You are on page 1of 66

摘要

i
ABSTRACT

This study focuses on investigating the motion and hydrodynamic performance of an


autonomous underwater helicopter (AUH). The research examines various aspects,
including the establishment of complete physical models and boundary conditions, the
effect of velocity on the hydrodynamic characteristics of the AUH during motion, and the
impact of angle characteristics on its hydrodynamics. Through the use of comprehensive
physical models and accurate boundary conditions, the study provides a solid foundation
for analyzing the hydrodynamic behavior of the AUH. The investigation into the effect of
velocity reveals that as the velocity increases, the drag experienced by the AUH also
increases, which affects its overall performance and maneuverability. Additionally,
variations in the angle of attack (AOA) significantly influence the lift generated by the
AUH, with higher angles leading to changes in lift forces that can affect stability and
control. The study further explores the effects of velocity on streamlines and pressure
distribution around the AUH during motion. Higher velocities result in more pronounced
vortices and alterations in flow patterns, while different velocities and angle characteristics
lead to variations in pressure distribution. Understanding these effects is crucial for
optimizing the AUH's design and control strategies, as they directly impact its lift, drag,
and stability. The findings contribute to a comprehensive understanding of the AUH's
hydrodynamic characteristics and provide valuable insights for engineers and researchers
to enhance the efficiency, maneuverability, and success of AUHs in underwater
applications. The simulation results show good agreement with experimental results,
validating the accuracy of the study's findings. The study concludes by highlighting the
potential for further research and development in the field of autonomous underwater
vehicles, aiming to advance underwater exploration, inspection, and surveillance
capabilities.
The study highlights the interplay between velocity, streamlined patterns, pressure
distribution, and the hydrodynamic performance of the AUH. By investigating these
effects, valuable insights have been gained, enabling the optimization of the AUH's motion,
stability, and overall performance. The findings provide a foundation for further research
and development in the field of autonomous underwater vehicles, contributing to the
advancement of underwater exploration, inspection, and surveillance capabilities. The
simulation results are in good agreement with the experimental results, with the difference
ii
between the two being less than 10%. This validation underscores the accuracy and
reliability of the numerical models and simulations used in the study. Such confidence in
the simulation results allows for meaningful interpretations and extrapolations of the data.
The investigation into the effect of angle characteristics on the AUH's hydrodynamics
reveals an important relationship between the angle of attack and lift generation. At lower
angles of attack, the lift coefficient gradually increases, but beyond 10 degrees, there is a
rapid increase due to the formation of vortices near the trailing edge of the disk-shaped
hull. This phenomenon significantly enhances the lift generation capability of the AUH.
Understanding and d the dd becomes critical for achieving the desired lift forces and
maintaining stability during the AUH's motion. The analysis of velocity streamline patterns
provides valuable insights into the fluid dynamics around the AUH. As the velocity of the
AUH increases, the patterns of the streamline undergo significant changes. Higher
velocities lead to the formation of more pronounced vortices and alterations in the flow
patterns. These observations have implications for the vehicle's stability, maneuverability,
and energy consumption. By optimizing the streamlined patterns through careful velocity
control, engineers can enhance the AUH's motion efficiency and reduce drag forces,
ultimately improving its overall performance.
Examining the pressure distribution around the AUH during motion further enhances
our understanding of its hydrodynamic behavior. The comparison of pressure contours at
different velocities and angle characteristics highlights variations in pressure magnitude
and distribution. Higher velocities and specific angles of attack result in distinct pressure
patterns, indicating areas of high and low pressure around the AUH. This knowledge is
crucial for optimizing the design of the AUH's hull and control surfaces to maximize lift,
minimize drag, and ensure stability.
In conclusion, this study provides comprehensive insights into an autonomous
underwater helicopter's motion and hydrodynamic performance. Engineers and researchers
can make informed decisions regarding the design, operation, and control of AUHs by
considering the effects of velocity, angle characteristics, streamlined patterns, and pressure
distribution. The findings contribute to improving the efficiency, maneuverability, and
mission success of AUHs in various underwater applications. Future research in this field
can build upon these insights to further advance the capabilities and performance of AUHs,
paving the way for new developments in underwater exploration, inspection, and
surveillance.
iii
ACKNOWLEDGEMENTS

First of all my sincere appreciation and deep respect go to my advisor, Chair Professor
... for his support and patient advice throughout my studies in… I can not get the
achievements in my research without his advice.

I would like to thank the members of my dissertation committee:... for many helpful
discussions about the experiments.

It is my pleasure to thank people in ...

Finally, a special thanks to my parents, my sisters, my brother, and also to my


relatives for their warm encouragement, sharing my difficulties over the years of my
studies in .... Life in ... has taught me a lot of new things and I thank all those persons who
made this unforgettable experience for me.

iv
TABLE OF CONTENTS

摘要 ............................................................................................................................ i

ABSTRACT .............................................................................................................. ii

ACKNOWLEDGEMENTS..................................................................................... iv

TABLE OF CONTENTS ......................................................................................... v

LIST OF FIGURES ................................................................................................ vii

LIST OF TABLES ................................................................................................... ix

NOMENCLATURE ................................................................................................. x

CHAPTER I: INTRODUCTION ............................................................................ 1

1.1. Motivation ........................................................................................................ 1

1.2. Objectives ........................................................................................................ 3

1.3. Organization of the thesis ................................................................................. 3

CHAPTER II: BACKGROUND.............................................................................. 5

2.1. The theoretical basis for determining profile..................................................... 5

2.2. Basic operating principles................................................................................. 6

2.3. The fundamental theory of Computational Fluid Dynamics .............................. 9

CHAPTER III: LITERATURE REVIEW ............................................................ 14

3.1. Theoretical and experimental investigations on motion and hydrodynamic


performance of the autonomous underwater helicopter .......................................... 14

3.2. Numerical investigations on motion and design optimization of the Autonomous


Underwater Helicopter .......................................................................................... 16

CHAPTER IV: DESCRIPTION OF PHYSICAL MODELS AND NUMERICAL


METHODOLOGY ................................................................................................. 18

4.1. Physical model ............................................................................................... 18

4.2. Mathematical formulation............................................................................... 19

4.2.1. Boundary layer ......................................................................................... 19

v
4.2.2. Boundary conditions for translational simulations .................................... 20

4.2.3. Mathematical model of motion stability ................................................... 22

4.3. Numerical methodology ................................................................................. 23

4.3. Computational process ................................................................................... 24

4.5. Mesh quality and convergence evaluation....................................................... 25

CHAPTER V: HYDRODYNAMIC CHARACTERISTICS OF THE


AUTONOMOUS UNDERWATER HELICOPTER ............................................ 29

5.1. The hydrodynamic characteristics of the AUH ............................................... 29

5.2. The effect of inclined angles ........................................................................... 29

5.3. The effect of velocities ................................................................................... 42

CHAPTER VI: CONCLUSION AND FUTURE WORKS .................................. 47

6.1. Conclusion ..................................................................................................... 47

6.2. Future works................................................................................................... 49

BIBLIOGRAPHIES ............................................................................................. 511

vi
LIST OF FIGURES

Figure 2.1. (a) Parameters of the disc-shaped hull; (b) 3D model of the hull;
(c) The photo of the disc-shaped hull [12]........................................................................ 5

Figure 2.2. AUH’s main working pattern and the differences between AUH and other
underwater vehicles [12] .................................................................................................. 6

Figure 2.3. The general layout of the prototype: (a) Top view layout concept diagram; (b)
3D model diagram [12] .................................................................................................... 7

Figure 2.4. Fixed coordinate system and moving coordinate system ............................... 8

Figure. 2.5. Control volume and control surface ........................................................... 12

Figure. 3.1. 3D model of AUH with horizontal propellers in through holes [16] ........... 17

Figure 3.2. Streamlines seeing from the rear of the vessel for the AUH geometry with two
horizontal propellers at the surging velocity of 2 m/s, for (a) HG1 and (b) HG3,
respectively; (c) the corresponding temporal evolution of the drag force with solid lines for
HG1 and dashed lines for HG3 from simulation; and (d) standard deviation of the
hydrodynamic force for HG1and HG3, respectively, calculated from (c) [16] ............... 17

Figure 4.1. (a) Parameters of the AUH; (b) 3D model of the AUH................................ 19

Figure 4.2. Mesh type in computational fluid domains translating in the horizontal and
vertical planes and thicknesses of the boundary layer .................................................... 20

Figure 4.3. Computational domain adopted to simulate unsteady, 3-D, turbulent flow for
a translating AUH .......................................................................................................... 21

Figure 4.4. The flow chart of the computational process. .............................................. 25

Figure 4.5. Typical mesh used in the computational domain for (a) an AUH and (b) the
area around the object. The mesh is finer near the interface and inside the liquid. .......... 28

Figure 5.1. The velocity streamlines at the tail of the AUH ........................................... 30

vii
Figure 5.2. Streamlines of AUH in plane xoz at a velocity of 2 m/s and an AOA of 0
degrees .......................................................................................................................... 31
Figure 5.3. Drag and lift force on the plane xoz value over time .................................. 31
Figure 5.4. Pressure distribution and streamlines of AUH in a plane at a velocity of 2 m/s
...................................................................................................................................... 32
Figure 5.5. a) Pressure on the AUH with the 50-mm aquatic transducer. b) Pressure on
the AUH with the 100-mm aquatic transducer. .............................................................. 33
Figure 5.6. a) Velocity around the AUH with 50-mm aquatic transducer and b) Velocity
around the AUH with 100-mm aquatic transducer ......................................................... 34
Figure 5.7. Drag force and AOA at vi = 2m/s ............................................................... 35
Figure 5.8. Lift force and AOA at vi = 2m/s ................................................................. 36
Figure 5.9. Streamlines at xoz plane (right) of AUH at vi = 2 m/s and AOA = 0o – 20o from
numerical simulation. .................................................................................................... 37
Figure 5.10. Pressure distribution and streamlines of AUH in plane xoz at vi = 2 m/s AOA
= 0o – 20o from numerical simulation............................................................................. 39
Figure 5.11. Comparison of simulation results and experimental results in terms of drag
force. ............................................................................................................................. 40
Figure 5.12. Comparison of simulation results and experimental results in terms of lift
force ............................................................................................................................. 41
Figure 5.13. The drag force at different surging velocities for hull geometries calculated
from numerical simulation ............................................................................................. 43
Figure 5.14. The lift force at different surging velocities for hull geometries was calculated
from numerical simulation. ............................................................................................ 44
Figure 5.15. Streamlines at xoz plane (right) of AUH at vi = 0,5 - 2 m/s from numerical
simulation. ..................................................................................................................... 45
Figure 5.16. Pressure distribution and streamlines of AUH in plane xoz at vi = 0,5 - 2 m/s
from numerical simulation ............................................................................................. 46

viii
LIST OF TABLES

Table 4.1. Main AUH parameters ........................................................................... 22

Table 4.2. The differences between five meshing objects ........................................ 26

ix
NOMENCLATURE

English symbols

γ Coefficient of kinematic viscosity

Re Reynolds number

L The diameter

x, y, x Body axis coordinates

h Design depth

H Height

m Mass

v Service speed

F The net force vector acting

Ar The relative tolerance

Aa The absolute tolerance

S The solution vector corresponding to the solution

E The error estimate

hn, ln The parameter of the edge point

Abbreviations

AUVs Autonomous Underwater Vehicles

AUH Autonomous Underwater Helicopter

CFD Computational Fluids Dynamics

UDF User-defined function

x
ALE Arbitrary Lagrangian Eulerian

CSF Continuum Surface Force

FEM Finite Element Method

EOM Equations of motion

6-DOF Six degrees of freedom

NURBS Non-Uniform Rational B-Spline

UUVs Unmanned Underwater Vehicles

FDM Finite Differnces Method

FVM Finite Volume Method

ROV Remote Operated Vehicle

BHS Bionic Hull Shape

RSM Response Surface Method

RANS Reynolds Averaged Navier–Stokes

S-AUH Small-Autonomous Underwater Helicopter

AOA Angle of attack

AOD Angle of drift

AOO Angle of orientation

LSF Level set function

NE Number of elements

ROV Remote Operated Vehicle

xi
CHAPTER I: INTRODUCTION

1.1. Motivation
In recent years, the study of motion stability in autonomous underwater vehicles
(AUVs), specifically a new type called “autonomous underwater helicopter” (AUH) with
a disc-shaped hull, has gained increasing importance within the field of deep-sea mobile
observation networks [1-7]. Motion stability analysis is crucial for improving the
performance of AUHs in complex underwater tasks, such as pipeline maintenance,
mobile observation networks, resource exploration, and depth navigation [8-10].
Notably, successful tests of unmanned disc-type submersibles at sea have marked
significant milestones and advancements in unmanned submersible technology [11].
Furthermore, innovative technical methods from the aircraft industry have been applied
to study the hydrodynamics of autonomous underwater ships, leading to new insights and
opportunities for the design, optimization, and operation of fluid problems. These
developments have generated excitement and interest in understanding the underlying
physical mechanisms driving motion stability in AUHs.
Compared to a torpedo-shaped AUV, the AUH's motion stability in the vertical
plane is better, which proves that the stability of the disk-shaped AUH hull is high [12].
Thus, the AUH vertical course-keeping ability is better, whereas the vertical
maneuverability is reduced. This is conducive to adequate landing navigation and
guidance in deep-sea conditions and over long distances [13]. Chen-Wei Chen et al. [14]
had the Computational Fluids Dynamics (CFD) results and the Routh stability criterion
applied to identify whether different AUH designs satisfied the motion stability criteria
in the horizontal and vertical planes. The analyzed hydrodynamic performance could
provide a meaningful reference for the future development of new AUVs, and based on
these results, verification studies on motion stability clouds are performed for vectoring
thrusters and buoyancy engines appended to AUHs in the future. Yuan Lin et al. [15]
Based on the numerical simulation with the overlapping mesh method through the use of
the dynamic mesh user-defined function (UDF), the hydrodynamic performance of AUH
basic hulls (the tailing shape on the hydrodynamic performance, two fore-aft asymmetric
geometries are proposed) is studied. When the AUH moves with an angle of attack
(AOA) less than 15 degrees and a speed of 2 m/s, the model with the rear part is extended
and sharpened by a straight line and has less resistance, greater lift, and a smaller
1
overturning moment than the recent one. This shows that this model has less energy loss
and higher motion stability than our model. Zhikun Wang at el. [12] the AUH, a new type
of disc-shaped AUV for seabed-to-seabed operating modes, is proposed. It has the
characteristics of maneuverability and vertical motion stability. The prototype was built,
and several groups of experiments were carried out on it. Based on this, we draw the
following conclusions: Because the special disc shape has the characteristics of
horizontal in-plane isotropy, large vertical resistance, and small horizontal resistance,
there is no need for over-actuation to achieve stable motion with high maneuverability
within acceptable limits. It is feasible to increase endurance by reducing the number of
propellers. The prototype of the AUH has the potential for high maneuverability. It has
the potential to cruise in a small area flexibly, land on the seabed, and take off. It is
indicated that the AUH may be a solution for increasingly complex undersea engineering
tasks, especially near-seabed operations. At present, although the AUH has the potential
for high maneuverability, only a simple quadrangle path test has been carried out, due to
the lack of a high-accuracy positioning system and navigation system. Next, we will try
to set up a more complex running path for the AUH in conjunction to further demonstrate
its excellent maneuverability. In addition, with the high-accuracy positioning system, we
will be able to more intuitively compare the performance of other submersibles, including
torpedo-type submersibles, to further verify the advantages of the high maneuverability
of the disc-shaped AUH.
In this scholarly article, we will examine and demonstrate the effectiveness of the
angle of attack and velocity in the context of AUH systems. To date, to our best
knowledge, some researchers [11, 16] have done the simulation works but no appropriate
simulation results are consistent with the previous experiment. Our works presented in
this dissertation show the quantitative agreement between the numerical and
experimental results. Furthermore, the UDF motion will be applied to illustrate precisely
the displacement of special points on the AUH surface.

This research addresses a crucial issue in the control and recognition of factors
affecting the high-pressure flow over the surface of AUH systems during their motion,
particularly in velocity fields. Our study meticulously investigates this problem,
considering various obstacles that arise when using numerical computation to analyze the
transient behavior of fluid problems associated with AUH motion and effective operation.

2
These obstacles include the coupling of three computational methods, namely the
arbitrary Lagrangian-Eulerian (ALE), the conservative level set, and the continuum
surface force (CSF); the complex phenomenon of dynamic contact angle of liquid during
the actuation process; the nonlinearity and dependence of physical functions of fluid
parameters on factors such as temperature, time, and level set value; the wetting ability
of liquid/solid interface; the displacement condition of liquid/gas interface; and the
utilization of SST and k-epsilon methods in CFD stimulation. However, with the
advancement of modern technology and computational methods, numerical simulations
offer a viable solution to tackle these challenges. The motivation behind this dissertation
is to comprehensively understand and solve these intriguing problems, ultimately
optimizing the profile of AUH. Currently, only a few simple models have emerged as
paradigms for investigating the hydrodynamic properties of AUH.

1.2. Objectives
The main objective of the present study is to use an autonomous underwater
helicopter with a disk-shaped hull geometry that has been developed by Wang et al. [11]
to the improvement of the hydrodynamic performance of the newly developed AUH to
enhance the hydrodynamic stability meanwhile reduce the drag force during the surge
motion. For this aim, the following goals are investigated such as:

- The motion and hydrodynamic performance of the autonomous underwater


helicopter
- Completely physical models and boundary conditions
- The effect of velocity on the hydrodynamic characteristics of the AUH during
motion
- The effect of characteristics of angles on the hydrodynamic of the AUH during
motion

1.3. Organization of the thesis


This thesis is divided into six chapters. The main outline of each chapter follows:
Chapter 1 introduces the field of autonomous underwater vehicles, including its
definition and scientific aspects. Chapter 1 also points out the motivation and objectives
of the study.

3
Chapter 2 provides the theoretical fundamentals required for basic operating
principles. Some of the fundamental theories of Computational Fluid Dynamics and the
theoretical basis for determining the profile.
Chapter 3 some theoretical experiments, numerical computation, and also related to
the hydrodynamic performance of the AUV influenced by the distance to the seabed
Chapter 4 describes the physical problems with the appropriate theoretical
formulations and computational methods. In this chapter, the physical models with the
governing equations, boundary conditions, and initial conditions solving the fluids flow
through the finite element method (FEM) are presented in detail.
Chapter 5 presents the numerical results of raised and lowered motion AUH. The
motion behavior of AUH effects of temperature and velocity gradients, inclined angle,
and viscosities are in detail analyzed and discussed. The results of numerical results of
AUH motion raised in Chapter 5 are also compared with previous experiments.
Finally, the thesis ends with chapter 6, highlighting the contributions and conclusion
from the present research. In addition, some recommendations for future works are also
given in this chapter.

4
CHAPTER II: BACKGROUND

2.1. The theoretical basis for determining the profile


In the CFD study on the motion stability of the Autonomous Underwater Helicopter,
the Routh stability criterion [17-20] was utilized to determine the dynamic stability of the
AUH in both horizontal and vertical planes. The criterion was derived from the well-
known equations of motion (EOM) [21] in six degrees of freedom (6-DOF) and expressed
in terms of translational and rotational hydrodynamic coefficients obtained from the
ANSYS-CFX CFD solver. The CFD solver accurately predicts the hydrodynamic forces
and moments on the AUH under various flow conditions and configurations, facilitating
hydrodynamic optimization. The CFD case studies involved different configurations of
the AUH hull, including variations in the lengths of aquatic transducers and depth gauges,
and without thrusters.
The disc-shaped hull is a crucial design aspect of the AUH, aimed at achieving high
maneuverability in the horizontal plane and stability in the vertical direction. To fulfill
this goal, the disc-shaped hull is defined by a curve, specifically a Non-Uniform Rational
B-Spline (NURBS), chosen for its curvature continuity and simple definitions that
facilitate design adjustments. The NURBS curve is defined by seven control points, as
shown in Figure 2.1a, with the main size of the disc denoted as L and set at a value of
1000 mm. H is the direct parameter of the shape height. h1 and l1 is the parameter of the
edge point. h2 and l2 is the parameter of the transition point. h3 and l3 is the parameter of
the optimized point. According to the conclusions, when H = 200 mm, h1 = 50 mm, h2 =
80 mm, h3 = 2(H - h1)/3 + h1 = 150 mm, l1 = L/8 = 125 mm, l2 = L/4 = 250 mm, l3 = 3L/8
= 375 mm, the movement in the horizontal plane has low resistance. In addition, in the
vertical direction, it has good stability.

Figure 2.1. (a) Parameters of the disc-shaped hull; (b) 3D model of the hull;
(c) The photo of the disc-shaped hull [12]
5
2.2. Basic operating principles
Figure 2.2 illustrates the primary operational pattern of the Autonomous
Underwater Helicopter, which is designed to perform tasks such as communication,
charging, and equipment maintenance, among others, from the seabed to the seabed. The
AUH initiates its mission from the surface and descends to the working depth using a
buoyancy adjustment system. With guidance from the positioning system, it maneuvers
from one Subsea Station to another, landing on or hovering over them to complete its
assigned tasks. To achieve accurate motion above the base station, the AUH requires
excellent maneuverability in the horizontal plane, while stable hovering over the base
station necessitates stability in the vertical direction. The Subsea Station serves as a base
station for the AUH, facilitating functions such as data interaction and power
transmission. The disc-shaped body of the revolution of the AUH allows for flexible
steering, with low damping in horizontal motion but high damping in a vertical motion.
These distinctive characteristics provide the AUH with unique advantages in precise
landing and stable hovering operations. Once the missions are completed, the AUH can
take off and navigate to the next base station using the positioning and navigation system,
and ultimately ascend to the surface using the buoyancy adjustment system after all tasks
are finished.

Figure 2.2. AUH’s main working pattern and the differences between AUH and
other underwater vehicles [12]
6
In order to fulfill the requirements of the main working pattern, the AUH needs to
possess both maneuverability in the horizontal plane and stability in the vertical direction.
Leveraging the high maneuverability of the AUH's disc shape and the low damping
characteristic of zero-radius rotation, stability can be achieved through symmetrical mass
distribution in the plane, and positioning a large mass near the center can help reduce the
moment of inertia, further enhancing the AUH's maneuverability. The layout concept
diagram and 3D model diagram of the AUH, as depicted in Figure 2.3, consist of two
symmetrical cylindrical cabins, a battery cabin, a control system cabin, four propellers,
iUSBL equipment for positioning and navigation, buoyancy material, a skeleton for
connecting and supporting, and a disc-shaped hull. Each cabin has successfully passed
the 100 m pressure seal test.

Figure 2.3. The general layout of the prototype: (a) Top view layout concept
diagram; (b) 3D model diagram [12]

The AUH exhibits low resistance in horizontal movement and good stability in the
vertical direction, as shown in Figure 2.3. The prototype has a diameter of 1 meter, a
height of 45 centimeters, and weighs 42 kilograms. It is designed to operate at a speed of
0.8 meters per second and a depth of 100 meters. The choice of a 100-meter depth for the
design is based on the need to keep costs low for prototype testing and to ensure ease of
finding suitable experimental environments. The design speed takes into consideration
the estimated average flow rate in the working environment, accounting for the need to
achieve hovering with sufficient resistance to flow.
To achieve high maneuverability, steady linear motion, fast turning, and a small
turning radius are crucial requirements. Typically, underwater vehicles with high
maneuverability rely on a six-degree-of-freedom system with over-actuation, such as the
7
hovering type AUV Cyclops, which has 8 thrusters. However, using too many propulsion
components can compromise the favorable hydrodynamic characteristics of the disc
shape and result in increased energy loss. Therefore, based on the dynamic analysis in
this section, efforts will be made to achieve high maneuverability in the AUH with as
few propulsion components as possible.
To describe the motion conveniently, the coordinate system is established as shown
in Figure 2.4. According to the momentum theorem, the linear motion description of
AUH can easily be described as:

dH d ( mVGE )
F  (2.1)
dt dt

where F is the resultant external force vector. H is the momentum of AUH. VGE is the
velocity of AUH relative to the fixed coordinate system E-xyz. Through the
transformation between coordinate systems, the equation of motion in moving coordinate
system o-pqr can be described as:

d (V   RGo )  V  
F  m  V  RGo   ( RGo )  (2.2)
dt  t t 

E x

q
p

Figure 2.4. Fixed coordinate system and moving coordinate system

According to Euler’s second law, the AUH rotational motion is described as:

8
dL  V
T  J   ( J  )  RGo m Go  RGo ( mVGo ) (2.3)
dt t t

where T is the resultant external moment vector relative to the origin of the moving
coordinate system. L is the Momentum moment of AUH relative to the fixed coordinate
system E-pqr. RGo and VGo are the distance and velocity of AUH’s center of gravity
relative to o-pqr, respectively. J is the moment of inertia. In that Equation, the first term
represents the moment of inertia of rotating motion; the second term represents the
moment of inertia of rotating unbalanced rotation of rotating axis; the latter is benign
because the origin is not in the center of gravity; the third term is the moment of inertia
of linear motion; and the fourth term is the moment of inertia of centrifugal motion. The
complete description of AUH motion can be obtained by both Equations.

2.3. The fundamental theory of Computational Fluid Dynamics


Computational fluid dynamics is a well-established numerical tool widely used in
fluid mechanics to solve a variety of problems. Le et al. [22-30] developed a CFD
simulation scheme to investigate fluid motion and hydrodynamic characteristics, such as
pressure and velocity, in microchannels or sterilization chambers. Numerous studies have
utilized CFD simulation methods to determine the hydrodynamic characteristics of
autonomous underwater vehicles. For instance, Jia et al. [31] calculated the hydro-drag
and lift of underwater gliders using kinematic equations and parameters' relations in
conjunction with CFD software. Pan et al. [32] analyzed the hydrodynamic
characteristics of unmanned underwater vehicles (UUVs) using CFD techniques, and
their numerical results were found to be in good agreement with engineering estimates.
Amory et al. [33] employed CFD to illustrate pressure and velocity distributions over
AUV bodies, showing that the streamlined SEMBIO hull has smaller hydrodynamic
parameters compared to the torpedo-shaped MONSUN hull. Sousa et al. [34] presented
and analyzed hydrodynamic characteristics of turbulent fluid flow over different AUV
hull shapes to optimize hull design, resulting in lower drag force and reduced energy
consumption for the optimized design. Randeni et al. [35] studied the hydrodynamic
interaction between the motion of two underwater bodies using CFD models and
simplified methods, obtaining critical hydrodynamic coefficients for predicting AUV
performance under varying conditions. Dantas et al. [36] considered the effect of control

9
surfaces on AUV maneuverability using CFD, establishing a linear relationship between
attack angle and control surface deflection for control surface stall occurrence. Tyagi et
al. [37] highlighted the importance of transverse hydrodynamic coefficients computed by
CFD simulations in the maneuverability analysis of AUV hulls. While most studies have
employed two-equation turbulence models (k-ε and k-ω) to optimize AUV design, recent
works are exploring Reynolds stress models due to limitations in turbulence physics
imposed by assumptions in two-equation models (e.g., eddy viscosity hypothesis,
gradient diffusion hypothesis). Interestingly, despite numerous studies on the
hydrodynamic characteristics of AUVs, there is limited research on torpedo-shaped
underwater gliders. Therefore, utilizing CFD to analyze hydrodynamic characteristics of
such gliders presents a promising idea for minimizing drag force and increasing
propulsion efficiency, potentially leading to optimized designs in the future.
In the present study, the CFD method is used to investigate the hydrodynamic
characteristics of a torpedo-shaped underwater glider. The Navier–Stokes equations and
the energy equation subjected to the prevalent boundary conditions are solved
numerically by Comsol Multiphysics software. The underwater glider motion with the
various velocities, different angles of attack, and the optimized shape that minimizes the
drag force is also considered in this study
In general, fluid problems can be reduced to three main groups of mathematical
equations:

 The Continuity Equation


 The Momentum Equation
 The Energy Equation.

The above equations refer to physical processes. They are mathematical statements
of the three fundamental physical principles on which all fluid dynamics is based:

1) Conservation of mass.
2) F = ma (Newton's second law).
3) Energy conservation.

From those bases, we build the basic equations of fluid activity:

 Conserved equation of continuity:

10
+ ∇. ⃗ =0 (2.4)

 Non-conservative continuity equation:

+ ∇. ⃗ = 0 (2.5)

 Momentum equation (full Navier-Stock equation, conserved form):

( ) ( ) ( ) ( )
+ + + =− + ∇. ⃗ + 2

+ + + + + (2.6)

( ) ( ) ( ) ( )
+ + + =− + +

+ ∇. ⃗ + 2 + ( + +

( ) ( ) ( ) ( )
+ + + =− + +

+ + + ∇. ⃗ + 2 +

 Energy equation:
- Unsecured form:

+ = ̇+ + + (2.7)
2
( ) ( ) ( ) ( ) ( )
− + + + + + +

( ) ( )
+ + + + + + ⃗. ⃗

- Guaranteed form:

+ +∇ + ⃗ = ̇+ (2.8)
2 2
( ) ( ) ( )
+ + − + +

11
( ) ( )
+ + + + + +

( ) ( )
+ + + + ⃗. ⃗

Solving the above equations in each specific problem directly by the analytical
method is very complicated. That's why the CFD method was born: Computational Fluid
Dynamic. CFD is a collection of methods to solve the above equations by computer,
including the following methods:

- Finite Difference Method (FDM): Use the difference diagram to approximate the
operators (Operators of partial derivatives, Integral operators, ..) to solve equations
without considering other physical properties.

- Finite Volume Method (FVM): We define the control volumes covered by the
control faces. These volume regions are large enough that the number of particles within
the region is infinitely large, but also small enough that the properties of the particles are
equivalent. After determining the control volumes, we proceed to calculate the defined
surfaces.

Figure 2.5. Control volume and control surface

- Finite Element Method (FEM): Using basis functions to build models, these
functions have finite definite spaces, so they are called finite.

Comments

The above methods are arranged in ascending order of complexity in programming


and use. The FDM method has a simple code, easy to evaluate the stability of the system.
But because this is a non-conservative method, the results do not always guarantee the
necessary physical properties (have unconfirmed solutions). This is the method for

12
beginners to learn about CFD. FVM is a conservative method, the code is not too complex
and easy to apply in practice, so it is used the most. However, FEM - the method with the
highest complexity - is the method with the most potential, which is being developed
very quickly by scientists.

Because CFD simulation software is modular, it is necessary to select the correct


modules before simulating. We consider our problem to have the following two
characteristics:

- The problem considers the interaction between the glider and the water
environment: this is the interaction between a solid body with structure and fluid.

- In addition to the movement of water, the glider has its motion.

From those comments, we see the need to use the modules: Laminar flow (laminar
flow, because the train is not running too fast and also to simplify the problem), Structure
(affecting the pressure of the water flow on the hull), and Moving Mesh (Ships move,
grid coordinates must also move).

13
CHAPTER III: LITERATURE REVIEW

The goal of this chapter is to provide a review of the literature related to Theoretical
and experimental investigations on the motion and hydrodynamic performance of the
autonomous underwater helicopter. Moreover, numerical investigations on motion and
design optimization of the Autonomous Underwater Helicopter are also presented.

3.1. Theoretical and experimental investigations on motion and hydrodynamic


performance of the autonomous underwater helicopter

Wang et al. [11] researched an autonomous underwater helicopter with a disk-shaped


hull geometry have been developed which can take movement in both horizontal and
vertical directions, and has a zero turning diameter. The motion of AUH is controlled by
two vertical propellers and four horizontal propellers. The long-distance cruise of the
AUH, acting like an AUV, is realized by the surge motion along the x-axis driving by
horizontal propellers. Meanwhile, with two vertical propellers, the AUH is capable of
taking a short-range heave motion along the z-axis for, e.g., offering and landing at a
certain spot on the seabed for energy replenishment, data exchange or close-up
investigation, behaving like the ROV (Remote Operated Vehicle). To summarize, the
AUH is designed for the condition when a high level of maneuverability is needed.

Zhikun Wang et al. [12] built a prototype that has the characteristics of
maneuverability and vertical motion stability, and several groups of experiments were
carried out on it. Based on this, they draw the following conclusions about the special
disc shape having the characteristics of horizontal in-plane isotropy, large vertical
resistance, and small horizontal resistance, there is no need for over-actuation to achieve
stable motion with high maneuverability within acceptable limits. It is feasible to increase
endurance by reducing the number of propellers. At least four propellers are necessary
for the maneuverability of the AUH. The pitch vibration in the horizontal linear motion
experiment is consistent with the prediction of dynamic analysis. A set of propellers that
can provide an active recovery moment is required to eliminate the harmful resistance
and angle of attack caused by this vibration. Together with at least two propellers needed
to complete the plane motion, four propellers in two groups are necessary for AUH to
complete the stable motion of high maneuverability

14
Chen-Wei Chen et al. [14] had the CFD results and the Routh stability criterion
applied to identify whether different AUH designs satisfied the motion stability criteria
in the horizontal and vertical planes. The analyzed hydrodynamic performance could
provide a meaningful reference for the future development of new AUVs, and based on
these results, verification studies on motion stability clouds be performed for vectoring
thrusters and buoyancy engines appended to AUHs in the future.

Yuan Lin et al. [15] based on the numerical simulation with the overlapping mesh
method through the use of the UDF, the hydrodynamic performance of two AUH basic
hulls (HG1 and HG3) is studied. When the AUH moves with an AOA less than 15 degrees
and a speed of 2 m/s, HG3 has less resistance, greater lift, and a smaller overturning
moment than HG1. This shows that HG3 has less energy loss and higher motion stability
than HG1. Zhikun Wang et al. [12], the AUH, a new type of disc-shaped AUV for seabed-
to-seabed operating modes, is proposed. It has the characteristics of maneuverability and
vertical motion stability.

Experimental works

To the best of my knowledge, many experimental works evaluated both the elevated
hydrodynamic performance to reduce the hydrodynamic drag and AUH's large
amplitude. Regarding the shape optimization of AUVs, Sun et al. proposed a bionic hull
shape (BHS) for AUVs according to the profile of the humpback whale, which was
optimized using a surrogate model and the response surface method (RSM). Honaryar
and Ghiasi proposed a body shape of catfish with sharp edges for AUV to improve the
turning rate; Alvarez et al. applied a simulated annealing algorithm to optimize the hull
shape of AUV, the head and tail of which were sharpened to minimize the wave resistance
near the free surface; Divsalar concluded that a hull shape with bullet nose and the sharp
tail has a better hydrodynamic performance compared to the traditional ones. As to the
hydrodynamic investigation of the disk-shaped AUH, Chen et al. had an analysis of
numerical simulation has also been applied to study the drag reduction performance and
the route motion excellence, Chen et al. demonstrate the hydrodynamic interaction of a
ship with the wave effect, the Chen and Lu also demonstrate the water entry impact force
as well as the Magnus effect. In a recent study, Lin et al found that the surging of the
disk-shaped AUH is hydrodynamically unstable, during which numbers of secondary

15
flow areas develop near the rear part of the vessel, which induces pulsation of the lift and
side force normal to the surge direction. The instability origins from the hydrodynamic
actuation give rise to a requirement for a high-performance controlling system based on
vertical propellers, which also results in high-power consumption. Therefore, to elevate
the hydrodynamic performance, an improvement on the hull geometry is essential to
reduce both the hydrodynamic drag and its large-amplitude fluctuation. Nesteruk et al.
prove that a more slender body shape makes marine species such as Medi-terranean
spearfish, Indo-Pacific sailfish, black marlin, and swordfish fast-swimming fishes. They
are also found to have long dorsal fins, extending along the body that may improve
hydrodynamic stability by Nguyen et al. and Wang et al. Therefore, for a disk-shaped
vessel such as the AUH, it is meaningful to study whether sharpening the hull geometry
is an effective way to improve the hydrodynamic performance, which is lacking at the
present stage. The present study aims to improve the hydrodynamic performance of the
autonomous underwater helicopter.

3.2. Numerical investigations on motion and design optimization of the Autonomous


Underwater Helicopter
In this dissertation, the numerical simulation scheme developed by Yuan Lin et al.
[15-16] focused on the improvement of the hydrodynamic performance of the newly
developed AUH to enhance the hydrodynamic stability meanwhile reduce the drag force
during the surge motion. Using computational fluid dynamics solving the Reynolds
Averaged Navier–Stokes (RANS) equation adopting the RNG k-ε model, we observed
that with proper extension and sharpening of the rear part, the hydrodynamic drag and
the drag force fluctuation are remarkably reduced compared to the original axisymmetric
hull geometry, which is due to suppression of boundary layer separation. It is observed
numerically that for the basic geometry of AUH at u = 2 m/s, the time average drag of
HG3 (65.3 N) is 26.5% lower than that of HG1 (88.9 N), and the fluctuation on the drag,
side, and lift force can be greatly eliminated. The elevation of the hydrodynamic
performance agrees with the findings from the water-channel experiment on scaled HG1
and HG3 models. The improved geometry also shows good stability against disturbance
by the aquatic cabin attached at the central bottom, as well as the propellers embedded in
the hull in a realistic working condition compared to the original HG1 geometry. It should
be noted that the proposed HG3 geometry is not the optimal hull geometry accounting

16
for the surge motion of AUH, although the hydrodynamic performance has been
considerably improved. Furthermore, an improvement in the hull shape for surge motion
leads to uncertainties when the AUH performs the heave motion since the hull is
somewhat extended, which becomes non-axisymmetric. Further investigation is
necessary to optimize the hull geometry considering Fig. 3.1. 3D model of AUH with
horizontal propellers in through holes. Fig. 3.2. Streamlines seeing from the rear of the
vessel for the AUH geometry with two horizontal propellers at the surging velocity of 2
m/s, for (a) HG1 and (b) HG3, respectively; (c) the corresponding temporal evolution of
the drag force with solid lines for HG1 and dashed lines for HG3 from simulation; and
(d) standard deviation of the hydrodynamic force for HG1and HG3, respectively,
calculated from the hydrodynamic performance at both surge and heave directions.

Figure 3.1. 3D model of AUH with horizontal propellers in through holes [16]

Figure 3.2. Streamlines seeing from the rear of the vessel for the AUH geometry with
two horizontal propellers at the surging velocity of 2 m/s, for (a) HG1 and (b) HG3,
respectively; (c) the corresponding temporal evolution of the drag force with solid lines
for HG1 and dashed lines for HG3 from simulation; and (d) standard deviation of the
hydrodynamic force for HG1and HG3, respectively, calculated from (c) [16].

17
CHAPTER IV: DESCRIPTION OF PHYSICAL MODELS AND
NUMERICAL METHODOLOGY
4.1. Physical model

Because CFD simulation software is modular, it is necessary to select the correct


modules before simulating. We consider our problem to have the following two
characteristics:

- The problem considers the interaction between the glider and the water
environment: this is the interaction between a solid body with structure and fluid.

- In addition to the movement of water, the glider has its own motion.

From those comments, we see the need to use the modules: Laminar flow (laminar
flow, because the train is not running too fast and also to simplify the problem), Structure
(affecting the pressure of the water flow on the hull), and Moving Mesh (Ships move,
grid coordinates must also move).

The AUH is placed in an aquatic medium with an area H x W x L where H is the


height, W is the length and L is the width of the aquatic medium. The disc-shaped hull
was originally considered to be disc-shaped with the following dimensions: The disc-
shaped hull is one of the key designs of the Small-Autonomous Underwater Helicopter
(S-AUH). To achieve the goal of high maneuverability in the horizontal plane and good
stability in the vertical direction, the disc-shaped hull must be designed as a body of
revolution, which is the reason the hull is defined by a curve. After considering various
curves, Non-Uniform Rational B-Splines (NURBS) were chosen because of the
continuity of their curvature and their simple definitions, which makes the design and
adjustment. According to the conclusions, when H = 200 mm, h1 = 50 mm, h2 = 80 mm,
h3 = 2(H - h1)/3 + h1 = 150 mm, l1 = L/8 = 125 mm, l2 = L/4 = 250 mm, l3 = 3L/8 = 375
mm, the movement in the horizontal plane has low resistance. In addition, in the vertical
direction, it has good stability.

18
a)

b)
Figure 4.1. (a) Parameters of the AUH; (b) 3D model of the AUH

4.2. Mathematical formulation


4.2.1. Boundary layer
The mesh parameters required to adequately model the boundary layer were
estimated with Eqs. (4.1) and (4.2) [46], and the first layer's thickness can be expressed
as:


13 (4.1)
y  L y 80 Re 14

where Re is the Reynolds number defined as VL/γ, where γ is the coefficient of kinematic
viscosity, L is the diameter of the AUH, and 20 ≤ Δy+ ≤ 200. The thickness of the
boundary layer for the AUH was estimated with the following equation:

1 (4.2)
  0.035LRe 7

where δ denotes the thickness in the boundary layer, the first layer's 1 mm thickness was
determined with Eq. (4.1). The total thickness of the boundary layer was estimated with

19
Eq. (4.2) to be approximately 5 mm, hence, it was layered by five encrypted grids, as
shown in Fig. 4.2 (a) and (b).

a)

b)

Figure 4.2. Mesh type in computational fluid domains translating in the horizontal and
vertical planes and thicknesses of the boundary layer

4.2.2. Boundary conditions for translational simulations


These meshes transition in Fig. 4.2 is from refined to coarse moving away from the
AUH and extending out to form a cuboid computational domain and two ring-shaped
computational domains, respectively. In Fig. 4.3, the positions of the AUH relative to the
boundaries in the computational domains were sketched as a function of the AUH's length
L.

Numerical simulations with steady-state inflow were performed at drift angles in


the horizontal plane (attack angles in the vertical plane) of 0°, ± 10°, ± 20°, and ± 30°. In
all cases except for 0°, two inlet, and two outlet boundaries were used. In the case of 0°,
only one inlet and outlet boundary were used.

20
L L
L
L 3L

3L
3L

Figure 4.3. Computational domain adopted to simulate unsteady, 3-D, turbulent flow
for a translating AUH.

When the AUH was translating in the horizontal plane, different boundary
conditions were defined as follows. First, the hull of the AUH was the only solid
boundary in the computational fluid domain and was approximated as a wall with a no-
slip condition in all case studies. Second, when the AUH translated without AOD or
AOA, the +x boundary, which was three body lengths upstream from the AUH, was set
as the inlet boundary condition with an inflow velocity of 1–3 knots and a flow turbulence
of 5% [46]. Third, the outlet boundary condition with zero relative pressure was located
at the –x boundary behind the AUH. Fourth, free slip wall boundary conditions were
applied to the remaining four boundaries. When the AUH was translated with an AOD,
the inlet boundary conditions were positioned two such that the +x and +y boundaries (–
y boundary) with different inflow velocities were dependent on the AOD. Comparatively,
the –x and –y boundaries (+y boundary) opposite to the inlet was set to the outlet
boundary conditions with zero relative pressure. Free slip wall boundary conditions were
applied to the remaining two boundaries. Similarly, when the AUH is translated with
different AOA in the vertical plane, the above method was applied to set the boundary
conditions.

21
Table 4.1. Main AUH parameters

Parameter Symbol Physical Value


Unit

Diameter L m 1.0

Design depth h m 1000

Height H m 0.43-0.48

Mass m kg 50-150

Service speed V knot 1-3

Distance between the CG BG mm 37


and BG

Number of thrusters 2-4

4.2.3. Mathematical model of motion stability

In the kinematic formulas, two coordinate systems and four hydrodynamic angles
are represented. One is a body-fixed Cartesian coordinate system with x, y, and z-axes
and the origin fixed on the CG of the AUH, and the sign conventions for roll, pitch, and
yaw follow the right-hand rule relative to the x, y, and z-axes, respectively. The other
system is an Earth-fixed coordinate system with x0,y0, and z0-axes referenced by the flow
domain and motion trajectory. The z0-axis points down, completing the right-handed
Cartesian system. Three hydrodynamic angles were used in the study of the
hydrodynamic performance of the AUH, including the angle of attack (AOA), α = tan-1

22
(w/u) , angle of drift (AOD), β = tan-1 ( −v/ u), and angle of orientation (AOO),Φ = tan-1
( −v/−w).

The EOM for the AUH is formulated in terms of a body-fixed coordinate system,
but the coordinate origin O does not necessarily coincide with the CG or CB of the AUH.
For the linear motion of the AUH, the acceleration is computed by evaluating the sum of
all applied forces, acting through the CG. The foundation for the equations of motion is
Newton's second law:

F  m uGv (4.3)
where F is the net force vector acting on the AUH, m is the mass of the AUH, and uGV is
the acceleration of the CG.

4.3. Numerical methodology

According to our best knowledge, there are a lot of computational methods that have
been used to solve the problems related to chuyển động của AUH. In this dissertation, we
report briefly three important numerical methods such as the conservative level set
method, the arbitrary Lagrangian-Eulerian method [38-40], and the continuum surface
force [41].

The level set method was first proposed by Osher and Sethian in [42] as a simple
method to analyze and compute the motion of an interface. This method represents a
contour or interface as the zero-level set of a signed distance function called the level set
function (LSF). The motion of the interface is matched with the zero level set of the LSF
and the resulting initial value partial differential equation for the evolution of the LSF
resembles a Hamilton-Jacobi equation. Although the level set method is highly flexible,
it is limited by two numerical issues. First, the level set method does not implicitly
preserve the LSF as a distance function, which is necessary to estimate accurately
geometry. Secondly, the level set algorithm is slow because the time step is limited. To
solve these problems, especially for incompressible two-phase flow, we used the
conservative level set method which was developed by Olsson et al. [43-44].

The arbitrary Lagrangian-Eulerian is a finite element formulation in which the


computational system is not a prior fixed in mesh or attached to the material. ALE method
[45] was developed in an attempt to combine the advantages of both the Lagrangian
23
description and the Eulerian description while alleviating a lot of the drawbacks that the
traditional Lagrangian and Eulerian finite element methods. By using the ALE method
in engineering simulations, the computational mesh inside the domains of the fluid can
move arbitrarily to optimize the shape of elements. The mesh on the boundary and
interface of the fluid can move along with materials to other regions of various materials.

To be able to describe the motion of the ship, we use the Moving Mesh Module,
which is built based on the Arbitrary Lagrangian-Eulerian method. The mathematical
formulas developed by two mathematicians Lagrange and Euler are the basis for the finite
element method applied in CFD. The method developed by Euler is based on the absolute
coordinate system, also known as the laboratory coordinate system. The equations
formulated by Euler are particularly useful in simulating the behavior of fluids, magnetic
fields, and acoustics.

The method developed by Lagrange is based on a relative coordinate system. He


divided the experimental model into several small parts. When solving equations on each
element, he applied his own frame of reference associated with that element. His method
is often used in the simulation of mechanical structures. Because many phenomena need
to describe the interaction between many fields, ALE was born, this is a method that
integrates the two methods of the two scientists mentioned above, by building two
coordinate systems Mesh Coordinates and Geometry Coordinates- these two coordinate
systems allow us to handle simulation problems using both methods above. In different
regions to be treated by the appropriate method, two new coordinate systems are
introduced to keep the model consistent when using the two methods simultaneously.
When applying Moving Mesh (actually ALE) to the glider model, we easily encounter
an error: Inverted Mesh. When the ship moves at high speed, the net is also deformed,
creating an area where the mesh is pulled and compressed, at some point the net will be
“inverted” - a form of error in Topology. This phenomenon increases the error gradually,
leading to the inability to simulate due to Inconverged error.

4.4. Computational process

The transient motion of AUH is studied by numerical simulation. Figure 4.4


illustrates the calculation process. The goal of process modeling here is to perform a

24
simulation of a physical model with input conditions and obtain accurate results. In this
study, the computational process consists of three steps, namely pre-processing,
processing, and post-processing steps. In the preprocessing step, we need to draw a
physical model and then select the appropriate application modes. The processing step
includes setting the physical conditions (fluid properties, subdomain conditions,
boundary conditions, and initial conditions), creating the mesh, initializing the level set
function, solving the parameters, physics, and convergence testing. In the final step (post-
processing), we export the simulation data to get the results. To predict the migration of
AUH, we have applied two application levels in Ansys including motion mesh (ALE),
and fluent flow.

Preprocessing

Processing

Post-processing

Figure 4.4. The flow chart of the computational process

4.5. Mesh quality and convergence evaluation

The mesh density in the finite element model is an important problem. The
dependency of the element number on the simulation results is determined to ensure the
accuracy of the solution and minimize computational time. In general, a finer mesh with
a higher number of elements can capture complex geometries, stress gradients, and
material behavior more accurately, leading to more precise results. However, using a very

25
fine mesh can also increase computational time and resources, as more elements require
more calculations, resulting in longer simulation times. On the other hand, using a coarser
mesh with fewer elements can reduce computational time, but may lead to less accurate
results, especially in areas where stress gradients or material behavior change rapidly.

The optimal mesh density depends on various factors, including the complexity of
the geometry, the type of analysis (e.g., static, dynamic, transient), the material
properties, and the desired level of accuracy. It is typically determined through mesh
convergence studies, where the finite element analysis is performed with different mesh
densities, and the results are compared to assess convergence and accuracy.

A common approach to determining an appropriate mesh density is to start with a


relatively coarse mesh and gradually refine it until the solution converges, i.e., the results
exhibit minimal changes with further mesh refinement. This helps ensure that the
accuracy of the solution is maintained without unnecessary computational overhead.
Techniques such as adaptive mesh refinement, where the mesh is refined in regions of
interest based on error indicators, can also be used to optimize the mesh density and
minimize computational time.

It is important to strike a balance between accuracy and computational efficiency in


selecting the appropriate mesh density for a finite element model. A mesh that is too
coarse may yield inaccurate results, while a mesh that is too fine may result in excessive
computational costs. Conducting mesh convergence studies and using techniques such as
adaptive mesh refinement can aid in determining an appropriate mesh density that
balances accuracy and computational efficiency for a given simulation.

Table 4.2. The differences between three meshing objects

1 2 3
Element size (mm) 250 800 2500
Number of nodes 240057 124910 30049
Number of
832835 430156 115206
elements

 Mesh Models

26
In order to obtain accurate numerical simulations of liquid migration, a finite
element method is employed utilizing a mesh model that is appropriately designed for
the specific application. As per the current state of knowledge, the density of the mesh
must be optimized in the vicinity of the free interface to ensure the accuracy of the
numerical results. Therefore, in the present study, a range of different mesh models have
been developed and rigorously tested to identify the optimal mesh model. The quality of
each mesh model was evaluated using a parameter called the number of elements (NE),
which is a measure of the level of mesh density employed. Through the evaluation of NE,
the efficacy of different mesh models in capturing the key features of the liquid migration
process was assessed, allowing the identification of the most suitable mesh model for this
application. Overall, the process of optimizing the mesh density for the specific
application is essential to achieve the desired level of accuracy in the numerical
simulations.

 Convergence Evaluation

Convergence is defined as the ability to obtain a solution within a given tolerance


criteria. In our computation, the relative tolerance (Ar) and the absolute tolerance (Aa) are
selected to be 0.01 and 10-7, respectively. The time step is selected to be 0.5.

The convergence criteria can be written as

21/2
  Ej  
 1
 
N 
  1, (4.4)
j  Aaj  A r S j  
   

where S is the solution vector corresponding to the solution at a certain time step and E
is the error estimate of the solver in S.

 Accuracy, Reliability of Numerical Results

The determination of mesh density in a finite element model is a crucial aspect of


achieving accurate simulation results while minimizing computational time. This process
involves establishing the appropriate number of elements required to capture the relevant
physical phenomena of the system under investigation. By doing so, the accuracy of the
solution is ensured while computational resources are efficiently allocated. To this end,
the relationship between the element number and simulation results is carefully evaluated
27
to identify any dependencies. This evaluation process includes analyzing the effects of
increasing or decreasing the element number on key parameters of interest, such as the
drag coefficient. The results of such evaluations are instrumental in guiding the choice of
appropriate mesh density for a given simulation. In the specific case of the drag
coefficient, it has been observed that varying the number of elements has a minimal
impact on the final value obtained (Fig. 4.5).

Figure 4.5. Typical mesh used in the computational domain for (a) an AUH and (b) the
area around the object. The mesh is finer near the interface and inside the liquid.

28
CHAPTER V: HYDRODYNAMIC CHARACTERISTICS OF
THE AUTONOMOUS UNDERWATER HELICOPTER

In this chapter, the hydrodynamic characteristics of an autonomous underwater


helicopter are discussed and numerically investigated. These characteristics refer to the
way the vehicle interacts with water and how its design affects its performance in an
underwater environment. The two main hydrodynamic characteristics that are important
to consider are drag and lift. Drag is the force that opposes the motion of the vehicle
through water and can be affected by factors such as the shape of the vehicle and the
speed at which it is traveling. Lift, on the other hand, is the force that allows the vehicle
to rise and stay afloat in the water. Both of these characteristics are important to consider
in the design and operation of an underwater helicopter. The simulation results shown in
this chapter correspond to the governing equations as well as the physical conditions.

5.1. The hydrodynamic characteristics of the AUH

The hydrodynamic characteristics of an autonomous underwater helicopter refer to


how the vehicle interacts with water and how its design affects its performance in an
underwater environment. Some of the key hydrodynamic characteristics of an underwater
helicopter include drag and lift force. To study the effect of drag and lift force on the
AUH body, the selected parameters are at the angle of attack 00 and the moving speed vin
= 2m/s in the horizontal direction. As shown in Figure 5.1, the time average is calculated
with an oscillation of 70N, which is due to the periodic vortex shedding at the tail of the
AUH. The average values of the lateral force are close to zero, both with obvious
fluctuations. Because the upstream area of the AUH in the heave direction is much larger
than that in the sway direction, the amplitude of the lift force is larger than that of the
side.

As shown in Figure 5.1. for streamlines at the tail of the AUH hull, there is an
obvious vortex area at the tail, which causes fluctuant drag and lift forces. At the inlet
position, the fluid flow rate is equal to the 2 m/s input speed. The flow rate decreases as
it goes deeper into the channel and reaches its lowest value in the middle of the channel.
At the final position of the channel, the flow rate increases slightly above the mid-channel
position but is still lower than the inlet position. From this, it can be inferred that the flow

29
rate of the liquid is affected by the structure of the channel, especially the curve of the
channel. These values can also be used to calculate other parameters such as drag or lift
acting on the vehicle moving in the fluid.

Figure 5.1. The velocity streamlines at the tail of the AUH

Drag is the resistance that a vehicle experiences as it moves through the water. The
shape, size, and surface roughness of the underwater helicopter's body, wings, and rotors
can significantly affect the drag it experiences. Minimizing drag is important for reducing
energy consumption and increasing the vehicle's speed and maneuverability. Lift is the
force that allows the underwater helicopter to generate upward or downward motion. The
design of the underwater helicopter's wings or rotors, as well as their orientation and
motion, affect the lift generated. Efficient lift characteristics are essential for the
underwater helicopter to achieve stable and controlled motion.

As shown in Figure 5.2, the velocity of the AUH begins to decrease over time, and
at about 8s, the velocity reaches its lowest point of about 0.25 m/s. This shows that drag
and air resistance are increasingly affecting the motion of the AUH as it moves across
the water. After that, the AUH continues to decelerate, but at a slower rate. This shows
that the air resistance and drag forces have been significantly reduced. At about 30s, the
velocity of the AUH decreases at a slower rate and is at 0.16 m/s. As the velocity of the
AUH is decreasing, the forces of resistance will have less and less influence on its motion,
and at some points, the velocity of the AUH will no longer be the same, this is known as
reaching steady state determined.
30
t = 10s t = 20s

t = 30s t = 40s

Figure 5.2. Streamlines of AUH in plane xoz at a velocity of 2 m/s and an AOA of 0 degrees

Figure 5.3 shows that drag increases linearly with velocity, while lift forces increase
faster with velocity. At 2 m/s, the max drag is about 250 N, while the lift is about 380 N.
This shows that, at this input speed, the lift still plays a major role in preventing
movement. displacement of AUH, and the drag force is still quite small. However, when
increasing the speed or changing the angle of attack, the force will increase significantly
and can become the dominant force affecting the movement of the AUH. This issue will
be discussed in sections 5.2 and 5.3.

Figure 5.3. Drag and lift force on the plane xoz value over time
31
In Figure 5.4, we can see a pressure graph representing the pressure at points on the
surface of the AUH. In general, the pressure will vary depending on the location and
shape of the AUH, as well as the velocity and direction of the flow. In front of the AUH,
we see an area of lower pressure, called the low-pressure area. This is the area where the
flow is moved through the AUH at a higher velocity than behind, resulting in a pressure
drop in this region. Meanwhile, at the rear of the AUH, we see an area of higher pressure,
called the high-pressure area. This is the area where the flow is blocked by the AUH,
leading to an increase in pressure in this region. This region of low and high pressure has
a significant effect on the performance of the AUH. The low-pressure zone at the front
can help pull the flow forward of the AUH, creating a forward drag, which increases the
flow rate. Meanwhile, the high-pressure zone at the back can create a forward thrust,
helping to maintain the flow rate and reduce drag.

Figure 5.4. Pressure distribution and streamlines of AUH in the plane at a velocity of 2 m/s

Understanding and optimizing the hydrodynamic characteristics of an autonomous


underwater helicopter is critical for its efficient and effective operation. Computational
fluid dynamics simulations, model testing, and empirical data analysis are commonly
used methods to study and optimize the hydrodynamic characteristics of underwater
vehicles, including helicopters, to ensure their reliable and safe performance in
underwater environments.

The current study aims to examine and evaluate the effectiveness of two aquatic
transducers with varying lengths that are mounted on the bottom of the Autonomous

32
Underwater Hull. A comprehensive three-dimensional Computational Fluid Dynamics
analysis was conducted to compare the performance of two different AUH design
configurations. The results revealed that the streamlines in the proximity of the hull of
the AUH with a 50-mm-long aquatic transducer were superior to those in the AUH with
a 100-mm-long aquatic transducer, particularly in pressure and velocity fields.
Consequently, this study includes an analysis of both AUH configurations equipped with
a 50-mm and 100-mm long aquatic transducer.

A full 3-D Computational Fluid Dynamics analysis was conducted to investigate


the flow patterns near the hull of the AUH with 50-mm and 100-mm-long aquatic
transducers. The results revealed that the streamlines near the hull of the AUH with a 50-
mm-long aquatic transducer were more favorable than those near the AUH with a 100-
mm-long aquatic transducer. Additionally, the pressure field showed significant
improvement in the AUH with a 50-mm-long aquatic transducer. Hence, both the AUHs
mounted with the 50-mm and 100-mm-long aquatic transducers were studied in this
investigation, and the performance of each was analyzed (Fig. 5.5). The findings of this
study have significant implications for optimizing the design of AUHs and improving
their performance in pressure fields.

a) b)

Figure 5.5. a) Pressure on the AUH with the 50-mm aquatic transducer. b) Pressure on
the AUH with the 100-mm aquatic transducer.

To this end, we conducted a full 3-D Computational Fluid Dynamics analysis of


two AUH design configurations equipped with a 50-mm and 100-mm long aquatic
transducer, respectively. The results revealed that the streamlines near the hull of the

33
AUH with a 50-mm-long aquatic transducer were superior to those near the hull of the
AUH with a 100-mm-long aquatic transducer in terms of velocity fields. Moreover, the
simulation results indicated that the use of a shorter transducer length leads to higher
velocity values compared to the longer transducer length (Figure 5.6). These findings
provide important insights for the design and optimization of aquatic vehicles equipped
with transducers for measuring water parameters.

a) b)

Figure 5.6. a) Velocity around the AUH with a 50-mm aquatic transducer and b)
Velocity around the AUH with a 100-mm aquatic transducer

5.2. The effect of inclined angles

The angle of attack of the underwater helicopter can impact the drag it experiences.
When the vehicle is inclined, the distribution of forces on its body, wings, or rotors
changes, resulting in different drag forces acting on the vehicle. This can affect the overall
resistance the vehicle experiences as it moves through the water, impacting its energy
consumption, speed, and maneuverability. Higher AOA can result in increased drag,
while lower inclined angles may result in reduced drag.

Figure 5.7 shows the variation of drag force with the angle of attack for an AUH
operating at a speed of 2 m/s. As the AOA increases, the drag coefficient also increases.
At higher AOAs, the flow over the surface of the AUH becomes more turbulent, resulting
in greater resistance to motion through the water. Figure 5.7 also shows that the drag
force reaches a peak value at an AOA of around 20 degrees before dropping slightly at
higher angles. This is due to the formation of a separated flow region behind the AUH,
which creates a low-pressure zone that reduces drag. However, at even higher AOAs, the
34
flow fully separated, resulting in a significant increase in drag once again. The drag of an
AUH is an important parameter to consider as it affects the vehicle's energy consumption
and overall performance. Higher drag requires more power to maintain the same speed,
which can limit the range and endurance of the vehicle. Therefore, it is important to
optimize the design of the AUH to minimize drag and improve its efficiency in operation.

Figure 5.7. Drag force and AOA at vi = 2m/s

The AOA can also affect the lift generated by the underwater helicopter. For
vehicles that rely on wings or rotors for lift, the AOA can impact the direction and
magnitude of the lift forces. Higher inclined angles can result in changes in lift forces,
potentially affecting the vehicle's stability, control, and performance. Properly managing
the lift forces is essential for maintaining stable and controlled motion.

Figure 5.8 shows the lift coefficient of the AUH at a speed of 2 m/s with the angle
of attack (AOA) varying from 0 to 20 degrees. The lift coefficient is a dimensionless
parameter that represents the lift generated by an object relative to the object's size and
fluid properties. As the AOA increases from 0 to 20 degrees, the lift coefficient initially
increases linearly. This is because, at small angles of attack, the airflow over the AUH is
relatively smooth and generates lift in a predictable manner. However, as the AOA
increases beyond 10 degrees, the lift coefficient starts to deviate from the linear trend and
reaches a maximum value of approximately 1200 N at an AOA of 20 degrees. The
35
deviation from the linear trend can be attributed to the formation of turbulent flow over
the AUH. As the AOA increases, the airflow over the vehicle becomes more complex,
with pockets of high and low pressure forming around the vehicle's surfaces. These
pressure variations can result in the formation of turbulent airflow, which can lead to
changes in the lift generated by the vehicle. Properly managing the lift generated by the
AUH is critical for maintaining stable and controlled motion. As the lift coefficient
increases, the AUH generates more lift, potentially impacting its stability and control.
Therefore, it is essential to design the vehicle with an appropriate AOA range that allows
for efficient lift generation while maintaining stability and control.

Figure 5.8. Lift force and AOA at vi = 2m/s

In Figure 5.9, which displays velocity streamlines for different angle of attack
(AOA) cases, it is evident that the vortex area varies significantly with changes in AOA.
The vortex area refers to the region of swirling fluid flow generated around the AUH.
Upon closer examination, it can be observed that as the AOA increases from 0 degrees
to 20 degrees, the size and intensity of the vortex area also increase. At lower AOAs,
such as 0 and 5 degrees, the vortex area appears relatively smaller and less pronounced.
However, as the AOA reaches 10 and 15 degrees, the vortex area expands and becomes
more distinct. At the highest AOA of 20 degrees, the vortex area is the largest and exhibits
a more complex pattern. Based on these observations, it can be concluded that a higher
36
AOA, specifically 20 degrees, generates the most prominent vortex area. This implies
that at this AOA, the AUH experiences stronger vortices and potentially higher levels of
turbulence. The presence of vortices can have significant effects on the motion and
stability of the AUH. While a certain level of controlled vortex shedding can be beneficial
for maneuverability, excessive turbulence, and unsteady flows can adversely impact the
vehicle's performance and controllability. Therefore, considering the goal of maintaining
a stable straight motion for better inspection near the seabed, it may be preferable to
operate the AUH at a lower AOA, such as 0 or 5 degrees. These lower AOAs exhibit
smaller and less intense vortex areas, indicating reduced turbulence and potentially more
stable motion. However, it is important to note that this conclusion is based solely on the
analysis of the vortex area, and further investigation, including other performance
parameters, is necessary to determine the optimal AOA for the motion of the AUH.

AOA = 0o AOA = 5o

AOA = 10o AOA = 15o

AOA = 20o

Figure 5.9. Streamlines at xoz plane (right) of AUH at vi = 2 m/s and AOA = 0o – 20o
from numerical simulation

37
In Figure 5.10, which depicts pressure distributions for different angle of attack
(AOA) cases, it is evident that the pressure patterns vary significantly with changes in
AOA. The pressure distribution provides valuable insights into the aerodynamic
characteristics of the AUH and can help determine the optimal AOA for its motion.
Analyzing the images, we can observe that as the AOA increases from 0 degrees to 20
degrees, the pressure distribution undergoes notable changes. At lower AOAs, such as 0
and 5 degrees, the pressure distribution appears relatively uniform, with minimal
variations across the AUH's surface. However, as the AOA increases to 15 and 20
degrees, distinct pressure variations emerge. Areas of higher pressure are observed near
the leading edge of the AUH, while regions of lower pressure are present toward the
trailing edge. Based on these observations, it can be concluded that a higher AOA,
specifically 20 degrees, generates the most pronounced pressure variations. This indicates
a significant change in the flow behavior around the AUH and suggests the presence of
a stronger pressure differential between the upper and lower surfaces of the vehicle. Such
pressure differentials contribute to the generation of lift forces that enable the AUH to
maintain its buoyancy and maneuverability in the water. Therefore, considering the goal
of achieving stable motion and control near the seabed, it may be advantageous to operate
the AUH at an AOA of 20 degrees. This AOA provides a distinct pressure distribution
that facilitates the generation of lift forces while maintaining a manageable level of
pressure variations. However, it is crucial to note that this conclusion is based solely on
the analysis of pressure distributions, and other factors, such as drag, lift-to-drag ratio,
and stability, should be considered in conjunction with the pressure data to determine the
optimal AOA for the motion of the AUH.

38
AOA = AOA =

AOA = 10o AOA = 15o

AOA = 20o

Figure 5.10. Pressure distribution and streamlines of AUH in plane xoz at vi = 2 m/s
AOA = 0o – 20o from numerical simulation

In the study conducted by Y. Lin et al., the hydrodynamic performance of a disk-


shaped autonomous underwater helicopter was investigated by experiments. The goal
was to improve the performance of the AUH in terms of lift and drag by modifying its
shape locally. In this analysis, we compare the numerical results obtained from
computational fluid dynamics with the experimental results in HG1. Specifically, we
compare the drag and lift coefficients of the AUH at a constant speed of 2 m/s with the
angle of attack varying from 0 to 20 degrees. The comparison allows us to evaluate the
accuracy of the simulation model and the effectiveness of the proposed shape
modification method in improving the hydrodynamic performance of the AUH.

Figure 5.11 shows the comparison of the drag force of the disk-shaped autonomous
underwater helicopter between experimental and simulation results. The results are
obtained by varying the angle of attack (AOA) from 0 to 20 degrees at a speed of 2 m/s.
The experimental results are represented by the blue line, while the red line represents

39
the simulation results. As can be observed, both the experimental and simulation results
show a similar trend, with the drag coefficient increasing as the angle of attack increases.
The comparison shows that the simulation results are in good agreement with the
experimental results for the entire range of AOA, with a maximum difference of around
6% at an AOA of 20 degrees. The agreement between simulation and experimental results
indicates that the simulation model used is reliable and accurate in predicting the
hydrodynamic performance of the AUH at different AOAs. The results can be useful for
optimizing the design of the AUH and improving its performance in practical
applications. However, the simulation results tend to overestimate the drag coefficient
compared to the experimental results. This discrepancy may be attributed to
simplifications or assumptions made in the numerical modeling, which may not fully
capture the complex fluid dynamics around the AUH. Nevertheless, the general trend
observed in both results is useful in understanding the hydrodynamic performance of the
AUH and optimizing its design for better efficiency and stability.

Figure 5.11. Comparison of simulation results and experimental results in terms of drag force

Figure 5.12 compares the lift force of the AUH obtained from simulations and
experiments conducted by Lin et al. [15]. The lift coefficient is a dimensionless parameter
that describes the amount of lift generated by an object for a given angle of attack. The
graph shows that the simulation results are in good agreement with the experimental
results. At lower angles of attack, the lift coefficient increases gradually and then rapidly

40
increases beyond 10 degrees. This rapid increase in lift coefficient is due to the formation
of vortices near the trailing edge of the disk, which enhances the lift generation. As the
AOA increases, the lift coefficient also increases for both simulation and experimental
results. However, the magnitude of the lift coefficient obtained from the simulation
results is generally higher than that of the experimental results for all AOAs. The
difference in the lift coefficients between the simulation and experimental results may be
attributed to several factors, such as the simplifications and assumptions made in the
numerical model, the limitations of the experimental setup, and the accuracy of the
measurement instruments used. For instance, the numerical model may not fully capture
the complex flow phenomena and turbulence that occur around the body, while the
experimental setup may not perfectly replicate the real-life conditions that the AUH
operates in. Overall, while there is some discrepancy between the simulation and
experimental results, the general trend and behavior of the lift coefficient with varying
AOA are captured by both methods. The comparison of these results provides valuable
insights into the performance of the AUH and can be used to improve the design and
optimization of future underwater vehicles.

Figure 5.12. Comparison of simulation results and experimental results in terms of lift force.

41
5.3. The effect of velocities

As the velocity of the underwater helicopter increases, the drag it experiences also
increases. The drag force acting on a body moving through a fluid, such as water, is
proportional to the square of its velocity, as given by the drag equation. This means that
as the velocity of the AUH increases, the drag force acting on it increases at a much faster
rate, making it more difficult for the AUH to move through the water. The increase in
drag can have a significant impact on the performance of the AUH. It can lead to
increased energy consumption, as the vehicle has to overcome the additional resistance
from the water. This can reduce the overall range and endurance of the vehicle. The
increase in drag can also reduce the maximum speed that the AUH can achieve, making
it slower and less efficient. In addition, higher drag can also affect the maneuverability
of the vehicle. It can make it more difficult to turn or change direction, as the drag force
acting on the vehicle resists the change in its motion. This can impact the ability of the
AUH to perform tasks such as underwater inspections or search and rescue missions.
Moreover, the increased drag can also result in increased wear and tear on the vehicle's
components, such as the propulsion system and control surfaces. This can lead to
increased maintenance costs and potentially shorten the lifespan of the AUH. Therefore,
it is important to consider the effect of drag when designing and operating underwater
vehicles like the AUH, in order to optimize their performance and efficiency

Figure 5.13 shows the relationship between velocity and drags force experienced
by an object moving through water. As the velocity increases, the drag force experienced
by the object also increases. This is in line with the drag equation, which states that drag
is directly proportional to the square of the velocity. The effects of velocity and drag are
critical in understanding the performance of objects moving through water, such as an
underwater helicopter. At low velocities, the drag force is relatively small, allowing for
efficient movement with lower energy consumption. As the velocity increases, however,
the drag force increases significantly, requiring more energy to maintain the same speed.
This relationship between velocity and drag force means that higher velocities can
significant impact the energy consumption and overall performance of the underwater
helicopter.

42
Furthermore, higher drag force can also impact the maneuverability of the
underwater helicopter. The drag force creates resistance to the vehicle's movement,
making it harder to change direction or make sudden movements. This resistance can also
result in increased wear and tear on the vehicle's components, reducing its lifespan and
increasing maintenance requirements. In conclusion, the graph above highlights the
critical role that velocity and drag play in the performance of objects moving through
water. As the velocity increases, the drag force also increases, impacting energy
consumption, speed, maneuverability, and the lifespan of the vehicle's components.
Understanding this relationship is crucial for optimizing the design and performance of
underwater vehicles.

Figure 5.13. The drag force at different surging velocities for hull geometries
calculated from numerical simulation

The velocity of the underwater helicopter can affect the lift it generates. For vehicles
that rely on wings or rotors for lift, higher velocities can result in increased lift due to the
Bernoulli principle, which states that as the velocity of a fluid increases, its pressure
decreases. This can result in increased lift forces, allowing the underwater helicopter to
achieve higher speeds or carry heavier payloads. However, it's important to note that the
relationship between velocity and lift is not always straightforward, and there can be other
factors that impact lift generation, such as the shape and orientation of the vehicle, as
well as the properties of the fluid it's operating in. Additionally, if the vehicle is moving
43
too quickly, it may experience negative effects from drag, which can reduce its overall
performance. Therefore, it's important to carefully consider the impact of velocity on lift
when designing and operating underwater vehicles. Figure 5.14 shows the relationship
between the velocity of the underwater helicopter and the lift force it generates. As the
velocity increases, the lift force also increases, which is consistent with the Bernoulli
principle. At lower velocities, the lift force is relatively constant, but as the velocity
increases, the lift force grows at an increasing rate. The Bernoulli principle is a
fundamental concept in fluid dynamics that explains the relationship between the velocity
of a fluid and its pressure. According to the principle, as the velocity of a fluid increases,
its pressure decreases. This can be seen in the case of an underwater helicopter, which
relies on lift forces generated by the motion of its wings or rotors through the water. As
the velocity of the vehicle increases, the pressure of the water decreases around the wings
or rotors, resulting in an increase in lift force.

Figure 5.14. The lift force at different surging velocities for hull geometries calculated
from numerical simulation

The effect of velocity on lift force has important implications for the performance
of an underwater helicopter. At higher velocities, the vehicle can generate more lift force,
which can be used to achieve higher speeds or carry heavier payloads. However, it is
important to note that there are limits to the amount of lift force that can be generated by

44
a given wing or rotor design, and increasing velocity beyond a certain point may result
in diminishing returns or even a decrease in lift force due to factors such as drag.

Figure 5.15 shows the velocity streamline around an underwater helicopter at


different velocities. As the velocity increases, the velocity streamline becomes more
tightly packed, indicating a higher velocity flow around the AUH. At low velocities, the
flow is relatively uniform, with no major changes in direction or speed. However, at
higher velocities, the flow becomes more complex, with eddies and vortices forming
around the AUH. The velocity streamline can affect the performance of the AUH in
several ways. At higher velocities, the flow around the AUH can become turbulent,
increasing the drag on the vehicle and reducing its efficiency. Turbulence can also lead
to instability and difficulty in controlling the vehicle. On the other hand, the formation of
vortices can create lift forces, which can be beneficial for the AUH's performance,
allowing it to achieve higher speeds or carry heavier payloads. Overall, the velocity
streamline is an important factor in understanding the effects of velocity on the
performance of an AUH. As the velocity increases, the flow around the vehicle becomes
more complex and can have both positive and negative effects on its performance.

vi = 2 m/s vi = 1,5 m/s

vi = 1 m/s vi = 0,5 m/s

Figure 5.15. Streamlines at xoz plane (right) of AUH at vi = 0,5 - 2 m/s from numerical simulation

45
Figure 5.16 shows the effects of velocity on pressure distribution along the surface
of an AUH. As the velocity of the AUH increases, the pressure distribution changes, with
certain areas experiencing a decrease in pressure and others experiencing an increase in
pressure. At low velocities, the pressure is relatively evenly distributed along the surface
of the AUH, with a slight increase in pressure near the nose of the vehicle. As the velocity
increases, the pressure at the nose of the vehicle said, while the pressure near the tail of
the vehicle increases. This change in pressure distribution is due to the formation of a
boundary layer of water flowing over the surface of the AUH. At higher velocities, the
boundary layer becomes thicker and the water flowing over it becomes turbulent,
resulting in a decrease in pressure at the nose and an increase in pressure at the tail. This
change in pressure distribution can have significant effects on the performance of the
AUH. If the pressure at the nose of the vehicle is too low, it may not be able to maintain
its desired depth or control its direction of movement. On the other hand, if the pressure
at the tail of the vehicle is too high, it may cause excessive drag, slowing the vehicle
down and increasing its energy consumption. Understanding the effects of velocity on
pressure distribution is therefore critical in designing and operating AUHs effectively and
efficiently. By carefully controlling the velocity of the vehicle, engineers can optimize
its performance and ensure it operates as intended.

vi = 2 m/sm/s vi = 1,5 m/s

vi = 1 m/s vi = 0,5 m/s


m/s

Figure 5.16. Pressure distribution and streamlines of AUH in plane xoz at vi = 0,5 - 2 m/s
from numerical simulation
46
CHAPTER VI: CONCLUSION AND FUTURE WORKS

6.1. Conclusion
A autonomous underwater helicopter with a disk-shaped hull that has been
developed by Wang et al. [11] to the improvement of the hydrodynamic performance of
the newly developed AUH to enhance the hydrodynamic stability while reducing the drag
force during the surge motion. The motion of AUH has been evaluated numerically. The
motion of the AUH is determined by the boundary conditions. These factors can interact
with and interact with the AUH, affecting its stability and performance. To model the
movement of the AUH, the AUH was placed in a computational domain, the positions of
the AUH relative to the boundaries in the computational domains were sketched as a
function of the AUH's length L. Numerical simulations with steady state inflow were
performed at drift angles in the horizontal plane (attack angles in the vertical plane) of
0°, ± 10°, ± 20°, and ± 30°. And with the velocity value inlet is 2 m/s, 1,5 m/s, 1 m/s, and
0,5 m/s.

In conclusion, this study focused on investigating the motion and hydrodynamic


performance of the Autonomous Underwater Helicopter. Several key aspects were
addressed, including the establishment of complete physical models and boundary
conditions, the effect of velocity on the hydrodynamic characteristics of the AUH during
motion, and the impact of angle characteristics on its hydrodynamics. Through the
establishment of comprehensive physical models and boundary conditions, this study
provided a solid foundation for analyzing the hydrodynamic behavior of the AUH. The
accurate representation of the physical environment and governing equations allowed for
a realistic simulation of the AUH's motion and interaction with the surrounding fluid.

The investigation into the effect of velocity on the hydrodynamic characteristics


of the AUH yielded valuable insights. It was observed that as the velocity increased, the
drag experienced by the AUH also increased. This finding is consistent with the drag
equation, which states that drag is proportional to the square of the velocity. The
increased drag can impact the AUH's overall performance, energy consumption, speed,
and maneuverability. Proper management of drag is crucial for optimizing the AUH's
operation.

47
Furthermore, the study explored the effect of angle characteristics on the
hydrodynamics of the AUH during motion. It was observed that variations in the angle
of attack (AOA) had a significant influence on the lift generated by the AUH. Higher
likely angles in changes in lift forces, potentially affecting stability, control, and overall
performance. Finding the optimal AOA is essential for maintaining stable and controlled
motion.

In addition to the aforementioned aspects, this study also investigated the effects
of velocity on the streamline patterns and pressure distribution around the AUH during
motion. The analysis of velocity streamline revealed that as the velocity of the AUH
increased, the patterns of the streamline underwent significant changes. Higher velocities
led to the formation of more pronounced vortices and alterations in the flow patterns.
This understanding is crucial as it provides insights into the flow behavior and helps in
assessing the stability and maneuverability of the AUH. Moreover, the examination of
pressure distribution offered valuable insights into the hydrodynamic performance of the
AUH. The comparison of pressure contours at different velocities and angle
characteristics highlighted variations in the distribution and magnitude of the pressure.
Higher velocities and specific angles of attack resulted in distinct pressure patterns,
indicating areas of high and low pressure around the AUH. Understanding these pressure
variations is essential for optimizing the AUH's design and control strategies, as they
directly impact the lift, drag, and overall stability of the vehicle. By considering the
effects of velocity on streamlined patterns and pressure distribution, this study provides
a comprehensive understanding of the hydrodynamic characteristics of the AUH. These
insights enable engineers and researchers to make informed decisions regarding the
design, operation, and control of AUHs. By manipulating velocity and angle
characteristics, it becomes possible to optimize the streamline patterns and pressure
distribution to enhance the AUH's motion, stability, and overall performance. Such
knowledge is invaluable for improving the efficiency, maneuverability, and mission
success of AUHs in various underwater applications.

The study highlights the interplay between velocity, streamline patterns, pressure
distribution, and the hydrodynamic performance of the AUH. By investigating these
effects, valuable insights have been gained, enabling the optimization of the AUH's

48
motion, stability, and overall performance. The findings provide a foundation for further
research and development in the field of autonomous underwater vehicles, contributing
to the advancement of underwater exploration, inspection, and surveillance capabilities.
the simulation results are in good agreement with the experimental results, with the
difference between the two being less than 10% [15]. At lower angles of attack, the lift
coefficient increases gradually and then rapidly increases beyond 10 degrees. This rapid
increase in lift coefficient is due to the formation of vortices near the trailing edge of the
disk, which enhances the lift generation.

Overall, this study sheds light on the crucial aspects of the AUH's motion and
hydrodynamic performance. The findings highlight the importance of considering
velocity and angle in optimizing the AUH's operation and enhancing its efficiency. The
knowledge gained from this research can contribute to the development of improved
design strategies, control algorithms, and operational guidelines for autonomous
underwater helicopters. Further research in this field holds great potential for advancing
the capabilities and applications of AUHs in various underwater missions and
inspections.

6.2. Future works

The study on the motion and hydrodynamic performance of the autonomous


underwater helicopter has provided valuable insights into its behavior and characteristics.
However, there are several areas that warrant further exploration and research. The
following are some potential avenues for future work in this field:

- Experimental Validation: While this study primarily focused on numerical


simulations, it is important to complement these findings with experimental data.
Conducting physical experiments to validate the numerical results would enhance the
reliability and accuracy of the findings. This could involve testing different AUH
prototypes in controlled water tank environments or even in real-world scenarios to
gather comprehensive data on its motion and hydrodynamic performance.

- Optimization of AUH Design: The findings of this study provide insights into
the hydrodynamic characteristics of the AUH, but there is potential for further
optimization of its design. Future research can explore innovative design modifications

49
to improve the AUH's efficiency, maneuverability, and performance. This may include
exploring alternative shapes, propulsion systems, or materials that can reduce drag,
enhance lift, and improve overall hydrodynamic efficiency.

In conclusion, the study on the motion and hydrodynamic performance of the


AUH opens up several avenues for future research. Further experimental validation,
advanced control strategies, optimization of AUH design, multi-objective optimization,
real-time monitoring, and environmental considerations are among the key areas that can
be explored. Continued research in these areas will contribute to the advancement of
autonomous underwater vehicles and their effective utilization in various underwater
applications.

50
BIBLIOGRAPHIES

[1] P.E. Hagen, N.J. Storkersen, K. Vestgard, “HUGIN-use of UUV technology in marine
applications,” OCEANS'99 MTS/IEEE (1999) 1-6.

[2] Bondaryk J. E., “Bluefin autonomous underwater vehicles: programs, systems, and
acoustic issues,” Journal of the Acoustical Society of America.15 (2004) 2615.

[3] P. Jagadeesh, K. Murali, V.G. Idichandy, “Experimental investigation of


hydrodynamic force coefficients over AUV hull form,” Ocean Engineering 36 (2009)
113-118.

[4] G. Neves, M. Ruiz, J. Fontinele, L. Oliveira, “Rotated object detection with forward-
looking sonar in underwater applications,” Expert Systems with Applications 140 (2020)
111870.

[5] J. Sherman, R.E. Davis, W.B. Owens, J. Valdes, “The autonomous underwater glider
Spray,” IEEE Journal of Oceanic Engineering 26 (2001) 437-446.

[6] H. Stommel, “THE SLOCUM MISSION,” Oceanography 02 (2015) 22-25.

[7] X. Xiang, C. Yu, Z. Niu, Q. Zhang, “Subsea Cable Tracking by Autonomous


Underwater Vehicle with Magnetic Sensing Guidance,” Sensors 16 (2016) 1335-1357.

[8] R. B. Wynn, V. A. I. Huvenne, T. P. L. Bas, B. J. Murton, D. P. Connelly, B. J. Bett,


H. A. Ruhl, K. J. Morris, J. Peakall, D. R. Parsons, E. J. Sumner, S. E. Darby, R. M.
Dorrell, J. E. Hunt, “Autonomous Underwater Vehicles (AUVs): Their past, present and
future contributions to the advancement of marine geoscience,” Marine Geology 352
(2014) 451-468.

[9] C. A. Woolsey, “Review of Marine Control Systems: Guidance, Navigation, and


Control of Ships, Rigs and Underwater Vehicles,” Journal of Guidance, Control, and
Dynamics 28 (2005) 574-575.

[10] A. Phillips, S. R. Turnock, M. Furlong, “The Use of Computational Fluid Dynamics


to Aid Cost-Effective Hydrodynamic Design of Autonomous Underwater Vehicles,”

51
Proceedings of the Institution of Mechanical Engineers, Part M: Journal of Engineering
for the Maritime Environment. 224 (2010) 1-16.

[11] P. Du, S. H. Huang, W. Yang , Y. Wang , Z. Wang, R. Hu, Y. Chen, “Design of a


Disc-Shaped Autonomous Underwater Helicopter with Stable Fins,” Journal of Marine
Science and Engineering 10 (2022) 67-83.

[12] Z. Wang, X. Liu, H. Huang, Y. Chen, “Development of an Autonomous Underwater


Helicopter with High Maneuverability,” Special Issue Underwater Robots in Ocean and
Coastal Applications 9 (2019) 4072.

[13] D. Sward, J. Monk, N. Barrett, “A Systematic Review of Remotely Operated Vehicle


Surveys for Visually Assessing Fish Assemblages,” Sec. Deep-Sea Environments and
Ecology 6 (2019) 134.

[14] C.-W. Chen, Y. Jiang, H.-C. Huang, D.-X. Ji, G.-Q. Sun, Z. Yu, Y. Chen,
“Computational fluid dynamics study of the motion stability of an autonomous
underwater helicopter,” Ocean Engineering 143 (2017) 227-239.

[15] Y. Lin, J. Guo, H. Li, H. Zhu, H. Huang, Y. Chen, “Study on the Motion Stability of
the Autonomous Underwater Helicopter,” Journal of Marine Science and Engineering
10 (2022) 60-80.

[16] Y. Lin, J. Guo, H. Li, Z. Wang, Y. Chen, H. Huang, “Improvement of hydrodynamic


performance of the disk-shaped autonomous underwater helicopter by local shape
modification,” Ocean Engineering 260 (2022) 1-12.

[17] H. Zhang, Y.-r. Xu, H.-p. Cai, “Using CFD software to calculate hydrodynamic
coefficients,” Journal of Marine Science and Application 9 (2010) 149-155.

[18] S. Tang, T. Ura, T. Nakatani, B. Thornton, T. Jiang, “Estimation of the


hydrodynamic coefficients of the complex-shaped autonomous underwater vehicle
TUNA-SAND,” Journal of Marine Science and Technology 14 (2009) 373-386.

[19] K. Kim, H. S. Choi, “Analysis on the controlled nonlinear motion of a test bedAUV–
SNUUV I,” Ocean Engineering 34 (2007) 1138–1150.

52
[20] Minnick, L. Marie, “A Parametric Model for Predicting Submarine Dynamic
Stability in Early Stage Design,” Virgnia Tech, Virginia, URN: etd-04262006-231952.

[21] Dajka, “Software Design Document for a Six DOF Unsteady Simulation Capability
in ANSYS-CFX,” ANSYS 2006 Canada Ltd.

[22] T.-L. Le, J.-C. Chen, B.-C. Shen, F.-S. Hwu and H.-B. Nguyen, “Numerical
investigation of the thermocapillary actuation behavior of a droplet in a microchannel,”
Int. J. Heat Mass Transfer 83 (2015) 721-730.

[23] T.-L. Le, J.-C. Chen, F.-S. Hwu and H.-B. Nguyen, Numerical study of the migration
of a silicone plug inside a capillary tube subjected to an unsteady wall temperature
gradient,” Int. J. Heat Mass Transfer 97 (2016) 439-449.

[24] T.-L. Le, J.-C. Chen, and H.-B. Nguyen, “Numerical study of the thermocapillary
droplet migration in a microchannel under a blocking effect from the heated wall,” Appl.
Thermal Eng. 122 (2017) 820-830.

[25] T.-L. Le, J.-C. Chen, and H.-B. Nguyen, “Numerical investigation of the forward
and backward thermocapillary motion of a water droplet in a microchannel by two
periodically activated heat sources,” Numerical Heat Transfer, Part A: Applications
79(2) (2021) 146-162.

[26] T.-L. Le, N. T. Tien, “A CFD study on hydraulic and disinfection efficiencies of the
body sterilization chamber,” Annals of the Romanian Society for Cell Biology 25(2)
(2021) 3998-4004.

[27] T.L. Le, T.D. Hong, “Computational fluid dynamics study of the hydrodynamic
characteristics of a torpedo-shaped underwater glider,” Fluids 6 (2021) 252.

[28] T.L. Le, T.T. Nghia, H.D. Thong and M.H.K. Son, “Numerical study of aerodynamic
performance and flow characteristics of a centrifugal blower,” International Journal of
Intelligent Unmanned Systems (2022).

[29] T.-H. Tran, T.-L. Le, D.-T. Hong and T.-P. Nguyen, “The effect of inlet velocities
on the droplet size in T-junction microfluidic devices,” JP Journal of Heat and Mass
Transfer 28 (2022) 1-14.
53
[30] T.-L Le, T.T. Nguyen, T.T. Kieu, “A CFD Study on the Design Optimization of
Airborne Infection Isolation Room,” Mathematical Problems in Engineering (2022)
5419671.

[31] L. J. Jia, Z. F. Qi, S. Zhang, Y. F. Qin, J. Shi, X. M. Zhang, X. J. Sun, “Dynamic


Analysis of the Acoustic Velocity Profile Obser-vation Underwater Glider,” Appl. Mech.
Mater. 475 (2013) 50–54.

[32] P. Guang, B. Hu, X. X. Du, Y. Y. Wang, “Research on Hydrodynamic


Characteristics of Underwater Gliding UUV Based on the CFD Technique,” Adv. Mater.
Res. 479 (2012) 729–732.

[33] A. Amory, E. Maehle, “Modelling and CFD Simulation of a Micro Autonomous


Underwater Vehicle SEMBIO,” In Proceedings of the OCEANS 2018 MTS/IEEE
Charleston, Charleston, SC, USA, 22–25 October 2018.

[34] J.V.N. Sousa, A.R.L. Macedo, Jr. Lima, “Numerical analysis of turbulent fluid flow
and drag coefficient for optimizing the AUV hull design,” J. Fluid Dyn. 4 (2014) 263–
277.

[35] P.S.A.T. Randeni, Z.Q. Leong, D. Ranmuthugala, A. Forrest, J. Duffy, “Numerical


investigation of the hydrodynamic interaction between two underwater bodies in relative
motion,” Appl. Ocean. Res. 51 (2015) 14–24.

[36] J. Dantas, E. Barros, “Numerical analysis of control surface effects on AUV


maneuverability,” Appl. Ocean. Res. 42 (2013) 168–181.

[37] A. Tyagi, D. Sen, “Calculation of transverse hydrodynamic coefficients using


computational fluid dynamic approach,” Ocean Eng. 33 (2006) 798–809.
[38] J. Donea, A. Huerta, J.P. Ponthot and A. Rodriguez-Ferran, “Arbitrary Lagrangian-
Eulerian methods,” Encyc. Comp. Mech. (2004) 1-38

[39] T. Uchiyama, "ALE finite element method for gas-liquid two-phase flow including
moving boundary based on an incompressible two-fluid model," Nuclear Engineering
and Design 205 (2001) 69-82.

54
[40] F. Duarte, R. Gormaz, and S. Natesan, "Arbitrary Lagrangian-Eulerian method for
Navier-Stokes equations with moving boundaries," Computer Methods in Appl. Mech.
Eng. 193 (2004) 4819-4836.

[41] J. U. Brackbill, D. B. Kothe, C. Zemach, “A continuum method for modeling surface


tension,” J. Comp. Phys. 100 (1991) 335-354.

[42] S. Osher, J.A. Sethian, “Fronts propagating with curvature dependent speed:
algorithms based on Hamilton-Jacobi formulations,” J. Comput. Phys.

[43] E. Olsson, G. Kreiss, “A conservative level set method for two-phase flow,” Journal
of Computational Physics 210 (2005) 225-246.

[44] C. Song, K. Kim, K. Lee, and H.K. Pak, "Thermochemical control of oil droplet
motion on a solid substrate," Appl. Phys. Lett. 93 (2008) 084102-1-3.

[45] J. Donea, A. Huerta, J.P. Ponthot and A. Rodriguez-Ferran, “Arbitrary Lagrangian-


Eulerian methods,” Encyc. Comp. Mech. (2004) 1-38

[46] A. Phillips, M. Furlong, S.R. Turnock, “The use of computational fluid dynamics to
assess the hull resistance of concept autonomous underwater vehicles,” Proceedings of
OCEANS. IEEE (2007) 1–6.

55

You might also like