Plane Thermoelastic Waves in Infinite Half-Space Caused Operational Research in Engineering Sciences: Theory and Applications Vol. 4, Issue 3, 2021, pp. 142-175 ISSN: 2620-1607 eISSN: 2620-1747 DOI: https://doi.org/10.31181/oresta111221142c * Corresponding author. cetinkaya@cankaya.edu.tr (F. C. Çetinkaya), mehmet.duman@neu.edu.tr (M. Duman) SCHEDULING WITH LOT STREAMING IN A TWO- MACHINE RE-ENTRANT FLOW SHOP Ferda Can Çetinkaya 1*, Mehmet Duman 2 1 Department of Industrial Engineering, Çankaya University, Ankara, Turkey 2 NERITA, Near East University, TRNC, Mersin 10, Turkey Received: 16 August 2021 Accepted: 24 November 2021 First online: 11 December 2021 Research paper Abstract: Lot streaming is splitting a job-lot of identical items into several sublots (portions of a lot) that can be moved to the next machines upon completion so that operations on successive machines can be overlapped; hence, the overall performance of a multi-stage manufacturing environment can be improved. In this study, we consider a scheduling problem with lot streaming in a two-machine re-entrant flow shop in which each job-lot is processed first on Machine 1, then goes to Machine 2 for its second operation before it returns to the primary machine (either Machine 1 or Machine 2) for the third operation. For the two cases of the primary machine, both single-job and multi-job cases are studied independently. Optimal and near-optimal solution procedures are developed. Our objective is to minimize the makespan, which is the maximum completion time of the sublots and job lots in the single-job and multi- job cases, respectively. We prove that the single-job problem is optimally solved in polynomial-time regardless of whether the third operation is performed on Machine 1 or Machine 2. The multi-job problem is also optimally solvable in polynomial time when the third operation is performed on Machine 2. However, we prove that the multi-job problem is NP-hard when the third operation is performed on Machine 1. A global lower bound on the makespan and a simple heuristic algorithm are developed. Our computational experiment results reveal that our proposed heuristic algorithm provides optimal or near-optimal solutions in a very short time. Key words: scheduling, lot streaming, two-machine, re-entrant flow shop, makespan. 1. Introduction Manufacturing systems vary from a simple one-stage environment to more complex environments, such as a general job shop system, where jobs have different routings through multiple stages. Since the well-known efficient scheduling algorithm for the basic two-machine flow shop system, in which the flow of each job Scheduling with lot streaming in a two-machine re-entrant flow shop 143 is the same, was proposed by Johnson (1954), scheduling problems in different and more complicated manufacturing environments have been extensively studied. The re-entrant flow shop, which is a relatively new flow shop manufacturing environment, has drawn researchers’ attention. In the re-entrant flow shop, a job has to re-visit some of the machines since the number of operations for each job is more than the number of machines (Lev & Adiri, 1984). We observe re-entrant Re-entrant flow shops can be observed mainly in textile and high-tech industries, such as printing printed circuit boards, wafer fabrications, and signal processing. In all of the studies in the literature for the re-entrant flow shops and most of the scheduling studies for the other multi-stage manufacturing systems, it is assumed that jobs are indivisible entities. Thus, an operation of a job-lot (i.e., a process batch) consisting of identical units must be finished before it this job lot is transferred to the next machine. However, in many industrial applications, a job lot can be split into several sublots (i.e., transfer batches), which are the partial batches of the process batch. Transfer of the processed sublots to downstream machines without waiting for the completion of the whole job-lot on a machine gives an opportunity of allows the operations overlapping to overlap. The process of simultaneously splitting a job- lot into sublots and scheduling those sublots by overlapping their operations is known as lot streaming, which was first mentioned as a scheduling technique by Reither (1966). In this paper, we consider a scheduling problem with lot streaming in a two- machine re-entrant flow shop where each job-lot is processed first on Machine 1, then goes to Machine 2 for its second operation before it returns to the primary machine (either Machine 1 or Machine 2) for the third operation. We first focus on the single-job problem where a job-lot is spilt split into a given number of consistent or variable sublots. Consistent sublots case is the case where the size of each sublot does not change over the machines. However, the sublot sizes may change vary over the machines when variable sublots are used. Next, we extend the problem to the multi-job case in which the size of sublots and the schedule of multiple sublots and job lots need to be determined simultaneously. Our objective is to minimize the makespan, equivalent to the time to complete the last sublot in the single-job case, whereas it is the time to complete the last job lot in the multi-job case. Makespan aims to increase the utilization of the machines in the shop. To the best of our knowledge, our study is the only one that applies lot streaming for single- and multi- job cases in the re-entrant flow shops. The organization of the remaining parts is as follows. The following section presents a literature review on lot streaming problems, especially those in two- and three-machine manufacturing shops. In Section 3, the single-job problem is studied, and the optimal schedules with consistent and variable sublots are developed. Section 4 considers the multi-job lot streaming problem and gives an exact algorithm that determines the optimal consistent-sublot sizes and job schedules when the primary machine is Machine 2. Next, the multi-job lot streaming problem, in which the primary machine is Machine 1, is proved to be strongly NP-hard, and a polynomial-time solvable case of the problem is provided. Moreover, a heuristic algorithm is provided for the multi-job problem where the primary machine is Machine 1, and its effectiveness is computationally tested. Finally, our brief conclusions and some issues for future research are summarized in Section 5. Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 144 2. Literature review Although lot steaming is known and used in practice, there were no analytical studies in the literature of scheduling problems until the late ‘80s. Since then, lot streaming has attracted significant interest from researchers dealing with scheduling problems. Based on the number of job lots, the studies in the literature can be divided into two categories. One category is called the single-job lot problem and deals with the sublot-sizing problem in which the sublot sizes are determined when there is a single job lot. The other category is called multi-job problem and deals with the sublot- sizing and job-sequencing subproblems simultaneously. Here, we limit our literature review to the lot streaming studies for the single- and multi-job lot cases in two- and three-machine manufacturing shops only to expose the proper place of our study in the literature. The study by Baker (1988) is the first one considering the lot streaming technique for a single-job problem in two- or three-machine flow shops to minimize the makespan. For the two-machine case, he provided a linear programming model and determined the sublot sizes optimally. He also determined the optimal sublot sizes for the three-machine case where the job-lot is split into two sublots only. Potts & Baker (1989) proved that optimal sublots are consistent in the two-machine flow shop and illustrated that the consistent sublots are not always optimal for the flow shops with three and more machines. Glass et al. (1994) developed an algorithm that determines the optimal consistent-sublot sizes for the three-machine flow shop to minimize the makespan. This study was extended to the cases with sequence- independent detached and attached setups by Chen & Steiner (1997) and Chen & Steiner (1998), respectively. In the attached setup case of a machine, the first sublot belonging to a job lot should be available before setting up this machine. However, in the detached setup case, no need to wait for the arrival of the job lot. In both cases, no setup is necessary between successive sublots of the same job lot. In both cases, no setup is necessary between successive sublots of the same job lot. On the other hand, variable sublots case of the single job-lot problem was first examined by Trietsch (1989). The optimal solution with variable sublots in the three-machine flow shop was proposed by Trietsch & Baker (1993). Alfieri et al. (2012) and Alfieri et al. (2021) proposed exact, and heuristic solution approaches based on dynamic programming for a single-job problem to minimize the makespan and total flow time, respectively, in a two-machine flow shop with attached setup times. Vickson & Alfredson (1992) considered the concept of lot streaming for scheduling multiple job lots in flow shops. They demonstrated that a modified Johnson’s algorithm with unit-sized sublots solves the two-machine makespan minimization problem when the number of sublots in each jot-lot is unlimited and proved that sublots in each job-lot should be processed successively without the intermingling of different job lots. Vickson & Alfredson’s study was extended by Çetinkaya & Kayalıgil (1992) to develop a unified algorithm that treats sequence- independent attached and detached setups. Çetinkaya (1994) considered the scheduling of multiple job lots in a two-machine flow shop with attached setup and removal times on the machines and proved that the optimal schedule of the job lots to minimize the makespan is obtained by determining the equal or unequal sublots sizes of each job-lot independently and sequencing the job lots by a modified Scheduling with lot streaming in a two-machine re-entrant flow shop 145 Johnson’s algorithm. Vickson (1995) provided an optimal solution for the multi-job problem in a two-machine flow shop with sequence-independent attached or detached setups and transfer times from the first machine to the second one. Pranzo (2004) extended Çetinkaya’s study by considering limited buffers between machines. Glass & Possani (2011) considered the two-machine flow shop problem with attached setup and transportation times to minimize the makespan of the multiple job lots. Their study showed that sublot-sizing and job-sequencing problems are solved independently, as in Çetinkaya (1994) and Vickson (1995), and provided an algorithm solvable in polynomial time. Baker (1995) considered the multi-job problem with equal-sized sublots and setup times in a two-machine flow shop and proposed an algorithm using the time-lag approach. Yang & Chern (2000) extended Baker’s study to where detached setup times, transportation times, and removal times exist. Sriskandarajah & Wagneur (1999) investigated the multi-job problem in a two-machine flow shop with no-wait constraint. Çetinkaya (2005) considered a two-machine flow shop with a single agent transferring a completed item from Machine 1 to Machine 2. Machine 1 is blocked while the transport agent is in transferring and returning. He provided an algorithm that determines the optimal schedule for the case with unit-sized sublots. From the early 2000s, researchers began to consider flow shops with more than three machines (stages). Several metaheuristic algorithms, such as genetic algorithms (Yoon & Ventura, 2002; Marimuthu et al., 2008; Martin, 2009; Defersha & Chen, 2010), discrete particle swarm optimization algorithm (Tseng & Liao, 2008), threshold accepting and ant-colony optimization algorithms (Marimuthu et al., 2009), discrete artificial bee colony algorithm (Pan et al., 2011), and migrating birds optimization algorithm (Devendra et al., 2014; Meng et al., 2018) have been proposed to solve the multi-job lot streaming problem by considering its different aspects. All the studies mentioned above for the multi-job scheduling with lot streaming in the flow shops with more than three machines assume that sublots in each job lot should be processed successively for each operation on each machine. i.e., the intermingling of the job lots is not allowed, and the schedules are permutation schedules where the sequence of job lots on all machines is the same. However, a more realistic case is when intermingling of the job lots and non-permutation schedules are allowed. Feldmann & Biskup (2008) investigated the permutation flowshop scheduling problem with lot streaming and intermingling and developed a mixed-integer programming model. Rossit et al. (2016) investigated this non- permutation flowshop scheduling problem with lot streaming. They proposed a mathematical model to minimize the makespan of the multiple job lots that are not allowed to be intermingled. Besides the studies mentioned above, lot streaming studies for other two- and three-machine manufacturing shops are scarce. There are studies considering open shops (Şen & Benli, 1999), hybrid flow shops (Kim et al., 1997; Zhang et al., 2003; Zhang et al., 2005; Liu, 2008; Defersha, 2011; Defersha & Chen, 2012a; Naderi & Yazdani, 2015; Cheng et al., 2016; Zhang et al., 2017; Wang et al., 2019; Li et al., 2020), and mixed shops (Çetinkaya & Duman, 2010), job shops (Buscher & Shen, 2009; Defersha & Chen, 2012b), and assembly shops (Sarin et al., 2011; Yao & Sarin, 2014; Nejati et al., 2016; Cheng & Sarin, 2020). Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 146 The comprehensive surveys by Chang & Chiu (2005), Sarin & Jaiprakash (2007), Gomez-Gasquet et al. (2013), and Cheng et al. (2013), and Salazar-Moya & Garcia (2021) are also available for lot streaming problems with job lots having more than one operation in multiple machines environments. As we can see from the lot streaming literature and to the best of our knowledge, there is no previous study dealing with lot streaming for single- and multi-job cases in the re-entrant flow shops. The main contributions of our study can be summarized as follows: • This study is the first one in the scheduling literature dealing with lot streaming in the re-entrant flow shops. • Our study proves that the single-job problem is polynomial-time solvable regardless of whether the third operation is performed on Machine 1 or Machine 2 and develops optimal schedules with closed formulae for the optimal consistent and variable sublot sizes. • Our study also proves that the multi-job problem is polynomial-time solvable when the third operation is performed on Machine 2 and develops optimal schedules with closed formulae for the optimal sublot sizes. However, the multi-job problem is NP-hard when the third operation is performed on Machine 1. • A global lower bound on the makespan and a simple heuristic algorithm providing optimal or near-optimal schedules have been developed. 3. Single-job case Our single-job problem in the two-machine re-entrant flow shop is explained as follows: A job-lot of U identical items has three operations to be performed. Each operation k )3,2,1( =k requires kp time units of processing. There are two machines 1M (Machine 1) and 2M (Machine 2) operating independently. The first and second operations of the job-lot are performed on 1M and 2M , respectively. A primary machine ( 1M or 2M ) is re-visited by each item of the job lot for its third operation; hence the shop is a re-entrant flow shop. The job-lot is split into s sublots, and kix , is the size of the i th sublot that completes its k th operation. Sublots of a job-lot can be immediately transferred from one machine to another for their next operation without waiting to complete other sublots. The goal is to determine the size and schedule of all sublots to minimize the makespan, which is the time to complete the third operation of the sublot processed as the last. The assumptions made for the single-job problem are summarized here: • The sublots of a job lot are processed without any interruption on every machine. i.e., pre-emption is not allowed. • Each machine is ready at the beginning, say time zero, of the planning horizon. i.e., machines are not batching machines. • At any time, only one item of a job lot can be processed by a machine. • An unlimited storage space exists between the machines. • An idle time on a machine may occur between processing sublots. Scheduling with lot streaming in a two-machine re-entrant flow shop 147 • Transfer times from one machine to another are negligibly so short and thus ignored. • Setup times before processing the job-lot on a machine are negligibly so short and thus ignored. • The number of sublots is known in advance and fixed from one machine to another. • Processing times are known and deterministic. 3.1. Machine 2 is the primary machine We first consider that 2M is the primary machine where the third operation is performed. We investigate the problem for cases with consistent and variable sublots. 3.1.1. Consistent sublots When the lot streaming is applied, sometimes there might be no advantage to change the size of the sublots after they have completed their processing on a machine. In this situation, it is reasonable to let the sublot sizes be constant (consistent) over all pairs of operations, i.e., iki xx =, for 1, 2,..., ; 1, 2, 3i s k= = where Ux si i = 1 . Sublot availability assumption is used when sublots are consistent. i.e., a sublot can be processed at the next machine if all items in this sublot are completed on the current machine. We start our analysis with the following lemma. Lemma 1. For the single-job problem where 2M is the primary machine, it is sufficient to consider schedules of sublots where the last two operations of the sublots are processed consecutively on 2M . Proof. Consider any schedule of the consistent sublots. Suppose that there is a pair of sublots u and v , where the second operation of sublot v is immediately processed before the third operation of sublot u on 2M . Then interchanging the positions of sublots u and v on 2M is feasible and does not increase makespan. On machine 2M , when all sublots, which are immediately processed before sublot u , are pair-wise interchanged with the second operation of sublot u , then it is possible to consecutively process the second and the third operations of sublot u on 2M . Similarly, it is possible to schedule consecutively the second and the third operations of all sublots on 2M . From Lemma 1, the single-job problem can be illustrated by a network, as shown in Figure 1. Let 1 2( , ,..., )sx x x x= denote the sublot sizes, and ( , )k i be a node for the pair with operation k ( 1, 2, 3)k = and sublot i ( 1,..., )i s= where k ip x is the processing time of the k th operation of sublot i . The vertical arc from node (1, )i to node (2, )i indicates that sublot i cannot be processed on 2M unless it is completed on 1M . The horizontal arc from node (1, )i to node (1, 1)i + indicates that 1M can Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 148 start to process sublot 1i+ upon the completion of sublot i on 1M . Similarly, the horizontal arc from (2, )i to (3, )i represents that the third operation of sublot i can be started when its second operation is completed on 2M . Figure 1. Network representation for the case where 2M is the primary machine Here, the goal is to determine the sublot sizes minimizing the critical path (longest path) length from node (1, 1) to (3, )s , in which the sum of the processing times of the nodes on the critical path gives the length of the critical path. The following theorem gives the optimal sublot sizes for the single-job problem where 2M is the primary machine. Theorem 1. For the single-job problem where 2M is the primary machine, the optimal consistent-sublot sizes are 2 1 1 /(1 ... ) s x U − = + + + + , 1 1 xx i i − = for si ,...,2= where ( ) 132 / ppp += , 1 ii sU x = , and the associated optimal makespan is ( )2 1max 1 1 2 3 1 2 31( ) /(1 ... ) ( ) s s ii C p x p p x p p p U − = = + + = + + + + + + . Proof. See the Appendix A. ■ Theorem 1 proves that the single-job problem where 2M is the primary machine is equivalent to the single-job problem in the basic two-machine flow shop with processing times 1p and 32 pp + on 1M and 2M , respectively. From Theorem 1, it is clear that the optimal consistent-sublot sizes can be determined in ( )O s time. Example 1. In this numerical example, we illustrate Theorem 1, in which the consistent sublots are optimal. Suppose that we have a job lot of 70 identical items that will be split into three sublots, and the processing times for its three operations are 2, 3, and 1 time-units, respectively. Then, from Theorem 1, we determine that the sublot sizes are found to be 10, 20, and 40 units for the first, second, and third sublots, respectively. That is, 2 1 3 1 1 /(1 ... ) 70/(1 2 2 ) 10 s x U − − = + + + + = + + = , 2x = 1 12 (2)(10) 20x = = , and 2 3 1 2 (4)(10) 40x x= = = , where 2 3 1( ) / (3 1) / 2p p p = + = + 2= . The optimal makespan is max 1 1 2 3( ) (2)(10) (3 1)(70) 300C p x p p U= + + = + + = , as illustrated in Figure 2. 3.1.2. Variable sublots In some cases, there might be some advantages to change the size of the sublots after they have completed their processing on a machine. Thus, variable sublots are 11xp 12xp 21xp 22xp Sxp2 Sxp3 13xp 23xp Sxp1 1M 2M 31xp 32xp Scheduling with lot streaming in a two-machine re-entrant flow shop 149 preferred to the consistent sublots, using the item availability assumption where an individual item in a sublot can be processed at the next machine when this item is completed on the current machine. Figure 2. Optimal schedule of the sublots in Example 1 Remark 1. There is no need to investigate the optimal solution for the single-job problem having variable sublots in the two-machine re-entrant flow shop where 2M is the primary machine. The solution of the single-job problem having consistent sublots is also optimal for the single-job problem having variable sublots since only one set of sublot transfers from 1M to 2M is needed when the second and third operations are performed on 2M . 3.2. Machine 1 is the primary machine We now consider that 1M is the primary machine where the third operation is performed. We again investigate the problem for cases with consistent and variable sublots. 3.2.1. Consistent sublots Similar to the analysis given by Wang et al. (1997) for the basic two-machine re- entrant flow shop makespan minimization problem without lot streaming, we present the following lemma and theorem for finding the optimal solution to our problem having consistent sublots when 1M is the primary machine. We first give the following definitions. Definition 1. A schedule is called a compact schedule if the first operations of all sublots are scheduled successively on 1M and then followed by the third operations of all sublots. Definition 2. A schedule is called a permutation schedule if all sublots are processed in the same order on both machines. Lemma 2. For the single-job problem where 1M is the primary machine, it is sufficient to consider only compact and permutation schedules of the sublots. Proof. Consider any feasible schedule with consistent sublots. We first show that can be transformed into a compact schedule on 1M without worsening the makespan. Suppose we have a pair of two sublots, u and v , where the last (third) operation of sublot u immediately precedes the first operation of sublot v on 1M , i.e., 1,3, vu OO . Then interchanging their positions does not worsen the makespan. 60 50 300 260 140 140 120 60 2 20 20 2M 1M 1 2 3 1 1 2 3 3 Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 150 When all sublots, which immediately follow sublot u , are pair-wise interchanged with sublot u , we assume that all first operations of the sublots are scheduled first on 1M . If any, we may eliminate the idle time between the first operations of any pair of successive sublots by moving the start time of the second sublot in the pair to the left. This movement does not affect the feasibility and does not change the makespan. Now, assume that an idle time exists between the last operations of any pair of successive sublots. Then we can eliminate it by moving all sublots except the last to the right. Again, this movement does not affect the feasibility and does not change the makespan. This shows that the first and the last operations of the job-lot are performed continuously without idle time between the sublots on 1M . Now, consider any optimal schedule , which is compact on 1M but in which the processing order of the sublots on the first pair ),( 21 MM of machines is different. Let sublots u and v be the first pair of sublots such that 2,1, vu OO and ,2 ,2 v uO Op . Let sublots u and v be the last pair of sublots such that 1,1, vu OO and 2,2, uv OO . Interchanging the order of 2,uO and 2,vO is possible and maintains compactness on 1M , and the makespan remains unchanged. This indicates that an optimal schedule, which is compact and the processing order of the sublots on the last pair ),( 12 MM of machines is the same, exists. Theorem 2. Let maxC and maxC denote the optimal makespan values of the single-job problem having consistent sublots in the two-machine re-entrant flow shop where 1M is the primary machine and the single-job problem having consistent sublots in the basic three-machine flow shop, respectively. If max 1 31 ( ) s ii C p p x = + = 1 3( )p p U+ then maxmax CC = ; otherwise, UppC )( 31max += . Proof. The single-job problem where 1M is the primary machine can be represented by a network, as illustrated in Figure 3. Let ),...,,( 21 sxxxx = denote the sublot sizes and ),( ik be a node for the pair with operation k )3,2,1( =k and sublot i ),...,1( si = , having a weight of ikxp that represents the sublot processing time. The vertical arc from node ) ,1( i to node ) ,2( i indicates that sublot i can be processed on 2M upon its completion on 1M . The horizontal arc from node ) ,1( i to node )1 ,1( +i indicates that 1M can start to process sublot 1i+ upon the completion of sublot i on 1M . Similarly, the vertical arc from ) ,2( i to ) ,3( i represents that the third operation of a sublot can be started when the second operation is completed on 2M . In view of Lemma 2, we observe that the precedence constraint of arc between ) ,1( s and )1 ,3( becomes redundant if UppC )( 31max + . Thus, an idle time on 1M exists between the first operation of the last sublot and the third operation of the first sublot. However, if the arc between ) ,1( s and )1 ,3( is removed from the network )(xN , the new network then represents the lot streaming problem in the three-machine flow shop. Therefore, it is clear that max max 1 3max , ( )= +C C p p U . It is obvious that max max =C C when max 1 3 ( ) +C p p U . Thus in this case, the lot streaming problems in Scheduling with lot streaming in a two-machine re-entrant flow shop 151 the basic three-machine flow shop and the two-machine re-entrant flow shop are equivalent. However, max 1 3 ( )= +C p p U when max 1 3 ( )C p p U + . ■ As a result of Theorem 2, an optimal solution to the single-job problem where 1M is the primary machine can be constructed by the optimal solution proposed by Glass et al. (1994) for the single-job problem having consistent sublots in the basic three- machine flow shop, as in the following corollary. Figure 3. Network representation for the case where 1M is the primary machine Corollary 2. When the makespan is minimized, the optimal consistent-sublot sizes for the single-job problem in the basic three-machine flow shop are also optimal for the single-job problem having consistent sublots in the two-machine re-entrant flow shop where 1M is the primary machine, and they are as follows: (i) If 31 2 2 ppp , then the optimal sublot sizes are: / )1/()1( 31 31 1 = −− = ppifsU ppifU x s , 1 1 xx i i − = for si 2 where 2 3 1 2( )/( )p p p p = + + . (ii) If 31 2 2 ppp , then there exists a crossover sublot h , which can be determined by a search algorithm in )(log sO time, for which the optimal sublot sizes are: = = −+−− −−+− −−−+−− = +− +− 3221 3221 3221 1 1 , , , })1/()1/{( )}1/()1(1/{ }1)1/()1()1/()1/{( ppppif ppppif ppppif hsU hU U x h hs hsh h , h ih i xx − = for 11 − hi , h hi i xx − = for sih where 21 / pp= and 23 / pp= . Example 2. Assume that we have a job lot of 70 identical items that will be split into three sublots, and the processing times for its three operations are 1, 4, and 2 time- units, respectively. This case corresponds to the second case in Corollary 2 since 2 2 p = 2 1 3 (4) 16 (1)(2) 2p p= = = . The size of the crossover sublot h is determined 11xp 21xp 11 −Sxp Sxp1 1M 12xp 22xp Sxp2 2M 12 −Sxp 13xp 23xp 13 −Sxp Sxp3 1M Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 152 by 1 /{( 1) /( 1) ( 1) /( 1) 1} h s h h x U − + = − − + − − − where 4/1/ 21 == pp and 3p = 2 / 1/ 2p = since 1 2 3 1 4 2p p p= = = . When 1=h , 1 1 3 1 1 1 1 1 (1/ 4) 1 (1/ 2) 1 / 1 70 / 1 1 1 (1/ 4) 1 (1/ 2) 1 h s h x U − + − + − − − − = + − = + − = − − − − 70 40 1.75 = , 20)40()2/1( 1 1 12 2 === − xx , 10)40()2/1( 2 1 13 3 === − xx , and 340max =C . When 2=h , 1 2 3 2 1 2 1 1 (1/ 4) 1 (1/ 2) 1 / 1 70 / 1 1 1 (1/ 4) 1 (1/ 2) 1 h s h x U − + − + − − − − = + − = + − = − − − − 70 40 1.75 = , 10)40()4/1( 1 1 12 1 === − xx , 20)40()2/1( 1 1 23 3 === − xx , and 330max =C (See Figure 4a). When 3=h , 1 3 3 3 1 3 1 1 (1/ 4) 1 (1/ 2) 1 / 1 70 / 1 1 1 (1/ 4) 1 (1/ 2) 1 h s h x U − + − + − − − − = + − = + − − − − − 53.33 , 33.3)33.53()4/1( 2 1 13 1 === − xx , 33.13)33.53()4/1( 1 1 23 2 === − xx , and 390max =C . It is clear that the optimal sublots are achieved when 2=h , and their sizes are 101 =x , 401 =x , and 203 =x . Note that the optimal makespan is 330maxmax == CC since max 1 3 330 ( ) (1 2)(70) 210 = + = + =C p p U , and the optimal schedule is obtained by carrying the third operations of the sublots in the three-machine flow shop problem to 1M , as illustrated in Figure 4b. Figure 4. Optimal schedules with consistent sublots for the single-job problem in Example 2: (a) Three-machine flow shop; (b) Two-machine re-entrant flow shop. 10 50 70 50 70 210 290 330 10 50 210 290 10 50 70 190 210 290 330 1M 2M 1 2 3 1 1 2 3 2 3 10 50 210 290 2M 1M 3M 3 1 2 3 (a) (b) 2 1 1 2 3 Scheduling with lot streaming in a two-machine re-entrant flow shop 153 3.2.2. Variable sublots As a result of Theorem 2, an optimal solution to the single-job problem having variable sublots where 1M is the primary machine can be constructed by the optimal solution proposed by Trietsch & Baker (1993) for the single-job problem with variable sublots in the basic three-machine flow shop, as in the following corollary. Corollary 3. When the makespan is minimized, the optimal variable-sublot sizes in the single-job problem in the basic three-machine flow shop are also optimal for the single- job problem having the variable sublots in the two-machine re-entrant flow shop where 1M is the primary machine, and they are as follows: (i) If 31 2 2 ppp , then the consistent sublots are optimal, and they are: 1 ( 1)/x U = − ( 1)s − , 1 1 xx i i − = for si 2 where )/()( 2132 pppp ++= . (ii) If 31 2 2 ppp , then the variable sublots are optimal, • the optimal sublot sizes between 1M and 2M are: )1/()1(1 −−= s Ux , 1 1 xx i i − = for si 2 where 12 / pp= , and • the optimal sublot sizes between 2M and 1M are: )1/()1(1 −−= s Ux , 1 1 xx i i − = for si 2 where 23 / pp= . From Corollary 3, it is clear that the optimal variable-sublot sizes can be determined in ( )O s time. Example 3. In this numerical example, we illustrate the second case of Corollary 3, in which the variable sublots are optimal since 31 2 2 ppp . Assume that we have a job lot of 15 identical items that will be split into two sublots, and the processing times for its three operations are 1, 2, and 1 time-units, respectively. From Corollary 2, the sublot sizes between 1M and 2M are found to be 5)12/()12(15 2 1 =−−=x and 10)5)(2( 1 2 ==x where 21/2/ 12 === pp and ).1)(1(4)2( 31 22 2 === ppp (See Figure 5a). Similarly, the sublot sizes between 2M and 1M are calculated as 2 1 15((1/ 2) 1) /((1/ 2) 1) 10x = − − = and 5)10)()2/1(( 1 2 ==x since 2/1/ 23 == pp . Thus, the optimal makespan of the problem having variable sublots in the basic three-machine flow shop is 40, and the optimal schedule of the problem in the two- machine re-entrant flow shop is obtained by carrying the third operations of the sublots in the optimal schedule of the problem in the three-machine flow shop to 1M , as illustrated in Figure 5b. 4. Multi-job case In this section, we extend our problem presented to the multi-job case. The multi- job problem is explained as follows: There is a job-lot set njjJ ,...,2,1== where each job-lot has jU identical items of type j . Let kjp , be the processing time for Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 154 k th ( )3,2,1=k operation of job-lot j . There are two machines 1M and 2M . Each job-lot is processed first on 1M , on 2M for its second operation before returning to the primary machine ( 1M or 2M ) for the third operation. Suppose that job-lot j is split into js sublots, then our goal is to determine the sublot sizes for each job-lot and the schedule of all sublots and job lots to minimize the makespan. Figure 5. Optimal schedules with variable sublots for the single-job problem in Example 3: (a) Three-machine flow shop; (b) Two-machine re-entrant flow shop. For the multi-job problem, we consider all assumptions made for the single-job problem. Furthermore, we do not allow the intermingling of different job lots. That is, once a sublot of a job-lot is started on a machine, all other job lots should wait until all of the remaining sublots of that job-lot are completed on the same machine. Note that intermingling the sublots belonging to different job lots may further reduce the makespan, but we focus on the no-intermingling case in our study. Intermingling can be considered in a future study as future research, as we point out in Section 5. As in the case of the single-job problem, we investigate two cases associated with the primary machine. 4.1. Machine 2 is the primary machine The following lemma gives the basic structure of the job-sequence for the multi- job problem where 2M is the primary machine. 25 (a) (b) 35 5 25 40 15 5 35 25 2M 1M 1 2 1 2 1 2 1 2 35 5 25 40 15 5 35 2M 1M 1 2 1 2 1 2 1 2 Scheduling with lot streaming in a two-machine re-entrant flow shop 155 Lemma 3. For an optimal solution of the multi-job problem where 2M is the primary machine, it is sufficient to consider the job-sequence in which the last two operations of each sublot belonging to a job-lot are processed successively on 2M . Proof. Omitted since it is similar to the proof of Lemma 1. ■ The multi-job problem where 2M is the primary machine can be decomposed into two sub-problems: (a) the job-sequencing and (b) the sublot-sizing. 4.1.1. Job-sequencing sub-problem The multi-job problem reduces to the determination of the optimal sequence of job lots when the sublot-sizing sub-problem for each job-lot has already been solved. That is, the job-sequencing sub-problem needs to be solved. Suppose that each job- lot j is independently split into sublots by Corollary 2 or Corollary 3. Let jRI = time lag between the start of the first operation on 1M and the latest start time of the second operation on 2M for job-lot j . i.e., the latest delay time, simply called run-in-delay for job-lot j on 2M without affecting the completion time of job-lot j (called as run-in delay for Operation 2). jRO = time lag between the completion times of the first and the second operations for job-lot j (called as run-out delay for Operation 2). jIR = time lag between the latest start of the second operation on 2M and the latest start time of the third operation on 1M for job-lot j . i.e., the latest delay time, simply called run-in-delay for job-lot j on 1M without affecting the completion time of job-lot j (called as run-in delay for Operation 3). jOR = time lag between the completion times of the second and the third operations for job-lot j (called as run-out delay for Operation 3). The following theorem provides the optimal solution to the job-sequencing sub- problem. Theorem 3. Given the solution of the sublot-sizing sub-problem, job-lot v precedes job-lot z in an optimal job-sequence if vzzv RORIRORI ,min,min , where jRI = 1 1 ,1 , ,2 ,3 ,1 1 min ( ) j i i i s j j r j j j rr r p x p p x − = = − + , and ,2 ,3 ,1( )j j j j j jRO RI p p p U= + + − for any job-lot j . Proof. See the Appendix B. ■ 4.1.2. Sublot-sizing sub-problem The sublot-sizing sub-problem needs to be solved for each job-lot when the job- sequencing sub-problem has already been solved. The following theorem provides the optimal solution to the sublot-sizing sub-problem. Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 156 Theorem 4. Given the solution of the job-sequencing sub-problem, the optimal sublot sizes in a job-lot processed in position j of the job-sequence are 2 ,1 /(1 = + + j j x U 1 ... ) − + + s j , 1, 1 , j i ij xx − = for jsi ,...,2= where 1,3,2, /)( jjj ppp += . Proof. See the Appendix C. ■ From Theorems 3 and 4, we have the following result. Corollary 4. Solving the sublot-sizing sub-problem using Theorem 4 and then solving the job-sequencing sub-problem using Theorem 3 provides the optimal solution to the multi-job problem where 2M is the primary machine. Based on Corollary 4, we propose the following exact-solution algorithm with a computational complexity of )log( nnO to solve the multi-job problem where 2M is the primary machine and n is the total number of job lots. Algorithm 1. Step 1: [Sublot-sizing] Calculate the size for each sublot of the job-lot in the set 1, 2,...,J j j n= = as 12 ,1 / (1 ... ) j s j jx U − = + + + + and 1, ,1 i j i jx x − = for 2,..., ji s= where ,2 ,3 ,1( ) /j j jp p p = + . Step 2: [Job-sequencing] (a) Set 12 ,1 /(1 ... ) j s j j jRI p U − = + + + + and ,2 ,3(j j j jRO RI p p= + + − ,1)j jp U . (b) To obtain the job-sequence 1, 2,..., = =j j n , consider all jobs in the job-lot set J and apply Johnson’s Algorithm with processing times on 1M and 2M replaced by jRI and jRO , respectively. (ii) Calculate the associated makespan as ( ) 1 max 1 ,2 ,31 1 1 max ( ) w w n w n j j j j jj j j C RI RO p p U − = = = = − + + 4.2. Machine 1 is the primary machine This section considers the multi-job problem where 1M is the primary machine. Unfortunately, this problem is more complicated than the multi-job problem where 2M is the primary machine discussed in Section 4.1. Theorem 5. The multi-job problem where 1M is the primary machine is NP-hard in the strong sense. Proof. Suppose that each job-lot has one sublot only. This special case of our multi- job problem is equivalent to the multi-job problem without lot streaming in the basic two-machine re-entrant flow shop where 1M is the primary machine. It has been Scheduling with lot streaming in a two-machine re-entrant flow shop 157 proven by Wang et al. (1997) that this special case is NP-hard in the strong sense. Thus, our multi-job problem where 1M is the primary machine is also strongly NP- hard. ■ The following lemma restricts our search to the compact and permutation schedules of the job lots for the optimal schedule of the multi-job problem where 1M is the primary machine. Lemma 4. For an optimal solution of the multi-job problem where 1M is the primary machine, it is sufficient to consider only compact and permutation schedules of the job lots. Proof. Omitted since it is similar to the proof of Lemma 2. ■ 4.2.1. A Polynomial-time solvable case Since the multi-job problem where 1M is the primary machine has been shown to be NP-hard in Theorem 5, an optimal algorithm solving the problem in polynomial- time is impossible. Therefore, we first examine a polynomial-time solvable case of the problem that corresponds to the case, in which the sublots of all job lots are independently determined by Corollaries 2 or 3 and the idle time between the first and the third operations of all jobs on 1M is zero. Let jI be the idle time on 1M between the first and the third operations of job-lot j when it is independently split into sublots by Corollaries 2 or 3. The following theorem provides the optimal schedule for this special case. Theorem 6. If 0=jI for every job-lot j , then arbitrarily sequencing all jobs as illustrated in Figure 6 is an optimal schedule for the multi-job problem where 1M is the primary machine. Proof. If 0=jI for every job-lot j , there will be no idle time between the first and the third operations of job-lot j on 1M when job lots are arbitrarily sequenced. Then, the time to complete all operations of job-lot j becomes ,1 ,3 ( ) j j j j T p p U= + ,1 ,3 ( ) j j j j I p p U+ = + when it is independently split into sublots, and max 1 n jj C T = = = ,1 ,31 ( ) n j j jj p p U = + becomes the optimal makespan. Figure 6. Optimal schedule obtained by Theorem 6 nT 2T 1T 1M 2M 1 1 2 2 n n 1 2 n Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 158 4.2.2. Proposed heuristic algorithm For a given job-sequence , we define the job-lot d such that ,11 ,2 ,21 − = n j jd dj C p U C where 2,dC is the time to complete the second operation of job-lot d , and (2,2 ,21 11,..,( ) max = + === − = − n j d j d ij d ij n C C p U RI ) ( ) 1 ,21 1 − = = + j n ji ji j RO p U where )(2 C is the completion time of all job lots in sequence on 2M , and 1 ,2 ,2 ,2d d d dC C p U − = − . Here job-lot d that will be used to develop the proposed algorithm is called a partition job-lot and is the first job whose second operation finishes later than the completion of all first operations on 1M (see Figure 7). From the definition of the job-lot d , it is clear that no idle time exists between job lots following job-lot d on 2M . Thus, permutation sequence is partitioned into three subsequences: | 1,...,BJ j j= = 1d − , djjJD == | , and ndjjJA ,...,,1| +== . We propose the following heuristic algorithm with a computational complexity of )log( nnO to solve the multi-job problem where 1M is the primary machine. Figure 7. Partition job-lot in position d of the job-sequence Algorithm 2. Step 1: For each job lot j , (a) If sublots are considered as consistent, then use the sublot sizes in Corollary 2; otherwise (variable sublots), use the sublot sizes in Corollary 3. (b) Compute the makespan jT , and idle time on 1M as ,1(j j jI T p= − + ,3 ) j j p U . Step 2: Divide the job-lot set J into two sets: 0|1 == jIjJ and 2 | 0 j J j I= . Step 3: If = 2 J , then any arbitrary sequence of job lots is optimal, and the optimal makespan is ,)( 1 3,1, * max j n j jj UppC = += stop; otherwise, go to Step 4. )(max C )(2 C 1M 2M 2,dC 2,1−dC )(1 C BJ DJ AJ BJ DJ AJ BJ AJ DJ Scheduling with lot streaming in a two-machine re-entrant flow shop 159 Step 4: Compute the run-in delays ( jRI and jIR ) and run-out delays ( jRO and jOR ). Schedule all job lots njjJ ,...,1| == by applying Johnson’s Algorithm with processing times on 1M and 2M replaced by jRI and jRO to obtain the job-sequence njj ,...,1| == . Step 5: In the job-sequence , determine the partition job-lot d such that ,11 ,2 ,21 − = n j jd dj C p U C where ( ) 1 ,2 1 11,.., max − = == = − + j j d i ii ij n C RI RO ( ),21 = d jjj p U , and dddd UpCC 2,2,2,1 −=− . Step 6: Compute the associated makespan )(max C as: 1 max ,1 ,31 1 11,.., ( ) max , max n d j j j j j ij j ij n C p U p U RI − = = == = + − 1 1 1 ,21 1 ,.., max j d j j i j j i ii j i d i dj d n RO p U RI RO − − − = = = == + + − ,3 n j jj d p U = + Step 7: If j n j jj UppC )()( 1 3,1,max = += , then the job-sequence is optimal, stop; otherwise, go to Step 8. Step 8: Consider the job lots in ndjjJJ AD ,...,| == , and apply Johnson’s Algorithm with processing times on 1M and 2M replaced by jIR and jOR to obtain the partial job-sequence 1,...,1| +−== dnjj . Step 9: Set the final sequence of job lots as ,= where 1,...,1| −== djj followed by 1,...,1| +−== dnjj . The associated makespan is 1 1 max ,1 ,31 1 1 11,.., ( ) max , max n d j j j j j j ij j i ij n C p U p U RI − − = = = == = + − 1 1 ,21 ,.., max d j j i j j i ij i d i dj d n RO p U RI RO − − = = == + + − + ,3 n j jj d p U = 4.2.2. Lower bounds on the makespan We now develop four lower bounds on the makespan, and then we take the maximum of these lower bounds as a global lower bound, which will be used to test Algorithm 2. Lower bound 1. This lower bound is obtained by assuming that there will be no idle time between the first and the third operations of each job lot on 1M . Thus, the natural lower bound is Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 160 = += n j jjj UppLB 1 3,1,1 )( , (1) which is equivalent to the total processing time of all job lots on 1M . Lower bound 2. We derive the second lower bound as ( ) ( )j nj j n j jjnj ORUpRILB ++= === ,..,11 2, ,..,1 2 minmin (2) by assuming that the job lot with the minimum run-in delay for Operation 2 is processed as the first job-lot in the job-sequence, 2M operates continuously without any idle time between job lots, and the job lot with the minimum run-out delay for Operation 3 is processed as the last job lot in the job-sequence. Lower bound 3. To derive our third lower bound, we assume that all job lots do not wait for their operations to be performed on 2M . For a sequence , we can establish the third lower bound as ( ) ( )j nj n ji ii j i i j i inj ORUpRILB + ++= == − === ,..,1 2, 1 1 2,1 1,..,1 3 min max (3) where i is the job-lot in position i of the job-sequence , and 2,1 i is the overlapping time for job lot i when both 1M and 2M are busy (i.e., operating continuously without idle time between sublots), as illustrated in Figure 8. Figure 8. Run-in and run-out delays We can rewrite (3) as 1 3 ,2 ,21 1 1,..,1,.., max ( ) min j j n ji i i i i ii i i j j nj n LB RI p U RO p U RO − = = = == = + − + + 1 1 ,2 ,21 1 1 1,..,1,.., max min j j j n ji i i i i ii i i i j j nj n RI p U RO p U RO − − = = = = == = + − + + 1 ,21 1 1 1,..,1,.., max min j j n ji i i ii i j j nj n RI RO p U RO − = = = == = − + + (4) Note that the last two terms in (4) are constant and independent from the sequence, and the first term, == j i inj RI 1,..,1 max − − = 1 1 j i i RO , gives the total idle time 3,2 i )iRO iOR 2,1 i iRI iIR 31 MM 1M 2M Scheduling with lot streaming in a two-machine re-entrant flow shop 161 on 2M before completing the second operations of all job lots. The first term equals Johnson’s expression in which the processing times on 1M and 2M are replaced by the run-in and run-out delays, respectively. Thus, the job-sequence is obtained by job-lot k preceding job-lot l if min , min ,k l k lRI RO RO RI . Therefore, we obtain the third lower bound as ( )( ) j nj ORRORIJACLB += = ,..,1 * max3 min, (5) where ( )( )RORIJAC ,*max is the makespan obtained by Johnson’s Algorithm (JA) with processing times on 1M and 2M replaced by jRI and jRO , respectively. Lower bound 4. For any job-sequence , we can establish the fourth lower bound as 1 2,3 4 ,31 11,.., 1,.., min max j j n j i i iii i i jj n j n LB RI RI p U − = = == = = + + + (6) where 3,2 i is the overlapping time of the second and third operations for job lot )(i as shown in Figure 8. Equation (6) can be rewritten as 1 4 ,3 ,31 11,.., 1,.., min max ( ) j j n j i i i i i ii i i jj n j n LB RI RI p U RO p U − = = == = = + + − + 1 ,31 1 11,.., 1,.., min max j j n j i i i ii i ij n j n RI RI RO p U − = = == = = + − + (7) Note that the second term in (7), 1 1 11,.., max j j i ii ij n RI RO − = == − , is minimized by job-lot k preceding job lot l in the job-sequence if min , min ,k l k lRI RO RO RI . Therefore, we obtain the fourth lower bound as ( )( )ORIRJACRILB j nj += = ,min * max ,..,1 4 (8) where ( )( )ORIRJAC ,*max is the makespan obtained by Johnson’s Algorithm with processing times on 1M and 2M replaced by jIR and jOR , respectively. The global lower bound GLB becomes the maximum of the lower bounds developed above. i.e., 1,...,4maxi iGLB LB== . Example 4. As an illustration of the proposed algorithm and the global lower bound, we consider the five-job problem in Table 1. Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 162 Table 1. Processing times, number of sublots, and lot sizes j 1,j p 2,j p 3,j p j s j U 1 3 2 3 4 40 2 1 2 2 3 30 3 1 2 7 2 20 4 1 4 2 3 70 5 2 2 1 3 35 When all job lots are independently split into their number of consistent sublots, the sublot sizes and jT and jI values are obtained, as illustrated in Table 2. Table 2. Sublot sizes with jT and jI values j 1,j x 2,j x 3,j x 4,j x j T j I 1 10 10 10 10 240 0 2 6 12 12 - 90 0 3 5 15 - - 160 0 4 10 40 20 - 330 120 5 14 14 7 - 105 0 Job-lot set J is decomposed into two job-lot sets, 5 ,3 ,2 ,11 =J and 42 =J . We continue with Step 4 of Algorithm 2 since the job-lot set 2 J is not empty. The run-in and run-out delays for each job lot are obtained as shown in Table 3. Table 3. Run-in and run-out delays j j RI j RO j IR j OR 1 30 20 20 30 2 6 24 12 24 3 5 105 10 105 4 10 220 180 40 5 28 28 42 7 The application of Johnson’s Algorithm with processing times on 1M and 2M replaced by jRI and jRO ; respectively, gives the job-sequence 1 ,5 ,4 ,2 ,3= . The partition job-lot is job 2, and it is in the second position (i.e., 2=d ) of the sequence . The associated makespan )(max C for the sequence is computed as 805. The global lower bound becomes 805,,,max 4321 == LBLBLBLBGLB where 1 ,1 ,31 ( ) (3 3)(40) (1 7)(20) (1 2)(70) (2 1)(35) 805 n j j jj LB p p U = = + = + + + + + + + = , ( ) ( ) 2 ,211,.., 1,..,min min min 30, 6, 5,10, 28 ((2)(40) (2)(30)== = = + + = + + n j j j jjj n j n LB RI p U RO (4)(70) (2)(35) min 30, 24,105, 40, 7 542+ + + = , ( )( ) ( ) ( ) * *3 max max 1,.., , min 4 2 3 1 5 min 30, 24,105, 40, 7 = = + = − − − − + j j n LB C JA RI RO RO C Scheduling with lot streaming in a two-machine re-entrant flow shop 163 535 7 542= + = , ( )( ) ( )* *4 max max 1,.., min , min 30, 6, 5,10, 28 3 1 4 2 5 = = + = + − − − − j j n LB RI C JA RI RO C 5 560 565= + = . The job-sequence 1 ,5 ,4 ,2 ,3= is optimal since the associated makespan equals the natural lower bound 1LB . Thus, we stop. 4.2.3. Computational experiments and results The efficiency measure of Algorithm 2 is the computational time required to solve the problem. However, its computational time is not measured provided since it is relatively very short, less than a few seconds. On the other hand, to test the effectiveness of Algorithm 2, we generate the parameters for our problem instances as follows: n : number of job lots , 5,10,15, 20, 25,50, 75,100n . ,j k p : k th operation’s processing time for job lot j is randomly generated form from a discrete uniform distribution over 1, 10, including the lower and upper limits. j s : number of sublots for any job lot j is randomly generated form from a discrete uniform distribution over 2, 10. j U : lot size for any job lot j is randomly generated form from a discrete uniform distribution over 2, 50. For each possible number of job lots from 5 to 100, we first generate 100 problem instances in which the. The processing times for all operations are randomly distributed without any dominance of a specific operation, and a total of 800 problem instances are tested. The following statistics are collected: Z NT = number of times percent deviation is zero (i.e., the heuristic makespan equals the global lower bound). ]1,0(NT = number of times the percent deviation is greater than zero but less than or equal to 1 (i.e., GLBC H 01.1max ). AVE = average percent deviation. MAX = maximum percent deviation. From the results of our experiments, we observed that Algorithm 2 finds the optimum makespan (i.e., the heuristic makespan equals the global lower bound) for all 800 test problems when all processing times are randomly generated without any dominance of a specific operation. That is, 800 Z NT = . However, to evaluate the effectiveness of Algorithm 2 when the same operation for all job lots is dominant, we repeated our computational experiments with three data sets. The first data set, D1, assumes that the maximum processing time is on the first operation for all job lots, i.e., 3,2,1, ,max jjj ppp for j . Similarly, the second Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 164 and third data sets D2 and D3 correspond to the cases where ,2 ,1 ,3max ,j j jp p p and ,3 ,1 ,2max ,j j jp p p for j , respectively. In each of the data sets, we assume that the processing time of the dominant operation for each job-lot is randomly generated form from a discrete uniform distribution over 6, 10, and the processing times of the other operations are randomly generated form from a discrete uniform distribution over 1, 5. Our experiments with data sets D2 and D3 show that the global lower bound equals the heuristic makespan in all 800 problem instances tested when either the first or third operation is dominant. This result means that the global lower bound is effective when the first or third operation is dominant. However, the same argument is not valid for the case, where the second operation is dominant, since the global lower bound equals the heuristic makespan in 336 problem instances out of 800, as illustrated in Table 4. From Table 4, it is clear that the heuristic makespan deviates 1 percent from the global lower bound in 790 problem instances out of 800 for the case where the second operation is dominant. In the remaining ten problem instances when 5n = , the average and maximum deviations from the global lower bound are 0.352 percent and 2.055 percent, respectively. It should also be noticed that the average and maximum deviations decrease as the number of jobs increases. Table 4. Performance of Algorithm 2 when the second operation is dominant n NPIS ZNT ]1,0(NT AVE MAX 5 100 45 45 0.352 2.055 10 100 30 70 0.136 0.656 15 100 40 60 0.076 0.569 20 100 37 63 0.050 0.231 25 100 36 64 0.036 0.116 50 100 37 63 0.014 0.044 75 100 58 42 0.006 0.026 100 100 53 47 0.005 0.021 Given the results of our computational experiments, we conclude that Algorithm 2 is quite effective in solving the multi-job problem where Machine 1 is the primary machine. 5. Conclusions and future research In this study, we considered a problem in a two-machine re-entrant flow shop in which lot streaming is used for scheduling the single-job and multi-job cases separately. When Machine 1 or Machine 2 is the primary machine on which the third operation is performed, we proved that the single-job problem is polynomial-time solvable. We have also proved that the multi-job problem can be solved optimally when the third operation is performed on Machine 2. However, we have also proved that the multi-job problem is NP-hard when Machine 1 is the primary machine, machine so that we have developed a simple heuristic algorithm. To examine the effectiveness of our algorithm, we have developed a global lower bound on the Scheduling with lot streaming in a two-machine re-entrant flow shop 165 makespan. The results of our computational experiments imply that our heuristic algorithm is quite effective in solving the multi-job problem optimally in reasonably short computational times. Our study has some limiting assumptions for the problem under study. As in most studies in the lot streaming literature, we assume that the number of sublots of each job lot is known in advance. However, a more realistic case is when the problem’s solution has to give its value along with the sublot sizes. In our study, as in almost all of the previous studies in the literature, we assume that sublots in each job lot should be processed successively for each operation on each machine. i.e., we do not allow the intermingling of the job lots for the multi-job case. We may or may not obtain a better schedule by relaxing this assumption, but it is worth investigating. Furthermore, in our study, sublot sizes may not be integral, so rounding them to the nearest integer numbers without violating the job-lot size may be needed. However, this approach may not provide the optimal makespan. Thus, this issue is also worthy of investigating. Through our study for two machine and re-entrant flowshops, we hope that researchers working on scheduling problems will be aware that there is no study other than ours for scheduling with lot streaming in re-entrant manufacturing systems. Our study will open a new direction in the literature of scheduling with lot streaming since it is the only study that applies lot streaming for single- and multi- job cases in the re-entrant flow shops. We hope that our work will form a basis for developing algorithms to solve the scheduling problems in more complex re-entrant manufacturing systems where lot streaming is allowed. There are several research extensions of this study that are open for future investigation: • The assumption of knowing the number of sublots in advance could be relaxed, and investigating the problem under consideration without this assumption could be a future study issue. • The sublot sizes obtained for the no-setup case studied in this paper may not be valid for the setup case The so that problem under consideration can be extended to a case where an issue with attached or detached setup is made before processing a job-lot. • Our study assumes that sublots in each job-lot should be processed successively for each operation on each machine. i.e., we do not allow the intermingling of the job lots for the multi-job case. Relaxing this the no-intermingling assumption could be another future research issue. • Our study could also be extended to the flow shops with more than two stages, having one machine at each stage, and a job lot may visit some stages more than once. and hybrid flow shops, which are the flow shops in which at least one of the stages has more than one machine, could also be studied. • Different measures of performance rather than makespan could also be considered. For instance, the total completion time of sublots and job lots in the Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 166 single- and multi-job cases, respectively, could be a more relevant performance measure than makespan if the inventory holding costs are minimized. Appendix A. Proof of Theorem 1. As illustrated in the network representation given in Figure 1, a lower bound on the makespan max C , which is based on the first sublot, is given as the first sublot’s processing time on 1M plus the total processing time of all sublots on 2M , i.e., max 1 1 1 2 3 1 2( )( ... )sC LB p x p p x x x = + + + + + . Similarly, the total processing time of sublots 1 and 2 plus the total processing time of sublots 2 through s gives another lower bound, i.e., max 2 1 1 2 2 3 2 3 ( ) ( )( ... ) s C LB p x x p p x x x = + + + + + + . Similarly, we can write max 1 2 31 ( ) i s i k kk k i C LB p x p p x = = = + + for each sublot si ,...,2= . It follows that max 1 2 311 1 max max ( ) i s i k kk k ii s i s C LB p x p p x = = = + + . Thus, the linear programming model below can formulate the single-job problem where 2M is the primary machine. ( P ) Minimize maxC (A.1) Subject to max 1 2 31 ( ) i s k kk k i C p x p p x = = + + for si ,...,2= (A.2) 1 s ii x U = = (A.3) max 0C (A.4) 0ix for si ,...,2= (A.5) Assuming that each of the inequalities in (A.2) is satisfied as equality, i.e., all sublots are critical to determining the makespan maxC ; we can obtain a feasible solution to problem P . In such a case, both 1M and 2M operate without idle time from one sublot to another, and the adjacent pair of sublots must satisfy the following relationship: 1 1 2 3 1 2 31 1 1 ( ) ( ) i s i s k k k kk k i k k i p x p p x p x p p x − = = = = − + + = + + for si ,...,2= (A.6) or equivalently, ( )1 2 3 1/i ix x p p p−= + for si ,...,2= (A.7) The idle time on 2M is only before the first sublot, and it equals 11xp . Then solving the set of simultaneous equations (A.3) and (A.7) yields )...1/( 12 1 − ++++= s Ux , (A.8) 1 1 xx i i − = for si ,...,2= (A.9) where ( ) 132 / ppp += . Scheduling with lot streaming in a two-machine re-entrant flow shop 167 Using (A.2), (A.3), (A.8) and (A.9), the makespan is obtained as 2 1 max 1 1 2 3 1 2 31 ( ) ( / (1 ... ) ( )) s s ii C p x p p x p p p U − = = + + = + ++ + + + + . (A.10) We have shown that the solution given by the theorem is feasible. Note that this solution is the solution for the single-job lot streaming problem in the two-machine flow shop with processing times 1 p and 2 3 p p+ on 1M and 2M , respectively. Now, to prove the optimality of this feasible solution, the problem P may be re- written as ( P ) Maximize maxC− (A.11) Subject to max 1 2 31 ( ) 0 i s k kk k i C p x p p x = = − + + + for si ,...,2= (A.12) 1 s ii x U = = (A.4) max 0C , 0ix for si ,...,2= (A.5) The dual of ( P ) is constructed as ( D ) Minimize 0 U y (A.13) Subject to 1 2 3 01 1 ( ) 0 s i k kk k p y p p y y = = + + − for si ,...,2= (A.14) 1 1 s ii y = − − (A.15) 0 i y for si ,...,2= (A.16) 0 y unrestricted in sign (A.17) We can find a feasible solution to problem D by assuming that all constraints defined in the dual problem D are satisfied as equalities. It follows that a feasible solution to problem D is achieved when 1 1 2 3 1 2 31 1 1 1 ( ) ( ) s i s i k k k kk k k i k p y p p y p y p p y − = = = − = + + = + + for 2,...,i s= (A.18) or equivalently, 1 1 2 3 / ( ) i i y y p p p − = + for 2,...,i s= . (A.19) Solving the set of equations in (A.18) and 1 1 ii s y = simultaneously yields 1 2 1 1 (1 / (1 ... )) s s y − − = + ++ + + , (A.20) 1 1 i i y y − = for 2,...,i s= , (A.21) 2 1 0 1 2 3 / (1 ... ) ( ) s y p p p − = + + + + + + + (A.22) where ( ) 132 / ppp += . Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 168 The objective function of the dual problem 1D is obtained as 2 1 0 1 2 3 ( /(1 ... ) ( )) s U y p p p U − = + + + + + + . (A.23) From equations (A.10) and (A.23), it is clear that the problem P and its dual D have the same objective function values. Thus, from the duality theory, it follows that equations (A.8)-(A.10) and (A.20)-(A.22) are the optimal solutions to the primal and dual problems, respectively. Therefore, we can conclude that the sublot sizes given in the theorem statement are optimal for the single-job problem where 2M is the primary machine on which the third operation is performed. ■ Appendix B. Proof of Theorem 3. For a job sequence , let [j] = job lot sequenced in the j th position of the sequence , jU = lot size of job lot [j], j s = number of sublots of job lot [j], i = index for sublots ( 1,..., ji s= ), ,j ix = size of the i th sublot in job lot [j], k = index for machines ( 1, 2, 3k = ), ,j kp = item (unit) processing time for the k th operation of job lot [j], , ,j i kC = completion time of the k th operation for the i th sublot of job lot [j], jRI = time lag between the start of the first operation on 1M and the latest start time of the second operation on 2M for job lot [j], jRO = time lag between the completion times of the first and the third operations for job lot [j]. The completion time of the first operation for the i th sublot of job lot [j] is given by 1, ,1 , ,1 ,1 ,1j i j i j s j j rr C C p x − = = + for 1, 2,..., ji s= (B.1) where 00 , ,1 0 s C = . From equation (B.1), the completion time for the first operation of job lot [j], which is the completion time of the last sublot in this job on 1M , is obtained as 1 1, ,1 1 ), ,1 ,1 , 1 , ,1 ,11 j j j j s j s j s j j i j s j ji C C p x C p U − − − −= = + = + . (B.2) The completion time of the third operation for the first sublot of job lot [j] on 2M is expressed as 1,1,3 ,1,1 1 , ,3 ,2 ,1 ,3 ,1max , jj j j s j j j jC C C p x p x−−= + + Scheduling with lot streaming in a two-machine re-entrant flow shop 169 1,1,1 1 , ,3 ,2 ,3 ,1max , ( )jj j s j j jC C p p x−−= + + . (B.3) Substituting ,1,1jC value from (B.2) into (B.3) yields 1 1,1,3 1 , ,1 ,1 ,1 1 , ,3 ,2 ,3 ,1max , ( )j jj j s j j j s j j jC C p x C p p x− −− −= + + + . (B.4) Similarly, we may obtain the completion time of the third operation for the second sublot of job lot [j] on 2M as 1 1 1 ,2,3 1 , ,1 ,1 , ,2 ,3 , 1 , ,31 11 2 max max ( ) , j j i i j j s j j r j j j r j sr ri C C p x p p x C − − − − −= = = + − + 2 ,2 ,3 ,1 ( ) j j j ii p p x = + + (B.5) Repeating the process yields the completion time for the last operation of job lot [j], which is the completion time of the last sublot of this job lot on 2M , 1 1 , ,3 1 , ,1 ,1 , ,2 ,3 ,1 11 max max ( ) , j j j i i j s j s j j r j j j rr ri s C C p x p p x − − − = = = + − + 1 1 , ,3 ,2 ,3 ,1 ( )j j s j s j j j ii C p p x − − = + + 1 1 1 , ,1 ,1 , ,2 ,3 ,1 11 max max ( ) , j j i i j s j j r j j j rr ri s C p x p p x − − − = = = + − + 11 , ,3 ,2 ,3( )jj s j j jC p p U−− + + (B.6) By successive application of (B.6) using (B.2), the time to complete the last job lot processed on 2M , , ,3nn s C , is obtained as follows: 1 max, ,3 ,2 ,31 1 11 max ( ) n w w n n s j j j j jj j jw n C C RI RO p p U − = = = = = − + + (B.7) where 1 ,1 , ,2 ,3 ,1 11 max ( ) j i i j j j r j j j rr ri s RI p x p p x − = = = − + , (B.8) 1 ,1 , ,2 ,3 , ,2 ,3 ,11 11 max ( ) ( ) j i i j j j r j j j r j j j j jr ri s RO p x p p x p p U p U − = = = − + + + − ,2 ,3( )j j j j jRI p p p U= + + − (B.9) Note that the second part, ,2 ,31 ( ) n j j jj p p U = + , in (B.7) giving the makespan value is a constant so that it is enough to minimize the first part, 1 1 11 max w w j jj jw n RI RO − = = − , which is equivalent to the total idle time on 2M . Note that the first part is similar to Johnson’s expression, where the processing times on 1M and 2M are replaced by the run-in and run-out delays, respectively. Therefore, Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 170 job v precedes job z in an optimal schedule of job lots when min , minv zRI RO ,z vRI RO . ■ Appendix C. Proof of Theorem 4. It is clear that the minimizing the first term in (B.7), 1 ,1 , ,2 ,3 ,1 11 max ( ) j i i j j r j j j rr ri s p x p p x − = = − + , minimizes the makespan for a given sequence of job lots. In other words, sublot sizes minimizing the makespan for any arbitrary sequence (hence the optimal sequence) are identical to those sublot sizes which are determined by solving the sublot-sizing problem for each job lot separately. Therefore, the following linear programming model should be solved for every job lot in position [j]: Minimize jZ (C.1) Subject to 1 ,1 , ,2 ,3 ,1 1 ( ) i i j j j r j j j rr r Z p x p p x − = = − + for 1, 2,..., ji s= (C.2) ,1 j s j i ji x U = = (C.3) , 0j ix for 1, 2,..., ji s= . (C.4) The optimal solution to this model is trivial due to its special structure. The minimum value of the objective function jZ is achieved when ,1 ,1j j jZ p x= and the constraints in (C.2) are satisfied as equalities. Thus, the sublot sizes must be 1 , , 1 ,2 ,3 ,1 ,1 ,2 ,3 ,1 ( ) / (( ) / ) i j i j i j j j j j j j x x p p p x p p p − − = + = + for 2,..., ji s= . (C.5) Substituting (C.5) into (C.3) yields 12 ,1 / (1 ... ) j s j j x U − = + + + + , (C.6) 1 , ,1 i j i j x x − = for 2,..., ji s= (C.7) where ,2 ,3 ,1( ) /j j jp p p = + . Note that (C.6) and (C.7) are the same expressions as given in Section 3 for the single-job problem, and substituting (C.6) and (C.7) into (B.8) and (B.9) yields 1 ,1 , ,2 ,3 , ,1 ,11 11 max ( ) j i i j j j r j j j r j jr ri s RI p x p p x p x − = = = − + = 12 ,1 / (1 ... ) j s j j p U − = + + + + , (C.8) ,2 ,3( )j j j j j jRO RI p p p U= + + − . (C.9) Scheduling with lot streaming in a two-machine re-entrant flow shop 171 Acknowledgements: We would like to thank the editor and anonymous referees for their valuable comments and suggestions that helped us to improve the quality and presentation of this paper. References Alfieri, A., Glass, C. A., & van de Velde, S. L. (2012). Lot streaming in a two-machine flow shop with attached setup times. IIE Transactions, 44(8), 695–710. https://doi.org/10.1080/0740817X.2011.649384 Alfieri, A., Zhou, S., Scatamacchia, R., & van de Velde, S. L. (2021). Dynamic programming algorithms and Lagrangian lower bounds for a discrete lot streaming problem in a two-machine flow shop. 4OR, 19(2), 265–288. https://doi.org/10.1007/s10288-020-00449-8 Baker, K. R. (1988). Lot streaming to reduce cycle time in a flow shop, Working Paper, The Amos Tuck School of Business Administration, Dartmouth College, Hanover, NH, USA. Baker, K. R. (1995). Lot streaming in the two-stage flow shop with setup times. Annals of Operations Research, 57(1), 1–11. https://doi.org/10.1007/BF02099687 Buscher, U. & Shen, L. (2009). An integrated tabu search algorithm for the lot streaming problem in job shops. European Journal of Operational Research, 199(2), 385–399. https://doi.org/10.1016/j.ejor.2008.11.046 Çetinkaya, F. C. (1994). Lot streaming in a two-stage flow shop with set-up, processing and removal times separated. Journal of Operational Research Society, 45(12), 1445–1455. https://doi.org/10.1057/jors.1994.221 Çetinkaya, F. C. (2005). Unit sized transfer batch scheduling in an automated two- machine flow-line cell with one transport agent. International Journal of Advanced Manufacturing Technology, 29(1–2), 178–183. https://doi.org/10.1007/s00170- 004-2493-9 Çetinkaya, F. C., & Kayalıgil, M. S. (1992). Unit sized transfer batch scheduling with setup times. Computers & Industrial Engineering, 22(2), 177–183. https://doi.org/10.1016/0360-8352(92)90045-L Çetinkaya, F. C., & Duman, M. (2010). Lot streaming in a two-machine mixed shop. International Journal of Advanced Manufacturing Technology, 49(9–12), 1161–1173. https://doi.org/10.1007/s00170-009-2473-1 Chang, J. H., & Chiu, H. N. (2005). A comprehensive review of lot streaming. International Journal of Production Research, 43(8), 1515–1536. https://doi.org/10.1080/00207540412331325396 Chen, J., & Steiner, G. (1997). Lot streaming with detached setups in three-machine flow shops. European Journal of Operational Research, 96(3), 591–611. https://doi.org/10.1016/S0377-2217(96)00091-4 Chen, J., & Steiner, G. (1998). Lot streaming with attached setups in three-machine flow shops. IIE Transactions, 30(11), 1075–1084. https://doi.org/10.1023/A:1007563814941 Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 172 Cheng, M., Mukherjee, N. J., & Sarin, S. C. (2013). A review of lot streaming. International Journal of Production Research, 51(23–24), 7023–7046. https://doi.org/10.1080/00207543.2013.774506 Cheng, M., & Sarin, S. C. (2020). Lot streaming in a two-stage assembly system. Annual Reviews in Control, 50(23–24), 303–316. https://doi.org/10.1016/j.arcontrol.2020.08.004 Cheng, M., Sarin, S.C., & Singh, S. (2016). Two-stage, single-lot, lot streaming problem for a 1+2 hybrid flow shop. Journal of Global Optimization, 66(2), 263–290. https://doi.org/10.1007/s10898-015-0298-z Davendra, D., Senkerik, R., Zelinka, I., Pluhacek, M., & Bialic-Davendra, M. (2014). Utilising the chaos-induced discrete self organising migrating algorithm to solve the lot-streaming flowshop scheduling problem with setup time. Soft Computing, 18(4), 669–681. https://doi.org/10.1007/s00500-014-1219-7 Defersha, F. M., & Chen, M. (2010). A hybrid genetic algorithm for flowshop lot streaming with setups and variable sublots. International Journal of Production Research, 48(6), 1705–1726. https://doi.org/10.1080/00207543.2011.574952 Defersha, F. M. (2011). A comprehensive mathematical model for hybrid flexible flowshop lot streaming problem. International Journal of Industrial Engineering Computations, 2(2), 283–294. https://doi.org/ 10.5267/j.ijiec.2010.07.006 Defersha, F. M., & Chen, M. (2012a). Mathematical model and parallel genetic algorithm for hybrid flexible flowshop lot streaming problem. The International Journal of Advanced Manufacturing Technology, 62(1–4), 249–265. https://doi.org/10.1007/s00170-011-3798-0 Defersha, F. M., & Chen, M. (2012b). Jobshop lot streaming with routing flexibility, sequence-dependent setups, machine release dates and lag time. International Journal of Production Research, 50(2), 2331–2352. https://doi.org/10.1080/00207543.2011.574952 Feldmann, M., & Biskup, D. (2008). Lot streaming in a multiple product permutation flow shop with intermingling. International Journal of Production Research, 46(1), 197–216. https://doi.org/10.1080/00207540600930065 Glass, C. A., Possani, E. (2011). Lot streaming multiple jobs in a flow shop. International Journal of Production Research, 49(9), 2669–2681. https://doi.org/10.1080/00207543.2010.532935 Gomez-Gasquet, P., Segura-Andres, R., & Andres-Romano, C. (2013). A review of lot streaming in a flow shop environment with makespan criteria. Journal of Industrial Engineering and Management, 6(3), 761–770. http://dx.doi.org/10.3926/jiem.553 Johnson, S. M. (1954). Optimal two and three-stage production schedules with setup times included. Naval Research Logistics Quarterly, 1(1), 61–68. https://doi.org/10.1002/nav.3800010110 Kim, J. S., Kang, S. H., & Lee, S. M. (1997). Transfer batch scheduling for a two-stage flowshop with identical parallel machines at each stage. Omega-International Journal of Management Science, 25(5), 547–555. https://doi.org/10.1016/S0305- 0483(97)00015-7 Lev, V., & Adiri, I. (1984). V-shop scheduling. European Journal of Operational Research, 18(1), 51–56. https://doi.org/10.1016/0377-2217(84)90260-1 Scheduling with lot streaming in a two-machine re-entrant flow shop 173 Li, J. Q., Tao, X. R., Jia, B. X., Han, Y. Y., Liu, C., Duan, P. Zheng, Z. X., & Sanga, H. Y. (2020). Efficient multi-objective algorithm for the lot-streaming hybrid flowshop with variable sub-lots. Swarm and Evolutionary Computation, 5, Article 100600. https://doi.org/10.1016/j.swevo.2019.100600 Liu, J. (2008). Single-job lot streaming in m-1 two-stage hybrid flowshops. European Journal of Operational Research, 187(3), 1171–1183. https://doi.org/10.1016/j.ejor.2006.06.066 Marimuthu, S., Ponnambalam, S. G., Jawahar, N. (2008). Evolutionary algorithms for scheduling m-machine flow shop with lot streaming. Robotics and Computer- Integrated Manufacturing, 24(1), 125–139. https://doi.org/10.1016/j.rcim.2006.06.007 Marimuthu, S., Ponnambalam, S. G., Jawahar, N. (2009). Threshold accepting and Ant- colony optimization algorithms for scheduling m-machine flow shops with lot streaming. Journal of Materials Processing Technology, 209(2), 1026–1041. https://doi.org/10.1016/j.jmatprotec.2008.03.013 Martin, C. H. (2009). A hybrid genetic algorithm/mathematical programming approach to the multi-family flowshop scheduling problem with lot streaming. Omega, 37(1), 126–137. https://doi.org/10.1016/j.omega.2006.11.002 Meng, T., Pan, Q. K., Li, J. Q., & Sang, H. Y. (2018). An improved migrating birds optimization for an integrated lot-streaming flow shop scheduling problem. Swarm and Evolutionary Computation, 38, 64–78. https://doi.org/10.1016/j.swevo.2017.06.003 Naderi, B, & Yazdani, M. (2015). Modelling and Scheduling Lot Streaming Flexible Flow Lines. Journal of Optimization in Industrial Engineering, 8(18), 61–69. http://www.qjie.ir/article_221.html Nejati, M., Mahdavi, I., Hassanzadeh, R., & Mahdavi-Amiri, N. (2016). Lot streaming in a two-stage assembly hybrid flow shop scheduling problem with a work shift constraint. Journal of Industrial and Production Engineering, 33(7), 459–471. https://doi.org/10.1080/21681015.2015.1126653 Pan, Q. K., Tasgetiren, M. F., Suganthan, P. N., & Chua, T. J. (2011). A discrete artificial bee colony algorithm for the lot-streaming flow shop scheduling problem. Information sciences, 181(12), 2455–2468. https://doi.org/10.1016/j.ins.2009.12.025 Potts, C. N., & Baker, K. R. (1989). Flow shop scheduling with lot streaming. Operations Research Letters, 8(6), 297–303. https://doi.org/10.1016/0167- 6377(89)90013-8 Pranzo, M. (2004). Batch scheduling in a two-machine flow shop with limited buffer and sequence independent setup times and removal times. European Journal of Operational Research, 153(3), 581–592. https://doi.org/10.1016/S0377- 2217(03)00264-9 Reither, S. (1966). A system for managing job shop production. Journal of Business, 39(3), 371–393. Rossit, D., Tohmé, F., Frutos, M., Bard, J & Broz, D. (2016). A non-permutation flowshop scheduling problem with lot streaming: A Mathematical model. Çetinkaya & Duman /Oper. Res. Eng. Sci. Theor. Appl. 4 (3) (2021) 142-175 174 International Journal of Industrial Engineering Computations, 7(3), 507–516. https://doi.org/ 10.5267/j.ijiec.2015.11.004 Salazar-Moya, A., & Garcia, M. V. (2021). Lot Streaming in Different Types of Production Processes: A PRISMA Systematic Review. Designs, 5(4), 67. https://doi.org/10.3390/designs5040067 Sarin, S. C., & Jaiprakash, P. (2007). Flow shop lot streaming. Springer. http://dx.doi.org/10.1007/978-0-387-47688-9 Sarin, S. C., Yao, L., & Trietsch, D. (2011). Single-batch lot streaming in a two-stage assembly system. International Journal of Planning and Scheduling, 1(1–2), 90–108. https://doi.org/10.1504/IJPS.2011.044604 Sriskandarajah, C., & Wagneur, E. (1999). Lot streaming and scheduling multiple products in two-machine no-wait flow shops. IIE Transactions, 31(8), 695–707. https://doi.org/10.1023/A:1007643210135 Şen, A., & Benli, Ö. S. (1999). Lot streaming in open shops. Operations Research Letters, 23(3–5), 135–142. https://doi.org/10.1016/S0167-6377(98)00026-1 Trietsch, D. (1989). Polynomial transfer lot sizing techniques for batch processing on consecutive machines. Working Paper, Naval Postgraduate School, Monterey, California, USA. https://core.ac.uk/download/pdf/36722285.pdf Trietsch, D., & Baker, K. R. (1993). Basic techniques for lot streaming. Operations Research, 41(6), 1065–1076. https://doi.org/10.1287/opre.41.6.1065 Tseng, C. T., & Liao, C. J. (2008). A discrete particle swarm optimization for lot- streaming flowshop scheduling problem. European Journal of Operational Research, 191(2), 360–373. https://doi.org/10.1016/j.ejor.2007.08.030 Vickson, R. G. (1995). Optimal lot streaming for multiple products in a two-machine flow shop. European Journal of Operational Research, 85(3), 556–575. https://doi.org/10.1016/0377-2217(93)E0366-6 Vickson, R. G., & Alfredson, B. E. (1992). Two-and three-machine flow shop scheduling problems with equal sized transfer batches. International Journal of Production Research, 30(7), 1551–1574. https://doi.org/10.1080/00207549208948107 Wang, M., Sethi. S. P., & Van De Velde, S. L. (1997). Minimizing makespan in a class of re-entrant shops. Operations Research, 45(5), 702–712. https://doi.org/10.1287/opre.45.5.702 Wang, S., (1997). Minimizing makespan in a class of re-entrant shops. Operations Research, 45(5), 702–712. https://doi.org/10.1287/opre.45.5.702 Wang, S., Kurz, M., Mason, S. J., & Rashidi, E. (2019). Two-stage hybrid flow shop batching and lot streaming with variable sublots and sequence-dependent setups. International Journal of Production Research, 57(22), 6893–6907. https://doi.org/10.1080/00207543.2019.1571251 Yang, D. L., & Chern, M. S. (2000). Two-machine flowshop group scheduling problem. Computers & Operations Research, 27(10), 975–985. https://doi.org/10.1016/S0305-0548(99)00070-2 Yao, L., & Sarin, S. C. (2014). Multiple-lot, lot streaming in a two-stage assembly system. Essays in production, project planning and scheduling (pp. 357–388). Springer. https://doi.org/10.1007/978-1-4614-9056-2_15 Scheduling with lot streaming in a two-machine re-entrant flow shop 175 Yoon, S. H., & Ventura, J. A. (2002). An application of genetic algorithms to lot- streaming flow shop scheduling. IIE Transactions, 34(9), 779–787. https://doi.org/10.1080/07408170208928911 Zhang, B., Pan, Q. K., Gao, L., Zhang, X. L., Sang, H. Y., & Li, J. Q. (2017). An effective modified migrating birds optimization for hybrid flowshop scheduling problem with lot streaming. Applied Soft Computing, 52, 14–27. https://doi.org/10.1016/j.asoc.2016.12.021 Zhang, W., Liu, J., & Linn, R. J. (2003). Model and heuristics for lot streaming of one job in m-1 hybrid flowshops. International Journal of Operations and Quantitative Management, 9(1), 49–64. http://www.ijoqm.org/v9no1.asp Zhang, W., Yin, C., Liu, J., & Linn, R. J. (2005). Multi-job lot streaming to minimize the mean completion time in m–1 hybrid flowhops. International Journal of Production Economics, 96(2), 189–200. https://doi.org/10.1016/j.ijpe.2004.04.005 © 2021 by the authors. Submitted for possible open access publication under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).