634
Views
0
CrossRef citations to date
0
Altmetric
Production & Manufacturing

A novel Fibonacci-inspired enhancement of the Bees Algorithm: application to robotic disassembly sequence planning

ORCID Icon & ORCID Icon
Article: 2298764 | Received 26 Aug 2023, Accepted 17 Dec 2023, Published online: 18 Jan 2024

Abstract

The Bees Algorithm (BA) is an intelligent, nature-inspired metaheuristic algorithm first introduced in 2005 and based on the foraging activity of honeybees. Enhancements to the BA have been made continually since its introduction, with some aimed at reducing the number of user-determined parameters. The goal is to achieve optimal results without extensive efforts required for parameter tuning. This article presents an enhanced version of the BA. The enhancement uses the Fibonacci sequence, which can be found in the family tree of the male bee (drone), to give the numbers of bees assigned to conduct local searches. The proposed BA is simple, having only four parameters that must be set by the user, as opposed to five in the basic combinatorial version. To evaluate the performance of the new algorithm in robotic disassembly sequence planning, a case study is conducted on gear pump disassembly, a commonly used benchmark problem in robotic disassembly. A comparison is made between the proposed Fibonacci Bees Algorithm and the Basic Bees Algorithm (BBA). The results are assessed using a statistical performance metric and a performance evaluation index. The findings demonstrate that the proposed algorithm is better than the BBA for more complex problems while exhibiting a similar performance on simpler tasks. Future research will explore broader applications for both continuous and combinatorial problems, demonstrating the versatility of the algorithm across different domains.

1. Introduction

The Bees Algorithm (BA) has demonstrated success in solving complex real-world problems (Castellani & Pham, Citation2023; De la Torre Gutiérrez & Pham, Citation2018). The BA, developed in 2005 (Pham et al., Citation2005, Citation2006), draws inspiration from the foraging behaviour of honeybees. The BA employs a population of artificial ‘bees’ to navigate and exploit the search space, thereby emulating the foraging behaviour exhibited by bees in their quest for nectar sources. Since its inception, the BA has garnered considerable attention from researchers and has been applied in numerous domains (Castellani & Pham, Citation2023).

However, a problem with most versions of the BA is that they have many parameters (more than five) which the user has to set, making the algorithm difficult for the uninitiated to apply. This article presents a new version of the BA with just four user-selected parameters. In the proposed algorithm, the numbers of bees assigned to conduct local searches are determined with the help of the Fibonacci sequence which can be observed in the family tree of a male bee (drone). This natural inspiration has not only expanded the landscape of optimisation techniques but also furnished researchers and practitioners with a versatile tool adaptable to a wide range of complex problems across many domains. This study introduces a novel approach that significantly contributes to the parameter configuration and enhancement of the BA.

The article is structured as follows: Section 2 provides an overview of the BA and its variants, highlighting applications in the context of robotic disassembly planning. Section 3 focusses on the proposed enhancement of the BA, drawing inspiration from the Fibonacci sequence to yield the Fibonacci Bees Algorithm or BAF. A review of other metaheuristics inspired by Fibonacci is provided to highlight the novelty of BAF. In Section 4, the chosen case study is presented along with a detailed justification for its selection. The evaluation of the performance of BAF is elaborated upon in Section 5. Section 6 presents the experiments conducted using both the BA and the enhanced version in RDSP. Section 7 discusses the results obtained. Finally, the study concludes by summarising the key findings and insights obtained from the research and suggesting areas for further investigation.

2. Bees Algorithm

The BA, originally developed for continuous problems, has six user-determined parameters (Pham et al., Citation2005) and is commonly referred to as the Basic Bees Algorithm (BBA). The pseudocode of the BBA (Pham et al., Citation2006) is given in Algorithm 0.

Algorithm 0:

Basic Bees Algorithm Pseudocode

Input: n: number of scout bees, m: number of selected sites, e: number of elite sites, nsp: recruited bees for other selected sites, nep: recruited bees for elite sites, ngh: initial size of neighbourhood

1 Function BBA(n, m, e, nsp, nep, ngh):

2 InitialisePopulation with random solutions

3 EvaluateFitness of the population

4 while stopping criterion not met do

5  Forming new population

6  SelectSites for neighbourhood search

7  RecruitBees for selected sites (more bees for the best e sites) and evaluate fitness

8  SelectFittestBees from each patch

9  AssignRandomBees to search randomly and evaluate their fitness

10 end

11 return BestBee

The neighbourhood (ngh) for combinatorial problems depends on the design of the local search, which may involve operators like swap, insert, reverse, mutation, 2-OPT and 3-OPT. In the combinatorial version of BBA (Ismail et al., Citation2020; Liu et al., 2018a, b; Zeybek & Koç, Citation2015; Zeybek et al., Citation2021), where the swap, insert, and reverse operators are used, the neighbourhood size is considered equal to the sequence length. Therefore, for the combinatorial version, only five parameters need to be set: n, m, e, nsp and nep.

The introduction of site abandonment and neighbourhood shrinking to BBA, resulting in two additional parameters that need to be configured (Castellani & Pham, Citation2023; Pham & Castellani, Citation2009, Citation2015), represents a noteworthy enhancement and gives rise to a variant of the algorithm known as the Standard Bees Algorithm (SBA). SBA requires 7–8 parameters to be set by the user. Alternative methods for recruitment, neighbourhood modification and site abandonment have been extensively explored in the literature (Castellani & Pham, Citation2023). For a comprehensive understanding of the variants of the BA developed until 2017, interested readers are encouraged to refer to the survey paper by Hussein, Sahran and Sheikh Abdullah, which provides an in-depth analysis of the various modifications and enhancements in the algorithm (Hussein et al., Citation2017).

The utilisation of metaheuristics has been the subject of critique owing to their dependence on parameter values. This reliance can result in a lengthy procedure, as the optimal settings for each problem must be determined while considering their unique attributes. Previous researchers opted for the application of Taguchi and Design of Experiments (DoE) methods to discern the optimal parameter setting. However, these techniques are time-consuming (Hartono, Ismail et al., Citation2023). Several efforts have been made to reduce the parameter tuning in the BA, such as the application of fuzzy selection for self-tuning (Pham & Darwish, Citation2008). The Ternary Bees Algorithm, introduced by Laili et al. (Citation2019), uses only three bees and incorporates a single parameter setting for the site abandonment threshold. Another reduced-parameter version of BA is BA2, which has two parameters. BA2 is inspired by the traplining foraging behaviour of honeybees (Ismail, Citation2021; Ismail & Pham, Citation2023). The achievement of an adaptive parameter-free version and the potential for further parameter removal remain uncertain in the pursuit of enhanced versions of the BA (Castellani & Pham, Citation2023). However, it is apparent that reducing the number of parameters significantly decreases the effort required for parameter tuning.

BA2, initially developed for continuous problems and later extended to address combinatorial problems by incorporating specialised local operators tailored for the vehicle routing problem (VRP), has been previously investigated by Ismail (Citation2021). However, the primary focus of this study is on application to robotic disassembly problems. Robotic Disassembly Sequence Planning (RDSP) entails planning the sequence and actions required to disassemble objects, considering factors such as interdependencies, sizes, shapes, connections, accessibility, tools and multiple dimensions. These inherent complexities differentiate robotic disassembly from the VRP, which typically deals with fewer dimensions. Therefore, direct comparison with both the continuous version of BA2 or its combinatorial version of VRP falls outside the scope of this research. While earlier research by Ismail (Citation2021) drew inspiration from the traplining behaviour of bees to reduce the parameter count of the BA, this study takes a different approach by drawing inspiration from the Fibonacci sequence-based family tree pattern of honeybees. Building upon this unique source of inspiration, the proposed enhancement to the algorithm introduces a new configuration with four parameters. By exploring this alternative parameter setting, the research aims to investigate potential benefits and performance improvements. This novel approach contributes to the field by offering a fresh perspective on parameter configuration in the BA for robotic disassembly problems.

Numerous methodologies and algorithms have been developed to address the challenges posed by the problem of optimally planning disassembly (Gao et al., Citation2020). A limited number of previous studies have utilised exact methods to address either simple or prototypical problems (Ming et al., Citation2019; Yin et al., Citation2022). While mathematical and exact methods can provide optimal solutions for small-scale instances, their applicability to larger problems is limited by the nondeterministic polynomial (NP)-hard nature of the problem (Cevikcan et al., Citation2020; Gao et al., Citation2020; Guo et al., Citation2022; McGovern & Gupta, Citation2011). The number of possible subassemblies (Nn), the number of complete disassembly sequences (cn) and the total number of disassembly sequences (Pn) in a disassembly problem can be determined theoretically using EquationEqs. (1)–(3) (Lambert, Citation1997). For instance, with four parts, there are 15 subassemblies, 15 complete disassembly sequences and a total of 41 disassembly sequences. In the case of 10 parts, the number of subassemblies increases to 1023, the complete disassembly sequences amount to 34,459,425, and the total number of disassembly sequences is 314,726,297. These theoretical calculations demonstrate the exponential growth in the number of disassembly sequences as the number of products increases. Such insights shed light on the combinatorial nature of disassembly problems and underscore the challenges associated with exploring all possible disassembly sequences. (1) Nn=(2.Nn1)+1(1) (2) cn=(2n3).cn1(2) (3) P1=1P2=12 (12).P1+1P3=(13).P2+1P10=(110).P9+(210).P2.P8+(310).P3.P7+(410).P4.P6+12.(510).P5.P5+1(3)

Another theoretical calculation suggests that in the context of planning the optimal sequence for a product with n parts, the exploration of the search tree involves examining n! nodes, and the collision checking operation requires k x n! computations, where k represents the average number of directions tested for component removal (Costa et al., Citation2018). These theoretical calculations, coupled with the findings from previous research, provide compelling evidence to support the assertion that the disassembly problem is inherently complex and exhibits exponential growth in complexity. Metaheuristic algorithms have garnered significant attention from researchers due to their effectiveness in navigating the expansive search space of disassembly problems and achieving near-optimal solutions (Cevikcan et al., Citation2020; Gao et al., Citation2020; Ma et al., Citation2011; McGovern & Gupta, Citation2011).

The prevalence of metaheuristic approaches is evident in robotic disassembly research, where many researchers have adopted them (Gao et al., Citation2020; Hartono, Citation2023). Notably, metaheuristic algorithms offer the advantages of relatively short computation times, meeting practical requirements (Gao et al., Citation2020) and facilitating efficient decision-making in real-world scenarios (Laili et al., Citation2022b). In a real-world context, the efficient discovery of near-optimal solutions holds greater significance than achieving exact solutions. Moreover, real products exhibit complexity that further amplifies the challenge. During the initial stages of research, many studies employed toy problems to explore the mathematical and conceptual aspects, which is understandable considering the inherent complexity of the disassembly problem.

The BA has demonstrated its robustness and effectiveness in a wide range of remanufacturing applications (Caterino et al., Citation2022, Citation2023; Kerin et al., Citation2023; Zeybek, Citation2023), establishing it as a reliable optimisation approach within this domain. In particular, in the field of robotic disassembly, the BA is recognised as one of the most popular metaheuristic algorithms (Chen, Xu, et al., Citation2020; Cui et al., Citation2023; Hartono et al., Citation2022a, Citation2022b, Citation2023; Laili et al., Citation2019, Citation2022a, Citation2022b; Liu, Zhou, Pham, Xu, Ji, et al., 2018a; Liu, Zhou, Pham, Xu, Yan, et al., 2018b; Liu et al., Citation2020; Liu, Liu, et al., Citation2023; Liu, Xu, et al., Citation2023; Yang et al., Citation2022; Ye et al., Citation2022). Consistently outperforming other algorithms, the BA is highly regarded for its exceptional performance and capabilities. Thus, it remains a compelling choice for addressing optimisation problems in robotic disassembly. However, so far, only one robotic disassembly study (Laili et al., Citation2019) has been conducted on parameter reduction.

To ensure a meaningful and relevant evaluation, this research appropriately employs a comparison with Enhanced Discrete Bees Algorithm (EDBA), which is the most commonly adopted BA for RDSP. This choice enables an assessment of the effectiveness of the proposed enhancements in addressing the specific challenges posed by robotic disassembly problems. By aligning with the objectives and scope of this study, this approach facilitates a comprehensive examination of the proposed enhancements and their impact on addressing the complexities associated with robotic disassembly.

As previously mentioned, the BA was initially developed for the continuous domain. However, when the algorithm is applied to combinatorial problems such as the Travelling Salesman Problem (TSP), the local search component needs to be modified accordingly. In the case of the TSP, the BBA incorporates local search operators such as swap, insert and reverse (Ismail et al., Citation2020), which are specifically designed to manipulate the sequence of cities. Similarly, in the context of robotic disassembly problems, the BBA’s local search is adapted to include swap, insert and mutation operators (Liu, Zhou, Pham, Xu, Ji, et al., 2018a; Liu, Zhou, Pham, Xu, Yan, et al., 2018b) that are tailored to address the specific requirements and constraints of these combinatorial problems. In this research, the BA and the proposed BA utilise the swap, insert and mutation operators in their local search, as described by Liu, Zhou, Pham, Xu, Ji, et al. (2018).

3. Fibonacci Bees Algorithm (BAF)

Leonardo Fibonacci, a mathematician from the 13th century, is widely recognised for his significant contributions to the field of mathematics. One of his most notable achievements was the introduction of the Fibonacci sequence (Scott & Marketos, Citation2014). This sequence is a series of numbers, where each number is the sum of the two preceding numbers. It begins with 0 and 1, and the subsequent numbers are generated by adding the previous two numbers. The sequence progresses as follows: 0, 1, 1, 2, 3, 5, 8, 13, 21, 34 and subsequent numbers. The pattern extends indefinitely, revealing a sequence that possesses unique mathematical properties and occurs in diverse natural phenomena. The pattern can be found in a variety of natural phenomena, including leaf arrangements, spirals in seashells, tree branches and flower petals (Omotehinwa & Ramon, Citation2013; Scott & Marketos, Citation2014). The bee lineage parallels the Fibonacci sequence, establishing a link between bee ancestry and the numerical pattern seen in the Fibonacci sequence. The family tree of a male bee or drone exhibits a fascinating relationship with the Fibonacci sequence, as the number of ancestors in each successive generation corresponds to the Fibonacci numbers (Omotehinwa & Ramon, Citation2013; Scott & Marketos, Citation2014). Drones originate from unfertilised eggs through a reproductive process known as parthenogenesis. Parthenogenesis involves the development of embryos without fertilisation, resulting in male offspring that possess a mother but do not have a father. In contrast, female bees, which include both worker bees and queen bees, originate from fertilised eggs. illustrates the family tree of a drone in a simple ancestry model (Koshy, Citation2017; Scott & Marketos, Citation2014). Interestingly, upon counting the total number of bees in each generation, it is revealed that they conform to the Fibonacci sequence. This correlation emphasises the prevalence of Fibonacci patterns in honeybee population and the drone’s mathematical structure in its genealogical lineage.

Figure 1. Fibonacci sequence in the family tree of a drone (adapted from Koshy, Citation2017).

Figure 1. Fibonacci sequence in the family tree of a drone (adapted from Koshy, Citation2017).

To investigate whether previous enhancements of the BA have incorporated the inspiration from the Fibonacci sequence, given its relevance to bees, the following steps were conducted. An extensive search was conducted using the Scopus database, employing the keywords “bee” and “Fibonacci”. This search yielded a total of nine articles; however, none of them were found to be directly relevant to the BA. Additionally, a search was performed using the Scite database, but no pertinent results were found regarding the utilisation of the Fibonacci sequence in the BA. These searches across multiple databases indicate that, as of the present, there is no scholarly evidence to suggest that prior versions of the BA have integrated the Fibonacci sequence as a means of enhancement.

Further exploration was conducted using the Scopus database, employing the keywords “Fibonacci” and “heuristic”, with the aim of identifying algorithms that draw inspiration from Fibonacci. The initial search yielded a total of 62 articles. After refining the search based on titles and language criteria (English only) and availability of content, 43 articles were retained for further analysis. Upon examining the contents, nine articles were deemed irrelevant and were excluded from the analysis. Additionally, 18 articles that were identified as improvements to the original algorithm were also removed. Of the remaining articles, 15 focussed on metaheuristic algorithms, many of which were improved or hybridised with other algorithms, such as the Genetic Algorithm (GA) (Gudla & Ganguli, Citation2005; Sha & Pan, Citation2018; Yeh, Citation2006), Salp Swarm Algorithm (SSA) (Tran-Ngoc et al., Citation2023), Wolf Pack Algorithm (WPA) (Zhao et al., Citation2020), Tabu Search (Aras & Aksen, Citation2008), Particle Swarm Optimisation (PSO) (Khosla & Verma, Citation2022) and Grey Wolf Optimiser (GWO) (Khosla & Verma, Citation2022). These articles covered a range of inspirations related to Fibonacci, including the Fibonacci search, Fibonacci heaps, Fibonacci trees, the golden ratio and the application of Fibonacci indicators in the stock market.

summarises the Fibonacci inspirations of these algorithms and their applications. The Automated Hybrid Genetic Algorithm (AHGA) is utilised for benchmark function optimisation, wherein a combination of a GA and a local optima avoidance mechanism is employed (Gudla & Ganguli, Citation2005). Fibonacci numbers are used to select the number of GA banks in AHGA. The Memetic Algorithm (MA) is utilised for solving the multi-stage supply chain network problem, with the option of Fibonacci number generation to select the best offspring regardless of its origin (Yeh, Citation2006). The Fibonacci Tree Optimisation (FTO) algorithm is designed specifically for power grid optimisation, utilising a data storage structure known as a Fibonacci tree (Da et al., Citation2018). The Fibonacci indicator algorithm (FIA) finds its application in benchmark function optimisation, incorporating Fibonacci retracement and time zone techniques commonly utilised in stock market trading (Etminaniesfahani et al., Citation2018). For wellbore trajectory optimisation in the oil and gas industry, the Fibonacci sequence-based quantum GA (FSQGA) is employed, reflecting the rotation angle step in the quantum GA (Sha & Pan, Citation2018). The Golden Ratio Optimisation Method (GROM) is applied to benchmark function optimisation, with the movement direction of the algorithm following the formulation of the golden ratio (Nematollahi et al., Citation2020). Similarly, the Improved Wolf Pack Algorithm (IWPA) is employed in benchmark function optimisation, where the step length of scout wolves is adjusted based on the Fibonacci sequence (Zhao et al., Citation2020). The Child Drawing Development Optimisation (CDDO) algorithm uses the golden ratio to calculate two solution points in benchmark function optimisation (Abdulhameed & Rashid, Citation2022). The Opposition-based Learning PSO GWO (Opp-PSOGWO) algorithm is applied to benchmark function optimisation, generating an opposite population for searching using the Fibonacci sequence (Khosla & Verma, Citation2022). In the area of bridge structural health monitoring, a Fibonacci sequence-based optimisation algorithm (FSOA) rearranges the population of the algorithm using the golden ratio (Tran-Ngoc et al., Citation2023). Additionally, the table includes algorithms employed for solving the collection centre location problem inspired by the Fibonacci search (Aras & Aksen, Citation2008), project scheduling problem using the Multi-Objective Fibonacci-Based Algorithm (MOFA) inspired by the Fibonacci sequence (Hosseinian & Baradaran, Citation2020), and VRP using Brand Local Search Algorithm (BLSA), Adaptive Variable Neighbourhood Search (AVNS) and Evolutionary Local Search (ELS) that inspired by Fibonacci heaps (Wei et al., Citation2014; Zachariadis & Kiranoudis, Citation2010; Zhang et al., Citation2015).

Table 1. Metaheuristics inspired by Fibonacci.

The findings presented in provide additional support for the distinctiveness of the proposed Fibonacci Bees Algorithm (BAF) compared to previous research. While previous algorithms utilised Fibonacci numbers in various ways, such as selecting the number of banks, reflecting rotation angles, adjusting step lengths and generating populations, BAF introduces a novel framework specifically inspired by the Fibonacci sequence. In the process of improving BA, various ideas were experimented with and tested to identify improvements. The chosen and most effective idea is outlined below.

BAF is based on the observed Fibonacci sequence-based family tree pattern in drones, where the number of drone ancestors follows the Fibonacci sequence, as mentioned earlier. However, BAF introduces a different approach. Instead of employing the Fibonacci sequence to count ancestors, BAF uses it to determine the number of bees sent to the flower patches. This decision is motivated by the inherent growth pattern represented by the Fibonacci sequence, which aligns with the objective of maximising the foraging efficiency of the algorithm. By employing a ranking-based approach that recruits bees according to the Fibonacci sequence for targeting the most promising patches, BAF aims to exploit the patches more effectively, potentially yielding improved results. This departure from the conventional interpretation of the Fibonacci sequence within the context of drone ancestry showcases an innovative adaptation of the concept to address the specific requirements of the proposed algorithm.

In addition, unlike the original BA, which employs “elite” and “other selected” sites, BAF focusses solely on the selected sites, eliminating the need to distinguish between “elite” and “other selected” categories in the selected sites. Within these selected sites (sites with the best fitness values), bees conduct their search, and the allocation of bees to exploit the selected sites (nr) is determined by a ranking system that follows the Fibonacci sequence. The top-ranked bee receives the maximum number from the specified Fibonacci sequence while lower-ranking bees are assigned decreasing numbers from the sequence. For example, with three patches, m=3 and maxnr=8. The 1st patch attracts 8 recruited bees (nr1 = maxnr=8), the 2nd patch attracts 5 recruited bees (nr2 = 5) and the 3rd patch attracts 3 recruited bees (nr3 = 3). Although it is not certain if the allocation of foragers in nature follows the Fibonacci sequence, this differential allocation does reflect the fact that more foragers are allocated to higher-quality flower patches (Shackleton et al., Citation2023; Visscher & Seeley, Citation1982). After a specified number of tries, if the fittest bee in a patch remains unchanged, nr for that patch becomes 0 and a fresh set of patches is initialised.

Both BBA and EDBA share the same parameters and structure, with the only difference being the incorporation of a check for disassembly sequence feasibility after random initialisation. By disregarding that minor point, comparing EDBA and BAF becomes equivalent to comparing BBA and BAF. This article will compare the proposed BAF and EDBA in the context of RDSP. The problem is the minimisation of disassembly time. The comparison will utilise the Statistical Performance Metric (SPM) and the Performance Evaluation Index (PEI) presented in Hartono (Citation2023). The objective function (disassembly time) and number of function of evaluations (NFE) will be recorded for comparison purposes. In this work, the input data is a disassembly information matrix based on the modified interference matrix proposed by Liu, Zhou, Pham, Xu, Ji, et al. (2018a). This matrix ensures that the disassembly process takes into consideration the presence of fasteners while being feasible by adhering to precedence constraints. Using the matrix thus produces feasible disassembly sequences. In Algorithm 1 (EDBA) and Algorithm 2 (BAF), the input is the robotic disassembly information matrix, denoted as dis_m.

Algorithm 1:

The pseudo-code of EDBA for RDSP

Input: n: number of scout bees, m: number of selected sites, e: number of elite sites, nep: recruited bees for elite sites, nsp: recruited bees for other selected sites, dis_m: robotic disassembly information matrix

Output: RDSP(sequence, direction, tool, f:cost)

1 Function EDBA(n, m, e, nsp, nep):

2  Start

3  initialRDSPGlobalFDS(dis_m: sequence, direction)  // Generate initial population of feasible disassembly sequences (FDS)

4  while stopping criterion not met do

5    Evaluate population fitness

6    f ← FVALUE(initialRDSP)

7    Sort population according to f

8    Select m sites for local search

     // Generate local sites with waggle dance

9   for EliteSite(1 to e) do

     // Assign best elite site bee

10    BestEliteSiteBee ← the scout bee that found the elite site

11    for RecruitedEliteSiteBee(1 to nep) do

      // Do feasibility check

12     while feasibility not met do

13       RecruitedEliteSiteBeeWaggleDance(dis_m : sequence, direction)

      //Mutate the disassembly direction

14     RecruitedEliteSiteBeeMutation(dis_m : direction)

15     Evaluate fitness of RecruitedEliteSiteBee

16     if RecruitedEliteSiteBee is better than BestEliteSiteBee then

        //Update BestEliteSiteBee

17       BestLocalBee = RecruitedEliteSiteBee

18   for OtherSelectedSite(1 to (m-e)) do

     // Assign best other selected site bee

19    BestOtherSelectedSiteBee ← the scout bee that found the other selected site

20    for RecruitedOtherSelectedSiteBee (1 to nsp) do

      // Do feasibility check

21     while feasiblity not met do

22        RecruitedOtherSelectedSiteBeeWaggleDance(dis_m : sequence, direction)

       // Mutate the disassembly direction

23     RecruitedOtherSelectedSiteBeeMutation(dis_m : direction)

24     Evaluate fitness of RecruitedOtherSelectedSiteBee

25     if RecruitedOtherSelectedSiteBee is better than

        BestOtherSelectedSiteBee then

         // Update BestOtherSelectedSiteBee

26        BestLocalBee = RecruitedOtherSelectedSiteBee

    // Assign remaining scout bees for global search

27   for RemainingScoutBee(1 to (n-m)) do

28     RemainingScoutBee ← GlobalFDS(dis_m : sequence, direction)

29   Evaluate fitness of the new population

30   Sort population according to f

      // Store the best RDSP with minimum cost f

31   Best RDSP = Bee with minimum cost f

32  return Best RDSP (Bee with minimum cost f)

Algorithm 2:

The pseudo-code of BAF for RDSP

Input: n: number of scout bees, m: number of selected sites, nr: number of bees recruited for selected sites using ranking based recruitment, max_rv: maximum number of re-visits before the nr is set to zero, dis_m: robotic disassembly information matrix

Output: RDSP(sequence, direction, tool, f:cost)

1 Function BAF(n, m, nr, max_rv):

2  Start

3  initialRDSPGlobalFDS(dis_m: sequence, direction)  // Generate initial population of feasible disassembly sequences (FDS)

4  revisit = 0

5  while stopping criterion not met do

6    Evaluate population fitness

7    f ← FVALUE(initialRDSP)

8    Sort population according to f

9    Select m sites for local search

    // Generate local sites with waggle dance

10   for SelectedSite(1 to m) do

      // Assign best local bee

11     BestLocalBee ← the scout bee that found the site

      // Allocation of bees to the selected sites

12     nrnumber from Fibonacci sequence

13     for RecruitedBee (1 to nr) do

        // Do feasibility check

14       while feasiblity not met do

15         RecruitedBeeWaggleDance(dis_m : sequence, direction)

        // Mutate the disassembly direction

16       RecruitedBeeMutation(dis_m : direction)

17       Evaluate fitness of RecruitedBee

18       if RecruitedBee is better than BestLocalBee then

          // Update BestLocalBee

19          BestLocalBee = RecruitedBee

20       else

21         revisit ← revisit + 1

22         if revisitmax_rv then

            // Abandon site and randomly generated a new site

23           NewSiteFDS(dis_m : sequence, direction)

    // Assign remaining scout bees for global search

24   for RemainingScoutBee(1 to (n-m)) do

25      RemainingScoutBee ← GlobalFDS(dis_m : sequence, direction)

26   Evaluate fitness of the new population

27   Sort population according to f

    // Store the best RDSP with minimum cost f

28   Best RDSP = Bee with minimum cost f

29   return Best RDSP (Bee with minimum cost f)

4. Case study description

The case study chosen for this research is focussed on industrial gear pumps, which have been frequently used in previous studies on robotic disassembly of EoL products (Belhadj et al., Citation2022; Chen, Xu, et al., Citation2020; Chen, Yao, et al., Citation2020; Frizziero et al., Citation2022; Hartono et al., Citation2022a, Citation2022b, Citation2023; Laili et al., Citation2022c; Liu et al., Citation2018a, Citation2018b, Citation2020; Liu, Liu, et al., Citation2023; Liu, Xu, et al., Citation2023; Liu & Wang, Citation2017; Ramírez et al., Citation2020). Furthermore, the broad industrial application and low wear of gear pumps make them an ideal candidate for demonstrating the optimisation results and studying complete disassembly without destruction, highlighting their suitability as a test subject in this research. The results of this study on the robotic disassembly of gear pumps contribute to the broader context of robotic disassembly research, as they demonstrate the potential of using optimisation algorithms to disassemble EoL products efficiently and sustainably.

The two gear pumps in this research have different flow rates. Gear Pump A has a flow rate of 7.5 l/min, while Gear Pump B has a flow rate of 10 l/min. The data utilised in this study were derived from academic sources, and the 3D models were acquired from Grabcad (Grabcad Community, Citation2020; Liu, Zhou, Pham, Xu, Ji, et al., 2018a; Ramírez et al., Citation2020). The assumption for this study, taken from Liu, Zhou, Pham, Xu, Ji, et al. (2018a), is as follows: (1) the disassembly product is the same with geometric information and components that are known and applicable for repetitive disassembly; (2) All parts can be disassembled completely. The properties of the gear pumps can be found in .

Table 2. Properties of Gear Pumps A and B.

5. Performance evaluation

The literature review reveals that prior research has not fully utilised the potential of statistics, as they were only applied to the final outcomes. In addition, contradictory results have been observed across performance metrics (Laili et al., Citation2021). To bridge these identified gaps, two approaches are used in this research. One approach involves the application of statistical tests, while the other utilises a straightforward yet versatile metric. These proposed methods aim to address shortcomings and provide valuable insights into the performance analysis of the algorithm. The specifics of the first approach are elaborated upon in Section 5.1, while the details of the second approach are provided in Section 5.2.

5.1. Statistical performance metric

The SPM, introduced by Hartono et al., employs a statistical test not only for evaluating algorithm performance but also for identifying optimal parameter settings (Hartono et al., Citation2022a). This method follows a well-defined selection process for the appropriate statistical test. The advantage of this approach lies in its ability to statistically analyse observed differences and determine the optimal parameter settings for the chosen algorithm. The methodology is outlined as follows: Initially, the results are visualised using descriptive statistics. Subsequently, the assumption checklist is performed. If the number of experiments conducted exceeds 30, it is necessary to test whether the data adhere to certain assumptions, such as normality and homogeneity. In the case of any assumption violations, a nonparametric test is carried out. Conversely, if the number of experiments is below 30, a nonparametric test is employed. Finally, if the results indicated statistical significance, a post hoc test was conducted. The decision process can be observed in Algorithm 2. The choice of software for conducting the statistical analysis is dependent on the researcher’s preference and may involve the use of either commercial or freely available statistical software. In this research, statistical tests were performed using IBM SPSS Statistics 29 and MATLAB 2020b.

By employing this method, researchers can rigorously analyse the performance of an algorithm and identify the optimal parameter settings in a statistically sound manner. This systematic approach adds credibility to the experimental results and contributes to the advancement of algorithm optimisation techniques.

Algorithm 3:

Statistical Performance Metric

Require: data

1  Function Perform Descriptive Statistics And Box plot (data):

2  Calculate  descriptive statistics for the data;

3  Generate a boxplot to visually represent the results;

4  Function PerformAssumptionChecklist(data):

5   Check  the data size;

6   if the data size is below the threshold (eg 30), then

7    Perform  a nonparametric test;

8    Choose an appropriate nonparametric test;

9    return;

10   end

11   Test  for normality;

12   Test  for homogeneity of variances;

13  Function DecisionStepOne(data):

14   assumptionFail   false;

15   PerformAssumptionChecklist(data);

16   if any assumption fails (data size, normality, or homogeneity) then

17    Perform a nonparametric test based on the failed assumption;

18    assumptionFail   true;

19   end

20   if all assumptions are met and assumptionFail is false then

21    Perform  a parametric test;

22   end

23  Function DecisionStepTwo (result):

24   if result of the test is statistically significant then

25    Conduct post hoc tests to determine specific group differences;

26    return;

27   end

28     return;

29  Main Algorithm;

30  PerformDescriptiveStatisticsAndBoxplot(data);

31  DecisionStepOne(data);

32  DecisionStepTwo (result);

5.2. Performance Evaluation Index

As previously mentioned, relying solely on diagrams or figures to illustrate these indicators is insufficient, as it neglects the possibility of conflicting indicators, an aspect that has not received sufficient attention in scholarly investigations. As previously mentioned, there is a possibility of conflicting results obtained with different indicators (Laili et al., Citation2021). This research uses the PEI, a new index for evaluating the performance of algorithms introduced by Hartono et al. (Citation2023). The PEI can be used for MO as well as SO optimisation. The concept behind the proposed PEI is derived from formulas commonly used in the Multiple-Criteria Decision-Making (MCDM) literature. The index provides decision makers with a valuable tool for expediently evaluating multiple criteria and facilitating prompt decision-making. To calculate the index, two approaches are discussed in the MCDM literature. The first approach involves assigning weights to different criteria and summing the scores to obtain an overall score. However, this approach has certain limitations, including the need for additional computations such as normalisation and the inadequate accommodation of conflicting objectives related to specific criteria maximisation or minimisation. To address these limitations, Tofallis (Citation2014) introduced the second approach, known as the multiplicative approach. This approach resolves the limitations of the first approach by considering the multiplicative combination of scores, resulting in a simplified computation process and improved accommodation of conflicting objectives. In contrast to the additive method, the multiplicative approach does not require the normalisation or re-scaling of criteria, as the final outcome is unaffected by these operations (Tofallis, Citation2014). The multiplicative approach is followed by the PEI methodology, whereby the PEI is derived through the multiplication of indicators that are desired to have higher values and the division of indicators that are preferred to have lower values. Equal weights (ω) were assigned to all the functions, as all the indicators were deemed to be of equal significance. It should be noted that the value of ω is subject to the discretion of decision makers, who may set it based on their individual preferences.

As previously mentioned, the PEI can be used for both SO and MO problems. The example provided here pertains to a MO problem, while the next section will show the use of PEI in an SO application. In this example, a common MO nondominated performance metric is considered: the Hypervolume Indicator (HI), the Pareto Optimal Solutions (POSs) and NFE. A higher HI is desirable because it signifies a wider range of POSs. Having a higher number of POSs is also desirable. On the other hand, NFE serves as a reliable measure of computational complexity and is independent of the computer system. In this case, a lower NFE is preferred. The mathematical expression is represented by EquationEquation (4) as follows: (4) PEI=[HIω1POSsω2]/NFEω3(4)

The PEI is a versatile metric that can be tailored by researchers to align with their preferred evaluation criteria through the modification of equations. The addition of supplementary metrics can be incorporated into the equations based on whether higher or lower values are desired. As previously mentioned, metrics with higher desired outcomes are included in the numerator, while those with lower desired outcomes are placed in the denominator. By consolidating multiple performance indices into a single metric, this method streamlines the evaluation process and offers valuable insights for decision-making in complex optimisation scenarios.

6. Experiments

The selected test problem for investigation is the RDSP problem with an SO approach. As previously explained, the Enhanced Discrete Bees Algorithm (EDBA) proposed by Liu, Zhou, Pham, Xu, Ji, et al. (2018a) serves as the BBA employed to address the challenges of the robotic disassembly problem. The objective in this RDSP problem is to minimise the disassembly time, as expressed in EquationEquation (5). (5) Z=i=0Np1tb(xi)+i=0Np2tz(xi,xi+1)+i=0Np2tt(xi,xi+1)+i=0Np2mt(xi,xi+1)(5) where

  • Z is the total disassembly time.

  • Np is the number of total parts.

  • tb(xi) is the basic time for disassembling part xi.

  • tz(xi, xi+1) is the penalty time for disassembly direction changes between part xi and xi+1.

  • tt(xi, xi+1) is the penalty time for disassembly tool changes between part xi and xi + 1.

  • mt(xi, xi+1) is the moving time between part xi and xi+1.

6.1. Experimental setup and metrics

There are two commonly employed stopping criteria in optimisation algorithms: iterations or NFE. In this research, the stopping criteria for the algorithm are based on the number of iterations. The parameter settings used for EDBA are based on the recommended settings from Liu, Zhou, Pham, Xu, Ji, et al. (2018a). In their work, the authors applied population sizes of 10, 20, 30, 40 and 50 for Gear Pump A for 100, 200, 300 and 400 iterations. Similarly, for Gear Pump B, the same population numbers were employed between iterations 100 and 600. The values for the numbers of selected sites (m), elite sites (e), recruited bees for other selected sites (nsp) and recruited bees for elite sites (nep) were assigned as follows: 4, 1, 1 and 2, respectively. According to their findings, the best results were obtained for Gear Pump A in 300 iterations using a population of 20 and for Gear Pump B in 500 iterations using a population of 40 (Liu, Zhou, Pham, Xu, Ji, et al., 2018a). In this research, to ensure a fair comparison, the iteration and population sizes for all algorithms have been set to the same values.

To ensure a meaningful comparison with the best results obtained from EDBA, it was decided to use 100, 200, 300, 400 and 500 iterations for both gear pumps, along with population sizes of 21, 31, 41 and 51. It is important to note that in the context of the EDBA article, the term “population” refers to the number of scout bees (n). Therefore, when calculating the total population using the EDBA parameter settings and applying EquationEquation (6), the corresponding values for EDBA are 21, 31, 41 and 51 for the population sizes. (6) BApop=(e * nep)+(( e)* nsp)+( m)(6)

6.2. Experimental parameter setting

In this section, the determination of the experimental parameters employed in the study is presented. The objective was to find the optimal parameter settings for BAF and assess its performance in comparison to EDBA. The steps taken to determine the optimal parameter settings for BAF are as follows:

  1. Experimental Design: Define the experimental parameters. The numbers of selected sites (m), the maximum number of recruited bees around the selected sites (maxnr) and the maximum number of re-visits (max_rv) are configured as depicted in .

  2. Initial Runs: Perform 10 runs for each scenario, E1 to E100.

  3. Top Results Identification: Identify the best results from the initial runs.

  4. Iterations and Performance Validation: Evaluate the performance of the selected best results from the initial runs across 50 runs using 100 iterations.

  5. Descriptive Statistics: Analyse the descriptive statistics of the results.

  6. Best Result Selection: Identify and select the best result.

Table 3. Experimental design for BAF.

Since BAF uses a ranking-based mechanism to give the number of bees sent to the selected sites, for comparison purposes, it is necessary to calculate the population, as shown in EquationEquation (7), to obtain comparable population sizes for BAF and EDBA. presents values calculated using m=5 and various maxnr with a target maximum population of 51 bees. The maximum value of nr is nr1. (7) BAFpopulation=(nm)+i=1m(nri)(7)

Table 4. Example of BAF for a population of 51.

The experiment was then conducted 10 times on Gear Pump B for a maximum of 100 iterations each. Gear Pump B was selected as it was more complex than Gear Pump A. The parameter settings and average results of BAF are presented in . Using the best parameter settings from Liu, Zhou, Pham, Xu, Ji, et al. (2018a), EDBA achieved a disassembly time of 150.113 with 5550 function evaluations.

Table 5. BAF results of the initial runs (10 independent runs).

The preliminary findings indicate that all scenarios require fewer NFEs compared to the benchmark algorithm, EDBA. Furthermore, the majority of results across all scenarios demonstrate lower average disassembly times compared to EDBA. These results suggest that BAF offers higher accuracy and efficiency. However, further experiments were necessary before definitive conclusions could be drawn. Based on the average results and NFE, experiments E48, 49, E50, E68, E69 and E70 were selected for further investigations through 50 independent runs. The statistical results obtained are shown in .

Table 6. Descriptive statistics of E48, E49, E50, E68, E69 and E70 (50 runs).

shows that scenario E69 yielded a lower standard error of the mean (SEM). This lower SEM signifies a more accurate estimate of the population mean. Moreover, the standard deviation is also smaller, indicating reduced data variability. This characteristic is desirable regardless of whether the objective is to minimise or maximise a certain parameter. The preference is for results to cluster closely around the mean, as it signifies greater precision in achieving the intended outcome. Therefore, scenario E69 with m=4, maxnr=13, n=26, max_rv=10 emerges as the optimal choice. Fifty independent runs were subsequently conducted with the above parameter settings to compare BAF and EDBA. The best parameter settings for BAF specified a minimum population size of 31. Consequently, the comparative assessment with a population size 21 was eliminated.

6.3. Experimental results

Experiments were run 50 times with the maximum number of iterations set at 100, 200, 300, 400 and 500, while the population sizes used were 31, 41 and 51. For ease of reference, a grouping code was employed, where “100 31” indicates the experiment with a maximum of 100 iterations and a population size of 31. This coding scheme helps to conveniently identify the specific parameter settings used in each experiment. The best disassembly results using EDBA and BAF are presented in and for Gear Pump A and Gear Pump B, respectively. These tables display the sequence, direction, tool and total time of disassembly. The best result aligns with the best-known values for the case study, as reported in Liu, Zhou, Pham, Xu, Ji, et al. (2018a), who conducted experiments using 1000 runs and obtained values of 87.5731 and 135.3167 for Gear Pumps A and B, respectively. and display the average values of the disassembly time, SEM, average NFE (Ave_NFE) and delta. Delta is the percentage difference between the average objective value (disassembly time) to the best-known value. The boxplots for Gear Pumps A and B are presented in and , respectively.

Figure 2. EDBA and BAF results (Gear Pump A).

Figure 2. EDBA and BAF results (Gear Pump A).

Figure 3. EDBA and BAF results (Gear Pump B).

Figure 3. EDBA and BAF results (Gear Pump B).

Table 7. Gear Pump A best results (EDBA and BAF).

Table 8. Gear Pump B best result (EDBA and BAF).

Table 9. EDBA and BAF average results for Gear Pump A.

Table 10. EDBA and BAF results for Gear Pump B.

6.4. Statistical performance metric results

The statistical assumptions, as discussed in Section 5.1, were checked prior to conducting the subsequent statistical test. From the boxplots, it is evident that the data do not follow a normal distribution. The statistical tests for normality and homogeneity, as presented in Appendix A, confirm that the results deviate from both normality and homogeneity assumptions. Therefore, as the data violate the statistical assumptions required for conducting parametric tests, nonparametric tests are employed. The significance of differences between the mean ranks of various iterations and population sizes on EDBA and BAF results is examined using the Kruskal–Wallis ANOVA test, as depicted in and . The test results indicate the rejection of the null hypothesis (p< 0.05), providing strong evidence of a difference between at least one pair of iterations and population sizes. Consequently, a post hoc test, namely, the Dunn–Sidak test, is conducted. The results of the Dunn–Sidak tests are presented in and .

Figure 4. Dunn–Sidak test results (Gear Pump A).

Group 1 = 100_31 EDBA, Group 2 = 100_31 BAF, Group 3 = 200_31 EDBA, Group 4 = 200_31 BAF, …, Group 29 = 500_51 EDBA, Group 30 = 500_51 BAF.

Figure 4. Dunn–Sidak test results (Gear Pump A).Group 1 = 100_31 EDBA, Group 2 = 100_31 BAF, Group 3 = 200_31 EDBA, Group 4 = 200_31 BAF, …, Group 29 = 500_51 EDBA, Group 30 = 500_51 BAF.

Figure 5. Dunn–Sidak test results (Gear Pump B).

Group 1 = 100_31 EDBA, Group 2 = 100_31 BAF, Group 3 = 200_31 EDBA, Group 4 = 200_31 BAF, …, Group 29 = 500_51 EDBA, Group 30 = 500_51 BAF.

Figure 5. Dunn–Sidak test results (Gear Pump B).Group 1 = 100_31 EDBA, Group 2 = 100_31 BAF, Group 3 = 200_31 EDBA, Group 4 = 200_31 BAF, …, Group 29 = 500_51 EDBA, Group 30 = 500_51 BAF.

Table 11. Kruskal–Wallis test results (Gear Pump A).

Table 12. Kruskal–Wallis test results (Gear Pump B).

The Dunn–Sidak results offer detailed insights into the groups that display variation in performance based on iteration within each group size. Specifically, for Gear Pump A, the results show significant differences of groups 1 and 11 (100_31 EDBA and 100_41 EDBA) to the other groups. It is noteworthy that BAF yielded consistent outcomes across all parameters, as evidenced by the groups with even-numbered labels. However, in the case of Gear Pump B, which has a larger dataset in comparison to Gear Pump A, both EDBA and BAF displayed more pronounced disparities across all groups.

These findings emphasise the importance of selecting optimal parameter settings for disassembly time optimisation, taking into account both iteration size and population size. The observed variability in performance highlights the need for careful parameter selection to achieve optimal results.

In the subsequent step, groups with significant differences from the previous statistical test are eliminated, and all iterations and population sizes are combined to focus on the groups that demonstrate similar performance with the lowest mean objective values. This step ensures that the final test results only include the groups that yield the best results. The outcomes of the final statistical tests can be found in and and .

Figure 6. EDBA and BAF boxplot final results.

Figure 6. EDBA and BAF boxplot final results.

Table 13. Kruskal–Wallis final results (Gear Pump A).

Table 14. Kruskal–Wallis final results (Gear Pump B).

6.5. PEI results

As previously highlighted, PEI is designed to serve as a versatile index. In this experiment, the objective is to minimise the disassembly time while also aiming for the lowest value of the NFE. To achieve this, both metrics are placed in the denominator during the PEI calculation, as shown in EquationEquation (8), with equal weight assigned to all the metrics. It is important to note that since none of the metrics are desirable to have the highest value, a value of 1 is assigned to the numerator. (8) PEI=1/[Objω1NFEω2](8)

To ensure a meaningful comparison, only the groups that exhibited no significant differences in means were considered for the calculation. and and Appendix B show the PEI calculation results, scaled to units of 10−6. The figures provide a visual representation of the PEI values for the selected groups, facilitating comparison of their performance. Note that, although this work employs both SPM and PEI, they can be used on their own as they are independent performance measures. SPM is more suitable for a single-objective (SO) problem, while PEI can be utilised for both SO and multi-objective (MO) optimisation.

Figure 7. PEI (histogram) and Average Disassembly Time (dot): Higher PEI and Lower Time are Better. Higher PEI represents highest performance and Lower time represents best disassembly time (desirable).

Note: The line shows the highest PEI; EDBA is represented by blue and BAF by purple colour

Figure 7. PEI (histogram) and Average Disassembly Time (dot): Higher PEI and Lower Time are Better. Higher PEI represents highest performance and Lower time represents best disassembly time (desirable).Note: The line shows the highest PEI; EDBA is represented by blue and BAF by purple colour

7. Discussion

The experimentation process for determining the optimal parameters setting for BAF was conducted systematically in six steps, including statistical tests for result analysis. This approach was implemented to ensure that the best parameters are obtained for the subsequent comparison with EDBA. The best parameters for the RDSP problem were found to be m=4, maxnr=13, max_rv=10. The study conducted by Liu, Zhou, Pham, Xu, Ji, et al. (2018a) used a population range of 21–51. The number of scout bees (n) in BAF was chosen in the range 6–26 to yield a similar population size of 31–51. As previously mentioned, the population size 21 was eliminated from the final experiments. After determining the optimal parameters settings, 50 independent runs were conducted for both EDBA and BAF.

The minimum disassembly time can be achieved using both EDBA and BAF, as presented in and for both gear pumps. These tables give the sequence, direction, tools and total time required to disassemble the gear pumps. However, the average results reveal differences between EDBA and BAF. The average results of EDBA and BAF, presented in and , indicate that BAF outperforms EDBA. Examination of the outcomes for Gear Pump A shows that BAF consistently exhibits superior performance in comparison to EDBA across all 15 parameter configurations. BAF achieves a perfect accuracy rate in 12 instances, as evidenced by a delta value of 0%, indicating that the accuracy is 100%. In contrast, EDBA only achieves 100% accuracy once with the same settings. The SEM and delta exhibit a decreasing trend as larger parameter values are used. EDBA performed best in 500 iterations and with a population size of 41, giving an average NFE of 22,540. In contrast, BAF demonstrated the best performance in almost all cases. The average NFE was 5526 that was the lowest NFE required to achieve optimal outcomes in 100 iterations and with a population size of 51. Similarly, on Gear Pump B, it is evident that BAF outperforms EDBA for all population sizes and iterations. displays the boxplot results for Gear Pump A, indicating that BAF achieves better results. The algorithm demonstrates the ability to find the minimum disassembly time with minimal data spread, particularly at higher iterations for all tested population sizes. In , the boxplot results for Gear Pump B exhibit improved data visualisation, allowing for better observation of data spread. In this boxplot, it becomes even clearer that BAF outperforms EDBA in terms of mean, minimum disassembly time and data spread across all tested iterations and population sizes.

As mentioned earlier, research by Liu, Zhou, Pham, Xu, Ji, et al. (2018a) indicates that EDBA performs best on Gear Pump A in 300 iterations and with a population of 21 (NFE = 7520), while Gear Pump B achieves optimal performance in 500 iterations with a population of 41 (NFE = 22,540). However, it is important to emphasise that the conclusions drawn in this research are not solely based on statistically descriptive results. To ensure the reliability and validity of the findings, additional statistical tests are imperative to validate the parameter settings that yield the best performance. Relying solely on the statistically descriptive results may not provide the conclusive evidence required to draw robust conclusions. Since the experiments involved 50 independent runs, it is necessary to assess whether parametric statistical tests can be employed, as discussed in Section 5.1. The statistical assumptions for parametric tests include having a sample size of more than 30, data following a normal distribution, and data belonging to the same populations. However, the boxplot results indicate that the data do not follow a normal distribution, and thus, normality and homogeneity tests were conducted. The results confirm that the data violate the assumptions of normality and homogeneity, necessitating the use of nonparametric tests despite the sample size exceeding 30. As previously mentioned, the Kruskal–Wallis test indicated a significant difference among the iterations within the same population size groups (see and ). To further investigate these differences, the Dunn–Sidak post hoc test was employed to identify specific groups that exhibited statistically significant variations in mean values.

and clearly demonstrate noticeable differences, particularly in the outcome obtained from EDBA. Subsequently, the groups with significant differences are removed, and the statistical tests are repeated. The boxplot results, as shown in , visually indicate that the data distributions in each group are similar. The Kruskal–Wallis test (see and ) confirmed that there was no statistically significant difference among the groups (p> 0.05). Therefore, post hoc tests are not needed, as the Kruskal–Wallis test has already confirmed the similarity among the groups.

The final step involves the utilisation of the proposed index, as detailed in Section 5.2. In this study, the index is calculated based on the disassembly time and NFE. The calculation for the index has been previously explained in the previous section, and the obtained results, depicted in , exhibit the PEI values in conjunction with the average disassembly time. By examining , the parameter settings that yield the highest PEI can be identified, while the average disassembly time has been graphically presented for analytical purposes. As mentioned previously, the PEI is designed to be interpreted such that a higher index indicates better performance. The PEI results reveal that the best PEI of 9.87 × 10−6 for Gear Pumps A was obtained in 100 iterations and with a population size of 41 for BAF while for EDBA the optimum PEI of 4.23 × 10−6 was found in 300 iterations and with a population size of 31. The results indicate that BAF with a smaller NFE can produce statistically similar outcomes to those with a higher NFE. displays the average disassembly time as a dot, indicating that BAF consistently outperforms EDBA in all instances. Directly comparing similar parameter settings for the two algorithms reveals that BAF consistently yields higher PEI values – a highly desirable outcome for both gear pumps. In terms of average disassembly time, BAF yields lower values, aligning with the desired objective of minimising the disassembly time. The results obtained from the PEI analysis of Gear Pump B () clearly indicate that BAF outperforms EDBA. The optimal results were achieved after 400 iterations and with a population size of 31, yielding a PEI value of 2.04 × 10−6.

These findings highlight the robustness of the chosen parameter settings, which cannot be easily discerned based on statistical description alone. It is evident from the results that the average disassembly time exhibits a pattern of decreasing values with higher population size and iteration. However, when considering the average disassembly time and the PEI together, it becomes apparent that the parameter settings yielding the lowest average disassembly time may not always be the optimal ones. This is because the NFE also plays a crucial role, as a higher NFE indicates that the algorithm requires more evaluations to achieve the best results. Therefore, a balance between the average disassembly time and the NFE needs to be considered to determine the best parameter settings.

By utilising SPM and PEI, it becomes possible to identify the optimal parameter settings and evaluate the performance of the algorithms. From a statistical standpoint, it is evident that BAF outperforms EDBA with BAF demonstrating a faster convergence (smaller NFE). This observation is further supported by the average results presented in and , where it can be seen that BAF consistently achieves lower disassembly times across various iterations and population sizes. BAF consistently outperforms EDBA for both gear pumps, particularly in larger datasets, such as Gear Pump B, demonstrating that it is capable of handling more complex problems. Moreover, the statistical tests and PEI calculations consistently indicate that BAF outperforms EDBA. These findings provide valuable insights into the comparative performance of the algorithms and highlight the advantages of employing BAF in terms of achieving lower disassembly times and higher PEI values.

Furthermore, the statistical analysis has revealed interesting findings, particularly evident in the statistical results presented in and the normality test detailed in , focussing on the case study of Gear Pump A. It is worth noting that the best-known solution for Gear Pump A stands at 89.5731. As illustrated in , BAF demonstrates remarkable accuracy across various combinations of iterations and population sizes, consistently yielding a mean value of 89.5731, indicating a 100% accuracy rate. Upon closer examination in , it becomes evident that the data cannot conform to a normal distribution due to its precise alignment with the best-known solution. This non-normality aligns with our analysis, which suggests that in optimisation problems focussed on minimising or maximising objectives, deviation from the normal distribution is to be expected. In this context, it is crucial to emphasise that in problems where the objectives are centred on minimisation or maximisation, truly accurate results align precisely with a single point, inherently deviating from a normal distribution. Similarly, when results are near-accurate, they tend to cluster closely around that optimal point, leading to a skewed distribution pattern. As a result, we recommend that future researchers incorporate normality tests and examine normal graphs. It is anticipated that such analyses will confirm the departure from a normal distribution, particularly when results are highly accurate. In such instances, nonparametric tests should be preferred over parametric tests due to the observed accuracy of the results.

8. Conclusion

The enhancement of the BA was achieved by reducing the number of parameters to 4 and by simplifying the algorithm steps, removing elite sites and incorporating the Fibonacci sequence to guide the recruitment of bees for exploring the local search space. The SPM and PEI analyses conducted in this study indicate that BAF exhibits better performance than the benchmark EDBA in this RDSP problem. EDBA has been compared against other algorithms, including EDBA variants (EDBA without mutation operator (EDBA-WMO)), GA with Precedence Preserving Crossover (GA-PPX), and Self-Adaptive Simplified Swarm Optimisation (SASSO), and has demonstrated better performance. Therefore, the fact that BAF outperforms EDBA in this study means that BAF also outperforms these other algorithms. This further strengthens the evidence supporting the effectiveness of BAF as a robust and efficient algorithm in the context of RDSP. This conclusion is supported by several key factors, including the lower NFE values and the ability to achieve a minimum disassembly time. BAF consistently outperforms EDBA on both gear pumps. These findings highlight the effectiveness of BAF as an optimal choice for achieving efficient and effective RDSP. The SPM and PEI are both valuable tools in the field of algorithm analysis. These tools serve as effective means to identify optimal parameter settings and evaluate the performance of algorithms. SPM allows for statistical comparison and analysis of algorithm performance based on specific metrics, providing insights into the differences among various parameter settings. The PEI is a versatile metric that combines multiple metrics, such as disassembly time and NFE, into a single index to provide a comprehensive evaluation of algorithm performance. By combining SPM and PEI, researchers and practitioners can make informed decisions regarding parameter settings and algorithm selection to optimise their outcomes in various domains. Moreover, the validation of the proposed approach through a case study provides evidence of the algorithm’s robustness and effectiveness in successfully solving the robotic disassembly planning problem. Although this research specifically focusses on addressing the challenges of the RDSP problem, it is important to consider the wider applicability of the proposed algorithm in future research. Future studies will explore and compare the proposed enhancements with the BA2 approach in a broader context, encompassing both continuous and combinatorial problems. The parameters used in this study are smaller compared to those in TSP and VRP problems, which typically employ population sizes of 200–400 or even larger. This suggests the need for further research on the potential applications of BAF in other contexts. This would enable a comprehensive evaluation and understanding of the algorithm’s performance across different problem domains. Furthermore, investigating analogies with other features of honeybees, such as their learning process, olfactory system and hive organisation, could lead to more improvements to the BA. The algorithm can also be utilised for parameter tuning in machine learning including deep learning, thereby increasing its usefulness as an advanced artificial intelligence tool. This extension and exploration of the BA in different contexts will contribute to its versatility and applicability in various optimisation domains. In addition, the findings shows that BAF demonstrated exceptional accuracy in optimising Gear Pump A, leading to a non-normal distribution due to precise results. Therefore, for such an accurate and near-accurate result, the null hypothesis that the distribution is normal is expected to be rejected, and non-parametric tests should be conducted. Notably, none of the previous research in this area has observed this phenomenon, underscores the contribution to existing knowledge that results from this research. The MATLAB code of BAF for RDSP is available in github. The github address is given in Appendix C.

Abbreviations
AHGA=

Automated Hybrid Genetic Algorithm

AVNS=

Adaptive Variable Neighbourhood Search

BA=

Bees Algorithm

BAF=

Fibonacci Bees Algorithm

BA2=

Bees Algorithm with 2 parameters

BBA=

Basic Bees Algorithm

BLSA=

Brand Local Search Algorithm

CDDO=

Child Drawing Development Optimisation

DoE=

Design of Experiments

EDBA=

Enhanced Discrete Bees Algorithm

ELS=

Evolutionary Local Search

EoL=

End-of-Life

FIA=

Fibonacci Indicator Algorithm

FSOA=

Fibonacci Sequence-based Optimisation Algorithm

FSQGA=

Fibonacci Sequence-based Quantum GA

FTO=

Fibonacci Tree Optimisation

GA=

Genetic Algorithm

GA-PPX=

GA with Precedence Preserving Crossover

GROM=

Golden Ratio Optimisation Method

GWO=

Grey Wolf Optimiser

HI=

Hypervolume Indicator

IWPA=

Improved Wolf Pack Algorithm

MA=

Memetic Algorithm

MCDM=

Multiple-Criteria Decision-Making

MO=

Multi-Objective

MOFA=

Multi-Objective Fibonacci-Based Algorithm

NFE=

Number of Function Evaluations

NP=

Non-Deterministic Polynomial

Opp-PSOGWO=

Opposition-based Learning PSO GWO

PEI=

Performance Evaluation Index

POSs=

Pareto Optimal Solutions

PSO=

Particle Swarm Optimisation

RDSP=

Robotic Disassembly Sequence Planning

SASSO=

Self-Adaptive Simplified Swarm Optimisation

SEM=

Standard Error of the Mean

SSA=

Salp Swarm Algorithm

SBA=

Standard Bees Algorithm

SO=

Single-Objective

SPM=

Statistical Performance Metric

TSP=

Travelling Salesman Problem

VRP=

Vehicle Routing Problem

WPA=

Wolf Pack Algorithm

Acknowledgements

The authors would like to extend their sincere gratitude to Dr Jiayi Liu for generously providing the EDBA code for our research. The author would also like to thank the editor and the reviewers for their constructive feedback, which helped to improve the manuscript.

Disclosure statement

The authors declare no competing interests.

Additional information

Funding

This work was partly funded by the Engineering and Physical Sciences Research Council (EPSRC), UK, grant no. EP/N018524/1. Natalia Hartono’s doctoral study is supported by the Indonesian Endowment Fund for Education (LPDP), Ministry of Finance, Republic of Indonesia, grant no. 201908222215155 under BUDI LN scheme.

Notes on contributors

Natalia Hartono

Natalia Hartono earned her doctoral degree in Mechanical Engineering from the University of Birmingham, UK, in 2023, specialising in intelligent optimisation, the bees algorithm, sustainability, robotic disassembly, and remanufacturing. She has made contributions to these fields, with publications in reputable journals, research poster and thesis presentation competitions, and presentations at international conferences. She is part of the Bees Algorithm Research Group and Autonomous Remanufacturing Group at the University of Birmingham. Dr Hartono is well-known as the co-chair of the International Workshop Series on the Bees Algorithm and Its Applications and the co-editor of the book “Intelligent Production and Manufacturing Optimisation – The Bees Algorithm Approach” published by Springer.

References

  • Abdulhameed, S., & Rashid, T. A. (2022). Child Drawing Development Optimization Algorithm based on child’s cognitive development. Arabian Journal for Science and Engineering, 47(2), 1337–1351. https://doi.org/10.1007/s13369-021-05928-6
  • Aras, N., & Aksen, D. (2008). Locating collection centers for distance-and incentive-dependent returns. International Journal of Production Economics, 111(2), 316–333. https://doi.org/10.1016/j.ijpe.2007.01.015
  • Belhadj, I., Aicha, M., & Aifaoui, N. (2022). Product disassembly planning and task allocation based on human and robot collaboration. International Journal on Interactive Design and Manufacturing (IJIDeM), 16(2), 803–819. https://doi.org/10.1007/s12008-022-00908-y
  • Castellani, M., & Pham, D. T. (2023). The Bees Algorithm—A gentle introduction. In D. T. Pham & Natalia Hartono (Eds.), Intelligent production and manufacturing optimisation—The Bees Algorithm approach (pp. 3–21). Springer International Publishing.
  • Caterino, M., Fera, M., Macchiaroli, R., & Pham, D. (2023). Task optimisation for a modern cloud remanufacturing system using the Bees Algorithm. In D. T. Pham & Natalia Hartono (Eds.), Intelligent production and manufacturing optimisation—The Bees Algorithm approach (pp. 365–382). Springer.
  • Caterino, M., Fera, M., Macchiaroli, R., & Pham, D. T. (2022). Cloud remanufacturing: Remanufacturing enhanced through cloud technologies. Journal of Manufacturing Systems, 64, 133–148. https://doi.org/10.1016/j.jmsy.2022.06.003
  • Cevikcan, E., Aslan, D., & Yeni, F. B. (2020). Disassembly line design with multi-manned workstations: A novel heuristic optimisation approach. International Journal of Production Research, 58(3), 649–670. https://doi.org/10.1080/00207543.2019.1587190
  • Chen, B., Xu, W., Liu, J., Ji, Z., & Zhou, Z. (2020). Robotic disassembly sequence planning considering robotic collision avoidance trajectory in remanufacturing [Paper presentation]. 2020 IEEE 18th International Conference on Industrial Informatics (INDIN) (Vol. 1, pp. 494–501). Warwick, UK. https://doi.org/10.1109/INDIN45582.2020.9442129
  • Chen, Q., Yao, B., & Pham, D. T. (2020). Sequence-dependent robotic disassembly line balancing problem considering disassembly path [Paper presentation]. International Manufacturing Science and Engineering Conference (Vol. 84263, p. V002T07A019). Virtual, Online. https://doi.org/10.1115/MSEC2020-8268
  • Costa, C. M., Veiga, G., Sousa, A., Rocha, L., Oliveira, E., Cardoso, H. L., & Thomas, U. (2018). Automatic generation of disassembly sequences and exploded views from solidworks symbolic geometric relationships. 2018 IEEE International Conference on Autonomous Robot Systems and Competitions (ICARSC) (pp. 211–218). Torres Vedras, Portugal.
  • Cui, J., Yang, C., Zhang, J., Tian, S., Liu, J., & Xu, W. (2023). Robotic disassembly sequence planning considering parts failure features. iET Collaborative Intelligent Manufacturing, 5(1), e12074. https://doi.org/10.1049/cim2.12074
  • Da, C., Wenchao, Z., Siqing, S., Yiqun, L., Yazhou, L., & Xianbo, K. (2018). A parameter identification algorithm for turbine-governor system in regional power grid [Paper presentation]. 2018 2nd IEEE Conference on Energy Internet and Energy System Integration (EI2) (pp. 1–5). Beijing, China. https://doi.org/10.1109/EI2.2018.8582270
  • De la Torre Gutiérrez, H., & Pham, D. T. (2018). Identification of patterns in control charts for processes with statistically correlated noise. International Journal of Production Research, 56(4), 1504–1520. https://doi.org/10.1080/00207543.2017.1360530
  • Etminaniesfahani, A., Ghanbarzadeh, A., & Marashi, Z. (2018). Fibonacci Indicator Algorithm: A novel tool for complex optimization problems. Engineering Applications of Artificial Intelligence, 74, 1–9. https://doi.org/10.1016/j.engappai.2018.04.012
  • Frizziero, L., Donnici, G., Santi, G. M., Leon-Cardenas, C., Ferretti, P., Pascucci, G., & Liverani, A. (2022). Disassembly sequence planning validated through augmented reality for a speed reducer. Cogent Engineering, 9(1), 1–31. https://doi.org/10.1080/23311916.2022.2061321
  • Gao, K.-Z., He, Z., Huang, Y., Duan, P.-Y., & Suganthan, P. N. (2020). A survey on meta-heuristics for solving disassembly line balancing, planning and scheduling problems in remanufacturing. Swarm and Evolutionary Computation, 57, 100719. https://doi.org/10.1016/j.swevo.2020.100719
  • Grabcad Community. (2020). Gear pump 10 l/min. Retrieved July 22, 2020, from https://grabcad.com/library/gear-pump-10l-min-1
  • Gudla, P. K., & Ganguli, R. (2005). An automated hybrid genetic-conjugate gradient algorithm for multimodal optimization problems. Applied Mathematics and Computation, 167(2), 1457–1474. https://doi.org/10.1016/j.amc.2004.08.026
  • Guo, J., Pu, Z., Du, B., & Li, Y. (2022). Multi-objective optimisation of stochastic hybrid production line balancing including assembly and disassembly tasks. International Journal of Production Research, 60(9), 2884–2900. https://doi.org/10.1080/00207543.2021.1905902
  • Hartono, N. (2023). Intelligent robotic disassembly optimisation for sustainability using the Bees Algorithm University of Birmingham. http://etheses.bham.ac.uk/id/eprint/14384
  • Hartono, N., Ismail, A. H., Zeybek, S., Caterino, M., Jiang, K., & Sahin, M. (2023). Parameter tuning for Combinatorial Bees Algorithm in travelling salesman problems [Paper presentation]. AIP Conference Proceedings, Vol. 2485, no. 1. Bandung, Indonesia.
  • Hartono, N., Javier Ramírez, F., & Pham, D. T. (2023). Optimisation of robotic disassembly sequence plans for sustainability using the multi-objective Bees Algorithm. In Intelligent production and manufacturing optimisation—The Bees Algorithm approach (pp. 337–363). Springer International Publishing.
  • Hartono, N., Ramírez, F. J., & Pham, D. (2022a). Optimisation of robotic disassembly plans using the Bees Algorithm. Robotics and Computer-Integrated Manufacturing, 78, 102411. https://doi.org/10.1016/j.rcim.2022.102411
  • Hartono, N., Ramírez, F. J., & Pham, D. (2022b). A sustainability-based model for robotic disassembly sequence planning in remanufacturing using the Bees Algorithm. IFAC-PapersOnLine, 55(10), 1013–1018. https://doi.org/10.1016/j.ifacol.2022.09.509
  • Hosseinian, A. H., & Baradaran, V. (2020). P-GWO and MOFA: Two new algorithms for the MSRCPSP with the deterioration effect and financial constraints (case study of a gas treating company). Applied Intelligence, 50(7), 2151–2176. https://doi.org/10.1007/s10489-020-01663-x
  • Hussein, W. A., Sahran, S., & Sheikh Abdullah, S. N. H. (2017). The variants of the Bees Algorithm (BA): A survey. Artificial Intelligence Review, 47(1), 67–121. https://doi.org/10.1007/s10462-016-9476-8
  • Ismail, A. H. (2021). Enhancing the Bees Algorithm using the traplining metaphor [Unpublished doctoral dissertation]. University of Birmingham.
  • Ismail, A. H., Hartono, N., Zeybek, S., & Pham, D. T. (2020). Using the Bees Algorithm to solve combinatorial optimisation problems for TSPLIB. IOP Conference Series: Materials Science and Engineering, 847(1), 012027. https://doi.org/10.1088/1757-899X/847/1/012027
  • Ismail, A. H., & Pham, D. T. (2023). Bees traplining metaphors for the vehicle routing problem using a decomposition approach. In D. T. Pham & Natalia Hartono (Eds.), Intelligent production and manufacturing optimisation—The Bees Algorithm approach (pp. 261–287). Springer International Publishing.
  • Kerin, M., Hartono, N., & Pham, D. (2023). Optimising remanufacturing decision-making using the Bees Algorithm in product digital twins. Scientific Reports, 13(1), 701. https://doi.org/10.1038/s41598-023-27631-2
  • Khosla, T., & Verma, O. P. (2022). An efficient particle swarm optimization with fusion of figurate opposition-based learning and Grey Wolf Optimizer [Paper presentation]. 2022 International Conference on Computing, Communication, and Intelligent Systems (ICCCIS) (pp. 208–213). Greater Noida, India. https://doi.org/10.1109/ICCCIS56430.2022.10037666
  • Koshy, T. (2017). Fibonacci and Lucas numbers with applications, one. Wiley Online Library.
  • Laili, Y., Li, X., Wang, Y., Ren, L., & Wang, X. (2021). Robotic disassembly sequence planning with backup actions. IEEE Transactions on Automation Science and Engineering, 19(3), 2095–2107.
  • Laili, Y., Tao, F., Pham, D. T., Wang, Y., & Zhang, L. (2019). Robotic disassembly re planning using a two-pointer detection strategy and a super-fast Bees Algorithm. Robotics and Computer-Integrated Manufacturing, 59, 130–142. https://doi.org/10.1016/j.rcim.2019.04.003
  • Laili, Y., Wang, Y., Fang, Y., & Pham, D. T. (2022a). Evolutionary optimisation for robotic disassembly sequence planning and line balancing. In Optimisation of Robotic Disassembly for Remanufacturing (pp. 85–110). Springer Series in Advanced Manufacturing. Springer, Cham. https://doi.org/10.1007/978-3-030-81799-2_7
  • Laili, Y., Wang, Y., Fang, Y., & Pham, D. T. (2022b). Robotic disassembly sequence re-planning. In D. T. Pham (Ed.), Optimisation of robotic disassembly for remanufacturing (pp. 131–142). Springer International Publishing.
  • Laili, Y., Wang, Y., Fang, Y., & Pham, D. T. (2022c). Solutions for robotic disassembly sequence planning with backup actions. In Optimisation of Robotic Disassembly for Remanufacturing (pp. 111–130). Springer Series in Advanced Manufacturing. Springer, Cham. https://doi.org/10.1007/978-3-030-81799-2_8
  • Lambert, A. (1997). Optimal disassembly of complex products. International Journal of Production Research, 35(9), 2509–2524. https://doi.org/10.1080/002075497194633
  • Liu, J., Liu, Q., Zhou, Z., Pham, D. T., Xu, W., & Fang, Y. (2023). Collaborative optimisation of robotic disassembly planning problems using the Bees Algorithm. In D. T. Pham & Natalia Hartono (Eds.), Intelligent production and manufacturing optimisation—The Bees Algorithm approach (pp. 305–335). Springer International Publishing.
  • Liu, J., & Wang, S. (2017). Balancing disassembly line in product recovery to promote the coordinated development of economy and environment. Sustainability, 9(2), 309. https://doi.org/10.3390/su9020309
  • Liu, J., Xu, Z., Xiong, H., Lin, Q., Xu, W., & Zhou, Z. (2023). Digital twin-driven robotic disassembly sequence dynamic planning under uncertain missing condition. IEEE Transactions on Industrial Informatics, 19(12), 11846–11855. https://doi.org/10.1109/TII.2023.3253187
  • Liu, J., Zhou, Z., Pham, D. T., Xu, W., Ji, C., & Liu, Q. (2020). Collaborative optimization of robotic disassembly sequence planning and robotic disassembly line balancing problem using improved discrete Bees Algorithm in remanufacturing. Robotics and Computer-Integrated Manufacturing, 61, 101829. https://doi.org/10.1016/j.rcim.2019.101829
  • Liu, J., Zhou, Z., Pham, D. T., Xu, W., Ji, C., & Liu, Q. (2018a). Robotic disassembly sequence planning using enhanced discrete Bees Algorithm in remanufacturing. International Journal of Production Research, 56(9), 3134–3151. https://doi.org/10.1080/00207543.2017.1412527
  • Liu, J., Zhou, Z., Pham, D. T., Xu, W., Yan, J., Liu, A., Ji, C., & Liu, Q. (2018b). An improved multi-objective discrete Bees Algorithm for robotic disassembly line balancing problem in remanufacturing. The International Journal of Advanced Manufacturing Technology, 97(9–12), 3937–3962. https://doi.org/10.1007/s00170-018-2183-7
  • Ma, Y.-S., Jun, H.-B., Kim, H.-W., & Lee, D.-H. (2011). Disassembly process planning algorithms for end-of-life product recovery and environmentally conscious disposal. International Journal of Production Research, 49(23), 7007–7027. https://doi.org/10.1080/00207543.2010.495089
  • McGovern, S., & Gupta, S. M. (2011). The disassembly line: Balancing and modeling. McGraw-Hill.
  • Ming, H., Liu, Q., & Pham, D. T. (2019). Multi-robotic disassembly line balancing with uncertain processing time. Procedia CIRP, 83, 71–76.https://doi.org/10.1016/j.procir.2019.02.140
  • Nematollahi, A. F., Rahiminejad, A., & Vahidi, B. (2020). A novel meta-heuristic optimization method based on golden ratio in nature. Soft Computing, 24(2), 1117–1151. https://doi.org/10.1007/s00500-019-03949-w
  • Omotehinwa, T., & Ramon, S. (2013). Fibonacci numbers and golden ratio in mathematics and science. International Journal of Computer and Information Technology, 2(4), 630–638.
  • Pham, D. T., & Castellani, M. (2009). The Bees Algorithm: Modelling foraging behaviour to solve continuous optimization problems. Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science, 223(12), 2919–2938.
  • Pham, D. T., & Castellani, M. (2015). A comparative study of the Bees Algorithm as a tool for function optimisation. Cogent Engineering, 2(1), 1091540. https://doi.org/10.1080/23311916.2015.1091540
  • Pham, D. T., & Darwish, A. H. (2008). Fuzzy selection of local search sites in the Bees Algorithm. In Proceedings of the 4th International Virtual Conference on Intelligent Production Machines and Systems (IPROMS 2008) (pp. 1–14). Scotland: Whittles Publishing.
  • Pham, D. T., Ghanbarzadeh, A., Koç, E., Otri, S., Rahim, S., & Zaidi, M. (2005). The Bees Algorithm. Technical Note, Manufacturing Engineering Centre, Cardiff University, UK.
  • Pham, D. T., Ghanbarzadeh, A., Koç, E., Otri, S., Rahim, S., & Zaidi, M. (2006). The Bees Algorithm—A novel tool for complex optimisation problems. In D. T. Pham, E. E. Eldukhri & A. J.Soroka (Eds.) Intelligent production machines and systems (pp. 454–459). Elsevier.
  • Ramírez, F. J., Aledo, J. A., Gamez, J. A., & Pham, D. T. (2020). Economic modelling of robotic disassembly in end-of-life product recovery for remanufacturing. Computers & Industrial Engineering, 142, 106339. https://doi.org/10.1016/j.cie.2020.106339
  • Scott, T. C., & Marketos, P. (2014). On the origin of the Fibonacci sequence. MacTutor History of Mathematics, 23, 1–46.
  • Sha, L., & Pan, Z. (2018). FSQGA based 3D complexity wellbore trajectory optimization. Oil & Gas Sciences and Technology – Revue d’IFP Energies Nouvelles, 73, 79. https://doi.org/10.2516/ogst/2018008
  • Shackleton, K., Balfour, N. J., Toufailia, H. A., James, E., & Ratnieks, F. L. W. (2023). Honey bee waggle dance facilitate shorter foraging distances and increased foraging aggregation. Animal Behaviour, 198, 11–19. https://doi.org/10.1016/j.anbehav.2023.01.009
  • Tofallis, C. (2014). Add or multiply? A tutorial on ranking and choosing with multiple criteria. INFORMS Transactions on Education, 14(3), 109–119. https://doi.org/10.1287/ited.2013.0124
  • Tran-Ngoc, H., Le-Xuan, T., Khatir, S., De Roeck, G., Bui-Tien, T., & Abdel Wahab, M. (2023). A promising approach using Fibonacci sequence-based optimization algorithms and advanced computing. Scientific Reports, 13(1), 3405. https://doi.org/10.1038/s41598-023-28367-9
  • Visscher, P. K., & Seeley, T. D. (1982). Foraging strategy of honeybee colonies in a temperate deciduous forest. Ecology, 63(6), 1790–1801. https://doi.org/10.2307/1940121
  • Wei, L., Zhang, Z., & Lim, A. (2014). An adaptive variable neighborhood search for a heterogeneous fleet vehicle routing problem with three-dimensional loading constraints. IEEE Computational Intelligence Magazine, 9(4), 18–30. https://doi.org/10.1109/MCI.2014.2350933
  • Yang, C., Xu, W., Liu, J., Yao, B., & Hu, Y. (2022). Robotic disassembly sequence planning considering robotic movement state based on deep reinforcement learning [Paper presentation]. 2022 IEEE 25th International Conference on Computer Supported Cooperative Work in Design (CSCWD) (pp. 183–189). Hangzhou, China. https://doi.org/10.1109/CSCWD54268.2022.9776113
  • Ye, F., Perrett, J., Zhang, L., Laili, Y., & Wang, Y. (2022). A self-evolving system for robotic disassembly sequence planning under uncertain interference conditions. Robotics and Computer-Integrated Manufacturing, 78, 102392. https://doi.org/10.1016/j.rcim.2022.102392
  • Yeh, W.-C. (2006). An efficient memetic algorithm for the multi-stage supply chain network problem. The International Journal of Advanced Manufacturing Technology, 29(7–8), 803–813. https://doi.org/10.1007/s00170-005-2556-6
  • Yin, T., Zhang, Z., Zhang, Y., Wu, T., & Liang, W. (2022). Mixed-integer programming model and hybrid driving algorithm for multi-product partial disassembly line balancing problem with multi-robot workstations. Robotics and Computer-Integrated Manufacturing, 73, 102251.
  • Zachariadis, E. E., & Kiranoudis, C. T. (2010). An open vehicle routing problem metaheuristic for examining wide solution neighborhoods. Computers & Operations Research, 37(4), 712–723. https://doi.org/10.1016/j.cor.2009.06.021
  • Zeybek, S. (2023). Prediction of the remaining useful life of engines for remanufacturing using a semi-supervised deep learning model trained by the Bees Algorithm. In D. T. Pham & Natalia Hartono (Eds.), Intelligent production and manufacturing optimisation—The Bees Algorithm approach (pp. 383–397). Springer.
  • Zeybek, S., Ismail, A. H., Hartono, N., Caterino, M., & Jiang, K. (2021). An improved vantage point Bees Algorithm to solve combinatorial optimization problems from TSPLIB. Macromolecular Symposia, 396(1), 2000299. https://doi.org/10.1002/masy.202000299
  • Zeybek, S., & Koç, E. (2015). The vantage point Bees Algorithm. In 2015 7th International Joint Conference on Computational Intelligence (IJCCI) (Vol.1, pp. 340–345). Lisbon, Portugal.
  • Zhang, Z., Wei, L., & Lim, A. (2015). An evolutionary local search for the capacitated vehicle routing problem minimizing fuel consumption under three-dimensional loading constraints. Transportation Research Part B: Methodological, 82, 20–35. https://doi.org/10.1016/j.trb.2015.10.001
  • Zhao, Q., Tao, R., Li, J., & Mu, Y. (2020). An improved wolf pack algorithm [Paper presentation]. 2020 Chinese Control and Decision Conference (CCDC) (pp. 626–633). Hefei, China. https://doi.org/10.1109/CCDC49329.2020.9164486

Appendix A.

Statistical test results

 

Table A1. Gear Pump A normality test.

Table A2. Gear Pump B normality test.

Table A3. Gear Pumps A and B homogeneity test.

Appendix B.

Final PEI values

Table B1. PEI values – Gear Pump A.

Table B2. PEI values – Gear pump B.

Appendix C.

MATLAB Code

https://github.com/NataliaHartonoFung/Fibonacci-Bees-Algorithm-disassembly-sequence.