Investigation of Thrust Deduction on Azimuth Thrusters in Bollard Pull Condition Using CFD
Investigation of Thrust Deduction on Azimuth Thrusters in Bollard Pull Condition Using CFD
Investigation of Thrust Deduction on Azimuth Thrusters in Bollard Pull Condition Using CFD
UNIVERSITY OF TECHNOLOGY
FOIVOS LEMONAKIS
FOIVOS LEMONAKIS
Typeset in LATEX
Printed by Chalmers Reproservice
Gothenburg, Sweden 2019
iv
Investigation of Thrust Deduction on Azimuth Thrusters in Bollard Pull Condition
using CFD
FOIVOS LEMONAKIS
Department of Mechanics and Maritime Sciences
Chalmers University of Technology
Abstract
This report is result of a master’s thesis project performed together with Caterpillar
Propulsion AB, which is a revisit of an older master’s thesis with title "Hydrodynam-
ics of Conventional Propeller and Azimuth Thruster in Behind Condition" (Matin,
2011). In the previous master’s thesis, the thrust deduction factors of open shaft
propellers and ducted azimuth thrusters, installed on an offshore vessel, are studied.
However, during the recent years the ducted azimuth thrusters have become more
and more common and especially for vessels such as tug boats, drilling vessels etc.
For such vessels, the bollard pull condition is really important, which indicates high
power operation with high loading (Funeno, 2009). Thus, the purpose of the current
master’s thesis is the investigation of thrust deduction on ducted azimuth thrusters
in bollard pull condition using Computational Fluid Dynamics (CFD).
In order achieve this type of study, a harbour tug boat with twin ducted azimuth
thrusters and a combination of steady state Multiple Reference Frame (MRF) and
transient Sliding Mesh Interface (SMI) modelling methodologies have been used.
MRF has been proven to be inaccurate in behind condition and at low or high ad-
vance coefficients, as it usually results in relatively strange flow fields in the vicinity
of the propeller while SMI has shown higher accuracy. However, SMI is a much more
computationally demanding methodology (Gullberg & Sengupta, 2011). Thus, SMI
has been used for critical regions and MRF for regions of lower interest.
Outcome of this study is that during the bollard pull condition the contribution of
the nozzle is higher than the propeller in terms of thrust and the contribution of
the gear case housing is higher than the hull in terms of resistance. The highest
thrust deduction appears on the gear case housing with 8.92% while the highest
merit deduction, which shows a lower performance, appears when placing the whole
propulsion unit behind the hull with 9.09%. Furthermore, the measured bollard pull
force of the open water case is about 8.23% higher than the one of the Wageningen
CD Series. However, this high difference is not realistic and the reason has to be
identified. Finally, by comparing the MRF and SMI methodologies, it is concluded
that the MRF approach gives appropriate results for initialization and quite rea-
sonable results for medium advance coefficients at low computational cost while the
SMI approach is needed for the critical regions of low and high advance coefficients.
v
Acknowledgements
This master’s thesis is part of the Naval Architecture and Ocean Engineering mas-
ter’s programme, which is part of the Department of Mechanics and Maritime Sci-
ences, Chalmers University of Technology.
The author of this thesis would like to thank and acknowledge the help and contri-
bution of the following:
• Family and friends for their support during the two intensive years of this
master’s programme.
vii
Contents
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
List of Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii
List of Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii
List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxi
1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Importance of this topic . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Purpose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4.1 Investigation cases . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4.2 Software packages used . . . . . . . . . . . . . . . . . . . . . . 3
2 Theory 5
2.1 Introduction to Computational Fluid
Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.1 Definition of the computational domain . . . . . . . . . . . . . 6
2.1.2 Grid generation . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.3 Identification of the physical and chemical phenomena that
need to be modelled . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.4 Definition of the fluid properties . . . . . . . . . . . . . . . . . 7
2.1.5 Definition of the boundary conditions . . . . . . . . . . . . . . 7
2.1.6 Integration of the governing equations of the flow over the
finite control volumes of the domain . . . . . . . . . . . . . . . 7
2.1.7 Discretisation of the integrated equations into algebraic equa-
tions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.1.8 Solving of the algebraic equations by the use of iterative method 10
2.1.9 Visualization of the simulation results . . . . . . . . . . . . . . 10
2.1.10 Process of the numerical results . . . . . . . . . . . . . . . . . 10
2.2 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1 The continuity equation . . . . . . . . . . . . . . . . . . . . . 11
2.2.2 The Navier-Stokes equations . . . . . . . . . . . . . . . . . . . 12
2.3 Similarity between model and full-scale . . . . . . . . . . . . . . . . . 13
2.3.1 The continuity equation . . . . . . . . . . . . . . . . . . . . . 13
ix
Contents
3 Geometry 33
3.1 Main particulars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2 Azimuth thruster . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3 Hull . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4 Methodology 37
4.1 Pre-processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.1.1 Definition of the computational domain . . . . . . . . . . . . . 37
4.1.2 Grid generation . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.1.3 Identification of the physical and chemical phenomena that
need to be modelled . . . . . . . . . . . . . . . . . . . . . . . 43
4.1.4 Definition of the fluid properties . . . . . . . . . . . . . . . . . 44
4.1.5 Definition of the boundary conditions . . . . . . . . . . . . . . 44
4.2 Solving . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2.1 Integration of the governing equations of the flow over the
finite control volumes of the domain . . . . . . . . . . . . . . . 49
4.2.2 Discretisation of the integrated equations into algebraic equa-
tions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2.3 Solving of the algebraic equations by the use of iterative method 50
4.3 Post-processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.3.1 Visualization of the simulation results . . . . . . . . . . . . . . 50
4.3.2 Process of the numerical results . . . . . . . . . . . . . . . . . 51
4.4 System used to run OpenFOAM . . . . . . . . . . . . . . . . . . . . . 53
x
Contents
5.1.6 Cavitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.2 Hydrodynamic interactions in bollard pull condition . . . . . . . . . . 77
5.3 Accuracy and reliability . . . . . . . . . . . . . . . . . . . . . . . . . 83
References 87
xi
List of Abbreviations
Abbreviation Description
2D Two-Dimensional space
3D Three-Dimensional space
AHTS Anchor Handling/Tug/Supply
AMI Arbitrary Mesh Interpolation
CAD Computer Aided Design
CAE Computer Aided Engineering
CFD Computational Fluid Dynamics
DNS Direct Numerical Simulation
EAR Expanded Area Ratio
FPP Fixed Pitch Propeller
HPC High-Performance Computing
LES Large Eddy Simulation
MRF Multiple Reference Frame
PIMPLE Pressure-Implicit Method for Pressure-Linked Equation
PISO Pressure-Implicit with Splitting of Operators
QUICK Quadratic Upstream Interpolation for Convective Kinematics
RANS Reynolds-Averaged Navier Stokes
RPM Revolutions Per Minute
RSM Reynolds Stress equation Model
SIMPLE Semi-Implicit Method for Pressure-Linked Equations
SIMPLEC Semi-Implicit Method for Pressure Linked Equations - Consistent
SIMPLER Semi-Implicit Method for Pressure-Linked Equations - Revised
SMI Sliding Mesh Interface
SST Shear Stress Transport
TDMA Tri-Diagonal Matrix Algorithm
xii
List of Variables
D m diameter
De , Dw , Dn , Ds , Dt , Db diffusion terms of the east, west, north, south,
top, bottom faces in finite volume theory
dm kg mass of the element
dV m3 infinitesimal fluid element
e, w, n, s, t, b east, west, north, south, top, bottom faces of
the control volume in finite volume theory
E(κ) m3 /s2 spectral energy content of the large eddies in
turbulence modelling
En − Euler number
F N force
→
−
F N force vector derived from the total force acting
on the volume element in finite volume theory
F N mean force
Fe , Fw , Fn , Fs , Ft , Fb convection terms of the east, west, north, south,
top, bottom faces in finite volume theory
Fn − Froude number
g m/s2 acceleration of gravity
J − advance coefficient
k m2 /s2 turbulent kinetic energy in turbulence modelling
KQ − non-dimensional torque coefficient, torque
coefficient in behind condition
KQ − mean torque coefficient
KQo − torque coefficient in open water condition
ks µm surface roughness
KT − non-dimensional thrust coefficient
KT − mean thrust coefficient
KT,net − net sum of all the thrust coefficients
KT n − nozzle thrust coefficient
KT p − propeller thrust coefficient
KT,R − sum of the thrust coefficients that are derived
from the resistance components
KT,T − sum of the thrust coefficients that are derived
from the thrust components
L m reference length (i.e., characteristic linear
dimension) that describes the travelled length
of the fluid, obstacle width
` m characteristic length of the larger scales in
turbulence modelling
M N·m moment
M N·m mean moment
xiii
Contents
mc − merit coefficient
mc,deduct % merit deduction
n rpm rotational speed
p Pa pressure, instantaneous pressure in turbulence
modelling
P Pa steady mean value of the pressure component
P, E, W, N, S, T, B central nodal point of the control volume and
the east, west, north, south, top, bottom central
nodal points of the neighboring volumes
in finite volume theory
p0 Pa fluctuating pressure component in turbulence
modelling
PB W brake power of main engine
pd Pa downstream pressure
PD W delivered power to the propeller via the shaft
phd Pa hydrodynamic pressure
pu Pa upstream pressure
Q N·m torque
Q N·m mean torque
R N resistance
R N mean resistance
Rn − Reynolds number
Rncrit − critical Reynolds number
SM N/m3 momentum source term
◦
T C, N temperature, thrust
t s, s, % time, characteristic time of the larger scales in
turbulence modelling, thrust deduction fraction
T N mean thrust
TBP ton bollard pull force
TBP,1995kw,CD ton bollard pull force at a fixed delivered power
obtained from the Wageningen CD Series
TBP,1995kw,CF D ton bollard pull force at a fixed delivered power
obtained from CFD results
Tnet N net sum of the total thrust
Tnet,deduct N deduction of the net sum of the total thrust
Tp N propeller thrust
Ttot N total thrust of a vessel that operates in full power
U∞ m/s reference velocity, steady mean value of the
velocity component in turbulence modelling
u m/s characteristic velocity of the Kolmogorov
microscales in turbulence modelling
u, v, w m/s scalars representing the velocity in x, y and
z-directions, instantaneous velocities in x, y and
z-directions in turbulence modelling
xiv
Contents
→
−
u ,→−v ,→
−
w m/s velocity vectors in x, y and z-directions
0
u m/s fluctuating velocity component in turbulence modelling
u+ − non-dimensional velocity
uτ m/s friction velocity
VA m/s advance velocity
VS m/s ship velocity
Wn − Weber number
wT − effective Taylor’s wake fraction
x, y, z m, − components of a vector in the x, y or z-directions,
coordinates of global system
x̄, ȳ, z̄, ζ̄, ū, v̄, w̄, p̄, t̄ − non-dimensional quantities in similarity equations
y+ − non-dimensional distance from the wall
z m distance of the upper part of the nozzle and the
water surface
Z − number of blades
→
−
α m/s2 acceleration vector
Γ generalized diffusion coefficient
γ N/m surface tension or over-speed ratio in channel
δ m boundary layer thickness
δx W P m distance between the nodes W and P in
finite volume theory
∆R N increase of the resistance
ε m2 /s3 dissipation rate of turbulent kinetic energy
ζ m wave elevation
η m characteristic length of the Kolmogorov
microscales in turbulence modelling
η0 − open water efficiency
η0 − mean open water efficiency
ηD − total propulsive efficiency
ηH − hull efficiency
ηR − relative rotative efficiency
ηS − shaft efficiency
ϑ m/s characteristic velocity of the larger scales in
turbulence modelling
κ 1/m wave number
µ Pa · s dynamic viscosity of the fluid
ν m2 /s kinematic viscosity of the fluid
νt m2 /s turbulent viscosity field
π − mathematical constant
ρ kg/m3 density
τ Pa, s shear stress, characteristic time of the Kolmogorov
microscales in turbulence modelling
τw Pa wall shear stress
τxx , τyy , τzz Pa Reynolds normal stresses in RANS equations
xv
Contents
τxy , τyx , τxz , τzx , τyz , τzy Pa Reynolds shear stresses in RANS equations
φ − flow property in finite volume theory
ω 1/s specific dissipation rate of turbulent energy in
turbulence modelling
xvi
List of Figures
2.1 A finite control volume (cell) and its neighbouring nodes (Versteeg &
Malalasekera, 2007). . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Part of the grid presented in 2D (Versteeg & Malalasekera, 2007). . . 9
2.3 Mass flow in x-direction through the fluid element (Larsson & Raven,
2010). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 Scale effects (Carlton, 2007). . . . . . . . . . . . . . . . . . . . . . . . 16
2.5 Regions of flow (Larsson & Raven, 2010). . . . . . . . . . . . . . . . . 19
2.6 Regions of development of the velocity boundary layer on a flat plate
(Bergman, Lavine, Incropera, & Dewitt, 2011). . . . . . . . . . . . . . 20
2.7 Comparison of laminar and turbulent velocity profiles of the boundary
layer on a flat plate (Bergman, Lavine, Incropera, & Dewitt, 2011). . 21
2.8 Development of the velocity boundary layer on a flat plate (Bergman,
Lavine, Incropera, & Dewitt, 2011). . . . . . . . . . . . . . . . . . . . 21
2.9 Regions of development of the velocity boundary layer on a flat plate
(Larsson & Raven, 2010). . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.10 Arrangement of an open water test in a towing tank (Dyne & Bark,
2005). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.11 Arrangement of an open water test in a cavitation tunnel. . . . . . . 27
2.12 Open water characteristics diagram. . . . . . . . . . . . . . . . . . . . 28
xvii
List of Figures
xviii
List of Figures
xix
List of Figures
xx
List of Tables
xxi
List of Tables
xxii
List of Tables
xxiii
List of Tables
xxiv
1
Introduction
In this chapter, a background of the propulsion systems; the propellers; the driving
reasons of the continuous development; and finally the computational fluid dynamics
is presented. In the rest of the chapter, it is described how this thesis contributes to
the field by explaining the purpose, the importance and the scope of this topic.
1.1 Background
Transportation through the water is one of the oldest means of transport during
humankind. Ships and boats are used for many different purposes such as trade,
fishing, defense and leisure. Therefore, the need of some kind of propulsion sys-
tem was already created since the ancient years of the stone age. For thousands of
years, the oars were the main and probably the only type of propulsion while during
the 7th century the sails were introduced. During the 16th century the number of
propulsion systems were limited and the use of sails was at its peak. At the end
of the 18th century the Industrial Revolution arrived, which brought a new era for
the shipping industry as well as for many other industries. In the beginning of the
19th century, paddle wheels were the prevailing propulsion system for steam vessels
while at the same time the screw propeller got invented, which was based on the
Archimedes’ screw.
Since then, propellers are the predominant marine propulsion system, which convert
rotational motion into thrust power. A pressure difference is generated between the
pressure side and the suction side of the propeller blades and as a result the fluid,
which in this specific case is water, is accelerated and the vessel moves. The propul-
sion of a vessel though is not any more the main goal as this is already solved. How-
ever, there is a continuous development of the marine propulsion systems, which is
mainly driven by the need of higher efficiency; both for environmental and econom-
ical reasons, noise reduction due to noise pollution and comfort reasons, operating
security, safe performance and maneuverability. In order to achieve those, the un-
derstanding of the physical phenomena around the propulsion unit and the hull of
the ship is vital. Hydrodynamics is the subject that is studying the flow of water
and the forces applied on structures immersed in water.
An important tool for the study of hydrodynamic phenomena and designs is Com-
putational Fluid Dynamics (CFD), which is the numerical analysis of fluid flow, heat
transfer and associated phenomena such as chemical reactions. The methodology of
1
1. Introduction
this kind of analysis is mostly used to run computer-based simulations by the use
of computer software packages in order to aid various engineering tasks, something
that makes CFD a part of the Computer Aided Engineering (CAE). CFD starts from
the 1960s where the aerospace industry developed CFD techniques and integrated
them into the design while from the 1990s a wider industrial usage starts (Versteeg
& Malalasekera, 2007).
1.3 Purpose
The purpose of the current thesis is the investigation of thrust deduction on ducted
azimuth thrusters in bollard pull condition and in full-scale using CFD, in order
to develop a better understanding of the physical phenomena around this type of
geometry, understand where losses occur as well as understand which parts of those
geometries are beneficial. In order to fulfil this purpose, there is the need to investi-
gate the hydrodynamic interaction between the propeller, the nozzle, the gear case
housing and the hull.
1.4 Scope
In order perform this type of study, a harbour tug boat with twin ducted azimuth
thrusters and a combination of steady state Multiple Reference Frame (MRF) and
transient Sliding Mesh Interface (SMI) modelling methodologies have been used.
Steady state MRF has been proven to be inaccurate in behind condition and at
2
1. Introduction
low or high advance coefficients, as it usually results in relatively strange flow fields
in the vicinity of the propeller while transient SMI has shown higher accuracy.
However, transient SMI is a much more expensive methodology in terms of com-
putational cost (Gullberg & Sengupta, 2011). Thus, transient SMI has been used
for critical regions such as for advance coefficients J = 0.00, that indicate bollard
pull condition, and steady state MRF for regions of lower interest. Finally, a com-
parison with the Wageningen CD Series is done in order to develop an evaluation
tool which can be used in the future in order to evaluate the early stages of a project.
Note: The gear case housing is considered to be part of the propulsion unit and
not part of the hull.
3
1. Introduction
4
2
Theory
In this chapter, a thorough explanation of the background theory of this thesis is pre-
sented. The first sections of this chapter demonstrate a fundamental theory of CFD,
governing equations, similarity between model and full-scale, propeller scaling, types
of fluid flow and turbulence modelling. Thus, these sections could be skipped from
a reader with basic knowledge in these topics. On the contrary, the last sections of
this chapter indicate a more specific theory of open water characteristics, propulsive
factors, bollard pull condition and modelling methodologies, which is directly related
to the current study.
CFD codes and software packages are using numerical algorithms in order to solve
fluid flow problems. All commercial CFD packages include input problem parame-
ters and most of them also include tools for analyzing the results. Thus, all CFD
packages include three main steps (Versteeg & Malalasekera, 2007):
• Pre-processing, which includes the input problem parameters and the con-
version of the inputs into a form that the solver can use. Pre-processing can
be divided into:
– Definition of the computational domain
– Grid generation
– Identification of the physical and chemical phenomena that need to be
modelled
– Definition of the fluid properties
– Definition of the boundary conditions
• Solving can be divided into various categories according to the numerical
techniques that are being used. However, the finite volume method is the
most widely used method among most of the CFD codes. Solving can be
divided into:
– Integration of the governing equations of the flow over the finite control
volumes of the domain
5
2. Theory
In the following sections, the already mentioned main steps of CFD and their sub-
categories are presented in the same order as above.
6
2. Theory
7
2. Theory
The integration of the governing equations of the flow over the finite control volumes
is more thoroughly presented in Sections 2.2 and 2.3.
Figure 2.1: A finite control volume (cell) and its neighbouring nodes
(Versteeg & Malalasekera, 2007).
The transport phenomena include the convection part, which is the transport due
to fluid flow, and the diffusion part, which is the transport due to variations of
the flow variable φ from node to node (i.e. net movement from regions with high-
concentration to regions with low-concentration). An example of a grid in 2D and
its nodes (i.e. E, W, N, S) are presented in the following figure.
8
2. Theory
Figure 2.2: Part of the grid presented in 2D (Versteeg & Malalasekera, 2007).
Schemes can generally be split into different orders of accuracy. Schemes with
lower order of accuracy have higher stability of the results with less accuracy (due
to the high level of diffusion error). On the other hand, schemes with higher order
of accuracy have lower stability (due to false oscillations when the Peclet number
is high) with more accuracy. Peclet number is evaluated at the face of the control
volume and is defined as follows (i.e. assuming a west face):
Fw (ρU )w
P ew = = (2.1)
Dw Γ/δxW P
where:
Fw is the convection term of the west face,
Dw is the diffusion term of the west face,
Γ is the generalized diffusion coefficient,
δxW P is the distance between the nodes W and P.
Thus, sometimes hybrid schemes can be ideal since they combine a wider range of
problems (i.e. diffusion or convection dominated problems). Such an example is the
hybrid differencing scheme which employs:
• Central differencing scheme (2nd order accurate) for Peclet numbers P e < 2
• Upwind scheme (1st order accurate) by taking into account only the trans-
portiveness (i.e. upwind for convection and setting the diffusion to zero) for
Peclet numbers P e ≥ 2
Finally, the evaluation of the order of accuracy of a discretisation is done by the use
of Taylor expansion.
9
2. Theory
The mathematical algorithms are using point-iterative methods for computing points
of iterated functions where a linear system of algebraic equations is solved (i.e.
Ax = b with unknown the x). Examples of such iterative linear solvers are:
• Gauss-Seidel
• Tri-Diagonal Matrix Algorithm (TDMA)
10
2. Theory
∂ρ
+ div(ρu) = 0 (2.2)
∂t
which can also be written as:
∂ρ ∂ ∂ ∂
+ ρu · dV + ρv · dV + ρw · dV = 0 (2.3)
∂t ∂x ∂y ∂z
where:
dV = dx · dy · dz,
u is a scalar representing the velocity in x-direction,
v is a scalar representing the velocity in y-direction,
w is a scalar representing the velocity in z-direction.
However, since the fluid is considered as incompressible, the continuity equation can
be re-written as:
∂u ∂v ∂w
+ + =0 (2.4)
∂x ∂y ∂z
11
2. Theory
where:
→
−
F is the force vector derived from the total force acting on the element (i.e. arrow
represents vector),
dm is the mass of the element,
→
−
α is the acceleration vector.
After the long derivation, the Navier-Stokes equations for incompressible flow are
obtained, assuming that the gravity g is the only body force along the z-axis (Ver-
steeg & Malalasekera, 2007):
x-momentum:
∂(ρu) ∂p
+ div(ρuu) = − + div(µ grad u) + SM x (2.6)
∂t ∂x
! !
∂u ∂u ∂u ∂u ∂p ∂ 2u ∂ 2u ∂ 2u
ρ +u +v +w =− +µ + + + SM x
∂t ∂x ∂y ∂z ∂x ∂x2 ∂y 2 ∂z 2
y-momentum:
!
∂v ∂v ∂v ∂v 1 ∂p ∂ 2v ∂ 2v ∂ 2v 1
+u +v +w =− +ν + + + SM y (2.8)
∂t ∂x ∂y ∂z ρ ∂y ∂x2 ∂y 2 ∂z 2 ρ
z-momentum:
!
∂w ∂w ∂w ∂w 1 ∂p ∂ 2w ∂ 2w ∂ 2w 1
+u +v +w =− +ν + + + SM z − g (2.9)
∂t ∂x ∂y ∂z ρ ∂z ∂x2 ∂y 2 ∂z 2 ρ
where:
g is the gravity in z-direction,
SM is the momentum source term.
By looking at Eq.(2.7), (2.8) and (2.9), the left-hand side of these equations repre-
sents the acceleration of the fluid element in the x, y and z-directions respectively,
thus the convection terms. The right-hand side represents the forces due to pressure
(i.e. pressure gradient), viscous (i.e. diffusion terms) and body (i.e. gravity).
12
2. Theory
In order to apply the similarity laws between the model and full-scale, the governing
equations and the boundary conditions need to be independent of dimension. Thus,
the following dimensionless quantities are introduced:
x y z ζ
x̄ = , ȳ = , z̄ = , ζ̄ =
L L L L
u v w phd tU∞
ū = , v̄ = , w̄ = , p̄ = 2
, t̄ =
U∞ U∞ U∞ ρU∞ L
where:
x, y, z are components of a vector in the x, y or z-directions,
ζ is the wave elevation,
L is the reference length (i.e. characteristic linear dimension) that describes the
travelled length of the fluid,
U∞ is the reference velocity,
phd is the hydrodynamic pressure,
t is the time.
U∞
By dividing this equation by L
, the following equation is obtained:
∂ ū ∂v̄ ∂ w̄
+ + =0 (2.11)
∂ x̄ ∂ ȳ ∂ z̄
13
2. Theory
∂ ū ∂ ū ∂ ū ∂ ū
+ ū + v̄ + w̄ = (2.13)
∂ t̄ ∂ x̄ ∂ ȳ ∂ z̄
!
∂ p̄ ν ∂ 2 ū ∂ 2 ū ∂ 2 ū
− + + + + SM x
∂ x̄ U∞ L ∂ x̄2 ∂ ȳ 2 ∂ z̄ 2
In a similar way, the equations for the other two directions are obtained as follows:
y-momentum:
∂v̄ ∂v̄ ∂v̄ ∂v̄
+ ū + v̄ + w̄ = (2.14)
∂ t̄ ∂ x̄ ∂ ȳ ∂ z̄
!
∂ p̄ ν ∂ 2 v̄ ∂ 2 v̄ ∂ 2 v̄
− + + + + SM y
∂ ȳ U∞ L ∂ x̄2 ∂ ȳ 2 ∂ z̄ 2
z-momentum:
∂ w̄ ∂ w̄ ∂ w̄ ∂ w̄
+ ū + v̄ + w̄ = (2.15)
∂ t̄ ∂ x̄ ∂ ȳ ∂ z̄
!
∂ p̄ ν ∂ 2 w̄ ∂ 2 w̄ ∂ 2 w̄
− + + + − g + SM z
∂ z̄ U∞ L ∂ x̄2 ∂ ȳ 2 ∂ z̄ 2
The governing equations are taken into account when similarity is about to be
achieved in order to see which equations are dependent on the scale. The only
parameter that appears in these equations is the Reynolds number Rn (which
appears as 1/Rn and is circled in the above equations) in the dimensionless Navier-
Stokes equations (see Eq.(2.13), (2.14) & (2.15)), which is defined as:
ρuL uL
Rn = = (2.16)
µ ν
where:
u is a scalar representing the velocity in x-direction in respect to the object,
µ is the dynamic viscosity of the fluid,
ν is the kinematic viscosity of the fluid.
14
2. Theory
pu − pd
En = 2
(2.17)
ρU∞
where:
pu is the upstream pressure,
pd is the downstream pressure.
U∞
Fn = √ (2.18)
gL
• the Weber number W n, which is used to determine the spray, the wave break-
ing and the surface waves,
2
ρU∞ L
Wn = (2.19)
γ
where:
γ is the surface tension or over-speed ratio in channel.
It should be mentioned that the above statement applies for open shaft arrangement
where a huge number of model testing has been done. On the other hand, this state-
ment is not true for azimuth thrusters due to the high model testing uncertainties
that are derived from the high viscous effects that appear at model scale. For this
reason, full-scale CFD might be a better alternative for azimuth thrusters compared
to model testing.
15
2. Theory
The viscous effects are the main reason for the differences between the model and
full-scale due to the phenomena that are dependent on the Reynolds number. The
dependency of these phenomena to the Reynolds number and consequently to the
viscosity is explained thoroughly in Sections 2.2 and 2.3. The Reynolds number
should be the same at both scales in order for the viscosity effects to be correct.
However, due to the methods of testing model propellers, the Reynolds number ends
up being different between model and full-scale and as a result a different boundary
layer structure to the flow over the blades and nozzle arises. A consequence of that
is the too high viscous resistance in model scale.
16
2. Theory
• Compressible - Incompressible
A flow is described as compressible when the density of the fluid ρ is dependent
on the pressure, thus the density is changing within the control volume that
moves with the flow velocity at regions with different pressure. On the other
hand, a flow is described as incompressible when the density of the fluid ρ
is independent on the pressure and thus, constant within the control volume
that moves with the flow velocity at regions with different pressure.
• Viscous - Inviscid
A flow is described as viscous when the fluid of the flow is viscous. On the
other hand, a flow is described as inviscid when the fluid of the flow has no
viscosity (i.e. viscosity is equal to zero).
• Rotational - Irrotational
A flow is described as rotational when vorticity, which is the curl of the velocity
vector, appears all over the flow field. On the other hand, a flow is described
as irrotational when vorticity is zero all over the flow field.
• 1D - 2D - 3D
A flow is described as 1D when the flow variables φ, such as velocity, pressure,
temperature and density are defined within one dimensional coordinate system
in space. A flow is described as 2D and 3D when the flow variables φ are defined
within two or three dimensional coordinate systems respectively in space.
• Potential flow
A potential flow is a flow that is described by the velocity potential φ(x, y, z)
for some scalar field φ, being a function of space and time and assuming
incompressible, inviscid and irrotational flow. The flow velocity → −v is a vector
field that is equal to the gradient of the velocity potential as shown below
(Larsson & Raven, 2010):
→
−v (x, y, z) = 5φ (2.20)
The Reynolds number associates the inertia forces that are related to the convec-
tive effects and viscous forces with the flow. A critical Reynolds number, Rncrit is
introduced, which determines the type of the flow. In cases where Rn < Rncrit the
flow is considered as laminar while when Rn ≥ Rncrit the flow is considered as tur-
bulent. At turbulent flows the instantaneous flow variables φ are decomposed into
a mean value and a fluctuating value. The instantaneous velocity u and pressure p
are presented below:
u = U + u0 and p = P + p0 (2.21)
where:
U is the steady mean value of the velocity component,
17
2. Theory
Turbulent flows include rotational flows that are called turbulent eddies. These ed-
dies bring the fluid particles closer to each other, which results in a high heat, mass
and momentum exchange. This yields to an effective mixing, which rises the diffu-
sion coefficients. Turbulent eddies that are large enough to interact with the mean
flow and extract energy from it are called vortices while vortex stretching shows how
much a fluid particle has been stretched.
The characteristic velocity ϑ and the characteristic length ` of the larger eddies (e.g.
vortices) is observed to be of the same order as the mean velocity U and length L
of the mean flow. Thus, by looking at Reynolds equation (see Eq.2.16), it can be
considered that vortices are highly influenced by inertia effects while viscous effects
can be neglected due to the same order of the velocity and length. Since vortices
are considered inviscid and the angular momentum is preserved, the rotation rate
gets increased while the rotation radius gets decreased. This yields to creation of
motion at the smaller length scales and within smaller time steps. As a result, vor-
tex stretching provides the required energy and the turbulence is maintained.
Smaller eddies are mainly streched by larger eddies while mean flow has less im-
pact on them. The so called energy cascade explains the transfer of kinetic energy
from the larger eddies to the smaller ones. The smallest eddies that are dominated
equally by inertia effects and viscous effects are called Kolmogorov microscales. The
energy within these microscales is dissipated and converted into thermal energy,
thus resulting in high energy losses. Kolmogorov microscales are expressed by the
dissipation rate of turbulent kinetic energy ε and the fluid’s kinematic viscosity ν.
Magnitude ratios of the larger scales and the Kolmogorov microscales are used in
order to proceed to a dimensional analysis. These magnitude ratios, that are ex-
pressed by the characteristic velocity ϑ, length ` and time t of the larger scales and
the characteristic velocity u, length η and time τ of the Kolmogorov microscales,
are shown below:
• Velocity scale ratio:
u −1/4
≈ Rn` (2.22)
ϑ
• Length scale ratio:
η −3/4
≈ Rn` (2.23)
`
• Time scale ratio:
τ −1/2
≈ Rn` (2.24)
t
As it is already mentioned, the large eddies are considered as inviscid while they
are highly dependent on the characteristic velocity ϑ and the characteristic length
`. As a result, the spectral energy content E(κ) of the large eddies should be
directly proportional to the characteristic velocity ϑ and the characteristic length
18
2. Theory
` as shown in Eq.(2.25) and (2.26). The large eddies are considered anisotropic as
the characteristic length ` is linked with components such as the boundary layer
thickness δ, the obstacle width L and the surface roughness ks .
E(κ) ∝ ϑ2 ` (2.25)
19
2. Theory
The boundary layer is the most important region of flow where the transition from
laminar to turbulent flow occurs. The simplest boundary layer is considered the one
that is developed on a flat plate, a plate with really small thickness and parallel
to the flow on which the pressure is considered constant. The flat plate consists
of three regions: the laminar boundary layer, the transition region from laminar
to turbulent flow and the turbulent boundary layer. The development of velocity
boundary layer on a flat plate within these regions is shown in Figure 2.6 (Larsson
& Raven, 2010).
Figure 2.6: Regions of development of the velocity boundary layer on a flat plate
(Bergman, Lavine, Incropera, & Dewitt, 2011).
Laminar flow is the only flow within the laminar boundary layer, which is char-
acterized by ordered streamlines, until the transition region where an intermediate
flow dominates (i.e. laminar flow within the viscous sublayer and intermediate flow
within the buffer layer). Turbulent flow follows within the turbulent boundary layer
where a turbulence flow dominates (i.e. laminar flow within the viscous sublayer,
intermediate flow within the buffer layer and turbulent flow within the turbulent
region) (Bergman, Lavine, Incropera, & Dewitt, 2011).
Moving streamwise from the laminar to the turbulent region (i.e. increasing x-
direction), the boundary layer thickness δ grows (see Figure 2.6) and the velocity
gradients on the flat plate decrease (see Figure 2.7).
20
2. Theory
The reduction of the velocity is associated with the shear stresses τ acting in parallel
to the velocity (see Figure 2.8) in order to conserve the particle motion at a distance
δ from the solid surface.
The velocity within the boundary layer needs to be expressed in terms of the fluid
density ρ, the dynamic viscosity µ of the fluid and the shear stresses at the wall
τw . It should be mentioned that the velocity should be kept independent of the
boundary layer thickness δ. Thus, the following equations are introduced (Larsson
& Raven, 2010):
• The non-dimensional velocity u+ :
u
u+ = (2.29)
uτ
where the friction velocity uτ is defined as
s
τw
uτ = (2.30)
ρ
• The non-dimensional distance from the wall y + :
yuτ
y+ = (2.31)
ν
Finally, the boundary layer can be split into four regions:
• The viscous sublayer: 0 ≤ y + ≤ 5
This region (i.e. region I) is part of the inner layer where a laminar flow
dominates.
21
2. Theory
Figure 2.9: Regions of development of the velocity boundary layer on a flat plate
(Larsson & Raven, 2010).
LES and DNS methods provide higher resolution and a more detailed description
22
2. Theory
of turbulence compared to RANS. However, both LES and DNS are much more
expensive methods and the level of resolution and details that both provide is usually
not needed. On the other hand, RANS yields a decent level of resolution and details,
which are the result of the averaged flow properties such as the mean values of the
velocities, pressures and stresses. Due to this fact, RANS method has been used the
most among the three.
∂ρ
+ div(ρu) = 0 (2.32)
∂t
which can also be written as:
∂ρ ∂ ∂ ∂
+ ρu · dV + ρv · dV + ρw · dV = 0 (2.33)
∂t ∂x ∂y ∂z
However, since the fluid is considered incompressible (i.e. ρ = ρ) and by taking into
account that div(u) = div(U ), the continuity equation can be re-written as:
div(U ) = 0 (2.34)
Then, by taking the average of each term of the Navier-Stokes equations (see
Eq.(2.7), (2.8) & (2.9)) for incompressible flow and assuming that the gravity g is
the only body force along the z-axis, the RANS equation are obtained:
x-momentum:
∂(ρu) ∂p
+ div(ρuu) = − + div(µ grad u) + SM x (2.35)
∂t ∂x
Then, the mean and fluctuating components (see Eq.(2.21)) are introduced into the
time-average of each term:
∂(ρU ) ∂P
+ div(ρU U ) + div(ρu0 u0 ) = − + div(µ grad U ) + SM x (2.36)
∂t ∂x
where:
∂(ρu) ∂(ρU )
= , div(ρuu) = div(ρU U ) + div(ρu0 u0 ) (2.37)
∂t ∂t
∂p ∂P
− =− , div(µ grad u) = div(µ grad U )
∂x ∂x
23
2. Theory
Rearranging Eq.(2.36) and moving the fluctuating components on the right hand
side, the following equation is obtained:
! !
∂U ∂U ∂U ∂U ∂P ∂ 2U ∂ 2U ∂ 2U
ρ +U +V +W =− +µ + + + (2.38)
∂t ∂x ∂y ∂z ∂x ∂x2 ∂y 2 ∂z 2
" #
∂(−ρu02 ) ∂(−ρu0 v 0 ) ∂(−ρu0 w0 )
+ + + SM x
∂x ∂y ∂z
Dividing by ρ, the following equation is obtained:
!
∂U ∂U ∂U ∂U 1 ∂P ∂ 2U ∂ 2U ∂ 2U
+U +V +W =− +ν + + + (2.39)
∂t ∂x ∂y ∂z ρ ∂x ∂x2 ∂y 2 ∂z 2
" #
1 ∂(−ρu02 ) ∂(−ρu0 v 0 ) ∂(−ρu0 w0 ) 1
+ + + SM x
ρ ∂x ∂y ∂z ρ
In a similar way, the equations for the other two directions are obtained as follows:
y-momentum:
!
∂V ∂V ∂V ∂V 1 ∂P ∂ 2V ∂ 2V ∂ 2V
+U +V +W =− +ν + + + (2.40)
∂t ∂x ∂y ∂z ρ ∂x ∂x2 ∂y 2 ∂z 2
" #
1 ∂(−ρu0 v 0 ) ∂(−ρv 02 ) ∂(−ρv 0 w0 ) 1
+ + + SM y
ρ ∂x ∂y ∂z ρ
z-momentum:
!
∂W ∂W ∂W ∂W 1 ∂P ∂ 2W ∂ 2W ∂ 2W
+U +V +W =− +ν + + + (2.41)
∂t ∂x ∂y ∂z ρ ∂x ∂x2 ∂y 2 ∂z 2
" #
1 ∂(−ρu0 w0 ) ∂(−ρv 0 w0 ) ∂(−ρw02 ) 1
+ + − g + SM z
ρ ∂x ∂y ∂z ρ
By looking at the Navier-Stokes equations (see Eq.(2.7), (2.8) & (2.9)) and the
RANS equations (see Eq.(2.39), (2.40) & (2.41)), nine extra terms can be observed,
which represent three normal stresses:
τxx = −ρu02 , τyy = −ρv 02 , τzz = −ρw02 (2.42)
and three double shear stresses:
τxy = τyx = −ρu0 v 0 , τxz = τzx = −ρu0 w0 , τyz = τzy = −ρv 0 w0 (2.43)
These stresses are knows at the Reynolds stresses. The Reynolds stresses are addi-
tional turbulent shear stresses within the fluid layers, which are result of momentum
exchange derived from the convective transport by the eddies. This momentum ex-
change results to a deceleration of the faster moving fluid layers and an acceleration
of the slower moving fluid layers. The symmetry of the double shear stresses is
explained by the identical net flux of the x, y and z-momentum accordingly and the
identical mean component of this net flux of the x, y and z-momentum out of the
control volume (Versteeg & Malalasekera, 2007).
24
2. Theory
Some of the most common turbulence models that are usually used in CFD are
presented below:
• κ − ε model
• Reynolds stress equation model (RSM)
• Wilcox κ − ω model
• Menter Shear Stress Transport (SST) κ − ω model
The turbulent kinetic energy k, which is derived from the Navier-Stokes equations,
is defined as the half of the sum of the fluctuating velocity components (Versteeg &
Malalasekera, 2007):
1
k = (u02 + v 02 + w02 ) (2.44)
2
The dissipation rate of turbulent kinetic energy ε is defined as:
ϑ = k 1/2 (2.47)
k 3/2
`= (2.48)
ε
25
2. Theory
For the k − ω model, the velocity scale ϑ is the same as above while the the length
scale ` is defined as:
k 1/2
`= (2.49)
ω
The Reynolds stress and the k-equation are identical to the original Wilcox k − ω
model:
!
2 ∂Ui ∂Uj 2
τij = −ρu0 v 0 = −ρu0i u0j = 2µt Sij − ρkδij = µt + − ρkδij (2.50)
3 ∂xj ∂xi 3
∂(ρk) µt
+ div(ρkU ) = div µ+ grad(k) + Pk − β ∗ ρkω (2.51)
∂t σk
while the ε-equation is transformed into an ω-equation by substituting ε = kω:
" ! # !
∂(ρω) µt 2 ∂Ui
+ div(ρωU ) = div µ+ grad(ω) + γ2 2ρSij · Sij − ρω δij −
∂t σω,1 3 ∂xj
ρ ∂k ∂ω
β2 ρω 2 + 2 (2.52)
σω,2 ω ∂xk ∂xk
α1 ρk
µt = (2.53)
max(α1 ω, SF2 )
where:
q
S= 2Sij Sij (2.54)
while the revised model constants are the following (Versteeg & Malalasekera, 2007):
26
2. Theory
Then, the non-dimensional thrust coefficient KT and torque coefficient KQ are ob-
tained:
T
KT = 2 4 (2.58)
ρn D
27
2. Theory
Q
KQ = (2.59)
ρn2 D5
The delivered power to the propeller via the shaft PD is defined as:
PD = 2πnQ (2.60)
and the thrust power that is created from the propeller PT is:
PT = T VA (2.61)
The results of an open water test are presented through an open water characteristics
diagram as shown in Figure 2.12.
28
2. Theory
where:
KQo is the torque coefficient in open water condition,
KQ is the torque coefficient in behind condition.
ηD = ηo · ηR · ηH (2.66)
R/T 1−t
ηH = = (2.67)
VA /VS 1 − wT
Another useful factor often connected to the hydrodynamic interactions, even if itself
is a mechanical interaction, is the shaft efficiency ηS , which is defined as:
PD
ηS = (2.68)
PB
where:
PB is the brake power.
Empirically, ηS is set to 0.97 for open shaft configurations and to 0.95 for azimuth
thrusters.
29
2. Theory
In bollard pull condition, the thrust deduction fraction t is mainly used to describe
the increase of the resistance due to the suction of the propeller and to determine
the generated thrust. On the other hand, some of the other conventional hydro-
dynamic interactions such as the effective Taylor’s wake fraction wT , the relative
rotative efficiency ηR and the propulsive efficiency ηD cannot be applied and some
alternative ways have to be used instead. The merit coefficient mc is introduced
as a replacement of the total propulsive efficiency ηD in order to be able to express
the performance of the thruster. The merit coefficient is defined as:
(KT /π)3/2
mc = (2.69)
KQ
As it has already been mentioned, a vessel operates in full power while it is connected
on a bollard on shore through a tow-line while bollard pull testing. This force applied
on the line is represented in tons and is defined as:
Ttot
TBP = (2.70)
g
Bollard pull condition is really important for tug boats among other types of vessels.
However, it should be acknowledged that tug boats usually operate in really low
power while high power is used only for a short period of time, which is usually
less than the full power that bollard pull condition indicates. Though, full power is
mainly used only during bollard pull testing during the vessel’s sea trials.
30
2. Theory
MRF is a steady state approach, which indicates a frozen rotor hypothesis. The
RANS equations are solved for both the stationary parts and the rotor while the
rotation effect of the rotor is taken into account by including the coriolis and cen-
trifugal forces in additional source terms in the momentum equations.
SMI is a transient approach, which indicates a rotation of the rotor relative to the
stationary parts. In a similar way as before, the RANS equations are solved for both
the stationary parts and the rotor while the rotation effect of the rotor and the un-
steady interactions between the stationary parts and the rotor are taken into account
by solving the conservation equations on a moving (i.e. sliding) mesh. Moreover,
the rotation effect of the moving volume mesh is taken into account by including
the mesh motion flux in the face mass flux computation in the convective terms of
the governing equations. Finally, the position of the mesh of the rotor is updated
after every time step.
Here it should be mentioned that for both the MRF and SMI approaches, an Ar-
bitrary Mesh Interpolation (AMI) takes place between the stationary parts and the
rotor.
31
2. Theory
32
3
Geometry
In this chapter, the main particulars of the vessel and the machinery that are used
within this project are presented. Moreover, an explanation and representation of
the geometries follows.
Propulsion
Make 2x Caterpillar
Type Simplified MTA627
Propellers
Make 2x Caterpillar
Type FPP
Diameter, DP 2.70 m
Expanded blade area ration, EAR 0.67
Design pitch ratio at 0.7r 1.136
No. blades, Z 4
33
3. Geometry
3.3 Hull
In Figure 3.2, the azimuth thrusters and the hull that are studied within this project
are presented. The lid that is mentioned above is removed and the gear case hous-
ing is connected directly to the hull. Moreover, the hull is split into two different
components: the main hull and the skeg, which is the "fin" under the hull that is
used for stability during sailing.
Moreover, the azimuth thruster is rotated 6◦ around x-axis and 6◦ around y-axis as
shown in Figure 3.3.
Finally, it is important to keep in mind that during the CFD simulations a symmetry
plane is used on top of the domain while the water surface in Figure 3.2 is used only
for visualization and understanding of the geometries.
34
3. Geometry
35
3. Geometry
36
4
Methodology
In this chapter, the methodology used in order to accomplish the CFD simulations
is presented. In the first sections the three main steps of CFD (i.e. pre-processing,
solving and post-processing) are presented while in the end the system used to run
OpenFOAM is demonstrated.
4.1 Pre-processing
4.1.1 Definition of the computational domain
The main geometries used within this study project are modelled prior to this study
by Caterpillar Propulsion AB, using Computer Aided Design (CAD) software pack-
ages. These geometries are later provided to the author and researcher of this
project. However, small geometrical changes are made by the author using the
Computer Aided Engineering (CAE) software ANSA pre-processor v.19.0.1. The
flow domain is then defined using ANSA pre-processor, which is built around the
main geometries. Here it should be noted that all the simulation models are defined
in full-scale.
In Figure 4.1, the geometry of Case 1 is presented. A shaft that is connected to the
hub and to the inlet of the domain is added in order to get rid of the stagnation
point that would be created at the center of the hub in case the shaft was not added.
37
4. Methodology
The velocity of the fluid would end up being zero at the stagnation point and thus,
strange flows at this region would be created.
In Figure 4.2, the domain of Case 1 is presented. The dimensions of the domain are:
• Diameter d = 20DP = 54m
• Length L = 30DP = 81m
38
4. Methodology
In Figure 4.3, the geometry of Case 2 is presented while the domain of Case 2 is
identical to Case 1. As already mentioned in Chapter 3, the lid is an additional
geometry that does not exist in reality since the azimuth thruster is connected
directly to the hull. However, its presence is important in order to have a closed
geometry so that no strange flow is created around this region.
(a) Azimuth thruster pressure side. (b) Azimuth thruster suction side.
39
4. Methodology
while in Figure 4.5, the domain of Case 3 is presented. The dimensions of the
domain are:
• Width W ≈ 20DP = 2.3LP P = 52m
• Height H ≈ 20DP = 2.3LP P = 52m
• Length L ≈ 58DP = 6.9LP P = 156m
Looking at both the geometry and the domain of Case 3, it is clear that the the
hull geometry is split into half and the reason is due to the geometrical symmetry.
Thus, there is no need of simulating the whole geometry since this would result
into a much more expensive computational simulation. A symmetry boundary layer
during the CFD simulations is set on the top and side boundaries instead.
Initial step of the meshing is the creation of a surface mesh with triangle shapes.
Then, layers are created around the boundaries, which are a type of volume mesh
that are created by the extrusion of the triangle surface mesh. In the next step, a
volume mesh is created within the whole domain, which consists of elements with
pyramid, tetrahedral and hexahedral shapes.
40
4. Methodology
In Figure 4.6, the different types of mesh used in Case 1 are presented as an example.
More precisely, the surface mesh appearing on the blades, the hub and the inner part
of the nozzle is highlighted with magenta color, the layers around the solid geometries
are highlighted with green and the volume mesh is highlighted with brown. Finally,
the surface mesh on the nozzle is highlighted with lilac color.
Finally, the triangle (i.e. trias) and the square (i.e. quads) elements of the surface
mesh are converted into polygons, the extruded triangle shapes of the layers are
converted into extruded polygon shapes and the pyramid, tetrahedral (i.e. tetras),
pentahedral (i.e. pentas) and hexahedral (i.e. hexas) elements of the volume mesh
are converted into polyhedral elements.
41
4. Methodology
The cell sizes vary from 0.027% · DP ≈ 0.75mm at regions that require really high
accuracy (i.e. leading and trailing edge of the propeller blades) to 124.000% · DP ≈
3348.00mm at regions that require lower accuracy (i.e. domain). Moreover, seven
layers are created around the boundaries with a growth factor of 1.5 and the first
layer having a height with an aspect ratio of 0.05 (i.e. first height = base length ·
0.05). In the following tables the total numbers of cells of all the three cases used
within this study are demonstrated. The naming of the following elements and
categories is presented as in ANSA Pre-processor v.19.0.1. The shell elements are
surface elements that are referring to the trias, quads and polygons. On the other
hand, the volume elements are referring to the the pyramids, tetras, pentas, hexas
and polyhedrals. The layers that have already been discussed are part of the volume
elements.
42
4. Methodology
The quality of the final mesh is highly connected to the quality of the surface mesh,
thus the surface mesh is the one with the highest importance and one should put a
lot of effort into it. A detailed presentation of the different surface mesh types and
cell sizes follows in Appendix B.
43
4. Methodology
Second step is to define the boundary conditions for each field that needs to be
solved (i.e. velocity U , pressure p). The boundary conditions are located in the
<case>/0/* directory (e.g. <case>/0/U ).
Final step is to define the wall functions for the turbulence models (i.e. turbulent
kinetic energy k, dissipation rate of turbulent kinetic energy ε, specific dissipation
rate of turbulent energy ω, turbulent viscosity field νt ). The wall functions are
treated in the same way as the boundary conditions, which means that the wall
functions are also applied on individual patches. The wall functions are located in
the <case>/0/* directory (e.g. <case>/0/nut).
44
4. Methodology
• Azimuth
– Azimuth body (MTA) inner
– Azimuth body (MTA) outer
– Stay right
– Stay left
• Domain (in Case 3)
– Domain inner
– Domain outer
– Domain top
– Domain bottom
It should be mentioned that the naming of the patches is free of choice. On the other
hand, the geometric patch types, boundary conditions and wall functions should be
set in openFOAM as they appear in the following tables. A thorough explanation
of the patch types and the wall functions is given by F. Liu (2017) for a full under-
standing of the wall functions.
The boundary conditions used within this project are explained in the following list
(OpenCFD Ltd, 2018):
• calculated: This boundary condition is not evaluated. It is rather assumed
that the value is assigned vie field assignment.
• fixedValue: This boundary condition sets a fixed value constraint.
• zeroGradient: This boundary condition sets a zero gradient condition from
the patch internal field to the patch faces.
• cyclicAMI: This boundary condition determines a cyclic condition between
a pair of boundaries where communication between the patches is achieved
using an Arbitrary Mesh Interpolation (AMI).
• movingWallVelocity: This boundary condition sets a velocity condition for
cases with moving walls.
• slip: This boundary condition sets a slip constraint.
Finally, the patches, patch types, boundary conditions and wall functions are pre-
sented in Table 4.4, Table 4.5 and Table 4.6. Here it should be mentioned that the
slip boundary condition is the same as symmetry boundary condition for the fields
over which the boundaries are imposed. However, symmetry boundary condition
alters also the fields internal to the solvers. Thus, a slip condition is used.
45
46
Patch Geometric patch type Boundary conditions
boundary U p p_rgh
Inlet patch fixedValue zeroGradient zeroGradient
Outlet patch zeroGradient fixedValue fixedValue
4. Methodology
47
4. Methodology
Patch Geometric patch type Boundary conditions
boundary U p p_rgh
48
Inlet patch fixedValue zeroGradient zeroGradient
Outlet patch zeroGradient fixedValue fixedValue
Domain∗ patch slip slip slip
Blades∗ wall movingWallVelocity zeroGradient zeroGradient
Hub wall movingWallVelocity zeroGradient zeroGradient
4. Methodology
4.2 Solving
49
4. Methodology
Moreover, two types of modelling regarding the mesh take place within the project:
• MRF, which is specified in <case>/system/fvOptions.
• SMI, which is specified in <case>/constant/dynamicMeshDict.
4.3 Post-processing
4.3.1 Visualization of the simulation results
Initial step of the post-processing is to look at flow variable fields, which are a
results of the CFD simulations. The following flow fields are observed thoroughly
in ParaView v.5.6.0:
• Velocity field U
• Pressure field p
• The turbulent kinetic energy field k
• The specific dissipation rate of turbulent energy field ω
Next step is to plot the convergence of the forces F and moments M of all the cases
50
4. Methodology
and for each advance coefficient J and part (e.g. blades, nozzle, hull) that are lo-
cated in <case>/postProcessing/<part>/0/forces.dat. This is done in MATLAB
R2018b.
Later on, by the use of these mean values the following variables are calculated:
• Mean thrust coefficient KT ,
• Mean torque coefficient 10 · KQ ,
• Mean open water efficiency η0 .
Note: Due to simplification purposes, from now on all the forces, moments, co-
efficients and open water efficiencies will not be referred as mean values and the
overline of these symbols will be neglected, even though they are all derived from a
mean value.
The thrust deduction factors t are then calculated in order to describe the in-
crease of the resistance due to the suction of the propeller, determine the generated
thrust and obtain the different contributions of the gear case housing and the hull
from a thrust perspective:
• Case 1
0
t= =0 (4.1)
Tblades + Tnozzle
• Case 2
−Rgear_case_housing
t= (4.2)
Tblades + Tnozzle
• Case 3
−(Rgear_case_housing + Rhull )
t= (4.3)
Tblades + Tnozzle
Next step is to calculate the merit coefficients mc , which are a replacement of the
total propulsive efficiency ηD :
• Case 1
51
4. Methodology
• Case 2
[(KT blades + KT nozzle + KT gear_case_housing )/π]3/2
mc = (4.5)
KQblades
• Case 3 (excluding hull)
[(KT blades + KT nozzle + KT gear_case_housing )/π]3/2
mc = (4.6)
KQblades
• Case 3 (including hull)
[(KT blades + KT nozzle + KT gear_case_housing + KT hull )/π]3/2
mc = (4.7)
KQblades
Here it should be mentioned that the merit coefficient of Case 3 is calculated by
both including and excluding the contribution of the hull. This is done in order to
obtain the flow field that is derived from the case simulation with the hull (i.e. Case
3) but also at the same time to include or exclude the contribution of the hull itself.
Then, the open water characteristic diagrams for all the cases are generated.
Both the MRF and SMI simulation results that have been performed in OpenFOAM
are plotted for comparison and evaluation purposes.
A complete open water diagram is plotted for Case 1 with both MRF and SMI
results in order to get an overall idea of the behaviour of this specific arrangement
in bollard pull and free sailing conditions and in both steady state and transient
approaches. The values that are plotted are the following:
• MRF results for J = 0.00, 0.10, 0.20, 0.30, 0.40, 0.50, 0.60, 0.65, 0.70, 0.75, 0.80
0.85, 0.90, 1.00,
• SMI results for J = 0.00, 0.20, 0.40, 0.60, 0.80.
The SMI results are chosen for J = 0.00 and the MRF results for all the other J val-
ues and a combined MRF-SMI open water characteristics diagram is created. This
diagram, that from now on is referred to as "hybrid diagram", is later used as Case 1.
A complete open water diagram is also plotted for Case 2 but this time with only SMI
results in order to get an overall idea of the behaviour of this specific arrangement
in bollard pull and free sailing conditions and in transient approach. The values
that are plotted are the following:
• SMI results for J = 0.00, 0.20, 0.40, 0.60, 0.70, 0.80, 0.90, 1.00.
In the end, only the values for J = 0.00 are plotted for Case 3 with SMI results but
this time with the only focus in bollard pull condition and in transient approach.
Finally, the open water case (i.e. Case 1) is compared with the Wageningen CD
Series in three different ways (i.e. brake power similarity, thrust power similarity
and geometrical similarity) in order to develop an evaluation tool which can be used
in the future in order to evaluate the early stages of a project.
It should be noted that all the above calculations and plots are done in MATLAB.
52
4. Methodology
Through the domain decomposition, the geometry and the flow fields are split and
then distributed into separate processors, which are then separately solved. After
the completion of the simulation the decomposed cases are reconstructed into one
case.
The steady state simulations take about two hours to be completed while the tran-
sient simulations are run overnight.
53
4. Methodology
54
5
Results and data analysis
In this chapter, the results of all the CFD simulations are presented. In addition,
discussion of the results and data analysis is achieved. In the first section of this
chapter a data analysis in both free sailing and bollard pull conditions is achieved
in order to get an overall idea of the behaviour of all the cases. The hydrodynamic
interactions are presented in the second section only in bollard pull condition since
this is the main focus of this study project.
In Figure 5.1, two open water characteristic diagrams derived from the SMI (see
Table 5.2) and MRF (see Table 5.3) results are presented. Since this comparison is
done only for Case 1, the hydrodynamic coefficients discussed are the thrust coeffi-
cients of the blades KT blades and nozzle KT nozzle , the torque coefficient of the blades
10 · KQblades and the open water efficiency η0 . Moreover, identical rotational speed
n = 191rpm for all the advance coefficients J has been used. In this way, the differ-
ence of the resulting thrust and torque can be observed.
In Figure 5.2, the already mentioned hybrid open water characteristics diagram that
combines the SMI results for J = 0.00 and the MRF results for J = 0.10 − 1.00 is
presented.
55
5. Results and data analysis
The values of the hybrid diagram are presented in Table 5.1. The complete results
are fully presented in Table 5.2 for the SMI and in Table 5.3 for the MRF.
56
Blades Nozzle Open water
efficiency
J T KT Q 10KQ T KT Q 10KQ η0
[−] [kN] [-] [kNm] [-] [kN] [-] [kNm] [-] [-]
0.00 190.98 0.3460 94.15 0.6317 208.20 0.3772 -0.09 -0.0006 0.0000
0.10 190.97 0.3460 93.22 0.6255 158.92 0.2879 -0.14 -0.0009 0.1613
0.20 188.07 0.3407 92.10 0.6179 126.03 0.2283 -0.13 -0.0009 0.2931
0.30 183.15 0.3318 90.20 0.6052 97.40 0.1764 -0.12 -0.0008 0.4010
0.40 176.08 0.3190 87.44 0.5867 72.84 0.1320 -0.11 -0.0008 0.4893
0.50 166.82 0.3022 83.78 0.5621 52.16 0.0945 -0.10 -0.0007 0.5616
0.60 155.19 0.2811 79.09 0.5306 35.00 0.0634 -0.08 -0.0005 0.6201
0.65 148.43 0.2689 76.31 0.5120 27.61 0.0500 -0.07 -0.0005 0.6444
0.70 139.82 0.2533 72.72 0.4879 17.96 0.0325 -0.06 -0.0004 0.6526
0.75 119.69 0.2168 64.10 0.4301 -0.37 -0.0007 -0.05 -0.0004 0.5999
0.80 105.18 0.1905 57.75 0.3875 -11.90 -0.0216 -0.05 -0.0003 0.5552
0.85 89.97 0.1630 51.14 0.3431 -23.47 -0.0425 -0.06 -0.0004 0.4749
0.90 73.63 0.1334 44.09 0.2958 -35.27 -0.0639 -0.06 -0.0004 0.3365
1.00 37.47 0.0679 28.35 0.1902 -59.44 -0.1077 -0.06 -0.0004 0.3330
57
5. Results and data analysis
5. Results and data analysis
The total forces and moments extracted from the OpenFOAM post-processing sim-
ulation files for both MRF and SMI, in bollard pull condition (i.e. J = 0.00) and
for all the three directions are presented in the following figures.
x-direction
y-direction
58
5. Results and data analysis
z-direction
59
5. Results and data analysis
Looking at Figure 5.1 and by comparing the complete results for SMI in Table
5.2 and the complete results for MRF in Table 5.3, it is obvious that most of the
MRF results are following the SMI results. However, a big difference is observed at
J = 0.00 that bollard pull condition occurs. Observing the convergence plots above,
it is obvious that the x-direction results seem to be converged for both MRF and
SMI. However, the MRF results for y-direction and z-direction are not converged
while the SMI results are. Moreover, it should be mentioned that an oscillation of
the results appears for the SMI. Nevertheless, this oscillation is constant and thus
an average value of the last 720 iterations is taken.
Considering that the results for bollard pull condition seem more realistic by taking
into account the CFD results of other similar studies and by observing the conver-
gence of SMI and the convergence issues of MRF, the already mentioned statement
that the transient SMI methodology gives much more accurate results
compared to the steady state MRF for low J values can be supported. Mean-
while, the results for medium J values computed with the MRF approach are quite
reasonable.
Since the transient SMI methodology gives much more accurate results, the ideal
would be to run every simulation with SMI. On the other hand, SMI is much more
computationally expensive compared to MRF.
60
Blades Nozzle Open water
efficiency
J T KT Q 10KQ T KT Q 10KQ η0
[−] [kN] [-] [kNm] [-] [kN] [-] [kNm] [-] [-]
0.00 190.98 0.3460 94.15 0.6317 208.20 0.3772 -0.09 -0.0006 0.0000
0.20 188.78 0.3420 93.29 0.6259 132.26 0.2396 -0.12 -0.0008 0.2958
0.40 176.08 0.3190 87.44 0.5867 72.84 0.1320 -0.11 -0.0008 0.4893
0.60 157.22 0.2848 80.61 0.5408 36.29 0.0657 -0.09 -0.0006 0.6190
0.80 105.18 0.1905 57.75 0.3875 -11.90 -0.0216 -0.05 -0.0003 0.5552
61
5. Results and data analysis
The values of Case 2 are presented in Table 5.4. It should be mentioned that while
the hydrodynamic coefficients of Case 1 are derived from the hybrid MRF/SMI re-
sults, the coefficients of Case 2 are obtained only from the SMI results since only
SMI was used in this case in order to have slightly better accuracy for the gear case
housing within all the open water diagram.
Looking at Figure 5.9, it is noticed that the propeller thrust is really similar in
both cases for J = 0.00 − 0.40 while the difference starts to increase after this point.
For J = 0.70 − 1.00 the thrust is much higher for Case 2. The difference of the pro-
peller torque appears to be quite large between J = 0.00 − 0.40 while after this
point the difference gets even larger with the largest difference occurring between
J = 0.70 − 1.00. Looking at the nozzle thrust, it is clear that in Case 2 the thrust
is much higher than in Case 1 within the whole diagram. Moreover, it can be seen
that the gear case housing thrust remains constant within the whole diagram.
Since the thrust is having a negative value, it is considered as resistance instead.
Finally, it is observed that the open water efficiency is lower for Case 2 between
J = 0.00 − 0.70. Looking specifically at J = 0.00, the open water efficiency looses
its meaning (i.e. open water efficiency is always zero since the velocity of the ship is
zero). In this moment, the merit coefficient is introduced, however this is further
discussed in Section 5.2.
62
Blades Nozzle Gear Case Housing Open water
efficiency
J T KT Q 10KQ T KT Q 10KQ T KT Q 10KQ η0
[−] [kN] [-] [kNm] [-] [kN] [-] [kNm] [-] [kN] [-] [kNm] [-] [-]
0.00 193.00 0.3496 94.99 0.6373 242.24 0.4388 -0.10 -0.0007 -38.81 -0.0703 -1.10 -0.0074 0.0000
0.20 191.09 0.3462 94.14 0.6316 160.72 0.2912 -0.10 -0.0007 -37.56 -0.0680 1.70 0.0114 0.2869
0.40 180.09 0.3262 89.84 0.6028 101.65 0.1841 -0.08 -0.0005 -33.40 -0.0605 0.49 0.0033 0.4751
0.60 162.42 0.2942 82.57 0.5540 59.15 0.1072 -0.06 -0.0004 -30.25 -0.0548 -2.72 -0.0183 0.5974
0.70 149.66 0.2711 77.14 0.5176 42.56 0.0771 -0.05 -0.0003 -28.49 -0.0516 -4.50 -0.0302 0.6384
0.80 121.62 0.2203 65.00 0.4361 14.84 0.0269 -0.07 -0.0005 -28.56 -0.0517 -3.09 -0.0207 0.5707
0.90 88.37 0.1601 50.03 0.3357 -9.49 -0.0172 -0.07 -0.0005 -29.70 -0.0538 -2.64 -0.0177 0.3801
1.00 55.02 0.0997 35.85 0.2405 -33.30 -0.0603 -0.13 -0.0009 -30.21 -0.0547 -3.53 -0.0237 -0.1018
63
5. Results and data analysis
5. Results and data analysis
In Figure 5.10, the pressure distribution in bollard pull condition of Case 1 is pre-
sented. The pressure distribution seems normal with higher pressure at the pressure
side and lower pressure at the suction side. Looking at Figure 5.11 that presents
the velocity field of Case 1, it is clearly demonstrated by the velocity magnitude
field that the flow is straight. A small decrease of the velocities in the centerline is
observed, which is due to the hub that is blocking the flow.
In Figure 5.12, the pressure distribution of Case 2 is presented. The pressure dis-
tribution seems normal. Furthermore, it is observed that Case 2 has an inflow with
lower pressure compared to Case 1 due to the gear case housing that is blocking
the flow. Thus, the pressure difference in Case 2 is higher, which indicates higher
acceleration of water from the pressure side to the suction side (i.e higher advance
velocity). As a result, the higher thrust power that is created from the propeller
towards the outflow is explained. Looking at Figure 5.13 that presents the velocity
field of Case 2, it is observed that some fluctuations have been introduced to the
flow. Most likely this result occurs due to the change of the quality of the mesh
within this region. However, further investigation should be done in order to answer
whether this result is dependent on the mesh or whether it occurs due to physical
phenomena.
64
5. Results and data analysis
In addition, looking at Figure 5.14 that presents a closer view of the velocity field of
Case 2, it can be observed that the inlet velocity at the nozzle leading edge is higher
than the inlet velocity at the blades, hence the higher increase of the nozzle thrust
compared to the blade thrust in Case 2. However, due to the increase of resistance
at the same time, the net thrust generated in Case 2 is 0.68% lower than Case 1
(see Table 5.12).
65
5. Results and data analysis
It is obvious that the results are presented in two different perspectives. By exam-
ining the first perspective (see Table 5.5), the values are derived from the actual
thrust and presented in respect to the propeller-axis that the thrust is created. By
examining the second perspective (see Table 5.6), the values are derived from the
x-component of the actual thrust and presented in respect to the x-axis. Looking at
the same tables, it could be noted that the hull values are calculated only for x-axis
since the resistance of the hull is created only in this direction. Thus, the hull values
are identical in both tables.
66
Blades Nozzle
(propeller-axis) (propeller-axis)
J T KT Q 10KQ T KT Q 10KQ
[−] [kN] [-] [kNm] [-] [kN] [-] [kNm] [-]
0.00 199.64 0.3617 97.81 0.6563 229.82 0.4163 -38.63 -0.2592
Blades Nozzle
(x-axis) (x-axis)
J T KT Q 10KQ T KT Q 10KQ
[−] [kN] [-] [kNm] [-] [kN] [-] [kNm] [-]
0.00 198.55 0.3597 97.27 0.6527 228.56 0.4140 -38.42 -0.2578
67
5. Results and data analysis
5. Results and data analysis
In a similar way as before, the open water efficiency looses its meaning and the total
propulsive efficiency ηD becomes zero. Hence, the merit coefficient is used as a re-
placement of the total propulsive efficiency, which is further discussed in Section 5.2.
In Figure 5.15, the pressure distribution of Case 3 is presented. The pressure dis-
tribution seems reasonable. Furthermore, it is observed that Case 3 has an inflow
similar to Case 2. The increase of the thrust in Case 3 compared to Case 1 is ex-
plained in a similar way as Case 2. Looking at Figure 5.16 that presents the velocity
field of Case 3, it is observed that a small "step" is observed at the top of the flow
and under the hull. Most likely this result occurs due to a similar reason as Case 2,
meaning the change of the quality of the mesh within this region. However, further
investigation should also be done.
Finally, in a similar way as Case 2, the inlet velocity at the nozzle leading edge is
higher than the inlet velocity at the blades, hence the higher increase of the nozzle
thrust compared to the blade thrust in Case 3. However, the net thrust generated
68
5. Results and data analysis
in Case 3 is 4.09% lower than Case 1 due to the increase of the resistance (see Table
5.12).
propCalcCD2.3.1 needs some input parameters, such as brake power PB , shaft effi-
ciency ηS , expanded blade area ratio, propeller diameter DP , ship velocity VS , wake
fraction wT , rotational speed n, design pitch ratio at 0.7r and blade number. By the
use of tabulated data from the Wageningen CD Series as well as from Caterpillar
Propulsion AB test data, this tool predicts the propeller thrust, nozzle thrust, pro-
peller torque, pitch ratio at 0.7r etc. These predicted values are used for evaluation
of the early stages of a project.
In Figure 5.17, the hydrodynamic coefficients of Case 1 and the Wageningen CD Se-
ries are demonstrated. A brake power similarity is achieved between the compared
cases by setting up the same brake power at the Wageningen CD Series as in Case
1. All the hydrodynamic coefficients are included while the pitch ratio at 0.7r is also
plotted, which shows the geometrical similarity between the compared cases.
The brake power PB is calculated by using the torque from the blades Q (see Table
5.2).
Table 5.7: Calculation of the brake power PB by using the SMI results of Case 1.
69
5. Results and data analysis
Then, the brake power PB together with the ship’s velocity VS as well as a series of
constant parameters such as:
• nS = 1
• EAR = 0.67
• D = 2.7m
• w=0
• n = 191rpm
• Design pitch ratio at 0.7r = 1.136
• No. blades = 4
are set as input parameters into the propCalcCD2.3.1 tool.
Looking at Figure 5.17, it is noticed that the propeller thrust of Case 1 is higher
than the one that the Wageningen CD Series suggests. For a better understanding,
it is recommended to look at Table 5.8 for the Wageningen CD Series and at Table
5.12 for Case 1. On the other hand, the propeller torque appears to be identical
(i.e. brake power similarity is achieved). The nozzle thrust behaves in a similar
70
5. Results and data analysis
way for most of the advance coefficients while there is a quite big difference between
J = 0.00−0.10 with Case 1 having a higher thrust at bollard pull condition. Finally,
the pitch ratio at 0.7r remains almost identical within the whole diagram for the
Wageningen CD Series, which has values really close to the constant pitch ratio of
Case 1, which is 1.136.
In this section, the second comparison of Case 1 and the Wageningen CD Series is
achieved.
In Figure 5.18, the hydrodynamic coefficients of Case 1 and the Wageningen CD Se-
ries are demonstrated. A thrust power similarity is achieved between the compared
cases by setting up the same thrust power at the Wageningen CD Series as in Case 1.
The hydrodynamic coefficients and the pitch ratio at 0.7r are included as previously.
The thrust power PT is calculated by using the thrust from the propeller blades Tp
(see Table 5.2).
71
5. Results and data analysis
Table 5.9: Calculation of the thrust power PT by using the SMI results of
Case 1.
Then, the thrust power PT together with the ship’s velocity VS as well as the same
constant parameters as previously are set as input parameters into propCalcCD2.3.1.
In this section, the third and last comparison of Case 1 and the Wageningen CD
Series is achieved.
72
5. Results and data analysis
Figure 5.19: Open water characteristics: Case1/CD Series with constant pitch
ratio at 0.7r comparison.
In this case, a geometrical comparison is achieved. Thus, the pitch ratio at 0.7r is
the only input parameter into propCalcCD2.3.1, which is same as before:
• Design pitch ratio at 0.7r = 1.136
73
5. Results and data analysis
Finally, it should be mentioned that there are other differences between the geo-
metrical properties of the propeller as well as differences between the geometrical
properties of the nozzle between Case 1 and the Wageningen CD Series, thus the
plotted Wageningen CD Series values might differ in reality. Thus, the
above comparisons can only provide an approximation of the behaviour.
74
5. Results and data analysis
Finally, comparing the MRF (see Figure 5.22) and SMI (see Figure 5.24) results for
the advance coefficient J = 0.80 where the flow separation has already occured, it is
obvious that the separation is of similar scale. On the other hand, the flow field of
MRF is slightly different from the one of the SMI, with the SMI flow field resulting
in a more straight flow compared to the one of MRF.
5.1.6 Cavitation
Looking at the three following figures, it is observed that a really low pressure in
Case 2 and Case 3 occurs at the connection point of the gear case housing and the
nozzle as well as at the tip of the blade that is behind the gear case housing. This
observation indicates that cavitation will first appear at these points.
75
5. Results and data analysis
In this low pressure region the pressure value derived from the CFD simulations,
which is read from ParaView, is equal to:
−200kP a
1atm = 101.325kP a
where:
z is the distance of the upper part of the nozzle and the water surface.
By summing up the three pressure values above, the actual pressure at the connec-
tion point of the gear case housing and the nozzle is calculated, which is equal to
−74.54kP a. This is an unrealistic pressure since the vapor pressure for sea water at
20◦ C is equal to 2.29kP a. In other words, this means that the minimum value that
the pressure can get is 2.29kP a, while when this minimum value is exceeded, cav-
itation will occur. Thus, this should be further investigated by running cavitation
CFD simulations.
76
5. Results and data analysis
In Table 5.12, the hydrodynamic coefficients of all the cases are decomposed and a
detailed presentation of the thrust of each component is presented. Furthermore,
the torque coefficients derived from the propeller blades are also presented. It can
also be observed that Case 3 is divided into two parts. This is done in order to
express the forces in respect to the x-axis as well as the axis that the propeller is
77
5. Results and data analysis
rotating.
The above table is then presented in percentages, which makes it easier to identify
the contribution of each component to the thrust and resistance accordingly.
Looking at Table 5.13, it is observed that the nozzle has the highest contribu-
tion in thrust for all cases. Besides that, the gear case housing has the highest
contribution to resistance for all cases.
In the following tables (see Table 5.14, Table 5.15, Table 5.16 & Table 5.17), a com-
parison among the cases is presented by demonstrating values derived from the CFD
simulation in OpenFOAM as well as by demonstrating the calculated hydrodynamic
interactions.
In Table 5.14, Tnet represents the net sum of the total thrust (i.e. T − R), Tnet,deduct
represents its deduction and TBP represents the bollard pull force. In Table 5.15,
KT,T represents the sum of the thrust coefficients that are creating thrust (i.e.
KT p + KT n ), KT,R represents the sum of the thrust coefficients that are derived
from the parts that are creating resistance (i.e. Case 2: KT gear_case_housing and Case
78
5. Results and data analysis
3: KT gear_case_housing + KT hull ), KT,net represents the net sum of all the thrust co-
efficients (i.e. KT,T − KT,R ), mc represents the merit coefficient and finally mc,deduct
represents its deduction.
In Table 5.16, n1995kw represents the rotational velocity of the propeller at a fixed
delivered power and n1995kw,deduct represents its deduction. Here it should be men-
tioned that a delivered power of PD = 1995kW is derived from the main engine’s
power (see Table 3.1) with a reduction factor of 5%. Looking again at the definition
of the variables, T1995kw represents the generated thrust at a fixed delivered power
and T1995kw,deduct represents its deduction. PD,191rpm represents the delivered power
at a fixed rotational speed and PD,191rpm,deduct represents its deduction. Finally,
T191rpm represents the generated thrust at a fixed rotational speed and T191rpm,deduct
its deduction. It should be mentioned that the 191 rpm is an estimation of another
CFD study that was conducted within Caterpillar Propulsion AB.
Note: All the calculations below are presented for half the vessel and one az-
imuth thruster unit. Moreover, the values presented in Table 5.14 and Table 5.15
are derived from the CFD results while in Table 5.16 and Table 5.17 the values are
calculated by assuming constant delivered power and rotational velocity accordingly.
79
80
Case Parts included T [kN] R [kN] t [%] Tnet [kN] Tnet,deduct [%] TBP [ton]
Case 1 Blades+Nozzle 399.17 0.00 0.00 399.17 0.00 40.69
Case 2 Blades+Nozzle+Gear Case Housing 435.24 38.81 8.92 396.44 0.69 40.41
Case 3 Blades+Nozzle+Gear Case Housing 427.11 35.52 8.32 391.58 1.90 39.92
excluding hull
Case 3 Blades+Nozzle+Gear Case Housing 427.11 40.33 9.44 386.50 3.17 39.40
5. Results and data analysis
Table 5.14: Case comparison: Presentation of the thrust, resistance, thrust deduction,
net forces, deduction of the net forces and bollard pull forces.
Case KT,T [-] KT,R [-] 10KQ [-] KT,net [-] mc [-] mc,deduct [%]
Case 1 0.7231 0.0000 0.6317 0.7231 1.7481 0.00
Case 2 0.7885 0.0703 0.6373 0.7182 1.7150 1.90
Case 3 0.7737 0.0710 0.6527 0.7027 1.6210 7.27
excluding hull
Case 3 0.7737 0.0797 0.6527 0.6935 1.5892 9.09
including hull
Table 5.15: Case comparison: Presentation of thrust, torque, net thrust coefficients,
merit coefficient and the deduction of the merit coefficient.
Case n1995kw [rpm] n1995kw,deduct [%] T1995kw [kN] T1995kw,deduct [%]
Case 1 194.71 0.00 414.82 0.00
Case 2 194.14 0.31 409.56 1.27
Case 3 192.60 1.08 394.45 4.91
excluding hull
Case 3 192.60 1.08 389.28 6.16
including hull
81
5. Results and data analysis
5. Results and data analysis
Looking at Table 5.14, it can be seen that 8.92% of the thrust deduction occurs
at Case 2 when the gear case housing is placed behind the propeller and the nozzle.
By placing the whole azimuth thruster unit behind the hull and by excluding the
resistance created by the hull the thrust is deducted by 8.32%. Finally, the thrust
is deducted by 9.44% by including the resistance of the hull. Thus, the highest
thrust deduction occurs at Case 2 by placing the gear case housing be-
hind the propeller and the nozzle while the hull has a smaller contribution to
the thrust deduction.
By reviewing Table 5.16, it can be identified that the highest decrease of rota-
tional speed of the propeller as well as the highest thrust decrease occurs
in Case 3. However, the decrease of both values is almost unnoticeable.
In the end, by reviewing Table 5.17, similar conclusions as in Table 5.16 can be
drawn.
In Table 5.18, TBP,1995kw,CF D is representing the bollard pull force that is derived
from the generated thrust at a fixed delivered power T1995kw (see Table 5.16). This
generated thrust is calculated based on the CFD simulation results. Moreover,
TBP,1995kw,CD is the bollard pull force that the Wageningen CD Series gives, which
is calculated in propCalcCD2.3.1 by using as inputs the delivered power PD =
1995kW and the rotational velocity at a fixed delivered power n1995kw among other
parameters.
82
5. Results and data analysis
A reduction factor of 7% is added at TBP,1995kw,CD for the hull (i.e. Case 3 including
hull) while the reductions that appear on Case 2 and Case 3 by excluding the hull
are unknown.
83
5. Results and data analysis
84
6
Conclusions and future work
In this chapter, a conclusion of this study project is drawn and the degree that the
initial purposes are fulfilled within this project is presented. In the end of the chap-
ter, suggestions for further investigations and things that should be kept in mind are
presented as part of the future work.
The CFD modelling of a propeller indicates a relative motion between the stationary
parts and the rotor of a propulsion unit and the vehicle that this propulsion unit
is mounted on. Two methodologies have been used within this project in order to
capture this motion: the steady state Multiple Reference Frame (MRF) and the
transient Sliding Mesh Interface (SMI). By comparing the simulation results of SMI
and MRF and by looking at their convergence plots, it is concluded that the tran-
sient SMI methodology gives much more accurate results compared to the steady
state MRF for low advance coefficient J values. Meanwhile, the results for medium
J values computed with the MRF approach are quite reasonable. Finally, MRF gives
appropriate results at low computational cost for initializing the flow properties of
the CFD simulations in a steady state so then these flow properties can be used to
proceed to transient SMI simulations in order to increase the accuracy of the results.
Looking into the results of interest of this study project, it is concluded that the
contribution of the nozzle is higher than the propeller in terms of thrust while the
contribution of the gear case housing is higher than the hull in terms of resistance.
Thus, the highest thrust deduction occurs on the gear case housing. Bollard pull
condition indicates advance coefficient J and velocity of the ship VS both equal to
zero. As a result, the open water efficiency looses its meaning and the propulsive
efficiency ηD becomes zero. Thus, the performance of the propulsion unit in bollard
pull condition needs to be expressed with the merit coefficient mc , which is a replace-
ment of the total propulsive efficiency ηD . Moving back to the results of interest, the
highest merit deduction occurs when placing the gear case housing behind the hull.
Finally, a higher bollard pull force of the open water case (i.e. Case 1) compared to
the Wageningen CD Series is observed.
During the current study, a symmetry boundary condition has been used at the top
and side of the domain during the CFD simulations. As part of the future work
would be to include the free surface effect at the top boundary instead. It would
also be suggested to run more transient SMI simulations, whether enough resources
are available to proceed to these computationally expensive simulations, in order to
complete the open water characteristic diagrams only with transient SMI results.
85
6. Conclusions and future work
CFD simulations in order to get the towing resistance of the hull for various J values
and investigation of the behavior of the gear case housing and hull in free running
condition would also be suggested. Furthermore, cavitation should be investigated
more by running cavitation CFD simulations. Moreover, comparison of numerical
methods and turbulence models should be included in order to obtain the reliability
of the results. Last but not least, a grid refinement study should be done.
86
References
Bergman, T. L., Lavine, A. S., Incropera, F. P., & Dewitt, D. P. (2011). Fundamen-
tals of Heat and Mass Transfer (7th ed.). Chichester, United Kingdom: John
Wiley and Sons Ltd.
Carlton, J. S. (2007). Marine Propellers and Propulsion (2nd ed.). Burlington, USA:
Elsevier Ltd.
Dyne, G., & Bark, G. (2005). Ship Propulsion: Compendium for part of the MSc
course "Ship resistance and propulsion". Gothenburg, Sweden.
Fletcher, C. A. J. (1991). Computational Techniques for Fluid Dynamics: Specific
Techniques for Different Flow Categories (2nd ed.). Berlin; New York: Springer-
Verlag.
Funeno, I. (2009). Hydrodynamic Optimal Design of Ducted Azimuth Thrusters.
In Proceedings of the First International Symposium on Marine Propulsors -
smp’09, Trondheim, Norway.
Gullberg, P., & Sengupta, R. (2011). Axial Fan Performance Predictions in CFD,
Comparison of MRF and Sliding Mesh with Experiments. doi:https : / / doi -
org.proxy.lib.chalmers.se/10.4271/2011-01-0652
Larsson, L., & Raven, H. C. (2010). The Principles of Naval Architecture Series:
Ship Resistance and Flow. Jersey City, USA: The Society of Naval Architects
and Marine Engineers.
Liu, F. (2017). A Thorough Description Of How Wall Functions Are Implemented
In OpenFOAM. In Proceedings of the CFD with OpenSource Software, 2016,
Edited by Nilsson. H. Gothenburg, Sweden. Retrieved from http://www.tfd.
chalmers.se/~hani/kurser/OS_CFD_2016
Liu, J., Lin, H., & Purimitla, S. R. (2016). Wake field studies of tidal current turbines
with different numerical methods.
Matin, F. (2011). Hydrodynamics of Conventional Propeller and Azimuth Thruster
in Behind Condition. (Master’s thesis, Chalmers University of Technology,
Department of Shipping and Marine Technology).
Menter, F. R., Kuntz, M., & Langtry, R. (2003). Ten Years of Industrial Experience
with the SST Turbulence Model. In Proceedings of the Fourth International
Symposium on Turbulence, Heat and Mass Transfer, Redding, USA.
OpenCFD Ltd. (2018). OpenFOAM User Guide v1812. Retrieved from https : / /
www.openfoam.com/documentation/user-guide/
Tabib, M., Siddiqui, M. S., Rasheed, A., & Kvamsdal, T. (2017). Industrial scale
turbine and associated wake development - comparison of RANS based Ac-
tuator Line Vs Sliding Mesh Interface Vs Multiple Reference Frame method.
87
References
In Proceedings of the 14th Deep Sea Offshore Wind R&D Conferencem EERA
DeepWind’2017, Trondheim, Norway.
Versteeg, H. K., & Malalasekera, W. (2007). An Introduction to Computational Fluid
Dynamics: The Finite Volume Method (2nd ed.). Harlow, England: Pearson
Education Limited.
88
References
89
References
90
Appendices
91
A
Convergence of the simulated
results
In this chapter, the total forces and moments extracted from the OpenFOAM simu-
lations are plotted and their convergence is observed.
The total forces and moments extracted from the OpenFOAM v1806 simulations for
all the cases used within this study project are plotted with MATLAB R2018b and
their convergence is observed. For a better understanding, all the simulated cases
are presented below:
• MRF Simulations at J = 0.00, 0.10, 0.20, 0.30, 0.40, 0.50, 0.60, 0.65,
0.70, 0.75, 0.80, 0.85, 0.90, 1.00,
• SMI Simulations at J = 0.00, 0.20, 0.40, 0.60, 0.80,
• SMI Simulations at J = 0.00, 0.20, 0.40, 0.60, 0.70, 0.80, 0.90, 1.00,
• SMI Simulations at J = 0.00.
All the convergence plots bellow are represented in all the three dimensions (i.e.
x-dir, y-dir and z-dir).
x-dir
(a) MRF.
93
A. Convergence of the simulated results
(b) SMI.
(a) MRF.
(b) SMI.
94
A. Convergence of the simulated results
(a) MRF.
(b) SMI.
95
A. Convergence of the simulated results
(a) MRF.
(b) SMI.
96
A. Convergence of the simulated results
(a) MRF.
(b) SMI.
97
A. Convergence of the simulated results
y-dir
(a) MRF.
(b) SMI.
98
A. Convergence of the simulated results
(a) MRF.
(b) SMI.
99
A. Convergence of the simulated results
(a) MRF.
(b) SMI.
(a) MRF.
100
A. Convergence of the simulated results
(b) SMI.
101
A. Convergence of the simulated results
(a) MRF.
(b) SMI.
102
A. Convergence of the simulated results
z-dir
(a) MRF.
(b) SMI.
103
A. Convergence of the simulated results
(a) MRF.
(b) SMI.
(a) MRF.
104
A. Convergence of the simulated results
(b) SMI.
(a) MRF.
(b) SMI.
105
A. Convergence of the simulated results
(a) MRF.
106
A. Convergence of the simulated results
(b) SMI.
107
A. Convergence of the simulated results
x-dir
108
A. Convergence of the simulated results
109
A. Convergence of the simulated results
110
A. Convergence of the simulated results
111
A. Convergence of the simulated results
y-dir
112
A. Convergence of the simulated results
113
A. Convergence of the simulated results
114
A. Convergence of the simulated results
115
A. Convergence of the simulated results
z-dir
116
A. Convergence of the simulated results
117
A. Convergence of the simulated results
118
A. Convergence of the simulated results
119
A. Convergence of the simulated results
y-dir
120
A. Convergence of the simulated results
z-dir
121
A. Convergence of the simulated results
122
B
Illustration of the different surface
mesh grids
In this chapter, all the initial surface meshes used within this study project are pre-
sented. First, the general mesh grids that are used in all cases are presented while
later on the specific meshes used for each of the three different cases follow.
The definition of the geometry and mesh generation is done in ANSA Pre-processor
v.19.0.1. In Section 2.1.2, the grid generation has been discussed. It is worth men-
tioning in this point that when a geometry is changed, the mesh of this geometry
ends up being different as well. Thus, it is really important to acknowledge that
the general mesh grids mentioned in Section B.1 are slight different from case to case.
A good example to support the statement above is the hub in Case 1 that has
slightly different mesh compared to the hub in Case 2 due to the small geometrical
differences that are needed in order to adapt the hub to the shaft in Case 1 (see
Figure B.5) or to the gear case housing in Case 2 (see Figure B.7a). However, it is
attempted that consistency of the meshes of the same parts between the different
cases is kept, thus the meshes are considered almost identical.
123
B. Illustration of the different surface mesh grids
124
B. Illustration of the different surface mesh grids
125
B. Illustration of the different surface mesh grids
Figure B.2: Hub geometry and its mesh: Isometric view of the hub.
126
B. Illustration of the different surface mesh grids
Here it should be mentioned that the domain of the rotating geometries, also known
as rotor, is the inner volume that is rotating within the outer volume (i.e. the outer
domain). Thus, this configuration is a SMI configuration. The outer boundaries of
the rotor are the interfaces (i.e. inlet and outer interface) and a surface attached to
the inner part of the nozzle. An Arbitrary Mesh Interpolation (AMI) takes place
between the rotor and the outer volume.
127
B. Illustration of the different surface mesh grids
128
B. Illustration of the different surface mesh grids
Here it should be mentioned that part of the inner part of the nozzle (see Figure
B.3a) is geometrically identical with the surface of the rotor that is attached to the
inner part of the nozzle (see Figure B.4a). Thus, almost identical mesh is attempted
to be achieved. Moreover, this part of the inner part of the nozzle as well as the
surface of the rotor that is attached to the inner part of the nozzle are including
some refined regions. These refined regions occur at the regions where the tip of the
blades is placed.
129
B. Illustration of the different surface mesh grids
(b) Inner side of the nozzle, focusing on the refined mesh of the inner part
of the nozzle.
130
B. Illustration of the different surface mesh grids
131
B. Illustration of the different surface mesh grids
Figure B.5: Shaft-hub geometries and their mesh: Isometric view of the shaft
including the hub.
132
B. Illustration of the different surface mesh grids
133
B. Illustration of the different surface mesh grids
The azimuth unit excluding the blades is presented below. On the top of the gear
case housing, a lid has been used during the simulations of Case 2 while in Case 3
the gear case housing is connected directly to the hull.
134
B. Illustration of the different surface mesh grids
Table B.7: Minimum and maximum cell size at the gear case housing.
135
B. Illustration of the different surface mesh grids
Figure B.8: Domain geometry and its mesh: Isometric view of the domain.
136
B. Illustration of the different surface mesh grids
Figure B.9: Hull-azimuth geometries and their mesh: Isometric view of the hull
including the azimuth unit.
137
B. Illustration of the different surface mesh grids
(b) Part of the domain, focusing on the refined mesh of the hull and the wake field
behind it.
138
B. Illustration of the different surface mesh grids
(c) Part of the domain, focusing on the refined mesh of the wake field behind the
hull.
(d) Part of the domain, focusing on the refined mesh of the skeg of the hull.
139
B. Illustration of the different surface mesh grids
(e) Part of the domain, focusing on the refined mesh of the bow of the hull.
140