the random variable of the lifetime of component Ci. This conditional reliability depends on effective
age of Ci at the beginning of mission m + 1 as well as on the duration Δ of the next mission. According to Equation
(7), the conditional probability that component Ci successfully completes the next mission at mth scheduled break can
be computed as follows:

PrðTi [ ð1  gli Þ  Bi ðmÞ þ DÞ


PrðTi [ Aðm þ 1Þ þ DjTi [ Aðm þ 1Þ; li Þ ¼ (8)
PrðTi [ ð1  gli Þ  Bi ðmÞÞ
The gli in Equation (8) is a random variable defined on the support [0,1] with cumulative distribution function Fðgli Þ, the
conditional probability can be further written as

R1 R1
PrðTi [ ð1  gli Þ  Bi ðmÞ þ hÞ PrðTi [ ð1  gli Þ  Bi ðmÞ þ DÞdFðgli Þ Ri ðð1  gli Þ  Bi ðmÞ þ DÞdFðgli Þ
¼ 0
R1 ¼ 0
R1 (9)
PrðTi [ ð1  gli Þ  Bi ðmÞÞ
0 PrðTi [ ð1  gli Þ  Bi ðmÞÞdFðgli Þ 0 Ri ðð1  gli Þ  Bi ðmÞÞdFðgli Þ

where Ri(t) is the reliability function that component Ci could survive at time t. The conditional reliability of whole sys-
tem successfully completing the next mission, i.e. R(m + 1|m), can be evaluated from component reliability Ri(m + 1|m,
li) based on the system block diagram. It should be emphasised that the deterministic age reduction coefficient and
stochastic age reduction coefficient will lead to different values of conditional reliability Ri(m + 1|m, li). Indeed, the con-
ditional reliability obtained by a random age reduction coefficient is either smaller or greater than the conditional relia-
bility determined by deterministic age reduction coefficient, which reflects the uncertainty of the maintenance effect.

4. A model for random failures


We assume that the random failures follow a non-homogeneous Poisson process (NHPP) with power law intensity func-
tions as NHPP is commonly used in repairable systems subject to minimal repairs. Let N(t) denote the number of ran-
dom failures in the interval (0, t], t ≥ 0.
For a NHPP with rate h(t), the number of arrivals in any interval is a Poisson random variable. The random variable
N(t + τ) − N(t) hasRa Poisson distribution with mean H(t + τ) − H(t), where H(t) is the cumulative hazard rate function
t
with form HðtÞ ¼ 0 hðxÞdx. More specifically, we can write

½Hðt þ sÞ  HðtÞk
PfN ðt þ sÞ  N ðtÞ ¼ kg ¼ expððHðt þ sÞ  HðtÞÞÞ (10)
k!
Suppose that a repairable system with lifetime T starts working from t = 0. The counting process N(t) represents the num-
ber of failures during the interval [0, t]. Between the scheduled break interval, N(t + Δ) − N(t) has a Poisson process with
the mean H(t + Δ) − H(t) = −logR(t + Δ) + logR(t), where R(t) is the reliability function at t for t ≥ 0. The process
N(t + Δ) − N(t) increases as R(t) decreases. Whenever a random failure occurs, a minimal repair is performed to bring the
failed system back to the previous operating state.

5. Formulation of the optimisation problem


In this section, we focus on formulating the mathematical model for maintenance optimisation. In reality, the maintenance
managers are always faced with different constraints of selective maintenance. Cassady, Pohl, and Paul Murdock (2001b)
proposed three different mathematical programming models: reliability-based maintenance, cost-based maintenance, and
availability-based maintenance for optimising selective maintenance for systems. In a safety-critical application, e.g.
maintaining aircraft or spacecraft, it is likely that reliability should be paramount due to the severe implications of a sys-
tem failure. In a profit-oriented application, e.g. maintaining manufacturing equipment, cost-based maintenance model is
likely to be chosen, since controlling cost is a primary concern. In a service-oriented application, e.g. maintaining
7166 C. Duan et al.

computer servers, breaks between missions may correspond to loss to service for customers, and availability-based main-
tenance model is likely to be the formulation of choice.
In fact, the three mentioned cases can be formulated in same manner. Once one of the selective maintenance models
is established, the other two cases can be completed through a simple modification. Hence, we choose the commonly
and extensively used criterion, i.e. cost-based maintenance objective, as our optimisation goal.
Before presenting the mathematical formulation of the cost function, we define the following decision variables:
8
< 1 If a maintenance action of level li is performed on ith component
½li 
Xim ¼ in the mth mission period
:
0 otherwise

8
< 1 If in the mth mission period, system downtime occurs due to the
Ym ¼ maintenance action
:
0 otherwise
Based on the definitions above, the mathematical model for the optimal selective maintenance planning may be stated
as:

!
X
M X
Li X
N
½l 
X
M X
M X
Li X
N
½l 
min Ximi MCðli Þ þ KðmÞMCf þ Ym  D Ximi (11)
m¼1 li ¼1 i¼1 m¼1 m¼1 li ¼1 i¼1

subject to:

Rðm þ 1jmÞ  Rmin ; 8m 1  m  M (12)

X
Li X
N
½l 
Ximi Tli  Tmax ; 8n; m 1  i  N 1  m  M (13)
li ¼1 i¼1

X
Li
½l 
Ximi  1; 8n; m 1  i  N 1  m  M (14)
li ¼1

where M is the number of break periods during the planning horizon; MCf is the cost of random failure of the system; N is the
number of components in the system; D is average cost per unit time of the system downtime loss due to the planned
maintenance actions; Λ(m) = −ln(R(m + 1 | m) − (−ln(R(m|m − 1)). In this mathematical model, the objective function is
composed of three terms as shown in (11). The first term is related to the maintenance cost, the second term is associated with
the random failure cost. Λ(m) is the expected average number of random failures during mth mission. Generally, this cost
increases as probability of random failure occurrence increases resulting from decreasing of system reliability. The
last term represents the cost related to the planned system downtime due to the maintenance activities.
The first constraint in Equation (12) endorses the minimum required conditional reliability of system during the
entire planning horizon. Constraint (13) guarantees that the total time for maintenance at each break during the entire
planning horizon does not exceed available time Tmax . Constraint (14) signifies that in each mission break, only one of
maintenance actions or no maintenance action can be performed on a given component.

6. Solution approach
After configuring the mathematical model, we work on the solution of the selective maintenance optimisation model.
The objective is to find the optimal maintenance parameters while the constraints (12) to (14) are satisfied. Due to
the large number of the involved parameters in the model which have strong and non-linear interdependence effects, the
heuristic algorithms can be utilised as the optimisation techniques for such a problem. Current comparisons among the
International Journal of Production Research 7167

heuristic algorithms show that for the problem with discrete variables under dynamic maintenance environment, where
the input parameters may change often with system operating, the simulated annealing algorithm is much more stable to
find global optimal result for dynamic scheduling problems (Bertsimas and Tsitsiklis 1993; Breedam 2001; Saraiva et al.
2011). Moreover, the simple principles of simulated annealing algorithm are also preferable for industrial partners to
understand and carry out. Thus, the simulated annealing algorithm is applied to find the optimal selective maintenance
plan for multi-component systems in this paper. The simulated annealing algorithm generates the solutions to find the
minimum function value in the minimisation problems by simulating the cooling process, a solid is first melted in a high
temperature and then slowly cooled to reach a thermodynamic equilibrium. It is proved that if the control parameters of
the cooling process are properly chosen, the simulated annealing can provide good quality solutions within reasonable
searching times, and a detailed proof of convergence is also shown in Bertsimas and Tsitsiklis (1993).
A flow chart is shown in Figure 2 to summarise the simulated annealing algorithm, which is applied to optimise the
selective maintenance problem in this paper. The idea of simulated annealing algorithm is to use a sample of randomly
generated solutions. The simulated annealing algorithm starts an initial selective maintenance schedule at a high initial
temperature. As the algorithm progresses, a new trial solution is generated by making a move from the current solution,
and after evaluating this new solution, a decision is taken to accept it or not. If the new solution is improved, i.e.
ΔCk < 0, the new solution is accepted. Otherwise, the new solution is accepted with the probability exp (−ΔCk/Tk ). Once
the criterion on the number of accepted policies is satisfied, the temperature is reduced by Tk+1 = ρTk, and the process is
repeated until it reaches the final temperature where the algorithm is terminated.
For the problem under consideration, the parameters of simulated annealing algorithm are defined as follows:

Figure 2. Flow chart of the simulated annealing algorithm.


7168 C. Duan et al.

 
DCk
PðSMprev ! SMnew Þ ¼ exp (15)
Tk

Tk ¼ qTk1 ; 0\q\1 (16)


where SMprev and SMnew denote previous selective maintenance schedule and new selective maintenance schedule,
respectively. ρ is the cooling rate and Tk is the temperature at the kth iteration. P(SMprev → SMnew) is acceptance proba-
bility of the new schedule. The acceptance probability of non-improving solutions decreases with the temperature. At
the start of the search, the temperature is relative high, and most of the worsening moves may be accepted. This helps
the algorithm jump out of a local minimum. However, at the end of the search, only improving solutions are likely to
be accepted. The main challenge for implementing simulated annealing algorithm to optimise the selective maintenance
scheduling problem is to define the solution trials. In our problem, the levels of maintenance actions for each component
are the solutions of the selective maintenance. At each iteration, the different levels of the maintenance actions with
respect to system constraints and specifications are assigned to the components randomly. A change in the levels of the
maintenance actions will result in the different expected average costs. If such a change reduces the expected average
cost, the new solution is accepted unconditionally. A non-improving solution may also be accepted with the probability
given by Equation (15).

7. Illustrative examples
7.1 Numerical examples
In this section, we consider multi-component systems for illustrative purpose. From the literature in reliability, the sys-
tem failure time is popularly described as the Weibull distribution (Najid, Alaoui-Selsouli, and Mohafid 2011; Wang,
Zhao, and Peng 2014; Wang, Wang, and Peng 2017). Thus, in our cases, lifetimes of components are assumed to be
independent, non-identical and follow the Weibull distribution with a cumulative distribution function in the form of

 t b
Fðt; a; bÞ ¼ 1  exp  ; t[0 (17)
a
where α and β are the scale and shape parameters, respectively. The planning horizon is assumed to be two years with
monthly break intervals.
Throughout this section, we considered the average cost of random failures to be 500 and average cost per unit time
of downtime caused by maintenance to be 50. For the intermediate maintenance, we considered three different levels of
maintenance actions whose average values of age reduction coefficients are 0.25, 0.50, 0.75, respectively. In the inspec-
tion and replacement cases, the average values of age reduction coefficients are 0 and 1, respectively. The standard devi-
ations of all the maintenance and inspection actions are set to be 0.01. The times needed to perform the three
intermediate maintenance actions and replacement are 0.5, 1, 1.5 and 2 time units for each component, respectively.
Moreover, for starting a new mission, the minimum required reliability is set to be 0.9. To show the generality of the
proposed model and solution procedure, two examples of different reliability block diagrams are presented which are
series-parallel system and k-out-of-n system.

7.1.1 A series-parallel system


We have considered a general series-parallel system from a published paper (Dao, Zuo, and Pandey 2014). The series-
parallel system consists of three series-connected subsystems in which comprise 3, 2 and 4 parallel-connected compo-
nents, respectively. The input parameters for the selective maintenance scheduling are provided in Table 1. To imple-
ment the simulated annealing algorithm for obtaining the optimal selective maintenance schedule, a computer code was
written in MATLAB software (Ver. 2015). The selective maintenance schedule was determined by maintaining the sys-
tem condition reliability R(m + 1|m) over the required level of 0.9 and the available time for maintenance should not
exceed 6 time units. Based on several trial runs, the best initial temperature and the cooling rates were found to be
60,000 and 0.95, respectively.
International Journal of Production Research 7169

Table 1. The input parameters for components in the series-parallel system.

Component No. 1 2 3 4 5 6 7 8 9

Scale parameter 2500 2500 2500 2500 2400 2400 2400 2400 2400
Shape parameter 2.5 2.5 2.5 2.5 2.4 2.4 2.4 2.4 2.4
Cost of maintenance 1 20 25 15 18 20 22 15 12 20
Cost of maintenance 2 45 50 40 42 40 45 35 24 40
Cost of maintenance 3 70 75 65 68 65 70 60 36 65
Cost of maintenance 4 95 100 85 92 90 95 80 50 95

In each period, the maintenance activities may vary for different components. To make sure that the final selective
maintenance schedule given by simulated annealing is, most probably, the global optimum, we have run the code from
10 different starting points. Given the sufficient searching times, similar results were obtained and the final selective
maintenance schedule for a 24-month period is presented in Table 2. In this table, symbol ‘0’ indicates inspection only, ‘1’,
‘2’, ‘3’ stand for intermediate maintenance 1, 2, 3, respectively, and ‘4’ denotes the replacement. In the whole planning time
horizon, the conditional reliability of completing the next mission is maintained greater than 0.9 and time spent for
maintenance is no more than 6. In the first two breaks, the conditional reliability remains at high level, required time for
maintenance is 0 and only inspection is suggested for all components. After the fourth scheduled break, the maintenance
actions become frequent as components age. These results reveal that under selective maintenance with multiple maintenance
options, the levels and frequencies of the maintenance actions differ for various components.
To evaluate the effects of the minimum reliability requirements on the related maintenance costs, we derived costs
of the optimal selective maintenance schedules for some selected values of minimal conditional reliability requirements
and presented them in Table 3. According to the results, when the conditional reliability requirement increases, the
downtime cost rises dramatically while the failure cost decreases. It can be observable that when the conditional reliabil-
ity requirement reaches to the value of 0.9, this trend becomes much more noticeable.

Table 2. Optimal selective maintenance schedule for the series-parallel system.

Component No.
Mission break 1 2 3 4 5 6 7 8 9 Conditional reliability at the end of each mission Required time

1 0 0 0 0 0 0 0 0 0 0.968 0
2 0 0 0 0 0 0 0 0 0 0.943 0
3 0 0 0 0 1 0 1 2 0 0.912 2
4 0 0 3 0 0 0 0 0 4 0.914 3.5
5 4 0 0 0 3 0 2 0 0 0.921 4.5
6 0 2 0 4 0 3 0 0 0 0.907 4.5
7 0 2 4 2 0 0 0 3 1 0.903 6
8 3 2 0 1 2 0 2 0 2 0.922 6
9 0 0 0 0 0 3 1 0 4 0.904 4
10 0 1 0 0 3 2 2 0 0 0.905 4
11 0 0 2 2 2 2 1 0 3 0.900 6
12 4 2 0 1 0 0 2 3 0 0.911 6
13 0 0 0 2 0 0 2 0 3 0.902 3.5
14 0 0 2 0 2 4 0 0 0 0.906 4
15 0 0 0 0 2 0 3 0 1 0.916 3
16 4 0 0 1 0 0 1 0 2 0.904 4
17 0 2 2 2 0 0 3 3 0 0.900 6
18 0 0 0 1 2 2 2 2 0 0.917 4.5
19 2 2 3 1 0 0 2 0 1 0.902 5.5
20 0 0 1 0 0 4 0 0 4 0.916 4.5
21 2 0 0 2 3 0 3 2 0 0.900 6
22 0 4 0 0 0 2 0 0 0 0.901 3
23 3 0 0 0 0 1 0 0 3 0.908 3.5
24 0 3 0 2 0 2 0 4 0 0.900 5.5
7170 C. Duan et al.

Table 3. The costs of the selective maintenance plans for various required conditional reliability levels.

Conditional reliability 0.75 0.80 0.85 0.90 0.95

Maintenance cost 1895 1986 2156 2784 4134


Failure cost 1832 1580 1320 1158 835
Downtime cost 3825 4175 4450 4975 6750
Total cost 7552 7741 7926 8917 11,719

Table 4. The input parameters for components in the k-out-of-n system.

Component No. 1 2 3 4 5

Scale parameter 2500 2500 2500 2400 2400


Shape parameter 2.5 2.5 2.5 2.4 2.4
Cost of maintenance 1 20 25 15 20 22
Cost of maintenance 2 45 50 40 40 45
Cost of maintenance 3 70 75 65 65 70
Cost of maintenance 4 95 100 85 90 95

We also investigated the various time limits of scheduled beaks under reliability requirement of 0.9, and the result is
shown in Figure 3. The result shows that the total cost reaches the minimal when the time limit is 6. When the time
limit is increased, the total cost will experience an increasing trend. That’s because the more time allocated to mainte-
nance activity will lead to huge downtime loss. However, when the time limit is equal to the value of 5, the total cost is
increased again due to the increase of maintenance cost. When the time constraint becomes stricter, the higher levels of
maintenance are more likely to be chosen as to satisfy the reliability requirement and to keep the safety (i.e. reduce the
failures).
In the example of the series-parallel system, the provided approach helps us find the most economic schedule within
the available maintenance time and reliability requirement. To further validate the proposed approach, we want to inves-
tigate what maintenance actions should be implemented when the system is a k-out-of-n type.

7.1.2 A consecutive k-out-of-n system


The consecutive k-out-of-n systems are used in various fields of applications such as oil pipelines and telecommunica-
tion systems. For illustrative purpose, a consecutive 2-out-of-5 system is presented. The input data for selective

Figure 3. The costs of the maintenance plans for various required maintenance times.
International Journal of Production Research 7171

Table 5. Optimal selective maintenance schedule for the k-out-of-n system.

Component No.
Mission break 1 2 3 4 5 Conditional reliability at the end of each mission Required time

1 0 0 0 0 0 0.972 0
2 0 0 0 1 0 0.948 0.5
3 0 0 2 0 2 0.922 2
4 2 0 3 0 0 0.910 2.5
5 3 1 0 3 0 0.913 3.5
6 0 2 0 0 3 0.904 2.5
7 0 2 4 0 0 0.913 3
8 4 2 0 2 0 0.918 4
9 2 0 0 0 3 0.914 2.5
10 0 4 0 3 2 0.908 4.5
11 0 0 2 2 2 0.905 3
12 4 2 0 0 0 0.908 3
13 1 0 4 0 2 0.901 3.5
14 0 0 2 2 4 0.905 3.5
15 2 0 0 2 0 0.915 2
16 4 0 1 0 2 0.903 3.5
17 0 2 2 1 0 0.901 2.5
18 0 0 0 2 2 0.919 2
19 2 3 3 0 0 0.904 4
20 0 0 1 0 4 0.915 2.5
21 2 0 0 3 0 0.901 2.5
22 0 4 0 0 2 0.904 3
23 3 0 3 0 1 0.912 3.5
24 0 3 0 0 2 0.903 2.5

scheduling is given in Table 4. The parameters are as same as previous example. The reliability requirement and avail-
able maintenance time are set to be 0.9 and 6 as well. The conditional reliability function of the system at time t can be
derived based on the block diagram.
Using the proposed algorithm, the optimal selective maintenance plan was derived and presented in Table 5. The
result demonstrates similar patterns as those in the previous example. We can observe that time spent for maintenance
in this case is less than the obtained value of the series-parallel system. This may be due to the configuration of the sys-
tem structure and a smaller number of components involved. Through two examples, the results prove the efficiency of the
proposed simulated annealing algorithm for solving selective maintenance scheduling problems of various multi-component
systems.

7.2 Case study


In this subsection, the presented approach is applied to an industrial case of production system which is composed of
12 parallel-connected chip mounting head units. The main function of the mounting head units is to achieve components
assembly by picking and placing the chips through the liner railways. Deterioration of the mounting head units will lead
to inaccurate position fixing, so the maintainers inspect the production system regularly at the end of each month. Upon
the inspection, three actions are optional to improve the condition of each mounting head unit, which are inspection

Table 6. The estimated parameters for the production system.

Component No. 1 2 3 4 5 6 7 8 9 10 11 12

Scale parameter 1250 1500 1540 1820 1400 1480 1320 1720 1200 1800 1350 1400
Shape parameter 2.8 3.3 3.5 3.9 2.4 3.2 2.8 2.6 3.4 2.5 4.0 3.3
Cost of maintenance 1 750 800 820 810 740 820 760 780 840 810 830 780
Cost of maintenance 2 1350 1400 1410 1405 1330 1420 1380 1400 1460 1410 1425 1390
7172 C. Duan et al.

Table 7. Optimal selective maintenance schedule for the production system.

Component No.
Mission break 1 2 3 4 5 6 7 8 9 10 11 12 Conditional reliability at the end of each mission Required time

1 0 0 0 0 0 0 0 0 0 0 0 0 0.972 0
2 0 0 0 0 0 0 0 0 0 0 0 0 0.957 0
3 0 0 0 0 0 0 0 0 0 0 1 0 0.932 0.25
4 0 0 1 0 0 0 0 0 1 0 0 1 0.925 0.75
5 1 0 0 0 1 0 1 0 0 0 1 0 0.920 1
6 0 1 0 1 0 1 0 0 0 1 0 0 0.912 1
7 0 0 1 0 0 0 0 2 0 0 0 0 0.914 0.75
8 0 0 0 0 1 0 1 0 2 0 0 0 0.904 1
9 1 0 0 0 0 0 0 0 0 0 0 2 0.908 0.75
10 0 1 0 0 0 1 0 0 0 0 1 0 0.911 0.5
11 0 0 1 2 0 0 0 0 0 1 0 0 0.913 1
12 1 0 0 0 0 0 1 2 0 0 0 0 0.912 1
13 0 0 0 0 2 0 0 0 0 0 0 1 0.907 0.75
14 0 1 0 0 0 1 0 0 0 0 1 0 0.911 0.75
15 0 0 1 0 0 0 0 0 1 0 0 0 0.905 0.75
16 1 0 0 1 0 0 1 1 0 0 0 0 0.908 1
17 0 1 0 0 0 0 0 0 0 1 0 1 0.902 0.75
18 0 0 0 0 2 0 0 0 0 0 2 0 0.907 1
19 1 0 2 0 0 0 0 0 0 0 0 0 0.911 0.75
20 0 0 0 0 0 2 0 0 2 0 0 0 0.913 1
21 0 0 0 2 0 0 2 0 0 0 0 0 0.905 1
22 0 1 0 0 0 0 0 1 0 1 0 1 0.906 1
23 1 0 0 0 1 1 0 0 0 0 1 0 0.909 1
24 0 0 1 2 0 0 0 0 1 0 0 0 0.905 1

(lubricating, adjusting/calibrating), intermediate maintenance (replacing railway with new one), replacement (replacing
mounting head).
According to the maintenance histories of the production system, the parameters of the mounting head units are
shown in Table 6, and average values and standard deviations of age reduction coefficients for the three maintenance
actions are 0 and 0.05, 0.6 and 0.1, 1 and 0.08, respectively. The times needed to perform intermediate maintenance
and replacement are 0.25 day and 0.5 day for each mounting head, respectively. The time available for
maintenance action are 1 day, and the minimum required reliability is 0.9.
We denoted the inspection, intermediate maintenance, and replacement as symbols ‘0’, ‘1’, and ‘2’ as well.
Table 7 shows the optimal selective maintenance schedule of the parallel production system for a 2-year length. It can
be seen that, despite the fact that system structure has changed, our presented approach still works effectively for
different struc-tures of multi-component systems.
We calculated the maintenance cost and time under the same constraints, when only one maintenance level, i.e.
replacement, is chosen at scheduled break in selective maintenance. We have found that the time spent for maintenance
is 22.5 and total maintenance cost is 87,600, which are greater than 18.75 and 69,400 obtained by multi-level mainte-
nance, respectively. This is because replacement is only performed at severe deterioration of the system, which might
not be efficient to prevent potential failure at early stage of deterioration. Multi-level maintenance combines the advan-
tages of different maintenance levels, and provides multiple options to the production system, where the most appropri-
ate maintenance can be done according to the system actual status under time and reliability constraints in selective
maintenance scheduling.

7.2.1 Comparison with previous approach


To see the performance of provided optimisation technique, we first compared our simulated annealing with the most
used branch and bound algorithm upon the three mentioned cases of multi-component systems. To apply the branch and
bound algorithm in our selective maintenance framework, Shih (1979) showed that the optimisation problem in Equa-
tions (11)–(14) can be rewritten as
International Journal of Production Research 7173
0 1
XLi
M N X
M X
N Li
min vLð1Þ;Lð2Þ XLð1Þ;Lð2Þ þ YLð1Þ;Lð2Þ  @ D XLð1Þ;Lð2Þ A þ C
Lð1Þ¼1 Lð1Þ¼1 Lð1Þ¼1

subject to:

Rðm þ 1jmÞ  Rmin ; 8m 1  m  M


NP
Li
aLð1Þ;Lð2Þ XLð1Þ;Lð2Þ  Tmax ; 8Lð1Þ; Lð2Þ 1  Lð1Þ  N  Li ; 1  Lð1Þ  N  Li
Lð1Þ¼1
P
Li
aLð1Þ;Lð2Þ XLð1Þ;Lð2Þ  1; 8Lð1Þ; Lð2Þ 1  Lð1Þ  Li ; 1  Lð1Þ  Li
Lð1Þ¼1

where XL(1),L(2) and YL(1),L(2) are relabelled variables. XL(1),L(2) = 0, 1; YL(1),L(2) = 0, 1. L(1) and L(2) indicate the respec-
tive orders of magnitude of the ratio of variable coefficient to constraint coefficient defined in P Equations (11)-(14). vL(1),
L(2) and aL(1),L(2) are the coefficients of the relabeled variable and constraint, respectively. C ¼ m¼1 KðmÞMCf .
M

Following the procedures in Shih (1979), we solved the brand and bound problem formulated above using three
steps as follows:
Step 1: Establishing the starting node
Compute optimal fractional solutions provided in Dantzig (1957) for each single constraint separately. Find the
maximum value of objective function using these solutions. The solution corresponding to maximum objective func-
tion is represented by the variables contained in Fd(k), where d records the constraint number corresponding to the
solution, and k denotes the node iteration number.
Step 2: Testing the optimality and selecting a variable for next branching.
Based on the feasibility test of solution, if the solution satisfies all the constraints, it is also an optimal solution,
stop; if not, select a variable from the Fd(k) such that the value of the dth subscript of that variable is the largest
among that of all variables in the same set, call it XL(1)*.
Step 3: Computing lower bounds
Set k = k + 1, XL(1)* = 0, then test the feasibility of the new solution, if not, go to step 2. Set k = k + 1, XL(1)
* = 1, then test the feasibility of the new solution, if not, go to step 2.

For more details about this algorithm, reader can refer to Shih (1979). This branch and bound method has been
coded in MATLAB software (Ver. 2015). Solution time in minutes and the three mentioned cases (i.e. series-parallel
system, k-out-of-n system, and parallel system) over each of the two computer algorithms were recorded in detail in
Table 8. According to the results, for the three different system structures, the superiority of the simulated annealing
over branch and bound method is clearly evident with regard to solution time and the optimal value.
To make the comparison strong, we randomly generated system scale parameters using a discrete uniform distribu-
tion between 1000 and 3000, and shape parameters using a discrete uniform distribution between 2 and 4.
The generated systems are in three different structures (i.e. series-parallel system, k-out-of-n system, and parallel system)
with increasing order of component number. These test results in Table 9 show that the branch and bound method is
extremely inefficient for multi-component system maintenance scheduling problems and no optimal solutions can be
reached within 60 min of computer time when the components number exceeds 10. The proposed simulated annealing
is much more stable and computational time-saving in selective maintenance scheduling with a larger number of compo-
nents, which is very attractive in practical applications.

Table 8. Comparison of two algorithms.

Solution time (min) Optimal solution


Example Simulated annealing Branch and bound Simulated annealing Branch and bound

Series-parallel system 1.2 48 8917 8917


k-out-of-n system 0.6 24 6226 6226
Parallel production system 1.5 55 69,400 69,400
7174 C. Duan et al.

Table 9. Comparison of generated cases.

Solution time (min) Optimal solution


Example Simulated annealing Branch and bound Simulated annealing Branch and bound

Series-parallel system (1,1,1)1 0.1 22 3232 3232


Series-parallel system (2,2,2) 0.6 28 6534 6534
Series-parallel system (4,4,4) 1.1 58 8956 8956
Series-parallel system (6,6,6) 1.8 >60 – 19,323
Series-parallel system (7,7,7) 2.2 >60 – 22,323
Series-parallel system (8,8,8) 2.6 >60 – 25,034
Series-parallel system (9,9,9) 4.2 >60 – 28,093
Series-parallel system (10,10,10) 6.2 >60 – 30,321
k-out-of-n system (1,3)2 0.3 19 4350 4350
k-out-of-n system (2,4) 0.5 23 5324 5324
k-out-of-n system (5,8) 1.6 46 8784 8784
k-out-of-n system (6,9) 1.9 58 9432 9432
k-out-of-n system (7,10) 2.6 >60 – 10,327
k-out-of-n system (8,11) 3.2 >60 – 11,032
k-out-of-n system (9,12) 3.9 >60 – 12,876
Parallel system (2)3 0.04 2 9235 9235
Parallel system (4) 0.2 8 18,946 18,946
Parallel system (8) 0.8 36 35,066 35,066
Parallel system (16) 2.5 >60 – 63,646
Parallel system (17) 2.9 >60 – 66,097
Parallel system (18) 3.2 >60 – 69,460
Parallel system (19) 4.5 >60 – 73,400
Parallel system (20) 5.6 >60 – 79,860
1
System structure is organised with 1,1,1 parallel-connected components;
2
1-out-of-3 system;
3
Parallel system with 2 components;
– No optimal solutions can be reached within 60 min of computer time.

Figure 4. Comparison of stochastic maintenance quality and deterministic maintenance quality in conditional reliability.
International Journal of Production Research 7175
Table 10. The stochastic maintenance quality case using maintenance plan derived from deterministic model.

Mission 1 2 3 4 5 6 7 8 9 10 11 12

Conditional reliability 0.972 0.957 0.928 0.922 0.920 0.906 0.910 0.898 0.895 0.906 0.910 0.914
Mission 13 14 15 16 17 18 19 20 21 22 23 24
Conditional reliability 0.919 0.921 0.915 0.918 0.926 0.912 0.915 0.925 0.912 0.913 0.918 0.911

Figure 5. Total maintenance cost for different expected age reduction coefficients.

7.2.2 Comparison with deterministic case


To investigate the impact of the random maintenance quality on the maintenance schedule, we compared our mainte-nance
example (i.e. production system) with case in which the quality of maintenance action is deterministic. All the parameters
are kept the same. The age reduction coefficients of deterministic maintenance are set to be the expected values of their
stochastic counterparts. Using Equation (9), the conditional reliability that system successfully completes the next mission
both at the end of current mission and after maintenance actions can be calculated, and the results are depicted in Figure 4.
The results demonstrate that the optimal schedules for those two types of maintenance quality are completely different. At
the beginning, the conditional reliability is greater than that of maintenance with deterministic quality. After the 11th mission
break, the conditional reliability of deterministic case becomes higher than the stochastic case. In the later period of
scheduling, the higher levels of maintenance actions are performed to improve the conditional reliability level of the
deterministic case. In such a case, a maintenance planning may not necessarily maintain the conditional reli-ability level too
high above the minimum reliability level.
If deterministic age reduction coefficient is used to determine the optimal selective maintenance in case study of the
production system, the maintenance schedule might not be optimal and even not feasible for practical applications.
Table 10 shows that conditional reliability at the 8th and 9th mission breaks have not met the minimum requirement for
safe production when stochastic maintenance quality is assumed to be deterministic in scheduling optimisation.
When the expected age reduction coefficient is small, the total maintenance costs for both stochastic and determinis-
tic models have very small difference, as shown in the sensitivity analysis provided in Figure 5. When the expected age
reduction coefficient is increased to the value of 0.5, this cost difference becomes much more significant.
Through this comparison, we have two conclusions. The first one is that the quality of the maintenance would
impact the evaluation of conditional reliability and maintenance schedule. The second one is that selective maintenance
planning under the considered stochastic maintenance quality in this paper leads to considerable cost saving especially
when the expected maintenance improvement is high.
7176 C. Duan et al.

8. Conclusions and future research


This paper addresses a selective maintenance scheduling problem under stochastic maintenance quality with multiple
maintenance options. The quality of maintenance is assumed to be random and follow a Beta distribution defined on interval
[0, 1]. The maintenance cost and time consumed are constructed according to the corresponding maintenance quality to
reflect the improvement effect. The system is operating to perform a set of missions with limited break durations, where the
optimum maintenance actions should be determined. At each scheduled break, various maintenance levels, which have
different effects on component’s age, are available to ensure the reliability of completing the next mission remaining at
desirable level. The optimisation problem involves determining the best set of maintenance actions for components at each
break period with respect to the minimum total cost, time constraints, and reliability requirement. Simulated annealing
algorithm is employed to solve the resulting optimisation problem, and a comparison is given between the proposed model
and the deterministic maintenance model where the maintenance quality is constant. The result demonstrates the importance
of considering the stochastic version of selective maintenance scheduling problem.
In general, the provided selective maintenance model in this paper can help the maintenance manager determine the
best maintenance strategy to get a reliable system and allocate the resources effectively. With some minor modifications,
the proposed approach can be adapted to solve any types of selective maintenance problem defined in Cassady, Pohl,
and Paul Murdock (2001b). Further extension of the model to consider different failure consequences, i.e. soft failure
and hard failure, in selective maintenance scheduling is a suitable topic for future research. Another interesting future
research topic could be the development of random or non-periodic mission interval in selective maintenance model
which takes into account stochastic dependence among components.

Acknowledgements
The authors would like to express their gratitude to associate editor and three anonymous referees for their valuable comments and
suggestions which contributed to a significant improvement of the original version of this paper.

Disclosure statement
No potential conflict of interest was reported by the authors.

Funding
The research was supported by China Scholarship Council (CSC) [grant number 201506160096]; the National Natural Science Foun-
dation of China (NSFC) [grant number 51475189]; the National Key Research and Development Program of China [grant number
2016YFE0121700]; the Science and Technology Development Fund of Macao S.A.R (FDCT) under MoST-FDCT [joint grant number
015/2015/AMJ].

References

Alaswad, S., and Y. Xiang. 2017. “A Review on Condition-Based Maintenance Optimization Models for Stochastically Deteriorating
System.” Reliability Engineering & System Safety 157: 54–63.
Batun, S., and M. Azizoğlu. 2009. “Single Machine Scheduling with Preventive Maintenances.” International Journal of Production
Research 47 (7): 1753–1771.
Ben Mabrouk, A., A. Chelbi, and M. Radhoui. 2016. “Optimal Imperfect Preventive Maintenance Policy for Equipment Leased dur-
ing Successive Periods.” International Journal of Production Research 54 (17): 5095–5110.
Bertsimas, D., and J. Tsitsiklis. 1993. “Simulated Annealing.” Statistical Science 8 (1): 10–15.
Breedam, A. 2001. “Comparing Descent Heuristics and Metaheuristics for the Vehicle Routing Problem.” Computers & Operations
Research 28 (4): 289–315.
Cassady, C. R., W. P. Murdock, and E. A. Pohl. 2001a. “Selective Maintenance for Support Equipment Involving Multiple Mainte-
nance Actions.” European Journal of Operational Research 129 (2): 252–258.
Cassady, C. R., E. A. Pohl, and W. Paul Murdock. 2001b. “Selective Maintenance Modeling for Industrial Systems.” Journal of Qual-
ity in Maintenance Engineering 7 (2): 104–117.
Dantzig, G. B. 1957. “Discrete-Variable Extremum Problems.” Operations Research 5 (2): 266–288.
Dao, C. D., M. J. Zuo, and M. Pandey. 2014. “Selective Maintenance for Multi-State Series-Parallel Systems under Economic Depen-
dence.” Reliability Engineering & System Safety 121: 240–249.
Dijoux, Y., M. Fouladirad, and D. T. Nguyen. 2016. “Statistical Inference for Imperfect Maintenance Models with Missing Data.”
Reliability Engineering & System Safety 154: 84–96.
International Journal of Production Research 7177
Doostparast, M., F. Kolahan, and M. Doostparast. 2014. “A Reliability-Based Approach to Optimize Preventive Maintenance Schedul-
ing for Coherent Systems.” Reliability Engineering & System Safety 126: 98–106.
Duan, C., C. Deng, and B. Wang. 2017a. “Optimal Multi-Level Condition-Based Maintenance Policy for Multi-Unit Systems under
Economic Dependence.” The International Journal of Advanced Manufacturing Technology. 91: 4299–4312. doi:10.1007/
s00170-017-0100-0.
Duan, C., C. Deng, and B. Wang. 2017b. “Multi-Phase Sequential Preventive Maintenance Scheduling for Deteriorating Repairable
Systems.” Journal of Intelligent Manufacturing.: 1–15. doi:10.1007/s10845-017-1353-z.
Khatab, A., and E. H. Aghezzaf. 2016. “Selective Maintenance Optimization When Quality of Imperfect Maintenance Actions Are
Stochastic.” Reliability Engineering & System Safety 150: 182–189.
Khatab, A., E. H. Aghezzaf, C. Diallo, and I. Djelloul. 2017. “Selective Maintenance Optimisation for Series-Parallel Systems Alter-
nating Missions and Scheduled Breaks with Stochastic Durations.” International Journal of Production Research 55 (10):
3008–3024.
Koochaki, J., J. A. Bokhorst, H. Wortmann, and W. Klingenberg. 2012. “Condition Based Maintenance in the Context of Opportunis-
tic Maintenance.” International Journal of Production Research 50 (23): 6918–6929.
Liu, Y., and H. Z. Huang. 2010. “Optimal Selective Maintenance Strategy for Multi-State Systems under Imperfect Maintenance.”
IEEE Transactions on Reliability 59 (2): 356367.
Liu, J., G. Li, D. Chen, W. Liu, and Y. Wang. 2010. “Two-Dimensional Equilibrium Constraint Layout Using Simulated Annealing.”
Computers & Industrial Engineering 59 (4): 530–536.
Liu, Y., Y. Li, H. Z. Huang, and Y. Kuang. 2011. “An Optimal Sequential Preventive Maintenance Policy under Stochastic Mainte-
nance Quality.” Structure and Infrastructure Engineering 7 (4): 315–322.
Liu, B., M. Xie, Z. Xu, and W. Kuo. 2016. “An Imperfect Maintenance Policy for Mission-Oriented Systems Subject to Degradation
and External Shocks.” Computers & Industrial Engineering 102: 21–32.
Najid, N. M., M. Alaoui-Selsouli, and A. Mohafid. 2011. “An Integrated Production and Maintenance Planning Model with Time
Windows and Shortage Cost.” International Journal of Production Research 49 (8): 2265–2283.
Nakagawa, T. 1979. “Optimum Policies When Preventive Maintenance is Imperfect.” IEEE Transactions on Reliability R-28 (4):
331–332.
Nakagawa, T. 2005. Maintenance Theory of Reliability. London: Springer Science & Business Media.
Nodem, F. D., A. Gharbi, and J. P. Kenné. 2011. “Preventive Maintenance and Replacement Policies for Deteriorating Production
Systems Subject to Imperfect Repairs.” International Journal of Production Research 49 (12): 3543–3563.
Pandey, M., M. J. Zuo, R. Moghaddass, and M. Tiwari. 2013. “Selective Maintenance for Binary Systems under Imperfect Repair.”
Reliability Engineering & System Safety 113: 42–51.
Pandey, M., M. J. Zuo, and R. Moghaddass. 2013. “Selective Maintenance Modeling for a Multistate System with Multistate Compo-
nents under Imperfect Maintenance.” IIE Transactions 45 (11): 1221–1234.
Pedamallu, C. S., and L. Ozdamar. 2008. “Investigating a Hybrid Simulated Annealing and Local Search Algorithm for Constrained
Optimization.” European Journal of Operational Research 185 (3): 1230–1245.
Pham, H., and H. Wang. 1996. “Imperfect Maintenance.” European Journal of Operational Research 94 (3): 425–438.
Rasmekomen, N., and A. K. Parlikad. 2016. “Condition-Based Maintenance of Multi-Component Systems with Degradation State-
Rate Interactions.” Reliability Engineering & System Safety 148: 1–10.
Rice, W., C. Cassady, and J. Nachlas. 1998. “Optimal Maintenance Plans under Limited Maintenance Time.” Proceedings of the
Seventh Industrial Engineering Research Conference, Alberta, Canada, 1–3.
Şahin, R., K. Ertoğral, and O. Türkbey. 2010. “A Simulated Annealing Heuristic for the Dynamic Layout Problem with Budget Con-
straint.” Computers & Industrial Engineering 59 (2): 308–313.
Saraiva, J. T., M. L. Pereira, V. T. Mendes, and J. C. Sousa. 2011. “A Simulated Annealing Based Approach to Solve the Generator
Maintenance Scheduling Problem.” Electric Power Systems Research 81 (7): 1283–1291.
Shih, W. 1979. “A Branch and Bound Method for the Multiconstraint Zero-One Knapsack Problem.” Journal of the Operational
Research Society 30 (4): 369–378. doi:10.2307/3009639.
Shim, S. O. 2009. “Generating Subproblems in Branch and Bound Algorithms for Parallel Machines Scheduling Problem.” Comput-
ers & Industrial Engineering 57 (3): 1150–1153.
Tsai, Y. T., K. S. Wang, and L. C. Tsai. 2004. “A Study of Availability-Centered Preventive Maintenance for Multi-Component Sys-
tems.” Reliability Engineering & System Safety 84 (3): 261–270.
Tsai, T. R., P. H. Liu, and Y. Lio. 2011. “Optimal Maintenance Time for Imperfect Maintenance Actions on Repairable Product.”
Computers & Industrial Engineering 60 (4): 744–749.
Wang, W., F. Zhao, and R. Peng. 2014. “A Preventive Maintenance Model with a Two-Level Inspection Policy Based on a Three-
Stage Failure Process.” Reliability Engineering & System Safety 121: 207–220.
Wang, C., D. Mu, F. Zhao, and J. W. Sutherland. 2015. “A Parallel Simulated Annealing Method for the Vehicle Routing Problem
with Simultaneous Pickup-Delivery and Time Windows.” Computers & Industrial Engineering 83: 111–122.
Wang, H., W. Wang, and R. Peng. 2017. “A Two-Phase Inspection Model for a Single Component System with Three-Stage Degrada-
tion.” Reliability Engineering & System Safety 158: 31–40.
7178 C. Duan et al.

Wong, C. S., F. T. S. Chan, and S. H. Chung. 2012. “A Genetic Algorithm Approach for Production Scheduling with Mould Mainte-
nance Consideration.” International Journal of Production Research 50 (20): 5683–5697.
Wu, S., and D. Clements-Croome. 2005. “Preventive Maintenance Models with Random Maintenance Quality.” Reliability Engineer-
ing & System Safety 90 (1): 99–105.
Xia, T., L. Xi, X. Zhou, and S. Du. 2012. “Modeling and Optimizing Maintenance Schedule for Energy Systems Subject to Degrada-
tion.” Computers & Industrial Engineering 63 (3): 607–614.
Xia, T., L. Xi, X. Zhou, and J. Lee. 2013. “Condition-Based Maintenance for Intelligent Monitored Series System with Independent
Machine Failure Modes.” International Journal of Production Research 51 (15): 4585–4596.
Xia, T., X. Jin, L. Xi, and J. Ni. 2015. “Production-Driven Opportunistic Maintenance for Batch Production Based on MAM–APB
Scheduling.” European Journal of Operational Research 240 (3): 781–790.
Xia, T., L. Xi, E. Pan, and J. Ni. 2016. “Reconfiguration-Oriented Opportunistic Maintenance Policy for Reconfigurable Manufactur-
ing Systems.” Reliability Engineering & System Safety 166: 87–98.
Xia, T., X. Tao, and Li-Feng Xi. 2017. “Operation Process Rebuilding (OPR)-Oriented Maintenance Policy for Changeable System
Structures.” IEEE Transactions on Automation Science and Engineering 14 (1): 139–148.
Xia, T., L. Xi, E. Pan, X. Fang, and N. Gebraeel. 2017. “Lease-Oriented Opportunistic Maintenance for Multi-Unit Leased Systems
under Product-Service Paradigm.” ASME Journal of Manufacturing Science and Engineering 139 (7): 071005-1–071005-10.
Zhang, M., O. Gaudoin, and M. Xie. 2015. “Degradation-Based Maintenance Decision Using Stochastic Filtering for Systems under
Imperfect Maintenance.” European Journal of Operational Research 245 (2): 531–541.
Zhao, X., and M. Xie. 2017. “Using Accelerated Life Tests Data to Predict Warranty Cost under Imperfect Repair.” Computers &
Industrial Engineering 107: 223–234.
Zhu, H., F. Liu, X. Shao, Q. Liu, and Y. Deng. 2011. “A Cost-Based Selective Maintenance Decision-Making Method for Machining
Line.” Quality and Reliability Engineering International 27 (2): 191–201.

You might also like