Next Article in Journal
Applications of Heat Exchanger in Solar Desalination: Current Issues and Future Challenges
Previous Article in Journal
Retrofitting of Existing Bar Racks with Electrodes for Fish Protection—An Experimental Study Assessing the Effectiveness for a Pilot Site
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization of Water Distribution Networks Using Genetic Algorithm Based SOP–WDN Program

1
Graduate School of Water Resources, Sungkyunkwan University, Suwon 16419, Korea
2
Smart Water Technology and Consulting, Co., Ltd., Incheon 21315, Korea
*
Author to whom correspondence should be addressed.
Water 2022, 14(6), 851; https://doi.org/10.3390/w14060851
Submission received: 28 January 2022 / Revised: 5 March 2022 / Accepted: 7 March 2022 / Published: 9 March 2022

Abstract

:
Water distribution networks are vital hydraulic infrastructures, essential for providing consumers with sufficient water of appropriate quality. The cost of construction, operation, and maintenance of such networks is extremely large. The problem of optimization of a water distribution network is governed by the type of water distribution network and the size of pipelines placed in the distribution network. This problem of optimal diameter allocation of pipes in a distribution network has been heavily researched over the past few decades. This study describes the development of an algorithm, ‘Smart Optimization Program for Water Distribution Networks’ (SOP–WDN), which applies genetic algorithm to the problem of the least-cost design of water distribution networks. SOP–WDN demonstrates the application of an evolutionary optimization technique, i.e., genetic algorithm, linked with a hydraulic simulation solver EPANET, for the optimal design of water distribution networks. The developed algorithm was applied to three benchmark water distribution network optimization problems and produced consistently good results. SOP–WDN can be utilized as a tool for guiding engineers during the design and rehabilitation of water distribution pipelines.

1. Introduction

A water distribution network (WDN) is comprised of various elements, such as reservoirs, pumps, pipes, tanks, and valves. Around 80% of the total cost of a water supply project is invested in its water distribution system [1]. Hence, the design of a cost-effective and reliable water distribution network is a must. Optimization of the WDN involves the design of a reliable, efficient, and cost-effective distribution network that fulfils the necessary water demands, while maintaining adequate pressure heads.
This is crucial for conservation of water resources, as well as for reducing energy requirements and maintenance costs. Optimization of WDNs can be categorized into many types, viz. design, operation, calibration, level-of-service, monitoring system and network testing. This paper deals with determining the optimal diameters of pipelines in a water distribution network with a predetermined layout.

Background and Related Work

Over the years, numerous researchers have presented many different methods for obtaining the optimal solution to the pipe network optimization problem. The Hardy cross method is considered as the oldest method for solving a pipe network. In this method, at any pipe junction, the algebraic sum of flow must be zero, and the algebraic sum of pressure drops at any loop must also be zero [2]. This method was improved upon by many other researchers. Alperovits and Shamir [3] proposed one of the most significant approaches for solving the problem of water distribution network design by utilizing the successive Linear Programming Gradient (LPG) method. This method was adopted and further expanded upon by other researchers [4,5].
However, deterministic methods, such as linear programming and non-linear programming, presented drawbacks, such as entrapment in local minima, and dependence on the starting point. Hence, they failed to obtain near optimal solutions for complex, multi-objective, real-world pipe network problems. It is crucial to escape local minima [6] and to overcome these drawbacks, researchers began to utilize meta-heuristic algorithms (genetic algorithms, simulated annealing, etc.) for water network design problems. These techniques include algorithms having some stochastic components. Goldberg and Kuo introduced stochastic methods for the optimization of water distribution networks using the principles of natural selection and genetics [7]. Simpson et al. used simple genetic algorithms (GA), and obtained a near optimal solution [8], while Simpson et al. [9] compared the GA technique with other methods, such as complete enumeration and non-linear optimization, and concluded that the GA technique generates multiple alternative solutions that are both practical and close to the optimum. The results obtained by Simpson et al. [8] were further improved upon by Dandy et al. [10] using the concept of variable power scaling of the fitness function, an adjacency mutation operator, and gray codes. Savic and Walters developed the computer model GANET [11] that utilizes GA for the least-cost design of pipe networks.
To avoid unfeasible solutions due to the violation of constraints, a penalty factor is necessary during the selection process of GA. Deb and Agrawal [12] developed a niched-penalty method to more effectively solve constrained optimization problems using GAs. Wu and Simpson [13] demonstrated significant improvements in efficiency and robustness for single-objective optimization utilizing a boundary search method. Liong and Atiquzzaman used the shuffled complex evolution (SCE) linked with EPANET hydraulic network solver [14] to obtain the least cost of some well-known water distribution networks in the literature. SCE was demonstrated to be a potential alternative to other optimization algorithms, due to its faster computational speed. Other algorithms, such the shuffled frog-leaping algorithm (SFLA) by Eusuff [15] and the harmony search Algorithm (HS) by Geem [16], have obtained comparable results, and have proven to be effective tools for the optimal design of water networks.
Some studies consider a single economic objective (least-cost) to formulate the network optimization and rehabilitation problem, whereas others consider a multi-objective optimization approach that compares interesting trade-offs (e.g., a slight pressure deficit can sometimes be outweighed by substantial cost reduction) [17]. To improve network reliability, Chandramouli and Malleswararao [18] used fuzzy logic based on the excess pressure available at demand nodes. Jin et al. analyzed additional objectives, such as considering both pressure and velocity violations [19]. Prasad and Park utilized genetic algorithms and considered both minimization of cost and maximization of network reliability [20]. More recent developments include improving algorithm convergence by using an engineered initial population, rather than a random one [21]; improvement of computational efficiency via the reduction of search space [22]; combining GA and mathematical programming with the inclusion of new elements such as pressure reducing valves [23]; using artificial neural networks (ANNs) rather than hydraulic and water quality simulation models together with differential evolution (DE) for optimization [24]; developing the Harris hawks optimization algorithm (HHO) for WDN optimization [25].
Additionally, Bilal and Pant utilized a hybrid metaheuristic algorithm (FA-PSO) [26] focusing on the Hanoi distribution network. Praneeth et al. demonstrated water cycle optimization algorithm [27], whereas Pankaj et al. utilized Cuckoo search [28] for least cost design of benchmark distribution networks. Surco et al. utilized a modified particle swarm optimization (PSO) algorithm for the optimization of distribution networks [29]. Cassiolato et al. proposed a deterministic mathematical programming approach without utilizing hydraulic simulators for cost minimization of looped WDNs, where generalized disjunctive programming is used to reformulate the discrete optimization problem to a mixed-integer nonlinear programming (MINLP) problem [30]. Pant and Snasel proposed a fuzzy C-means adaptive differential evolution (FCADE) for optimizing well-known benchmark WDN problems [31]. Bi et al. compared the searching behavior of evolutionary algorithms on water distribution system design optimization [32]. Zhao et al. proposed an in-sync optimization model for network layout and pipe diameter determination of a self-pressurized drip irrigation system [33]. Shao et al. utilized genetic algorithm for optimal placement of flow meters and valves in a distribution system [34].
Moreover, many researchers have considered operational optimization of WDNs, since minimization of operational energy costs during pumping must also be accounted for together with the construction costs. Electricity consumption during operation is one of the biggest marginal expenditures for water utilities [35]. Operational optimization can be achieved by controlling times when pumps operate, also called pump scheduling [36,37], flow rates [38], pump speeds [39] and tank-water trigger levels [40]. Furthermore, other factors such as real-time control and water quality can also be considered [41,42,43]. Metaheuristics such as GA have been abundantly utilized for operational optimization of WDNs, primarily with the objective of minimizing the overall cost of the pumping operation [44,45,46]. Research works in optimization of WDNs have been comprehensively reviewed by Mala et al. for design and rehabilitation [47] and system operation [35]. GA have also been abundantly applied for optimization in multiple fields of research [48,49,50]. Katoch et al. [51] have elaborated the advancements made in the field of GA.

2. Materials and Methods

2.1. Problem Formulation

Cost-effective WDN design is a discrete optimization problem, as the individual pipe sizes are to be selected from a list of available commercial size diameters. The search space can be determined as the number of available diameters, raised to the power of the number of pipes in the network [52]; e.g., if 8 different commercial pipe sizes are available for the design of a WDN having 10 pipelines, the search space size would be 810, i.e., 1,073,741,824 different pipe combinations. Hence, even for a relatively small pipe network, the search space is large. The design of an economically optimal water distribution network is a difficult task, because it involves solving many complex, non-linear, and discontinuous hydraulic equations, while simultaneously optimizing pipe sizes and other network components [53,54].
Optimization of a water distribution network aims to find the optimal pipe diameters in the network for the given layout and demand requirements. The optimal pipe sizes that satisfy all implicit constraints (conservations of mass and energy), and explicit constraints (hydraulic and design constraints) are selected in the final network.
The continuity equation is given as:
i = 1 n q i = 0
The continuity equation is applied to each node, with q i being the flow rate (flow into and flow out of the node), and n is the number of pipes connected at the node.
The energy equation is given as:
i = 1 m h i = 0
The energy equation is applied to each loop in the distribution network, where h i is the head loss in each pipe, and m is the number of pipes in the loop.
The objective function is the total cost of the given network. The total cost C T is calculated as:
C T = i = 1 N p C i D i   ·   L i
where, N p is the total number of pipes, C i D i is the cost per unit length of pipe i with diameter D i , and L i is the length of pipe i . The objective function is to be minimized under the implicit constraints and explicit constraints.
The head loss is the sum of the local head losses and the friction head losses. The equation used to calculate the head loss is the Hazen–Williams equation. This equation is an empirical equation that relates the flow of water in a pipe with the physical properties of the pipe and the energy loss due to friction. The Hazen–Williams coefficient, abbreviated as C , is a dimensionless number used in the Hazen–Williams equation [55]. The equation can be expressed as:
h f = 4.72 C 1.85 Q 1.85 D 4.87 L  
where, h f is the head loss, Q is the flow rate, C is the Hazen–Williams coefficient, D is the pipe inside diameter, and L is the pipe length.

2.2. Genetic Algorithm

A genetic algorithm (GA) is a search algorithm based on the mechanics of natural selection and natural genetics [56]. Although stochastic at certain aspects, a genetic algorithm is not entirely random, as it utilizes historical information to determine new search points. GAs have been widely utilized to solve optimization problems in multiple fields [57]. Following the concept of ‘survival of the fittest’, improvements in solutions evolve from past generations, until a near optimal solution is obtained. In genetic algorithms, the candidate solutions are represented by chromosomes (e.g., binary strings), and are collectively known as the population. The chromosomes are then evolved in each subsequent generation, according to their fitness. The fitness evaluation of each candidate solution depends upon how well it the meets the requirements of a pre-defined objective function (e.g., lowest cost). The more fit the candidate solution, the greater probability it will have of being selected for reproduction. Hence, the more fit chromosomes replace the less fit chromosomes, and the process continues until a near optimal solution is found.
The general idea of GA in a pipe network optimization problem is to select a population of initial solution points, scattered randomly in the optimization space, and then converge iteratively to better solutions, until the desired criteria for stopping are achieved. The steps for using GA for pipe network optimization can be briefly described as follows [8]:
1.
Generation of initial population
The GA randomly generates an initial population of coded strings (binary) representing pipe network solutions of population size N. Each of the N strings represents a possible combination of pipe sizes.
2.
Computation of network cost
For each N string in the population, the GA decodes each substring into the corresponding pipe size and computes the total network cost (material cost, construction cost, etc.).
3
Hydraulic analysis of each network
A steady state hydraulic network solver computes the heads and discharges under the specified demand patterns for each of the network designs in the population. The actual nodal pressures are compared with the minimum allowable pressure heads, and any pressure deficits are noted. Similarly, the actual water velocities at pipes are compared with the desired velocity of the water distribution network and any deviation in velocity are also noted.
4.
Computation of penalty cost
The GA assigns a penalty cost for each individual network design in the population if a pipe network does not satisfy the pressure and velocity constraints (for example, pressure violation at a particular node if the pressure in the node is less than or greater than the desired pressure).
5.
Computation of total network cost
The total cost of each network in the current population is then taken as the sum of the network cost (Step 2) and the penalty cost (Step 4).
6.
Computation of the fitness
The fitness of the coded string is taken as some function of the total network cost. For each proposed pipe network in the current population, fitness can be computed as the inverse or the negative value of the total network cost (Step 5).
7.
Generation of a new population using the selection operator
The GA generates new members for the next generation by a selection scheme that depends on the fitness of the initial members.
8.
The crossover operator
Crossover occurs with some specified probability for each pair of parent strings selected in Step 7. A uniform type of crossover operator is commonly used to accompany the comparatively large string size for pipe network optimization.
9.
The mutation operator
Mutation occurs with some specified probability of mutation for each bit in the strings that have undergone crossover. The purpose of the mutation operator is to maintain genetic diversity from one generation of a population to another.
10.
Production of successive generations
The use of the three operators described above produces a new generation of pipe network designs using Steps 2 to 9. The GA repeats the process to generate successive generations. The final costs and pipe network designs are stored, and the cheaper cost alternatives that meet the required constraints are updated.

2.3. EPANET

EPANET is a hydraulic simulator that can perform extended hydraulic and water quality simulations for pressurized water distribution networks (Rossman, [58]). Generally, a water distribution network consists of many elements, such as pipes (links), pipe junctions (nodes), pumps, control valves, and tanks/reservoirs. EPANET solves the water distribution network for the flow of water in each pipe, pressure at each junction, water height in each tank, concentration of chemical species, etc. During the hydraulic analysis of the water distribution network, EPANET solves both the conservation of mass and energy equations.
EPANET–MATLAB Toolkit is a software for interfacing a drinking water distribution system simulation library, EPANET, with the MATLAB technical computing language developed by Eliades [59]. The Toolkit allows users to access EPANET and EPANET-MSX through their shared object libraries, as well as their executables. EPANET can be called and used through a programming interface by an external software, which can be written in different programming languages (such as C/C++, Python, or MATLAB). Generally, a large number of commands have to be written to achieve specific results, such as extracting the node pressures, pipe diameters, pipe roughness coefficients, or specifying demand patterns. However, in the EPANET–MATLAB toolkit, a significant part of the repetitive code is already included in the toolkit functions and can be used directly.

2.4. SOP–WDN

Smart optimization program for water distribution networks (SOP–WDN) is an algorithm that has been developed by Smart Water Grid (SWG) research works for water distribution network optimization by using a genetic algorithm. The algorithm is written in MATLAB programming language. It interfaces with EPANET-MATLAB toolkit [59] (hydraulic solver integrated within SOP-WDN) for steady state hydraulic simulation and solution. The algorithm imports the network layout and data from EPANET. Design parameters, such as available pipe sizes, respective cost of pipes, roughness coefficient, and required pressure and velocities for the network, are to be added. GA optimization parameters, such as population size, crossover probability, and mutation rate, are prerequisites for the algorithm, and can be set by the user. Figure 1 shows a flowchart of the overall algorithm:

2.4.1. Encoding Scheme, Interpretation and Redundancy

A sequence of binary numbers is used to represent a water distribution network in SOP–WDN, as it enables a relatively extensive exploration of the search space. The coding scheme is also simple in implementation for the given task. Reflected binary code (RBC) or grey ode is utilized since it assures that two successive values differ by only one bit (binary digit). The number of bits required to represent an individual pipe in the network depends upon the number of available pipe diameters. The total binary sequence represents the entire pipe network. An example for the process of interpretation of a water distribution network by the algorithm is demonstrated.
Consider, a water distribution network having 1 reservoir, 1 tank, 5 demand nodes and 6 pipelines (all 100 m in length); there are 8 available pipe sizes for this network as shown in Table 1.
As there are 8 available pipe sizes in the example, the number of bits required to represent each individual pipe is 3 bits (starting from 000 to 100).
Any 18-character-long binary string can, hence, represent all 6 pipelines in this distribution network. Consider a randomly generated binary string of length 18 (exe. 101110011001000111). Since there are 6 pipes in the distribution system layout, the generated binary string of length 18 can be divided into six binary strings, each a length of 3 bits. Each binary string of 3 bits then represents a unique pipe size and the position of that binary string represents the position of the pipe in the overall network layout. The resulting pipe network and its respective cost can be obtained as shown in Table 2 and the network layout can be visualized as shown in Figure 2.
When a parameter belonging to a finite discrete set is encoded with binary notations, some of the codes may become redundant. For example, consider the design of a water distribution system where the number of commercially available diameters is 13. In this case, a minimum of 4 bits are required for coding the elements of the set of diameters, and this gives 24 = 16 discrete values. With 13 diameter options, three of the substrings become redundant (do not represent any diameters). The challenges of dealing with redundant binary codes have been overlooked in the literature and in published methods [60]. Saleh and Tanyimboh proposed representing redundant codes with closed pipes of fictitious diameters and low fitness values, assuming their extinction through evolution and natural selection [61]. However, this approach was found to prematurely lose valuable genetic information.
Alternatively, the redundant codes may be mapped to valid codes. In this case, some diameter options will be represented more than once. The mapping options may be fixed or random. In fixed mapping, the redundant codes are assigned to valid pipe diameter options prior to optimization. Random mapping assigns the redundant codes to the valid diameter options randomly. Tanyimboh [62] observed that, during mapping, over-representation of the largest pipe diameter performed better than over-representation of the smallest pipe diameter. It was also overserved that a balanced, unbiased allocation achieved good results. SOP-WDN handles redundant mapping by linearly scaling the probability of allocation of pipe diameters prior to optimization, where the largest pipe has 2 times the allocation chance than the smallest.

2.4.2. Genetic Algorithm Operators

Genetic algorithms differ from conventional optimization algorithms and search procedures as they work with a coding of the solution set and not the solutions themselves [56]. GA mainly consists of four basic operators: selection, crossover, mutation and elitism and utilizes these operators together to produce new generations.
1.
Selection
Selection is a crucial step in GA that determines whether a particular candidate will participate in the reproduction process or not. The selection operators give preference to the fitter chromosomes (candidate solutions), allowing them to pass on their ‘genes’ (information) to the next generation. Hence, worse chromosomes (with poor fitness) get eliminated. SOP-WDN algorithm computes fitness of an individual as the reciprocal of (the total cost times the imposed penalty). The selection operator then randomly picks chromosomes out of the population according to fitness. Roulette wheel, rank, and tournament selection are well-known techniques for selection and are commonly utilized in the literature.
In general, the probability of selection of an individual is calculated as the fitness values of the individual divided by fitness value of the population. Selection pressure is defined as the degree to which the better individuals are favored in the population; it drives the GA into improving the fitness over successive generations. In this study, power-law based probability selection was utilized, where P i is the preserved probability based on rank of the ith ranked individual (ranked based on fitness) in the population a having total of n individuals. The value of τ = 1.1 ~ 1.2 (controlling selection pressure) was found ideal after numerous simulations.
P i   = i τ i = 1 n i τ
2.
Crossover
Crossover is a genetic operator that combines two chromosomes (parents) in order to produce new chromosomes (children). Some common crossover techniques utilized in the literature are one-point crossover, two-point or k-point crossover, uniform crossover, shuffle crossover, three-parent crossover. In one point crossover, a crossover point is selected at random through the length of the chromosome. Summation of genes from the first parent, put before crossover point and second parent after the crossover point creates the new child. In two-point or k-point crossover, genetic material is exchanged between two or more random positions along the length of chromosome. Uniform crossover operates in individual genes of the selected chromosome, rather than on blocks.
Since GAs are problem specific, crossover is applied considering the chromosomes as group of distribution pipelines. Here, change of any single gene in a chromosome means changing and replacing the current pipe with another one of a different diameter. It is preferred to alter and test many different pipe combinations when searching for the optimal solution. After experimenting with varying crossover operators, k-point crossover was selected as the preferred crossover method. The number of crossover points is taken as (0.8 ×   N p ) rounded to the nearest integer, where N p is the total number of pipelines in the distribution system. The locations of crossover points are taken randomly. For example, the number of crossover points for a distribution network with 6 pipes is shown in Figure 3. The higher mixing ratio showed faster convergence. Uniform crossover also showed good results whereas the worst results were obtained from one-point crossover. The applied crossover probability was 85–90%.
3.
Mutation
After crossover, the chromosomes are subjected to mutation. Mutation prevents the algorithm from being trapped in a local optimum (premature convergence of the GA). It involves the modification of the value of each ‘gene’ of a chromosome with a small probability, called the mutation probability. Mutation plays the role of re-discovering lost or unexplored genetic materials, as well as for maintaining genetic diversity [56]. Mutation operators produce random changes in various chromosomes of the population.
Generally, a mutation probability of ( 1 / l ), where l is the chromosome length considered appropriate in the literature. In this study, mutation rate was kept around (4–6)%, which is considered higher, especially for networks with large number of pipelines, and a hill-climbing mutation operator was applied for the final 10% of the generations, which would permit mutation only if it improved the quality of the solution. The mutation strategy utilized in the study, shown in Figure 4, was bit-flip mutation since it complements the encoding scheme of the algorithm.
4.
Elitism
Elitism involves selecting a small proportion of the fittest chromosomes of one generation and copying them directly into the next generation. Elitism is used to protect the fittest chromosomes from crossover and mutation. The least-cost distribution networks, without violating any pressure and velocity constrains per generation are considered as the elite chromosomes. The 2 best performing individuals from each generation are saved prior to crossover and mutation for the next generation.

2.4.3. Penalty Functions

The following Equations (6) and (7) are used to calculate the penalty cost [63]. The penalty imposed will be higher, the larger the violations of pressure and velocity from the target values. The penalty equation for violation of pressure constraint in the water distribution network implemented by SOP–WDN can be represented as:
P P = 1 + j = 1 N n T P P j   ·   P P 1 + j = 1 N n T P P j   ·   P P 2
where, N n is the number of nodes in the network, PP the pressure penalty, P j is the pressure of node j , T P is the target pressure, P P 1 is the pressure penalty coefficient if the pressure at the node is above the target pressure, and P P 2 is the pressure penalty coefficient if the pressure at the node is below the target pressure.
The penalty equation for the violation of velocity constraint in the distribution network implemented by SOP–WDN can be represented as:
V P = 1 + i = 1 N p T V V i   ·   V P 1 + i = 1 N p T V V i   ·   V P 2
where, N p is the number of pipes in the network, V P   is the velocity penalty, V i is the flow velocity at link i , T V is the target velocity, V P 1 is the velocity penalty coefficient if the velocity at a given link is above target velocity, and V P 2 is the velocity penalty coefficient if the velocity at the link is below target velocity.

2.4.4. Sensitivity Analysis

Optimal solutions are not guaranteed in evolutionary algorithms; hence, maximizing ‘near optimal’ solutions is essential [64]. Genetic algorithm parameters are user input parameters that significantly affect the overall performance and speed of SOP-WDN. These parameters interact in a complex way [65] and must be tuned to obtain better solutions and faster convergence.
A sensitivity analysis was carried out to determine the most influential parameters and to obtain the best parameter combinations for effective execution of the algorithm. The parameters tested were:
  • Population Size
  • Crossover Probability
  • Mutation Rate
  • Velocity Penalty 1
  • Velocity Penalty 2
  • Pressure Penalty 1
  • Pressure Penalty 2
A total of 100 individual runs were performed (2000 iterations each) for any one alteration of any one of the seven parameters. Crossover rate and mutation rate were further tuned in combination since a proper balance between these operators is essential to ensure global optima. In total, 10–12 different values were tested in total for any one parameter and the value assigned per alteration was based on the literature and the preceding results. The final set of parameters selected, and the results obtained are given in Table 3 and Figure 5. It is to be noted that larger population size and iterations may be required for large water distribution networks.

3. Results and Discussion

The conventional approach when testing the functionality, validity, and efficiency of a developed optimization algorithm is to choose some benchmark water distribution network problems and obtain their solution. Benchmark WDNs have provided a common testbed for newly developed optimization algorithms and design approaches. To prove their significance, the developed optimization algorithms are applied to benchmark WDN problems and are compared to the existing algorithms. Using SOP–WDN, some benchmark networks of the literature have been examined.

3.1. Example 1: Two-Loop Network

The two-loop network is an imaginary network introduced by Alperovitz and Shamir [3] that consists of 8 pipelines and 7 nodes (with reservoir), all fed by gravity flow from a single reservoir with an elevation of 210 m. The layout of the two-loop network is given in Figure 6. All pipes in the layout are 1000 m in length, and the Hazen–Williams coefficient is 130. The minimum head requirement in each node is 30 m above ground level. The commercially available diameters for the distribution network are described below on Table 4. The details of the distribution network have been provided in Appendix A:
Table 5 gives the solution obtained by SOP–WDN for the two-loop network. Figure 7 shows the EPANET network layout of the solved network and Figure 8 shows the pressure heads obtained at the nodes of the distribution network. Table 6 compares the solution obtained from SOP–WDN with the solution obtained by other research reports:
The optimal cost of USD 419,000 was obtained for the Two-loop network, and the minimum pressure requirement of 30 m was fulfilled for all nodes. Table 6 shows the results obtained by other research reports for comparison. The optimal cost of USD 419,000 obtained by SOP-WDN for the two-loop network, is same as the solution obtained by Van Dijk et al. and Geem.

3.2. Example 2: Hanoi Network

Hanoi network, located in Vietnam, was first presented by Fujiwara and Kang [66]. It consists of 32 nodes, 34 pipes, and 3 loops, and is fed by gravity from a reservoir with a 100 m fixed head. The layout of the Hanoi network is given in Figure 9. All pipes available for this distribution network have a Hazen–Williams coefficient C of 130. The elevation of all nodes is 0 m, and minimum head limitation is 30 m above ground level. The commercially available diameters for the distribution network are described below on Table 7. The details of the distribution network have been provided in Appendix A.
Table 8 gives the solution obtained by SOP–WDN for the Hanoi network, while Table 9 compares the solution obtained from SOP–WDN with the solution obtained by other research reports. Figure 10 shows the EPANET network layout of the solved network and Figure 11 shows the pressure heads obtained at the nodes of the distribution network.
The optimal cost of USD 6.081 million was obtained for Hanoi network, and the minimum pressure constraint of 30 m was fulfilled for all nodes. It was observed to be the best solution (lowest cost) without the violation of any constraints. The solution obtained by Geem [16] has a lower cost. However, this solution failed to meet the pressure constraint of 30 m at five nodes.

3.3. Example 3: GoYang Network

The GoYang water network is located in South Korea, and consists of 30 pipes, 22 nodes, and 9 loops. This network was first introduced by Kim et al. [67] and is fed by a single fixed pump of 4.52 kW from a 71 m constant head reservoir. The layout of the GoYang network is given in Figure 12. The Hazen–Williams coefficient for all pipes in the network is 100. The minimum head limitation for this network is 15 m above ground level. The commercially available diameters for the distribution network are described below on Table 10. The details of the distribution network have been provided in Appendix A.
Table 11 gives the solution obtained by SOP–WDN for the GoYang network. Figure 13 shows the EPANET network layout of the solved network and Figure 14 shows the pressure heads obtained at the nodes of the distribution network. Table 12 compares the solution obtained from SOP–WDN with the solution obtained by other research reports.
The lowest cost obtained by SOP–WDN for the GoYang network was 177,010,359 Won, which compared to the other studies, is the cheapest cost. The obtained solution also has no nodes containing pressure violations, as all nodes in the distribution network have fulfilled the minimum pressure requirement of 15 m.

4. Conclusions

In this study, the developed Genetic Algorithm based optimization algorithm SOP–WDN was tested on three benchmark water distribution networks, and in comparison, to the other studies, it was able to produce competitive results. EPANET software, which was used for the hydraulic analysis and calculations of the water distribution systems is a well-accepted and utilized software. EPANET–MATLAB toolkit enabled the SOP–WDN algorithm to perform EPANET based calculations directly in the MATLAB environment, which greatly improved the overall computational speed, performance, and efficiency of the algorithm. Hence, the EPANET–MATLAB toolkit can prove to be an important tool that enables the facile use of EPANET software in the MATLAB environment for many different research purposes. SOP–WDN can be used as a reliable algorithm and can easily be implemented and adapted to aid engineers and designers during the design process of new water distribution networks, or the rehabilitation of existing water distribution networks.

Author Contributions

U.S. developed SOP–WDN algorithm and wrote the manuscript of this study; K.-H.H., K.-M.K. and K.G. made recommendations during the writing of the manuscript; K.-T.Y. provided significant suggestions on the methodology and structure of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Korea Environment Industry & Technology Institute (KEITI) through Smart Water City Research Program of 2019002950004 and Development of Water and Sewage Innovation Technology Program of ARQ202001302001, funded by Korea Ministry of Environment (MOE).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Datasets that are restricted and not publicly available.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Node data for the Two-loop network.
Table A1. Node data for the Two-loop network.
Node No.Elevation (m)Demand (m3/h)
1210Reservoir
2150100
3160100
4155120
5150270
6165330
7160200
Table A2. Pipe data for the Two-loop network.
Table A2. Pipe data for the Two-loop network.
Pipe No.Begin NodeEnd NodeLength (m)
1121000
2231000
3241000
4451000
5461000
6671000
7351000
8571000
Table A3. Node data for the Hanoi network.
Table A3. Node data for the Hanoi network.
Node No.Demand (m3/h)
1Reservoir
2890
3850
4130
5725
61005
71350
8550
9525
10525
11500
12560
13940
14615
15280
16310
17865
181345
1960
201275
21930
22485
231045
24820
25170
26900
27370
28290
29360
30360
31105
32805
Table A4. Pipe data for the Hanoi network.
Table A4. Pipe data for the Hanoi network.
Pipe No.Begin NodeEnd NodeLength (m)
112100
2231350
334900
4451150
5561450
667450
778850
889850
9910800
101011950
1111121200
1212133500
131014800
141415500
151516550
1617162730
1718171750
181918800
19319400
203202200
2120211500
222122500
2320232650
2423241230
2524251300
262625850
272726300
281627750
2923281500
3028292000
3129301600
323031150
333231860
342532950
Table A5. Node data for the GoYang network.
Table A5. Node data for the GoYang network.
Node No.Elevation (m)Demand (m3/d)
171.0Reservoir
256.4153.0
353.870.5
454.958.5
556.075.0
657.067.5
753.963.0
854.548.0
957.942.0
1062.130.0
1162.842.0
1258.637.5
1359.337.5
1459.863.0
1559.2445.5
1653.6108.0
1754.879.5
1855.155.5
1954.2118.5
2054.5124.5
2162.931.5
Table A6. Pipe data for the GoYang network.
Table A6. Pipe data for the GoYang network.
Pipe No.Begin NodeEnd NodeLength (m)
112165
223124
334118
44581
556134
6612135
71215202
8222135
9221170
102122113
112220335
122019115
13219345
141917114
15316103
161617261
17171872
18718373
193798
2078110
214898
2289246
23511174
241011102
2561092
2669100
271013130
28121390
291314185
30151490

References

  1. Swamee, P.K.; Sharma, A.K. Design of Water Supply Pipe Networks; John Wiley & Sons: Hoboken, NJ, USA, 2008. [Google Scholar]
  2. Cross, H. Analysis of Flow in Networks of Conduits or Conductors; University of Illinois at Urbana Champaign, College of Engineering, Engineering Experiment Station: Urbana, IL, USA, 1936. [Google Scholar]
  3. Alperovits, E.; Shamir, U. Design of optimal water distribution systems. Water Resour. Res. 1977, 13, 885–900. [Google Scholar] [CrossRef]
  4. Goulter, I.C.; Lussier, B.M.; Morgan, D.R. Implications of head loss path choice in the optimization of water distribution networks. Water Resour. Res. 1986, 22, 819–822. [Google Scholar] [CrossRef]
  5. Quindry, G.E.; Liebman, J.C.; Brill, E.D. Optimization of looped water distribution systems. J. Environ. Eng. Div. 1981, 107, 665–679. [Google Scholar] [CrossRef]
  6. Bifulco, I.; Cirillo, S. Discovery Multiple Data Structures in Big Data through Global Optimization and Clustering Methods. In Proceedings of the 2018 22nd International Conference Information Visualisation (IV), Fisciano, Italy, 10–13 July 2018; pp. 117–121. [Google Scholar]
  7. Goldberg, D.E.; Kuo, C.H. Genetic algorithms in pipeline optimization. J. Comput. Civ. Eng. 1987, 1, 128–141. [Google Scholar] [CrossRef]
  8. Simpson, A.; Murphy, L.; Dandy, G. Pipe Network Optimisation Using Genetic Algorithms; American Society of Civil Engineers: Seattle, WA, USA, 1993. [Google Scholar]
  9. Simpson, A.R.; Dandy, G.C.; Murphy, L.J. Genetic algorithms compared to other techniques for pipe optimization. J. Water Resour. Plan. Manag. 1994, 120, 423–443. [Google Scholar] [CrossRef]
  10. Dandy, G.C.; Simpson, A.R.; Murphy, L.J. An improved genetic algorithm for pipe network optimization. Water Resour. Res. 1996, 32, 449–458. [Google Scholar] [CrossRef]
  11. Savic, D.A.; Walters, G.A. Genetic algorithms for least-cost design of water distribution networks. J. Water Resour. Plan. Manag. 1997, 123, 67–77. [Google Scholar] [CrossRef]
  12. Deb, K.; Agrawal, S. A niched-penalty approach for constraint handling in genetic algorithms. In Artificial Neural Nets and Genetic Algorithms; Springer: Vienna, Austria, 1996; pp. 235–243. [Google Scholar]
  13. Wu, Z.Y.; Simpson, A.R. A self-adaptive boundary search genetic algorithm and its application to water distribution systems. J. Hydraul. Res. 2002, 40, 191–203. [Google Scholar] [CrossRef]
  14. Liong, S.Y.; Atiquzzaman, M. Optimal design of water distribution network using shuffled complex evolution. J. Inst. Eng. Singap. 2004, 44, 93–107. [Google Scholar]
  15. Eusuff, M.; Lansey, K.; Pasha, F. Shuffled frog-leaping algorithm: A memetic meta-heuristic for discrete optimization. Eng. Optim. 2006, 38, 129–154. [Google Scholar] [CrossRef]
  16. Geem, Z.W. Optimal cost design of water distribution networks using harmony search. Eng. Optim. 2006, 38, 259–277. [Google Scholar] [CrossRef]
  17. Di Pierro, F.; Khu, S.T.; Savić, D.; Berardi, L. Efficient multi-objective optimal design of water distribution networks on a budget of simulations using hybrid algorithms. Environ. Model. Softw. 2009, 24, 202–213. [Google Scholar] [CrossRef]
  18. Chandramouli, S.; Malleswararao, P. Reliability based optimal design of a water distribution network for municipal water supply. Int. J. Eng. Technol. 2011, 3, 13–19. [Google Scholar]
  19. Jin, X.; Zhang, J.; Gao, J.L.; Wu, W.Y. Multi-objective optimization of water supply network rehabilitation with non-dominated sorting genetic algorithm-II. J. Zhejiang Univ. Sci. A 2008, 9, 391–400. [Google Scholar] [CrossRef]
  20. Prasad, T.D.; Park, N.S. Multiobjective genetic algorithms for design of water distribution networks. J. Water Resour. Plan. Manag. 2004, 130, 73–82. [Google Scholar] [CrossRef]
  21. Kang, D.; Lansey, K. Revisiting optimal water-distribution system design: Issues and a heuristic hierarchical approach. J. Water Resour. Plan. Manag. 2012, 138, 208–217. [Google Scholar] [CrossRef]
  22. Cisty, M.; Bajtek, Z.; Celar, L. A two-stage evolutionary optimization approach for an irrigation system design. J. Hydroinform. 2017, 19, 115–122. [Google Scholar] [CrossRef]
  23. Martínez-Bahena, B.; Cruz-Chávez, M.A.; Ávila-Melgar, E.Y.; Cruz-Rosales, M.H.; Rivera-Lopez, R. Using a genetic algorithm with a mathematical programming solver to optimize a real water distribution system. Water 2018, 10, 1318. [Google Scholar] [CrossRef]
  24. Bi, W.; Dandy, G. Optimization of water distribution systems using online retrained metamodels. J. Water Resour. Plan. Manag. 2014, 140, 04014032. [Google Scholar] [CrossRef]
  25. Khalifeh, S.; Akbarifard, S.; Khalifeh, V.; Zallaghi, E. Optimization of water distribution of network systems using the Harris Hawks optimization algorithm (Case study: Homashahr city). MethodsX 2020, 7, 100948. [Google Scholar] [CrossRef]
  26. Bilal; Pant, M. Parameter Optimization of Water Distribution Network—A Hybrid Metaheuristic Approach. Mater. Manuf. Process. 2020, 35, 737–749. [Google Scholar] [CrossRef]
  27. Praneeth, P.; Vasan, A.; Srinivasa Raju, K. Pipe size design optimization of water distribution networks using water cycle algorithm. In Harmony Search and Nature Inspired Optimization Algorithm; Springer: Singapore, 2019; pp. 1057–1067. [Google Scholar]
  28. Pankaj, B.S.; Naidu, M.N.; Vasan, A.; Varma, M.R. Self-adaptive cuckoo search algorithm for optimal design of water distribution systems. Water Resour. Manag. 2020, 34, 3129–3146. [Google Scholar] [CrossRef]
  29. Surco, D.F.; Vecchi, T.P.; Ravagnani, M.A. Optimization of water distribution networks using a modified particle swarm optimization algorithm. Water Sci. Technol. Water Supply 2018, 18, 660–678. [Google Scholar] [CrossRef]
  30. Cassiolato, G.; Carvalho, E.P.; Caballero, J.A.; Ravagnani, M.A. Optimization of water distribution networks using a deterministic approach. Eng. Optim. 2021, 53, 107–124. [Google Scholar] [CrossRef]
  31. Pant, M.; Snasel, V. Design optimization of water distribution networks through a novel differential evolution. IEEE Access 2021, 9, 16133–16151. [Google Scholar]
  32. Bi, W.; Xu, Y.; Wang, H. Comparison of searching behaviour of three evolutionary algorithms applied to water distribution system design optimization. Water 2020, 12, 695. [Google Scholar] [CrossRef]
  33. Zhao, R.H.; He, W.Q.; Lou, Z.K.; Nie, W.B.; Ma, X.Y. Synchronization optimization of pipeline layout and pipe diameter selection in a self-pressurized drip irrigation network system based on the genetic algorithm. Water 2019, 11, 489. [Google Scholar] [CrossRef]
  34. Shao, Y.; Yao, H.; Zhang, T.; Chu, S.; Liu, X. An improved genetic algorithm for optimal layout of flow meters and valves in water network partitioning. Water 2019, 11, 1087. [Google Scholar] [CrossRef]
  35. Mala-Jetmarova, H.; Sultanova, N.; Savic, D. Lost in optimisation of water distribution systems? A literature review of system operation. Environ. Model. Softw. 2017, 93, 209–254. [Google Scholar] [CrossRef]
  36. Bene, J.G.; Selek, I.; Hős, C. Comparison of deterministic and heuristic optimization solvers for water network scheduling problems. Water Sci. Technol. Water Supply 2013, 13, 1367–1376. [Google Scholar] [CrossRef]
  37. Pasha, M.F.K.; Lansey, K. Optimal pump scheduling by linear programming. In World Environmental and Water Resources Congress 2009: Great Rivers; American Society of Civil Engineers (ASCE): Kansas City, MO, USA, 2009; pp. 1–10. [Google Scholar]
  38. Martin-Candilejo, A.; Santillan, D.; Iglesias, A.; Garrote, L. Optimization of the design of water distribution systems for variable pumping flow rates. Water 2020, 12, 359. [Google Scholar] [CrossRef]
  39. Cimorelli, L.; Covelli, C.; Molino, B.; Pianese, D. Optimal regulation of pumping station in water distribution networks using constant and variable speed pumps: A technical and economical comparison. Energies 2020, 13, 2530. [Google Scholar] [CrossRef]
  40. Broad, D.R.; Maier, H.R.; Dandy, G.C. Optimal operation of complex water distribution systems using metamodels. J. Water Resour. Plan. Manag. 2010, 136, 433–443. [Google Scholar] [CrossRef]
  41. Brentan, B.M.; Luvizotto, E., Jr.; Montalvo, I.; Izquierdo, J.; Pérez-García, R. Near real time pump optimization and pressure management. Procedia Eng. 2017, 186, 666–675. [Google Scholar] [CrossRef]
  42. Jung, D.; Kang, D.; Kang, M.; Kim, B. Real-time pump scheduling for water transmission systems: Case study. KSCE J. Civ. Eng. 2015, 19, 1987–1993. [Google Scholar] [CrossRef]
  43. Tu, M.Y.; Tsai, F.T.C.; Yeh, W.W.G. Optimization of water distribution and water quality by hybrid genetic algorithm. J. Water Resour. Plan. Manag. 2005, 131, 431–440. [Google Scholar] [CrossRef]
  44. Mackle, G.; Savic, G.A.; Walters, G.A. Application of genetic algorithms to pump scheduling for water supply. In Proceedings of the First International Conference on Genetic Algorithms in Engineering Systems: Innovations and Applications IET, Sheffield, UK, 12–14 September 1995; pp. 400–405. [Google Scholar]
  45. Van Zyl, J.E.; Savic, D.A.; Walters, G.A. Operational optimization of water distribution systems using a hybrid genetic algorithm. J. Water Resour. Plan. Manag. 2004, 130, 160–170. [Google Scholar] [CrossRef]
  46. de Oliveira Turci, L.; Sun, H.; Bai, M.; Wang, J.; Hu, P. Water pump station scheduling optimization using an improved genetic algorithm approach. In Proceedings of the Congress on Evolutionary Computation (CEC), Wellington, New Zealand, 10–13 June 2019; pp. 944–951. [Google Scholar]
  47. Mala-Jetmarova, H.; Sultanova, N.; Savic, D. Lost in optimisation of water distribution systems? A literature review of system design. Water 2018, 10, 307. [Google Scholar] [CrossRef]
  48. Khan, S.; Grudniewski, P.; Muhammad, Y.S.; Sobey, A.J. The benefits of co-evolutionary Genetic Algorithms in voyage optimisation. Ocean. Eng. 2022, 245, 110261. [Google Scholar] [CrossRef]
  49. Caruccio, L.; Deufemia, V.; Polese, G. A genetic algorithm to discover relaxed functional dependencies from data. In Proceedings of the 25th Italian Symposium on Advanced Database Systems; University of Salerno: Fisciano, Italy, 2017; Volume 2037, pp. 1–12. [Google Scholar]
  50. Mayer, M.J.; Szilágyi, A.; Gróf, G. Environmental and economic multi-objective optimization of a household level hybrid renewable energy system by genetic algorithm. Appl. Energy 2020, 269, 115058. [Google Scholar] [CrossRef]
  51. Katoch, S.; Chauhan, S.S.; Kumar, V. A review on genetic algorithm: Past, present, and future. Multimed. Tools Appl. 2021, 80, 8091–8126. [Google Scholar] [CrossRef]
  52. Van Dijk, M.; van Vuuren, S.J.; Van Zyl, J.E. Optimising water distribution systems using a weighted penalty in a genetic algorithm. Water SA 2008, 34, 537–548. [Google Scholar] [CrossRef]
  53. Lansey, K.E.; Mays, L.W. Optimization models for design of water distribution systems. In Reliability Analysis of Water Distribution Systems; Mays, L.R., Ed.; ASCE: New York, NY, USA, 1989; pp. 37–84. [Google Scholar]
  54. Keedwell, E.; Khu, S.T. A hybrid genetic algorithm for the design of water distribution networks. Eng. Appl. Artif. Intell. 2005, 18, 461–472. [Google Scholar] [CrossRef]
  55. Lamont, P.A. Common pipe flow formulas compared with the theory of roughness. J. Am. Water Work. Assoc. 1981, 73, 274–280. [Google Scholar] [CrossRef]
  56. Goldberg, D.E. A Gentle Introduction to Genetic Algorithms; Addison-Wesley Publishing Company: Boston, MA, USA, 1989. [Google Scholar]
  57. Abuiziah, I.; Nidal, S. A Review of Genetic Algorithm Optimization: Operations and Applications to Water Pipeline Systems. Int. J. Math. Comput. Phys. Electr. Comput. Eng. 2013, 7, 341–347. [Google Scholar]
  58. Rossman, L.A. EPANET 2: User’s Manual; United States Environmental Protection Agency: Washington, DC, USA, 2000. [Google Scholar]
  59. Eliades, D.G.; Kyriakou, M.; Vrachimis, S.; Polycarpou, M.M. EPANET-MATLAB toolkit: An open-source software for interfacing EPANET with MATLAB. In Proceedings of the 14th International Conference on Computing and Control for the Water Industry, Amsterdam, The Netherlands, 7–9 November 2016. [Google Scholar]
  60. Czajkowska, A.M. Maximum Entropy Based Evolutionary Optimization of Water Distribution Networks under Multiple Operating Conditions and Self-Adaptive Search Space Reduction Method. Ph.D. Thesis, University of Strathclyde, Glasgow, UK, 2016. [Google Scholar]
  61. Saleh, S.H.; Tanyimboh, T.T. Optimal design of water distribution systems based on entropy and topology. Water Resour. Manag. 2014, 28, 3555–3575. [Google Scholar] [CrossRef]
  62. Tanyimboh, T.T. Redundant binary codes in genetic algorithms: Multi-objective design optimization of water distribution networks. Water Supply 2021, 21, 444–457. [Google Scholar] [CrossRef]
  63. Güç, G. Optimization of Water Distribution Networks Using Genetic Algorithm. Master’s Thesis, Middle East Technical University, Ankara, Turkey, 2006. [Google Scholar]
  64. Mora-Melia, D.; Iglesias-Rey, P.L.; Martinez-Solano, F.J.; Fuertes-Miquel, V.S. Design of water distribution networks using a pseudo-genetic algorithm and sensitivity of genetic operators. Water Resour. Manag. 2013, 27, 4149–4162. [Google Scholar] [CrossRef]
  65. Eiben, A.E.; Michalewicz, Z.; Schoenauer, M.; Smith, J.E. Parameter control in evolutionary algorithms. In Parameter Setting in Evolutionary Algorithms; Springer: Berlin/Heidelberg, Germany, 2007; pp. 19–46. [Google Scholar]
  66. Fujiwara, O.; Kang, D.B. A two-phase decomposition method for optimal design of looped water distribution networks. Water Resour. Res. 1990, 26, 539–549. [Google Scholar] [CrossRef]
  67. Kim, J.H.; Kim, T.G.; Kim, J.H.; Yoon, Y.N. A study on the pipe network system design using non-linear programming. J. Korean Water Resour. Assoc. 1994, 27, 59–67. [Google Scholar]
  68. Menon, K.K.; Narulkar, S.M. Application of heuristic based algorithm in water distribution network design. J. Water Resour. Pollut. Stud. 2016, 1, 1–14. [Google Scholar]
Figure 1. General Flowchart of the SOP–WDN.
Figure 1. General Flowchart of the SOP–WDN.
Water 14 00851 g001
Figure 2. Example: Pipe network showing pipe position and diameter.
Figure 2. Example: Pipe network showing pipe position and diameter.
Water 14 00851 g002
Figure 3. k-point crossover.
Figure 3. k-point crossover.
Water 14 00851 g003
Figure 4. Bit flip mutation.
Figure 4. Bit flip mutation.
Water 14 00851 g004
Figure 5. Sensitivity analysis results.
Figure 5. Sensitivity analysis results.
Water 14 00851 g005
Figure 6. The Two-loop network layout.
Figure 6. The Two-loop network layout.
Water 14 00851 g006
Figure 7. The Two-loop network solution obtained (pressure and velocity).
Figure 7. The Two-loop network solution obtained (pressure and velocity).
Water 14 00851 g007
Figure 8. The Two-loop network nodal pressure heads.
Figure 8. The Two-loop network nodal pressure heads.
Water 14 00851 g008
Figure 9. The Hanoi network layout.
Figure 9. The Hanoi network layout.
Water 14 00851 g009
Figure 10. The Hanoi network solution obtained (pressure and velocity).
Figure 10. The Hanoi network solution obtained (pressure and velocity).
Water 14 00851 g010
Figure 11. The Hanoi network nodal pressure heads.
Figure 11. The Hanoi network nodal pressure heads.
Water 14 00851 g011
Figure 12. The GoYang network layout.
Figure 12. The GoYang network layout.
Water 14 00851 g012
Figure 13. The GoYang network solution obtained (pressure and velocity).
Figure 13. The GoYang network solution obtained (pressure and velocity).
Water 14 00851 g013
Figure 14. The GoYang network nodal pressure heads.
Figure 14. The GoYang network nodal pressure heads.
Water 14 00851 g014
Table 1. Example: Available pipe sizes.
Table 1. Example: Available pipe sizes.
S.N.Available Pipe Sizes (mm)Unit Cost (per m)3-Bit (Binary) Representation3-Bit (Grey) Representation
18100000000
210120001001
312150010011
414180011010
516200100110
618250101111
720300110101
824350111100
Table 2. Example: Interpretation of binary to pipe networks.
Table 2. Example: Interpretation of binary to pipe networks.
Randomly Generated
Binary String
101110011001000111
6 Individual Pipes101, 110, 011, 001, 000, 111
Pipe Position in WDNPipe No.1Pipe No.2Pipe No.3Pipe No.4Pipe No.5Pipe No.6
Chromosome (Binary)101110011001000111
Pipe Diameter (mm)20161210818
Unit Cost (per m)300200150120100250
Length (m)100100100100100100
Cost of individual pipe30,00020,00015,00012,00010,00025,000
Total Cost of Network112,000
Table 3. Set of Genetic Algorithm parameters.
Table 3. Set of Genetic Algorithm parameters.
S.N.GA ParametersValues
1Population Size80–100
2Crossover Probability (%)85–90
3Mutation Rate (%)4–6
4Velocity Penalty 10.3
5Velocity Penalty 20.06
6Pressure Penalty 10.02
7Pressure Penalty 21.9
Table 4. Available pipes for selection for the Two-loop network.
Table 4. Available pipes for selection for the Two-loop network.
Diameter (in)Diameter (mm)Unit Cost (USD/m)
125.42
250.85
4101.611
6152.416
10254.032
14355.660
16406.490
18457.2130
Table 5. SOP–WDN results for the Two-loop network.
Table 5. SOP–WDN results for the Two-loop network.
Pipe No.Pipe Diameter (mm)Pipe Length (m)Cost (USD)Node No.Nodal Pressure (m)
1457.21000130,0001Reservoir
2254.0100032,000253.25
3406.4100090,000330.46
4101.6100011,000443.45
5406.4100090,000533.81
6254.0100032,000630.44
7254.0100032,000730.55
825.410002000--
Total Cost:419,000CheckOK
Table 6. Comparison of SOP–WDN results to past studies for the Two-loop network.
Table 6. Comparison of SOP–WDN results to past studies for the Two-loop network.
StudiesAlperovitz and ShamirSavic and WaltersGeemVan Dijk et al.SOP–WDN
Least cost obtained (USD)479,525420,000419,000419,000419,000
Table 7. Available pipes for selection for the Hanoi network.
Table 7. Available pipes for selection for the Hanoi network.
Diameter (in)Diameter (mm)Unit Cost (USD/m)
12304.845.73
16406.470.40
2050898.38
24609.6129.33
30762180.75
401016278.28
Table 8. SOP–WDN results for the Hanoi network.
Table 8. SOP–WDN results for the Hanoi network.
Pipe No.Pipe Diameter (mm)Pipe Length (m)Cost (USD)Node No.Nodal Pressure (m)
1101610027,8281100 (Reservoir)
210161350375,678297.14
31016900250,452361.67
410161150320,022456.92
510161450403,506551.02
61016450125,226644.81
71016850236,538743.35
81016850236,538841.61
91016800222,624940.23
10762950171,712.51039.20
11609.61200155,1961137.64
12609.63500452,6551234.21
1350880078,7041330.01
14406.450035,2001435.52
15304.855025,151.51533.72
16304.82730124,842.91631.30
17406.41750123,2001733.41
18609.6800103,4641849.93
1950840039,3521955.09
2010162200612,2162050.61
215081500147,5702141.26
22304.850022,8652236.10
2310162650737,4422344.52
247621230222,322.52438.93
257621300234,9752535.34
2650885083,6232631.70
27304.830013,7192730.76
28304.875034,297.52838.94
29406.41500105,6002930.13
30304.8200091,4603030.42
31304.8160073,1683130.70
32406.415010,5603233.18
33406.486060,544--
34609.6950122,863.5--
Total Cost:6,081,115.4CheckOK
Table 9. Comparison of SOP–WDN results to past studies for the Hanoi network.
Table 9. Comparison of SOP–WDN results to past studies for the Hanoi network.
StudiesSavic and WaltersLiong and AtiquzzamanGeemVan Dijk et al.SOP–WDN
Least cost obtained (Million USD)6.1876.2206.0566.1106.081
Table 10. Available pipes for selection for the GoYang network.
Table 10. Available pipes for selection for the GoYang network.
Diameter (mm)Unit Cost (Won/m)
8037,890
10038,933
12540,563
15042,554
20047,624
25054,125
30062,109
35071,524
Table 11. SOP–WDN results for the GoYang network.
Table 11. SOP–WDN results for the GoYang network.
Pipe No.Pipe Diameter (mm)Pipe Length (m)Cost (Won)Node No.Nodal Pressure (m)
1200165.07,857,960115.62
2125124.05,029,812229.33
3125118.04,786,434328.73
410081.03,153,573426.58
580134.05,077,260524.20
680135.05,115,150621.51
780202.07,653,780727.72
880135.05,115,150826.70
980170.06,441,300921.20
1080113.04,281,5701016.17
1180335.012,693,1501116.03
1280115.04,357,3501218.16
1380345.013,072,0501317.46
1480114.04,319,4601415.33
1580103.03,902,6701515.48
1680261.09,889,2901628.31
178072.02,728,0801726.75
1880373.014,132,9701826.44
198098.03,713,2201927.36
2080110.04,167,9002026.68
218098.03,713,2202119.74
2280246.09,320,9402219.36
2380174.06,592,860--
2480102.03,864,780--
258092.03,485,880--
2680100.03,789,000--
2780130.04,925,700--
288090.03,410,100--
2980185.07,009,650--
308090.03,410,100--
Total Cost:177,010,359CheckOK
Table 12. Comparison of the SOP–WDN results to those of past studies of the GoYang network.
Table 12. Comparison of the SOP–WDN results to those of past studies of the GoYang network.
StudiesOriginal NetworkKim et al.GeemMenon et al. [68]SOP–WDN
Least cost obtained (Million Won)179.428179.142177.135177.417177.010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sangroula, U.; Han, K.-H.; Koo, K.-M.; Gnawali, K.; Yum, K.-T. Optimization of Water Distribution Networks Using Genetic Algorithm Based SOP–WDN Program. Water 2022, 14, 851. https://doi.org/10.3390/w14060851

AMA Style

Sangroula U, Han K-H, Koo K-M, Gnawali K, Yum K-T. Optimization of Water Distribution Networks Using Genetic Algorithm Based SOP–WDN Program. Water. 2022; 14(6):851. https://doi.org/10.3390/w14060851

Chicago/Turabian Style

Sangroula, Uchit, Kuk-Heon Han, Kang-Min Koo, Kapil Gnawali, and Kyung-Taek Yum. 2022. "Optimization of Water Distribution Networks Using Genetic Algorithm Based SOP–WDN Program" Water 14, no. 6: 851. https://doi.org/10.3390/w14060851

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop