You are on page 1of 54

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/303676909

Effects of Propeller Geometry on Cavitation

Research · May 2016


DOI: 10.13140/RG.2.1.2863.8325

CITATIONS READS

0 3,030

1 author:

Oliver Ayris

2 PUBLICATIONS   0 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Oliver Ayris on 30 May 2016.

The user has requested enhancement of the downloaded file.


University of Strathclyde

4th Year Individual Project


MEng Naval Architecture with Ocean Engineering

Effects of Propeller Geometry on


Cavitation

Oliver Aristotelis Airis Asimakopoulos

supervised by

Dr. Panagiotis Kaklis

May 30, 2016


Abstract
In this project, a study and evaluation of the relation between geometrical char-
acteristics of marine propellers and the phenomenon of sheet cavitation during their
operation has undertaken. A review of current literature was conducted to establish
the known correlations for propeller optimisation with regards to cavitation perfor-
mance. A validation procedure was conducted to assess the accuracy of the selected
software. Subsequently, the modelling of 6 groups of propellers was achieved, to-
talling 40 models, with each group varying from a parent model in either diameter,
blade number, blade area ratio, pitch, skew or rake through a prescribed range of
values. The cavitation tests were simulated using a limited version of the Propeller
Panel Method (PPB) code provided by the Hamburg Ship Model Basin (HSVA).
The Boundary Element Method (BEM) implemented into the software produced fair
results in open water and poor results for cavitation. The results of the systematic
study are in close agreement with design recommendations in industry, suggesting a
consideration for balance between efficiency and long term costs in the interdepen-
dent selection of the diameter, blade number and blade area ratio. Pitch increases
cavitation and must be cautiously chosen along with the diameter. Skew can indeed
be used to favour cavitation performance effectively, whereas rake has no notable
impact on cavitation.

iii
iv
Acknowledgements
First and foremost, I must thank my supervisor Dr Panagioti Kakli for his sup-
port, understanding and advice in every step throughout the process. I would also
like to express special thanks to Sotiri Chouliara, who repeatedly went out of his
way to assist me when I was in need. This project was also supported by Dr Heinrich
Streckwall and the Hamburg Ship Model Basin who provided me with their useful
tools and guidance. Finally, I would like to show my appreciation towards my family
for their unconditional support and faith in me.

v
Contents

Abstract iii

Acknowledgements v

Table of Contents v

List of Figures viii

List of Tables x

Nomenclature xi
Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

1 Introduction 1
1.1 Cavitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Types of Cavitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Propeller Performance Evaluation Methods . . . . . . . . . . . . . . . 2
1.4 Numericals Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.5 Project Aims and Objectives . . . . . . . . . . . . . . . . . . . . . . . 4

2 Literature Review 5
2.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Diameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.3 Blade Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.4 Blade Area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.5 Pitch Angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.6 Skew . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.7 Warp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.8 Rake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.9 Blade Section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.10 Propeller Hub . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 Propeller Modelling 16
3.1 Validation Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2 Comparative Study Models . . . . . . . . . . . . . . . . . . . . . . . 16

vi
4 Propeller Simulations 21
4.1 Numerical Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2 Open Water Validation . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3 Cavitation Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.4 Comparative Study Setup . . . . . . . . . . . . . . . . . . . . . . . . 24

5 Results 27
5.1 Diameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.2 Blade Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.3 Blade Area Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.4 Pitch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.5 Skew . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.6 Rake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

6 Conclusions and Discussion 36

References 38

vii
List of Figures

1.1 Water phase diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.2 Various types of cavitation . . . . . . . . . . . . . . . . . . . . . . . . 3

2.1 Blade section with back cavitation . . . . . . . . . . . . . . . . . . . . 6


2.2 Optimum diameter selection graph . . . . . . . . . . . . . . . . . . . 7
2.3 Correlation between optimum diameter, blade number and blade area
ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.4 Burrill’s cavitation diagram for uniform flow . . . . . . . . . . . . . . 8
2.5 Blade section pitch P, nose-tail pitch angle θ and inflow angle β . . . 9
2.6 Left – Velocity distributions at (a) positive, (b) ideal and (c) negative
angle of attack. Right – Cavitation bucket diagram . . . . . . . . . . 10
2.7 Pressure distribution for various cavitation numbers at 5◦ angle of
incidence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.8 Observation from experimental tests . . . . . . . . . . . . . . . . . . 11
2.9 Unwrapped view of section showing the relationship between warp
and skew . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.10 Comparison between undistorted, 36◦ and 72◦ skew, and 36◦ and 72◦
warp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.11 Novel camber distribution compared to the NACA08 camberline . . . 13
2.12 Right – The influence of thickness on cavitation inception. Left – The
influence of camber on cavitation inception . . . . . . . . . . . . . . . 14
2.13 Hub shape and its effect on cavitation . . . . . . . . . . . . . . . . . 14

3.1 Solid and wire representation of the parent model . . . . . . . . . . . 18


3.2 CAD drawing of the parent propeller . . . . . . . . . . . . . . . . . . 20

4.1 Grid and mesh of the DTNSRDC Propeller 4381 . . . . . . . . . . . . 22


4.2 Open water performance comparison between PPB and experimental
results by Boswell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3 Total cavitation inception as predicted by PPB compared to the ex-
perimental prediction . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.4 Flowchart demonstrating the derivation of the PPB input variables . 26

5.1 Cavitation with diameter variation at σ = 0.15 . . . . . . . . . . . . . 28


5.2 Sensitivity of cavitation at lower cavitation number with diameter
variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3 Cavitation with blade number variation at σ = 0.15 . . . . . . . . . . 29
5.4 Sensitivity of cavitation at lower cavitation number with blade num-
ber variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.5 Optimum blade number suggestion by MAN Diesel & Turbo . . . . . 30

viii
5.6 Cavitation with BAR variation at σ = 0.25 (left) and σ = 0.15(right) 31
5.7 Sensitivity to cavitation at lower cavitation number with BAR variation 31
5.8 Cavitation with pitch variation at σ = 0.15 . . . . . . . . . . . . . . . 32
5.9 Sensitivity to cavitation at lower cavitation number with pitch variation 33
5.10 Cavitation with skew variation at σ = 0.15 . . . . . . . . . . . . . . . 34
5.11 Sensitivity to cavitation at lower cavitation number with skew variation 34
5.12 Cavitation with rake variation at σ = 0.25 (left) and σ = 0.15 (right) 35
5.13 Sensitivity to cavitation at lower cavitation number with rake variation 35

ix
List of Tables

2.1 Literature Review Summary . . . . . . . . . . . . . . . . . . . . . . . 15

3.1 Blade Section Thickness and Camber Distribution . . . . . . . . . . . 17


3.2 DTNSRDC Propeller 4381 . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3 Comparative Study Models . . . . . . . . . . . . . . . . . . . . . . . 19

x
Nomenclature

Symbols

AE Expanded blade area in2


AP Projected blade area in2
c Chord length m
Cp Pressure coefficient
D Propeller diameter m
f Camber m
fmax Section maximum camber m
g Gravitational acceleration m/s2
h Submersion of 0.8R of blade at 12 o’clock m
J Advance coefficient
KQ Torque coefficient
KT Thrust coefficient
n Rotational speed rps
p Pressure Pa
P Propeller Pitch m
p0 Free stream pressure Pa
patm Atmospheric pressure Pa
ps Static pressure at the shaft centreline Pa
pv Vapour pressure Pa
q Dynamic flow pressure Pa
r Radius of a propeller section m
R Propeller radius m
t Thickness m
T Thrust lb
ti Time
tm ax Section maximum thickness m
TP Thrust N
V Free stream velocity m/s
x Non-dimensional radius (r/R)
Section chord-wise position m
xs Section skew induced rake m
Z Blade number
∆p Propeller blade pressure difference Pa
β Section inflow angle degrees (◦ )
η Open water efficiency
θ Section nose-tail pitch angle degrees (◦ )

xi
θs Section skew angle degrees (◦ )
ρ Water density kg/m3
σ Cavitation number
σ0 Free stream based cavitation number
σm Mean cavitation number
σn Rotational speed based cavitation number
σ n0.8 Local cavitation number at 0.8R
φ Perturbation potential

Abbreviations
BAR Blade Area Ratio
BEM Boundary Element Method
CFD Computational Fluid Dynamics
CPP Controllable Pitch Propeller
HSVA Hamburg Ship Model Basin
MIT Massachusetts Institute of Technology
NACA National Advisory Council for Aeronautics
PPB Propeller Panel Based method
USS United States Ship

xii
Chapter 1

Introduction

1.1 Cavitation
Cavitation, also referred to as “cold-boiling”, is the phenomenon of the formation
of vapour pockets – cavities – within a fluid, caused by pressure reduction below
a certain value named “vapour pressure”. When the cavities find themselves back
in a higher pressure environment they implode, causing the liquid to rush towards
its centre to fill it, hence generating large pressures (up to 1 GPa). These are
subsequently propagated in the form of pressure waves to their surroundings.[1]

Figure 1.1: Water phase diagram

This phenomenon occurs in many fields such as chemical engineering, biomedi-


cal applications, cleaning technologies, the automotive industry and, of course, has
developed a dedicated research branch under fluid mechanics. In marine hydrody-
namics this phenomenon is especially present in pumps, turbines and propellers,
which often comprise of multiple hydrofoils. As pressure is a decreasing function of
velocity, according to Bernoulli, the flow around each propeller blade section expe-
riences rapid pressure changes, which lead to the inception of cavities around them.
This effect is amplified at high inflow speeds but is also dependent on the body
geometry.[2]

1
Its occurrence has detrimental impact on its surrounding environment. First
and foremost, the pressures exerted by the imploding cavities cause erosion of the
propeller blades, the encompassing hull plates and other nearby equipment such as
the rudder etc. Also the effects cause vibrations to be radiated towards the vessel
and transmitted structurally, which not only endanger the vessel and its machinery,
but also is uncomfortable for the crew and passengers on-board. On that note, the
collapsing vapour pockets produce noise which is not only undesirable to the people
on-board, but is also an environmental concern. Additionally, it is an important
subject of concern to engineers in the naval industry, where stealth is a primary
aspect of a successful product.[3]
Finally, cavitation causes loss of thrust in marine propellers due to the reduced
contact area between the blades and water and thus efficiency degradation is ob-
served. There are certain instances when the propeller is required to operate at high
rotational speeds and cavitation is unavoidable. In these cases, propellers can be
designed to take advantage of the flow characteristics, for example by reducing the
curvature of the leading edge of the blade and making it sharper. Propellers that
operate with fully developed cavitation are called “super-cavitating” propellers and
are employed exclusively for high speed scenarios, as their operation at lower speeds
is sub-optimal.[2]

1.2 Types of Cavitation


Propeller cavitation has been found to manifest itself in different locations of the pro-
peller and with certain unique forms. Thin string-like cavities at the tip of the blade
and from the hub of the propeller are called vortex cavities and are fully developed.
On the other hand, we also encounter sheet cavities, which are relatively stable and
present the least modelling difficulties in numerical methods. Sheet cavities start
from the leading edge of the blade whereas if they start from mid-chord they tend
to turn into bubble cavitation from there and aftward. Root cavitation may appear
just forward of the hub at certain rotation angles of the propeller and often joins the
hub vortex. Finally, cloud cavitation is visually similar to sheet cavitation, however
is very unstable and presents many modelling difficulties.[4]

1.3 Propeller Performance Evaluation Methods


Experimental tests have been the prime method for the performance evaluation of
propellers in both full and model scale settings. The former is achieved in ship trials
and involves a non-uniform inflow wake due to the hull shape, while the latter can be
done with or without the presence of a dummy hull model. Performance charts are
produced by undergoing open water propulsion tests, conducted ideally in a deep
water basin, to avoid wall interferences. The tests are carried out by attaching a
propeller on a moving carriage via a horizontal shaft, selecting a fixed rotational
speed, and then conducting a number of runs, each with different forward speed
until the desired range of advance coefficients J is achieved.
Cavitation tests on the other hand are conducted in cavitation tunnels which have
the ability to increase and reduce the pressure to recreate the needed environmental
setting. Standard procedures require the propeller to be mounted on a shaft, a

2
Figure 1.2: Various types of cavitation

constant inflow speed and gradually increasing the rotational speed up to the motor’s
capabilities, usually achieving a range of at least 0 to 50 percent slip for inception
tests. For studying the cavitation extent and behaviour on a propeller, a certain
advance coefficient is selected, normally according to a thrust identity between model
and full scale, while also retaining cavitation number similarity.[5][6]
These evaluation methods have been implemented with wide acceptance by re-
searchers and the industry, however they involve a certain number of disadvantages.
Firstly, satisfying flow similarity between model and full scale is impossible, and
thus various techniques are needed to transform tank test results. This demands
a certain amount of empiricism and also allows for technique variation between
different testing facilities. Additionally, the time requirements and costs involved
are not always feasible, which has led to a restricted amount of experimental data
to become available. A combination of the above, and the complex geometry of
propellers, which are tailor made on a case-by-case basis, make the process quite
restrictive and prone to errors.

1.4 Numericals Methods


With rapid improvement of computational power, numerical methods have been
a focal point for hydrodynamic analyses, including propeller testing. Both viscous
and inviscid methods are constantly being developed and improved, becoming useful
tools for initial design stages. Due to the high demands of viscous CFD in both
time and computer resources, simplifications of the fluid flow can be introduced
with acceptable accuracy. One such idealised scenario is potential flow solvers, i.e.
a flow which is irrotational, incompressible and inviscid. The time savings gained by
replacing the Navier-Stokes equations and continuity equation with a simple linear
equation, the Laplace equation, while still producing results close to experimental,
have made such methods popular amongst naval architects.
Boundary Element Methods (BEM) implement the potential flow assumption
and discretise the fluid domain over the boundaries rather than the whole fluid,

3
yielding surfaces as opposed to volumes, and the velocities are computed on each
individual panel. Depending on the nature of the grid generator and solver, this
vastly accelerates the computation while retaining great accuracy due to its semi-
analytical nature. They also have benefits over lifting-line methods with regards to
errors near the leading edge and root of the blades.[5] The main issues that arise from
these methods is the complex programming required and the flow at the propeller
tip is not effectively captured, which in the case of cavitation, does not allow the
research of tip vortex cavitation.[7]

1.5 Project Aims and Objectives


Propeller design optimisation for cavitation has been ongoing since the beginning
of the 20th century and research has been applied in practical cases. An example
of such work can be seen with the USS Pennsylvania submarine, which employed a
greater number of propeller blades amongst other modifications to reduce the radi-
ated noise.[8] There has been an accumulation of valuable information concerning
the effect of propeller geometry on cavitation, yet it is not clear how they have been
implemented in the design process as part of recommendations.
This project aimed to review existing conclusions derived particularly from ex-
perimental observation, but also from applications of numerical solving methods,
with regards to correlations between propeller geometry and cavitation. Addition-
ally, taking advantage of the availability of current computational resources, a study
was conducted using a BEM solver to investigate the accuracy and subsequently, its
agreement with current findings. The ultimate objectives were defined as follows:

• Review literature and reports concerning propeller geometry, the phenomenon


of cavitation and cases of how the former influences the latter.

• Tabulate the above interrelations for use in propeller design and also the con-
firmation study that took place.

• Undergo a validation study that evaluates the accuracy of the available nu-
merical method.

• Perform a systematic process that involves the generation of model propellers


of varying geometrical characteristics and their simulation in cavitation tests.

• Study the results and derive conclusions based on the above findings.

A summary of the review, methodology used to accomplish the above mentioned


objectives and the conclusive thoughts are described in the ensuing sections of this
report.

4
Chapter 2

Literature Review

2.1 Background
Cavitation in fluid flows is quantified using the cavitation number σ , which is defined
as the ratio of the static pressure head over the dynamic pressure head. The static
head is usually defined at the shaft centreline or either a radial position of 0.7 to
0.9R depending on the needs of the study. The dynamic head is based on either
a single or vector sum of the axial and rotational velocities of the propeller, once
again, possibly at an intermediate radial position of the blade. The following are
various forms in which it can be defined.[5]
p 0 − pv p0 − pv p0 − p v
σ0 = 1 σn = 1 σm = 1
2
ρV 2 2
ρ(πxnD)2 2
ρ(V 2+ (πxnD)2 )
This parameter is a means of measuring the difference between the inflow stream
pressure compared to the vapour pressure, the point at which vapour cavities will
start to form. When compared to the pressure coefficient obtained from the pressure
distribution on the propeller blades it allows us to identify possible locations at which
cavitation will occur due to significant loads and pressure imbalance. The pressure
coefficient at a certain point is defined as follows:
p − p0
Cp = −
q
When the resultant velocity of the blade increases, the pressure imbalance on
the face and back of the propeller blade increases, leading to the low pressure side to
possibly fall below vapour pressure. This pressure limit, although not an absolute
requirement for the inception of cavitation, is identified for a certain cavitation num-
ber and it is to be avoided during operation by manipulation the propeller geometry
amongst other methods, and therefore the change of the pressure distribution.
The geometrical parameters for the review are identified as:

• Diameter

• Blade number

• Blade area

• Pitch angle

5
• Skew

• Rake

• Warp

• Blade section

• Propeller hub

A summary of the findings is presented in Table 2.1.

Figure 2.1: Blade section with back cavitation

2.2 Diameter
A larger diameter is favoured by naval architects, as it means that the vessel can
achieve higher thrust, and so higher forward speed for the same rotational speed
of the shaft. Usually, the design limitation is the tip clearance from the bottom
of the hull and also high-speed planning vessels are an exception. For optimum
efficiency, the propeller diameter determination is a complex procedure involving
various empirical formulae.[9]
With regards to cavitation, a larger diameter is beneficial as the rotational speed
of the propeller could be reduced and still achieve the required forward movement.
A reduced rotational speed would mean that the pressure imbalance on the blades
would be reduced with the reduction of the inflow forces and so cavitation would be
decreased or alleviated. So for a certain engine output or desired forward speed, a
large diameter propeller would allow for slower rotation and reduced cavitation, as
seen in Figure 2.2.[10]

6
Figure 2.2: Optimum diameter selection graph

2.3 Blade Number


The primary effect that is to be avoided with propeller number selection is resonance,
as the number of blades affects the frequency of vibrations and the strength of
vibrations that occur during operation. There is also a strong interrelation between
diameter, blade area and blade number, with a greater diameter requiring fewer
blades and a greater area requiring more blades. An increase in propeller number can
reduce sheet cavitation of the suction side due to a reduced load per blade, however
it can increase root cavitation due to reduced clearance between each blade.[5]
It is recommended that an area ratio of 16-18% should be assigned to each blade,
with a decrease of 4% in the diameter per each additional blade. This guideline has
been obtained with both performance and cavitation efficiency in mind.[10]

2.4 Blade Area


There are several cases in which the blade area is subjected to constraints. For
example, in the case of controllable pitch propellers there must be enough space
between the blades to allow for reversibility. An increase in blade area can also lead
to an increase in drag, and hence a balance must be found between it and thrust
for optimum efficiency. For cavitation purposes, according to early studies it was
shown that cavitation on the back and suction side that caused thrust breakdown
could be reduced by increasing the blade area. There are two methods to determine
the blade area of the propeller, namely Burrill’s and Keller’s.[5]

7
Figure 2.3: Correlation between optimum diameter, blade number and blade area
ratio

Figure 2.4: Burrill’s cavitation diagram for uniform flow

8
Burrill’s method can provide recommendations for fixed pitch propellers using
Figure 2.4. For a certain “permissible” level of cavitation, the graph provides the
thrust loading coefficient at the 0.7R section of the blade. This estimate was deter-
mined from cavitation tunnel tests in uniform axial load.
The projected and expanded propeller areas would be calculated using the fol-
lowing two formulae respectively:
T AP
AP = 1 AE =
τ ρ(V 2
2 c
+ (0.7πnD)2 ) 1.067 − 0.229P
D
where τc is obtained from Figure 2.4.
Keller’s method is based on the expanded area ratio:

AE (1.3 + 0.3Z)TP
= +K
A0 (ps − pv )D2
The value K varies from 0.2 for single screw propellers, to 0.1 for slow twin screw
merchant ships, and 0 for fast twin screw naval ships.

2.5 Pitch Angle


The pitch angle of the propeller is a crucial parameter, determining the angle of
attack of the blades with the incoming flow and largely dictates the variation of
the pressure exerted on the blades. The determination of the pitch of a propeller
is made in conjunction with diameter and rotational speed to achieve the desired
forward speed. A reduced angle of attack (θ–β→0 in Figure 2.5) leads to a reduction
in back cavitation by creating a more even distribution of pressure, when combined
with appropriate camber to produce lift. On the other hand, increasing the angle of
attack, often with the combination of decreased camber, reduces cavitation on the
face of the blade and the risk of negative angles of attack occurring.[11]
The pitch of the propeller has been researched extensively by testing blade sec-
tions at various angles of attack. The results are presented in various forms, in-
cluding velocity distributions, bucket diagrams and pressure distributions, as seen

Figure 2.5: Blade section pitch P, nose-tail pitch angle θ and inflow angle β

9
Figure 2.6: Left – Velocity distributions at (a) positive, (b) ideal and (c) negative
angle of attack. Right – Cavitation bucket diagram

in Figures 2.6 and 2.7. It can be observed with these graphs how the angle of at-
tack can be modified to move cavitation inception from the back to the face of the
propeller. It has been deduced from studies that cavitation inception is limited to
lower numbers of the cavitation number for small angles of incidence, however as it
increases, the susceptibility spreads to the wider range.[5]
Bucket diagrams can show the type of cavitation occurring at a certain com-
bination of angle of attack and cavitation number while also expressing the non-
cavitating capabilities of the 2-dimensional section in the width of the “bucket”.
The fact that this refers to a section rather than a 3-dimensional blade leaves room
for uncertainty with regards to tip and root cavitation, and so other design methods
must be used to optimise those areas in particular. Nevertheless, they are often used
during the design process.[5]

Figure 2.7: Pressure distribution for various cavitation numbers at 5◦ angle of


incidence

10
Figure 2.8: Observation from experimental tests

2.6 Skew
Skew is considered to be the most effective geometrical parameter that can be ad-
justed in order to manage the extent and inception of cavitation, due to not having
a simultaneous effect on thrust.[12] The optimum amount to be applied has been
suggested to be in the range of 45-60◦ , where the greatest amount of blade unloading
occurs. Angles in excess of 60◦ have been shown to cause the increase of negative
pressure at the tip region, nevertheless moderate angles of skew have been shown to
be beneficial in normal operations.[13]
Skew weakens sheet cavitation and moves it from the blade tip to the leading
edge, however this can be aggravated at low cavitation numbers and turn into leading
edge vortices.[14] The resultant pressure fluctuation produced for skewed propellers
is 50-70% less than that commonly found.[15]

2.7 Warp
Warp is defined as the angular displacement equal to the projected skew angle mea-
sured after the skew induced rake is made up with opposite rake.[16] Warp has been
proven to have a very similar effect to skew, with a good comparison test showing
that warp might even be more effective for moderate cavitation numbers, while, only
for very low values of the cavitation parameter, skew might be superior.[12]

11
Figure 2.9: Unwrapped view of section showing the relationship between warp and
skew

Figure 2.10: Comparison between undistorted, 36◦ and 72◦ skew, and 36◦ and 72◦
warp

2.8 Rake
Evidence for the use of rake for cavitation reduction was not sufficient to confirm
its use for cavitation management purposes. The main purpose for which it is
introduced is for increasing the clearance between the blade tip and the hull bottom.
There is some evidence that slight backwards rake can reduce cavitation with small
reduction of propeller efficiency.[17] Since rake also allows for the use of a larger
diameter, it might implicitly lead to cavitation improvement.
On the other hand, it leads to an increase of the load experienced by the propeller
blades and therefore a greater section thickness is used, the effects of which will be
discussed in the next section. Although blade rake does not seem to be used for
cavitation mitigation purposes, blade tip rake has been found to be of great interest
and benefit.

12
2.9 Blade Section
The blade section geometry comprises of the section thickness, camber and chord
length. For better cavitation characteristics, NACA sections have become a popular
choice for newer ships with marked improvements.[18] The vast amount of experi-
mental data on NACA foils makes them an attractive option for engineers, who will
go on to modify them to fit their particular case. The flat pressure distribution on
the suction side provides a wide bucket diagram, on the pressure side however the
phenomenon is not so well controlled. To solve the problem of high loading at the
leading edge of the blade, which is the source of severe cavitation, the maximum
thickness is suggested to be positioned forward of 50% of the chord length with an
S-shaped camber.
Thicker foils also allow for the use of a wider range of angles of attack, however
the increased drag and reduced lift that result may not be welcomed. An increase
in camber can be regarded as an increase of the initial angle of attack, which would
eventually cause cavitation, but with less impact on the drag of the foil. Therefore,
cavitation-free operation can be carried out at larger positive angle and smaller
negative angle with a small increase in the width of the range.[19] The thickness is
also limited by the cord length, as a thickness-to-chord ratio above 0.35 can lead to
cavitation.
In the radial direction, although thicker blades are more prone to cavitation,
combining it with greater camber can help avoid the low pressure peak at the lead-
ing edge and hence reduce the possibility of sheet cavitation. On the contrary, a
thin blade with small camber is better for avoiding bubble cavitation.[20] Overall,
the blade section requires a significant amount of attention and an iterative and sys-
tematic process to accomplish a satisfactory design. More detailed characteristics,
such as the blade leading edge radius and shape are also important, but their study
could not be accommodated in this project.

Figure 2.11: Novel camber distribution compared to the NACA08 camberline

13
Figure 2.12: Right – The influence of thickness on cavitation inception. Left – The
influence of camber on cavitation inception

2.10 Propeller Hub


Design processes require the hub to be as small as possible, for various hydrodynamic
reasons including cavitation, while still being able to serve its purpose, especially
for controllable pitch propellers (CPP) which have to house the mechanism. Also, it
must not be too small, as the spacing of the blades would be significantly reduced,
increasing the possibility of root cavitation. The hub ratio is defined as the diameter
of the hub over the diameter of the propeller equivalent actuator disk. For control-
lable pitch propellers this ranges from 0.24-0.32D, while for fixed pitch propellers
from 0.16-0.25D.[5]
The form of the hub is interestingly important and two designs, with either a
convergent or a divergent design being used. Convergent hubs are used for slow
merchant ships and see face and back root cavitation more easily. Slightly diver-
gent hubs are used for high-speed ship, patrol vessels and warships, while strongly
divergent hubs are less popular. Finally, parallel or divergent hub cones are often
employed in an attempt to reduce the root vortices generated.[5]

Figure 2.13: Hub shape and its effect on cavitation

14
Table 2.1: Literature Review Summary

Parameter Recommendation Primary Limitation


Diameter Increase (usually with rotational Blade tip clearance
speed reduction)
Blade Number Increase Blade spacing at the hub
Blade Area Increase Lift-drag balance
Pitch Angle Low to moderate angles are Highly dependent on blade
recommended section
Skew Moderate angles between 45-60◦ Negative effects after 75
Warp Moderate angles between 45-60◦
Rake Unclear Strength requirements
Thickness Increase Optimisation of pressure
distribution for lift and drag
Camber Increase – Negligible effect on
range
Chord Length Increase Rotational ability of blades in
CPPs
Hub Decrease Fit mechanism for CPPs

15
Chapter 3

Propeller Modelling

The modelling of the propellers was completed with the assistance of a parametric
modeller, with the capability of reproducing various series of widely known and
used propellers, such as the Wageningen B-series, Gawn, Kaplan and others. It also
allows for custom rules to be applied on the blades, hub and individual cross section,
making it a very useful tool for custom design.

3.1 Validation Model


Due to a restriction with regards to the flow simulation software, blades had to be
designed with cross-sections according to a NACA66-MOD thickness distribution
and NACA08 camber line. They are regarded as some of the most commonly used
forms and are designed for cavitation reduction, however they usually serve as a
starting point.[5] The distributions can be seen in Table 3.1.
First, the model of a propeller with such blade sections, i.e. not modified further,
had to be identified in literature, for which experimental data was also available.
This would allow for the validation of PPB. One such propeller was used by Boswell
in both open water experiments and cavitation tests as part of a study on the effect
of skew on cavitation.[21] Propeller 4381 has simple geometry in the form of sections
at various radial positions and thus the input parameters for the simulation were
manually reproduced using Excel.[5][21]

3.2 Comparative Study Models


For this project a total of 40 propellers were produced, with one being the “parent”
propeller and the rest varying with respect to a certain characteristic for a range
of magnitudes. The initial aspiration was to be able to take advantage of the Wa-
geningen B-series wide range of experimental results, however due to the constraint
imposed by PPB, a modification was made to incorporate the NACA sections. All
propellers were defined according to a full pitch distribution, with the exception of a
second 4-blade propeller which followed a non-linear pitch distribution due to a 20%
pitch reduction at the hub. This is standard for B-series propellers, to assist with
adapting to the velocity distribution behind a hull.[6] All propellers had a hub-to-
diameter ration of 0.167, with the exception of the 3-blade propeller, which required
a ratio of 0.18.[22] The outline of the blades, thickness and skew distribution were

16
Table 3.1: Blade Section Thickness and Camber Distribution

NACA08 NACA66-MOD Thickness


Camber-line Distribution
x/c f /fmax t/tmax
0 0 0
0.0025 0.02 0.085
0.005 0.035 0.14
0.0075 0.05 0.18
0.0125 0.071 0.2248
0.025 0.1263 0.3094
0.0499 0.2236 0.4271
0.0749 0.3147 0.5122
0.0999 0.3975 0.582
0.1498 0.54 0.693
0.1998 0.6576 0.7804
0.3002 0.8338 0.9046
0.4001 0.9453 0.9763
0.5 1 1
0.5999 0.9967 0.9724
0.6998 0.9318 0.8787
0.8 0.7769 0.7016
0.9 0.4361 0.4228
0.95 0.2181 0.2344
1 0 0

Table 3.2: DTNSRDC Propeller 4381

Number of blades: 5
Hub diameter ratio: 0.2
Expanded area ratio: 0.725
Section mean line: NACA08
Section thickness distribution: NACA66-MOD
Design advance coefficient: 0.889
r/R c/D P/D θs xs /D tmax /D fmax /c
0.2 0.174 1.332 0 0 0.0434 0.0351
0.25 0.202 1.338 0 0 0.0396 0.0369
0.3 0.229 1.345 0 0 0.0358 0.0368
0.4 0.275 1.358 0 0 0.0294 0.0348
0.5 0.312 1.336 0 0 0.0240 0.0307
0.6 0.337 1.280 0 0 0.0191 0.0245
0.7 0.347 1.210 0 0 0.0146 0.0191
0.8 0.334 1.137 0 0 0.0105 0.0148
0.9 0.280 1.066 0 0 0.0067 0.0123
0.95 0.210 1.031 0 0 0.0048 0.0128
1.0 0 0.995 0 0 0.0029 -

17
Figure 3.1: Solid and wire representation of the parent model

according to B-series polynomials, while the rake distribution was linear. Finally,
the maximum camber was set at 0.01 of the section chord length.
The parent propeller consisted of 5 blades, a blade area ratio of 0.75, a 7m
diameter and pitch, 15◦ rake and the skew was determined by the modeller at 16.3◦ .
The first group of propellers consisted of 11 models that varied in diameter, starting
from 4 meters up to 9 meters in increments of 0.5 meters. The second group consisted
of 6 models, ranging from 3 to 7 blades, including the second 4-bladed model with the
80% pitch. Next, were 6 propellers with BAR ranging from 0.3 to 1.05 in increment
of 0.15. Group 4 varied in pitch from 4.2 to 9.8 meters in 1.4 meter steps and an
additional model with 3.5 meter pitch. The group of skew-varying propellers started
at 0◦ , then 16.3◦ and from 25◦ to 75◦ in 10◦ steps. Finally, rake began from -5◦ up
to 30◦ by 5◦ increments.

18
Table 3.3: Comparative Study Models

Propeller Difference Propeller Difference


Number from Number from
Propeller Propeller
No.1 No.1
Diameter (m)
2 7.5 7 6
3 8 8 5.5
4 8.5 9 5
5 9 10 4.5
6 6.5 11 4
Blade Number
12 6 15 4 (80% pitch)
13 7 16 3
14 4
Blade Area Ratio
17 0.9 20 0.45
18 1.05 21 0.3
19 0.6
Pitch (m)
22 8.4 25 4.2
23 9.8 26 3.5
24 5.6
Skew (degrees)
27 25 31 65
28 35 32 75
29 45 33 0
30 55
Rake (degrees)
34 20 38 5
35 25 39 0
36 30 40 -5
37 10

19
Figure 3.2: CAD drawing of the parent propeller

20
Chapter 4

Propeller Simulations

4.1 Numerical Method


The program used for this project was a limited version of the PPB code created and
distributed by the HSVA. It utilises a surface panel method, meaning it discretises
only the surface of the boundary of the body into panels, and with this particular
pre-processor, quadrilateral or, if need be, triangular panels are generated. The
trailing edge of the blades is forced to have triangular panels, as this has been
shown to be beneficial for the Kutta condition. One of the limitations imposed was
that the hub was not modelled to avoid issues brought up by the vorticity generated
at the junction of the blade and hub. Hence, all results are due to the blades
only. Additionally, as mentioned before, the blade sections were all forced to have
NACA66-MOD thickness distribution and NACA08 camber-lines.
To generate the mesh a geometry control file is needed, which includes the num-
ber of blade sections, their non-dimensional radial position, chord, maximum thick-
ness and camber, rake, pitch-over-diameter ratio and the distance of the leading
edge of the section to the generator line. These were obtained through spreadsheet
formulations of the modeller output, and, together with the diameter and blade
number specification, composed the whole pre-processor control file. The grid gen-
erator produces the mesh of one blade and then the solver uses a key-blade approach,
i.e. takes advantage of the rotational symmetry of the propeller and by solving for
one blade produces the solution for the whole propeller.[23] Each blade consists of
21 chord-wise and 11 radial ordinates, yielding a total of 200 panels on each side.
Modification of the grid density was not possible with the limited version of PPB.
PPB is also potential based, meaning it incorporates the assumptions of irrota-
tionality, incompressibility and absence of viscosity. The basic unknown which is
to be solved at the centre of each panel is the local perturbation potential φ(x,ti ),
which is solved in the frequency domain and satisfies the Laplace equation ∇2 φ= 0,
replacing the continuity and momentum equations to yield a faster calculation.[23]
The computation is derived directly for steady state, meaning the solver takes
into account only the zeroth harmonic order and saves computational time. The
derivation of the surface pressures is achieved through two steps: first the flow
potential is deduced for the whole blade and then surface strengths are evaluated
for each panel to give the pressures. Viscous corrections are also implemented on the
basis of the whole panel domain by adding up the shear force contribution of each
panel, which is estimated for every element from friction lines.[23] More detailed

21
Figure 4.1: Grid and mesh of the DTNSRDC Propeller 4381

modelling was not disclosed, however the method is loosely based on the panel
method approach used by Kinnas et al. and Lee et al.
The input file for the solver requires the following parameters: the advance
coefficient, the propeller rotational speed, the diameter, a wake parameter (set to
zero throughout the project as it relevant only to simulations of in-behind condition)
and finally the local cavitation number at 0.8R of the blade at the 12 o’clock position,
and is defined as follows:
patm + ρgh − pv
σn0.8 = 1
2
ρ(πnD)2
The file can be configured to run consecutive simulations, thus providing a table
of results. The output file provides the pressure coefficient on the suction and
pressure side for each radial ordinate at a position of 1/7th of the total chord. It
also provides a non-dimensionalised circulation and cavity volume, estimated from
the mean surface pressure, both for all 10 radially distributed panels. For this code,
only sheet cavitation is predicted.[24]

4.2 Open Water Validation


First, the pre-processor input was evaluated to produce the mesh of the blade. For
the simulation, the advance coefficient was increased in increments of 0.1 from 0
to 1.3. To derive the open water results with as little discrepancy as possible, the
propeller was set to a constant rotating speed of 7.8 rps and the inflow speed varied,
to achieve the desired J, from 3 to 10 feet per second, following the same process
as the experiment itself.[21] The report does not mention the submergence of the
propeller and so it is assumed that the shaft centre of the propeller is one diameter
deep under the waterline. With D,n,σ n0.8 and J in hand the open water results were
calculated, tabulated and compared to the experimental data, as shown in Figure
4.2.

22
Figure 4.2: Open water performance comparison between PPB and experimental
results by Boswell

Overall the open water results do follow the same trend as the experimental
results, however the discrepancy is on the higher end of the spectrum for engineering
applications; at the design point of J= 0.889 the percentage difference between
results is 8.8% for both thrust and torque. It is not specified in the experimental
test procedures if a run of the tank carriage was made in order to determine the
system friction and dummy hub torque and thrust. Therefore, it is unknown if
the results are solely due to blade action or not, and potential flow solvers tend to
predict a slightly higher thrust at the design point when modelled without a hub.[25]
In general it has been shown that below the design point BEMs are often unable
to properly predict torque when compared to other numerical methods.[26][27] The
efficiency is under-predicted by less than 1.5%.
This propeller was designed by a lifting surface procedure with additional correc-
tions based on thrust performance criteria. The overall performance of the models
tested was lower than the design condition, with 2.3% thrust, 0.4% torque and -1.8%
efficiency at J= 0.889. When these results are compared with the numerical results
the error of the method is reduced to 6.3% for thrust and 8.3% for torque, whilst
the error in efficiency increases to 3.2%.

23
4.3 Cavitation Validation
The study conducted by Boswell included cavitation results and thus the opportunity
to validate the performance of PPB in that aspect was utilised. The test procedure
was completed in a variable pressure water tunnel and so the cavitation number,
selected at the shaft centre-line, could be controlled independently. First, a value
for the advance coefficient was selected and then values for the inflow and rotational
speed were adjusted, always remaining between 10 and 20 feet per second for the
former, and 14 and 20 rps for the latter. Initially the propeller would operate in a
non-cavitating condition and a reduction of pressure was introduced until cavitation
would appear or change.[21]
To convert the free stream cavitation number to the local value required by
PPB, a systematic procedure was required for each advance coefficient. First the
advance coefficient was selected together with the highest possible inflow speed, not
exceeding the previously mentioned range. The rotational speed was now adjusted
by definition J=V /(nD). The static head needed to be converted from that at the
propeller shaft, to one at 0.8R of the blade at 12 o’clock, to achieve the correct local
number. If the two definitions can be expressed as:

patm + ρgh − pv patm + ρg(h − 0.8 D2 ) − pv


σ0 = 1 and σn0.8 = 1
2
ρV 2 2
ρ(πnD)2

Then, by combining the two:


1 2 1
ρV σ 0 = ρ(πnD)2 σn0.8 + 0.4ρgD ⇒
2 2

1
2
ρV 2 σ 0 − 0.4ρgD
⇒ σn0.8 = 1
2
ρ(πnD)2
From here the cavitation number could be derived for the range of 1 to 12 of
σ 0 , and the whole process repeated to cover the range of 0.5 to 1 for the advance
coefficient. The results for cavitation inception are only an estimate as it would
require an extensive procedure to identify the exact value of J of its initiation. It
has been noticed that PPB detects the sheet at the very top ordinate and, as it
becomes larger, it spreads to lower parts of the blade. The two lines in Figure 20
represent the tip and 0.4R ordinate of the blade.
The results are not encouraging as they indicate a severe under-prediction of
cavitation at higher cavitation numbers. The cavity volume estimation is based on
the blade pressures, nevertheless there seems to be significant difference. However,
due to uncertainty with regards to the processes followed in the model experiments
it is not possible to draw a definite conclusion. For the comparison study lower
cavitation numbers will be used in order to avoid any potential issues with the
prediction.

4.4 Comparative Study Setup


The comparison of propeller cavitation sensitivity is done on the basis of the volume
of the cavity. The comparison of two propellers is done for same advance coeffi-

24
Figure 4.3: Total cavitation inception as predicted by PPB compared to the
experimental prediction

cient and local cavitation number. There is substantial discussion regarding the
ideal definition of the cavitation number for performance comparison in model test-
ing.[28][29][30] There is no definite consensus on the “correct” method, however it
is suggested that for model tests the local cavitation number is used, especially for
inception studies.[5] This is not relevant in this case, as all propellers are in full-scale
and no scaling of the model or the effect needs to be done.
Nevertheless, the strategy followed was to test all 40 propellers at 3 different local
cavitation numbers: 0.15, 0.2 and 0.25. First, the parent propeller was used, along
with an inflow speed of 10 m/s to determine the free stream cavitation number when
the shaft was submerged by one diameter. The result was kept throughout all the
tests and was approximately 3.3. For every new propeller the new static pressure was
determined always at a shaft submergence of one diameter and, using the free-stream
cavitation number the inflow velocity was determined for the particular diameter.
Every propeller was tested at advance coefficients ranging from 0 to 1.5, and for
each one a unique rotational speed was derived by definition of J=V /(nD). Finally,
the desired local cavitation number would be selected and this would eventually
re-adjust the static pressure, changing the free-stream cavitation number for each
propeller.
The reason for keeping the local cavitation constant is mainly due to the fact
that it is part of the input of PPB, and so greater consistency is achieved for the
comparison. The initial derivation of the inflow speed is justified as a “prediction”
of what forward speed the propeller would have to achieve the free-stream cavitation
number. The correction of the inflow velocity is an attempt to adjust for the pres-
sure change due to the modified submergence that would occur if the tests were not
performed in a pressure controlled environment, such as normal ship operation. Fi-

25
∆p ∼
∆p = patm + ρgD − pv σv = 1 = 3.3
2
ρV 2
where D = 7m
where V = 10 m/s

Propeller Geometry i.e. D ∆p = patm + ρgD − pv

s
∆p
V = 1
J = 0, 0.1, 0.2 . . . 1.5 2
ρσv

where σv ∼
= 3.3

V V
J= ⇒n=
nD JD

Figure 4.4: Flowchart demonstrating the derivation of the PPB input variables

nally, since the ability to de-/pressurise the environment exists, it seems appropriate
to take advantage of it to keep other variables constant.
Overall this process is not necessary, as one could just keep either the local
or free-stream cavitation number constant and not take advantage of the pressure
adjustability. This would set either the inflow or rotational speed, depending on the
definition of the cavitation number, and for the range of advance coefficients, the
other speed would be determined. Yet the above process seemed like a justifiable
extra step in the process.

26
Chapter 5

Results

For all geometrical characteristics, two types of figures are presented. First, the
value of the volume of the cavity is shown as both the geometry and the advance
coefficient are varied through the prescribed range. The second type is an attempt
to investigate how the sensitivity to cavitation changes with the geometry, as the
cavitation number decreases. To achieve this, the increase in volume is measured
between consecutive cavitation number reductions. Then the difference between the
two increases is derived and plotted against the geometrical change and J.

5.1 Diameter
Larger diameter propellers generate more cavitation at a certain advance coeffi-
cient, with cavitation increasing almost linearly as J reduces. At the same J, the
growth of the volume is quadratic with the diameter and this is amplified at low
values of advance. This is not surprising, as the expanded blade area also increases
quadratically with diameter, leading to a greater surface area. However, the fact
the this effect is amplified at lower values of J seems to suggest that the optimum
rotational speed reduces as a propeller becomes larger, which is in agreement with
industry practice.[9] Therefore, an engineer must identify this reduction of optimum
rotational speed and compare it to the gains in forward speed and thrust. The
cavitation increase per unit diameter could be linear, however the slope of this line
depends on the above residual parameters.
Figure 23 also confirms the previous statement for optimum rotational speed, as
the local cavitation number reduces quadratically with n. For larger propellers, cavi-
tation increases drastically at low J, whereas smaller ones seem to reach a stagnation
point. The reduction of increase at low J - high D suggests a similar stagnation in
further growth. An important note, however, is that cavitation is conceived almost
at the same advancement coefficient for all diameters, therefore this is a matter of
tolerance of total cavity volume.

5.2 Blade Number


The increase of propeller blades presents a rapid reduction to the total volume, while
all other geometrical characteristics remain constant. This could confirm the advice
found in literature regarding the blade area ratio; the BAR of 0.75 is 15% per blade

27
Figure 5.1: Cavitation with diameter variation at σ = 0.15

Figure 5.2: Sensitivity of cavitation at lower cavitation number with diameter


variation

28
Figure 5.3: Cavitation with blade number variation at σ = 0.15

for the 5-blade propeller and 25% for the 3-blade, which greatly exceeds the recom-
mendation of 16-18%. The 80% pitch reduction for the 4-blade propeller suggested
for B-series models does seem to have a small favourable impact, but cannot negate
significant errors in design decisions. The further reduction in cavitation upwards
of 5 blades may not come along with other performance benefits or could be due to
the inability to detect root cavitation due to close blade proximity.
Although the difference in increase of cavitation when lowering the cavitation
number seems to be somewhat slightly affected by the number of blades, the 3-
bladed propeller exhibits an unanticipated vulnerability. In industry it is suggested
that lower blade number propellers have a higher efficiency but for strength reasons
cannot be utilised when high thrust is a requirement. However, the above advantage
is hindered by the diameter, which seems to fall below the advisable range for 3-blade
designs.[9]

5.3 Blade Area Ratio


The blade area affects cavitation in an interesting manner. At higher cavitation
numbers, a large blade area is beneficial, with an onset occurring at lower values of
J and reduction of the volume compared to the parent propeller. Benefit can also be
seen at low blade area ratios, but onset is not postponed as before. Comparatively,
intermediate values are unexpectedly not beneficial. As the cavitation number re-
duces, the benefit of using higher blade areas is diminished for lower J, even though
the delay of onset is still present to an extent.
Additionally, low blade area ratio propellers do not experience increases in growth

29
Figure 5.4: Sensitivity of cavitation at lower cavitation number with blade number
variation

Figure 5.5: Optimum blade number suggestion by MAN Diesel & Turbo

30
Figure 5.6: Cavitation with BAR variation at σ = 0.25 (left) and σ = 0.15(right)

Figure 5.7: Sensitivity to cavitation at lower cavitation number with BAR


variation

31
Figure 5.8: Cavitation with pitch variation at σ = 0.15

of cavitation at high values of J as the number reduces. On the contrary, as also


deduced from the previous figure, high areas show a significant disadvantage, which
contradicts literature.[5] At more reasonable operational values, however, this effect
is opposite. Since the rotational speeds at these values of the advance coefficient are
more possible for an actual operating condition, it would be sensible to agree with
current design procedures regarding BAR, namely Burrills method.

5.4 Pitch
Pitch demonstrates a linear response to both growth of the total volume. It also
appears earlier for propellers with higher pitch in a linear fashion, which seems like
a reasonable outcome since a linear pitch distribution was used for all propellers and
the variation of pitch effectively alters the angle of attack of the blade to flow. This
results in the effective pressure on the blade to increase with the angle of attack.
In an attempt to arrive at a useful conclusion, Figure 30 shows that as the
cavitation number reduces, at the speed of inception, all propellers experience the
same increase. Therefore, pitch, together with diameter, must be adjusted to fit the
rate of revolution that is to be selected. For lower rates of revolution, pitch-over-
diameter should be larger, and so should the diameter. Therefore, the losses due to
an increased pitch must be counteracted by the benefits gained from the increase in
diameter. This is in agreement with research findings and industry practice.[5][9]

32
Figure 5.9: Sensitivity to cavitation at lower cavitation number with pitch
variation

5.5 Skew
The use of skew for preventing cavitation is confirmed by the comparative study.
The effect is not of great significance, however given that skew does not have a
consequence on thrust, it is a useful tool for this purpose. A delayed onset rotational
speed and a reduced rate of growth is observed. In a numerical study [13] it was
found that optimum skew is around 45-60◦ , however the results from PPB indicate
that the benefits continue at higher values, which agrees with studies that used both
numerical and experimental methods.[12][14]
An increase in skew also reduces the rate of growth of cavitation as the cavitation
number reduces, meaning that it can benefit high rotating speed propellers. One
important note is that, although skew has beneficial effects, the manufacturing of
the blades may require other geometrical changes, such as an increase in thickness,
which could negate the positive gains.

5.6 Rake
The study reveals that cavitation is almost totally independent of rake. The effect
is negligible, however it is interesting to note that it is positive for large angles at
higher cavitation numbers, whereas it is negative for larger angles at lower cavitation
numbers. Thus, it seems unlikely that rake would be used with this goal in mind,
considering the additional measures that are required to cater for the strength of the
blades. This also confirms the implementation of warp in marine propellers, where,
when skew is implemented, skew induced rake is removed by applying negative rake.

33
Figure 5.10: Cavitation with skew variation at σ = 0.15

Figure 5.11: Sensitivity to cavitation at lower cavitation number with skew


variation

34
Figure 5.12: Cavitation with rake variation at σ = 0.25 (left) and σ = 0.15 (right)

Figure 5.13: Sensitivity to cavitation at lower cavitation number with rake


variation

35
Chapter 6

Conclusions and Discussion

The validation of PPB produced mixed results that were not re-assuring in regards
to accuracy for its purpose. Open water results were reasonably accurate, with a
deviation of 8.8% compared to experimental methods. Over-estimation of torque
and thrust is a normal behaviour of potential flow solvers; however, in this case
results could be improved significantly by increasing the mesh density.[23]
Cavitation prediction was underwhelming in the validation process, indicating
that the cavity volume estimation from the blade pressures might not be sufficient.
A late onset of cavitation was found; still there are various differences between the
experimental and numerical tests that might occur and lead to such deviation. First,
the exact values of inflow and rotational velocity were not provided in the experi-
ment description by Boswell, and therefore the precise values of the local cavitation
number, required as input for PPB, could be affected slightly. Additionally, there
are environmental parameters, such a water density, water air content and tank wall
effects, that are a significant factor for cavitation inception.[5][28][29][30] Finally,
there was no mention of a dummy hub test, which means the effect of the hub on
the results could be a reason for the discrepancy.
Overall, the results from the comparative study are in agreement with the knowl-
edge applied to marine propeller design. Although the diameter produces higher
volumes of cavitation, it increases linearly per meter of diameter, and therefore the
designer must take into account operational speed and thrust requirements. A re-
duction of the number of blades causes a large spike, as the area per blade becomes
excessive and the width of the blade exceeds its root-to-tip height. For a 5-blade pro-
peller, the increase of BAR is welcomed, as it delays the onset of cavitation, however
due to the conclusions from blade number variation, one must be cautious with the
reaction with different geometry. In general, the above three geometrical properties
are interconnected, yielding an optimum design based on requirements.[10]
Although a larger pitch is usually beneficial for performance, the onset of cavi-
tation moves to lower rotational speeds in a linear fashion with pitch increase. Skew
was confirmed to be successful in reducing cavitation throughout the range of 0◦ to
75◦ . This does not agree with certain numerical studies, which show that pressure
loading at the tip increases further from 60◦ [13], and this could be due to the in-
ability of BEMs to identify tip vortex cavitation. Finally, rake was not found to
be useful in alleviating cavitation, especially when the negative consequences of its
implementation could harm a propellers service life.
Throughout the study there were a number of decisions and techniques involved

36
that could be a source of eventual faults in the outcome and conclusions. Potential
flow introduces various simplifications to a real flow, and hence may not necessarily
lead to representative results. The mesh generated on the blades, although refined
at the leading and trailing edge may not be dense enough to capture the flow in
detail. The above, combined with viscous corrections applied to the results may
even lead to further inaccuracies.
An additional source of potential error can be found in the validation with the
absence of a more detailed description of the experimental procedures. Combined
with an uncertainty for the accuracy of the cavity volume estimation through the
surface pressure prediction, a deviation of the results for cavitation is not entirely
unlikely. Finally, the absence of the propeller hub, although a common technique in
such studies, is still another source of discrepancies between the real and simulated
scenario.
This study could be extended to provide similar insight on the effects of the
blade section on cavitation inception and extent. It is possible that camber could
be particularly useful for controlling cavitation, utilised appropriately with pitch. A
collection of the all the above could be coded into a program similar to “OpenProp”,
which started at MIT and since moved to Dartmouth College. It uses a generalised
lifting line method for model design and analysis, while it can also perform basic
stress calculations on the blades and output geometries for further work.[31] It would
form as a useful optimisation tool for cases where cavitation is a significant issue,
and would give suggestions on which geometrical parameter one should change, to
control sheet cavitation in the most efficient way.

37
References

[1] B. S. Massey and A. J. Ward-Smith, Mechanics of Fluids. London; New York:


Taylor & Francis, 2006.
[2] J. N. Newman, Marine Hydrodynamics. MIT Press, 1977.
[3] P. Eisenberg, H. S. Preiser, and A. Thiruvengadam, On the Mechanism of
Cavitation Damage and Methods of Protection. Hydronautics, Incorporated,
1965.
[4] S. Kinnas. (1996). Photographs of Different Types of Cavitation, [Online].
Available: http://cavity.ce.utexas.edu/kinnas/cavphotos.html.
[5] J. Carlton, Marine Propellers and Propulsion. Oxford: Butterworth-
Heinemann, 2012.
[6] W. P. A. van Lammeren, J. D. van Manen, and M. W. C. Oosterveld, “The
Wageningen B-Screw Series”, Society of Naval Architects and Marine Engi-
neers Transactions, vol. 77, no. 8, 1969.
[7] B. Volker, Practical Ship Hydrodynamics. Butterworth-Heinemann, Oxford,
2000.
[8] P. F. Johnston. (2000). The Taming of the Screw, [Online]. Available: http:
//americanhistory.si.edu/subs/anglesdangles/taming.html.
[9] MAN Diesel & Turbo, “Basic Principals of Ship Propulsion”, MAN Diesel &
Turbo, Copenhagen, Denmark, 2011.
[10] AB Volvo Penta, “Propellers - Inboard Propellers and Speed Calculation”, AB
Volvo Penta, Göteborg, Sweden, 1998.
[11] D. Bertetta, S. Brizzolara, S. Gaggero, M. Viviani, and L. Savio, “CPP Pro-
peller Cavitation and Noise Optimization at Different Pitches with Panel Code
andVvalidation by Cavitation Tunnel Measurements”, Ocean Engineering, vol.
53, pp. 177–195, 2012.
[12] J. E. Kerwin, S. D. Lewis, and S. Kobayashi, “Systematic Experiments to
Determine the Influence of Skew and Rake on Hull Vibratory Excitation Due to
Transient Cavitation”, presented at the Ship Vibration Symposium, SNAME,
1978.
[13] M. A. Mosaad, M. Mosleh, H. El-Kilani, and W. Yehia, “Propeller Design
for Minimum Induced Vibrations”, presented at the Port Said Engineering
Research Journal, Port Said University, 2011.
[14] Z. F. Zhu, “Numerical Study of the Effect of Propellers Skew on Cavitation
Performance”, Advanced Materials Research, vol. 705, pp. 405–409, 2013.

38
[15] B. Ji, X. Luo, and Y. Wu, “Unsteady Cavitation Characteristics and Al-
leviation of Pressure Fluctuations Around Marine Propellers with Different
Skew Angles”, Journal of Mechanical Science and Technology, vol. 28, no. 4,
pp. 1339–1348, 2014.
[16] D. R. Smith and J. E. Slater, “The Geometry of Marine Propellers”, Defence
Research Establishment Atlantic, Dartmouth, Nova Scotia, 1988.
[17] Y. Ukon, K. Kume, J. Fujisawa, H. Takeshi, Y. Kawanami, J. Hasegawa, R.
Fukasawa, S. Yamasaki, J. Ando, and T. Kanai, “Research on Improvement of
Propulsive Performance of a High-Speed Ship Equipped with a High-Powered
Propeller”, National Maritime Research Institute, Tokyo, Japan, 2008.
[18] J. Dang, “Improving Cavitation Performance with New Blade Sections for
Marine Propellers”, International Shipbuilding Progress, vol. 51, no. 4, pp. 353–
376, 2004.
[19] S.-W. Chau, K.-L. Hsu, J.-S. Kouh, and Y.-J. Chen, “Investigation of Cavita-
tion Inception Characteristics of Hydrofoil Sections via a Viscous Approach”,
Journal of Marine Science and Technology, vol. 8, no. 4, pp. 147–158, 2004.
[20] G. Kuiper, “Cavitation Inception on Ship Propeller Models”, TU Delft, Delft
University of Technology, 1981.
[21] R. J. Boswell, “Design, Cavitation Performance, and Open-Water Performance
of a Series of Research Skewed Propellers”, Department of the Navy Naval Ship
Research and Development Center, Washington, D.C., 1971.
[22] R. Biven, “Interactive Optimization Programs for Initial Propeller Design”,
University of New Orleans, New Orleans, 2009.
[23] S. Gaggero, J. Gonzalez-Adalid, and M. Perez Sobrino, “Design of Contracted
and Tip Loaded Propellers by Using Boundary Element Methods and Opti-
mization Algorithms”, Applied Ocean Research, vol. 55, pp. 102–129, 2016.
[24] H. Streckwall, “User Manual for HSVA’s Propeller Panel Method ”ppbfdom””,
Hamburgische Schiffbau-Versuchsanstalt GmbH, Hamburg, Germany, 2001.
[25] S. Brizzolara, S. Gaggero, and D. Grassi, “Hub Effect in Propeller Design
and Analysis”, presented at the Third Internationals Symposium on Marine
Propulsors smp’13, 2013.
[26] L. Greco, R. Muscari, C. Testa, and A. Di Mascio, “Marine Propellers Perfor-
mance and Flow-Field Prediction by a Free-Wake Panel Method”, Journal of
Hydrodynamics, vol. 26, no. 5, pp. 780–795, 2014.
[27] S. Gatchell, D. Hefermann, and H. Streckwall, “Open Water Test Propeller
Performance and Cavitation Behaviour Using PPB and FreSCo+”, presented
at the Second International Symposium on Marine Propulsors smp’11, Ham-
burg, Germany, 2011.
[28] “Comparative Tests of Propellers in Cavitation Tunnels”, presented at the
Fifth International Conference of Ship Tank Superintendents, London: The
International Towing Tank Conference, 1948.
[29] “Comparative Cavitation Tests”, presented at the Sixth International Confer-
ence Ship Tank Superintendents, Washington: The International Towing Tank
Conference, 1951.

39
[30] “Comparative Cavitation Tests of Propellers”, presented at the 8th Interna-
tional Towing Tank Conference, Madrid: The International Towing Tank Con-
ference, 1957.
[31] B. Epps and R. Kimball. (2013). OpenProp v3: Open-Source Software for
the Design and Analysis of Marine Propellers and Horizontal-Axis Turbines,
[Online]. Available: http://engineering.dartmouth.edu/epps/openprop.

40

View publication stats

You might also like