Skip to main content
Log in

A condition-based maintenance policy for multi-component systems with a high maintenance setup cost

  • Regular Article
  • Published:
OR Spectrum Aims and scope Submit manuscript

Abstract

Condition-based maintenance (CBM) is becoming increasingly important due to the development of advanced sensor and information technology, which facilitates the remote collection of condition data. We propose a new CBM policy for multi-component systems with continuous stochastic deteriorations. To reduce the high setup cost of maintenance, a joint maintenance interval is introduced. With this interval and the control limits of components as decision variables, we develop a model for the minimization of the average long-run maintenance cost rate of the systems. Moreover, a numerical study of a production system consisting of a large number of non-identical components is presented, including a sensitivity analysis. Finally, our policy is compared to a failure-based policy and an age-based policy, in order to evaluate the cost-saving potential.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  • Barata J, Soares CG, Marseguerra M, Zio E (2002) Simulation modelling of repairable multi-component deteriorating systems for ‘on condition’ maintenance optimisation. Reliab Eng Syst Saf 76(3):255–264

    Article  Google Scholar 

  • Berg M, Epstein B (1976) Modified block replacement policy. Naval Res Logist 23(1):15–24

    Article  Google Scholar 

  • Bouvard K, Artus S, Berenguer C, Cocquempot V (2011) Condition-based dynamic maintenance operations planning & grouping: application to commercial heavy vehicles. Reliab Eng Syst Saf 96(6, SI):601–610

    Article  Google Scholar 

  • Callister WD, Rethwisch DG (2003) Materials science and engineering: an introduction, 6th edn. Wiley, New York

    Google Scholar 

  • Castanier B, Grall A, Berenguer C (2005) A condition-based maintenance policy with non-periodic inspections for a two-unit series system. Reliab Eng Syst Saf 87(1):109–120

    Article  Google Scholar 

  • Cho DI, Parlar M (1991) A survey of maintenance models for multiunit systems. Eur J Oper Res 51(1):1–23

    Article  Google Scholar 

  • Daniel C (1973) One-at-a-time plans. J Am Stat Assoc 68(342):353–360

    Article  Google Scholar 

  • Dekker R, Wildeman RE, van der Schouten FA (1997) A review of multi-component maintenance models with economic dependence. Math Methods Oper Res 45(3):411–435

    Article  Google Scholar 

  • Dieulle L, Berenguer C, Grall A, Roussignol M (2003) Sequential condition-based maintenance scheduling for a deteriorating system. Eur J Oper Res 150(2):451–461

    Article  Google Scholar 

  • Elwany AH, Gebraeel NZ, Maillart LM (2011) Structured replacement policies for components with complex degradation processes and dedicated sensors. Oper Res 59(3):684–695

    Article  Google Scholar 

  • Gebraeel NZ, Lawley MA, Li R, Ryan JK (2005) Residual-life distributions from component degradation signals: a Bayesian approach. IIE Trans 37(6):543–557

    Article  Google Scholar 

  • Grall A, Berenguer C, Dieulle L (2002) A condition-based maintenance policy for stochastically deteriorating systems. Reliab Eng Syst Saf 76(2):167–180

    Article  Google Scholar 

  • Haurie A, L’ecuyer P (1982) A stochastic-control approach to group preventive replacement in a multicomponent system. IEEE Trans Autom Control 27(2):387–393

    Article  Google Scholar 

  • Jardine AKS, Lin D, Banjevic D (2006) A review on machinery diagnostics and prognostics implementing condition-based maintenance. Mech Syst Signal Process 20(7):1483–1510

    Article  Google Scholar 

  • Jiang R, Kim MJ, Makis V (2013) Availability maximization under partial observations. OR Spectr 35(3):691–710

    Article  Google Scholar 

  • Kharoufeh JP, Solo CJ, Ulukus M (2010) Semi-markov models for degradation-based reliability. IIE Trans 42(8):599–612

    Article  Google Scholar 

  • Liao HT, Elsayed A, Chan LY (2006) Maintenance of continuously monitored degrading systems. Eur J Oper Res 175(2):821–835

    Article  Google Scholar 

  • Lin DM, Makis V (2003) Recursive filters for a partially observable system subject to random failure. Adv Appl Probab 35(1):207–227

    Article  Google Scholar 

  • Lu CJ, Meeker WQ (1993) Using degradation measures to estimate a time-to-failure distribution. Technometrics 35(2):161–174

    Article  Google Scholar 

  • Marseguerra M, Zio E, Podofillini L (2002) Condition-based maintenance optimization by means of genetic algorithms and Monte Carlo simulation. Reliab Eng Syst Saf 77(2):151–165

    Article  Google Scholar 

  • Nicolai RP, Dekker R (2007) A review of multi-component maintenance models. In: Risk, reliability and societal safety, Proceedings and monographs in engineering, water and earth sciences vol 1–3, pp 289–296

  • Ozekici S (1988) Optimal periodic replacement of multicomponent reliability systems. Oper Res 36(4):542–552

    Article  Google Scholar 

  • Park KS (1988) Optimal continuous-wear limit replacement under periodic inspections. IEEE Trans Reliab 37(1):97–102

    Article  Google Scholar 

  • Peng Y, Dong M, Zuo MJ (2010) Current status of machine prognostics in condition-based maintenance: a review. Int J Adv Manuf Technol 50(1–4):297–313

    Article  Google Scholar 

  • Radner R, Jorgenson DW (1963) Opportunistic replacement of a single part in the presence of several monitored parts. Manag Sci 10(1):70–84

    Article  Google Scholar 

  • Si XS, Wang W, Hu CH, Zhou DH (2011) Remaining useful life estimation–a review on the statistical data driven approaches. Eur J Oper Res 213(1):1–14

    Article  Google Scholar 

  • Tian ZG, Jin TD, Wu BR, Ding FF (2011) Condition based maintenance optimization for wind power generation systems under continuous monitoring. Renew Energy 36(5):1502–1509

    Article  Google Scholar 

  • Tian ZG, Liao HT (2011) Condition based maintenance optimization for multi-component systems using proportional hazards model. Reliab Eng Syst Saf 96(5):581–589

    Article  Google Scholar 

  • van der Schouten FA, Vanneste SG (1993) Two simple control policies for a multicomponent maintenance system. Oper Res 41(6):1125–1136

    Article  Google Scholar 

  • van Dijkhuizen G, van Harten A (1997) Optimal clustering of frequency-constrained maintenance jobs with shared set-ups. Eur J Oper Res 99(3):552–564

    Article  Google Scholar 

  • van Noortwijk JM (2009) A survey of the application of gamma processes in maintenance. Reliab Eng Syst Saf 94(1, SI):2–21

    Article  Google Scholar 

  • Vlok PJ, Coetzee JL, Banjevic D, Jardine AKS, Makis V (2002) Optimal component replacement decisions using vibration monitoring and the proportional-hazards model. J Oper Res Soc 53(2):193–202

    Article  Google Scholar 

  • Wang W (2000) A model to determine the optimal critical level and the monitoring intervals in condition-based maintenance. Int J Prod Res 38(6):1425–1436

    Article  Google Scholar 

  • Wang W, Christer AH (2000) Towards a general condition based maintenance model for a stochastic dynamic system. J Oper Res Soc 51(2):145–155

    Article  Google Scholar 

  • Wijnmalen DJD, Hontelez JAM (1997) Coordinated condition-based repair strategies for components of a multi-component maintenance system with discounts. Eur J Oper Res 98(1):52–63

    Article  Google Scholar 

  • Wildeman RE, Dekker R, Smit ACJM (1997) A dynamic policy for grouping maintenance activities. Eur J Oper Res 99(3):530–551

    Article  Google Scholar 

  • Zhu Q, Peng H, Houtum GJJAN (2013) An opportunistic maintenance policy for components under condition monitoring in complex systems. In: Working paper, BETA research school. Eindhoven University of Technology, Eindhoven

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hao Peng.

Appendices

Appendix A: Description of two comparison policies

1.1 Failure-based policy

When the degradation of one component \(X_i(t)\) in the system reaches \(H_i\), a CM action is taken. For each component \(i\in I\), the failure-based policy implies that there is no PM action taken, so that no control limit \(C_i\) is set on the degradation before \(H_i\) is reached (see Fig. 1). Or equivalently, \(C_i=H_i\). The optimization algorithm of our model in Sect. 3.2 remains unchanged in essence. Equations (6), (7), (11), (13) and (12) are derived as follows,

$$\begin{aligned} \hbox {Pr}\{\mathrm{PM} \ \hbox {at} \ n\tilde{\tau }\}&= 0 \\ \hbox {Pr}\{\mathrm{CM} \ \hbox {at} \ n\tilde{\tau }\}&= \hbox {Pr}\{(n-1)\tilde{\tau } \le T_{H_i} < n \tilde{\tau } \} \\ \end{aligned}$$

1.2 Age-based policy

Unlike the failure-based policy, PM actions are taken at joint maintenance time point according to a threshold \(A_i\) on the age of component i. It is almost the same as our proposed policy in Sect. 2, except the ages of components are observed, instead of the condition. Notice the assumptions in Sect. 2.2 are also valid. Since PM actions are taken at a joint maintenance time point to save setup costs, the decision variable \(A_i\) should be a multiple of \(\tilde{\tau }\), i.e., \(A_i=k_i \tilde{\tau }\). Hence, if there is no failure before \(A_i\), a PM action will be performed at \(A_i\) which is also a joint maintenance point. Otherwise, if there is a failure, a CM action will be performed at the next closest joint maintenance point, similarly to the maintenance policy proposed in this paper. The optimization algorithm of this age-based policy is also similar to the one proposed in Sect. 3.2, except Eqs. (11)–(13) are derived as follows,

$$\begin{aligned} \mathbb {E}\left[ K_i(\tilde{\tau }, A_i) \right]= & {} \int _{k_i \tilde{\tau }}^{\infty }{ f_{T_{H_i}}(x)\,\hbox {d}x} \ C_{\mathrm{PM},i} + \int _{0}^{k_i \tilde{\tau }}{ f_{T_{H_i}}(x)\,\hbox {d}x} \ C_{\mathrm{CM},i}\nonumber \\&+\,\mathbb {E}\left[ D_i(\tilde{\tau }, A_i)\right] C_{p,i} ,\\ \mathbb {E}\left[ L_i(\tilde{\tau }, A_i) \right]= & {} \int _{k_i \tilde{\tau }}^{\infty }{k_i \tilde{\tau } f_{T_{H_i}}(x)\,\hbox {d}x} + \sum _{n=1}^{k_i} \int _{(n-1)\tilde{\tau }}^{n\tilde{\tau }} {n \tilde{\tau } f_{T_{H_i}}(x)\,\hbox {d}x} ,\\ \mathbb {E}\left[ D_i(\tilde{\tau }, A_i) \right]= & {} \sum _{n=1}^{k_i} \int _{(n-1)\tilde{\tau }}^{n\tilde{\tau }} {(n \tilde{\tau } -x) f_{T_{H_i}}(x)\,\hbox {d}x} , \end{aligned}$$

and \(f_{T_{H_i}}(x)\) is the probability density function of the failure time [\(C_i=H_i\) in Eq. (14)]. Notice that the distribution of the failure time is the same as the distribution of the passage time of \(H_i\), because a soft failure occurs when the degradation process crosses the threshold \(H_i\).

Appendix B: The average cost rate of single component over two decision variables \(C_i\) and \(\tau \)

To show how the objective function varies with two decision variables \(C_i\) and \(\tau \), we plot the the average cost rate of component 1, which is a function of both \(C_1\) and \(\tau \), in Fig. 5 as an example.

Fig. 5
figure 5

Average cost rate on component 1 over \(\tau \) and \(C_1\) (left 3D plot; right contour plot)

Appendix C: Optimization algorithm

The procedure of the nested enumeration algorithm can be summarized in Algorithm 1.

figure a

Notice different grid sizes can be used for optimizing \(C_i\) and \(\tau \) , which will also affect the computational duration. In this paper, we use the grid size \(H_i/500\) and \(M_{\tau }/500\) for \(C_i\) and \(\tau \) respectively. The upper bound \(M_{\tau }\) is a very large value (at least larger than \(\max _{i\in I}\{G_i\}\).). In this paper, we choose \(M_{\tau }=300 \) days.

To determine the grid sizes for \(C_i\) and \(\tau \) on \(H_i\) and \(M_{\tau }\), we suggest the decision makers first to select a large grid size, e.g., \(H_i/100\) and \(M_{\tau }/100\), in order to have a brief overview of the objective function. Then if the company is sensitive to the cost difference between different sub-optimal solutions incurred by the grid sizes, a smaller grid size can be set for searching the optimal \(C_i\) and \(\tau \). Notice that while changing the grid sizes, we should also observe the changes of the objective function. If the objective function is not sensitive to the changes of the grid sizes, we can stop further decreasing the grid sizes.

Appendix D: Computation performance

Instead of optimizing \(C_i(\tau )\) and \(\tau \) simultaneously, we used a nested approach. Namely, we (i) optimize \(C_i(\tau )\) for each component under a given \(\tau \) and then (ii) optimize \(\tau \) for the system. The motivation of such a decomposition is to reduce the computation time of large-scale problems. When the amount of components in a system is large, the solution space of decision variables increases dramatically, also known as “curse-of-dimensionality”.

For example, a system consisting of two components (\(i\in \{1,2\}\)) is considered in our optimization model. For each component, we have to optimize the \(C_i(\tau ) \in (0, H)\). Suppose we discretise the degradation range (0, H) into 10 grids with a grid size H / 10. The size of the solution space \((C_1,C_2)\) at a given \(\tau \) value is \(10^2\). In the case of this two-component system, it is plausible to optimize \(\tau \) and \(C_i\) simultaneously. However, if a system consists of 50 components, then the size of its solution space will be \(10^{50}\) under each given \(\tau \), which is nearly impossible to solve within a short period. Therefore, it is not efficient to optimize \(\tau \) and \(C_i\) simultaneously. To solve such a large-scale problem within a reasonable computation time, we propose a nested approach to decompose the problem at system level into component level (see Sect. 3.2). This approach will reduce the solution space to \(10\times 50\) under a given \(\tau \). Regarding the numerical example in Sect. 4, the code is built in MATLAB with the runtime of \(4.6\times 10^3\) seconds (by a computer with a 2.5 GHz processor and 4 G RAM).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhu, Q., Peng, H. & van Houtum, GJ. A condition-based maintenance policy for multi-component systems with a high maintenance setup cost. OR Spectrum 37, 1007–1035 (2015). https://doi.org/10.1007/s00291-015-0405-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00291-015-0405-z

Keywords

Navigation