You are on page 1of 22

Journal of

Marine Science
and Engineering

Article
Numerical Study on Hydrodynamic Coefficient Estimation of
an Underactuated Underwater Vehicle
Lin Hong , Xin Wang * , Desheng Zhang and Hang Xu

School of Mechanical Engineering and Automation, Harbin Institute of Technology, Shenzhen 518055, China;
20B953023@stu.hit.edu.cn (L.H.); desheng_zhang@foxmail.com (D.Z.); 17853080111@163.com (H.X.)
* Correspondence: wangxinsz@hit.edu.cn

Abstract: Hydrodynamic coefficient estimation is crucial to the shape design, dynamic modeling,
and control of underwater vehicles. In this paper, we conduct a numerical study on the hydrodynamic
coefficient estimation of an underactuated underwater vehicle (actuated only in the surge, heave, and
yaw degrees of freedom) by adopting the computational fluid dynamics (CFD) approach. Firstly, the
mechanical structure of an underactuated underwater vehicle is briefly introduced, and the dynamic
model of the underwater vehicle with hydrodynamic effects is established. Then, steady and unsteady
Reynolds Averaged Navier–Stokes (RANS) simulations are carried out to numerically simulate the
towing test, rotating arm test, and Planar Motion Mechanism (PMM) test of the underwater vehicle
numerically. To simulate unsteady motions of the underactuated underwater vehicle, a sliding mesh
model is adopted to simulate flows in the computational fluid domain that contain multiple moving
zones and capture the unsteady interactions between the underwater vehicle and the flow field.
Finally, the estimated hydrodynamic coefficients of the underwater vehicle are validated in a physical
experiment platform, and the results show that the numerical estimates are in good agreement with
the experimental data.

Keywords: numerical study; hydrodynamic coefficient estimation; computational fluid dynamics


Citation: Hong, L.; Wang, X.; Zhang, (CFD); sliding mesh model underwater vehicles
D.; Xu, H. Numerical Study on
Hydrodynamic Coefficient
Estimation of an Underactuated
Underwater Vehicle. J. Mar. Sci. Eng. 1. Introduction
2022, 10, 1049. https://doi.org/
Underwater vehicles, including human-occupied vehicles (HOV), remotely operated
10.3390/jmse10081049
vehicles (ROV), and autonomous underwater vehicles (AUV), have been widely used in
Academic Editor: Alessandro Ridolfi military, civilian, and academic fields [1]. In recent years, towards different underwater
Received: 25 June 2022
tasks, various underwater vehicles have been developed, including open-frame underwater
Accepted: 28 July 2022
vehicles [2], revolving-body shaped underwater vehicles [3], and bio-inspired underwater
Published: 30 July 2022
vehicles [4,5]. These underwater vehicles have been applied in a wide range of tasks (e.g.,
oceanographic research [6], mine exploration [7], pipe inspection [8], and underwater search
Publisher’s Note: MDPI stays neutral
and rescue [9]). Underwater vehicles often work in hazardous, complex, and unknown
with regard to jurisdictional claims in
environments, and the maneuverability and controllability of underwater vehicles are
published maps and institutional affil-
unavoidably affected by hydrodynamic loads (i.e., hydrodynamic forces and moments). In
iations.
particular, the hydrodynamic loads, arising from water particle velocity and acceleration,
can be expressed in terms of a set of hydrodynamic coefficients corresponding to the
generic dynamic model of underwater vehicles [10]. Therefore, obtaining the accurate
Copyright: © 2022 by the authors.
values of these hydrodynamic coefficients is essential for minimum-resistance shape design,
Licensee MDPI, Basel, Switzerland. advanced control strategy development, and optimal path planning of underwater vehicles.
This article is an open access article Towards this end, many insightful investigations on the hydrodynamic performance of
distributed under the terms and underwater vehicles have been carried out in the past few decades [1], most of which can
conditions of the Creative Commons be divided into three categories: experiment-based methods, analytical and semi-empirical
Attribution (CC BY) license (https:// (ASE) methods, and numerical methods.
creativecommons.org/licenses/by/ The experiment-based methods for hydrodynamic performance investigation are
4.0/). usually implemented in some standard physical experiment platforms (e.g., towing tank,

J. Mar. Sci. Eng. 2022, 10, 1049. https://doi.org/10.3390/jmse10081049 https://www.mdpi.com/journal/jmse


J. Mar. Sci. Eng. 2022, 10, 1049 2 of 22

rotating arm mechanism, and planar-motion-mechanism (PMM)). Experiment-based meth-


ods can offer real flow dynamics passing underwater vehicles and the data collected from
these experiment platforms are entirely reliable. However, experimental difficulties and
costs associated with such experiments are very high, preventing the widespread applica-
tion of these methods. The ASE method, which relies on physical concepts and empirical
results collected over the years from typical geometries, can derive approximations of the
hydrodynamic coefficients of simple-shape underwater vehicles through various system
identification methods. However, the ASE methods cannot guarantee satisfactory perfor-
mance of hydrodynamic coefficient estimation when the underwater vehicles have complex
structures or a novel shape that was first-ever designed.
The numerical method is an efficient and practical way to obtain hydrodynamic coeffi-
cients of underwater vehicles with precisely constructed mathematical models. In recent
years, due to the widespread availability of powerful computers and the development of
computational fluid dynamics (CFD) software, many effective numerical hydrodynamic
coefficient estimation methods [11–19] have been proposed for shape design and control of
underwater vehicles with fluid dynamics efficiency. Phillips et al. [20] used ANSYS CFX to
determine the dynamic stability of a torpedo-shaped AUV. This work successfully repli-
cated the AUV’s towing tank and rotating arm tests to derive the hydrodynamic coefficients.
Takahashi et al. [21] tried to inspect the potential of the numerical method as an effective
tool for the hydrodynamic performance investigation of underwater vehicles. Based on the
well-known DARPA SUBOFF model, the hydrodynamic forces and moment exerted on
the hull in steady translation and steady turn with a drift angle were conducted following
the published methodology and procedure. Amini et al. [22] adopted the commercial CFD
software ANSYS Fluent to estimate the hydrodynamic coefficients of the SUBOFF model
through the rotating arm test simulations. Islam et al. [23] used OpenFOAM to estimate
hydrodynamic derivatives of a container ship using PMM simulation. The numerical
estimated hydrodynamic coefficients were verified by comparing the numerical estimated
hydrodynamic coefficients with the experimental data in the published literature [24].
Gao et al. [25] proposed an efficient numerical method to determine hydrodynamic coeffi-
cients in underwater vehicles’ dynamic model with only one simulation by developing a
spatial captive motion simulation method. This paper employed linear regression to derive
the unknown hydrodynamic coefficients. Since most hydrodynamic coefficient estimation
studies ignore shallow water operating environments, Mitra et al. [26] tried to investigate
the hydrodynamic performance of underwater vehicle hulls at different Reynolds numbers
over sloped channel beds by combining the experiment-based and numerical methods. To
investigate the hydrodynamic performance of an underwater towed system, Wu et al. [27]
proposed a numerical method with a new hydrodynamic model. In this work, the towing
cable in the flow field was discredited into a series of elastic catenary segments connected
by nodes. The Morrison equation was adopted to determine the hydrodynamic force acting
on the cable.
In terms of unsteady motion simulation of underwater vehicles, Sasan et al. [28]
developed a mathematical model based on 2D+T theory for simulating PMM tests of
planing hulls when boat is free in heave, pitch and roll directions. In another work [29], the
authors used a mathematical model to analyze the roll of a planing hull by coupling surge,
heave, pitch, and roll motions using 2D+T theory to study the effects of roll-induced vertical
motions on roll coefficients and response. The dynamic mesh methods are widely used
to deal with unsteady flow problems in deformable domains due to boundary movement
over time [30]. Xu et al. [31] combined various CFD techniques (e.g., overlapping grids and
sliding meshes) to investigate the maneuverability and hydrodynamic performance of a
tethered underwater vehicle. In [32], the maneuverability and hydrodynamic performance
of a tethered underwater vehicle in a uniform flow field were numerically investigated. The
numerical simulations, based on the multiple CFD techniques (overlapping grid, sliding
mesh, and finite volume method), were carried out to solve the Reynolds Averaged Navier-
–Stokes (RANS) equations. Lin et al. [33] carried out numerical simulations using a finite
J. Mar. Sci. Eng. 2022, 10, 1049 3 of 22

volume code for a hydrodynamic performance investigation of the SUBOFF model with
full appendages. In this method, unsteady RANS equations were coupled with a shear
stress transport (SST) k − ω turbulence model. To simulate unsteady motions of SUBOFF
and capture unsteady interactions between the SUBOFF model and the flow field, a sliding
mesh model was employed to simulate flows in the fluid domain containing multiple
moving zones. In [34], based on the RANS solver with SST k − ω turbulence closure, a
dynamic mesh model has been successfully applied to the PMM test simulations of an
underwater vehicle. Through a series of user-defined functions (UDFs), the dynamic mesh
model was employed to assign oscillation motions of the underwater vehicle in the yaw,
pitch, and roll modes. Javanmard et al. [35] applied RANS computations to investigate the
unsteady flow around an optimized surface-piercing propeller in various shaft inclination
angles with the CFX software. In their work, the sliding mesh technique was used to
simulate the rotational motions of the propeller.
This paper presents a numerical study on hydrodynamic coefficient estimation of an
underactuated underwater vehicle (actuated only in the surge, heave, and yaw degrees of
freedom) for effective control strategy design. Our previous research work [36] designed the
mechanical structure of an AUV and carried out a preliminary hydrodynamic performance
investigation of it. However, this work did not consider its underactuated characteristics.
The RNG k − e turbulence model we employed in [36] cannot capture the hydrodynamic
performance of underwater vehicles in the near-wall region. In addition, the dynamic
mesh method we proposed often suffers negative cell volume, for there are some small-
volume grids generated on the appendages of the underwater vehicle model. To address
the aforementioned problems and improve the accuracy of hydrodynamic coefficient
estimation method, this paper adopts a numerical method for hydrodynamic coefficient
estimation of the underactuated underwater vehicle by its steady and unsteady motion
simulations. Specifically, using Fluent on ANSYS™ Workbench software. The SST k − ω
turbulence model is employed to improve the simulation accuracy and computational
stability, and a sliding mesh method is designed to simulate the unsteady motion of the
underactuated underwater vehicle. The contributions of this work are as follows
(1) Steady motion simulations (i.e., towing test simulation and rotating arm test simu-
lation) of the underwater vehicle are carried out to obtain velocity-dependent and
rotary-dependent hydrodynamic coefficients. The SST k − ω model is employed in
our CFD method to simulate the turbulence for its good prediction capability at a
reasonable computational cost.
(2) Unsteady motion simulations (i.e., PMM test simulation) of the underwater vehicle
are carried out to obtain acceleration-dependent hydrodynamic coefficients. The
oscillation motions of the pure surge mode, the pure sway mode, the pure heave
mode, and the pure yaw mode of the underactuated underwater vehicle are simulated
by adopting a sliding mesh model and a series of UDFs.
(3) Based on benchmark models such as the standard SUBOFF model and the ellipsoid
model, extensive simulations are conducted to validate the effectiveness of our numer-
ical method of hydrodynamic coefficient estimation. In addition, based on a physical
experiment platform, the numerical estimates of our method are in good agreement
with the experimental data.
The remainder of this paper is organized as follows. The mechanical structure in-
troduction of the underactuated underwater vehicle and its dynamic model, including
hydrodynamic effects, are covered in Section 2. The numerical method designed for hy-
drodynamic coefficient estimation of the underactuated underwater vehicle is presented
in Section 3. Section 4 shows the details of towing test, rotating arm test, and PMM test
simulations for hydrodynamic coefficient estimation. In Section 5, extensive experiments
are conducted to validate the effectiveness of our proposed CFD method and the estimated
hydrodynamic coefficients. Finally, this paper concludes in Section 6.
J. Mar. Sci. Eng. 2022, 10, 1049 4 of 22

2. Mechanical Structure and Modeling


2.1. Introduction of Mechanical Structure
Figure 1a shows the overall mechanical structure of the underactuated underwater
vehicle. The underwater vehicle is torpedo type, equipped with an LED light and two
cameras. The actuated degrees of freedom (surge, heave, and yaw) are driven by four
thrusters. The overall configuration of the underwater vehicle is symmetrical about the
xoy and xoz planes. The size of the underwater vehicle is 761.5 mm × 435 mm × 220 mm
(L × W × H), and its dry weight is approximately 16.5 kg. The basic parameters of the
underwater vehicle are provided in Table 1. In this paper, the underwater vehicle is assumed
to operate at a depth that is far from the water surface.

Figure 1. (a) Mechanical structure introduction of the underactuated underwater vehicle. (b) Refer-
ence frames of the underactuated underwater vehicle. (OE -XYZ ) represents the earth-fixed reference
frame, (ob -xyz) is the body-fixed reference frame. Red color indicates the actuated degrees of freedom
(surge, heave, and yaw).

Table 1. Basic parameters of the underactuated underwater vehicle.

Parameters Value Parameters Value


Total mass 16.5 kg Submerge depth 50 m
Total gravity 162 N Operation time 3h
Total buoyancy 165 N Body length (L) 761.5 mm
Speed 0–2 knot Body width (W) 435 mm
DOF 3 Body height (H) 220 mm

2.2. Mathematical Modeling


In this paper, earth-fixed reference frame (OE -XYZ ) and body-fixed reference frame
(ob -xyz) are established to describe the motion of the underactuated underwater vehicle
(only in the surge, heave, and yaw degrees of freedom), as shown in Figure 1b. The linear
and angular velocities of the underwater vehicle are expressed in the body-fixed reference
frame, which is denoted by

u = [u, v, w] T ∈ R3 , ω = [ p, q, r ] T ∈ R3 (1)
T
where u, ω denote the linear and angular velocities, respectively. ν = u T , ω T is the


velocity vector of the underwater vehicle.

η1 = [ x, y, z] T ∈ R3 , η2 = [φ, θ, ψ] T ∈ T 3 (2)
T
where η1 , η2 denote the position and altitude, respectively. η = η1T , η2T is the position


and altitude vector of the underwater vehicle.


The origin of the body-fixed reference frame is located at the vehicle‘s center of mass,
and the vehicle’s center of buoyancy coincides with its center of gravity. The underwater
J. Mar. Sci. Eng. 2022, 10, 1049 5 of 22

vehicle is symmetric about the xoz plane and xoy plane and close to symmetric about the
yoz plane. The angles θ and φ and angular velocities p and q can be considered equal to zero
in this paper. Therefore, it can safely assumed that the motion of the underwater vehicle
in heave, roll, and pitch modes is decoupled [10]. Based on the considerations mentioned
above, the dynamic model of the underwater vehicle represented in state variables can be
expressed as follows.

ẋ = u cos ψ − v sin ψ
ẏ = u sin ψ + v cos ψ
ż = w
ψ̇ = r
1 h i
u̇ = m22 vr + Xu u + Xu|u| |u|u + τX
m11 (3)
1 h i
v̇ = −m11 ur + Yv v + Yv|v| |v|v
m22
1 h i
ẇ = (W − B) + Zw w + Zw|w| |w|w + τZ
m33
1 h i
ṙ = (m11 − m22 )uv + Nr r + Nr|r| |r |r + τN
m44
where m11 = (m − Xu̇ ), m22 = (m − Yu̇ ), m33 = (m − Zu̇ ), and m44 = (m − Nu̇ ) are the
added mass components; Xu̇ , Yu̇ , Zu̇ , and Nu̇ are the hydrodynamic coefficients in the
added mass term; W is the vehicle’s weight and B is the buoyancy force; Xu , Xu|u| , Yv ,
Yv|v| , Zw , Zw|w| , Nr , and Nr|r| are negative hydrodynamic damping coefficients; τX , τZ , and
τN are the control inputs of the underactuated underwater vehicle generated by the four
thrusters. As shown in Table 1, the gravity force of the underwater vehicle is almost equal
to its buoyancy force. Therefore, in CFD simulations, (W − B) is set to zero. It is worth
noting that the underwater vehicle moves under the influence of external disturbances. In
our work, the model uncertainty of the underwater vehicle regarding the sway direction
is considered.

3. Methodology
3.1. Governing Equations
The CFD software ANSYS Fluent is widely used for hydrodynamic coefficient estima-
tion of underwater vehicles for its accuracy and reliability. The underactuated underwater
vehicle is designed to be fully immersed in water when in operation. The Navier–Stokes
(NS) equation is adopted as the governing equation in this work to simulate the flow fields
around the underwater vehicle model. The NS equation consists of a continuity equation
and momentum conservation equation, which can be described in the Cartesian coordinates
as follows
∂ρ ∂(ρui )
+ = 0, i = 1, 2, 3 (4)
∂t ∂xi
 
∂ ( ui u j )
h  ∂u j
i
∂p 2
ρ ∂u
∂t
i
+ ∂x j = − ∂xi + ∂
∂x j v ∂ui
∂x j + ∂xi − δ ∂ui
3 ij ∂xi
  (5)
+ ∂x −ρui0 u0j i, j = 1, 2, 3

j

where ui and u j are the fluid’s velocities in i and j directions, respectively; xi is the rect-
angular coordinate of space; ρ is the density of the fluid; p is the pressure term; t is the
physical time of motion, v is the dynamic viscosity, δ is the Kronecker delta; −ρui0 u0j is the
Reynolds stress tensor based on the Boussinesq hypothesis [37].
J. Mar. Sci. Eng. 2022, 10, 1049 6 of 22

The SST k − ω model [38] is adopted as the turbulence model in our numerical
simulations. The SST k − ω employs the k − ω and k − e model in the near-wall region and
far-field, respectively. The transport equations for the SST k − ω model are expressed as
!
∂ ∂ ∂ ∂k
(ρk) + (ρkui ) = Γk + Gk − Yk (6)
∂t ∂xi ∂x j ∂x j
!
∂ ∂ ∂ ∂ω
(ρω ) + (ρωui ) = Γω + Gω − Yω + Dω (7)
∂t ∂xi ∂x j ∂x j

where Γk and Γω are the effective diffusivity of k and ω, respectively. Gk and Gω represent
the generation of k and ω, respectively. Yk and Yω are the dissipation of k and ω, respectively.
Dω is the cross-diffusion term. The working fluid in the computational domain is water-
liquid with ρ = 998.2 kg/m3 .

3.2. 3D Model Simplification


Figure 2a shows the 3D model of the underwater vehicle. The underwater vehicle is
equipped with many thrusters, sensors and connectors, causing many small appendages
and cavities (i.e., external thrusters, cameras’ window) in its 3D model, which increase
the difficulty of mesh generation and the cost of computation. In this work, to meet the
requirements of Fluent on ANSYS™ Workbench software and improve the reliability of
numerical results derived by the proposed CFD method, the 3D model of the underwater
vehicle is simplified. As shown in Figure 2b, small appendages and cavities that have little
effect on the overall hydrodynamic performance of the underwater vehicle are removed,
while its shape features are well retained. In addition, the four thrusters are simplified as
four static cylinders.

(a) (b)

Figure 2. 3D model simplification of the underwater vehicle. (a) Original 3D model of the underwater
vehicle; (b) Simplified 3D model of the underwater vehicle.

3.3. Mesh Generation


In this paper, the curvature and the proximity size functions are applied to mesh
generation for the underwater vehicle model and the fluid domain. In our CFD method,
unstructured meshes are generated to fit the model’s shape, because they can be skewed
or stretched without degrading simulation performance. Furthermore, the face-sizing
meshing method is adopted to control the mesh size on the underwater vehicle model’s
surface and the body-sizing meshing method to control the mesh size of the computational
fluid domain. Finally, the inflation function is applied to capture the flow properties of the
boundary layers around the underwater vehicle model. Specifically, as shown in Figure 3a,
prism layer meshes are generated on the surface of the underwater vehicle model by
inflation function. The first layer thickness of the prism layer is determined based on the
formulas summarized in [39] and the y+ to be 1. The value of growth rate and the number
of prism layers are determined empirically by ensuring that the volume of the grid of the
final inflation layer is similar to that of freestream cells.
J. Mar. Sci. Eng. 2022, 10, 1049 7 of 22

Figure 3. (a) Prism layers generated on the underwater vehicle model. (b) Visualization of y+ value
over the surface of the underwater vehicle model.

3.4. Grid Independence Study


An appropriate mesh density is critical to achieving a cost-effective solution since the
computational load and the accuracy of the solution are highly dependent on the number
of cells. In this work, the grid independence study is conducted based on towing test
simulation of the underwater vehicle. Three different mesh generation methods (setting
the relevance center option to coarse, medium, and fine) are adopted to compute the drag
of the underwater vehicle model. Three grids with element numbers 880,975, 1,222,996,
and 1,596,664 are generated for the underwater vehicle model and computational fluid
domain to study grid independence. The sensibility of the drag estimation to the grid
density is presented in Table 2. According to the grid independence study results, the
relevance center option with medium (1,222,996 grid cells) is considered suitable for our
CFD method to ensure the accuracy of hydrodynamic coefficient estimates at a reasonable
computational cost.

Table 2. Grid independence study of the underwater vehicle.

Mesh Density Options Number of Elements Run Time (min) Drag (N)
Coarse 880,975 31.2 17.57
Medium 1,222,996 43.1 16.91
Fine 1,596,664 57.7 16.69

3.5. Numerical Method


Figure 4 shows the overall process of the proposed numerical method for hydrody-
namic coefficient estimation. The finite volume method is employed as the solver of our
numerical method to discretize the governing equations of the fluid motion based on a
conservation equation of an integral form. The pressure–velocity coupling algorithm, a
semi-implicit method for pressure-linked equations-consistent (SIMPLEC), is used as the
pressure-based segregated solver to improve the convergence. Thus, the second-order
implicit transient formulation is used to achieve a transient flow solution. The second-order
upwind scheme is used to discretize the convective terms of momentum equations and
turbulence quantities transport equations. The second-order central difference scheme is
adopted to discretize the diffusive fluxes. The y+ is set to approximate 1 for the SST k − ω
model belonging to the low Reynolds number turbulence model, Figure 3b shows the distri-
bution of y+ value over the surface of the underwater vehicle model. The residuals of 10−7
is fixed to satisfy the convergence criteria for both continuity and momentum equations.
J. Mar. Sci. Eng. 2022, 10, 1049 8 of 22

Figure 4. Procedure of numerical study on hydrodynamic coefficient estimation of the underactuated


underwater vehicle.

4. CFD Simulation
In this paper, the velocity-dependent, rotary-dependent, and acceleration-dependent
hydrodynamic coefficients of the underwater vehicle can be numerically estimated from
two kinds of CFD simulations: steady motion simulations (i.e., towing test simulation and
rotating arm test simulation) and unsteady motion simulation (i.e., PMM test simulation).
Table 3 lists the unknown hydrodynamic coefficients in Equation (3) corresponding to
different CFD simulations.

Table 3. Hydrodynamic coefficients corresponding to different test modes and motions of the
underwater vehicles.

Test Mode Motion Hydrodynamic Coefficients


Towing test Steady straight-line motion Xu|u| , Yv|v| , Zw|w|
along x, y, z direction
Rotating arm test Steady turning motion with Nr|r|
different radius
Pure surge motion Xu̇ , Xu
PMM test Purge sway motion Yv̇ , Yv
Purge heave motion Zẇ , Zw
Pure yaw motion Nṙ , Nr

4.1. Steady Motion Simulation


To simulate the underactuated underwater vehicle’s steady motion, towing test and
rotating arm test simulations are carried out to obtain the velocity-dependent and rotary-
dependent hydrodynamic coefficients.

4.1.1. Towing Test Simulation


Towing test simulations of the underwater vehicle includes straight-line motion along
the x, y, and z-direction. Figure 5a shows the schematic of the straight-line motion of the
underwater vehicle along the x-direction. Underwater vehicles are static in the computa-
tional fluid domain, facing flows with variable velocities (−1.2 m/s to 1.2 m/s), F X is the
longitudinal force, and F Y is the lateral force. To simulate the straight-line motion in the
towing test of the underwater vehicle, the computational fluid domain is designed with
the ITTC guidelines as shown in Figure 5b. In addition, the length of the computational
fluid domain is set at 23 L (17.5 m), the width is set at 40 W (11.84 m), and the height is set
at 20 H (4.4 m).
J. Mar. Sci. Eng. 2022, 10, 1049 9 of 22

(a) (b)

Figure 5. (a) Schematic of towing test of the underwater vehicle along the x-direction. (b) Computa-
tional fluid domain designed for the towing test simulation.

The underwater vehicle model is placed at the origin of the fluid domain, 10 H below
the upper wall, 20 W to the left wall, 10 L to the inlet, and 12 L to the outlet. In this way,
the numerical results of the flow field around the underwater vehicle model can be safely
guaranteed to be free of boundary effects. Table 4 shows parameter settings of the CFD
method for steady motion simulation of the underwater vehicle. In the simulation of
the straight-line motion of the underwater vehicle along the x direction, the underwater
vehicle model is statically located in the origin of the computational fluid domain, the inlet
velocity increased from −1.2 m/s to 1.2 m/s, and the increasing step is 0.1 m/s. From the
visual point of view, the underwater vehicle model is not strictly symmetrical about the yoz
plane. Therefore, we conduct a straight-line motion simulation of the underwater vehicle
along the positive and negative x-directions to investigate their respective hydrodynamic
performance. Figure 6 compares the fluid field streamlines around the underwater vehicle
moving along positive and negative x-directions, and Table 5 shows the resistance force
variables in Reynolds. It can be seen that when the underwater vehicle moves in a straight
line with different speeds along the positive and negative x-directions, the resistance of the
underwater vehicle is almost equal. As shown in Figure 7a, the two curves representing
resistance versus Reynolds are fitted using the least-squares algorithm. Then, by extracting
the quadratic coefficients of the corresponding fitted curves, the value of hydrodynamic
coefficients (Xuu+ and Xuu− ) of the underwater vehicle can be derived.

(a) (b)

Figure 6. Comparison of flow field streamlines around the underwater vehicle model. (a) Flow field
streamlines diagram around the underwater vehicle moving in a straight line along the positive
x-direction; (b) flow field streamlines diagram around the underwater vehicle moving in a straight
line along the negative x-direction.
J. Mar. Sci. Eng. 2022, 10, 1049 10 of 22

(a) (b) (c)

Figure 7. Change of resistance with respect to Re in towing test simulation. (a) Straight-line motion
along x-direction; (b) straight-line motion along y-direction; (c) straight-line motion along z-direction.
Table 4. Parameter settings of CFD method for steady motion simulation of the underwater vehicle.

Parameters Settings
Face sizing Element size is 0.02
Turbulence model SST k − ω model
Discrete format Second-order upwind discrete scheme
Solution method SIMPLEC
Fluid type Water-liquid with density of 998.2 kg/m3
Boundary conditions Inlet velocity increased from −1.2 m/s to 1.2 m/s
Residual value 10−7

Table 5. Numerical results of towing test simulation when the underwater vehicle conducts straight-
line motion in x-direction.

Speed (m/s) Reynolds Resistance x+ (N) Resistance x− (N)


0.2 150,520 0.692 0.662
0.3 225,780 1.464 1.495
0.4 301,041 2.597 2.640
0.5 376,301 4.275 4.165
0.6 451,561 6.131 5.978
0.7 526,822 8.367 8.085
0.8 602,082 10.976 10.780
0.9 677,342 13.432 13.346
1.0 752,603 17.524 16.458
1.1 827,863 21.419 20.133
1.2 903,123 25.223 24.335

Similarly, with the same settings, the proposed numerical method is used to simulate
the straight-line motion of the underwater vehicle along the y-direction and z-direction.
Figure 7b,c shows the fitted curves of the numerical estimated data with respect to Re.
Based on the two fitted curves, the velocity-dependent hydrodynamic coefficients Yv|v| and
Zw|w| can be obtained.

4.1.2. Rotating Arm Test Simulation


The purpose of rotating arm test simulation of the underwater vehicle is to obtain the
value of rotary-dependent hydrodynamic coefficient Nrr . Figure 8a shows the schematic of
the rotating arm test, the rotational motion of the underwater vehicle is implemented with
a constant angular velocity ω and a varying radius R. Figure 8b shows the computational
fluid domain of the rotating arm test simulation, the radius of the computational domain
is R. In this paper, it is considered that a rotational pattern of the fluid motion can be
used for reproduction of the rotating arm. The underwater vehicle model is static in the
computational fluid domain, R to the upper slip wall, 1.5 L arc length to the inlet wall,
the inlet velocity is 0.5 m/s, the radius of the computational fluid domain increased from
2.20 m to 6.60 m, with an increasing step of 0.55 m. Figure 9 shows the flow field streamlines
around the underwater vehicle in the rotating arm test simulation. It can be seen that the
J. Mar. Sci. Eng. 2022, 10, 1049 11 of 22

flow field streamlines around the underwater vehicle are uniformly distributed, and the
flow velocity can be regarded as a constant of 0.5 m/s. The numerical results of the rotating
arm test simulation of the underwater vehicle are shown in Table 6. The least-square
algorithm is adopted to fit the change of moment concerning radii. As shown in Figure 10,
the slope of the fitted line is the value of hydrodynamic coefficient Nr|r| .

(a) (b)

Figure 8. (a) Schematics for the rotating arm test of the underwater vehicle. (b) Computational fluid
domain designed for the rotating arm test simulation.

Figure 9. Flow field streamlines around the underwater vehicle in rotating arm test simulation.

Figure 10. Change of moment with respect to angular velocities in rotating arm test simulation.

Table 6. Numerical results of the rotating arm test simulation of the underwater vehicle model at
various radii when the inlet velocity is 0.5 m/s.

R (m) 2.20 2.75 3.30 3.85 4.40 4.95 5.50 6.05 6.60
Moment
0.0905 0.0685 0.0702 0.0577 0.0466 0.0422 0.0417 0.0407 0.0309
(N · m)
J. Mar. Sci. Eng. 2022, 10, 1049 12 of 22

4.2. Unsteady Motion Simulation


4.2.1. Sliding Mesh Method
A computational fluid domain with multiple regions (inner zone and outer zone) is
designed to simulate the unsteady motions of the underwater vehicle model in the PMM
test, as shown in Figure 11a. Since the designed computational fluid domain needs to be
suitable for capturing the flow field characteristics of fluid–structure interaction, the size of
the computational outer zone and the underwater vehicle model’s location are the same
as that of the towing test simulation. In this way, the numerical results of the flow field
around the underwater vehicle model can be safely guaranteed to be free of boundary
effects. More specifically, the inner zone is designed as a circular cylinder to fit the shape of
the underwater vehicle, the diameter of the inner zone is 6 H, and the length of the inner
zone is 1.6 L. It is also indicated that the inlet velocity and the pressure outlet are defined
as the inlet and outlet boundary conditions at both ends of the outer zone. The boundary
of the outer zone is determined as a free sliding wall. However, since the influence of
viscous flow on the underwater vehicle model can not be ignored, the non-slide wall
boundary condition is applied to the surface of the underwater vehicle model in the PMM
test simulation. Figure 11b,c shows the mesh generation of the underwater vehicle model
and the fluid domain. This paper uses the sliding mesh method to simulate the unsteady
flow field, for it is a time-accurate solution for the moving underwater vehicle in the fluid.
In addition, the sliding mesh method can help solve the problem of negative mesh volume
of the thin mesh at the start of the inflation layer caused by using the SST k − ω model. In
the sliding mesh method, two or more cell zones are used. Each cell zone is bounded by at
least one “interface zone”, where it meets the opposing cell zone. The interface zones of
adjacent cell zones are associated with one another to form a “mesh interface”. The two cell
zones will move relative to each other along the mesh interface, as shown in Figure 11d.
During the calculation, the cell zones slide (i.e., rotate or translate) relative to one another
along the mesh interface in discrete steps. As the rotation or translation occurs, node
alignment along the mesh interface is not required. Since the flow is inherently unsteady, a
time-dependent solution procedure is required. The settings of some initial parameters of
the PMM test simulation are shown in Table 7.

Table 7. Parameter settings of CFD method for unsteady motion simulation.

Parameters Settings
Face sizing Element size is 0.02 m
Body sizing Element size is 0.04 m
Turbulence model SST k − ω model
Discrete format Second-order upwind discrete scheme
Solution method SIMPLEC
Fluid type Water-liquid, density is 998.2 kg/m3 , others are default
Calculation setup Number of time steps is 500, time step size is 1/ f ,
maximum iteration/time step is 20

Figure 11. (a) Computational fluid domain designed for the PMM test simulation; (b) meshing
generation of the underwater vehicle model for PMM test simulation; (c) closer look of the interface
zone; and (d) multi-zone fluid domain designed for sliding mesh model.
J. Mar. Sci. Eng. 2022, 10, 1049 13 of 22

4.2.2. PMM Test Simulation


PMM test simulation is one of the primary methods to simulate unsteady motions
of the underactuated underwater vehicle, which includes pure surge motion, pure sway
motion, pure heave motion, and pure yaw motion, as shown in Figure 12. In this work, a
sliding mesh model and a series of UDFs are designed to simulate these motions by the
ANSYS Fluent software. Then acceleration-dependent hydrodynamic coefficients of the
underactuated underwater vehicle can be numerically estimated. Figure 12a is the diagram
of pure surge motion. The underwater vehicle model acts the oscillatory motion with surge
velocity u along the x-direction, sway velocity v and acceleration v̇ are zero. In the pure
surge motion simulation, the motion of the underwater vehicle can be expressed as


 x = x0 sin ωt
u = u̇ = x0 ω cos ωt

(8)

 u̇ = − x0 ω 2 sin ωt
v = v̇ = 0

where x, x0 , and ω are the displacement, amplitude, and frequency of the underwater
vehicle in pure surge motion, u and u̇ represent the surge velocity and acceleration of the
underwater vehicle.

(a) (b)

(c) (d)

Figure 12. Schematics of four different unsteady motions of the underactuated underwater vehicle.
(a) Pure surge motion; (b) pure sway motion; (c) pure heave motion; and (d) pure yaw motion.

Figure 12b show the diagram of the pure sway motion of the underwater vehicle. The
pure sway motion in the PMM test simulation is that the underwater vehicle oscillates in
the y-direction and has a constant surge speed of Uc . In the pure sway motion simulation,
the motion of the underwater vehicle can be expressed as


 y = y0 sin ωt
 v = ẏ = y0 ω cos ωt


v̇ = −y0 ω 2 sin ωt (9)
u = Uc = −U0




u̇ = 0

where y, y0 and ω are the displacement, amplitude, and frequency of the underwater
vehicle in pure sway motion, v and v̇ represent the sway velocity and acceleration of the
underwater vehicle.
The pure heave motion of the underwater vehicle is similar to its pure sway motion,
except that the pure heave motion acts on the vertical plane of the body-fixed reference
frame rather than on the horizontal plane. As shown in Figure 12c, in the pure heave
motion simulation, the underwater vehicle acts in an oscillatory motion in the z-direction
J. Mar. Sci. Eng. 2022, 10, 1049 14 of 22

and has a constant surge speed U0 . In the pure heave motion simulation, the motion of the
underwater vehicle can be expressed as


 z = z0 sin ωt
 w = ż = x0 ω cos ωt


ẇ = − x0 ω 2 sin ωt (10)
u = Uc = −U0




u̇ = 0

where z, z0 , and ω are the displacement, amplitude, and frequency of the underwater
vehicle in pure heave motion, w and ẇ are the heave velocity and acceleration of the
underwater vehicle.
In pure yaw motion, both sway velocity v and sway acceleration v̇ are set to be zero,
and the speed of an underwater vehicle is always tangent to its path, as shown in Figure 12d.
In addition, the underwater vehicle model moves in a sinusoidal motion with a yaw angle
ψ, and the constant surge speed Uc can be replaced by an opposite direction inflow velocity
U0 . The rotary-dependent hydrodynamic coefficients can be derived from the pure yaw
motion simulation. In the pure yaw motion, the motion of the underwater vehicle can be
described below


 ψ = ψ0 sin ωt
r = ψ̇ = −ψ0 ω cos ωt

(11)

 ṙ = −ψ0 ω 2 sin ωt
v = v̇ = 0

where ψ, r, ṙ are the yaw angle, yaw rate, and yaw rate acceleration.
In the pure heave motion simulation, assuming that the resistance force on the under-
water vehicle model is Z, combined with the Equation (10), the hydrodynamic resistance of
the underwater vehicle model can be expressed as

Z = Zẇ ẇ + Zw w + Z0
(12)
= − aω 2 Zw sin ωt + aωZw cos ωt + Z0

In this paper, the pure heave motion simulation of the underwater vehicle is selected to
show how to estimate the values of hydrodynamic coefficients. In heave motion simulation,
the UDF defines the underwater vehicle’s oscillatory motion. The setting of other param-
eters of this simulation is shown in Table 7. Figure 13a–c shows the mesh deformation
and movement of the computational fluid domain for pure heave motion simulation of
the underwater vehicle with the motion stroke of 0.04 m during a period T0 of 20 s. It
can be seen that the inner zone containing the underwater vehicle model moves up and
down along the z-direction harmonically. Figure 13a shows the initial positions of the
underwater vehicle model, whereas Figure 13b,c show the maximum and minimum extents
of deformed meshes, respectively. The sliding mesh method we used in the PMM test
simulation works well as desired. The meshes around the inner zone move smoothly,
and no invalid meshes or negative volume cells are generated in the computational fluid
domain. Figure 13d–f shows the flow field streamlines around the underwater vehicle for
pure heave motion at t/T0 = 0.25, 0.50, 0.75. Figure 14 shows the change of resistance of
underwater vehicles at f = 0.6 Hz, and the Fourier function is adopted to fit the change of
resistance with time. From the fitted curves, the numerical results ( f = 0.2, 04, 0.6, 0.8, 1.0)
of pure heave motion simulation of the underwater vehicle can be derived, as shown in
Table 8. Based on the Table 8, the heaving resistance versus the motion parameters a × ω
and − a × ω 2 are fitted in Figure 15a,b, and the slope of the fitted lines are the estimated
hydrodynamic coefficients Zẇ and Zw . The hydrodynamic coefficients of the underactuated
underwater vehicle corresponding to the other three unsteady motions (i.e., pure surge
motion, pure yaw motion, and pure sway motion) can be numerically estimated similarly.
J. Mar. Sci. Eng. 2022, 10, 1049 15 of 22

(a) (b) (c)

(d) (e) (f)

Figure 13. Mesh deformation and movement for pure heave motion of the underwater vehicle at
(a) t/T0 = 0.25; (b) t/T0 = 0.50; and (c) t/T0 = 0.75. Flow field streamlines around the underwater
vehicle for pure heave motion at (d) t/T0 = 0.25; (e) t/T0 = 0.50; and (f) t/T0 = 0.75.

Figure 14. Change of heaving force of underwater vehicle at f = 0.6 Hz and its Fourier fitting.

(a) (b)

Figure 15. Change of heaving resistance Z with respect to motion parameters a and ω in the PMM
test simulation. (a) Change of resistance Z with respect to a · w; and (b) change of resistance Z with
respect to − a · w2 .
J. Mar. Sci. Eng. 2022, 10, 1049 16 of 22

Table 8. Numerical results of pure heave motion of the underwater vehicle by our CFD method.

f (Hz) w (1/s) − a · w2 a·w


0.2 1.256 −0.06317 0.05027
0.4 2.513 −0.25266 0.10053
0.6 3.769 −0.61685 0.15708
0.8 5.026 −1.01065 0.20106
1 6.283 −1.57914 0.25133

Through the towing test simulations, rotating arm test simulations, and PMM test
simulations, the unknown hydrodynamic coefficients in the Equation (3) can be numerically
estimated, as listed in Table 9. To present the main particulars of the underwater vehicle, the
estimated hydrodynamic coefficients in a non-dimensional form are also provided. The hy-
drodynamic forces and moments are normalized by 1/2ρU 2 L2 and 1/2ρU 2 L3 , respectively.
(ρ is the fluid density, U is the freestream velocity, L is the length of the underwater vehicle.)

Table 9. Numerical results of hydrodynamic coefficients estimation by towing test simulation, rotating
arm test simulation, and PMM test simulation. The corresponding non-dimensional hydrodynamic
coefficients are marked with 0.

Coefficients Value Coefficients Value Coefficients Value Coefficients Value


Xu −20.40 Xu0 −0.1407 Zẇ −26.39 Zẇ0 −0.1196
Yv −13.51 Yv0 −0.0932 Nṙ −0.48 Nṙ0 −0.0037
Zw −42.88 0
Zw −0.2958 Xu | u | −18.20 Xu0 |u| −0.0606
Nr −1.55 Nr0 −0.0184 Yv|v| −34.25 Yv0 |v| −0.1181
Xu̇ −6.48 Xu̇0 −0.0294 Zw|w| −87.33 0
Zw |w| −0.3012
Yv̇ −19.27 Yu̇0 −0.0873 Nr|r| −8.24 Nr0|r| −0.0644

5. Experiments and Analysis


5.1. Validation of the Proposed Numerical Method
5.1.1. Validation of the Towing Test Simulations
In this study, the SUBOFF model is used as the standard model to verify our towing
test simulation, because the experimental data of the SUBOFF towing test can be obtained
in [24]. The experimental data can be used as a benchmark to verify the effectiveness
of towing test simulation method. The resistance estimated by our method is compared
to the reported experimental data. As shown in Figure 16a,b, the settings of towing test
simulation of the SUBOFF model in this paper is consistent with that of the standard
experiment reported in [24]. From Table 10, it can be seen that the errors range from 0.72%
to 3.63%, which demonstrates that the hydrodynamic coefficients obtained by our towing
test simulation are in good agreement with the experimental data, and the numerical results
derived from our towing test simulation are reliable.

Figure 16. (a) Standard SUBOFF model; (b) computational fluid domain for towing test simula-
tion of SUBOFF model; and (c) computational fluid domain for rotating arm test simulation of
SUBOFF model.
J. Mar. Sci. Eng. 2022, 10, 1049 17 of 22

Table 10. Validation of our CFD method by comparing the numerical results with the experimental
data reported in [24].

Experimental Results
Speed (Knot) Our Results (N) Error (%)
(N)
5.92 87.4 90.6 3.63
10.00 242.2 238.7 1.45
11.84 332.9 330.5 0.72
13.92 451.5 447.1 0.97
16.00 576.9 582.9 1.04
17.99 697.0 716.4 2.82

5.1.2. Validation of the Rotating Arm Test Simulations


To validate the effectiveness of the rotating arm test simulations, our CFD method
added is applied to the SUBOFF model. The numerical results of the estimated hydro-
dynamic coefficients obtained by rotating arm test simulation are compared with the
published experimental data [40]. The rotating arm test simulation settings are similar
to the standard experiment reported in [41]. As shown in Figure 16c, the radius of the
computational fluid domain is set to 20.32 m, and the drift angle ranges from 0 °C to
16 °C. The comparison between the numerical results and the reported experimental data
is shown in Figure 17. The rotating arm test simulations can provide reliable performance
in hydrodynamic coefficient estimation.

(a) (b) (c)

Figure 17. Validation of rotating arm test simulation. (a) Comparison of the numerical result X 0
with the reported experimental data; (b) comparison of the numerical result Y 0 with the reported
experimental data; and (c) comparison of the numerical result N 0 with the reported experimental data.

5.1.3. Validation of the PMM Test Simulations


To validate the effectiveness of the PMM test simulations, we apply our CFD method
to the standard ellipsoid. In addition, the numerical results of the estimated hydrody-
namic coefficients obtained by the PMM test simulation are compared with the published
experimental data. The pure heave motion of the standard ellipsoid is used to validate
the effectiveness of the PMM test simulations. The motion of the underwater vehicle is
simulated by dynamic mesh and UDFs: the oscillation amplitude of the standard ellipsoid
is 0.04 m, the inlet velocity of the flow field is 0.5 m/s, and the oscillation frequency of
the pure heave motion of the standard ellipsoid is 0.2, 0.4, 0.6, 0.8, and 1.0, respectively.
Table 11 shows numerical results of pure heave motion of the standard ellipsoid by our
CFD method. Table 12 shows the results of our simulation experiments are compared
with the reported data in [42]. It can be seen that the hydrodynamic coefficients obtained
by our CFD method are in good agreement with the reported experimental data, which
demonstrates the reliability of our PMM test simulations.
J. Mar. Sci. Eng. 2022, 10, 1049 18 of 22

Table 11. Numerical results of pure heave motion of the standard ellipsoid by our CFD method.

Za Zb
f (Hz) w (1/s) − aLw2 /V 2 aw/V 0.5ρL2 V 2 0.5ρL2 V 2

0.2 1.256 −0.37899 0.10053 −0.0014 0.0096


0.4 2.513 −1.51597 0.20106 −0.0031 0.0398
0.6 3.769 −3.41094 0.30159 −0.0064 0.0983
0.8 5.026 −6.06388 0.40212 −0.0099 0.1617
1 6.283 −9.47482 0.50265 −0.0151 0.2528

Table 12. Validation of PMM test simulation by comparing the numerical solution with the reported
data [42].

Hydrodynamic Coefficient 0
Zẇ 0
Zw 0
Mẇ 0
Mw
Numerical results −0.0268 −0.0257 7 × 10−5 0.0242
Experimental data −0.0268 −0.0186 0 0.0186
Error (%) 0 38.2 0 30.1

5.2. Towing Tank Test


To validate the estimated velocity-dependent hydrodynamic coefficients of the un-
derwater vehicle, the towing tank test of the SUBOFF model is conducted in a physical
experiment platform. Figure 18 shows the design of the physical experiment platform.
It is a tank pool with a length of 4.12 m, a width of 2.01 m, and a height of 1.22 m. The
synchronized linear module, consisting of a synchronized belt, a servo motor, and a guide
rail, is used to drive the slider to move on the guide rail with a minimum speed of 0.003 m/s
and a maximum speed of 6 m/s. The underwater vehicle is connected to the slider through
a sting so that the SUBOFF can get the same surge speed as the slider. In the towing test
experiment, the load cell used is the Optoforce six component load cell. The load cell can
measure the force (moment) in (around) the x-axis and y-axis, and the z-axis.

Figure 18. Introduction of the physical experiment platform.

5.2.1. Experiment Platform Calibration


To calibrate the established experiment platform, towing test experiments of the
standard SUBOFF model in the physical experimental platform are carried out, as shown in
Figure 19. The SUBOFF model is connected to the slider through the sting connector. The
surge speed of the SUBOFF model is increased from 0.6 m/s to 1.3 m/s with an increasing
step of 0.09 m/s. The experimental results are compared with the published data reported
in [24], as shown in Figure 20. It can be seen that the experimental data collected from
the physical experiment platform are in good agreement with the reported data, which
demonstrates the effectiveness of the physical experiment platform.
J. Mar. Sci. Eng. 2022, 10, 1049 19 of 22

(a) (b)

Figure 19. (a) 3D SUBOFF model. (b) Experiment platform calibration based on the 1/8 SUBOFF model.

Figure 20. Comparison between physical experimental results and numerical results of hydrodynamic
coefficient estimation of the 1/8 SUBOFF model.

5.2.2. Experiments and Analysis


The physical experiment platform is used to validate the estimated velocity-dependent
hydrodynamic coefficients by comparing the numerical results derived from our CFD
method with the experimental results. Figure 21a,b are the experimental settings of the un-
derwater vehicle’s straight-line motion along x-direction and y-direction. The surge speed
of the underwater vehicle’s straight-line motion is increased from 0.09 m/s to 0.51 m/s,
with an increasing step of 0.03 m/s. The load cell is used to collect the hydrodynamic resis-
tance of the underwater vehicle. Figure 22a,b shows the resistance versus surge speed along
positive and negative x-direction, respectively. The sway speed of straight-line motion in
the experiment is increased from 0.09 m/s to 0.3 m/s with an increasing step of 0.03 m/s.
The collected experimental data (resistance) versus speed is fitted in Figure 22c. It can
be seen that the velocity-dependent hydrodynamic coefficients obtained by the proposed
numerical method are not strictly consistent with the experimental data from the physi-
cal experiment platform. However, only a constant exists between the numerical results
and the experimental data, which demonstrates that the established physical experiment
platform can be used to get the hydrodynamic coefficients of the underwater vehicle.

(a) (b)

Figure 21. Straight-line motion test of the underwater vehicle on the physical experiment platform.
(a) Straight-line motion along x-direction; and (b) straight-line motion along y-direction.
J. Mar. Sci. Eng. 2022, 10, 1049 20 of 22

(a) (b) (c)

Figure 22. Resistance comparison of the underwater vehicle obtained by the CFD simulations and
the physical experiment platform. (a) Resistance comparison along positive x-direction; (b) resistance
comparison along negative x-direction; and (c) resistance comparison along y-direction.

6. Conclusions
This work adopts the CFD method to estimate essential hydrodynamic coefficients
of an underactuated underwater vehicle through the towing test simulation, rotating arm
test simulation, and PMM test simulation. Based on the underactuated characteristics of
the underwater vehicle, the dynamic model of the underwater vehicle is established while
considering the hydrodynamic effects. Then, the unknown hydrodynamic coefficients in
the underwater vehicle’s dynamic model are estimated by towing test simulations, rotating
arm test simulations, and PMM test simulations in the underwater environment with
ANSYS Fluent software. To simulate unsteady motions of the underwater vehicle and
capture unsteady interactions between the underwater vehicle and the flow field, the sliding
mesh model is employed to simulate flows in the computational fluid domain consisting
of multi-zones. The SST k − ω model is employed to simulate the turbulence for it has
excellent prediction capability at a reasonable computational cost. Extensive experiments
are conducted to verify the effectiveness of the estimated hydrodynamic coefficients based
on the SUBOFF model and the standard ellipsoid. The results of comparative experiments
demonstrate that estimated hydrodynamic coefficients of our method are close to ground-
truth values. In addition, based on a physical experiment platform, the effectiveness of
the estimated hydrodynamic coefficients of the underwater vehicle is verified. In the
future, our method will be extended to the shape design of underwater vehicles to reduce
hydrodynamic resistance on the dynamic performance of underwater vehicles and use the
numerical estimated hydrodynamic coefficients to advanced control strategy design and
optimal path planning of underwater vehicles.

Author Contributions: Conceptualization, L.H.; methodology, L.H. and D.Z.; software, L.H., D.Z.
and H.X. ; validation, L.H., D.Z.; formal analysis, L.H., D.Z. and H.X.; investigation, L.H.; resources,
X.W.; data curation, L.H.; writing—original draft preparation, L.H.; writing—review and editing,
X.W.; visualization, L.H., D.Z. and H.X.; supervision, X.W.; project administration, X.W.; funding
acquisition, X.W. All authors have read and agreed to the published version of the manuscript.
Funding: This work was supported by the Joint Funds of the National Natural Science Foundation of
China (No.U1913206) and Shenzhen Science and Technology Program (No. JSGG20211029095205007).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.
J. Mar. Sci. Eng. 2022, 10, 1049 21 of 22

References
1. 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. [CrossRef]
2. Jain, A.; Chandra, N.R.; Kumar, M. Design and Development of an Open-frame AUV: ANAHITA. In Proceedings of the 2018
IEEE/OES Autonomous Underwater Vehicle Workshop (AUV), Porto, Portugal, 6–9 November 2018; pp. 1–5.
3. Chao, S.; Guan, G.; Hong, G.S. Design of a finless torpedo shaped micro AUV with high maneuverability. In Proceedings of the
OCEANS 2017-Anchorage, Anchorage, AK, USA, 18–21 September 2017; pp. 1–6.
4. Honaryar, A.; Ghiasi, M. Design of a bio-inspired hull shape for an AUV from hydrodynamic stability point of view through
experiment and numerical analysis. J. Bionic Eng. 2018, 15, 950–959. [CrossRef]
5. Dong, H.; Wu, Z.; Chen, D.; Tan, M.; Yu, J. Development of a whale-shark-inspired gliding robotic fish with high maneuverability.
IEEE ASME Trans. Mechatronics 2020, 25, 2824–2834. [CrossRef]
6. Eriksen, C.C.; Osse, T.J.; Light, R.D.; Wen, T.; Lehman, T.W.; Sabin, P.L.; Ballard, J.W.; Chiodi, A.M. Seaglider: A long-range
autonomous underwater vehicle for oceanographic research. IEEE J. Ocean Eng. 2001, 26, 424–436. [CrossRef]
7. Fernandez, R.A.S.; Parra R., E.A.; Milosevic, Z.; Dominguez, S.; Rossi, C. Design, Modeling and Control of a Spherical
Autonomous Underwater Vehicle for Mine Exploration. In Proceedings of the 2018 IEEE/RSJ International Conference on
Intelligent Robots and Systems (IROS), Madrid, Spain, 1–5 October 2018; pp. 1513–1519.
8. Palomer, A.; Ridao, P.; Ribas, D. Inspection of an underwater structure using point-cloud SLAM with an AUV and a laser scanner.
J. Field Robot. 2019, 36, 1333–1344. [CrossRef]
9. Venkatesan, S. AUV for Search & Rescue at sea–An innovative approach. In Proceedings of the 2016 IEEE/OES Autonomous
Underwater Vehicles (AUV), Tokyo, Japan, 6–9 November 2016; pp. 1–9.
10. Fossen, T.I. Guidance and Control of Ocean Vehicles. Ph.D. Thesis, University of Trondheim, Trondheim, Norway, 1999;
ISBN 0-471-94113 1.
11. Gibson, S.B.; Stilwell, D.J. Hydrodynamic Parameter Estimation for Autonomous Underwater Vehicles. IEEE J. Ocean Eng. 2018,
45, 385–394. [CrossRef]
12. Sajedi, Y.; Bozorg, M. Robust estimation of hydrodynamic coefficients of an AUV using Kalman and H∞ filters. Ocean Eng. 2019,
182, 386–394. [CrossRef]
13. Sabet, M.T.; Daniali, H.M.; Fathi, A.; Alizadeh, E. Identification of an autonomous underwater vehicle hydrodynamic model
using the extended, cubature, and transformed unscented Kalman filter. IEEE J. Ocean Eng. 2017, 43, 457–467. [CrossRef]
14. Allotta, B.; Costanzi, R.; Pugi, L.; Ridolfi, A. Identification of the main hydrodynamic parameters of Typhoon AUV from a
reduced experimental dataset. Ocean Eng. 2018, 147, 77–88. [CrossRef]
15. Jagadeesh, P.; Murali, K.; Idichandy, V. Experimental investigation of hydrodynamic force coefficients over AUV hull form. Ocean
Eng. 2009, 36, 113–118. [CrossRef]
16. de Barros, E.A.; Dantas, J.L.; Pascoal, A.M.; de Sá, E. Investigation of normal force and moment coefficients for an AUV at
nonlinear angle of attack and sideslip range. IEEE J. Ocean Eng. 2008, 33, 538–549. [CrossRef]
17. Battista, T.; Woolsey, C.; Perez, T.; Valentinis, F. A dynamic model for underwater vehicle maneuvering near a free surface.
IFAC-PapersOnLine 2016, 49, 68–73. [CrossRef]
18. Wu, J.; Dou, Y.; Lv, H.; Ma, C.; Zhong, L.; Xu, S.; Han, X. Thrust Characteristics of Ducted Propeller and Hydrodynamics of an
Underwater Vehicle in Control Motions. J. Mar. Sci. Eng. 2021, 9, 940. [CrossRef]
19. Qin, D.; Huang, Q.; Pan, G.; Shi, Y.; Li, F.; Han, P. Numerical simulation of hydrodynamic and noise characteristics for a
blended-wing-body underwater glider. Ocean Eng. 2022, 252, 111056. [CrossRef]
20. Phillips, A.; Furlong, M.; Turnock, S.R. The Use of Computational Fluid Dynamics to Determine the Dynamic Stability
of an Autonomous Underwater Vehicle. 2007. Available online: https://eprints.soton.ac.uk/48786/1/ABP_NUTTS.pdf
(accessed on 26 July 2022).
21. Takahashi, K.; Sahoo, P.K. Fundamental CFD Study on the Hydrodynamic Performance of the DARPA SUBOFF Submarine.
In Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering, Glasgow, UK, 9–14 June 2019;
Volume 58776, p. V002T08A052.
22. Amini Foroushani, J.; Sabzpooshani, M. An approach for the estimation of hydrodynamic coefficients of an underwater vehicle in
off-design velocities. J. Mar. Sci. Technol. 2021, 26, 368–381. [CrossRef]
23. Islam, H.; Soares, C.G. Estimation of hydrodynamic derivatives of a container ship using PMM simulation in OpenFOAM. Ocean
Eng. 2018, 164, 414–425. [CrossRef]
24. Liu, H.L.; Huang, T.T. Summary of DARPA SUBOFF Experimental Program Data; Technical Report; Naval Surface Warfare Center
Carderock Div Bethesda Md Hydromechanics: Bethesda, MD, USA, 1998.
25. Gao, T.; Wang, Y.; Pang, Y.; Chen, Q.; Tang, Y. A time-efficient CFD approach for hydrodynamic coefficient determination and
model simplification of submarine. Ocean Eng. 2018, 154, 16–26. [CrossRef]
26. Mitra, A.; Panda, J.P.; Warrior, H.V. Experimental and numerical investigation of the hydrodynamic characteristics of autonomous
underwater vehicles over sea-beds with complex topography. Ocean Eng. 2020, 198, 106978. [CrossRef]
27. Wu, J.; Yang, X.; Xu, S.; Han, X. Numerical investigation on underwater towed system dynamics using a novel hydrodynamic
model. Ocean Eng. 2022, 247, 110632. [CrossRef]
J. Mar. Sci. Eng. 2022, 10, 1049 22 of 22

28. Tavakoli, S.; Dashtimanesh, A. Mathematical simulation of planar motion mechanism test for planing hulls by using 2D+T theory.
Ocean Eng. 2018, 169, 651–672. [CrossRef]
29. Tavakoli, S.; Dashtimanesh, A.; Mancini, S.; Mehr, J.A.; Milanesi, S. Effects of vertical motions on roll of planing hulls. J. Offshore
Mech. Arct. Eng. 2021, 143, 041401. [CrossRef]
30. Huang, S.; Su, X.; Yang, F. Numerical Simulation of 3D Unsteady Flow in Centrifugal Pump by Dynamic Mesh Technique. Sci.
Technol. Rev. 2013, 61, 270–275.
31. Xu, S.; Wu, J.; Zhang, S.; Yang, X. Hydrodynamic analysis of a tethered underwater robot with control equipment subjected to
cnoidal waves. Ocean Eng. 2022, 243, 110264. [CrossRef]
32. Xu, S.; Wu, J.; Lv, H.; Ma, C.; Yang, X.; Wang, H.; Dou, Y. Numerical simulation and hydrodynamics analysis of a tethered
underwater robot with control equipment. J. Mar. Sci. Technol. 2022, 27, 368–382. [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. [CrossRef]
34. Lin, Y.H.; Chiu, Y.C. The estimation of hydrodynamic coefficients of an autonomous underwater vehicle by comparing a dynamic
mesh model with a horizontal planar motion mechanism experiment. Ocean Eng. 2022, 249, 110847. [CrossRef]
35. Javanmard, E.; Yari, E.; Mehr, J.A. Numerical investigation on the effect of shaft inclination angle on hydrodynamic characteristics
of a surface-piercing propeller. Appl. Ocean Res. 2020, 98, 102108. [CrossRef]
36. Hong, L.; Fang, R.; Cai, X.; Wang, X. Numerical Investigation on Hydrodynamic Performance of a Portable AUV. J. Mar. Sci. Eng.
2021, 9, 812. [CrossRef]
37. Hinze, J. Turbulence; McGraw-Hill Publishing Co.: New York, NY, USA, 1975.
38. Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 1994, 32, 1598–1605. [CrossRef]
39. Sahoo, P.; Takahashi, K. Numerical Study on the Hydrodynamic Performance of the DARPA Suboff Submarine for Steady
Translation. In Proceedings of the ASME 2020 39th International Conference on Ocean, Offshore and Arctic Engineering, Virtual,
3–7 August 2020.
40. Toxopeus, S.; Atsavapranee, P.; Wolf, E.; Daum, S.; Pattenden, R.; Widjaja, R.; Zhang, J.T.; Gerber, A. Collaborative CFD exercise
for a submarine in a steady turn. In International Conference on Offshore Mechanics and Arctic Engineering; American Society of
Mechanical Engineers: New York, NY, USA, 2012; Volume 44922, pp. 761–772.
41. Roddy, R.F. Investigation of the Stability and Control Characteristics of Several Configurations of the DARPA SUBOFF Model (DTRC
Model 5470) from Captive-Model Experiments; Technical Report; David Taylor Research Center Bethesda MD Ship Hydromechanics
Department: Bethesda, MD, USA, 1990.
42. Lee, S.K.; Joung, T.H.; Cheo, S.J.; Jang, T.S.; Lee, J.H. Evaluation of the added mass for a spheroid-type unmanned underwater
vehicle by vertical planar motion mechanism test. Int. J. Nav. Archit. Ocean Eng. 2011, 3, 174–180. [CrossRef]

You might also like