Next Article in Journal
A Spiral-Propulsion Amphibious Intelligent Robot for Land Garbage Cleaning and Sea Garbage Cleaning
Next Article in Special Issue
Shape-Informed Dimensional Reduction in Airfoil/Hydrofoil Modeling
Previous Article in Journal
A Review on the Antimicrobial Activity of Chitosan Microspheres: Milestones Achieved and Miles to Go
Previous Article in Special Issue
Machine-Learning-Enabled Foil Design Assistant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sequential Design-Space Reduction and Its Application to Hull-Form Optimization

1
Key Laboratory of High Performance Ship Technology, Wuhan University of Technology, Ministry of Education, Wuhan 430063, China
2
School of Naval Architecture, Ocean and Energy Power Engineering, Wuhan University of Technology, Wuhan 430063, China
3
China Ship Development and Design Center, Wuhan 430063, China
*
Authors to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2023, 11(8), 1481; https://doi.org/10.3390/jmse11081481
Submission received: 14 June 2023 / Revised: 20 July 2023 / Accepted: 24 July 2023 / Published: 25 July 2023
(This article belongs to the Special Issue Machine Learning and Modeling for Ship Design)

Abstract

:
Hull-form optimization is a complex engineering problem. Owing to the several numerical simulations and complex design-performance spaces, hull-form optimization is considered an inefficient process, which makes determining the global optimum difficult. This study used rough set theory (RST) to acquire knowledge and reduce the design space for hull-form optimization. Furthermore, we studied one of the hull-form optimization problems by practically applying RST to the appropriate number of sampling points. To solve this problem, we proposed the RST-based sequential design-space reduction (SDSR) method that uses interval theory to calculate subspace intersections and unions, as well as test calculations to choose an appropriate stopping criterion. Finally, SDSR was used to optimize a KRISO container ship to minimize the wave-making resistance. The results were compared to those of direct optimization and one-time design-space reduction, thus proving the feasibility of this method.

1. Introduction

Owing to the increasing need for energy resources and the growing ubiquity of marine resource exploitation, many countries and international organizations have imposed stringent environmental requirements on shipping industries. The Energy Efficiency Design Index (EEDI), which was made mandatory by the International Maritime Organization (IMO), has directly affected the core competitiveness and international market share of the shipbuilding industry, resulting in new challenges in the development of green ships. Therefore, extensive efforts must be made to develop energy-efficient ships to address this challenge.
Considering the rapid development of computational fluid dynamics (CFD) and computer technology over the past few decades, CFD-based hull-form optimization has become an effective method for developing energy-efficient ships. Gammon [1] used a multiobjective genetic algorithm to perform multiobjective optimization of the resistance performance, seakeeping performance, and stability of fishing boats. Feng et al. [2,3] used the potential flow method to optimize the resistance performance of a variety of hull forms. Cheng et al. [4] used Shipflow to optimize an S60 hull form. Serani et al. [5] optimized the hull form of the DTMB 5415 model to reduce the resistance for Fn (Froude number) = 0.25. Yang et al. [6] used response surface models and multiobjective optimization algorithms. Kim and Yang [7] used radial basis function (RBF) to vary the hull forms during the optimization process. Zhang et al. [8,9] presented a ship hull-form optimization loop using the surrogate model to reduce the wave-making resistance of the Wigley ship. Jafaryeganeh et al. [10] applied a multiobjective genetic algorithm to deal with the optimization of the internal layout of oil tankers under uncertainties. Feng et al. [11] selected the Wendland ψ3,1 function as the basis functions of RBF interpolation, and the modification method was used to optimize a trimaran model.
In hull-form optimization, techniques such as parametric hull-surface modification and CFD numerical simulation and optimization are directly used to obtain optimal ship designs that satisfy a given set of constraints. However, considering hull-form optimization is a complex engineering problem, it involves several numerical simulations and has a complex design-performance space. Consequently, the hull-form optimization process is highly inefficient, making it difficult to determine the global optimum. The most commonly used approaches to solve these problems are as follows:
(1) High-efficiency optimization algorithms: The approach is used to develop an algorithm that can determine the global optimum using a small number of iterations. (2) Approximate modeling: This method uses a set of numerically simulated samples to construct an approximate model that can replace CFD computations, significantly reducing computational costs. (3) Using high-performance computing clusters: Modern computing hardware and parallel computing technologies can be used to increase computational speed and decrease computing time [12].
Although significant progress has been made in the aforementioned techniques, their efficiency and the quality of the resulting solutions are insufficient for practical application.
Therefore, many studies have been performed on design-space reduction. In studies conducted by Wang [13,14,15], Chu [16], Cheng et al. [17], Zhou et al. [18], Li et al. [19], Chi [20], Tseng et al. [21], Qiu et al. [22], Ye et al. [23], and D’Agostino et al. [24], design-space reduction was combined with optimization theory, which significantly improved the optimization efficiency and product quality.
This study applies design-space reduction to hull-form optimization by simulating a set of samples prior to hull-form optimization and using rough set theory (RST) to study the internal features of the design space to identify the design space that contains the optimal ship. Figure 1 shows an example of design-space reduction, where circle represents sample. Figure 1a shows the initial design space. The optimization space was large and exhibited complex internal characteristics. Therefore, the initial design space was sampled to obtain a uniform distribution of simulated sample points in the design space. Design-space reduction was then used to analyze and delineate the design space to locate attractive subspaces (i.e., R1, R2, and R3 in Figure 1b), which significantly reduced the range of values considered by the design variables.
RST has various applications in data preprocessing and design-space reduction and is often combined with other methods. However, several limitations hinder the use of design-space reduction in design optimization. For example, the number of sampling points are usually chosen using a subjective, experience-based approach; the larger the number of samples, the greater the likelihood that it captures the characteristics of the design space. However, a large number of sampling points significantly increase the required number of numerical simulations and computational costs. While reducing the number of sampling points might reduce the computational costs, it increases the risk that the sampled points may not be representative of the design space. To solve these problems, we propose an RST-based sequential design-space reduction (SDSR) method for ship design optimization. First, interval theory is used to compute the intersection and union of subspaces. Then, a stopping criterion suitable for hull-form optimization is proposed to determine the termination conditions of the iteration.
The remainder of this paper is organized as follows. Section 1 presents the introduction. Section 2 introduces RST-based design-space reduction. Section 3 describes the proposed RST-based SDSR method and discusses its application to some problems. In Section 4, a KCS ship is optimized using the genetic algorithm, RST-based design-space reduction, and RST-based SDSR. The results are then analyzed and summarized in Section 5.

2. Design-Space Reduction Based on Rough Set Theory

RST, a mathematical tool that requires no a priori information other than samples that represent the problem at hand, was proposed by Pawlak in 1982. Therefore, RST is an objective method that characterizes or processes a problem. The aim of RST is to extract the rules that govern a set of elements through element classification and set-logic operations. In particular, RST uses indiscernibility relations, lower approximations, upper approximations, and reductions to characterize and express an information system. Interested readers can refer to the works by Pawlak [25] for a more detailed explanation of RST.
Figure 2 describes the procedure for RST-based design-space reduction.
Step 1: Construction of the initial information system. First, the design variables and their range of values are ascertained depending on the design optimization problem to determine the initial design space. Then, the design space is sampled and the target value of each sample is calculated to construct the data sample set.
Step 2: Discretization of continuous data. All continuous target values and variables in the decision table are transformed into discrete data. First, the target values are discretized using defining boundary values based on subject area knowledge, experience, or rules to classify the target values into two to three grades, which are expressed using different numbers or symbols. Then, the variables are discretized by defining a number of cuts followed by partitioning the attribute space into finite subspaces based on these cuts to ensure the values of the objects in all the subspaces are divided into the same intervals.
Step 3: Reduction in design variables. The discretized decision table is simplified by removing attributes that do not affect the classification power and knowledge of the decision table.
Step 4: Identifying attractive spaces. After attribute discretization and variable reduction, the simplified decision table is used to derive causal relations between the decision attributes and condition attributes. The rules that contain the minimum decision value (target value) are particularly important, considering the range of values of their condition attributes (variables) constitute an attractive space. The sum of all attractive spaces corresponds to the new design space obtained from design-space reduction.

3. RST-Based Sequential Design-Space Reduction

RST is applied to ship simulation samples in this application of RST-based design-space reduction. The larger the number of samples, the greater the probability of the characteristics space to be captured by RST. However, a larger number of samples significantly increase the computational cost associated with numerical simulations. Conversely, a low number of samples could fail to represent the design space despite being computationally efficient. Therefore, a sequential sampling strategy, called the sequential design-space reduction (SDSR) method, was used.

3.1. Optimization with Sequential Design-Space Reduction

In sequential sampling, the sampling process is divided into a number of iterations. In each iteration, the current to-be-sampled region is determined from the previous iteration. The sampling points are then added accordingly to provide the maximum amount of new information about the original model.
Research about sequential sampling began in the 1990s when Jones et al. [26] proposed the DIRECT (Dividing Rectangles) algorithm. In this algorithm, the search space is partitioned into hyperrectangles and the centerpoint of each hyperrectangle is evaluated. The hyperrectangle that potentially contains the optimum solution is selected based on the previously evaluated centerpoints, before being divided into another set of hyperrectangles to create the next search space. This process continues until the convergence criterion is satisfied.
SDSR is a technique that uses a relatively small number of sampling points to reflect the characteristics of the design space. In this technique, the number of sampling points and search space in each iteration are determined based on the information from the previous sampling iteration. In this study, SDSR was used to progressively shrink the design space while adding sampling points in the subdesign spaces to ultimately identify the optimal ship parameters.
Figure 3 shows the SDSR optimization procedure. Module 1 constructs the sample set of the design space, whereas Module 2 performs the design-space reduction according to the sample set to obtain the attractive spaces Sa. If the stopping condition is satisfied, the iteration stops. This step is followed by the optimization or selection of the optimal solution. If the stopping condition is not satisfied, it indicates that the number of sampled points are insufficient. In this case, more sampling points are added to the design space (while retaining all existing sampling points), and the simulated sample set is updated accordingly by repeating the iterative process (Modules 1 and 2).
During this iterative process, the attractive design spaces Sai+1 and Sa gradually overlap each other while approaching the optimal design space So, as shown in Figure 4. Therefore, the relationship between Sai+1 and Sai (especially the intersection between these subspaces) can be used to adjudge whether the design space is sufficiently reduced. This relationship will be analyzed in Appendix A.

3.2. Solution for Subspace Intersections

The BIAS for Matlab (b4m) interval arithmetic toolbox was used to solve subspace intersections and calculate the overlap coefficient C. The procedure is as follows: (1) The minimum interval of each attribute (as defined by its cuts) is expressed in vector form. (2) The minimum intervals of each attribute are merged to form minimum subspaces. (3) The volume of all minimum subspaces is computed. (4) Containment relations between the minimum subspaces and Sai are adjudged, and Sai is expressed as a set of minimum subspaces to determine its volume Vol (Sai). (5) The intersection and union vectors of Sai and Sai+1 are obtained by adjudging the equivalence of their minimum subspaces. (6) Finally, the overlap coefficient C (the stopping criterion) is computed using the union and intersection volumes of Sai and Sai+1.
C = V o l ( S a i S a ( i + 1 ) ) V o l ( S a i S a ( i + 1 ) )
where Vol is the volume of the design space.
Appendix B shows the exact procedure of the aforementioned process.

3.3. Setting the Stopping Condition for Sequential Design-Space Reduction

The stopping condition (i.e., C > C0) is a criterion that determines whether the iteration should stop or continue. In other words, it is an indirect solution to the question of “how many sampling points should be sampled”. As the iteration progresses, the Sa gradually approaches the optimal design space. However, owing to the high computational cost, it is not viable to allow Sa to be infinitely close to So. Herein, we use common test functions to study how C changes when Sai and Sai+1 are reduced during the iteration to determine C0. Appendix A shows the validation work to determine C0.
By analyzing the effect of changes in the design subspaces during SDSR on the SC and Hartman functions, it was found that each additional iteration reduced the design space and increased the overlap coefficient C up to or close to 1, as shown in Equation (2).
lim C = lim i V o l ( S a i S a ( i + 1 ) ) V o l ( S a i S a ( i + 1 ) ) = 1
However, the total number of samples increased when SDSR was applied to problems of greater complexity. Considering the complexity and computational cost of hull-form optimization problems, C > 0.85 was chosen as the stopping criterion for these problems to ensure SDSR was reliably performed within a reasonable amount of time.

4. Application of SDSR to Hull-Form Optimization and Design

Herein, a KRISO container ship (KCS) (see Figure 5) model is optimized to minimize its wave-making resistance. The main parameters of the KCS model are shown in Table 1.

4.1. Optimization Object and Constraints

  • Optimization Object
The aim of the optimization is to minimize the wave-making resistance at a Froude number (Fr) of 0.26, as shown below.
min f = R w , F r = 0.26
Note: Fr = 0.26 is the design speed of a KCS. The wave-making resistance was calculated using the Shipflow fluid dynamics software suite and the optimization was performed in the fully constrained conditions, without considering the heave and trim. The computational domain is 1 times the length (Lpp) in width, 0.5 times the length in height forward of the bow, and 1.2 times the length in height behind the stern. There are a total of 7783 surface elements in the mesh of the free surface and hull surface. Figure 6 shows the hull surface and free-flow surface element mesh.
  • Constraints
To prevent excessively large changes in deadweight and buoyancy, the hydrostatic condition was imposed, where the displacement and longitudinal center of buoyancy Lcb were restricted to changes within ±1%. The grid nodes at the deck edge line and keel and near the midship cross-section were fixed.

4.2. Method for Hull-Surface Modification and To-Be-Optimized Variables

In this study, radial basis function (RBF) interpolation was used to perform hull-surface modification [4,7,11,27]. Five representative nonuniform rational basis spline (NURBS) control nodes (numbered 1–5) were used as movable control nodes (see Figure 7) on the hull. The direction and range of movement of the control nodes were configured based on our experience.
As shown in Figure 7, control node 1 was movable along the x (length of the ship) and z (draught) directions, whereas nodes 2–5 were movable along the y (breadth of the ship) direction. Control node 1 determines the length and height of the bulbous bow, whereas control node 2 sets the breadth of the bulbous bow and the shape of the bilge. Control nodes 3 and 5 adjust the entrance angle and shape of the entrance at the water plane, respectively, whereas control nodes 2 and 4 adjust the shape of the bow and bilge, respectively. Control nodes 2–5 are also used to adjust the degree of U- or V-shape of the hull.
Therefore, there are a total of six variables: X1, Z1, Y2, Y3, Y4, and Y5, and the value range of these variables (shown in Table 2) constitutes the initial design space for hull-form optimization. These value ranges were chosen based on our experience to preclude excessively large changes in the hull. Based on our analysis, the chosen value ranges were sufficient to cause significant changes in the shape of the hull and minimize its wave-making resistance.

4.3. Results and Analysis

4.3.1. Case 1: CFD-Based Hull-Form Optimization

The multi-island genetic algorithm was used to directly optimize the wave-making resistance within the initial design space. Figure 8 shows the convergence graph of the algorithm over 1000 CFD calculations. Table 3 shows the parameters giving the lowest wave-making resistance, whereas Figure 9 shows the lines plan of the optimized ship design.
As shown in Table 3, displacement increased by 0.4%, whereas the wave-making resistance decreased by 9.7% in the optimized ship. Considering the wave-making resistance only accounts for a small proportion of the total resistance, the total resistance decreased by 0.4%. In the optimized ship, the bulbous bow extended forward and upward in a “wing-like” shape, effectively increasing the ship length below the waterline and the entrance length, thereby reducing the wave-making resistance. Furthermore, the changes made to control nodes 2, 3, and 4 increased in breadth at the bow. However, the breadth decreased around control node 5.
According to the dynamic pressure distributions shown in Figure 10, the area of the low-pressure area behind the bulbous bow became smaller after optimization, which reduced the steepness of the pressure gradient across the bow. The longitudinal wave cuts at different parts of the KCS before and after optimization are shown in Figure 11. The wave amplitude at the first 1/4 of the ship (from the bow) decreased significantly after optimization, indicating that the optimized ship was wasting less energy on making waves and undertaking less work to overcome the wave-making resistance, thus explaining the improved wave-making resistance in the optimized ship.

4.3.2. Case 2: RST-Based Design-Space Reduction

First, an initial information system was constructed based on the aforementioned to-be-optimized variables and their value ranges. Then, the system was discretized to facilitate the extraction of attractive spaces using RST. Finally, gradient optimization was used to determine the optimal solution within each of the subspaces and produce the optimal ship design.
Firstly, a uniform design was used to generate a uniform distribution of sampling points in the initial design space (61 sampling points). Then, RBF interpolation was used to generate the corresponding ship models. Finally, CFD was used to compute the drag performance of each ship design. Using the procedure described in Section 2, 15 rules were obtained, as shown in Table 4. In this table, the “support” column shows the number of objects supporting the rule. Rules may be interpreted as a certain target value being obtained when the variables take values in certain ranges.
For example, Rule 1 may be interpreted as follows: if Y2, Y3, Y4, and X1 take values within (−Inf, 0.0719), (0.097, Inf), (0.13415, Inf), and (−Inf, −0.06335), respectively, the decision attribute is then 0.
Y5 and Z1 did not appear in any of the rules. According to RST, this can be interpreted as Y5 and Z1 not affecting the rules, regardless of their values. In theory, Y5 and Z1 can be treated as initial coordinates (i.e., the values of the original ship) in the hull-form optimization problem. However, if these coordinates are fixed, the variations in the other coordinates could result in unsmooth hull surfaces. Therefore, the optimization ranges of both these variables were still used as their initial range of values in the optimization.
As shown in Table 4, there are five target values with the rule “d = 0” (rules 1–5); hence, five attractive spaces were generated after design-space reduction (Table 5).
If any pair of subspaces have a variable with adjacent value ranges and all their other variables have the same value ranges, this pair of subspaces can be merged into one subspace. For example, subspaces 1 and 2 have adjacent Y4 value ranges and all their other variables have identical value ranges. Therefore, subspaces 1 and 2 can be merged by combining the value ranges of Y4 to create a new subdesign space. Two subdesign spaces were obtained by merging the five subspaces in Table 5, as shown in Table 6. Furthermore, the last two columns in this table show the volume of each subspace and its percentage of the initial design space, indicating that the total volume of the two subspaces is 4.867 × 10−10, i.e., 11.55% of the size of the initial design space. Therefore, there was a significant reduction in the design space.
The nonlinear programming by quadratic lagrangian (NLPQL) method was used to directly search the two subspaces. The convergence graph is shown in Figure 12. A total of 347 CFD calculations were performed, wherein 286 calculations were performed during the optimization and 61 CFD calculations were performed to obtain the initial designs. The solution that minimized the target value was selected as the optimal solution, which is analyzed below.
Figure 13 shows the lines plan before and after optimization. The bulbous bow was extended forward and upward. Furthermore, the surfaces at stations 19 and 20 expanded outward, while the surfaces at stations 17 and 18 became more concaved. Therefore, a ship that has Fr = 0.26, a slightly raised bulbous bow, slightly convexed surfaces at stations 18 and 19, and slightly concaved surface at station 16 can significantly reduce the wave-making resistance.
Based on the dynamic pressure distribution shown in Figure 14, the low-pressure area at the bulbous bow became smaller after optimization, and the maximum pressure decreased. As a result, the pressure gradient across the bow became more uniform. As shown in Figure 15, the wave amplitudes decreased in the first 1/2 of the ship (from the bow) after optimization, indicating that less energy was wasted on wave-making on the optimized ship.

4.3.3. Case 3: RST-Based SDSR

1st reduction: Two subspaces were obtained by reducing the initial set of simulated samples (see Table 7). To facilitate comparison, the initial sample set in Case 3 was kept identical to that in Case 2. As a result, the subspaces obtained after one reduction were identical to those in Case 2.
2nd reduction: Uniform design was used to add new ship-type sample points in subspaces 1 and 2 (from the 1st reduction). CFD calculations were then performed to calculate their wave-making resistances and form two sets of sample points with the pre-existing sample points in each subspace. Data discretization, variable simplification, and rule extraction were performed on these sample points to finish the second round of design-space reduction, which resulted in two subspaces (see Table 8).
3rd reduction: First, a set number of sample points were added to subspaces 3 and 4 (from the 2nd reduction), which were then combined with the original sample points in these subspaces to create two sets of sample points. RST was then used to reduce the two subspaces based on these sample point sets. Thus, subspace 3 was reduced and divided into “subspace 5” and “subspace 6”, whereas subspace 4 was reduced to “subspace 7”. The three subspaces obtained from the 3rd reduction are shown in Table 9.
4th reduction: A set number of sample points were added to the subspaces obtained from the 3rd reduction to create three sets of sample points, which were then combined with the pre-existing sample points in these subspaces to construct three sets of sample points. RST was then used to perform design-space reduction on these sample point sets. During this process, subspace 5 was divided into “subspace 8” and “subspace 9”, while subspaces 6 and 7 were reduced to “subspace 10” and “subspace 11”, respectively. The four subspaces obtained from the 4th reduction are shown in Table 10.
Table 11 shows how the reduction efficacy, overlap coefficient, and other relevant factors changed with each iteration in the SDSR method. Herein, “reduction efficacy” refers to the ratio of the reduced design space to the initial design space in terms of percentage, which is negatively correlated with the overlap coefficient. Figure 16 shows how the overlap coefficient and reduction efficacy changed with each iteration. The overlap coefficient increased with each iteration, and considering the overlap coefficient was already 0.88 by the 4th iteration (i.e., greater than 0.85), no further reductions were performed. Considering the volume of the final design space was only 3.61% of the initial design space, the reduction in the design space was significant.
A total of 113 CFD calculations were performed in Case 3. The ship with the smallest target value among these 113 samples was selected as the optimal ship, and its hydrostatic parameters and wave-making resistance values are shown in Table 12. Considering the wave-making resistance of the optimal ship was 10.12% lower than the original, we can conclude that the optimization was effective.
Figure 17 shows how the lines plan of the optimal ship differs from the original. In the optimal ship, the bulbous bow extended forward and upward in a “wing-like” shape. The surface at stations 19 and 20 expanded outward, whereas that at station 17 became concaved. Therefore, in a KCS with a ship speed of Fr = 0.26, a slightly raised bulbous bow, slight convexing at the hull surfaces adjacent to the bulbous bow, and some degree of concaving at the surface just behind the bow can help the wave-making resistance.
Figure 18 shows the distribution of hydrodynamic pressures on the bow of the original ship and the optimal ship in Case 3. After hull-form optimization, the low-pressure region at the bulbous bow became smaller and the maximum pressure decreased. As a result, the pressure distribution at the bow became more uniform after optimization. Figure 19 shows the longitudinal wave cuts at different parts of the KCS before and after optimization. The wave amplitude decreased significantly in the first 1/2 of the optimal ship, indicating that less energy was wasted on wave-making in the optimal ship.

4.3.4. Comparative Analysis

All optimal ships from Cases 1–3 (see Table 13) showed approximately a 10% decrease in the wave-making resistance. Therefore, we can conclude that the effectiveness of hull-form optimization was roughly the same in all three cases. In terms of efficiency, Case 1 required 1000 CFD calculations, the highest of all cases. Case 2 required a total of 346 CFD calculations, considering the optimizations were performed immediately after one reduction. Multiple reductions were performed in Case 3, reducing the total number of CFD calculations to 113 (the lowest of all cases). Therefore, hull-form optimization was least efficient in Case 1 and most efficient in Case 3. Therefore, using SDSR in hull-form optimization gives the same optimization efficacy while significantly reducing the computational cost.
By comparing Cases 1–3, it was observed that the design-space reduction (Case 2) reduced computational complexity compared to direct optimization, allowing for the use of gradient-based optimization methods to directly search for the optimal solution, which significantly reduced the computational cost. However, the one-time space reduction did not provide a natural answer to the number of samples that should be considered. As a result, the samples did not adequately represent the design space, and the reduction was insignificant. Furthermore, the optimization converged to local optima. SDSR, involving multiple reductions, effectively solved all of the problems above and gave smaller optimal design spaces. Therefore, considering SDSR makes hull-form optimization much more feasible and allows for the optimal ship design to be selected without any further optimization, it is considered an excellent method for reducing the computational cost of hull-form optimization.

5. Conclusions

Hull-form optimization is an effective method for improving ship speeds and cost efficiency, making it an important tool for green ship design. With the advent of hull-surface-modification techniques, CFD numerical simulations, optimization techniques, and continuously improving computing hardware, CFD-based hull-form optimization has become very common. However, this approach exhibits significant practical flaws owing to its low computational efficiency and difficulty in identifying the global optimum. To solve these issues, this study proposed the RST-based SDSR technique that uses interval theory to compute the intersections and unions of the subdesign spaces. Furthermore, a suitable stopping criterion for hull-form optimization was proposed to facilitate the application of RST-based SDSR to hull-form optimization problems. Lastly, a KCS model was optimized using our method to minimize the wave-making resistance. The results were then compared to those of direct optimization and one-time design-space reduction, thus proving the feasibility of our method. Cases 1–3 showed approximately a 10% decrease in the wave-making resistance. In terms of efficiency, direct optimization (Cases 1) required 1000 CFD calculations. One-time design-space reduction (Case 2) required a total of 346 CFD calculations. Multiple reductions (Case 3) reduced the total number of CFD calculations to 113 (the lowest of all cases). Therefore, hull-form optimization was least efficient in Case 1 and most efficient in Case 3. Therefore, using SDSR in hull-form optimization gives the same optimization efficacy while significantly reducing the computational cost.

Author Contributions

Z.-Y.L.: Project administration, Funding acquisition, Writing—Reviewing and Editing. Q.Z.: Conceptualization, Methodology, Software, Data curation, Writing—Original draft preparation. H.-C.C.: Visualization, Investigation. B.-W.F.: Conceptualization, Software, Writing—Reviewing and Editing. X.W.: Software. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China [grant Numbers 51979211, 52271327, 52271330], Equipment research Joint Fund of Ministry of Education (Young Talents) project [8091B032201], Key Research and Development Plan of Hubei Province (2021BID008), 111 Project (BP0820028).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article (Tables and Figures).

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

To date, sequential design-space reduction was performed on the test functions based on the procedure shown in Figure 3. A uniform experimental design was used with a fixed decision threshold, while the overlap coefficient C was used to decide whether the iteration should stop or continue. The procedure is as follows:
(1)
Determine the to-be-optimized subject and its variables to determine the initial design space.
(2)
Construct the initial sample set: A uniform experimental design is used to sample the initial variable space, and the target values of the sampled points are calculated to construct the initial sample set.
(3)
Discretize the target values and variables: A suitable threshold base is chosen on the initial sample set and the target function is discretized into two values: 0 and 1. Then, the continuous condition attributes are discretized by discretizing their attribute values into attribute intervals.
(4)
Reduce the design space to obtain Sa: RST is used to perform attribute reduction and rule extraction to obtain the variable subspaces corresponding to rules with d = 0, i.e., the attractive spaces Sa.
(5)
Use the method described in Section 3.3 to calculate the overlap coefficient C: to study how C changes during the iteration, C0 is left undefined in order to perform many iterations.
(a) The subspace obtained in the previous step is used as the new sampling space, and the uniform experimental design is used to add new sampling points. The target values of these sampling points are calculated and combined with the original sample points to create a new sample set. (b) The decision attributes (according to the decision threshold) and condition attributes are discretized to obtain the discretized decision table. RST is used to perform attribute reduction and rule extraction to obtain Sai. (c) C is calculated using Sai and Sai+1. The steps above are repeated until the overlap coefficient approaches 1.
The Sai given by the last step was treated as the optimal design space. To validate the correctness of SDSR, we checked whether the optimal design space contains the global optimal solution. Furthermore, we analyzed the variations of C to select a reasonable value for the stopping criterion C0.
The 2D six-hump camel-back (SC) function and 6D Hartman function (shown below) were used as test functions.
  • Six-hump camel-back (SC) function
f ( x ) = ( 4 2.1 x 1 2 + x 1 4 3 ) x 1 2 + x 1 x 2 + ( 4 + 4 x 2 2 ) x 2 2 ,   x i [ 2 , 2 ] , i = 1 , 2
min ( f ( x ) ) = f ( 0.0898 , 0.7127 ) = f ( 0.0898 , 0.7127 ) = - 1.0316
  • Hartman function
f H a r t m a n ( X ) = i = 1 4 c i exp [ j = 1 6 a i j ( x j p i j ) 2 ] ,   0 x j 1
In this equation,
a i j = ( 10 3 17 3.5 1.7 8 0.05 10 17 0.1 8 14 3 3.5 1.7 10 17 8 17 8 0.05 10 0.1 14 ) ;   p i j = ( 0.1312 0.1696 0.5569 0.0124 0.8283 0.5886 0.2329 0.4135 0.8307 0.3736 0.1004 0.9991 0.2348 0.1415 0.3522 0.3522 0.3047 0.6650 0.4047 0.8828 0.8732 0.5743 0.1091 0.0381 )
This function contains three local optima, one being the global optimum, as shown in Equation (A4).
f ( 0.4047 , 0.8824 , 0.8462 , 0.5740 , 0.1388 , 0.0385 ) = 3.203
f ( 0.2017 , 0.1500 , 0.4769 , 0.2753 , 0.3116 , 0.6573 ) = 3.322
f ( 0.4046 , 0.8823 , 0.8537 , 0.5739 , 0.2262 , 0.0387 ) = 3.203
Tests 1 and 2 were performed on the SC function. During test 1, a relatively small number of sampling points were added during the iteration. Conversely, in test 2, a relatively large number of sampling points were added. Test 3 was performed on the 6D Hartman function.
  • Test 1
Design-space reduction was performed on the SC function. In Test 1, a relatively small number of samples were added during the iteration. Table A1 summarizes the processes and results of the test. The second column shows the number of samples that were added in each iteration: 80 sample points were selected in the first iteration (initial design space), and 2 attractive spaces were found. In the second iteration (1 round of reduction), 16 sample points were added to the 2 design spaces, resulting in 3 attractive spaces (see Figure A1). This process continued until convergence was reached.
Table A1. Results of Test 1 (SC function).
Table A1. Results of Test 1 (SC function).
IterationNumber of SamplesNumber of RulesOverlap Coefficient CReduction Efficacy
18020.04844.84%
21630.54762.65%
31530.88622.35%
41430.93012.18%
59312.18%
Figure A1. Distribution of sample points in each iteration of Test 1. (a) Initial design space, (b) 2nd iteration, (c) 3rd iteration, (d) 4th iteration, (e) 5th iteration.
Figure A1. Distribution of sample points in each iteration of Test 1. (a) Initial design space, (b) 2nd iteration, (c) 3rd iteration, (d) 4th iteration, (e) 5th iteration.
Jmse 11 01481 g0a1
Figure A2 shows the change in the overlap coefficient C during the iteration. C progressively increased up to its maximum theoretical value, i.e., 1.
Figure A2. Trend of the overlap coefficient C in Test 1.
Figure A2. Trend of the overlap coefficient C in Test 1.
Jmse 11 01481 g0a2
The optimal design spaces in Test 1 are shown in Table A2. The total volume of the optimal design spaces was only 2.18% (0.3495) compared to that of the initial design space, which significantly reduced the design space. Considering both global optima are contained by the optimal design spaces, we can conclude that SDSR can significantly and accurately reduce design spaces.
Table A2. Optimal design space in Test 1.
Table A2. Optimal design space in Test 1.
x1x2VolumeComments
Subspace 1(−0.405, 0.22985)(0.53145, 0.81)0.17684Contains an optimal solution
(−0.0898, 0.7127)
Subspace 2(−0.23375, −0.015)(−0.59845, −0.5065)0.02011-
Subspace 3(−0.015, 0.3795)(−0.89325, −0.5065)0.15257Contains an optimal solution
(0.0898, −0.7127)
Initial design space(−2, 2)(−2, 2)16Contains two optimal solutions
(0.0898, −0.7127);
(−0.0898, 0.7127)
  • Test 2
The main difference between Tests 2 and 1 is that the former has a larger number of sampling points. The aim of this test is to analyze the effect of the number of samples on the iteration. The processes and results of this test are shown in Table A3. Figure A3 shows the 2D projection of the sample points in each iteration.
Figure A3. Distribution of sample points in each iteration of Test 2. (a) Initial design space, (b) 2nd iteration, (c) 3rd iteration, (d) 4th iteration, (e) 5th iteration, (f) 6th iteration.
Figure A3. Distribution of sample points in each iteration of Test 2. (a) Initial design space, (b) 2nd iteration, (c) 3rd iteration, (d) 4th iteration, (e) 5th iteration, (f) 6th iteration.
Jmse 11 01481 g0a3
As shown in Figure A4, as the number of iterations increase, the overlap coefficient C gradually increases to 1. Considering a larger number of sampling points were added to each iteration in Test 2, the increase in C during the iteration looks “smoother” in Test 2 than in Test 1.
Table A3. Results of Test 2 (SC function).
Table A3. Results of Test 2 (SC function).
IterationNumber of SamplesNumber of RulesOverlap Coefficient CReduction Efficacy
18020.04844.84%
25650.66033.20%
34050.87752.80%
42850.96592.71%
52150.98632.67%
61750.99902.67%
Figure A4. Trend of the overlap coefficient C in Test 2.
Figure A4. Trend of the overlap coefficient C in Test 2.
Jmse 11 01481 g0a4
Five optimal design spaces were obtained at the end of Test 2 (see Table A4), whose total volume was only 2.67% (0.4270) compared to that of the original design space. Furthermore, the two optimal solutions were located in subspaces 1 and 3. Therefore, the design space was significantly and accurately reduced.
Table A4. Optimal design spaces in Test 2.
Table A4. Optimal design spaces in Test 2.
x1x2VolumeComments
Subspace 1(−0.013, 0.3795)(−0.871, −0.5065)0.14307Contains an optimal solution
(0.0898, −0.7127)
Subspace 2(−0.183, −0.013)(−0.89325, −0.5065)0.06575-
Subspace 3(−0.2405, 0.2222)(0.4445, 0.81)0.16912Contains an optimal solution
(−0.0898, 0.7127)
Subspace 4(−0.405, −0.2405)(0.5195, 0.81)0.04779-
Subspace 5(−0.19425, −0.19065)(−0.871, −0.5065)0.00131-
Initial design space(−2, 2)(−2, 2)16Contains two optimal solutions
(0.0898, −0.7127);
(−0.0898, 0.7127)
  • Test 3
In Test 3, SDSR was performed on the 6D Hartman function, and the changes in C were analyzed. The subspace volume and size of the reduced design space in each iteration are shown in Table A5. The design space was ultimately reduced to a mere 1.22% of the initial design space. The trends of the intersection set (grey), union set (yellow), and C (red) indicated that C approached 1 very quickly (Figure A5). By the 4th iteration, C was already 0.9163.
Table A5. Results of SDSR on the Hartman function.
Table A5. Results of SDSR on the Hartman function.
IterationNumber of SamplesNumber of RulesOverlap Coefficient CReduction Efficacy
16040.08988.98%
25470.22542.02%
37180.65551.33%
46880.91631.22%
Figure A5. Trend of C in Test 3.
Figure A5. Trend of C in Test 3.
Jmse 11 01481 g0a5
The eight optimal design spaces obtained by SDSR are shown in Table A6 (S1, S2, …, S8), with So being the initial design space. S1 contained one global optimum, whereas S7 contained two local optima. The total subspace volume was 0.0122, only 1.22% of the volume of the original design space.
Table A6. Optimal design spaces for the Hartman function.
Table A6. Optimal design spaces for the Hartman function.
x1x2x3x4x5x6
S1(0, 0.4153)(0, 0.4656)(0, 1)(0.0561, 0.3814)(0.2813, 0.4661)(0.5419, 1)
S2(0, 0.4153)(0, 0.194)(0, 1)(0, 0.0561)(0.2813, 0.3302)(0.5419, 1)
S3(0, 0.4153)(0.3492, 0.4656)(0, 1)(0, 0.0561)(0.2813, 0.3302)(0.5419, 1)
S4(0.1074, 0.4153)(0, 0.4656)(0, 1)(0, 0.3814)(0.1527, 0.2813)(0.5419, 0.9506)
S5(0.1074, 0.1146)(0.4656, 1)(0, 1)(0, 0.3814)(0.2813, 0.4661)(0.5419, 1)
S6(0.2077, 0.2225)(0, 0.4656)(0, 1)(0, 0.3814)(0, 0.1527)(0.1949, 0.5419)
S7(0.1615, 0.4153)(0.6112, 1)(0.2457, 1)(0.3814, 0.8428)(0, 0.4661)(0, 0.1949)
S8(0.428, 0.4788)(0.5990, 1)(0.2457, 0.7643)(0.3814, 1)(0.4661, 1)(0, 0.1949)
So(0, 1)(0, 1)(0, 1)(0, 1)(0, 1)(0, 1)

Appendix B

(1)
Input the cuts and boundaries of the ith and i + 1th iterations: the boundary values and cuts of each variable are arranged in increasing order, with the result expressed in the form of an array. For example, given a set of cuts c1 = [0, 0.046, 0.4745, 1], 0 and 1 are the boundary values while 0.046 and 0.4745 are the cuts.
(2)
Record the minimum interval set of each variable: the ”interval” command in b4m is used to convert the array into interval variables, which gives the minimum interval set of each variable, i_1, i_2, …, i_n. The number of minimum intervals is recorded as n1, n2, …, nn.
for i = 1:n_c1
i_1(i) = interval(c1(i), c1(i + 1));
end
where n_c1 is the number of intervals of variable 1, and n_c1 = length(c1-1).
(3)
Record the minimum subspace set q(1,:): The minimum subspace set is recorded by cycling through the minimum intervals (i_1, i_2, …, i_n) of each variable and expressed as a row vector q(l,:). As there are n1 × n2 × … × nn minimum subspaces:
q(l,:) = [i_1(i),i_2(j),…,i_n(t)];
(4)
Calculate the volume of the minimum subspaces vol(i): The design-space volume is equal to the sum of the interval lengths of all variables. The ”sup” and ”inf” commands in b4m are used to cyclically compute the volume of all minimum subspaces and output a n1 × n2 × … × nn-dimensional column vector.
for j = 1:n
vol(i) = vol(i) × (sup(q(i,j))-inf(q(i,j)));
end
(5)
Input the two attractive spaces Sa1 and Sa2 as row vectors. For example:
Sa1(i,:) = [interval(0,0.491), interval(0,0.414), interval(0.163,0.466), interval(0.3345,1), interval(0.2895,1)];
(6)
Adjudge whether the attractive spaces intersect with each other: The ”intersection” command in b4m is used to check for intersections between Sa1 and Sa2. If an intersection is found, the following error message is output: “The attractive spaces have an intersection, please check the inputs!”, and do not proceed to the next step.
(7)
Use the minimum subspaces to record attractive spaces Sa1 and Sa2 as t_1 and t_2. Adjudge whether the minimum subspace q(l,:) is contained in an attractive space Sa. If yes, record this minimum subspace in t. This process produces expressions for the attractive spaces in terms of minimum subspaces, t_1 and t_2. The volumes of t_1 and t_2 are calculated using the cyclic computations described in Step 3.
(8)
Calculate the intersection between Sa1 and Sa2, Ints, and its volume, S_ints: For each row vector in t_1 and t_2, if t1(i,:) = t2(j,:), then it is recorded as a row in the new matrix. This step yields the intersection between Sa1 and Sa2, Ints.
(9)
Calculate the volume of Ints:
The volume of the i-th row: S_ints0 = S_ints0 × (sup(Ints(i,k))-inf(Ints(i,k)));k = 1:n
Total volume: S_ints = Sints + S_ints0;
(10)
Calculate the union between the i-th and i + 1-th attractive spaces: S_union = vol_1 + vol_2-S_ints.
Calculate the overlap coefficient: C = S_ints/S_union.
Figure A6. Computation of subspace intersections.
Figure A6. Computation of subspace intersections.
Jmse 11 01481 g0a6

References

  1. Gammon, M.A. Optimization of fishing vessels using a Multi-Objective Genetic Algorithm. Ocean Eng. 2011, 38, 1054–1064. [Google Scholar] [CrossRef]
  2. Feng, B.; Liu, Z.; Zhan, C.; Chang, H. Application of Multi-Disciplinary Design Optimization Techniques in Ships’ Preliminary Design. Shipbuild. China 2009, 50, 109–116. [Google Scholar]
  3. Feng, B.; Liu, Z.; Zhan, C.; Chang, H. A New method of Modifying Hull Surface and Its Application in Ship Hull Form Optimization. Shipbuild. China 2013, 54, 30–39. [Google Scholar]
  4. Cheng, X.; Feng, B.; Liu, Z.; Chang, H. Hull surface modification for ship resistance performance optimization based on Delaunay triangulation. Ocean Eng. 2018, 153, 333–344. [Google Scholar] [CrossRef]
  5. Serani, A.; Fasano, G.; Liuzzi, G.; Lucidi, S.; Iemma, U.; Campana, E.F.; Stern, F.; Diez, M. Ship hydrodynamic optimization by local hybridization of deterministic derivative-free global algorithms. Appl. Ocean Res. 2016, 59, 115–128. [Google Scholar] [CrossRef] [Green Version]
  6. Yang, Z.; Pang, Y.; Wang, J.; Song, L. Application of a response surface model in multi-objective optimization for submersible shapes. J. Harbin Eng. Univ. 2011, 407–410. [Google Scholar]
  7. Kim, H.; Yang, C. A new surface modification approach for CFD-based hull form optimization. J. Hydrodyn. Ser. B 2010, 22, 520–525. [Google Scholar] [CrossRef]
  8. Zhang, S.; Zhang, B.; Tezdogan, T.; Xu, L.; Lai, Y. Computational fluid dynamics-based hull form optimization using approximation method. Eng. Appl. Comput. Fluid Mech. 2018, 12, 74–88. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, S.; Tezdogan, T.; Zhang, B.; Lin, L. Research on the hull form optimization using the surrogate models. Eng. Appl. Comput. Fluid Mech. 2021, 15, 747–761. [Google Scholar] [CrossRef]
  10. Jafaryeganeh, H.; Ventura, M.; Soares, C.G. Robust-based optimization of the hull internal layout of oil tanker. Ocean Eng. 2020, 216, 107846. [Google Scholar] [CrossRef]
  11. Feng, B.; Zhan, C.; Liu, Z.; Cheng, X.; Chang, H. Application of Basis Functions for Hull Form Surface Modification. J. Mar. Sci. Eng. 2021, 9, 1005. [Google Scholar] [CrossRef]
  12. Li, S. Research on Hull Form Design Optimization Based on SBD Technique; China Ship Research and Development Academy: Beijing, China, 2012. [Google Scholar]
  13. Wang, G.G.; Simpson, T. Fuzzy clustering based hierarchical metamodeling for design space reduction and optimization. Eng. Optim. 2004, 36, 313–335. [Google Scholar] [CrossRef]
  14. Wang, G.G.; Shan, S. Design space reduction for multi-objective optimization and robust design optimization problems. SAE Trans. 2004, 113, 37–46. [Google Scholar]
  15. Shan, S.; Wang, G.G. Space exploration and global optimization for computationally intensive design problems: A rough set based approach. Struct. Multidiscip. Optim. 2004, 28, 427–441. [Google Scholar] [CrossRef]
  16. Chu, X. Research on Design Space Exploration and Coordinated Decomposition Methods for Complex Product; Huazhong University of Science and Technology: Wuhan, China, 2010. [Google Scholar]
  17. Cheng, S.; Shi, Y.; Qin, Q. Dynamical exploitation space reduction in particle swarm optimization for solving large scale problems. In Proceedings of the IEEE Congress on Evolutionary Computation, Brisbane, Australia, 10–15 June 2012; pp. 1–8. [Google Scholar]
  18. Zhou, S.; Li, D.; Tang, G. Design space reduction based on the metamodeling and clustering method. Chin. J. Comput. Mech. 2012, 2, 242–248. [Google Scholar]
  19. Li, Y.; Liu, L.; Long, T.; Dong, W. Metamodel-based global optimization using fuzzy clustering for design space reduction. Chin. J. Mech. Eng. 2013, 26, 928–939. [Google Scholar] [CrossRef]
  20. Chi, L. Research and Applications on Design Space Reduction Methods Based on SOM and Fuzzy Clustering; Huazhong University of Science and Technology: Wuhan, China, 2013. [Google Scholar]
  21. Tseng, H.H.; Wang, S.W.; Chen, J.Y.; Liu, C.N.J. A novel design space reduction method for efficient simulation-based optimization. In Proceedings of the IEEE International Symposium on Circuits and Systems (ISCAS), Melbourne, Australia, 1–5 June 2014; pp. 381–384. [Google Scholar]
  22. Qiu, H.; Xu, Y.; Gao, L.; Li, X.; Chi, L. Multi-stage design space reduction and metamodeling optimization method based on self-organizing maps and fuzzy clustering. Expert Syst. Appl. Int. J. 2016, 46, 180–195. [Google Scholar] [CrossRef]
  23. Ye, P.; Pan, G.; Dong, Z. Ensemble of surrogate based global optimization methods using hierarchical design space reduction. Struct. Multidiscip. Optim. 2018, 58, 537–554. [Google Scholar] [CrossRef]
  24. D’Agostino, D.; Serani, A.; Diez, M. Design-space assessment and dimensionality reduction: An off-line method for shape reparameterization in simulation-based optimization. Ocean Eng. 2020, 197, 106852. [Google Scholar] [CrossRef]
  25. Pawlak, Z. Rough sets. Int. J. Comput. Inf. Sci. 1982, 11, 341–356. [Google Scholar] [CrossRef]
  26. Jones, D.R.; Perttunen, C.D.; Stuckman, B.E. Lipschitzian optimization without the Lipschitz constant. J. Optim. Theory Appl. 1993, 79, 157–181. [Google Scholar] [CrossRef]
  27. Buhmann, M. A new class of radial basis functions with compact support. Math. Comput. 2001, 70, 307–318. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Illustration of design-space reduction. (a) Initial design space, (b) Reduced design space.
Figure 1. Illustration of design-space reduction. (a) Initial design space, (b) Reduced design space.
Jmse 11 01481 g001
Figure 2. Procedure for design-space reduction.
Figure 2. Procedure for design-space reduction.
Jmse 11 01481 g002
Figure 3. Optimization by SDSR.
Figure 3. Optimization by SDSR.
Jmse 11 01481 g003
Figure 4. Convergence of attractive spaces to the optimal design space.
Figure 4. Convergence of attractive spaces to the optimal design space.
Jmse 11 01481 g004
Figure 5. A 3D model of the KCS.
Figure 5. A 3D model of the KCS.
Jmse 11 01481 g005
Figure 6. Hull surface and free-flow surface element mesh.
Figure 6. Hull surface and free-flow surface element mesh.
Jmse 11 01481 g006
Figure 7. Distribution of movable control nodes (hull form of bow from 16 to 20 section).
Figure 7. Distribution of movable control nodes (hull form of bow from 16 to 20 section).
Jmse 11 01481 g007
Figure 8. Convergence graph of the multi-island genetic algorithm during the optimization of wave-making resistance.
Figure 8. Convergence graph of the multi-island genetic algorithm during the optimization of wave-making resistance.
Jmse 11 01481 g008
Figure 9. Lines plan before (solid lines) and after (dotted lines) optimization (from 16 to 20 section).
Figure 9. Lines plan before (solid lines) and after (dotted lines) optimization (from 16 to 20 section).
Jmse 11 01481 g009
Figure 10. Comparison between dynamic pressure distributions of the original (top) and optimized (bottom) ships.
Figure 10. Comparison between dynamic pressure distributions of the original (top) and optimized (bottom) ships.
Jmse 11 01481 g010
Figure 11. Wave cuts before (solid lines) and after (dotted lines) optimization. (a) y/L = 0.1509, (b) y/L = 0.20.
Figure 11. Wave cuts before (solid lines) and after (dotted lines) optimization. (a) y/L = 0.1509, (b) y/L = 0.20.
Jmse 11 01481 g011
Figure 12. Convergence graph.
Figure 12. Convergence graph.
Jmse 11 01481 g012
Figure 13. Comparison between lines plans before (solid lines) and after (dotted lines) hull-form optimization.
Figure 13. Comparison between lines plans before (solid lines) and after (dotted lines) hull-form optimization.
Jmse 11 01481 g013
Figure 14. Comparison between dynamic pressure distributions of the original (top) and optimized (bottom) ships.
Figure 14. Comparison between dynamic pressure distributions of the original (top) and optimized (bottom) ships.
Jmse 11 01481 g014
Figure 15. Wave cuts before (solid lines) and after (dotted lines) optimization. (a) y/L = 0.1509, (b) y/L = 0.20.
Figure 15. Wave cuts before (solid lines) and after (dotted lines) optimization. (a) y/L = 0.1509, (b) y/L = 0.20.
Jmse 11 01481 g015
Figure 16. Changes in the overlap coefficient (orange line) and the effectiveness of design-space reduction (blue line) with each iteration.
Figure 16. Changes in the overlap coefficient (orange line) and the effectiveness of design-space reduction (blue line) with each iteration.
Jmse 11 01481 g016
Figure 17. Lines plans before (solid lines) and after (dotted lines) optimization.
Figure 17. Lines plans before (solid lines) and after (dotted lines) optimization.
Jmse 11 01481 g017
Figure 18. Comparison between dynamic pressure distributions of the original (top) and optimized (bottom) ships.
Figure 18. Comparison between dynamic pressure distributions of the original (top) and optimized (bottom) ships.
Jmse 11 01481 g018
Figure 19. Wave cuts before (solid lines) and after (dotted lines) optimization. (a) y/L = 0.1509, (b) y/L = 0.20.
Figure 19. Wave cuts before (solid lines) and after (dotted lines) optimization. (a) y/L = 0.1509, (b) y/L = 0.20.
Jmse 11 01481 g019
Table 1. Main dimensions and parameters of the standard KCS model.
Table 1. Main dimensions and parameters of the standard KCS model.
Length between Perpendiculars
Lpp/m
Waterline Length
Lwl/m
Breadth at Loaded Waterline
Bwl/m
Moulded Depth
D/m
Draught
T/m
Block Coefficient
Cb
Displacement   /m3
7.27867.29441.01940.60130.34180.65051.6478
Table 2. Initial range of variable values (in units of m).
Table 2. Initial range of variable values (in units of m).
VariableY2Y3Y4Y5X1Z1
Lower bound0.060000.082000.100000.24500−0.080000.17000
Initial value0.073580.094470.132990.26038−0.002560.19463
Upper bound0.085000.106000.145000.275000.000000.23500
Table 3. Variable values, hydrostatic parameters, and target values of the optimized ship.
Table 3. Variable values, hydrostatic parameters, and target values of the optimized ship.
Y2/mY3/mY4/mY5/mX1/mZ1/mLcb /kg Rw/NRt/N
Optimized ship0.08360.10570.14110.2587−0.07850.22650.5081.69711.04880.155
Original0.07360.09450.13300.2604−0.00260.19460.5091.69112.23780.476
Table 4. Decision rules.
Table 4. Decision rules.
No.Decision RuleSupport
1(Y2 = “(−Inf,0.0719)”)&(Y3 = “(0.097,Inf)”)&(Y4 = “(0.13415,Inf)”)&(X1 = “(−Inf,−0.06335)”) = >(d = 0[1])1
2(Y2 = “(−Inf,0.0719)”)&(Y3 = “(0.097,Inf)”)&(Y4 = “(−Inf,0.13415)”)&(X1 = “(−Inf,−0.06335)”) = >(d = 0[3])3
3(Y2 = “(0.0719,Inf)”)&(Y3 = “(0.097,Inf)”)&(Y4 = “(0.13415,Inf)”)&(X1 = “(−0.06335,Inf)”) = >(d = 0[3])3
4(Y2 = “(0.0719,Inf)”)&(Y3 = “(0.097,Inf)”)&(Y4 = “(0.13415,Inf)”)&(X1 = “(−Inf,−0.06335)”) = >(d = 0[1])1
5(Y2 = “(0.0719,Inf)”)&(Y3 = “(0.097,Inf)”)&(Y4 = “(−Inf,0.13415)”)&(X1 = “(−Inf,−0.06335)”) = >(d = 0[1])1
6(Y2 = “(−Inf,0.0719)”)&(Y3 = “(−Inf,0.097)”)&(Y4 = “(0.13415,Inf)”)&(X1 = “(−0.06335,Inf)”) = >(d = 1[4])4
7(Y2 = “(0.0719,Inf)”)&(Y3 = “(0.097,Inf)”)&(Y4 = “(−Inf,0.13415)”)&(X1 = “(−0.06335,Inf)”) = >(d = 1[6])6
15(Y2 = “(0.0719,Inf)”)&(Y3 = “(−Inf,0.097)”)&(Y4 = “(−Inf,0.13415)”)&(X1 = “(−0.06335,Inf)”) = >(d = 1[14])14
Table 5. Subspaces after design-space reduction.
Table 5. Subspaces after design-space reduction.
SubspaceVariableY2Y3Y4Y5X1Z1
1Lower bound0.060.0970.134150.245−0.080.170
Upper bound0.07190.1060.1450.275−0.063350.235
2Lower bound0.060.0970.10.245−0.080.170
Upper bound0.07190.1060.134150.275−0.063350.235
3Lower bound0.07190.0970.134150.245−0.063350.170
Upper bound0.0850.1060.1450.27500.235
4Lower bound0.07190.0970.134150.245−0.080.170
Upper bound0.0850.1060.1450.275−0.063350.235
5Lower bound0.07190.0970.10.245−0.080.170
Upper bound0.0850.1060.134150.275−0.063350.235
Table 6. Merged subspaces and the effectiveness of design-space reduction.
Table 6. Merged subspaces and the effectiveness of design-space reduction.
SubspaceVariableY2Y3Y4Y5X1Z1Volume
(×10−10)
Percentage
1Lower bound0.060.0970.10.245−0.080.1703.2877.80%
Upper bound0.0850.1060.1450.275−0.063350.235
2Lower bound0.07190.0970.134150.245−0.063350.1701.5803.75%
Upper bound0.0850.1060.1450.27500.235
Table 7. Two subspaces obtained from the first reduction.
Table 7. Two subspaces obtained from the first reduction.
1st ReductionVariableY2Y3Y4Y5X1Z1
Subspace 1Lower bound0.060.0970.10.245−0.080.170
Upper bound0.0850.1060.1450.275−0.063350.235
Subspace 2Lower bound0.07190.0970.134150.245−0.063350.170
Upper bound0.0850.1060.1450.27500.235
Table 8. Two subspaces obtained after the 2nd reduction.
Table 8. Two subspaces obtained after the 2nd reduction.
2nd ReductionVariableY2Y3Y4Y5X1Z1
Subspace 3Lower bound0.060.0970.12010.245−0.080.170
Upper bound0.0850.1060.1450.2675−0.063350.235
Subspace 4Lower bound0.07190.0970.134150.245−0.063350.19705
Upper bound0.0850.1060.1450.27500.235
Table 9. Three subspaces obtained from the 3rd reduction.
Table 9. Three subspaces obtained from the 3rd reduction.
3rd ReductionVariableY2Y3Y4Y5X1Z1
Subspace 5Lower bound0.060.0970.12010.245 −0.080.2229
Upper bound0.0850.1060.1450.2675−0.063350.235
Subspace 6Lower bound0.060.0970.12010.245 −0.080.170
Upper bound0.0850.1060.1450.2675−0.06750.2229
Subspace 7Lower bound0.07190.0970.134150.245−0.063350.19705
Upper bound0.0850.1060.1450.275−0.01960.235
Table 10. Three subspaces obtained from the 4th reduction.
Table 10. Three subspaces obtained from the 4th reduction.
4th ReductionVariableY2Y3Y4Y5X1Z1
Subspace 8Lower bound0.060.0970.12010.245−0.080.2296
Upper bound0.0850.1060.1450.2675−0.063350.235
Subspace 9Lower bound0.06620.0970.12010.245−0.080.2229
Upper bound0.0850.1060.1450.2675−0.063350.2296
Subspace 10Lower bound0.061250.0970.12010.245−0.080.17
Upper bound0.0850.1060.1450.2675−0.06750.2229
Subspace 11Lower bound0.07190.0970.134150.245−0.063350.19705
Upper bound0.0850.1060.1450.269−0.01960.235
Table 11. Effect of the design-space reductions and changes in the overlap coefficient with each iteration.
Table 11. Effect of the design-space reductions and changes in the overlap coefficient with each iteration.
Initial Space1st Reduction2nd Reduction3rd Reduction4th Reduction
Number of points60221813——
Number of design spaces12234
Volume × 10−1042.124.867582.286861.724681.52069
Number of good samples in the subspaces——9193148
Reduction efficacy100.00%11.56%5.43%4.09%3.61%
Overlap coefficient——0.11560.46980.75420.8817
Table 12. Optimal solution for Case 3.
Table 12. Optimal solution for Case 3.
Y2/mY3/mY4/mY5/mX1/mZ1/mLcb /kg Rw/NRt/N
Case 30.0850.1060.1450.255−0.06890.2350.5081.69710.99980.210
Original0.07360.09450.13300.2604−0.00260.19460.5091.69112.23780.476
Table 13. Comparison between the optimal solutions from Cases 1–3.
Table 13. Comparison between the optimal solutions from Cases 1–3.
Y2/mY3/mY4/mY5/mX1/mZ1/mRw/NImprovement in Wave-Making ResistanceNumber of CFD Calculations
CASE10.08360.10570.14110.2587−0.07850.226511.048−9.71%1000
CASE20.08500.10600.14500.2602−0.06330.235010.976−10.30%346
CASE30.08500.10600.14500.2550−0.06890.235010.999−10.12%113
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, Z.-Y.; Zheng, Q.; Chang, H.-C.; Feng, B.-W.; Wei, X. Sequential Design-Space Reduction and Its Application to Hull-Form Optimization. J. Mar. Sci. Eng. 2023, 11, 1481. https://doi.org/10.3390/jmse11081481

AMA Style

Liu Z-Y, Zheng Q, Chang H-C, Feng B-W, Wei X. Sequential Design-Space Reduction and Its Application to Hull-Form Optimization. Journal of Marine Science and Engineering. 2023; 11(8):1481. https://doi.org/10.3390/jmse11081481

Chicago/Turabian Style

Liu, Zu-Yuan, Qiang Zheng, Hai-Chao Chang, Bai-Wei Feng, and Xiao Wei. 2023. "Sequential Design-Space Reduction and Its Application to Hull-Form Optimization" Journal of Marine Science and Engineering 11, no. 8: 1481. https://doi.org/10.3390/jmse11081481

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