Next Article in Journal
A Wave Drift Force Model for Semi-Submersible Types of Floating Wind Turbines in Large Waves and Current
Previous Article in Journal
Transcriptome Analysis Reveals the Regulatory Mechanism of Lipid Metabolism and Oxidative Stress in Litopenaeus vannamei under Low-Salinity Stress
Previous Article in Special Issue
CFD-Powered Ship Trim Optimization: Integrating ANN for User-Friendly Software Tool Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrodynamic Analysis and Drag-Reduction Design of an Unmanned Underwater Vehicle Based on Computational Fluid Dynamics

1
China Offshore Fugro Geosolutions (Shenzhen) Co., Ltd., Shenzhen 518065, China
2
School of Mechanical Engineering and Automation, Harbin Institute of Technology, Shenzhen 518055, China
3
Guangdong Provincial Key Laboratory of Intelligent Morphing Mechanisms and Adaptive Robotics, Shenzhen 518055, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(8), 1388; https://doi.org/10.3390/jmse12081388
Submission received: 1 July 2024 / Revised: 3 August 2024 / Accepted: 12 August 2024 / Published: 13 August 2024
(This article belongs to the Special Issue CFD Applications in Ship and Offshore Hydrodynamics)

Abstract

:
In order to establish a proper geometry of an Unmanned Underwater Vehicle (UUV) for stable motion control and energy usage reduction, this paper analyzes the hydrodynamic performance of a complex shape underwater vehicle and develops a systematic Computational Fluid Dynamics (CFD) simulation method to solve the hydrodynamic parameters of the system. Based on the simulation method and their results, the streamlines and pressure distributions of the water flow around the underwater vehicle are analyzed, and the geometric model design is improved based on the drag characteristics. Also, a comparison scheme is designed to evaluate the vehicle model before and after the geometry improvement. Simulation result shows that the design schemes brings 18% drag reductions in surge direction and 32% in heave direction. Moreover, by establishing the physical and mathematical models of the UUV on a physical simulation platform, a complete model of the underwater vehicle is constructed, laying the foundation for further simulations and experiments.

1. Introduction

With the introduction of ocean energy policies and the development of related industries, marine-related industries have begun to focus on deep-sea mining and transportation of energy resources such as natural gas and oil. In addition, the installation and construction of submarine cables and other facilities that support communication between continents have also become a major hotspot [1]. However, due to the complex and variable environments of the ocean, it is difficult for divers to perform complex tasks through saturation diving, hence underwater vehicles has been widely applied to underwater exploration and operations. Among them, unmanned underwater vehicles (UUVs), which do not require divers to dive with them and can be remotely controlled from the surface, have gradually become the mainstream equipment for executing deep-sea operation projects [2]. The design and development of UUVs is a comprehensive and complex engineering task, with a high degree of technological intensity. UUVs can generally be divided into Remotely Operated Underwater Vehicles (ROVs) and Autonomous Underwater Vehicles (AUVs) [3]. ROVs can be further categorized into open-frame structures or enclosed structures. A frame with buoyancy block structure is widely adopted in most ROVs geometry designs, due to its various advantages. This structure enables the carrying and installation of large payloads, and exploration and operation equipment can be conveniently mounted on or removed from the frame-structure for further analysis and maintenance [4]. In the early stages of development, UUV system design placed greater emphasis on drag-reduction design to enhance navigation speed and energy efficiency. With the diversification of the UUVs’ application and functional requirements, loading capacity and motion stability are also be took into account. Thus, the hydrodynamic characteristics of the UUVs, especially drag-reduction performance, become an important research field [5].
For UUVs with complex geometries, such as ROVs, the structural integrity of the vehicles are important to withstand the influence from sea currents and vortices, and it is the cornerstone for ensuring the vehicle’s capability to effectively carry out tasks in deep-sea environments [6]. Accomplishing this objective entails a meticulous design of the UUVs’ geometry, along with the integration of well-engineered hydrodynamic components both on attachments and the main body, ensuring dynamic stability and drag reduction. Good geometry design not only entails maintaining robust structural stability, but also ensuring consistent hydrodynamic performance. Additionally, considerations must be given to endurance requirements, as well as the specific operational conditions and objectives [7]. Consequently, optimizing the geometry of underwater vehicles serves as the bedrock of the entire vehicle design process and represents a critical step.
To have a further knowledge of the hydrodynamic characteristic of the UUVs, the hydrodynamic model of the system should be established first. And it is fundamental to establish the dynamic mathematical models for UUVs, which play crucial roles in their maneuvering control and motion prediction. Traditional hydrodynamic model construction is based on a fundamental motion, typically uniform straight-line forward movement, to which a small perturbation is added. The multivariate function is then expanded using a Taylor series, with the fundamental motion serving as the reference point for the expansion. The coefficients obtained from the Taylor series expansion are known as hydrodynamic coefficients. Fossen, from the Norwegian University of Science and Technology, has proposed a simplified mathematical model for underwater robots [8]. This model considers the ROV to be symmetrical along three orthogonal planes: vertical, lateral, and fore/aft. In this model, the viscous hydrodynamic forces and moments acting on the underwater robot are also significantly simplified. It is assumed that the viscous hydrodynamic force (torque) in a particular direction is only related to the velocity (angular velocity) in that direction. This reduction in complexity significantly decreases the required hydrodynamic coefficients, thus easing the acquisition process. Chin et al. emphasized a systematic modeling of hydrodynamic damping using the CFD software ANSYS-CFXTM on a complex-shaped ROV [9]. Also from NTNU, Eidsvik established hydrodynamic model and conducted experiments to identify hydrodynamic parameters of several ROVs [10]. Yang et al. [11] proposed a control-oriented modeling approach for a low-speed semi-AUV which has complex-shaped structures. They also used cost-efficient CFD softwares (ANSYS-CFXTM, STAR-CCM+TM to predict the two hydrodynamic key parameters.
The common methods for obtaining hydrodynamic coefficients currently in use are experimental methods, semi-empirical formula calculations, and numerical simulation methods based on CFD technology [12]. The experimental method is simple, intuitive, easy to implement, and has high reliability. Many researchers have predicted hydrodynamic coefficients using this method [13,14]. However, the high cost of experiments and the long cycle of the experimental process lead to difficulties using physical experimental method; the semi-empirical formula calculation method is suitable for calculating the hydrodynamic parameters of underwater vehicles with simple shapes and single geometric structures. However, the shapes of AUVs and ROVs are usually more complex, making it difficult for the semi-empirical formula calculation method to accurately obtain all of the added mass and added inertia moments and viscous hydrodynamic forces in the kinematic equations of underwater vehicles. In recent years, numerical simulation methods based on potential flow theory and finite element theory, which have developed from the principles of mass conservation, momentum conservation, and energy conservation, have advanced rapidly. Many researchers used CFD method to estimate hydrodynamic coefficients of AUVs with complex shapes, such as an axisymmetric body, a body with appendages, and a spherical body designed for amphibious tasks [15,16,17,18,19]. Also, the flow status of open-frame vehicles are studied numerically, which proves the possibility of analyzing the hydrodynamic characteristics of UUVs with complex geography accurately, using CFD method [20]. The co-researchers in Harbin Institute of Technology, Shenzhen, also conduct numerical studies on hydrodynamic characteristics for AUVs and ROVs, they proposed systemic methods and reviewed the CFD method of hydrodynamic analysis [21,22,23].
The primary strategy for enhancing the efficiency of UUVs involves refining the flow dynamics around the vehicle. Vortices and turbulent flow patterns can significantly diminish the ROV’s hydrodynamic capabilities. Altering the vehicle’s geometry while preserving its fundamental stability represents the most effective approach. The advancement of computer simulation technology has facilitated comprehensive optimization analyses of underwater vehicles. The iterative design paradigm, integrating ROV analysis with CFD simulation, is extensively utilized in underwater vehicle design processes [24]. For AUVs and other torpedo-shaped underwater robots, researchers focusing on reducing the drag in surge directions to lower the energy consumption [25,26,27,28,29]. And for ROVs and other work-class underwater robots, enhancing the suitability becomes the most important goal [6,30,31]. A comprehensive review on recent trends and applications of simulation-based design optimization (SBDO) of underwater vehicles [32], shows that the studies in SBDO still remains a predominant reliance on simpler, deterministic single-objective formulations. To model and navigate the uncertainties inherent in marine engineering, including factors like wave dynamics and ocean currents more accurately, advanced optimization approaches which are more sophisticated, multi-objective, and stochastic should be applied in to SBDO.
This paper initially establishes a six-degree-of-freedom (6-DOF) dynamics and kinematics equations for a ROV. Concurrently, it employs CFD simulation methods to solve for the most challenging hydrodynamic parameters within the dynamics. A more systematic CFD simulation method for complex-shape vehicles is proposed to calculate these hydrodynamic parameters, thereby establishing the vehicle’s dynamic model. By analyzing the motion state of the vehicle in the simulation, optimization schemes for the overall vehicle model based on hydrodynamic parameters are proposed, and the optimized model parameters are determined. Finally, comparative simulation experiments of the model’s motion before and after optimization are presented to demonstrate that the optimization scheme can effectively enhance the vehicle’s motion capabilities. To facilitate the simulation of underwater operation scenarios, an underwater simulation physical platform for the vehicle has been developed based on the Robot Operating System (ROS). This platform simulates the underwater environment and interference, and integrates various sensors for perception within the simulation environment, providing a channel for algorithm validation for different types of underwater vehicles.

2. Dynamic Model Establishment

To facilitate the determination of the motion state of underwater vehicles, two commonly used system coordinate frames are introduced, namely the Inertial Coordinate System (North East Down) and the Body Fixed Frame. For general types of underwater vehicles, the position and attitude of the vehicle can be described by six different variables, as shown in Figure 1.
For an underwater vehicle, six different motion components are conveniently defined as heave, sway, surge, roll, pitch, and yaw. The symbol definitions for motion parameters are defined by their direction and orientation, where clockwise rotation and movement along the positive direction of the coordinate system are considered positive, otherwise are negative. The descriptions and symbols for the relevant motions under each degree of freedom for underwater vehicles are shown in Table 1.
For underwater vehicles, the general motion of a six-degree-of-freedom (6-DOF) underwater vehicle with the origin O as the coordinate reference is defined as follows.
The velocity of the underwater vehicle can be decomposed into linear and angular velocities. In terms of linear velocity, the vehicle facilitates coordinate transformations through a rotation matrix, as shown in the following equation:
η = η B / N N , Θ B / N N T = x , y , z , ϕ , θ , ψ T v = v B / N B , ω B / N B T = u , v , w , p , q , r T τ = f B , τ B T = X , Y , Z , K , M , N T
η B / N N = R ( z ^ , ψ ) R ( y ^ , θ ) R ( x ^ , ϕ ) v B / N B = J B / N V ( Θ B / N N ) v B / N B
In the given context, R ( z ^ , ψ ) represents the coordinate transformation matrix for the rotation of the underwater vehicle about the Z-axis, R ( y ^ , θ ) is the coordinate transformation matrix for rotation about the Y-axis, and R ( x ^ , ϕ ) denotes the coordinate transformation matrix for rotation about the X-axis. The aforementioned equation represents the linear velocity rotation matrix between the body-fixed coordinate system and the global (world) coordinate system, with the computation method as illustrated in the equation.
J B / N V ( Θ B / N N ) = c ψ c θ s ψ c ϕ + c ψ s θ s ϕ s ψ s ϕ + c ψ s θ c ϕ s ψ c θ c ψ c ϕ + s ψ s θ s ϕ c ψ s ϕ + s ψ s θ c ϕ s θ c θ s ϕ c θ c ϕ
The expressions for angular velocity and linear velocity are consistent, both derived from transformation matrices. Vectors are used to represent the kinematics, and the following equation illustrates the 6-DOF kinematic equations for the underwater vehicle:
η ˙ = J θ ( η ) v
η ˙ B / W W Θ ˙ B / W W = J B / W V 0 3 × 3 0 3 × 3 J B / W ω ( θ B / W W ) v B / W B ω B / W B
Considering the analysis object of this paper is a general underwater vehicle model, the shape of the underwater vehicle does not change under normal circumstances, and its mass and center of gravity position are fixed. The body coordinate system is established based on the center of gravity of the underwater vehicle, and the three degrees of freedom inertial axes of the underwater vehicle are used as the coordinate axes of the body coordinate system. Additionally, due to the low-speed operational state of the underwater vehicle and the deep-sea working environment, the influence of surface waves does not need to be considered. The dynamic equations of the underwater vehicle can be described using relevant mechanics formulas, as shown in the following equation.
F = M R B V ˙ + C R B ( V ) V = m I 3 × 3 0 3 × 3 0 3 × 3 I g V ˙ + m S ( V 2 ) 0 3 × 3 0 3 × 3 S ( I g V 2 ) V
In the equation, M R B represents the mass distribution of the UUV, while the Coriolis matrix, denoted as C R B , represents the Coriolis and centrifugal forces acting on the UUV, and it satisfy the relationship C R B ( V ) = C R B T ( V ) ; F represents the total force and moment acting on the underwater vehicle. In summary, by solving the mass matrix and the Coriolis and centrifugal force matrix, the general kinematic equations for the underwater vehicle can be obtained.

3. Hydrodynamic Coefficients Estimation for the Underwater Vehicle

To establish the dynamic model of the vehicle, this paper employs CFD simulations to solve for its hydrodynamic coefficients. CFD simulation requires meshing of the vehicle’s model, necessitating certain simplifications to the shape of the underwater vehicle before computation: (1) for minor structures, which have a negligible impact on drag, simplifications are made by filling in or removing them; (2) the propeller blades and external attachments, which have complex shapes that may cause issues with meshing, are simplified; and (3) for the minor recesses on the aluminum profile structure that have a minimal impact on hydrodynamics, they are optimized using solid square steel. For the load-bearing structures and other appendages that have a minor impact on drag, they are simplified by removal.
The basic numerical simulation process adopted in this paper is depicted in Figure 2. In the pre-processing stage, the underwater vehicle model is first simplified, followed by the meshing of the fluid domain. An unstructured meshing scheme is utilized to avoid meshing failures that may arise due to the irregular shape of the vehicle. In the numerical computation stage, the fluid domain of the vehicle is iteratively solved through CFD simulation. The fluid domain changes are studied frame by frame, and the simulation is iterated to the specified number of iterations to determine if the results have converged. If convergence is achieved, the process proceeds to the post-processing stage. Otherwise, the parameters in the pre-processing stage need to be readjusted. Finally, in the post-processing stage, the simulation software outputs parameters such as streamlines and pressure maps. These data are analyzed and solved using the aforementioned formulas to ultimately determine the hydrodynamic parameters and establish the model of the underwater vehicle.

3.1. Model Simplification and Fluid Domain Design

The basic model of the underwater vehicle obtained after applying the aforementioned simplification rules is shown in Figure 3. After the simplification, the model maintained the main structure of the original model, while it also improve the operability of the meshing process.
The CFD simulation defines the fluid domain as shown in Figure 4. Following the ITTC guidelines and baseline tests, by defining the length of the underwater vehicle in the direction of flow as L, the length from the entry surface to the body of the vehicle is designed as 4 L , and the length from the rear of the body to the exit surface is 8 L . The computational domain is characterized by a height and width of 8 L each, and a length of 12 L .
The surrounding boundaries of the fluid domain are set as no-slip walls, the inlet face is set as a velocity control face for fluid inflow, the outlet face is set as a pressure control face for fluid outflow, and the surface of the UUV is set as a no-slip wall in order to specify the flow velocity and state of the fluid outside the UUV model. The definitions of each boundary are shown in Figure 5.

3.2. Mesh Generation

By obtaining the model of the underwater vehicle and the fluid domain, the mesh of the model is designed and generated using mesh generation software ANSYS Meshing, the mesh structure and the transition crossing the inflation layer are shown in Figure 6. The SST k ω model is chosen as the turbulence model due to its greater accuracy by separately employing k ω and k ϵ model in the near-wall inflation zone and the developing zone of the mesh grids. According to these features of SST k ω model, the value of y + can be set to 1 to obtain a first layer of the inflation layer with lower height, which can ensure a smooth boundary layer transition with finer height changes. Additionally, controlling the surface mesh size of the ROV is essential to minimize boundary layer mesh flatness and enhance its quality. Drawing on the research of Lin et al. [33], and considering the scale of small structural components like the UUV framework, the first layer mesh size for this model is set at 0.025 m. This size ensures that the short-axis surface of the UUV framework is covered by at least five grids, maintaining calculation accuracy. It also allows for approximately 20 boundary layers, balancing computational cost and precision.
After simplifying the 3D model of the UUV, it is necessary to generate separate meshes for steady and unsteady motion simulations. The main parameters and settings for these two types of meshes are detailed in Table 2.

3.3. Simulation Settings and Parameters

After proper mesh generation, the hydrodynamic parameters of the vehicle can be calculated through numerical solvers in CFD simulation software ANSYS Fluent 2021. Once the model simplification and the computational fluid domain are established, mesh generation is performed using the built-in methods of the CFD software, and the momentum exchange approach for the simulation process is determined. The CFD simulation parameters and settings for the underwater vehicle are shown in Table 3. For steady-state simulations, the convergence residuals are set to 10 6 , but the number of iterations is limited to 2000, which can satisfy our convergence criterion that the drag coefficient output in last 100 iterations stay stable (difference less than 0.5 % ), and decrease the computation load at the same time.

3.4. Mesh Independence Study

During mesh generation, variations in parameters such as face sizing, body sizing, and grid transition smoothness can affect grid refinement, resulting in differences in the quantity of grids for the same model. It is crucial to evaluate simulation accuracy across various grid scales and ensure that adjustments in grid density within an appropriate range maintain the precision of simulation outcomes.
To verify the grid’s independence within a specific range and its impact on hydrodynamic simulation results for ROV model with different mesh scales, experiments are conducted under five different face sizing settings. These experiments are conducted under the condition of Y-direction straight-line motion at 1.0 m/s. Drag in Y direction under different grid scale at R e = 1,482,475.247 are shown in Table 4.
Analysis of the drag data reveals that increasing the face sizing of the ROV surface to 0.1 m results in the inability to generate the boundary layer of the framework properly. This occurs because the size of individual grids exceeds the width of some thin rods of the ROV frame, leading to the failure of the calculation. However, upon calculating the Y-directional drag under other mesh sizes, it is observed that the relative error of damping remains within 2.8 percent within the mesh quantity range of 2.9 to 5.0 million. This indicates that the influence of mesh quantity on the results begins to converge. The trend of Y-directional damping with the change in mesh number is depicted in Figure 7.
Therefore, balancing the accuracy and calculation load, 0.025 m was chosen as the face sizing value.

3.5. Validation of the Simulation Method

To validate the CFD method employed in the simulations in this work, A standard model called SUBOFF, which designed as a standard underwater vehicle model for testing from a past research conducted by DARPA, is simulated under same process and parameter decision method with this research. The original towing test data for the SUBOFF model can be sourced from a comprehensive series of experiments detailed in the experiment report [34]. By comparing the simulation results with these experimental data under identical operating conditions, we can ascertain the accuracy of the CFD method. Further details regarding the validation procedures can be found in our earlier publication [21]. Figure 8 shows the geometry of the SUBOFF model and also the simulation flow zone. And the errors between the simulations and physical experiments are shown in Table 5.
To further validate the simulation method with a open-framed model, we simplified the open-source BlueROV2 model as another validation model. The simplified model and the flow zone for the CFD simulation are shown in Figure 9. By conducting towing test simulations in surge direction under the same velocity gradient comparing with a experimental towing test for BlueROV2 from Li et al. [35], the drag force results and the errors between the simulations and the physical experiments are shown in Table 6.

3.6. Steady and Unsteady Motion Simulation

By employing the aforementioned CFD method, the hydrodynamic coefficients of the UUV can be obtained. These coefficients can be primarily divided into three parts, corresponding to the coefficients solved for straight-line motion in the X, Y, and Z directions, respectively. Using least squares method for the fitting, hydrodynamic coefficients can be calculated from the simulation results (force and moment). Detailed calculating process are presented in our previous work [21]. As examples, Z-direction straight-line motion and pure heave motion simulations settings are as follows.

3.6.1. Z-Direction Straight-Line Motion Simulation

In this simulation set, a total of 10 velocity conditions were simulated in the positive and negative directions for the Z-directional straight-line navigation. The experimental plan is shown in Table 7.

3.6.2. Pure Heave Motion Simulation

In this simulation set, a total of five motion frequency conditions were simulated for the pure heave motion. The experimental plan is shown in Table 8.

3.6.3. Simulation Results

The final normalized hydrodynamic coefficients of the UUV model are presented in Table 9. Based on these obtained coefficients, the complete dynamic model of the UUV can be described and imported to the physical simulation platform in Section 7.

4. Baseline Parameter Selection of Underwater Vehicle Geometry Design

To identify the geometry of the underwater vehicle body with lower drag, a CFD simulation on the basic body model of the ROV is conducted to analyzed the flow pattern and the pressure distribution around the model, as Figure 10 shows. It can be observed from the velocity streamline and the pressure distribution contour that the highest pressure mainly distributed on the upstream surface of the buoyancy block. Thus, when designing the geometry of the vehicle, emphasis should be placed on optimizing the front part of the vehicle. The design should ensure that after the optimization of the parameters, the flow distribution around the vehicle is even, separations generation is minimized, and the pressure on the upstream surface of the buoyancy block is reduced.
The principle of the ROV’s geometry design is to optimize the front part of the buoyancy block while ensuring the mid and rear body of the buoyancy block remains in same volume to provide adequate buoyancy, as shown in Figure 11. By studying existing design schemes from some of the work-class ROV buoyancy block design in the industry, the shape of the front part of the buoyancy block is altered to streamlined forms in to reduce the drag in surge direction. Two geometry models, concave and convex model were selected, as shown in Figure 12.
By conducting simulations on the baseline and optimized models of the buoyancy block, as shown in Figure 13, it can be observed that the streamlines above and below the buoyancy block are relatively more sensitive to the cross-section shape at their same side. In light of this, main geometry parameters were designed to decide the cross-section shape of the upstream side in surge direction of the buoyancy block, as Figure 14 shows.
In both cross-section shape conditions, parameter 1 is used to optimize the flow state on the upper side independently. For the concave shape design, two exclusive parameters are used for cross-section shape constraint, while for convex shape design, one exclusive parameter is used. Each parameter is assigned with a range to obtain optimized drag-reduction results, as Table 10 shows.
These parameter ranges are designed reasonably considering avoiding excessive simulation conditions and be able to cover optimal result interval.

5. Underwater Vehicle Drag-Reduction Design Simulation

The general design strategy of the buoyancy block in this research is to reduce the drag from the upstream surface by introducing chamfer and curved surface that can guide the fluid flows smoother. Here, in this section, X-direction refers to the direction of surge motion, and Z-direction to the heave direction. And the complete process of the design strategy is shown in Figure 15.
Firstly, a CFD simulation is conducted for the design parameter 1 of the buoyancy block model of the underwater vehicle to observe the optimal results of parameter 1. With the underwater vehicle’s speed around 1 knot and 0.5 m/s, the results obtained from the simulation are presented in Table 11.
Using mathematical software tools, the results are fitted and analyzed. Figure 16 shows the fitting plot of the drag in the X direction. To aid in visualizing the optimization outcomes, polynomials are employed for fitting within the figure. As parameter 1 is set to 27.5°, the drag in the X direction of the buoyancy block achieves the lowest value. This phenomenon occurs due to the inclination angle of the buoyancy block’s windward surface, which induces the formation of water streamlines along the windward surface. Consequently, this reduces the impact force and stabilizes the streamlines in the upper region, thereby mitigating drag in the X direction.
The pressure distribution contour, velocity streamline, and comparison of the drag-reduced models are shown in Figure 17. When adjusting parameter 1, the angle of approach, the optimization results are best when the angle is between 20 and 25°. If the angle is too large, it will produce undesired forces in the Z direction, leading to counter-optimization results for the underwater vehicle in both directions. The conclusions are consistent with the aforementioned fitting results. In summary, it is concluded that when performing convex shape or concave shape design, the optimal result for the resistance of the buoyancy block’s windward surface is achieved when parameter 1 is around 27.5°, within the given parameter range.
After the determination of parameter 1, further simulations is conducted combining the concave design parameter 2, which is the approach length, and also parameter 3, which is the approach radius. With the premise that the optimal result of 27.5° has already been used for parameter 1, the simulation results for concave design are presented in Table 12.
The combined fitting result for parameter 2 and 3 is shown in Figure 18. Polynomial fitting is used for demonstrating the variation trend, and the results were unexpected. The combined optimization of parameters 2 and 3 always showed an increasing trend in the drag in the X direction within the given range, which did not achieve the purpose of drag-reduction design.
The obtained pressure contour and velocity streamline are as Figure 19 shows. It can be observed that during the concave design process, the pressure at the bottom of the buoyancy block gradually increases with the optimization parameters. Additionally, due to the high-speed movement of the fluid along the surface of the vehicle, a larger separation is formed at the rear of the vehicle. Clearly, the concave design scheme has provided results in the opposite direction. The increasing separation acts on the rear of the buoyancy block, generating an oblique downward pull, which significantly increases the drag in both X and Z directions. This does not meet the purpose of this study.
In summary, under the condition of concave design, the shape optimization does not produce a positive effect. Due to the increased windward surface area, it leads to larger drag in X and Z directions. Thus, the concave design scheme should not be adopted.
For the convex design, with the premise that the optimal result of parameter 1 is 27.5°, the simulation results are as shown in Table 13.
The convex design pressure distribution contour and velocity streamline are shown in Figure 20. In this case, the pressure points on the model decrease, leading to a balanced and stable flow velocity distribution, thereby reducing the generation of separations. This results in a significant decrease in resistance for the vehicle in both the X and Z directions. However, when parameter 2 reaches a threshold value, the separations generated at the model’s tail gradually increase, leading to a reverse optimization outcome.
Figure 21 illustrates the fitting results of convex design simulations. When parameter 1 is set to 27.5°, an increase in parameter 2 leads to a quadratic distribution of resistance in the X direction, while resistance in the Z direction shows a linear trend. The growth in the Z direction is gradual, and the Z direction resistance gradually approaches zero within the parameter range. The irregular Z direction resistance generated due to water flow impact on the irregular exterior is mitigated through shape optimization. Moreover, the Z-direction resistance has minimal impact compared to the X direction. As depicted in Figure 12, convex optimization reaches its minimum value around 290 mm. Hence, for parameter 2, 290 mm is selected as the final value.
After obtaining our final simulation results, in order to verify the impact of different speeds on shape optimization and resistance, resistance experiments are conducted at different speeds for optimization parameter 1. The purpose was to confirm the independence of speed from the optimization results. The simulation results are presented in Table 14.
As Figure 22 shows, the final data were compared with the condition where the velocity is 0.5 m/s, as shown in Figure 15. It was observed that there was not a significant difference in the resistance distribution at this velocity. The overall trend remained consistent. Consequently, the conclusion was drawn that at different velocities, there are no significant factors affecting the drag-reduction of the vehicle.

6. Result of Drag-Reduction Shape Design

Firstly, comparing the concave and convex design results by analyzing the velocity streamline of the these two strategy under their optimal parameter set, as Figure 23 shows. The concave strategy increase the flow velocity below the buoyancy block, but does not give much contribution to the reduction of the separations in the lower and rear area as the convex strategy.
Following the shape optimization of the buoyancy block at the vehicle’s front end, we have obtained optimized parameters that ensure the normal operation of the underwater vehicle in its current configuration. To assess the impact of locally optimizing the buoyancy block on the overall model, the partially designed model is re-integrated into the ROV model. The resulting comparison between the final model and the original model is shown in Figure 24.
Similarly, the optimized model of the ROV is simulated in the condition of velocity equals to 0.5 m/s. The obtained pressure contour and velocity streamline are as Figure 25 shows. A comparison between before and after optimization reveals that the optimized model has great characteristic to reduce pressure on the upstream side. The number of areas with high pressure decreases, leading to a more uniform distribution. Furthermore, the flow velocity field post-optimization indicates that the presence of streamlined surfaces results in the near elimination of separations at the rear of the vehicle. Consequently, with the absence of separations behind the vehicle, resistance in both the X and Z directions decreases.
And as Table 15 shows, The simulation numerical results indicate that after drag-reduction design, the resistance in the X direction has decreased by approximately 18%, while the resistance in the Z direction has decreased by approximately 32%. The localized results have a significant overall impact, leading to a notable reduction in drag on the underwater vehicle.
After analyzing the simulation results and contours before and after the optimization design, what can be found is that the major factors that influencing the drag performance of the ROV are the streamlined cross-section shape of its main structure which has big volume proportion (e.g., buoyancy blocks), due to its open-frame design. By adopting convex shape design of the buoyancy block, the pressure at the front end are greatly reduced, and the volume of the separations at the rear end are also reduced, which decrease the drag from the tail.

7. Establishment and Validation of Underwater Physical Simulation Model

Since one of the objectives of the shape optimization of underwater vehicles is to enhance their mobility, it is essential to design a reasonable motion scheme to verify the improvement in the vehicle’s mobility before and after optimization. In this paper, a physical simulation platform is constructed for underwater vehicles, which is based on ROS and built using C++ and Python languages. It integrates Gazebo and Rviz for data visualization operations. In terms of the establishment of the vehicle model, the platform provides a sufficient number of interfaces and offers two different types of underwater vehicles for the current simulation.
The simulation platform uses Gazebo for construction. Although Gazebo itself provides very few underwater-specific modules, the development of Gazebo packages can rely on a continuously improving platform. In addition, integration with ROS, developed by OSRF, has been ensured through the Gazebo/ROS package. The simulation environment is mainly divided into the following parts: ocean environment, underwater vehicle body, thruster model, and sensors. The overall operation process design of the simulation platform is shown in Figure 26. After the underwater vehicle obtains waypoint data and the current state, the heading angle error required by the vehicle at this moment is sent to the controller through a path-tracking algorithm. At this time, the controller calculates the error to obtain the total thrust required by the vehicle. The underwater vehicle thruster model is responsible for decoupling the total thrust required by the underwater vehicle to the relevant thrusters to achieve the independent control of each thruster. The modeling of the underwater vehicle body is mainly divided into three parts: physical model establishment, physical parameter setting, and hydrodynamic parameter estimation. At this time, the thruster thrust can move the vehicle through the body model of the underwater vehicle. The sensors carried by the vehicle will record the movement state at this time and capture and send the current state to the path-tracking algorithm to complete the next cycle.
Figure 27 shows the laboratory underwater vehicle reproduced in the simulation platform, which is a fully actuated vehicle composed of seven thrusters. The control in the vertical plane is driven by three motors, and the control in the horizontal plane is driven by four motors. The simulation model is constructed based on the solved kinematic and dynamic models of the underwater vehicle.
To facilitate intuitive observation of the model establishment and drag-reduction design, the primary movement objective of the vehicle in the simulation platform is to follow a predetermined trajectory within a two-dimensional plane. The simulation mainly consists of four components: the preset trajectory, the controller, the vehicle model parameters, and the effect presentation. Five coordinate points are selected within the two-dimensional plane and connected. The vehicle starts from a given position and traverses the five coordinate points, ensuring that the control parameters remain unchanged between the original model and the final model, i.e., a consistent controller and control strategy are determined. The model parameters of the vehicle before and after the drag-reduction design have been solved, as mentioned earlier. The dynamic and kinematic models of the vehicle before and after optimization are written into the simulation, and the results are shown in Figure 28.
By examining the path tracking plot for the five given way-points, it can be observed that with all other variables remaining constant, the optimized model achieves better results with the specified control scheme. This allows the underwater vehicle to complete the set objectives more easily, with more accurate tracking of the established route. The vertical error of the vehicle from the preset trajectory is minimized at each navigation position, as shown in Figure 29. Since the vehicle needs to stably track the path, which involves converging its vertical error from the trajectory onto the vehicle’s path, the comparison of the vertical error from the optimized model before and after optimization with the preset trajectory indicates that the vehicle has acquired superior mobility capabilities after optimization. When tracking the trajectory, it can ensure that the vehicle better achieves the set objectives.

8. Conclusions

This work primarily focuses on modeling and drag-reduction design of a remotely operated underwater vehicle. By establishing the dynamic model of the vehicle, a comprehensive vehicle model that can represent the hydrodynamic characteristics of the vehicle is developed. CFD simulation methods are employed to determine the hydrodynamic parameters and analyze the fluid performance of underwater vehicle. Building upon this analysis, hypotheses concerning drag-reduction design are proposed, aiming to enhance the vehicle’s shape without compromising its overall operational effectiveness. The paper presents two distinct types of design strategy, which are evaluated against baseline parameters to derive the target shape for the underwater vehicle. Results demonstrate a reduction in resistance by 18% in the X direction and 32% in the Z direction during vehicle movement. To further validate the efficacy of the design scheme, a independence verification is conducted across different velocities. Based on the analysis of the simulation results and contours before and after the drag-reduction design, it is evident that the primary factors influencing the drag performance of the UUV are related to the streamlined cross-section shape of its main structure, particularly the buoyancy blocks, due to the open-frame design. By adopting a convex shape for the buoyancy blocks, the pressure at the front end is significantly reduced, and the volume of separations at the rear end is minimized, thereby decreasing the drag from the tail. Moreover, the paper establishes an underwater physical simulation platform to verify the mobility of the vehicle both before and after the drag-reduction design. The findings confirm that the proposed design scheme effectively enhances the mobility of underwater vehicles. Due to the limitations of the simulations, physical experiments should be conducted in the following studies to validate the effectiveness of the design strategy. Also, flow influence from waves and vortexes that applied on underwater vehicles can be tested in real-world to measure the positive impact on motion stabilization control of the vehicles.

Author Contributions

Conceptualization, X.L.; methodology, X.L. and M.Z.; software, M.Z. and Y.S.; validation, X.L. and M.Z.; formal analysis, X.L. and M.Z.; investigation, X.L.; resources, M.Z.; data curation, X.L.; writing—original draft preparation, X.L. and D.Z.; writing—review and editing, X.W. and D.Z.; visualization, X.L. and M.Z.; supervision, X.W.; project administration, X.W.; funding acquisition, X.W. and X.L. All authors have read and agreed to the published version of the manuscript.

Funding

The work is supported by the Shenzhen Science and Technology Program (JSGG20211029095205007 and JCYJ20210324115812034), the Shenzhen Science and Technology Major Program (KJZD20231023100459001) and the Guangdong Provincial Key Laboratory of Intelligent Morphing Mechanisms and Adaptive Robotics Program (2023B1212010005).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

Authors would like to give special thanks to China Offshore Fugro Geosolutions (Shenzhen) Co., Ltd. for supporting experimental resources and sites.

Conflicts of Interest

Author Xuecheng Li was employed by the company China Offshore Fugro Geosolutions (Shenzhen) Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Dalhatu, A.A.; de Azevedo, R.C.; Udebhulu, O.D.; De Tomi, G. Recent developments of remotely operated vehicle in the oil and gas industry. Holos 2021, 3, 1–18. [Google Scholar]
  2. Kabanov, A.; Kramar, V.; Ermakov, I. Design and modeling of an experimental rov with six degrees of freedom. Drones 2021, 5, 113. [Google Scholar] [CrossRef]
  3. Wu, Y.; Sima, C.; Zhu, Z. Key Development Directions of Marine Science and Technology. Sci. Technol. Foresight 2022, 1, 20–35. [Google Scholar]
  4. Shen, K.; Yan, Y.; Yan, H. Research Status and Development Trend of Deep-sea Work Class ROV in China. Control Inf. Technol. 2020, 3, 7. [Google Scholar]
  5. Zhang, K.; Yang, H.; Sun, T. Research on drag reduction structure design of ROV for high speed. Ship Sci. Technol. 2021, 43, 76–80. [Google Scholar]
  6. Purzycki, M.; Komorowska, A.; Ilnicka, A.; Papież, J.; Szymańska, E. From ROVs to AUVs–Optimization and Analysis of Underwater Vehicles Design. In Proceedings of the 2022 45th Jubilee International Convention on Information, Communication and Electronic Technology (MIPRO), Opatija, Croatia, 23–27 May 2022; pp. 1517–1521. [Google Scholar]
  7. Singh, A.P.; Paul, A.; Singh, Y.; Mukherjee, K. Mechanical design of a modular underwater rov for surveillance and cleaning purpose. In Proceedings of the Recent Advances in Mechanical Engineering: Select Proceedings of ICRAME 2020; Springer: Berlin/Heidelberg, Germany, 2021; pp. 779–793. [Google Scholar]
  8. Fossen, T.I. Handbook of Marine Craft Hydrodynamics and Motion Control; John Wiley & Sons: Hoboken, NJ, USA, 2011. [Google Scholar]
  9. Chin, C.; Lau, M. Modeling and testing of hydrodynamic damping model for a complex-shaped remotely-operated vehicle for control. J. Mar. Sci. Appl. 2012, 11, 150–163. [Google Scholar] [CrossRef]
  10. Eidsvik, O.A. Identification of Hydrodynamic Parameters for Remotely Operated Vehicles. Master’s Thesis, NTNU, Trondheim, Norway, 2015. [Google Scholar]
  11. Yang, R.; Clement, B.; Mansour, A.; Li, M.; Wu, N. Modeling of a complex-shaped underwater vehicle for robust control scheme. J. Intell. Robot. Syst. 2015, 80, 491–506. [Google Scholar] [CrossRef]
  12. Panda, J.P.; Mitra, A.; Warrior, H.V. A review on the hydrodynamic characteristics of autonomous underwater vehicles. Proc. Inst. Mech. Eng. Part M J. Eng. Marit. Environ. 2021, 235, 15–29. [Google Scholar] [CrossRef]
  13. Avila, J.P.J.; Adamowski, J.C. Experimental evaluation of the hydrodynamic coefficients of a ROV through Morison’s equation. Ocean Eng. 2011, 38, 2162–2170. [Google Scholar] [CrossRef]
  14. Zarei, A.; Ashouri, A.; Hashemi, S.; Bushehri, S.F.; Izadpanah, E.; Amini, Y. Experimental and numerical study of hydrodynamic performance of remotely operated vehicle. Ocean Eng. 2020, 212, 107612. [Google Scholar] [CrossRef]
  15. Ray, A.; Singh, S.; Seshadri, V. Evaluation of linear and nonlinear hydrodynamic coefficients of underwater vehicles using CFD. In Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering, Honolulu, HI, USA, 31 May–5 June 2009; Volume 43444, pp. 257–265. [Google Scholar]
  16. Tang, S.; Ura, T.; Nakatani, T.; Thornton, B.; Jiang, T. Estimation of the hydrodynamic coefficients of the complex-shaped autonomous underwater vehicle TUNA-SAND. J. Mar. Sci. Technol. 2009, 14, 373–386. [Google Scholar] [CrossRef]
  17. Jagadeesh, P.; Murali, K. RANS predictions of free surface effects on axisymmetric underwater body. Eng. Appl. Comput. Fluid Mech. 2010, 4, 301–313. [Google Scholar] [CrossRef]
  18. Jinxin, Z.; Yumin, S.; Lei, J.; Jian, C. Hydrodynamic performance calculation and motion simulation of an AUV with appendages. In Proceedings of the 2011 International Conference on Electronic & Mechanical Engineering and Information Technology, Harbin, China, 12–14 August 2011; Volume 2, pp. 657–660. [Google Scholar]
  19. Hou, X.; Guo, S.; Shi, L.; Xing, H.; Liu, Y.; Liu, H.; Hu, Y.; Xia, D.; Li, Z. Hydrodynamic analysis-based modeling and experimental verification of a new water-jet thruster for an amphibious spherical robot. Sensors 2019, 19, 259. [Google Scholar] [CrossRef]
  20. Katsui, T.; Kajikawa, S.; Inoue, T. Numerical Investigation of Flow around a ROV with Crawler Based Driving System. In Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering; American Society of Mechanical Engineers: New York, NY, USA, 2012; Volume 44946, pp. 23–30. [Google Scholar]
  21. Hong, L.; Wang, X.; Zhang, D.; Xu, H. Numerical study on hydrodynamic coefficient estimation of an underactuated underwater vehicle. J. Mar. Sci. Eng. 2022, 10, 1049. [Google Scholar] [CrossRef]
  22. Zhang, D.; Wang, X.; Zhao, M.; Hong, L.; Li, X. Numerical Investigation on Hydrodynamic Characteristics and Drag Influence of an Open-Frame Remotely Operated Underwater Vehicle. J. Mar. Sci. Eng. 2023, 11, 2143. [Google Scholar] [CrossRef]
  23. Hong, L.; Wang, X.; Zhang, D.S. CFD-based hydrodynamic performance investigation of autonomous underwater vehicles: A survey. Ocean. Eng. 2024, 305, 117911. [Google Scholar] [CrossRef]
  24. Baital, M.S.; Mangkusasmito, F.; Rahmawaty, M.A. Specification Design and Performances Using Computational Fluid Dynamics for Mini-Remotely Operated Underwater Vehicle. Ultim. Comput. J. Sist. Komput. 2022, 14, 20–27. [Google Scholar] [CrossRef]
  25. Wang, Y.; Gao, T.; Pang, Y.; Tang, Y. Investigation and optimization of appendage influence on the hydrodynamic performance of AUVs. J. Mar. Sci. Technol. 2019, 24, 297–305. [Google Scholar] [CrossRef]
  26. Ignacio, L.C.; Victor, R.R.; Del Rio, R.; Pascoal, A. Optimized design of an autonomous underwater vehicle, for exploration in the Caribbean Sea. Ocean Eng. 2019, 187, 106184. [Google Scholar] [CrossRef]
  27. Lin, G.; Yang, Y.; He, Z.; Jiao, P. Hydrodynamic optimization in high-acceleration underwater motions using added-mass coefficient. Ocean Eng. 2022, 263, 112274. [Google Scholar] [CrossRef]
  28. Moonesun, M.; Korol, Y.; Dalayeli, H. CFD analysis on the bare hull form of submarines for minimizing the resistance. Int. J. Marit. Technol. 2015, 3, 1–16. [Google Scholar]
  29. Luo, W.; Guo, X.; Dai, J.; Rao, T. Hull optimization of an underwater vehicle based on dynamic surrogate model. Ocean Eng. 2021, 230, 109050. [Google Scholar] [CrossRef]
  30. Lee, M.H.; Park, Y.D.; Park, H.G.; Park, W.C.; Hong, S.; Lee, K.S.; Chun, H.H. Hydrodynamic design of an underwater hull cleaning robot and its evaluation. Int. J. Nav. Archit. Ocean. Eng. 2012, 4, 335–352. [Google Scholar] [CrossRef]
  31. Lin, Y.; Guo, J.; Li, H.; Wang, Z.; Chen, Y.; Huang, H. Improvement of hydrodynamic performance of the disk-shaped autonomous underwater helicopter by local shape modification. Ocean Eng. 2022, 260, 112056. [Google Scholar] [CrossRef]
  32. Serani, A.; Scholcz, T.P.; Vanzi, V. A Scoping Review on Simulation-based Design Optimization in Marine Engineering: Trends, Best Practices, and Gaps. Arch. Comput. Methods Eng. 2024, 1–29. [Google Scholar] [CrossRef]
  33. Lin, Y.H.; Li, X.C. The investigation of a sliding mesh model for hydrodynamic analysis of a SUBOFF model in turbulent flow fields. J. Mar. Sci. Eng. 2020, 8, 744. [Google Scholar] [CrossRef]
  34. Liu, H.L.; Huang, T.T. Summary of DARPA Suboff Experimental Program Data; Technical Report; Naval Surface Warfare Center, Carderock Division: Bethesda, MD, USA, 1998. [Google Scholar]
  35. Li, Q.; Cao, Y.; Li, B.; Ingram, D.M.; Kiprakis, A. Numerical modelling and experimental testing of the hydrodynamic characteristics for an open-frame remotely operated vehicle. J. Mar. Sci. Eng. 2020, 8, 688. [Google Scholar] [CrossRef]
Figure 1. The establishment of coordinate systems for underwater vehicles.
Figure 1. The establishment of coordinate systems for underwater vehicles.
Jmse 12 01388 g001
Figure 2. Numerical simulation process.
Figure 2. Numerical simulation process.
Jmse 12 01388 g002
Figure 3. Comparison diagram of underwater vehicle model before and after simplification. (a) Basic model of underwater vehicles prior to simplification; (b) basic model of underwater vehicles post-simplification.
Figure 3. Comparison diagram of underwater vehicle model before and after simplification. (a) Basic model of underwater vehicles prior to simplification; (b) basic model of underwater vehicles post-simplification.
Jmse 12 01388 g003
Figure 4. Fluid domain of CFD computation.
Figure 4. Fluid domain of CFD computation.
Jmse 12 01388 g004
Figure 5. Boundary definition of the computational domain.
Figure 5. Boundary definition of the computational domain.
Jmse 12 01388 g005
Figure 6. Mesh details in cross-section view (a) Overall left view; (b) Inflation layer.
Figure 6. Mesh details in cross-section view (a) Overall left view; (b) Inflation layer.
Jmse 12 01388 g006
Figure 7. Trend of Y-direction drag with the change in grid number.
Figure 7. Trend of Y-direction drag with the change in grid number.
Jmse 12 01388 g007
Figure 8. SUBOFF model. (a) SUBOFF geometry; (b) simulation flow zone.
Figure 8. SUBOFF model. (a) SUBOFF geometry; (b) simulation flow zone.
Jmse 12 01388 g008
Figure 9. BlueROV2 model. (a) Simplified BlueROV2 geometry; (b) Simulation flow zone.
Figure 9. BlueROV2 model. (a) Simplified BlueROV2 geometry; (b) Simulation flow zone.
Jmse 12 01388 g009
Figure 10. Simulation results of 1 knot surge motion of the ROV basic model. (a) Velocity streamline; (b) pressure distribution.
Figure 10. Simulation results of 1 knot surge motion of the ROV basic model. (a) Velocity streamline; (b) pressure distribution.
Jmse 12 01388 g010
Figure 11. Position of the buoyancy blocks.
Figure 11. Position of the buoyancy blocks.
Jmse 12 01388 g011
Figure 12. Comparison of buoyancy block geometry of drag-reduction design. (a) Basic geometry; (b) concave design; (c) convex design.
Figure 12. Comparison of buoyancy block geometry of drag-reduction design. (a) Basic geometry; (b) concave design; (c) convex design.
Jmse 12 01388 g012
Figure 13. Velocity streamline of the basic shape.
Figure 13. Velocity streamline of the basic shape.
Jmse 12 01388 g013
Figure 14. Geometry parameters of the buoyancy block. (a) Concave shape; (b) convex shape.
Figure 14. Geometry parameters of the buoyancy block. (a) Concave shape; (b) convex shape.
Jmse 12 01388 g014
Figure 15. Flowchart of the design strategy.
Figure 15. Flowchart of the design strategy.
Jmse 12 01388 g015
Figure 16. Fitting curve of simulation results of parameter 1.
Figure 16. Fitting curve of simulation results of parameter 1.
Jmse 12 01388 g016
Figure 17. Partial pressure maps, velocity flow field diagrams, and comparison of optimization models.
Figure 17. Partial pressure maps, velocity flow field diagrams, and comparison of optimization models.
Jmse 12 01388 g017
Figure 18. Fitting curve of simulation results of concave design.
Figure 18. Fitting curve of simulation results of concave design.
Jmse 12 01388 g018
Figure 19. Pressure distribution contour, velocity streamline, and comparison of concave shape models.
Figure 19. Pressure distribution contour, velocity streamline, and comparison of concave shape models.
Jmse 12 01388 g019
Figure 20. Pressure distribution contour, velocity streamline, and comparison of convex shape models.
Figure 20. Pressure distribution contour, velocity streamline, and comparison of convex shape models.
Jmse 12 01388 g020
Figure 21. Fitting curve of simulation results of convex shape.
Figure 21. Fitting curve of simulation results of convex shape.
Jmse 12 01388 g021
Figure 22. Fitting curve of simulation results under different velocities.
Figure 22. Fitting curve of simulation results under different velocities.
Jmse 12 01388 g022
Figure 23. Velocity streamline of concave and convex optimization. (a) Concave optimization; (b) convex optimization.
Figure 23. Velocity streamline of concave and convex optimization. (a) Concave optimization; (b) convex optimization.
Jmse 12 01388 g023
Figure 24. Comparison of underwater vehicle model before and after drag-reduction design. (a) Original model; (b) final model.
Figure 24. Comparison of underwater vehicle model before and after drag-reduction design. (a) Original model; (b) final model.
Jmse 12 01388 g024
Figure 25. Pressure distribution contour and velocity streamline of the final ROV model. (a) Pressure distribution; (b) velocity streamline.
Figure 25. Pressure distribution contour and velocity streamline of the final ROV model. (a) Pressure distribution; (b) velocity streamline.
Jmse 12 01388 g025
Figure 26. Components of the simulation platform.
Figure 26. Components of the simulation platform.
Jmse 12 01388 g026
Figure 27. Establishment of the ROV model in the physical simulation platform. (a) With map; (b) without map.
Figure 27. Establishment of the ROV model in the physical simulation platform. (a) With map; (b) without map.
Jmse 12 01388 g027
Figure 28. Path following simulation of the underwater vehicle model before and after drag-reduction design. (a) Original model; (b) final model.
Figure 28. Path following simulation of the underwater vehicle model before and after drag-reduction design. (a) Original model; (b) final model.
Jmse 12 01388 g028
Figure 29. Vertical error between the model and the trajectory before and after drag-reduction design.
Figure 29. Vertical error between the model and the trajectory before and after drag-reduction design.
Jmse 12 01388 g029
Table 1. The motion definitions for the degrees of freedom of underwater vehicles.
Table 1. The motion definitions for the degrees of freedom of underwater vehicles.
DOFMotionForceVelocityCoordinate
1SurgeXux
2SwayYvy
3HeaveZwz
4RollKp ϕ
5PitchMq θ
6YawNr ψ
Table 2. Mesh parameters and settings.
Table 2. Mesh parameters and settings.
ParametersSteady MotionUnsteady Motion
Mesh typeUnstructured
Size FunctionCurvature and Proximity
Element size0.99 m
Max size1 m
Inflation layerWith inflation
Face sizing0.025 m
Body sizingNone0.3 m
GeometryFlow ZoneSub Zone
BoundaryROV Surface
Inflation optionFirst, Layer Thickness
Y plus1
Table 3. Simulation parameters of CFD.
Table 3. Simulation parameters of CFD.
ItemsSteady MotionUnsteady Motion
Numerical solution methodFVM
Turbulence modelSST k ω
Solver algorithmSIMPLEC
Discretization schemeSecond-order upwind scheme
Boundary conditionsCustom velocity inlet,
no-slip wall
Grid typeUnstructured
Number of iterations2000100/T 1
1T stands for the period of the sine motion in unsteady-state motion simulations.
Table 4. Drag in Y-directional with variation in mesh scale.
Table 4. Drag in Y-directional with variation in mesh scale.
Face Sizing (m)Mesh ScaleDrag (N)
0.0225,062,8802108.3
0.0254,493,5772105.168
0.033,511,0022095.363
0.0352,919,0022046.842
0.051,985,6621632.463
0.0751,200,3851738.659
0.1abnormal grid/
Table 5. Comparison of the numerical results with the experimental data reported in [34].
Table 5. Comparison of the numerical results with the experimental data reported in [34].
Velocity (Kn)Experimental Results of Drag-X (N)Numerical Results of Drag-X (N)Errors (%)
5.9287.490.63.63
10.00242.2238.71.45
11.84332.9330.50.72
13.92451.5447.10.97
16.00576.9582.91.04
17.99697.0716.42.82
Table 6. Comparison of the numerical results with the experimental data reported in [35].
Table 6. Comparison of the numerical results with the experimental data reported in [35].
Velocity (m/s)Experimental Results of Drag-X (N)Numerical Results of Drag-X (N)Errors (%)
0.21.1691.2093.45
0.46.2596.6215.78
0.616.16716.3661.23
0.829.16231.1426.79
1.044.70145.2482.67
Table 7. Experiment condition of Z-direction straight-line motion.
Table 7. Experiment condition of Z-direction straight-line motion.
Motion TypeMotion DirectionExperiment ConditionNumber of Simulation
Z-direction
straight motion
Z-positiverelative velocity
0.1–1.0 m/s
10
Z-negativevariation interval
0.1 m/s
10
Table 8. Experiment condition of pure heave motion.
Table 8. Experiment condition of pure heave motion.
Motion TypeExperiment ConditionNumber of Simulation
Pure heave motionrelative velocity: 0.5 m/s
amplitude: 0.02 m
frequency: 0.2–1.0 Hz
variation interval: 0.2 Hz
5
Table 9. Normalized hydrodynamic coefficients.
Table 9. Normalized hydrodynamic coefficients.
SteadyCoefficientsNormalized ValueUnsteadyCoefficientsNormalized Value
X straight X u u 0.2326heave Z w −0.4045
X u | u | 0.2087 M w −0.0142
Z u u 0.0396 Z w −0.1223
Z u | u | −0.00826 M w −0.1059
M u u 0.0260sway Y v −0.1461
M u | u | 0.00397 K v 0.018
Y straight Y v v 0.4131 N v 0.0077
Z straight Z w w −0.4740 Y v −0.4097
Z w | w | −0.0791 K v 0.0109
XOY drift X u v 0.0381 N v 0.1412
Y u v −0.416yaw Y r −0.0044
Z u v −0.0144 K r 0.0009
K u v 0.0121 N r 0.008
M u v −0.00928 Y r −0.1208
N u v 0.00625 K r 0.004
XOZ drift X u w −0.00427 N r 0.0269
X u | w | −0.00041pitch X q −0.0056
Z u w −0.501 Z q −0.0032
Z u | w | 0.0811 M q −0.0084
M u w 0.00625 X q −0.0338
M u | w | −0.00607 Z q −0.0013
Surge X u −0.1573 M q −0.0117
Table 10. Ranges of the geometry parameters.
Table 10. Ranges of the geometry parameters.
ParametersConcave ShapeConvex Shape
Parameter 10°–30°
Parameter 2100–200 mm100–332 mm
Parameter 3150 m–250 mm/
Table 11. Simulation results of parameter 1 gradient under 0.5 m/s.
Table 11. Simulation results of parameter 1 gradient under 0.5 m/s.
Parm1 (°)Drag-X (N)Drag-Z (N)
0 (Baseline)249.42290.8633
5235.10881.8906
10220.86555.1278
15203.187610.7220
20179.067931.8028
25165.046422.261
26168.031615.3322
27163.53137.929
28163.53127.9326
29166.7049−8.6101
30169.7156−16.599
Table 12. Simulation results combining parameters 2 and 3.
Table 12. Simulation results combining parameters 2 and 3.
Parm1 (°)Parm2 (mm)Parm3 (mm)Drag_X (N)Drag_Z (N)
27.50 (Baseline)0 (Baseline)163.53137.929
27.5100100176.2649118.1247
27.5100150180.1904120.8813
27.5100200187.2354126.9848
27.5100250192.7954125.1602
27.5150100171.6083111.8356
27.5150150176.0831114.4418
27.5150200182.5689116.6661
27.5150250189.7901117.7527
Table 13. Simulation results of convex design.
Table 13. Simulation results of convex design.
Parm1 (°)Parm2 (mm)Drag_X (N)Drag_Z (N)
27.50143.7221−129.0891
27.5100133.6255−108.5722
27.5150128.4418−95.4861
27.5200122.7489−81.0143
27.5250119.3638−71.6148
27.5260117.0644−67.5918
27.5270117.9888−67.8878
27.5280116.7907−64.4911
27.5290114.5748−58.5837
27.5300116.7188−61.3501
27.5310115.9669−58.5325
27.5320116.6530−56.7416
27.5332116.7736−53.8360
Table 14. Simulation results under 1 m/s.
Table 14. Simulation results under 1 m/s.
Parm1 (°)Parm2 (mm)Drag_X (N)Drag_Z (N)
5290940.01611.6734
10290883.952719.0173
15290813.020241.2795
20290717.9258128.3172
25290658.521392.1224
30290673.7852−55.7875
Table 15. Simulation resistance values before and after drag-reduction design.
Table 15. Simulation resistance values before and after drag-reduction design.
Drag_X (N)Drag_Z (N)
Pre338.013956.4946
Post276.029838.3697
Percentage18.338%32.083%
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

Li, X.; Zhang, D.; Zhao, M.; Wang, X.; Shen, Y. Hydrodynamic Analysis and Drag-Reduction Design of an Unmanned Underwater Vehicle Based on Computational Fluid Dynamics. J. Mar. Sci. Eng. 2024, 12, 1388. https://doi.org/10.3390/jmse12081388

AMA Style

Li X, Zhang D, Zhao M, Wang X, Shen Y. Hydrodynamic Analysis and Drag-Reduction Design of an Unmanned Underwater Vehicle Based on Computational Fluid Dynamics. Journal of Marine Science and Engineering. 2024; 12(8):1388. https://doi.org/10.3390/jmse12081388

Chicago/Turabian Style

Li, Xuecheng, Desheng Zhang, Ming Zhao, Xin Wang, and Yu Shen. 2024. "Hydrodynamic Analysis and Drag-Reduction Design of an Unmanned Underwater Vehicle Based on Computational Fluid Dynamics" Journal of Marine Science and Engineering 12, no. 8: 1388. https://doi.org/10.3390/jmse12081388

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