Sunteți pe pagina 1din 88

Short Introduction to Ship

Resistance and Propulsion

Jerzy Matusiak

Copyright 2008 by Jerzy Matusiak


2/88

1. INTRODUCTION .................................................................................. 4

4. SHIP RESISTANCE .............................................................................. 5


4.1. General 5
4.2. Definition of ship resistance 5
4.3. The effect of propulsor on hull flow and resistance 7
4.4. Similitude and order of magnitude 8
Non-dimensional form of resistance ...............................................9
Froude hypothesis ...........................................................................9
4.5. Flow disturbances generated by a ship 9
4.6. Some ‘rules of thumb’ in hull form design 13
4.7. Restricted water effects 19
4.8. Evaluation of ship resistance 22
The methods based on former experimental evidence ..................22
Theoretical methods ......................................................................22
Evaluation of ship resistance with an aid of model tests ..............25
4.9. Main features of different ship types 27
Slow ships, Fn<0.18......................................................................27
Ships operating at a moderate speed range, 0.18<Fn<0.25 ..........28
High speed displacement mono-hull ships, 0.25 < Fn <0.36 ........30

5. PROPULSION AND PROPELLERS .................................................. 31


5.1. General 31
5.2. Features of a good propulsor 32
5.3. Momentum theory of propeller action 32
Efficiency of ideal propulsor ........................................................34
Bollard pull ...................................................................................35
5.4. Geometry of a marine screw propeller 36
Blade sections ...............................................................................37
Propeller pitch and pitch angle .....................................................39
Propeller Plane and Propeller Reference Line ..............................40
Rake 40
Blade Contours and Areas ............................................................41
5.5. Open water characteristics of propeller 43
Effect of propeller pitch on the open water characteristics...........45
The effect of other propeller parameters on the open water
characteristics ...................................................................46
5.6. Controllable pitch propeller 46
Features of CPP.............................................................................47
Efficiency of CPP .........................................................................47
5.7. Propeller-hull steady interaction 50
Self-propulsion model test ............................................................53

Jerzy Matusiak
3/88

5.8. On propeller selection and design 56


Number of blades ..........................................................................56
Optimum shaft speed ....................................................................57
Optimum propeller diameter .........................................................60
5.9. Propeller-main engine interaction 62
Bollard pull and propeller – main engine interaction ...................63
Controllable pitch propeller and main engine ...............................64
5.10. Propeller cavitation, vibration excitation and noise 65
Types of propeller cavitation ........................................................65
Cavitation and propeller performance ..........................................68
Propeller induced pressures and noise ..........................................70
Cavitation types and broadband pressure spectrum ......................71
5.11. Special types of propulsors 75
Ducted propellers ..........................................................................75
Z-drive units and podded propellers .............................................77
Counter-rotating propellers ...........................................................80
Hybrid propulsors .........................................................................81
Water-jets ......................................................................................82

6. REFERENCES ..................................................................................... 86

Jerzy Matusiak
4/88

1. INTRODUCTION
This document comprises the lecture-notes aimed for people studying at or graduated
from a Technical University or Polytechnic Institution. The presented material covers,
what is regarded by the author most important in Ship Theory, especially for the
people working with the propulsion systems of ships. The lecture-notes are based
mainly on three Finnish language lecture-notes of the author (Matusiak, 2007, 2008,
1993). These are used in three courses at the Helsinki University of Technology as a
part of Naval Architecture study program.

Jerzy Matusiak
5/88

2. SHIP RESISTANCE

2.1. General

Ship resistance is a very important factor with a direct influence on the costs of
operating the vessel. That is the knowledge of it is very important already at the early
stage of ship design. For a ship designer it is a big challenge to predict accurately the
resistance and to design a vessel with resistance being at the minimum. Apart low
resistance, the design has to fulfil other requirements such as: safety of the vessel,
cargo capacity, good general arrangement, low construction and operation costs, etc.

2.2. Definition of ship resistance

There are several matters that affect ship resistance in a real sea environment. These
are among the others: waves, wind, currents, shallow water and other vessels.
However, ship hydrodynamic design starts from an idealized situation where ship is
proceeding with a constant speed with no drift in calm deep water with no wind
action. The forces acting on a ship are:
• The weight of a ship
• The environmental (hydro- and aerodynamic) forces acting on a hull
with no propulsor
• The environmental (hydrodynamic) forces acting on a propulsor.

The environmental (hydro- and aerodynamic) forces can be expressed as a distribution


of stress q as presented in Figure 4.1.

Jerzy Matusiak
6/88

F Fz

M G V
x Fx r dS
!

q n

Figure 4.1 Hydrodynamic forces acting on a ship proceeding with a constant


speed V.

Stress q is a vector quantity having a normal component to the surface being the
pressure p (acts along the unit normal vector n) and shear stress component t. When
ship is at rest, shear stress is zero and the pressure equals the hydrostatic pressure.
Shear stress is associated with the friction between the flow and hull plating.
Integration of the stress over the hull surface yields the force and moment

F= q dS , M = r x q dS (4.1)
S S

in respect to the co-ordinate system which origin is located in COG of the ship and x-
axis is oriented astern. The x-component of the force F, that is F , is called ship’s total
x

resistance with operating propulsor. This is somewhat bigger than total resistance R T

of hull with no propulsor with the difference


∆Rp= F - Rx T (4.2)

being called the thrust deduction. This name is somewhat misleading because actually
it means an increase of resistance due to the action of the propulsion unit.
The product of total resistance and ship speed, that is
PE = R T V (4.3)

is called effective power. Total resistance can be obtained by integrating both the
pressure and the shear stress as follows

(4.4)

Jerzy Matusiak
7/88

where first of the integrals yields the so-called pressure resistance and the second
integral gives the frictional resistance. The wave-making of hull has a big effect on
the pressure resistance while the frictional component is governed by the shear
stresses ruled by friction of water particles and hull resulting in flow local deceleration
and in the formation of boundary layer.

2.3. The effect of propulsor on hull flow and resistance

Propulsor, usually screw propeller, located at ship stern accelerates the flow.

RT

RT + !RP

Figure 4.2 Propeller accelerates the flow at stern.

As a result the pressure at stern decreases and shear stresses increase. This yields an
increase of resistance as both, the pressure and the frictional resistance component
increase. In order to have a ship proceeding with a constant speed, thrust T of the
propeller has to be equal to the total resistance, that is
. (4.5)

If we write (4.5) as follows

(4.6)

additional resistance DR seemingly decreases the thrust delivered by the propulsion


P

unit (propeller in particular). For this reason this additional resistance is called thrust
deduction. Non-dimensional term t of (4.5), being a measure of it, is known as thrust
deduction factor. A typical value of it is 0.1.

Jerzy Matusiak
8/88

2.4. Similitude and order of magnitude

Ship hydrodynamics makes much use of the dimensional analysis. According to


Wikipedia (2008), dimensional analysis is a conceptual tool often applied in physics,
chemistry, and engineering to understand physical situations involving a mix of
different kinds of physical quantities. It is routinely used by physical scientists and
engineers to check the plausibility of derived equations and computations. It is also
used to form reasonable hypotheses about complex physical situations that can be
tested by experiment or by more developed theories of the phenomena.

In hydrodynamics, dimensional analysis is the tool to conducting model tests. It also


allows to weigh the significance of different factors on the considered flow. For
instance, it allows to judge whether the effects of flow viscosity or these of the
deforming water free surface have to be taken into account when conducting an
experiment or when constructing a mathematical model of a flow in question. Usually,
the analysis starts with making the governing equations, which are based on mass and
momentum conservation conditions, non-dimensional. This yields:

Reynolds number Re=VL/n (4.7)

Froude number (4.8)

Euler number (4.9)

Strouhal number (4.10)

where the characteristic global quantities are:

V Characteristic flow velocity; ship speed

L Characteristic length; ship waterline length when ship resistance is


considered

n Kinematic viscosity coefficient; physical property of the fluid (water)

g Gravitational constant

p0 Reference pressure such as atmospheric or cavitation pressure value

r Water density

T Characteristic time such as period of harmonically varying flow initial or


boundary conditions

Jerzy Matusiak
9/88

Non-dimensional form of resistance

It is customary and very useful to express the resistance or the components of it in a


non-dimensional form as follows

, (4.11)

where S is so-called wetted surface, that is hull surface, which is in contact with water
while ship is at rest. Resistance coefficient is related to the Euler number presented
above.

Froude hypothesis

William Froude (1810 - 1879) has divided total resistance coefficient of a surface
vessel into two coefficients as follows
, (4.12)

where C is the so-called residual resistance coefficient and C frictional resistance


R F

coefficient. His hypothesis was that the residual resistance coefficient is dependent
upon the Froude number (4.8) while the latter depends on the Reynolds number (4.8)
solely. If we present ship resistance definition (4.4) in a non-dimensional form as

(4.13)

we notice the resemblance of (4.13) to the resistance as presented by Froude.

2.5. Flow disturbances generated by a ship

Surface vessel generates a complex flow disturbance to the sea. Firstly, the waves are
generated. Roughly, these can be divided into transverse and divergent waves, as
shown in Figure 4.3. Generated waves are limited to a certain angular sector and they
are stationary in respect to ship. The length of the waves increases with Froude
number. According to the Froude hypothesis, the generated wave pattern is
independent of the scale and depends solely on the Froude number. Heights and
lengths of the waves are linearly related with the geometrical scale.

Jerzy Matusiak
10/88

Figure 4.3 Ship generated flow disturbances

In the immediate vicinity of a hull viscosity of the flow creates a thin decelerated flow
region. This thin sheet of flow having lower speed is called the boundary layer. The
thickness of boundary layer increases downstream. When passing aft shoulders and at
ship’s stern the region of decelerated flow increases rapidly the thickness and results
in a flow region called wake. The pressure distribution, such as shown in Figure 4.4, is
associated with a wave-making. High pressure at hull causes a raise of water surface,
that is a crest in the created wave. Trough in wave pattern is associated with a low
pressure region at hull.

Jerzy Matusiak
11/88

Figure 4.4 An example of computed pressure distribution on a hull and wave-making.


Red color means high pressure and wave elevation. Blue color means low pressure
region and wave trough (computed with Finflo by Jussi Martio).

Computationally evaluated wave pattern of a fast Ro-Pax vessel is shown in Figure


4.5.

Jerzy Matusiak
12/88

Figure 4.5 Waves generated by KRISO containership. Fn=0.26. Computed with


Finflo.

Viscosity of a flow, apart decelerating the flow, causes shear stresses on hull plating
and results in frictional resistance. An example of shear stress distribution at a stern of
a fast Ro-Pax vessel, as computed by Finflo, is shown in Figure 4.6

Figure 4.6 Shear stress distribution at a stern of a fast Ro-Pax vessel, as computed by
Janne Niittymäki (2001) using CFD-tool Finflo. Red color means high stress values.

Jerzy Matusiak
13/88

Quite often it is feasible to use a bow bulb in order to decrease the wave created by
ship bow. For fast and moderately fast vessels, bow bulb is designed as a streamlined
body having well concentrated volume located ahead of stem (see Figure 4.7).

Figure 4.7 Pressure distribution, selected streamlines and wave elevation at bow of a
fast Ro-Pax. Computed by Jussi Martio with an aid of Finflo.

Bow bulb creates own waves, which interact with the waves generated by bow part of
a hull and results in a decrease of a total wave system. A decrease of wave-making has
a very beneficial effect on the associated part of the resistance.
For relatively slow vessels, such as tankers or bulk carriers, bow bulb design
aims at decreasing the wave-breaking at bow. This has also a good effect on the
resistance of a ship. In such a case bulb form is different. The volume of it is smaller
and it is better integrated into the hull form.

2.6. Some ‘rules of thumb’ in hull form design

The flow disturbance generated by a hull is very complex. That is why it is difficult to
give generally applicable rules for a good hull form. However, some observations can
be made. These are based mainly on the model tests, analytical considerations
concerning wave-making and sophisticated calculations conducted using
Computational Fluid Dynamics.
Wave-making resistance is related to the height of the generated wave squared.
Thus it is crucial to have wave pattern generated by a hull as low as possible. Regular

Jerzy Matusiak
14/88

sine-form wave of a high amplitude means high resistance. Especially high bow wave
contributes much to the resistance. High wave at stern is good as it decreases the
resistance. In a sense high pressure at stern ‘propels’ ship. However, too high wave at
stern may result in a deceleration and separation of a flow. Flow separation should be
avoided. Apart increasing the frictional resistance, separation has several negative
effects. In particular it may affect propeller efficiency, cause vibration and noise
problems. It can also be a cause of propeller ventilation (see Figure 4.8), which further
worsens propeller operation.

Figure 4.8 Propeller ventilation

Large scale separation of flow at stern results in a low pressure region. This low
pressure and propeller action may cause a suction of air from a nearby free surface. As
a result, a mixture of water and air enters propeller disc. This results in a decrease of
thrust and torque, vibration and noise.
It is commonly believed that ship’s bow shape has a most important effect on
hull wave-making resistance. From Figures 4.5 and 4.9 it can be seen that whole hull
contributes to wave-making (Figure 4.5) and to the associated resistance (Figure 4.9).

Figure 4.9 Accumulation of wave-making resistance of ro-pax vessel. Fn = 0.277.

Jerzy Matusiak
15/88

From Figure 4.9 it is seen that at bow both stern-wards oriented normal pointing
inside the hull and raised water surface (wave crest) contribute to resistance increase.
High wave at stern, where normal points bow-wards, decreases the resistance.

The effect of bow shape and that of the bow bulb is easy to observe in model tests.
The effect of stern form on wave-making is far more difficult to observe, although it is
nearly of the same importance. Stern forms of the twin-screw vessels are shown in
Figure 4.10.

Figure 4.10 Stern types of twin-screw vessels.

The flow at stern is very complex. Both wave-making and the viscous effects are
equally important. Apart trying to minimize the resistance the quality of flow at stern
should be taken care for. A form of stern has to be sufficiently streamlined, so there is
no flow separation.

An example of flow separation as detected by model tests (the so-called paint test) and
computations is shown in Figure 4.11.

Jerzy Matusiak
16/88

Figure 4.11 The result of paint-tests (upper Figure) and computed shear stresses at
the stern of the so-called HSVA tanker (Schweighofer&Helllsten, 1999).

Separation of flow is visible at stern as the regions of the rapidly diverging


streamlines (changing directions of shear stresses). A strong bilge vortex is visible,
too.

Moreover, for a single screw or so-called twin-skeg vessel (refer to Figure 4.10) we
try to form the stern so that the wake at the propeller plane is as much as possible
axially symmetric (see Figure 4.12).

Figure 4.12a More homogeneous and Figure 4.12b Worse nominal wakefield.
nearly axially symmetric nominal Propeller blades encount strongly
model wake varying flow

Jerzy Matusiak
17/88

The wake is presented as the fraction of the local flow axial velocity component
related to the ship’s speed. So for instance the isoline 0.5 means that the axial velocity
of the flow in the propeller plane is half the ship speed. The so-called nominal model
wake of Figure 4.13b is worse than the one of Figure 4.13a. The in-flow velocities of
the b-case change rapidly with angle q. This means that the flow at propeller blades
changes much in time causing possibly cavitation, vibration and noise problems. The
words ‘nominal’ and ‘model’ mean that the wake was measured or computed for a
model with no propeller being in place.

The result of nominal model wake measurement and an attempt to scale the wake to
the ship’s scale are presented in Figure 4.13. Model tests of surface vessels are nearly
always conducted according to the Froude’s scaling law. This, in practice, means that
speed of the model is smaller than that of the ship, where l is model scale. As a
consequence of this, Reynolds number of the experiment is much lower than the one
of a ship flow. This means that the role of viscosity is too high in model experiments.
The scaling of model results to ship scale of Figure 4.13 attempts to take the
difference in viscous effects into account. The scaling results in stronger contraction
of the boundary layer (thinner layer of decelerated flow) in prototype scale.

Jerzy Matusiak
18/88

Figure 4.14 Nominal model axial wake as measured (lower Figure) in model tests and
scaled to full-scale (upper Figure).

Jerzy Matusiak
19/88

2.7. Restricted water effects

Shallow water affects ship resistance and running position. It can be thought that
proximity of sea bottom and channel banks or other nearby passing vessels restrict the
flow. As a result flow velocities increase. This in turn decreases the pressures and
results in squat and increased resistance. Squat means an increase of running draft of
the vessel by DT. Three types of restricted water are shown in Figure 4.15 after
Huuska (1976).

Figure 4.15 Three types of restricted water (Huuska, 1976).

The so-called Froude number based on water depth is a good measure of squat and
increased resistance. This number is defined by

, (4.14)

where h is water depth. A plot presenting the limiting values of depth based Froude
number and water depth in relation to ship draft is presented in Figure 4.16 after
Harvald (1983).

Jerzy Matusiak
20/88

Figure 4.16 Approximate guidance on shallow water effect (Harvald, 1983)

An approximate formula, based on two-dimensional simple flow model, of Tuck &


Taylor (1970) yields an estimate of squat

, (4.15)

where L is length between the perpendiculars and Ñ volumetric displacement. The


pp

formula 15 can be applied up to Fn = 0.8. A schematic presentation of a change of the


h

running position with depth based Froude number is shown in Figure 4.17.

Increase of a Froude number means a growth of sinkage and a slower increase of trim.
Ship resistance increases. Maximum of sinkage and trim is reached close to the
critical Froude number, which is Fn ≈ 1. If this critical speed is overcome, sinkage
h

changes rapidly a sign and ship elevates. Resistance decreases.

Jerzy Matusiak
21/88

Figure 4.17 A change of running position with depth based Froude number.

Proximity of a running ship from a quay or from a closely passing other ship involves
acceleration of a flow velocity. This is associated with a decrease of pressure, which
involves a suction effect. The situation is sketched in Figure 4.18. Ship does not have
necessarily to be running to experience this effect. Even with a stationary ship,
operating propeller accelerates the flow bounded by hull and quay and causes a
suction used to berth the ship.

Jerzy Matusiak
22/88

Figure 4.18 A suction effect of a quay on a ship (a rough sketch showing a


principle of the phenomenon).

2.8. Evaluation of ship resistance

Evaluation of ship resistance is done in a number of ways of different accuracy and


required effort.

The methods based on former experimental evidence

The simplest and often a very reliable way is a simple estimation based on the
experience gained from previous similar projects. This approach works well when we
have in our databank model and full-scale test results for a similar vessel to the
designed one. There is also fast and easy to use methodology based on using large
number of model tests conducted for a large variety of vessels. In this approach either
published systematic model experiments for a certain ship type are utilized or a well
processed and organized large data bank of model tests is used (Holtrop, 1977&1984).
Unfortunately using these publically available databases yields usually a result, which
can be used as a preliminary and not very accurate estimate, only.

Theoretical methods

Theoretical methods of evaluating ship resistance can be put into two catagories.

The methods based on the potential flow approximation

The so-called potential approximation to the flow means that an assumption of an in-
viscid and irrotational flow is made. This means that the viscous effects are neglected.
The approach results in an evaluation of wave-making. The wave pattern created by a
ship is evaluated. Also the wave-making resistance is produced by integrating the
computed pressure distribution. The accuracy of the estimated wave-making

Jerzy Matusiak
23/88

resistance may be questioned, especially for low Froude numbers. The method is very
efficient when ranking the versions of the same hull and when optimizing the hull
form in respect to the wave-making. Numerical methods solving the problem are
usually called as Rankine panel method or boundary element method. The method
requires a discretization of hull and that of the free surface. An example of the
numerical model is shown in Figure 4.15.

Figure 4.15 Discrete (panel) model of Series 60 ship used for wave-making
evaluation.

The result of computations compares well with the more sophisticated (viscous effects
including) methods when the pressures at bow are compared (Figure 4.16). At stern,
where viscous effects play important role, a comparison gives much worse agreement.

Jerzy Matusiak
24/88

Figure 4.16 Pressure distribution on a bow of cruise-liner. Upper Figure as


evaluated by Finflo (including viscosity) and using a potential approximation of
Shipflow (lower Figure), (Lönnberg, 2000)

Computational Fluid Dynamics

Genuine CFD, with an allowance for wave-making, has made a breakthrough in naval
hydrodynamics approximately 10 years ago. At present, there are a dozen of methods
capable to evaluate steady viscous flow with a deforming free surface. Accuracy of
the computations in terms of total resistance is usually better than 5% when compared
to the model test results. Apart resistance, computations yield a vast data on
distributions of pressure and shear stresses, vortices, streamlines and other important
flow details. Computations can be conducted for both the model and the full-scale
yielding important information on scaling effects.
For the time being, CFD is not used as a routine tool in ship design practice. The
reasons are: tedious input data preparations (computational mesh comprising millions
of cells), slow computation times, high costs of required investments in trained
personnel, hardware and software.

Jerzy Matusiak
25/88

Evaluation of ship resistance with an aid of model tests

Theoretical models and computational methods are not sufficiently accurate and fast
enough to rely on when evaluating still water resistance. Especially evaluation of the
required power by the prototype vessel requires more traditional and safe approach,
namely model tests. William Froude set the foundations of conducting and analysis of
model tests. Tests are conducted with a model representing a hull of the ship in a
geometrical scale l. Model is towed with a velocity l times lower than the ship
0.5

velocity, that is V =V /l . The measured quantities are model resistance and running
M S
0.5

position. Tests are photographed. Photograph taken during a resistance model test of
MV Uikku is shown in Figure 4.17.

Figure 4.17 Model of MV Uikku in model basin of TKK. Students’ project laboratory
exercise.

The ITTC-57 method

The so-called ITTC-57, which was formally accepted as a standard method by the
International Towing Tank Conference in 1957, is still a common method used to
evaluate resistance of a ship. In the method:

• Measured model resistance RTM, is made non-dimensional as

(4.16)

Jerzy Matusiak
26/88

• Frictional resistance coefficient is estimated both for a model and for a ship
using the so-called "ITTC-57” correlation (friction) line formula

, where i = M stands for model and i = S for ship. (4.17)

• Residual resistane is obtained by subtracting frictional resistance coefficient


from the coefficient of the total model resistance, that is

CRM = CTM ! CFM (4.18)

• It is assumed that the residual resistance coefficient is the same for a model
and for ship, that is C =C =C and the total resistance coefficient of a ship is
R RS RM

CTS = CFS + CRM + CA , (4.19)

where CA is so-called additional resistance coefficient, which is based on the


correlation between the model test predictions and sea-trials of many ships.
Ship resistance is calculated from

. (4.20)

• Effective power is calculated using definition (4.3), that is

. (4.21)

The scaling principle is shown in Figure 4.18. It is clearly seen from the Figure 4.18
that in model scale the viscous effects have much more significance than in the ship’s
scale.

Jerzy Matusiak
27/88

0.01

CTM

CR

C CTS
CA
C FM
CR

CFS

0.001
1.E+05 1.E+06 1.E+07 1.E+08 1.E+09
Re

Figure 4.18 ITTC 1957 method of scaling the model resistance to the full-scale.

If ship has appendages, such as propeller shaft, supporting them struts, bossings etc.
then in some research institutions resistance test is conducted twice. First test is done
for a so-called naked hull, that is hull without appendages and the second tests is run
with model equipped with the appendages. Only a part (for instance 60%) of a
difference in total resistance coefficient is taken into account when evaluating the
resistance due to appendages. In this way, a difference in Reynolds numbers of model
tests and the full-scale is taken into account.

2.9. Main features of different ship types

In the following, displacement mono-hull ships are divided into three categories
according to their speed. A rough guidance concerning main features of them is given.
This is based on few recently designed and constructed merchant ships.

Slow ships, Fn<0.18

The resistance of slow ships is governed by a viscous part. That is why when
designing them, we try to prevent flow separation and keep the frictional stresses low.
In the same time, we to try keep breaking of bow-wave as low as possible. An
example of such a vessel is shown in Figure 4.19.

Jerzy Matusiak
28/88

Figure 4.19 Shuttle tanker. Wave profile at Fn=0.16. Courtesy of Mr. R. Hämäläinen
of Aker Yards Turku.

As seen from the cross-sectional area distribution (upper solid curve), the longitudinal
Centre of Buoyancy (lcb) is located forward of the amidships bow-wards.
Recommended location is 2-3% of the ship length. Usually this type of ships have a
single propeller. Prevention of flow separation at stern and securing good in-flow into
propeller are the main concerns in stern design. As there is no much wave-making, the
form of cross-sectional area distribution recalls that of a submersible. The role of a
bow bulb is to prevent breaking of the bow wave and to homogenize pressure
distribution in the bow area. Bow shoulders can’t be too sharp. Otherwise, flow may
separate from them or a rapid acceleration of flow may result in a deep trough in wave
pattern, low pressure region and trimming bow down.

Ships operating at a moderate speed range, 0.18<Fn<0.25

The wave-making starts to play role at this speed range. Hull hydrodynamic design
aims at keeping wave-making resistance low. The prismatic coefficient of this type of
vessels is 0.80>C >0.60. An example of such a vessel is shown in Figure 4.20.
p

Jerzy Matusiak
29/88

Figure 4.20 Cruise-liner. Wave profile at Fn=0.22. Courtesy of Mr. R. Hämäläinen of


Aker Yards Turku.

The wave generated at Fn=0.22 is much longer. It is recommended to have a slender


and longer fore-part of a ship than the aft-part. As a result the lcb moves aft as
presented in Figure 4.21.

Figure 4.21 A guidance of the longitudinal location of centre of buoyancy in respect


to amidships (Hämäläinen&van Herd, 1998).

Normally for ships operating at speed Fn<0.25, bow bulb does not have much
influence. However, for a high beam to draft ratio (B/T > 3.5), it’s influence is
important. Bow bulb lowers the entrance angles of waterlines and smoothens the bow
shoulder. Thus it has a big influence on wave-making. Another means of decreasing
wave-making resistance in this ship type is the so-called Wave Damping Aftbody
(Hämäläinen&van Herd, 1998). Just before reaching transom the verticals at stern
bend down decreasing wave generated by stern without much influence on pressure at
stern.

Jerzy Matusiak
30/88

High speed displacement mono-hull ships, 0.25 < Fn <0.36

The prismatic coefficient is kept low, but higher then C =0.58. Usually vessel has a
p

twin-screw propulsion arrangement. Stern is of transom type, often with the so-called
duck tail, which controls a separation of flow, lengthens the effective (wave-making)
ship length and thus lowers the effective Froude number and results in a lowering the
wave-making. Well designed bow bulb has a big effect on wave-making resistance.
Hull does not have a parallel mid-part. An example of a fast mono-hull vessel is
shown in Figure 4.22.

Figure 4.22 Fast Ro-Pax vessel. Frouden number is Fn=0.33. Courtesy of Mr. R.
Hämäläinen of Aker Yards Turku.

Jerzy Matusiak
31/88

3. PROPULSION AND PROPELLERS

3.1. General

When defining ship resistance the stresses acting on hull were considered and
integrated into the forces and moments acting on a ship. This was done excluding the
surfaces representing a propulsor. In Figure 5.1 most common type of ship propulsors,
that is a marine screw propeller, is sketched propelling a ship.

Figure 5.1 The forces acting on a propeller.

Loading acting on a propulsor, propeller in this case, is considered. The stress p


distribution comprises pressure p acting normally to the propeller surface and shear
stress component t tangential to the surface. Integration of stress p yields the propeller
force F and resulting moment M . Usually integration is conducted using the co-
p p

ordinate system fixed with the propeller as it is at rest. In other words, the forces are
expressed in relation with the co-ordinate system which does not rotate with propeller.
Projection of the total force F on the x-axis of a ship is called thrust and marked by
p

symbol T. Usually propulsor generates thrust by accelerating water or air passing


through it. An increase of momentum is associated with a generation of a reaction,
which propels a ship. Hydrodynamic reaction felt by a propulsor as moment M has top

be balanced by a same magnitude moment delivered to the propulsion unit by on-


board machinery.

Jerzy Matusiak
32/88

3.2. Features of a good propulsor

• Steady thrust
• Good efficiency
• Thrust is easily controlled and directed (good acceleration and stopping
qualities)
• Propulsion is well suited to hull and vice versa
• Reliability
• Small investment and operational costs
• Does not cause vibration nor noise
• Operates well in a variety of conditions (ice, shallow water, etc.)

3.3. Momentum theory of propeller action

The simplest representation of propulsor action is given by the momentum theory.


Propulsor is represented by a disk accelerating the on-coming flow. The generated
thrust is uniformly distributed over the disk area. The flow is frictionless and
irrotational. Sometimes a term ideal propulsor is used to describe this kind of flow
machine, which efficiency is higher than that of a real type propulsor such as marine
screw propeller. The action of a disk accelerating oncoming flow is shown in Figure
5.1. The disk is moving in the right direction with a so-called advance velocity V . IfA

we fix the co-ordinate system with the disk, then the flow velocity upstream of the
disk is V . The pressure upstream is denoted by p . The same pressure is far
A 0

downstream.

Jerzy Matusiak
33/88

Figure 5.1 Propeller disk imparts acceleration on oncoming flow.

We use Bernoulli equation and the momentum equation to derive the velocities in the
propeller plane and far downstream. Propeller brings energy to the flow. That is why
we apply Bernoulli equation separately for the flow upstream and before the propeller
disc (5.1) and for the flow leaving the disc and extending downstream (5.2).

(5.1)

, (5.2)

where U and U are additional velocities due propeller action downstream of the
A0 A

propeller and in the propeller plane respectively. p’ and p’’ are the pressures just
before and behind the propeller disc. The pressure jump over the propeller disk is

. (5.3)

Pressure difference (5.3) and disc area A yield thrust


0

. (5.4)

In order to get relation between the velocities U and U the momentum equation is
A0 A

used. The momentum difference downstream and upstream equals the thrust, that is
, (5.5)

where is mass flux thru the propeller disk, which equals

Jerzy Matusiak
34/88

. (5.6)

Comparing equations 4 and 5 yields the relation

, (5.7)

which means that propeller induced velocity in the propeller plane is twice smaller
than the induced velocity far downstream.

Efficiency of ideal propulsor

Next we derive the efficiency of the propulsor by dividing the work done by it in unit
time, that is

, (5.8)

by the energy lost in a form of kinetic energy of the water passing through the disc,
that is

. (5.9)

As a result the efficiency of propulsor idealized as a disc accelerating the flow in axial
direction is obtained as

. (5.10)

Another way of expressing the efficiency of the idealized propulsor is to use thrust
loading coefficient defined by

. (5.11)

Inserting thrust (5.4) into (5.11) yields

(5.12)

and the relation of velocities upon the thrust loading coefficient in the form

(5.13)

Jerzy Matusiak
35/88

Using (5.13) yields the efficiency in the form

(5.14)

which is presented in Figure 5.2.

Figure 5.2 The efficiency of the idealized propulsor as a function of the thrust loading
coefficient.

It is clearly seen from Figure 5.2 and thrust definition (5.4) that propeller disc area has
to be sufficiently large in order to secure sufficiently high efficiency. This explains
why a large single propeller is usually better than the two small ones.

Bollard pull

The bollard pull means thrust delivered by a stationary propulsor. This can be
obtained by setting advance velocity to zero, that is V =0 in (5.4). This results in
A

(5.15)

The propulsion power is related to by the following formula (refer to (5.9))

(5.16)

Solving induced velocity U from formula (5.16) yields


A0

, (5.17)

Jerzy Matusiak
36/88

which substituted into the thrust expression (5.15) yields the bollard pull that is the
maximum attainable thrust

. (5.18)

Bollard pull given by (5.18) is approximately 30% higher than the one given by a
semi-empirical expression for a bollard pull of propeller with no duct

. (5.19)

The difference can be attributed to other power loses associated with propeller
operation. In particular the effect of rotational velocity and friction on the propeller
efficiency are not taken into account by a simple momentum theory model.

3.4. Geometry of a marine screw propeller

This Chapter of the lecture notes is mainly based on the book of Kuiper (1992). The
book was published on the occasion of the MARIN’s 60 anniversary.
th

A sketch of a propeller is given in Figure 5.3.

Figure 5.3 Sketch of the right-handed propeller.

Jerzy Matusiak
37/88

The propeller blades are attached to the hub, which is fitted at the end of the propeller
shaft. The propeller rotates about the shaft center line. The direction of rotation is
viewed from behind, that is towards the shaft. In normal forward operation a right
handed propeller rotates in clockwise direction when viewed from behind. The
propeller in Figure 5.3 is right-handed. The front edge of the blade is called the
leading edge. The other edge of the blade is called the trailing edge. The outermost
position, where leading and trailing edges meet, is called the blade tip. The radius at
which the tip is found is the propeller radius. The propeller diameter is, of course,
twice the outer radius.

The surface of the blade which is at the side of the shaft is called the propeller back.
The other side is the face of the propeller. (When the ship has forward speed the
propeller moves with its back forward) Because in forward speed the back side has a
low pressure and the face side has a high pressure (which difference generates the
thrust), the back is also called the suction side and the face the pressure side. These
names are less ambiguous than face and back and are therefore to be preferred. The
propeller hub is of course rotational symmetric because it should not disturb the flow.
The attachment of the propeller blade to the hub is gradual, which is done in the fillet
area or blade root. A streamlined cap is generally fitted to the hub.

Blade sections

We consider a cylindical cut of a screw propeller, as drawn in Figure 5.4.

Figure 5.4 Cylindrical cross section of a propeller blade.

Jerzy Matusiak
38/88

The intersection of a cylinder with radius r and a propeller blade, the blade section,
has the shape of an airfoil. Such a shape is also called just a foil or a profile. Some
characteristic parameters of a foil will be defined first. A general shape of a profile is
shown in Figure 5.5.

Figure 5.5 Blade profile and geometry definition. (paksuus=thickness,


kaarevus=camber, imu=suction, paine=pressure).

The side, which meets the flow is the leading edge of the profile. The trailing edge is
generally sharp. Let us first assume a sharp trailing edge, because this facilitates the
definition of a coordinate system in which the profile coordinates are defined. The
leading edge is found as the point on the contour with the largest distance from the
trailing edge. Other names for leading and trailing edge are nose and tail. The straight
line between the leading and the trailing edge of the profile is the chord of the profile
and the distance between nose and tail is the chord length c. The chord line is also
called the nose-tail line. The trailing edge is not always sharp, however. In that case
the chord-line is defined as the direction of the maximum distance between two points
on the contour. This direction has to be found iteratively in such a case. (A different
definition of leading and trailing edge will be given below). Generally, the origin of
the local coordinate system of a profile is taken at the leading edge. The x-direction is
towards the tail, the y-direction upwards, perpendicular to the chord. The angle
between the nose-tail line and the undisturbed flow is the angle of attack a. Its
positive direction is given in Figure 5.5. At a positive angle of attack the pressure at
the upper side of the profile is lower than the pressure in the undisturbed flow and this
side is therefore called the suction side. The pressure at the lower part is higher than
the pressure in undisturbed flow over most of the chord and is therefore called the
pressure side. These names match with the names of the corresponding blade surfaces.
The distance between the suction side and the pressure side, measured perpendicular
to the chord, is the thickness distribution t(x) of the profile (see Figure 5.5). The line
through the middle of the thickness is the camber line of a profile. The vertical
distance of the camber line to the nose-tail line is the camber distribution f(x). The
camber and thickness distributions are often made non-dimensional with their
maximum values, so that the camber and thickness distributions are given in values up
to 1. When the same camber and/or thickness distribution is used for the blade
sections at all radii, as is often the case, the blade sections can simply be described by
the maximum thickness and camber. The maximum thickness and maximum camber

Jerzy Matusiak
39/88

are often given as percentages of the chord length. Also the positions where these
values occur, expressed in percentages of the chord, are sometimes used as
parameters. (A section is described then e.g. as having 2% maximum thickness and
1% maximum camber, with the position of maximum camber and thickness at 35%
from the leading edge. The distributions of camber and thickness are then assumed to
be known).

Propeller pitch and pitch angle

The cylindrical cross section of a propeller blade, as shown in Figure 5.4, is now
developed into a plane. (Figure 5.6).

Figure 5.6 Cylindrical cut of propeller (Figure 5.4) expanded into the plane.

In this figure the left and right edge are the cut in the cylinder and the width of the
developed plane is 2pr. In this developed plane a number of parameters can be
defined.

The chord-line or nose-tail line of the blade section changes from a helix on the
cylinder into a straight line, and its extension is called the pitch line. I, The propeller
pitch "P" is defined as the increase in axial direction of the pitch line over one full
revolution 2pr. The dimension of the pitch is a length. The pitch angle f is the angle
between the pitch line and a plane perpendicular to the propeller shaft. The relation
between pitch and pitch angle is

. (5.20)

Jerzy Matusiak
40/88

The pitch distribution is given in a pitch diagram, which is simply a graph of the pitch
at every radius. The pitch diagram is given in the propeller drawing. If pitch has a
variable distribution it is given at least at the radius r/R=0.7, which is regarded as the
most significant blade section.

Propeller Plane and Propeller Reference Line

A coordinate system in which the geometry of a propeller is expressed is always


chosen as a cylindrical coordinate system (x, r, q) fixed to the propeller, with the
positive x-axis as the shaft center line in the direction of the propeller back. (see
Figure 5.3). The origin of the coordinate system is chosen arbitrarily on the shaft
center line and the plane through the origin and perpendicular to the x-axis is called
the propeller plane. The coordinate q = 0 is chosen arbitrarily also in the propeller
plane. The line q = 0 in the propeller plane is called the propeller reference line
(Figure 5.6).

In Figure 5.6 the intersection of the propeller plane with the expanded cylinder is
drawn. The x-axis is the intersection of the plane q = 0 with the expanded cylinder.
These two intersections reflect the propeller coordinate system.

In practice the coordinate system is often chosen arbitrarily. In principle, an additional


point in space defines the propeller reference line. A possible point is the mid-chord
location of the blade root section. This definition is also used by the ITTC. In that case
the coordinate system is completely defined for arbitrary propeller geometry.

Rake

Having defined the coordinate system some other parameters can be defined in Figure
5.6. The x-axis intersects the pitch line at a point on the generator line and the
distance between the generator line at a certain radius and the propeller plane is called
the rake. Rake therefore has the dimension of a length. When the rake is away from
the ship hull (in the direction of the negative x-axis and increasing the tip clearance) it
is called positive rake or also backward rake. This direction is the common direction
for propellers. When there is no rake the propeller reference line coincides with the
generator line.

Only in case of a linear rake distribution from root to tip the generator line is a straight
line in the plane q = 0. In that case the angle between the generator line and the
propeller reference line is called the rake angle. The rake angle is positive in case of
backward rake. An example of linear backward rake (without skew, see later) is
shown in Figure 5.7.

Jerzy Matusiak
41/88

Figure 5.7 Definitions of propeller blade rake and skew.

Backward rake is used to increase the tip clearance, the distance between a propeller
tip in top position and the hull. When this is the only purpose of the rake the rake
distribution is mostly linear.

The mid-chord of the blade section in Figure 5.6 does not coincide with the generator
line. Relative to the generator line the section is shifted along the pitch line. The
location of the midchord of the propeller section is now called the blade reference
point and its position is indicated in Figure 5.6. The projection of the distance between
generator line and blade reference line at a certain radius on the propeller plane is
called skew. When skew is in the negative direction of q it is called backward skew.

Since skew moves the blade reference point along the pitch line, the blade reference
point also moves in axial direction when skew is changed. The axial displacement of
the blade reference point due to skew is called skew induced rake. For a propeller
without skew the generator line coincides with the blade reference line.

Skew is applied to reduce the unsteady forces on the blades when a propeller blade
passes a wake peak.

Blade Contours and Areas

The projection of the blade contour on the propeller plane gives the projected blade
contour. An example is given in Figure 5.8.

Jerzy Matusiak
42/88

Figure 5.8 Blade areas.

The propeller reference line and the generator line are the same vertical axis in this
projection. So rake is not visible in the projected blade contour. In combination with
the projected blade contour the developed blade contour is sometimes given. The
intersection of the blade with the axial cylinder, as represented by a circle in the
projected blade contour, is rotated along the blade reference line (mid-chords of the
sections) into a plane parallel to the propeller plane. The amount of rotation is equal to
the pitch angle at every radius. The contour thus found is the developed blade contour
(Figure 5.8). The main flow component along a propeller, however, is along the chord
in the circular cross section of Figure 5.4. In the expanded cylinder of Figure 5.6 this
is the direction of the pitch line. The expanded blade contour is found when the
expanded blade sections in Figure 5.6 are rotated over the pitch angle into a plane
parallel to the propeller plane. The projection on the propeller plane results in the
expanded blade contour. An example is given in Figure 5.8. The blade sections in that
contour are straight lines and the section geometry is often given in this propeller
contour. These are the sections that matter hydrodynamically, since the flow passes
the blade along these sections. An important parameter of the propeller is the area of
the blades. The blade area A is given as a ratio between the area of the blade contour
of all blades and the area of the propeller plane A =pD /4. Two blade area ratios are
0
2

used: the projected blade area ratio A /A and the expanded blade area ratio A /A or
P 0 E 0

EAR. The latter ratio is physically most significant. The areas of propeller skewed
blades are shown in Figure 5.9.

Jerzy Matusiak
43/88

Figure 5.9 Skew definitions using projected and expanded bled contours.

An example drawing of propeller is shown in Figure 5.10.

Figure 5.10 Example drawing of propeller.

3.5. Open water characteristics of propeller

Important non-dimensional variables used when describing propeller performance in


open water are:
Advance number

Jerzy Matusiak
44/88

, (5.20)

thrust coefficient

, (5.21)

and torque coefficient

. (5.22)

The angular velocity of propeller in terms of revolutions per second is denoted by n.

Open water characteristic of propeller is usually obtained with an aid of a model test,
namely test called open water test of propeller. In this test, propeller constructed to a
geometrical scale l is fixed to the shaft of propeller dynamometer, which rotates the
model and measures both thrust and torque of it (5.Figure 5.11).

Figure 5.11 Open water test of a model propeller.

Whole set-up is immersed deep enough, so that the inflow into the propeller is not
affected by the water surface and by the wave-making in particular. Propeller is
rotated with high revolutions (15-20 rps) to ensure turbulent flow regime on the
blades. The speed V of water inflow is adjusted yielding a desired advance coefficient
A

(5.20). Measured thrust and torque are made non-dimensional according to (5.21) and
(5.22). Results are plotted as shown in Figure 5.12.

Jerzy Matusiak
45/88

Figure 5.12 Open water characteristics of the four bladed propeller of the
Wageningen B-series. Expanded blade area ratio is 0.7 and pitch ratio is 1.

The open water efficiency of propeller is obtained from

. (5.23)

At the bollard pull, where J=0, efficiency is zero because advance velocity is zero as
well. Propeller absorbs power without work being done. At bollard pull the angles of
attack of the in-flow into the blades are at the maximum. If flow separates as a result
of this, there is a danger of thrust breakdown. Propellers are designed to operate in
normal conditions with a maximum efficiency, that is in the advance number range
corresponding to high values of h0-curve.

Increasing advance coefficient lowers propeller loading. At certain advance


coefficient value thrust coefficient approaches zero. This point of operation is called
wind-milling. Propeller set in the in-flow will rotate absorbing energy similarly to
turbine. Propeller efficiency is zero at this point, too.

Effect of propeller pitch on the open water characteristics

Pitch has a big effect on the propeller open water characteristics (see Figure 5.13).

Jerzy Matusiak
46/88

Figure 5.13 Open water characteristics of two five bladed propeller of the
Wageningen B-series. Expanded blade area ratio is 0.7. Pitch ratios are 1 and 1.4.

Larger pitch means larger angles of attack of the inflow into the propeller blades. This
in turn means higher lift values and results in higher thrust and torque values.
Increasing pitch moves the maximum of the efficiency towards higher advance
coefficient values.

The effect of other propeller parameters on the open water


characteristics

Increasing camber of propeller blades is followed by an increase of thrust and torque


coefficients. However, this increase is not that fast as in the case of pitch.

Increasing blade area of propeller also increases thrust and torque. However, larger
blade area means increased frictional and induced drag losses, which result in a
lowering of propeller efficiency.

Increasing blade thickness results also in a lowering of the propeller efficiency.

3.6. Controllable pitch propeller

Fixed blade propellers operate efficiently in a limited range of advance ratio only. If
the revolutions of propeller shaft are kept more or less constant, varying the speed of
advance yields a situation where propeller efficiency decreases rapidly. This means

Jerzy Matusiak
47/88

that propeller’s thrust is low compared to the power it uses. Propeller having
controlled pitch can operate more efficiently in varying conditions.
Controllable pitch propeller (CPP) consists of blades, which can be turned along
the generator line. All blades are turned by the same angle. As a result pitch angle of
propeller changes by a certain amount. The logic controlling pitch setting is called
combinator curve.

Features of CPP

Good features of CPP are:


• Good stopping, acceleration and maneuvering qualities
• Thrust does not change much with loading
• Good fuel economy in case of ships operating with different speeds and
varying loading
• There is no need to have machinery with the reverse revolutions
• A change of ship’s speed does not require changing revolutions of the
main engine
• Generator can be connected to the main engine, which revolutions do
not change much.
Minuses of CPP are:
• Structural complexity
• High cost
• Need of maintenance
• Hydrodynamic efficiency lower than that of the fixed pitch propeller
(FPP)

Efficiency of CPP

Although blades of CPP are designed using the same principles as in the case of FPP,
efficiency of controllable pitch propellers is lower. There are several reasons for this.
The most important is the fact that the control mechanism requires space in the hub.
For this reason the relative hub diameter is d/D=0.30 to 0.35 while in the case of FPP
it is typically d/D=0.17 to 0.20.
When designing the blades of CPP two factors should be taken into account.
Firstly, the mechanism controlling the blades has to withstand the so-called spindle
torque, which is a torsional moment of a blade along the axis of pitch changing. The
expanded blade area ratio of propeller can’t be too high. If the value of it is higher
than 0.8 it may result in overlapping of blades and prevent their turning. Skewed
blades may have this problem even at lower expanded blade area ratio.

Jerzy Matusiak
48/88

A change in pitch angle affects a distribution of pitch and the form of propeller
profiles. CPP having in design conditions constant pitch and the corresponding pitch
angle values for three propeller radii are shown on the left of Figure 5.14.

Figure 5.14 The effect of changing the propeller pitch on it’s distribution (Dudziak
1988).

When turning the blades by angle Q, new pitch values P are obtained. These are
i

given by formula
, (5.24)

where r and f are the considered propeller radii and the corresponding pitch angles
i i

before changing of pitch. It is clearly seen that pitch distribution changes. A peculiar
blade loading develops when a decrease Q of pitch angle exceeds pitch angle at blade
tip. In this situation the loading at the tip is reversing while rest of the blade is
thrusting. This may result in pressure side cavitation and if used frequently and for
long periods may cause erosion damage to the blades.

Instead of higher price and lower efficiency at the design point, controllable pitch
propellers are very popular, especially in traditional (no-pod) propulsion arrangements
with machinery running at nearly constant rate. Open water characteristics of a CPP
propeller for different pitch settings are presented in Figure 5.15.

Jerzy Matusiak
49/88

Figure 5.15 Open water characteristics (K &K -curves) of the four bladed propeller
T Q

having expanded blade area A /A =0.48 and design pitch ratio P /D=0.7 (Dudziak,
E 0 0.7

1988).

Jerzy Matusiak
50/88

3.7. Propeller-hull steady interaction

Propeller located at ship stern accelerates the flow. Accelerated flow increases
resistance of ship hull by the amount
, (5.25)

called thrust deduction. This is expressed by a non-dimensional thrust deduction


coefficient
, (5.26)

In Figure 5.7 the power and efficiency coefficients used in ship propulsion are
presented.

Figure 5.16 Most important power and efficiency coefficients used in ship propulsion.

The definitions and symbols used when describing propeller-hull steady interaction
and propulsion are given in Table 1.

Table 1 The definitions and symbols used when describing propeller-hull steady
interaction and propulsion

Symbol Definition

Effective power PE PE = RT V

Thrust power PT PT = T VA
Brake power PB Nominal power of the main
engine
Delivered power at propeller PD PD = 2p Q n

Indicated power PI Calculated from the mean


value of the cylinder pressure

Shaft power PS Power at bow end of


propeller shaft
Propulsive efficiency hD !D = PE = PE PT = !H !0 !R
PD PT PD

Jerzy Matusiak
51/88

Hull efficiency !H !H = RT V = RT/T = 1 " t


T VA VA/V 1 " w
Propeller efficiency h0 Formula 5.23
!
Relative rotative efficiency hR !R = B
!0
Propeller efficiency behind ship hB !B = !0 !R = PT = T VA
PD 2"Qn
P D
Shafting efficiency hS !S =
PS
Thrust deduction factor t t = T ! RT
T
Wake fraction w w = 1 ! VA
V
Thrust T
Torque Q
Total resistance RT
Thrust deduction DRP DRP = T - RT
Speed of advance of propeller VA

Propeller operates in a flow decelerated by hull. The effective inflow velocity of this
inflow, that is advance velocity V , is obtained from the wake factor w as follow
A

. (5.27)

The differences of propeller operation in open water and in the behind condition are
sketched in Figure 5.17.

Jerzy Matusiak
52/88

Figure 5.17 Propeller-hull interaction.

Propeller power and efficiency in open water are marked with a sub-index 0 and they
are

PT0 = T0 VA, PD0 = 2 ! n Q0 , "0 = T0 VA . (5.28)


2 ! n Q0

The same quantities for a propeller working at stern are

PT = T VA, PD = 2 ! n Q , "B = T VA . (5.29)


2!nQ

Propulsive efficiency is defined as a ratio effective and delivered power at propeller,


that is

!D = PE = RT V . (5.30)
PD 2 " n Q

Making use of the relations (5.26) and (5.27) the propulsive efficiency is obtained

!D = 1 " t T VA = !H !B = !H !0 !R . (5.31)
1"w 2#nQ

Propulsive efficiency is a very important quantity in ship propulsion. It can be


presented as a product of three efficiencies. The first one, that is hull efficiency h ,
H

tells how good is the selected propeller to operate behind the hull in question. It tells
also on the selected location of the propeller. For a beneficial propeller-hull
interaction hull efficiency has a value exceeding unity. This is often the case for a

Jerzy Matusiak
53/88

single screw vessel with a properly selected propeller. From the definition of hull
efficiency it is seen that it is beneficial to locate propeller in the region of decelerated
flow (wake). On the other hand propeller location should not lead to high acceleration
of hull flow velocities because this causes an increase of thrust deduction. Propeller
open water efficiency was discussed earlier. The so-called relative rotative efficiency
is a measure of propeller efficiency change caused by a non-uniformity of inflow at
ship stern. This factor as well can have a value slightly exceeding unity (by approx. 1-
2%).

Self-propulsion model test

Self-propulsion model test is the last test of the series aiming at evaluating the
required propulsive power of ship. Hull resistance and open water test of propeller
precede the self-propulsion test.

Test procedure

In self-propulsion test model is equipped with operating model propeller. Model


propeller is rotating and thrust and torque delivered by it are measured. Tests are
conducted, similarly as the resistance test, according to the scaling law of Froude,
which means that model velocity is

. (5.32)

As it was discussed when ship resistance was dealt with, because of large difference in
Reynold’s numbers, frictional resistance coefficient in model scale is much too high.
In order to have a proper loading of the propeller in self-propulsion test, the model is
towed with a force compensating for a difference in frictional resistance and taking
into account additional resistance coefficient of the ship. This force equals

!RTM = 0.5 " M V2M SM CFM # CFS + CA . (5.33)

Test analysis

Thrust deduction coefficient is obtained from the known model resistance and thrust
as follows

t = TS ! RTS = 1 ! RTM ! "RTM . (5.34)


TS TM

It is customary to assume that thrust deduction factor is the same for the ship and the
model.
Wake factor w is analyzed as follows.
Thrust measured in the self-propulsion test is made non-dimensional and
presented as

Jerzy Matusiak
54/88

TM .
KTM = (5.35)
! M n2M D4M

Using the propeller open water characteristics, the advance number corresponding to
this thrust coefficient is found as shown in Figure 5.18 and wake factor is calculated
from

wTM = 1 ! JTM DM nM . (5.36)


VM

10 KQ KT, !0

1.0

0.8

KQTM
KQM

0.4

KTM
0.2

0.2 0.4 0.6 JTM JQM 1.2 1.4


J = VA
nD

Figure 5.18 Using open water propeller characteristics when analyzing self-
propulsion test.

Using the so-called thrust identity principle, that is equating thrust coefficient (5.35)
to the one obtained in the open water test, the relative rotative efficiency is obtained as
a ration of torque coefficients
KQTM
!R = . (5.37)
KQM

Jerzy Matusiak
55/88

Extrapolating test results to ship scale

Normally it is assumed that wake fraction, thrust deduction factor and propeller open
water characteristics are same for the ship and for the model scale. The extrapolation
method called ITTC-78 attempts some additional scaling effects into account. In
particular wake factor and propeller characteristics are scaled by taking in a simplified
manner into account a difference in frictional effects. An older and more frequently
used method, that is ITTC-57 extrapolating method, is straightforward in presenting
the propulsion qualities for a ship. The formulae summarizing the method are
presented in Table 2.

Table 2 Model test scaling of the ITTC-57 method.

Quantity Symbol or furmula

Residual resistance CRM = CTM ! CFM

Total resistance of ship CTS = CFS + CRM + CA

Wake factor wTS = w TM = w

Propeller open water characteristics JTS = JTM = JT


KTS = KTM = KT
KQS = KQM = KQ

Propeller revolutions nS = nM/ !

Propulsive power PDS = 2 ! KQ "S D5 n3S

Propeller thrust TS = !S D4 n2S KT

Propeller torque QS = !S D5 n2S KQ

Effective power PES = 0.5 ! S V2S SS CTS

Propulsive efficiency !D = PES


PDS

Hull efficiency !H = 1 " t


1 " wT

Some research institutions apply extra corrections to the derived propeller revolutions
and propulsive power. The corrective factors are of an order of magnitude 1-2%. The
propulsive quantities obtained from model tests analyzed as presented above are for
the idealized, so-called tank conditions.
The installed power takes into account the lowering of ship performance caused
by winds, waves, aging deterioration and biological fouling. To secure the speed in

Jerzy Matusiak
56/88

actual environmental conditions and taking also ageing into account the installed
power is increased by a certain ratio when compared to the tank condition. This
additional power is called the sea margin. The component of sea margin divides into
two groups. One is a weather effect and the other is a deterioration effect, which
progresses with time.

3.8. On propeller selection and early design

This Chapter is based mainly on the book of Tornblad (1987).

There are no formulas based entirely on theory for direct calculation of the main
dimensions of a propeller, such as diameter, pitch and blade area. Formulas, empiric
or based on model tests, are available but they do not provide a general understanding
of the function of the propeller and are used only for preliminary estimates.

The main dimensions of a propeller are determined by means of model test results.
These results are presented, as described earlier, in the form of diagrams in which the
torque coefficient K the thrust coefficient K and the efficiency h are given as
Q T

functions of the advance number J, see Figure 5.12. A diagram suitable for designing
propellers must include results for a number of propeller models with different pitch
ratios but otherwise equal, see Figure 5.13. Results for propellers with different blade
area ratios are presented in separate diagrams.

Number of blades

The blade number is the first "dimension" to determine when starting the design of a
propeller. For small vessels the choice is normally 3 or 4 blades, for larger ones 4 or 5
-but there are cases with 3-b1aded propellers for high engine power, and 5-b1aded
ones for motorboats.

The larger the number of blades is, the less the risk of vibration -and the less the
efficiency, due to increased blade area in order to maintain the cavitation margin.

The optimum diameter decreases as well at increasing blade number. But in case the
same diameter is chosen for a 3-b1aded and a 4-b1aded alternative, the diameter of the
3-bladed design then being somewhat below optimum, the efficiency of the two
alternatives will be nearly the same.

In case there is a risk of resonant vibration in some part of the vessel with a certain
number of blades this should be avoided. The character of the wake field may also
lead to a larger risk of vibration with a certain blade number. In order to determine the
correct blade number it may be necessary to perform a model test with two similar

Jerzy Matusiak
57/88

propeller models, although with different blade numbers, in a cavitation tunnel with
the propeller running in a simulated wake field, at which test the pressure pulses
produced by the propellers are measured.

Optimum shaft speed

Let us first consider the academically correct method of designing a propeller. The
premises are that a given vessel shall be driven at a certain speed. The resistance R at
this speed must be overcome by the propeller producing the thrust T = R /(l-t). The
T

size of the vessel determines the maximum propeller diameter D -leaving a sufficient
clearance between blade tip and hull above the propeller, and avoiding the propeller
protruding below the keel.

Thus, V, T and D (= D ) are known, and assuming the wake fraction to be known as
max

well, also V = V(l-w) is known. The values to be calculated are a suitable shaft
A

speed n, required propeller power P and the pitch-diameter ratio P/D of the propeller.
D

Considering the definitions of thrust coefficient and advance number the uknown
revolutions n are eliminated yielding

KT/J2 = T n2 D2 = T tai KT = T J2 ,
2 4 2 2 2 2 2 (5.38)
! n D VA ! D VA ! D VA

The relation between K and J is second-degree polynomial (parabola), which can be


T

plotted in a diagram for a propeller with a suitable blade area ratio. The blade area
must for a start be based on a rough estimate and checked later on when the main
dimensions of the propeller have been calculated.

Jerzy Matusiak
58/88

Figure 5.19 Propeller diagram with K /J parabola.


T
2

At the intersection points of the parabola and the K curves for the different pitch
T

ratios P/D, J is read off. For these J values K and h are read off on their respective
Q

curves. For each J the shaft speed n and power P are calculated:
D

n = VA , PD = T VA ja Q = PD . (5.39)
JD ! 2"n

Jerzy Matusiak
59/88

The obtained values can be tabulated as

P/D J KQ h n [rps] n [RPM] PD [kW]

and plotted

Figure 5.20 Propeller revolutions and efficiency as a function of pitch ratio.

The maximum of efficiency h corresponds to the optimal revolutions n and the


max opt

pitch ratio (P/D) .opt

In case the delivered power is given and optimum propeller pitch ratio and revolutions
are sought torque and advance coefficients are used producing
PD PD
KQ/J3 = tai KQ = J3 , (5.40)
2 ! " D2 V3A 2 3
2 ! " D VA

Which when plotted on the propeller diagram

Jerzy Matusiak
60/88

Figure 5.21 Propeller diagram with K /J polynomial.


Q
3

Similarly as in the previous case optimum revolutions of the propeller are found from
the pitch ratio of propeller corresponding to maximum open water efficiency.

Optimum propeller diameter

In practice, one seldom has a free choice of shaft speed. Once the power is determined
one has to choose a standard engine which implies that the shaft speed is more or less
standard as well, at least with large low- speed direct drive diesels. On medium- speed
engines, however , the reduction gear presents a certain freedom of choice of propeller
shaft speed. Suppose that the shaft speed n is fixed, and so is the propeller power P '
D

the ship's speed V and the wake fraction w.

The optimum propeller diameter, rendering the maximum efficiency, is to be


calculated (of course, the diameter must not exceed the maximum possible with
respect to the space available).

Using again torque and advance ratio definitions the sought variable, that is diameter
D is reduced yielding K as a function of the fifth order polynomial in terms of
Q

advance ratio J

Q n5 D5 Q n3 Q n3
KQ/J5 = = tai KQ = J5 (5.41)
! n2 D5 V5A ! V5A ! V5A

Jerzy Matusiak
61/88

Proceeding similarly as in the previous cases, yields optimum propeller diameter at


pitch ratio corresponding to the maximum efficiency (see Figure 5.22).

Figure 5.22 Propeller diameter and efficiency as a function of pitch ratio.

The maximum value of h determines the optimum diameter for the given power, shaft
speed and ship's speed. It appears from the Figure 5.22 that the efficiency curve has a
very flat peak. That implies that a small reduction of the diameter below the optimum
value does not lead to any considerable drop in efficiency. In addition, the optimum
diameter of the propeller when fitted behind a vessel - the diameter rendering
optimum behind efficiency h - is somewhat smaller than the optimum one calculated
B

from a diagram, which is based on tests in homogeneous, parallel flow. Therefore, the
actual diameter chosen is normally about 5 per cent below optimum for single screw
vessels, and 2-3 per cent below for twin-screw vessel. The reduction in diameter is, of
course, connected to a corresponding increase in pitch and pitch ratio, as can be seen
in Figure 5.22. As a rule of thumb, the sum of diameter and pitch is constant for small
changes of diameter, D + P = const, - all other quantities as power, speed etc being
unchanged. If D is reduced by 3 per cent, P increases by 3 and P/D by 6 per cent. In
case the space in the propeller well is limited – the clearance between blade tips and
hull plating should be 25-30 per cent of the diameter - one may have to accept a
further reduction of the diameter. 10 per cent below optimum is acceptable, as an
exception down to 12-14 per cent. Once the optimum diameter has been calculated,
one would think that the shaft speed is optimum at the same time. However,
calculation shows that the optimum shaft speed calculated with the above optimum
diameter is lower than the initial shaft speed.

Again, the optimum diameter based on the new shaft speed is larger than the first
optimum diameter. Further calculations along the same line show the diameter to
continue increasing and the shaft speed decreasing. But there must be an end to this.

Jerzy Matusiak
62/88

Looking at the pitch ratio one finds that it is increasing all the time. In a K -n-J
Q

diagram the maximum efficiency increases for increasing pitch ratio, reaching an
overall maximum about P/D = 2.2 corresponding to a pitch angle of 45 degrees. In
practice, pitch ratio exceeding P/D = 1.4-1.5 is seldom used.

3.9. Propeller-main engine interaction

A proper propeller-main engine interaction means that the power required by the
propeller, in all operational conditions, falls into the region covered by the engine’s
nominal power-revolutions region. The matter is illustrated in Figure 5.23, where the
propeller power requirement as a function of shaft revolutions is shown. This curve
corresponds to a single operational case (single load condition and certain
environmental conditions). The relation is called propeller law. Ship speed is a
parameter on the curve with a maximum attainable speed being V . Higher speed can
5

not be achieved because of the limit in available power.

Figure 5.23 Power (teho) required by the propeller (potkurin) and that delivered by
the main engine (pääkone) as function of shaft revolutions (pyörimisnopeus).

Propeller law is approximately dependent upon the revolution to the power three.

Jerzy Matusiak
63/88

If propeller law does not fall into the delivered power-revolutions region, two cases
can be found. These are presented in Figure 5.24.

Figure 5.24 Too light and too heavy propellers.

Too light propeller, having too small diameter or too low pitch, attains maximum
revolutions of the engine without utilizing all power delivered by the engine. Too
heavy propeller, on the contrary, reaches the power limit at relatively low revolutions.
In this case propeller can’t either utilize engine’s delivered power.

Bollard pull and propeller – main engine interaction

In Figure 5.25 two different propeller law curves are presented. Thin line represents
‘free running’ operation mode while thicker lines corresponds to the bollard pull
condition.

Jerzy Matusiak
64/88

Figure 5.25 Diesel engine power – revolutions nominal region and Fixed pitch
propeller is designed for: a) ‘free running’ and for b) bollard pull condition.

In Figure 5.25a) propeller designed for ‘free running’ is too heavy at the bollard pull
condition. Propeller in Figure 5.25b), which was designed for a bollard pull operation,
is too light in the ‘free running’ condition.

Controllable pitch propeller and main engine

Ship operating at different conditions (free running in calm water, ice, head waves,
bollard pull) is subjected to varying resistance versus speed conditions. Fixed pitch
propeller, with a limited range of propeller shaft revolutions, results easily in a
situation where propeller is too heavy. In order to better utilize the power of the
machinery, controllable pitch propellers are often used. In Figure 5.26 a set of
propeller law curves is shown. Each of them corresponds to a certain pitch setting.

Jerzy Matusiak
65/88

Figure 5.26 Propeller law curves for a controllable pitch propeller.

Nominal propeller setting (P/D [%]=100) curve runs nicely through the nominal
power-revs region and meets the Maximum Continuous Rate (MCR) point utilizing
all power of the main engine. In this normal operational condition decreasing pitch
settings yields ‘too light propeller’. However, when accelerating the vessel or for the
increased resistance, due to head waves, ice etc., it is feasible to decrease the pitch.
Otherwise a propeller curve gets easily to heavy, propeller can’t utilize all power of
the engine and as a result ship does not accelerate sufficiently and does not achieve a
target speed.

3.10. Propeller cavitation, vibration excitation and noise

Types of propeller cavitation

Propeller cavitation can be divided into two types (Blake, 1986 ) - fixed and traveling
bubble cavitation. Sheet cavitation is more or less stationary with respect to the
propeller blade, glassy in appearance thin layer of gas and vapour mixture extending
from the leading edge to a certain chord-wise location (Figure 5.27).

Jerzy Matusiak
66/88

sheet cavitation

bubble
cavitation

tip-vortex
cavitation

Figure 5.27. Propeller cavitation types.

This type of cavitation is typical for hydrofoils and lifting surfaces. The mechanism of
sheet cavitation inception in the case of a hydrofoil is illustrated in Figure 5.28.

U, po
"

cp

-!

cavitation

Figure 5.28. Illustration of a cavitating hydrofoil.

This is a very typical cavitation type where low pressure develops on the upper
(suction) side of a hydrofoil, set in a flow of velocity U and an ambient pressure p0, at
an angle of attack a. If this pressure drops below the vapor pressure value pv then
cavitation occurs. In Figure 5.28 this is illustrated in terms of non-dimensional
quantities

(5.42)

and

. (5.43)

Jerzy Matusiak
67/88

The pressure coefficient cp is a non-dimensional measure of the pressure change pd


due to dynamic effects of fluid motion caused by the presence of the hydrofoil in the
flow. pd is called the dynamic pressure. Cavitation number s is a measure of the
ambient pressure closeness to the vapor pressure pv, which is critical for the cavitation
to occur. Fluid (water) density is marked by r . A definition of the pressures is shown
f

in Figure 5.29. Pressure p is the total pressure on the hydrofoil. In Figure 5.29 g is the
acceleration due to gravity, h is the depth of submergence and pa is atmospheric
pressure.

Figure 5.29. Definition of pressures.

Observed cavitation is fixed to the hydrofoil and extends over the suction side region
given approximately by an inequality relation cp ! "#. The modifying effect of the
cavitation on the pressure distribution is normally small and thus it is usually
neglected. Turbulence of the inflow or regular non-uniformity of it in a form of ship
wake cause a certain instability at the downstream end of sheet cavitation. A uniform
cavity film breaks off into free bubbles, which are shed downstream. This collapse of
the uniform cavitation sheet into a cavitation bubble cloud causes the erosion of
propeller blades and is associated with high-frequency noise generation.

Clearly distinguishable cavitation bubbles on propellers may occur as a result of two


flow phenomena. The first is associated with the instability of sheet cavitation as
described above. The second is associated with inflow turbulence. Figure 5.30, quoted
from the Blake (1986) illustrates this phenomenon.

Jerzy Matusiak
68/88

Figure 5.30. Illustration of unsteady bubble cavitation generation.

At high Reynolds numbers, so that the flow is turbulent, a randomly fluctuating


pressure is superimposed on a static component. The peak values of total pressure fall
randomly both in time and space below the critical vapor pressure value. These fast
oscillations of pressure drive the growth and collapse of the micro-bubbles (air nuclei)
contained by the inflow. As a result a significant erosion of the blades and noise
develop.

Another situation at which bubble cavitation may develop is a controllable pitch


propeller set to a low pitch value and rotating at high speed. The flow may locally, i.e.
at certain blade positions, reach negative values of angle of attack, resulting in
pressure side cavitation of the bubble type.

Cavitation and propeller performance

The effect of cavitation on propeller performance is sketched in Figure 5.31.

Jerzy Matusiak
69/88

Figure 5.31 The effect of cavitation on propeller performance.

Decreasing cavitation number yields cavitation on propeller blades, especially at low


advance ratio values that is at high propeller loading. At the beginning, this does not
have any visible effect on propeller thrust and torque. When sheet cavitation covers a
substantial part of blade suction side, thrust starts to decrease. Further decrease of
cavitation number is followed by a decrease in torque. As a decrease of thrust is faster,
efficiency of propeller decreases as well.

A simple mean of decreasing propeller cavitation is to increase blade area. In this way
load is distributed over a larger blade area and a drop in dynamic pressure is
decreased. There are simple empirical formulas giving a minimum blade area, which
ensures that there will be no serious problems with cavitation. One of the formulas is
the one of Tornblad (1987)
AE = 1.3 + 0.3 Z T + k
A0 (5.44)
p0 ! pv D2

where Z is blade number, hydrostatic pressure at the shaft immersion h is p =rgh and
0 0 0

k is

Vessel type k

Fast navy vessel 0


Twin screw ship 0.1
Single screw ship 0.2

When applying formula 44 the SI units should be used.

Jerzy Matusiak
70/88

Propeller induced pressures and noise

A marine screw propeller located at the stern of a ship operates in a flow affected by
the hull. The ship wake, which is a decelerated, non-uniform inflow into the propeller
disc, has certain negative effects on its operation. Apart from the desired steady thrust
component developed by the propeller, an unwanted unsteady loading is experienced
by propeller and ship. These periodic loads, called "bearing forces", are transferred
through the shaft and the bearings to the ship structure (see Figure 5.32).

Figure 5.32 Propeller induced loading at ship stern (Matusiak, 1993).

The magnitude of these unsteady loads depends mainly on the non-uniformity of the
wake and usually does not exceed 5% of the steady components, i.e. thrust and torque.
These concentrated propeller forces and moments can cause vibration of the main
propulsion machinery and shafting system, but are not important, in general, as a
source of hull vibratory excitation (Matusiak, 1992).

A rotating propeller induces an unsteady pressure field which affects the submerged
part of the hull as "surface pressures and forces". In general, propellers of surface
vessels are designed to carry the loading at which cavitation is unavoidable. The
occurrence and extent of this cavitation depends on the angular position of the blades.
The main reasons for cavitation unsteadiness are a non-homogeneous inflow into the
propeller disc (ship wake – see Figure 5.33), resulting in unsteadiness of blade
loading, and hydrostatic pressure variation experienced by the rotating blades.

This is unsteady cavitation that usually the main cause of ship vibration problems. The
pressure induced by unsteady cavitation decreases much slower with distance than the
contributions of the non-cavitating propeller. Moreover, this pressure also comprises
high-frequency components and thus it is the main source of the underwater noise
generated by vessels. The primary source of high-frequency noise is the collapse of
free cavitation bubbles. This noise contributes to crew discomfort, may interfere with
the sonar systems of a research vessel and may promote the detection of a naval vessel
by enemy sensors (Matusiak, 1992).

Jerzy Matusiak
71/88

Figure 5.33 Sketches of ship wake. Deceleration of longitudinal, that is x-directional,


velocity is shown only. On the left symmetric wake of a single screw vessel and on the
right a typical wake of a twin-screw vessel is sketched. (Matusiak, 1993).

Cavitation types and broadband pressure spectrum

Power spectral density of a pressure induced by a marine screw propeller is shown in


Figure 5.34. The spectrum starts with a blade frequency, which is propeller shaft
revolutions times the number of blades, and extends to tens of kilohertz. At the top of
the figure already discussed types of propeller cavitation are shown. Four small
profiles, with adjacent to them formulas, represent different contributors to the
propeller induced pressures.

Non-cavitating propeller induces pressure, which is represented by two upper profiles.


This pressure is mainly limited to the blade frequency component and it is caused by
the co-called displacement effect and pressure loading of the rotating propeller blades.
This pressure decays rapidly with a distance to power three.

Two lower profiles represent the effect of cavitation on the induced pressure. The
effect of sheet cavitation varying with a blade position is shown on the left. Sheet
cavitation area (volume when integrated over the blade span) induces the pressure
displacing in a similar way as a volume of blade material- 1 term in the formula p .
st
c

This effect is again restricted to a vicinity of propeller as it is decaying with a distance


to power three (1/R ). Time variations of sheet cavity volume (term
3
) is
usually more significant as its effect decreases in space much slower, as 1/R. The
effect of sheet cavitation is visible as pressure with a frequency content starting from
the blade frequency and extending to hundreds of Hertz.

Jerzy Matusiak
72/88

Figure 5.34 Contributors to the propeller induced pressure (Matusiak, 1993).

The profile on the right depicts the effect of bubble cavitation on the pressure. The
rapid collapse (rapid volume change) of small-size cavitation bubbles is primarily
responsible for propeller-induced high-frequency noise. The bubble diameter is small
compared to its distance from the rigid boundaries. The pressure peak induced by the
rebounding bubble, that is a bubble subjected to an increase of ambient pressure, is
extremely short. This and the fact that there is a large number of bubbles of different
sizes, is responsible for a high frequency content of the pressure caused by a cloud of
cavitation bubbles (Matusiak, 1993).

Jerzy Matusiak
73/88

Judging the propeller induced pressures

Usually the biggest concern when judging the propeller-induced pressure is ship
vibration. In particular low frequency components of excitation are of interest. For
this reason model tests or theoretical investigations dedicated to the evaluation of
propeller-induced pressures are mainly concerned with a blade frequency component.
Dedicated model tests are conducted in the so-called cavitation tunnels (see Figure
5.35) or in the under-pressurized towing tank of the Maritime Research Institutions
Netherlands (Marin).

Figure 5.35 Model of a propeller in the measuring section of Taiwanese cavitation


tunnel.

In Table 2 there are the values of blade frequency amplitutdes given with grading.
This judgment is based on the sparse and somewhat outdated information provided by
literature and personal experience of the author.

Table 2 Judgment of the amplitude of blade frequency component of the


propeller-induced pressure.
Pressure amplitude [kPa] Grading
0 - 1.2 Very good
1.2 - 2.5 good
2.5 - 4.0 satisfactory
4.0 - 8.0 bad
8.0 - Very bad

Jerzy Matusiak
74/88

What affects unsteady propeller cavitation and induced pressure

The main reason for the unsteadiness of cavitation and most serious cause of the
propeller-induced pressures is ship’s wake. Propeller operating with a varying blade
inflow experiences unsteady loading and intermittent cavitation. In order to smoothen
growth and decrease of sheet cavitation two measures can be used.

The first and most efficient one is to form ship stern so that wake is as much as
possible axially symmetric (Figure 5.36 – left). This secures small variations of the
inflow into the sections of propeller blades and results in a less rapid cavitation
growth and collapse compared to sharper wake shown on the right of the same figure.

Figure 5.36 Axial wake of stern bulb form on the left and V-stern on the right.

Second measure aimed at decreasing propeller cavitation induced vibration excitation


is associated with propeller design. Moderately or highly skewed propellers are
known to be generating less hull vibration. The effect of skew can be explained with
an aid of Figure 5.37.

Figure 5.37 Effect of blade skew on the intermittent sheet cavitation (Matusiak,
1993).

Jerzy Matusiak
75/88

For the propeller shown on the left the entire length of blade with no skew enters the
peak region of wake at the same time. Sheet cavitation grows at the same time at
different blade sections. As a result, sheet cavitation volume variations are rapid and
induce high pressures at stern of a ship. Skewed blades (propeller on the right)
smoothen the entrance of blade into the wake peak and decrease speed of cavitation
growth and decrease.

Model and full-scale observations and measurements confirm a better performance in


terms of vibration of the skewed propellers. Usually skew is not the only parameter
introduced to improve propeller cavitation performance. The so-called ‘tip off-
loading’ is another measure. This means a decrease pitch at the blade tip region. As a
result loading at a tip decreases, tip-vortex cavitation is reduced and an overall
cavitation performance of propeller improves as well. Quite often both measures, that
is introducing skew to blades and ‘tip-offloading’, are utilized together.

3.11. Special types of propulsors

There are many ship propulsors that differ from a normal marine screw propeller. All
of them we may call as special types of propulsors.

Ducted propellers

Ducted propeller, called also propeller in nozzle, consists of a normal screw propeller,
which can be both fixed- or controllable pitch propeller, and an axially symmetric
nozzle encircling the propeller. There are two types of ducts. The first and most
common one is which uses flow-accelerating nozzle. Less frequently used duct is the
one equipped with a nozzle decelerating the flow. Example of a ducted propeller is
shown in Figure 5.38.

Jerzy Matusiak
76/88

Figure 5.38 Ducted propeller.

Flow accelerating nozzle increases the flow velocity of the flow entering the propeller
disk, especially at low advance numbers. It also smoothens non-homogeneity of the
inflow. As a result thrust of the propeller-nozzle combination increase, especially at
low advance speeds. That is why ducted propellers are frequently used in tugs and in
ice-breaking vessels despite of a danger of nozzle being locked by ice blocks.

The nozzle propeller is less sensitive to load variations than the open propeller. This is
also seen in a form of a relatively flat torque, thrust and efficiency curves (Figure
5.39).

Jerzy Matusiak
77/88

Figure 5.39 Open water characteristics of the ducted propeller Ka 4-70.

At low advance numbers, nozzle contribution to the total thrust (denoted by KTn) is
high. At high advance speeds nozzle has a breaking effect on total thrust.

Z-drive units and podded propellers

Z-drive propulsion unit consists at least of a marine screw propeller, sometimes in a


nozzle, supporting strut and a pod (see Figure 5.40).

Jerzy Matusiak
78/88

Figure 5.40 Z-drive propulsion unit of IB Fennica. D =4.2 m, P = 7.5 MW.


D

In 1980’s a new rotatable propulsion unit type was developed having primarily in
mind ice-going vessels. The concept, named Azipod, was developed jointly by
Strömberg Ltd (ABB since 1988) and Wärtsilä Marine (presently STX Europe). In
this propulsion unit type, an electric motor is mounted in a pod situated ahead of the
propeller. One of the first Azipod installations is shown in Figure 5.41. The units have
developed, both in hydrodynamic and electro engineering sense, within the last 20
years. In Figure 5.42 three units installed at the stern of Freedom class ship are shown.
It is worth noting that while the side units are rotatable and of the pulling type
(propellers are located ahead of the pods) and the centre unit is fixed with a pushing
propeller.

Jerzy Matusiak
79/88

Figure 5.41 Azipod propulsion unit installed at stern of MS Uikku. D = 5.6 m, PD


= 11.4 MW.

Figure 5.42 Azipod propulsion units installed at the stern of Freedom class
passenger vessel. PD = 3*14 MW.

There are several benefits associated with using the rotatable thrusters, sometimes
called simply as pods or podded propulsion

Jerzy Matusiak
80/88

Propeller shafts, supporting struts, bossing and tunnels of stern steering thrusters
contribute to resistance of ship hull by approximately 10-15%. Refer to Figure 5.43
for a normal stern arrangement with propeller shaft and brackets.

Figure 5.43 Stern arrangements of a typical RoPax vessel.

Although the open water efficiency of podded propulsion is somewhat lower than the
one of a normal screw propeller, the gain in reduced resistance means that usually
pods require less power than a normal twin-screw propulsion arrangement. Vibration
and noise problems may be also reduced using pods. Especially pulling propeller
operating in an inflow not affected by shaft and brackets is a good choice in this
respect. The turning capabilities of rotatable propellers are very good. However, open
pram type stern with no skeg and equipped with turning thrusters may easily result in
directionally unstable ship.

Counter-rotating propellers

Contra-rotating propellers (CRP) have been known since 1836. They are commonly
used in torpedoes and are often met in fast motorboats. Although their good
hydrodynamic properties were well recognized they did not gain much popularity as
ship propulsors because of the mechanical complexity associated with long shafts,
their bearings and sealing. However, the CRP concept was successfully used in some
merchant ships, mainly in Japan. Power savings as high as 16% have been reported
for a ship retrofitted with a mechanically driven CRP propulsor.

It was the introduction of the azimuthing thruster as the propulsion unit that brought
the concept of the contra-rotating propeller back into the daylight. This started with
the mechanically driven units in the 1980s. What was mechanically complex for fixed
propellers was no longer difficult to implement in the azimuthing thruster. In Figure
5.44 mechanically driven CRP rotatable propulsion unit is shown.

Jerzy Matusiak
81/88

Figure 5.44 Mechanically driven CRP rotatable propulsion unit.

The basic idea behind the contra-rotating propeller arrangement is to recover the
slipstream rotational energy of the forward propeller. The hydrodynamic axial losses
of the CRP are similar to those of a traditional single propeller arrangement. However,
thanks to lower propeller loading, propeller diameter can be increased as the pressure
pulse level will go down. This will have a beneficial effect on the axial losses of the
CRP and yield an increased propulsive efficiency. Dividing loading between two
propeller units yields additional benefits in the propulsor’s efficiency. The blade area
of the propellers can be lower when compared with a single propeller arrangement.
The aspect ratio of the blades increases and as a result the propeller efficiency
increases. Lower blade loading also results in improved resistance to cavitation.

Hybrid propulsors

Hybrid propulsion means that at least two propulsion units of different types are used
to propel ship. Interesting newcomer to this category is a combination of a fixed
propeller and turnable podded propulsor forming the CRP arrangement (refer to
Figure 5.45).

Jerzy Matusiak
82/88

Figure 5.45 Model of a ship equipped with hybrid propulsion (Konsin, 2003)

The unit combines all good features of counter rotating propeller and these of the
rotatable thruster. The hybrid concept is especially attractive when the main
propulsion is very heavily loaded due to the large ship speed or power requirements of
the vessel. Adding a pod to the main propulsion will raise the overall speed and
power. The design of both units requires special attention for optimisation of
cavitation and noise (Matusiak, 2004).

Water-jets

Water-jet is an old propulsion type. It was patented in England already in 1661. The
unit consists of an impeller, which imparts momentum on (accelerates) the flow.
Water enters the unit through the bottom opening and leaves through a nozzle (Figure
5.46).

Figure 5.46 Water-jet propulsion unit.

Jerzy Matusiak
83/88

Discharged water-jet impacts water and creates reaction force that is thrust. Nozzle
can turned along the vertical axis. This makes it possible to steer the vessel by
changing the orientation of thrust in horizontal plane. The so-called reversing bucket
(not shown in the Figure) turns water-jet downwards and forward making it possible
to stop the vessel or to run astern.

The hydrodynamic efficiency of water-jet propulsion can be estimated similarly as it


was done in case of ideal propulsor. Thrust of the unit stems from the momentum
equation

T = m Vj ! V , (5.45)

where mass flux m = ! Vj An with A being nozzle cross-section area and V water-jet
n j

velocity. We apply Bernoulli equation separately for the flow upstream (station 1, at
ship bottom level in front of the bottom opening) and before the impeller disc (1 st

Equation of 5.46) and for the flow leaving the impeller and extending downstream as
water-jet – station 2 (2n Equation of 5.46). The pressures at the considered stations
are:

At station 1; upstream p1 = pa + ! g (h " hj )

In front of the impeller pe

Just downstream of the impeller pt = pe + !pP

Downstream that is at station 2 pa

The Bernoulli equations are

pa + ! g (h " hj ) + 1 ! V2 + ! g 0 = pe + 1 ! V2P + ! g h + #pL


2 2 (5.46)
pa + ! Vj + ! g h = pe +$pP + 1 ! V2P + ! g h .
1 2
2 2

where V is flow velocity at the impeller and !pL = 1 " k V2 are pressure losses, which
P
2
primarily are associated with viscosity effects. These effects are represented by a non-
dimensional factor k. The pressure jump at the impeller is

!pp = 1 " V2j # V2 + " g hj + 1 " k V2 . (5.47)


2 2

The energy flux delivered from the pump to the water-jet is

Pp = Vj An !pp = Vj An 1 " V2j # V2 + " g hj + 1 " k V2 . (5.48)


2 2

The efficiency of water-jet can be obtained by dividing the work done by it in unit
time by power P , that is
P

Jerzy Matusiak
84/88

" Vj An Vj # V V
! = TV =
Pp " Vj An 1 V2 # V2 + g hj + 1 k V2
2 j 2 (5.49)
2 Vj # V V
= .
V2j # V2 + 2 g hj + k V2

If, for the sake of simplicity, we disregard the height h the efficiency can be presented
j

as a function of velocity ratio V /V as shown in Figure 5.47.


j

1
0,9 k=0
!
0,8
k = 0,1
0,7
0,6
0,5
0,4
k = 0,5 k=1
0,3
0,2
0,1
0
1 1,5 2 2,5 3 3,5 4
Vj /V

Figure 5.47 Efficiency of the water-jet as a function of relative velocity.

It is clearly seen, that similarly as in the case of ideal propulsor, the efficiency is at the
maximum (h=1) for water-jet velocity equal to ship speed. However, in this situation
there is no thrust (refer to 5.45). In order to have efficiency sufficiently high and thrust
as required, cross-section of nozzle has to be sufficiently large. Bottom opening an
inlet have to be designed so that there is no separation of flow at intake and no
cavitation. When deriving efficiency we have disregarded hull and pump efficiencies.
These have to be included when detailed evaluation of delivered power is done.
However, this requires dedicated model tests, which are not elaborated in this course.
The method presented above can be used when a fast evaluation of the required power
is needed. Application of the method is illustrated in Figure 5.48. Ship resistance R is
plotted as a function of speed along with the thrust T of water-jet for different power
values. The points where resistance meets the thrust curves are the self-propulsion
points. The shadowed area is the heavy cavitation zone, which should be avoided.

Jerzy Matusiak
85/88

2000
R [kN] R [kN]
P = 30 MW
T [kN]
1500 P = 20 MW
pumpun kavitaatioalue P = 10 MW

1000

500

0
0 5 10 15 20 V [m/s] 25

Figure 48 Water-jet thrust T as a function of ship speed and delivered power.


Cross-section area of nozzle A = 5 m . Coefficient of flow losses K=0.3.
n
2

Jerzy Matusiak
86/88

4. REFERENCES
Biran, A.B. 2003 Ship Hydrostatics and Stability. Butterworth-Heinemann 2003,
ISBN 0 75064988 7
Blake, W. K. 1986 Mechanics of flow-induced sound and vibration. Academic Press ,
Inc., 974 p.

Brix, J. 1993 Manoeuvring Technical Manual, Seehafen Verlag GmbH, Hamburg

Clayton B.R. and Bishop R.E.D, 1982 Mechanics Of Marine Vehicles, London, ISBN
0 419 12110-2.
Dudziak, J. 1988 Ship theory, Gdansk 1988, Wydawnictwo Morskie Gdansk, in
Polish, pp 619.
Faltinsen, O. M. 1990 Sea Loads on Ships and Offshore Structures, Cambridge
University Press, 328 p.

Fossen, T.,I. 1994 Guidance And Control Of Ocean Vehicles, J. Wiley&Sons, ISBN
0 471 94113 1.

Garrison, C.J. 1974 Dynamic response of floating bodies. Houston. Proceedings in


Offshore Technology Conference, paper No. 2067, pp. 365 – 377.

Hämäläinen, R., van Heerd, J. (1998) Hydrodynamic development for a large fast
monohull passenger ferry, Transactions of SNAME, vol. 106,
pp. 413 – 441.

Harvald, S.A. (1983) Resistance and Propulsion of Ships. New York, John Wiley &
Sons, Inc. 353 s.

Holtrop, J. (1977) A statistical analysis of performance test results. Interntional


Shipbuilding Progress vol. 24, February 1977

Holtrop, J. (1984) A statistical re-analysis of resistance and propulsion data.


Interntional Shipbuilding Progress vol. 31, 1984.

Huuska, O. (1976) On the Evaluation of Underkeel Clearances in Finnish Waterways.


Helsinki University of Technology, Ship Hydrodynamics
Laboratory, Otaniemi, Report No. 9. 128 p

ITTC, 1999 The Specialist Committee on Stability, Final Report and


Recommendations to the 22 ITTC, Seoul and Shanghai 1999
nd

Journee J. M. Strip Theory Algorithms, report MEMT 24, Delft University of


Technology, Ship Hydrodynamics Laboratory (1992).

Jerzy Matusiak
87/88

Konsin, J. 2003 An experimental study of forces on a pod propulsor in hybrid


propulsion. Master Thesis. Helsinki Universituy of Technology. Mechanical
Engineering Faculty. In Finnish.

Kreyszig, E. 1993 Advanced Engineering Mathematics, John Wiley & Sons, Inc. 7th
edition
Kuiper, G. 1992 The Wageningen Propeller Series. Marin Publication 92-001. May
1992. Pp. 110.

Lloyd, A.R.J.M 1989 Seakeeping: Ship Behaviour in Rough Weather, Ellis Horwood
Ltd.

Lönnberg, Lars (2000) Computation of hull flow of cruise-ship using FINFLO-


software, Helsinki University of Technology 2000 (Master
thesis).
Matusiak, J. 1992 Pressure and noise induced by a cavitating marine screw propeller.
Espoo 1992. VTT Publications. 80 p. + app. 22 p.
Matusiak, J. 1993 Ship Propulsion. Helsinki University of Technology. Ship
Laboratory. Report M-176. Otaniemi 1993. (in Finnish)

Matusiak, J. 1995 Ship Buoyancy and Stability. Otatieto Oy. Helsinki 1995. (textbook
in Finnish), ISBN 951-672-293-8

Matusiak, J. 2000 Dynamics of cargo shift onboard a ship in irregular beam waves,
International Shipbuilding Progress, 47, No 449 pp. 77-93
Matusiak, J. 2004 Hydrodynamic Aspects of Hybrid Propulsors using the Contra-
Rotating Propeller Concept. Wärtsilä Marine News, 2004. February.

Matusiak, J. 2007 Ship Dynamics. Helsinki University of Technology. Ship


Laboratory. Unpublished lecture-notes.

Matusiak, J. 2008 Ship resistance. Helsinki University of Technology. Ship


laboratory, report M-289, 3rd edition, in Finnish.

Naito, S. 1995, Generation and absorption of waves. Symposium on Wave


Generation, Analysis and Related Problems in Experimental Tanks-
Especially on Directional Wave. Yokohama National University, 25
September 1995. Pp. 1-27.

Niittymäki, J. 2000 Effects of the stern form on hydrodynamic properties of a fast


ROPAX- vessel. Helsinki University of Technology 2000
(Master thesis).

Nuotio, N. 1995 Relation between the wave profile and the wave-making resistance.
Helsinki University of Technology 1995 (Master thesis).

Ohkusu, M. editor 1996 Advances in Marine Hydrodynamics, Computational


Mechanics Publications, Southampton Boston

Ruponen, P. 2007 Progressive Flooding of a Damaged Passenger Ship. Department of


Mechanical Engineering. Doctor Thesis. ISBN 978-951-22-9012-3

Jerzy Matusiak
88/88

Schweighofer, J. & Hellsten, A. 1999 Computation of viscous flow around the


HSVA-1 tanker using two versions of the k-e turbulence
model, Helsinki University of Technology, Laboratory of
Aerodynamics, Series-B, Report B-51.
Tornblad, J. 1987 Marine Propellers and Propulsion of Ships. Kristinehamn 1987,
Marine Laboratory KaMeWa AB. 238 s.

Triantafyllou Michael S. & Hover Franz S. 2003 Maneuvering and control of marine
vehicles, Department of Ocean Engineering, Massachusetts Institute
of Technology, Cambridge, Massachusetts USA

Tuck, E.O., Taylor P.J. 1970 Shallow water problems in ship hydrodynamics.
Pasadena 1970, 8th SNH.

Jerzy Matusiak

S-ar putea să vă placă și