Vous êtes sur la page 1sur 367

Resistance and Propulsion of Ships

Technical University Delft


Course MT512

Prof. Dr. Ir. G.Kuiperl

January 3, 1994
Re-v w7. @L%-74K

1MARIN, Maritime Research Institute Netherlands, Wageningen, Technical


University Delft
Contents

1 Hull forms 13
1.1 Displacement Hulls. 13
1.1.1 Efficiency 13
1.1.2 Typical Speeds 14
1.1.3 Hull Forms . 14
1.1.4 Form Parameters. 14
1.1.5 .Considerations for the Stern Form 16
1.1.6 COnsiderations for the Bow Form 16
1.1.7 Bulbs. 17
1.2 High Speed Ships. 18
1.2.1 Planing Hulls 19
1.2.2 Hydrofoils. 21
1.3 Air as Carrier. 24
1.3.1 Air Cushion Vehicles. 24
1.3.2 Surface Effect Ships. 25
1.4 Multi Hulls 25
1.4.1 Catamarans. 25
1.4.2 Swath 26

2 Propulsors 29
2.1 Propellers 30
2.1.1 Propeller Arrangements 30
2.1.2 Trusters 31
2.1.3 Controllable Pitch Propellers. 32
2.1.4 Overlapping Propellers. 34
2.1.5 Contra Rotating Propellers. 34
2.1.6 Surface Piercing Propellers. 34
2.2 Special Types of Propellers. 35
2.2.1 Supercavitating Propellers. 35
2.2.2 Agouti Propellers. 35
2.2.3 Tipplates . 36
2.2.4 Vane Wheels. 37
2.3 Ducted Propellers. 37

1
2

2.3.1 Ringpropellers. 40
2.3.2 Mitsui Duct. 40
2.4 Other Propulsors 41
2.4.1 Voight-Shneider Propellers 41
2.4.2 Paddle-Wheels. 43
2.4.3 Pump Jets. 43
2.4.4 Sails 45
2.4.5 Other Types of Propulsion. 47

3 Intermezzo: Resistance of Simple Bodies 51


3.1 Non-dimensional Coefficients. 51
3.2 Drag of a Flat Plate . 52
3.3 Boundary Layer Flow. 53
3.3.1 Laminar and Turbulent Flows. 54
3.3.2 Effects of the Pressure Gradient. 55
3.4 Drag of a Two-dimensional Cylinder.. 56
3.5 Drag Components. 59
3.6 Additional References. 59
3.7 Additional Data 60
3.8 Summary.. 61

4 Resistance, Wake and Wake Distribution 63


4.1 Resistance and Wake. 63
4.2 Flow along a Ship Hull. 65
4.3 Cross Flow. 65
4.4 Separation. 67
4.5 The Wake behind Simple Ship like Bodies . 69
4.6 Horse-Shoe Vortices. 71
4.7 Visualisation of the Flow around the Hull. 72
4.8 Ship Wake. 73
4.8.1 Representation of the wake 75
4.8.2 Relation between hull form and wake distribution 76
4.8.3 Wake Fraction 78
4.9 Design Considerations . 79

5 Wave Resistance 81
5.1 SurfaCe Waves. 81
5.2 Properties of Surface Waves 82
5.2.1 The Dispersion Relation 82
5.2.2 Energy in a Wave. 82
5.2.3 The Group Velocity. 83
5.3 The Kelvin Wave System. 83
5.3.1 The Froude Number 85
3

5.3.2Resistance due to a Kelvin Wave System 85


5.3.3The Wave System of a Ship 86
5.4 Wave Interference. 89
5.4.1 A Two-dimensional Simplified Hull Form. 90
5.5 Economical Speed. 91
5.6 Hull Speed 93
5.6.1 High Speed Ships. 93
5.7 Bulbous Bows. 94
5.8 Shallow Water Depth. 94

6 Intermezzo: Scaling Rules 97


6.1 Dimension Analysis. 98
6.2 Physical Meaning of Non-dimensional Parameters 101
6.3 Scaling Rules 102
6.4 Scale Effects. 102

7 Resistance Prediction using model tests 103


7.1 Elements of Ship Resistance 103
7.2 Scaling Laws for Model Tests. 104
7.3 Froudes Hypothesis. 106
7.4 Determination of Resistance Components 106
7.4.1 Determination of the Frictional Resistance . 108
7.4.2 Determination of the Form Resistance. 109
7.4.3 Determination of the Wave Resistance 112
7.5 Extrapolation of Resistance Tests 112
7.5.1 Froude's Extrapolation Method 113
7.6 Effects of Surface Roughness. 113
7.6.1 Equivalent Sand Roughness 115
7.7 Appendage Drag 115
7.8 Effective Power 116
7.9 Effects of Laminar Flow 116
7.10 Wake Scale Effects 116
7.11 Example of Resistance Extrapolation 118

8 Resistance Prediction using Statistical or Systematic Data123


8.1 General Considerations for Hull Design. 123
8.2 Systematic Series 126
8.3 Regression of Available Data. 128
8.4 Design of Curve of Sectional Areas by Lap 131
8.5 The method of Holtrop and Mennen. 132
8.6 Example of Resistance Prediction. 135
8.7 Resistance of Small Vessels. 137
4

9 Intermezzo:Equations of Motion 139


9.1 The Continuity Equation. 140
9.2 The Equations of Motion. 141
9.2.1 Rotation and Deformation. 143
9.2.2 Relation between Stresses and Strain. 144
9.2.3 Navier-Stokes Equations 145
9.3 A Simple Example. 146
9.4 The Euler Equations. 148
9.5 The Bernoulli Equation 148
9.6 Summary.. 149

10 Intermezzo:Potential flow 151


10.1 Singularities in Potential Flow. 152
10.1.1 Uniform Flow. 152
10.1.2 Source. 152
10.1.3 Vortex. 153
10.1.4 Dipole. 155
10.2 A Simple Example of Potential Flow. 156
10.3 Forces on a Vortex 158
10.4 Panel Methods . 159
10.4.1 The Lifting Problem 161
10.5 Summary.. 163

11 Intermezzo: Boundary Layers 165


11.1 The non-dimensional Navier Stokes Equations 166
11.2 The Boundary Layer Equation 166
11.2.1 Scaling the Thickness of the Boundary Layer. . . 168
11.3 Solutions of the Boundary Layer Equations: Blasius. . . 168
11.4 Turbulence. 169

12 Flow Calculations without Waves 173


12.1 Potential Flow Calculations 175
12.1.1 Panel Methods without Free Surface . 175
12.1.2 Assessment of Various Bulb Designs. 176
12.1.3 Knuckles and Bulge Keels 178
12.1.4 Assessment of the Afterbody. 179
12.2 Navier-Stokes Solutions. 184

13 Flow Calculations with a Free Surface 187


13.1 The Linearized Free Surface Condition 187
13.2 Kelvin Sources. 190
13.3 Applications of the Kelvin Sources. 190
13.3.1 The Michell Theory. 190
5

13.4 Kelvin Sources for Catamaran Hulls, an Example 193


13.5 Dawson's Method. 194
13.5.1 Applications of Dawson's Method. 196
13.6 General_Considerations to Assess Programs. 198

14 Axial Momentum Theory 201


14.1 Axial Mc;Mentum Theory. 201
14.1.1 Efficiency 204
14.2 Optimum Radial Loading Distribution 205

15 The Propeller Geometry 208


15.1 General Outline. 208
15.2 Blade Sections. 209
15.2.1 NACA Definition of Thickness and Camber 212
15.2.2 Root and Tip. 213
15.3 Pitch and Pitch Angle 213
15.4 Propeller Plane and Propeller Reference Line 214
15.5 Rake. 215
15.6 Skew. 216
15.7 Blade Contours and Areas 216
15.8 Warped Propellers . 219
15.9 The Propeller Drawing. 219
15.10Description of a Propeller 220
15.11 Controllable Pitch Propellers. 222

16 Systematic Propeller Series 224


16.1 Open Water Diagram 224
16.2 The Quality Index . 227
16.3 Systematic Propeller Series. 227
16.4 Propeller Hull Interaction 229
16.5 Propeller Design Requirements. 231
16.6 Choice of Number of Blades and Blade Area Ratio. 232
16.7 Propeller Design using B-Series Charts 234
16.8 Elimination of Variables 235
16.8.1 Known Power and Diameter. 236
16.8.2 Known Power and Rotation Rate 237
16.8.3 Known Thrust and Diameter 237
16.8.4 Known Thrust and Rotation Rate 238
16.9 Optimization using the Open Water Diagrams. 239
16.10Example. 241
16.11Four Quadrant Measurements 243
16.12Propeller Design using the Optimized Data. 246
16.130ther Series. 246
6

17 Profile Characteristics 247


17.1 The Pressure Distribution 247
17.2 The Loading Distribution. 249
17.2.1 The Lift Curve 250
17.3 The Zero Lift Angle. 251
17.4 The Leading Edp Suction Peak 252
17.4.1 The Ideal Angle of Attack 252
17.4.2 Profile Drag. 254
17.5 Profile Series. 256
17.5.1 Thickness Distributions. 257
17.5.2 Camber Distributions. 257
17.5.3 Derivation of the Local Pressure of a Profile 260
17.5.4 Considerations to Choose or Design a Profile. 261

18 Cavitation 265
18.1 The Cavitation Number 266
18.2 Types of Cavitation. 266
18.2.1 Bubble Cavitation 266
18.2.2 Sheet Cavitation 267
18.2.3 Root Cavitation. 268
18.2.4 Tip Vortex Cavitation 268
18.2.5 Propeller Hull Vortex Cavitation . 269
18.2.6 Unsteady Sheet Cavitation. 269
18.2.7 The Mechanism of the Development of Cloud Cavitation271
18.3 Noise and Erosion. 274
18.3.1 The Implosion of a Single Bubble Cavity 275
18.3.2 Noise Radiation. 976
18.3.3 Thrust Breakdown 978
18.4 The Cavitation Bucket. 980
18.5 Cavitation Tests. 284

19 Lifting Line Propeller Design 287


19.1 Lifting Line Theory. 289
19.1.1 Two-dimensional Lifting Lines 289
19.1.2 Lifting Lines in Three Dimensions. 289
19.1.3 Lifting Line Theory for a Propeller 292
19.2 Optimum Radial Loading Distribution 993
19.3 Induction Factors. 994
19.4 Propeller Design using the Induction Factors. 295
19.4.1 Determination of the Inflow 295
19.4.2 Determination of the Blade Sections. 296
19.4.3 Strength of the Propeller. 299
19.4.4 Stresses due to Loading. 299
7

19.4.5 Stresses due to Centrifugal Forces. 300


19.4.6 Approximate Methods 301
19.4.7 Lifting Surface Corrections. 301
19.4.8 Viscous Forces. 308

20 The Propulsion Test 311


20.1 The Additional Towing Force 311
20.1.1 Self Propulsion Test with an Additional Towing Force. 312
20.2 Overload Tests. 313
20.3 Scaling Laws. 313
20.3.1 Scale Effects. 314
20.4 Propeller Hull Interaction 314
20.4.1 Thrust Deduction. 314
20.4.2 Taylor Wake Fraction. 315
20.5 Extrapolation of the Interaction Effects. 317
20.6 Extrapolation of the Open Water Characteristics. 317
20.6.1 The Equivalent Blade Section 318
20.6.2 Extrapolation of the Drag Coefficient of the Equiva-
lent Blade Section 319
20.7 Extrapolation of the Propulsion Test Results. 322
20.8 Trial Condition and Service Condition. 323
20.9 Efficiencies. 323
20.10Variations on the Extrapolation Method 325
20.11Example of Extrapolation of the Propeller Open Water Dia-
gram 326
20.12Example of the Extrapolation of the Propulsion Test 328
20.12.1 Comparison with Resistance Test 330
20.13Extrapolation of the Example using the Marin Method. . 333

21 Propulsion Calculations. 337


21.1 Statistical Prediction of the Model Wake Fraction. 337
21.2 Statistical Prediction of the Full Scale Wake Fraction 340
21.3 Statistical Prediction of the Thrust Deduction 340
21.4 Statistical Prediction of the Relative Rotative Efficiency. . 341

22 TABLES 343
A DICTIONARY 355
B WOORDENLIJST 362
8 G.Kuiper, Resistance and Propulsion, January 3, 1994
Preface
This is an introductory course on ship resistance and propulsion for the
Maritime Technology Department of the Delft University of Technology.
The text is written for students who have only basic knowledge of mathe-
matics and fluid dynamics. Vector and tensor notation is therefore avoided.
The propeller inflow is averaged in time and space to an average uniform
inflow, and the propeller loading is consequently assumed to be steady. The
unsteady conditions will be treated in an advanced course.

The intention of the course is to describe the models which are used.
This means that this course does not contain the complete diagrams, data
and formula's necessary for the actual application of the methods. These
will nowadays often be contained in a computer program. The use of com-
puter programs in routine calculations makes it even more necessary that
the user understands the model which is used and the restrictions which are
inherent to such a model. For an engineer it is risky to refer only to "a
formula" without knowing the basics behind this formula. It is even more
risky to refer to a computer program, which may contain fudge factors, even
errors, and which_will be changed over time.

It is also essential for an engineer to be able to formulate rapidly a crude


approximation of the problem and to grasp the main variables involved. For
this and for a proper use of complicated programs understanding the basic
approach is more important than the detailed development of a theory. This
understanding is the aim of this course.

Structure of the course


The first part of the course is on resistance, the second part on propellers.
Only in the last chapters on the propulsion test the interaction between hull
and propeller is accounted for. This sequence has been chosen because it
allows a gradual introduction of the concepts involved.

9
10 G.Kuiper, Resistance and Propulsion, January 3, 1994

The prediction of resistance, propeller and propulsion characteristics


each are treated in three ways:

By extrapolation from model test results.

By systematic or sfatistical data

By flow calculations

This sequence is natural since model tests form the basis for many
systematic data sets. The development of computational fluids dynam-
ics (CFD) has been rapid in the last decade, so these methods have become
a considerable help in the prediction of the behavior of ships at full scale.

Model tests and computations are often complementary, both having


their advantages and diadvantages. Model tests have the disadvantage of
possible scale effects, but have the advantage that complex flow phenomena
can be simulated.

Calculations have the advantage that the flow can be calculated in detail
and that variations can be made rather easily. However, drastic simplifica-
tions such as inviscid flow are used in the calculations. An important aim of
this course is to explain the complementary role of calculations and model
tests.

Textbooks
It is not intended to provide a full inventory of practical methods for ship
design or for the prediction of resistance and propulsion. For this the Prin-
ciples of Naval Architecture [38] is more suitable. The basis of the math-
ematical description of marine hydrodynamics can be found in Newman's
book with the same title [33]. Related specialized books are Lighthill's book
on waves [26]and the books of Knapp [21] and Young [46] on cavitation. An
introduction in basic aerodynamics with numerical solutions of potential
flow problems can be found in Katz and Plotkin [19]
The emphasis in this course is on the practical application of first princi-
pies to the prediction of the behavior of ships and propellers. Insight in
these first principles increases understanding of the complex phenomena
and forms a basis for intelligent problem solving.
January 3, 1994, Preface 11

Intermezzo's
The basic knowledge on fluid mechanics, required for the understanding
of the introductory course, may not always be available. Therefore some
chapters on the basics of fluid mechanics, such as the equations of motion
and its simplificatkins, notably potential flow and boundary layer flow, have
been included. These chapters, which are not a part of the introductory
course, are indicated as Intermezzo in the title. The reader who is familiar
with these topics can skip these chapters, others can read them without
going into great detail. The objective of these chapters is again to show
the basic approach, not the details. Some equations are therefore used in
one direction only. This makes it possible to avoid vector analysis, which is
nearly unavoidable for the full three dimensional equations.

Additional data
In the text some additional data, such as formula's which are often used,
are printed in smaller print. This text is only given for convenience when
the reader is going to use the material for his own purposes. It is not a part
of the text and does not add to the understanding of the problems.

Important formula's or statements


Some conclusions,- formula's or definitions are important throughout the
text. In order to recognize and retrieve these more easily, a box has been
placed around the text involved.

References
Since this course is aimed at an understanding of the basic approach of a
topic, a limited use of references has been made. In most cases users of this
course will not yet study literature of the subjects in depth. No efforts have
been made to refer to the most recent literature. For that the references
in the Proceedings of the International Towing Tank Conference [16] or
textbooks can be used. When names are linked to formulations or theories,
these names are given with the year, but without the reference. E.g. the
Betz condition for optimum efficiency is mentioned to date back to 1929,
but is not referred in the references. Only when full sets of diagrams, data
or formula's can be found in the literature full references are made. This
makes the list of references less dependent on the most recent publications.
12 G.Kuiper, Resistance and Propulsion, January 3, 1994

Acknowledgements
Many students and colleagues from Marin have given comments, corrections
and material for this course. The help of Mrs. Raven, Hoekstra, van Gent,
van Wijngaarden, Holtrop-, de Koning-Gans is gratefully acknowledged. The
text will be developed furtlier in the future. Therefore the date of printing
is present on all pages aiid on the title page. Any comment can be helpful
to improve it and will be very welcome.

G.Kuiper. January 3, 1994


Chapter 1
Hull forms
Objective: Introduction of the hydrodynamic features of some ship types.

1.1 Displacement Hulls.


The most common purpose of a ship is transport of cargo. For such ships
the displacement hull is the most appropriate concept. The weight of the
cargo, stores and fuel is called the (deadweight). The deadweight and the
weight of the empty ship together are equal to the displacement of the ship
(Achimedes' law).

1.1.1 Efficiency.
The movement of such a displacement ship requires little energy in com-
parison with other means of transport, at least when the required speed
of transport is low This is because the friction of water is low as long
.

the generated wave height is small. To give an idea: a ship of 100 meters
lenght, 12.5 meters breadth and 5 meters draft at a speed of 20 km/hr re-
quires approximately *** kW. Still the deadweight of such a ship is about
3500 tons. Compared to road transport, where a 20 tons truck requires
some 100 kW to drive at 80 km/hr the amount of fuel, required for water
transport is low. This can be expressed kWh per tonkm. A characteristic
feature of displacement ships is therefore that a large amount of cargo is
moved at low speeds. The restriction of a low speed is important, at increas-
ing speeds the required power increases very rapidly due to wave generation.

13
14 G.Kuiper, Resistance and Propulsion, January 3, 1994

1.1.2 Typical Speeds.


At higher speeds the influence of waves becomes large and the waves be-
come responsible for most of the resistance. So the speed is an important
parameter for the type of ships. For bulk transport (oil, ore, coal, etc)
speeds of up to 16 knots ade common.
1 LNG is transported* special tankers at a speed of about 20 knots. For
large, fast displacement ships such as containerships and reefers, speeds of
up to 25 knots are found, with extremes of over to 30 knots for e.g.passenger
vessels and up to 40 knots for navy ships.

The displacement hull is by far the most common type of ship and in
this introductory course most attention will be devoted to this type of ship.

1.1.3 Hull Forms.


Since transport of cargo is_the basic purpose of most merchant displacement
ships the most efficient hull form , both from a viewpoint of cargo stowage
and from the viewpoint of building costs, would be a square box. (Fig. 1.1)
. However, the consequences of such a simple shape in terms of resistance

are too large. So bow and stern are shaped such that the volume remains
but the resistance is decreased. Efforts have been made to design hullforms
with chines, preferably with surfaces which could be developed into flat
plates (Fig. 1.2). When the chines are accurately in the direction of the
flow such a ship can be as-good as a faired hull. Most of the hullforms are
faired, however.

1.1.4 Form Parameters.


The form coefficients of a ship are given in a non-dimensional form.

An important coefficient is the block coefficient C2 , defined asCB =


7 with v=volume of the displacement, L=length at the waterline 2
LBT
B=breadth en T=draft. The block coefficient is an indication for the full-
'Although SI units should be used the speed of a ship is expressed in knots, which
is an international or U.S. nautical mile of 1852 meter per hour. A nautical mile is 1
arcminute of a greatcircle. The U.K. nautical mile of 1853.184 meter is also used. Note
that the nautical mile is different from the statute mile of 1609.344 meter, which is in
use for distances over land.
2Hydrodynamically the length is the length of the waterline L1 although the differ-
,

ence between the waterline length and the length between perpendiculars Lpp will mostly
be negligible from a hydrodynamic point of view.
January 3, 1994, Hull forms 15

Figure 1.1: Transport of containers

Figure 1.2: Hull form with chines in the afterbody

ness of the hullform. It is also indicated by S.

The vertical prismatic coefficient Cp, defined as Cp, = -22-r


A,T-, where Au,
16 G.Kuiper, Resistance and Propulsion, January 3, 199.4

is the area at the waterline. It indicates the vertical distribution of the dis-
placement. The longitudinal prismatic coefficient Cp is similarly defined as
A.L where A, is the area of the maximum tranverse section, which gener-
ally is the midship section.. The longitudinal prismatic coefficient indicates
the moment of inertia of the displacement around the midship section. It
is also indicated by O.

The midship section coefficient indicates the fullness of the midship sec-
tion and is defined as Cm =---m-ABT. It is also indicated by 0.

The longitudinal position of the center of buoyancy is given as a per-


centage of the ship length L and is indicated as LCB .

The waterline coefficient Cui P - LB in dicates the fullness of the water-


line. Cwp is also indicated by a.

These coefficients can be formed similarly for the fore- and afterbody.
When a parallel middlebody is present these coefficients can also be formed
for the entrance and run, but because the length of entrance or run is diffi-
cult to determine that is not common.

1.1.5 Considerations for the Stern Form.


The form of the stern is predominantly determined by the requirement of
attached flow and proper inflow to the propeller. When the afterbody is too
blunt the flow will detach from the hull, a phenomenon called separation.
This drastically increases the resistance. In principle it is simple to avoid
this by making a slender ship. But this requires a larger and more expensive
ship for the same deadweight. So the main topic of ship hull design is
to optimise the conflicting requirements of economics and resistance. In
practice it means that a ship's hull has to be designed such that the flow
is on the verge of separation. This makes the hydrodynamics of the ship's
hull very complicated.
The form of the afterbody and the stern is also strongly dependent on the
type of the propulsor. Some forms will be shown in Chapter 2.

1.1.6 Considerations for the Bow Form.


The shape of the bow is predominantly determined by the generation of
waves and therefore depends on the ship speed. The fullness of the bow
January 3, 1994, Hull forms 17

Figure 1.3: Fast cargoship with a bulb

decreases with inc-reasing speed. Tankers and bulkcarriers have a block co-
efficient up to 0.85, a slender fast containership has a block of 0.6 or lower.
As a consequence, fast ships (Fig. 1.3) have a slender bow, slow ships such
as tankers have a very full bow (Fig. 1.4). For very full ships with block
coefficients over 0.80 a cylindrical bow has been applied sometimes (See
Fig. 1.5a). Poor ballast performance and high resistance in waves have
made this type of bow obsolete.

1.1.7 Bulbs.
On many fast ships a bulb is applied, as shown in Fig. 1.3. A bulb is applied
to decrease the generation of waves around the ship. Many different shapes
have been designed, as shown in Fig. 1.5.
Note that the bulb is mostly designed for one draft. In Fig. 1.3 the
ship is in slightly loaded condition, where the bulb is partly above water
and therefore less effective or even counterproductive. For tankers, which
operate frequently at ballast draft, a bulb which is effective at various drafts
18 G.Kuiper, Resistance and Propulsion, January 3, 1994

41,

A--74
AL,

-
-
_Lc

.136%-7, -

-;t:o.pu"
--orr-Ar
4s`

Figure 1.4: Loaded tanker

is used, as shown in Fig. 1.6. Application of a bulb on a tanker is not very


effective because the wave fesistance is low (of the order of the air resistance)
and the increased frictional resistance of the bulb dominates.
As will be discussed later, a bulb is only effective in a certain speed range.

For ships with very high speeds the bulb loses its effect because the wave
system changes (see Chapter 5 ) and a sharp bow is applied. (Fig. 1.7)

1.2 High Speed Ships.


As metioned the generation of waves causes a high resistance at high speeds.
When these high speeds are required, the speed can also be used to create
lift. This reduces the displacement and strongly affects the wave generation
and thus the resistance. There are several ways to create lift.
January 3, 1994, Hull forms 19

07-",f3TRAM
-organ

AIM
I Ia

"'A:1:0-m.

Figure 1.5: Various bulbs

1.2.1 Planing Hulls.


In case of planing there is a pressure build-up on the bottom of the ship,
such that an upward force is generated. The displacement is thereby re-
20 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 1.6: Bulbcontour for various drafts

Figure 1.7: High speed displacement ship

duced, but not eliminated entirely. The upward force is obtained by a flat
bottom. The water flowing along the hull is mainly flowing along the bot-
tom and it is displaced downwards.
A flat bottom is very sensitive to incoming waves. High loads occur, a phe-
nomenon called slamming. To reduce this sensitivity deadrise is used in the
midbody, in combination with a sharp bow. In the stern region the bottom
January 4, 1994, Hull forms 21

is nearly flat (see Fig. 1.7) and ends in a cut-off stern, the transom stern.
This causes the flow to separate smoothly from the hull.

To increase the vertical force of the water in the afterbody and to con-
trol the trim a trim wedge can be applied at very high speeds. The bottom
of the trim wedge is a continuation of the transom stern. The trim wedges
are also made as adjustable flaps extending from the flat bottom.

Planing may generates excessive spray, which causes additional resis-


tance (see Fig. 1.8). Spray rails are therefore used to reduce the spray.

Figure 1.8: Spray generated by a planing hull

These rails are a kind of longitudinal spoilers on the hull, which deflect the
spray downward.

The amount of planing can vary. A small planing force can be generated
by using chines. Planing is common for luxury yachts (Fig. 1.9). Extreme
planing is used in speedboats and racing boats.

1.2.2 Hydrofoils.
Displacement can be eliminated entirely by using foils to carry the whole
ship. Such ships are called hydrofoil ships. These ships are designed specif-
ically for high speeds. The foils can pierce through the water surface to
22 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 1.9: Planing luxury yacht

ensure stability. In such a case these are called surface piercing hydrofoils.
(Fig. 1.10).
These hydrofoils have io operate close to the surface, a condition where the

kf.;

Figure 1.10: Surface Piercing Hydrofoil

lift is reduced. The transition from water to air also causes additional spary
resistance . This is avoided by the fully submergedhydrofoil ship (Fig. 1.11).
In that case stability and trim have to be maintained by actively controlled
fins.
January 4, 1994, Hull forms 23

ft

Figure 1.11: Submerged hydrofoil

Hydrofoils are applied in the speed range up to 40 knots. Especially the


fully submerged hydrofoils are rather insensitive to waves, as long as the
hull is not hit by green water.

Hydrofoils are genererally driven by propellers. 3 The shafts, which ex-


tend from the hull into the water, are a source of high resistance. Moreover,
the propeller thrust is not exactly in the forward direction due to the rake
of the shafts. This is improved when propellers in front of or behind the
main hydrofoil are used, driven with Z-drives (see chapter 2). In that case
the propulsor is integrated in the hydrofoil.

The distribution of the load over the front and rear hydrofoils can differ.
When the front foil carries only a very small part of the load this is called a
3Sometimes these propellers are supercavitating propellers, see Chapter 2.
24 G.Kuiper, Resistance and Propulsion, January 4, 1994

canard arrangement, as in the case of airplanes with the stabiliser in front


of the wing instead of at the tail.

1.3 Air as Carrier.


Instead of lift also air can be used to create an upward force. This is used
in air-cushion vehicles.

1.3.1 Air Cushion Vehicles.

a is
tinzintvp=e----

Figure 1.12: Air Cushion Vehicle

Air can carry the weight of the ship when maintained at a high enough
pressure. This pressure is built up in an air cushion, which is maintained
below the vessel by skirts around the ship (Fig. 1.12). In such a case the
ship is called an air cushion vehicle or ACV. Loss of air will occur in waves
or due to forward speed, so the air pressure has to be maintained with air
compressors.

An ACV still has a displacement which is equal to the weight of the total
ship. The pressure inside the cushion times the area of the cushion has to
be equal to the total weight or displacement. An ACV therefore does not
float above the water, as it does on land. The total resistance of an ACV
is lower than that of a displacement ship due to the lower friction over the
bottom and partly because of the more favorable wave forms.
Typical for an ACV is its amphibious character: it can operate both on land
January 4, 1994, Hull forrns 25

and in the water. They are therefore generally propelled by air propellers.
ACV's can also operate over a wide speed range.

1.3.2 Surface Effect Ships.


When the amphibious character is not required the loss of air under the
skirts can be reduced by using fixed side walls. These also improve the
behavour in waves (Fig. 1.13). Of course, the side walls have frictional
resistance, but the shape of the walls can be better streamlined than skirts
and the resistance is therefore lower. Such ships are called Surface Effect
Ships or SES. They can be used for very high speeds of up to 60 knots, but
a more common speed range is between 25 and 35 knots.

Figure 1.13: Surface Effect Ship

The largest SES vessels nowadays have a length of about 50 meter and
a speed of almost 100 knots (US-Navy). In waves the speed reduction,
however, is larger than e.g. with hydrofoils and it occurs at lower sea states.

1.4 Multi Hulls.


1.4.1 Catamarans.
Long slender ships have a low wave resistance and are therefore good at
high speeds. A very slender displacement ship, however, is very narrow and
26 G.Kuiper, Resistance and Propulsion, January 4, 1994

has no deck space nor stability. This can be countered by using two hulls: a
twin hull ship or catamaran .. An example is a passenger ferry (Fig. 1.14),

- -1:'''
I iL.
, -44,-4 _
;,.., -1.74111 *sin
7-- iressisoirsiarir 1, N

Figure 1.14: Catamaran

where a large deck space is required for a relatively small displacement.

The slender hulls have a low wave resuistance, although the wetted area
is almost doubled in comparison with a mono hull, which increases the fric-
tional resistance. So a catamaran is typically used for higher speeds, where
the wave resistance becom- es important. Catamarans operate satisfatorily
in calm water. Its response to waves is still a problem. In such conditions
it behaves uncomfortably and there is a risk of hitting the water with the
superstructure.
Efforts. have been made to improve the riding qualities of a catamaran by
special bow shapes, such as the wave piercer. The effects still have to be
proven.

1.4.2 Swath.
A variation on a catamaran is a Small Waterline Area Twin Hull or SWATH
ship (Fig. 1.16).
In that case the displacement is brought far below the waterline, thus
reducing the waterline area to a minimum (Fig. 1.17). As a result the vessel
will react only sligtly on waves, so it offers a stable platform in waves. A
January 4, 1994, Hull forms 27

Figure 1.15: Wavepiercing catamaran

Figure 1.16: SWATH

disadvantage is of course that its stabvility is very poor, so it is very sen-


sitive to changes in loading or even to forward speed. A SWATH therefore
must have active fins to control trim and stability. These fins can also be
used for further roll reduction.
28 G.Kuiper, Resistance and Propulsion, January 4, 1994

DWL
STA 111
17
le
11
20
21
22

AFT FWD

Figure 1.17: Cross section of SWATH hull

Literature for Further Reading.


Several afterbody forms and their relative merit are given by Vossnack and
Voogd [45]
A very rough estimate of the relative power requirements of various high
speed concepts is given by Dorey [6].
Chapter 2
Propulsors

Ob jective: Introduction of various types of pro pulsors and their main


proper- ties

The basic action of a propulsor is to bring water into motion. The force
required for that is the thrust force. The energy of the water behind the
propeller is lost energy.

The amount of concepts for ship propulsion is large. The most impor-
tant criterion for a propulsor is its efficiency. The efficiency varies widely
between various types of propulsors, but the screw propeller has not yet
been equalled in most cases.
The propulsor is generally mounted behind the hull. This is because of
efficiency: the water which is set into forward motion by the friction along
the ship is reversed by the propeller action. As a result less energy is left
behind in the water.

A risk for every propulsor operating at high rotational velocities is cav-


itation . This occurs when the local pressure in the fluid is lower than the
vapor pressure due to local high velocities. Regions with vapor occur e.g.
on the propeller blades, such as occurs extensively in Fig. 2.8. When these
vapor filled (not air filled) cavities arrive in regions with a higher pressure
they collapse violently, causing erosion (Fig. 2.1). Strong dynamic behav-
iour of large cavities also generate vibrations in the ship structure.

29
30 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.1: Example of erosion due to cavitation

2.1 Propellers.
The most common propulsor is the screw propeller. A propeller generates
a force by lift on the blade sections. These blade sections are similar to
airfoils, operating at an angle of attack in the flow. The geometry of the
propeller blades is quite critical due to the occurrence of cavitation, as
described below. Therefore a separate propeller is generally designed for
each ship to accomodate the specific circumstances behind the ship. The
geometry of the propeller blades has to be very accurate too. A propeller
is therefore a delicate piece of equipment. An example of a finished set of
Navy propellers is shown in Fig. 2.2. The propeller will be treated in some
detail in this course.

2.1.1 Propeller Arrangements.


The propeller is located behind the hull. The traditional afterbody shape is
such that the hull ends in a screw aperture in front of the propeller, while
the rudder stock forms the after part of the propeller aperture.
January 4, 1994, Propulsors 31

Figure 2.2: Newly finished Navy Propellers (courtesy Esscher Wyss)

When the hull is cut away both above and below the propeller shaft
the stern is called an open stern , as shown in Fig. 2.3. An arrangement
with the propeller shaft extending under a flat stern under a small angle is
typical for Navy ships and twin screw ships. The shaft can be supported by
brackets (Fig. 2.4). In large twin screw ships like passenger ships the shafts
are covered by bossings .

2.1.2 Trusters.
A propeller can be driven from above by a vertical shaft. This makes it pos-
sible to rotate the propeller along the vertical axis and to generate thrust
in all directions. These configurations are called thrusters. . An example is
shown in Fig. 2.5. Thrusters are common for dynamic positioning , as illus-
trated in Fig. 2.13. The use of such thrusters for normal propulsion is still
limited because the shaft close to the propeller decreases the efficiency and
because of the more complicated construction. In fast ships or in hydrofoils
a thruster arrangement can also be used, as shown in Fig. 2.6. In this figure
a special arrangement with shafts is also shown. This arrangement reduces
32 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.3: propeller with open stern

r!t, .anwen
ar......orr
emil

.
'.
f.
::',,, .
--
-- - 4..:..:1-= ,------!
,
i

Figure 2.4: Shafts with brackets

the shaft angle as much as possible.

2.1.3 Controllable Pitch Propellers.


In case of a fixed pitch propeller the thrust, and consequently the speed of
the ship, is controlled by the propeller revolutions. In case of a controllable
January 4, 1994, Propulsors 33

Figure 2.5: Thruster configuration

NUPE! -'11;

A111111111

111111.110111111116110,10,111MINIMO
171:3 tinn5111.15._Aii

Figure 2.6: Different shaft arrangements (Courtesy Hydromarine, Italy)

pitch propeller or CPP the thrust is controlled by changing the pitch of


the blades. In tha t case the shaft is at a constant rotation rate. This is
often used when the propeller has to operate in more than one condition,
e.g. free running and towing. It is also effective when rapid manoeuvring is
34 G.Kuiper, Resistance and Propulsion, January 4, 1994

required. Reversing the thrust occurs by changing the pitch with constant
revolutions in the same direction. This decreases significantly the time re-
quired to change the direction of the trust. A CP propeller is specifically
favourable in case of high_ skew, because a highly skewed fixed pitch pro-
peller will experience extremely large moments on the blades.

The hub of a CPP is of course more complicated and expensive, while


the hub diameter is also larger than that of a fixed pitch propeller. This is
a disadvantage for hub and blade root cavitation.

2.1.4 Overlapping Propellers.


For large ships with high speeds the thrust is distributed over two or more
propellers. Such a twin screw configuration has a lower efficiency because
the propellers operate outside the region of the highest wake. To increase the
efficiency the twin propellers can be brought together as close as possible,
with one propeller slightly -ahead of the other. The blades can than overlap
and the twin screw arrangement approaches a single screw arrangement.
The propellers can in principle rotate in opposite direction, so that also a
contra rotating arrangement is approached.

2.1.5 Contra Rotating Propellers.


A rotating propeller also induces a rotating motion in its wake. This is
lost energy. In order to gain this energy two propellers behind each other
are used at the same shaft (Fig. 2.7). These propellers turn in opposite
directions, thus eliminating each others' rotating wake. The diameter of
the front propeller is often slightly larger than that of the behind propeller,
to account for the contraction of the propeller wake.
Efficiency gains of over 10 percent are claimed for such configurations, al-
though the lost rotational energy in the wake is less.
The construction of the shaft is complicated and costly. Some prototypes
have been build.

2.1.6 Surface Piercing Propellers.


When the draft is too small for a normal propeller to operate tunnels are
applied in the hull to lead the water upwards to the propellers. When this
is also insufficient the propellers are allowed to operate partly submerged.
Such propellers are called surface piercing propellers. The geometry of the
blades is generally skewed to soften the impact of the blades on the water
surface and the exit from the water. Air suction may decrease the efficiency
January 4, 1994, Propulsors 35

'
..`"" .
-

Figure 2.7: Contra rotating propeller arrangement

of the submerged blades, but the efficiency drop compared to submerged


propellers is not very large. [4]

2.2 Special Types of Propellers.


2.2.1 Supercavitating Propellers.
When cavitatio-n occurs extensively, e.g. at very high rotation rates of
the propeller, it is advantageous to use blade sections which generate a long
sheet cavity at one side of the blade, about two times the chordlength of
the blades. These propellers are called supercavitating propellers (Fig. 2.8).
Because cavitation implodes far behind the blades the danger of erosion is
absent, at the cost of a drastically reduced efficiency.

2.2.2 Agouti Propellers.


A special way to control cavitation is to supply air to the cavity. The cavity
will then contain air together with vapor and on implosion the air will cush-
ion the collaps. As a result the radiated noise of the cavitation is lower than
without air supply. The amount of air supplied is very critical, because an
overdose of air will increase the cavity volume drastically.
36 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.8: Supercavitating Propeller

The air is supplied through small holes at the leading edge of the blades.
A restricted supply of air will not affect the efficiency of the propeller.
Agouti systems are used only for navy ships.

2.2.3 Tipplates.
An increased loading of the blade tips would be beneficial to efficiency, but
the flow around the tips prevent such a heavy loading. In order to prevent
such a flow around the tips tipplates have been applied. Because this would
also reduce the strength of the tip vortex these propellers have been called
Tip Vortex Free Propellers or TVF Propellers. It is, however, very difficult
to locate the tip plates properly in the wake behind the hull and the tip
vortex does not disappear in general.. Moreover, the tip vortex tends to
occur in the corners between the tip plates and the blades.

These propellers have mostly been applied in combination with a duct.


A new develop- ment with more sophisticated design techniques is being
developed by de Jong [17] He also developed new shapes of the tip plates,
based on numerical calculations (Fig. 2.9).No full scale applications are
available yet.
January 4, 1994, Propulsors 37

Figure 2.9: Propeller model with tipplates

2.2.4 Vane Wheels.


A large propeller diameter is often benificial for efficiency. When an
increase of an existing diameter is benificial or when the diameter of the
main propeler is restricted a vane wheel can be applied. This is a kind of
propeller, which runs freely downstream of the main propeller. (Fig. 2.10).
The inner part of the vane wheel, the impeller part or turbine part, has a
pitch such that the vane wheel is driven by the wake of the main propeller.
The outer part of the blades of the vane wheel, the propeller part, has a
different pitch, which causes the vane wheel to generate thrust at these
radii. The rotation rate of the vane wheel will be lower than that of the
main propeller. (In German the vane wheel is called after its designer the
"Grimmse Leitrad"). The concept is patented.

2.3 Ducted Propellers.


At high propeller loadings a duct can increase efficiency (Fig. 2.11).
A duct generates part of the total thrust due to its interaction with the
38 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.10: Vane wheel

propeller. This is the case with an accelerating duct (Fig. 2.12a) , in which
the flow velocity is increased due to the duct. The duct shape can also
cause the flow to be decelerated (Fig 2.12b). This suppresses cavitation,
but decreases the efficiency. A decelerating duct is therefore suitable for
navy ships only, and there it is rarely applied.

Ducted propellers are used in a wide range of applications of heavily


loaded propellers, such as for tugs and in applications for dynamic posi-
tioning (Fig. 2.13) . These thrusters can freely rotate over the full circle
and are therefore also called azymuthing thrusters. The power of these sys-
tems is increasing rapidly with increasing availability of appropriate gears
(Fig. 2.14.
A special application is on active rudders (Fig 2.15).

The flow along heavily loaded ducts may separate from the duct, which
decreases their effect and increases their resistance. A method to reduce
this type of separation is the application of slots at the exit of the duct (see
Fig. 2.16).

The gap between the blade tips and the duct has to be small for a proper
interaction of propeller and duct. This makes the construction of the duct
January 4, 1994, Propulsors 39

'Nara

--
Figure 2.11: Ducted Propeller

decelerating acr-elerating

Figure 2.12: Accelerating and decelerating duct

more difficult, especially the very large ducts on e.g. tankers.


For manufacturing reasons the duct is also generally rotational symmetric:
it has the same cross section at every position. A-symmetrical ducts have a
different angle over the circumference to make the wake distribution more
uniform.
40 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.13: Dynamic positioning

Instead of an a-symmetric duct other types of fins or ducts can be applied


to make the propeller inflow more uniform. These ducts or fins are applied
at some distance upstream of the propeller. Generally they accelerate the
retarded flow in the upper part of the propeller plane. A patented concept
is the "Schneekluth Duct", as shown in Fig. 2.17 . Variations are possible,
as in Fig. 2.18.

2.3.1 Ringpropellers.
A variation on the ducted- propeller is the ringpropeller indexringpropeller.
This is a duct similar to the normal duct, but now the duct is connected
to the propeller blades and rotates with it (Fig. 2.19). This eliminates the
gap between blades and duct, but at the cost of a greatly increased viscous
resistance. The efficiency of a ringpropeller is therefore relatively low.

2.3.2 Mitsui Duct.


The position of the propeller of a ducted propeller is generally inside the
duct. The propeller can also be moved towards the exit of the duct without
too much loss of efficiency. Such a position is appropriate when a duct is
used as a retrofit , that is an improvement afterwards. It should be kept in
mind, however, that the application of a duct in front of an existing propeller
will change the propeller loading and may require another propeller design.
Mitsui has patented such retrofits with a duct, the combination is therefore
January 4, 1994, Propulsors 41

Figure 2.14: Thrusters for dynamic positioning

also called a "Mitsui Duct".

2.4 Other Propulsors.


2.4.1 Voight-Schneider Propellers.
A very special propulsor is the Voight-Schneider Propeller (Fig. 2.20).
a number of "knifes" on a rotating plate. These "knifes" can rotate on this
plate and their position is such that they are always perpendicular to the
radials from a moving centerpoint P, as shown in Fig. 2.21. When this
centerpoint is in the center of the blade circle there is no resulting force
(Fig. 2.21c). When this centerpoint is moved a thrust is generated perpen-
dicular to the direction in which the centerpoint is shifted.
The main asset of a Voight-Schneider Propeller is that in that way the thrust
can be applied in all directions, just by moving the centerpoint. Rudders
and shafts can be omitted. This can be used e.g. for tugs or supply boats,
for which manoeuvring is important. Its efficiency,however, is lower than
that of an open propeller due to the fact that the blades generate thrust
over part of the revolution only, while the viscous resistance is present over
42 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.15: Active rudders

.iwansa
C

Figure 2.16: Duct with trailing edge slots (Courtesy v.Gunsteren and
Gelling, Delft, the Netherlands)
January 4, 1994, Propulsors 43

'

Figure 2.17: Flow improving fins (Schneekluth)

the whole revolution.

Voigth Schneider propellers can be mounted under a flat bottom. For


protection some cover is sometimes applied (Fig. 2.20).

2.4.2 Paddle-Wheels.
The oldest form of mechanical propulsion after the sails is the paddle
wheel . Contrary to the propeller, which uses lift for propulsion, a paddle

wheel uses drag, which at higher speeds is less efficient. The blades of a
paddle wheel are most effective in the lowest position, in other positions
they also generate a vertical force. So a paddle wheel has to be large, with
only a small immersion. In order to improve the entrance and exit of the
blades in and from the water, the blades have been made rotating by a
system of rods. This made the wheel very complicated, however.

2.4.3 Pump Jets.


44 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.18: Flow improving fins

Figure 2.19: Ringpropeller

The basic mechanism of propulsion is acceleration of water. This cannot


only be done by e.g. a propeller outside the hull, but also by a pump inside
the hull. The water is sucked in from the bottom of the ship, is accelerated
inside the ship by a pump and leaves the ship at the stern. This has many
advantages when a propeller is too sensitive to damage, or when a propeller
is dangerous (e.g. rescue vessels). Also in shallow waters a pumpjet can
January 4, 1994, Pro pulsors 45

Genera!plan

Figure 2.20: Voight-Schneider Propeller

Figure 2.21: Blade positions of Voight-Schneider Propeller

be useful. The inner surface of the pump system is large and the velocities
inside are high, so the viscous losses are high too. The efficiency is therefore
lower than that of an open propeller.

A special version of a pumpjet is the rotational pumpjet, as shown in


Fig. 2.24. The water goes into the pumpjet at the center of the jet and
is blow out tangentially. Rotation of this pumpjet along the vertical axis
makes it possible to control the direction of the thrust.

2.4.4 Sails.
The oldest sails were square rigged, using drag as the thrust force,just
46 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.22: Paddle wheel

Figure 2.23: Pumpjet

as the paddle wheels later (Fig. 2.25). Sailing towards the wind is not
possible with this rigging. Before the steam engine took over longitudinal
sails were also used. When the energy crisis hit, some modern sail designs
were made of both form, either as additional power or as main propulsor.
(Figs. 2.26 and 2.27). The use of computer controlled settings of the sails
can highly improve their operation. The development of racing yachts as
the 12 meters, used for the America's Cup, can provide more experimental
January 4, 1994, Propulsors 47

Impression of
pumpjet

Figure 2.24: Rotational pumpjet

and theoretical experience with sails. Sails will only become attractive when
the fuel price rises again considerably.

2.4.5 Other Types of Propulsion.


When a cylinder rotates in wind a thrust force is generated. This effect
resembles sailing. The rotating cylinders are called Flettner Rotors , after
their original designer. Flettner rotors have been applied on experimental
ships only (Fig 2.28). A major problem is the mechanical connection of
the rotor to the hull, where the ship motions cause very large forces and
moments.

Even more esoteric types of ship propulsion are ramjets , which are an
analogy of jet engines. In a water jet hot compressed air is injected in a
water stream, and the expanding air accelerates the flow in the engine.

Nature has often been an example for technology (although the airplane
only became practical after the bird wing motion was abandoned). Fish
propulsion has also been an example. In this case a flat plate makes sinu-
soidal motions perpendicular to the direction of the ship. The construction
is very complicated, of course.

A variation on the fish propulsion is Weiss-Fogh propulsion . This is


simply a flat plat which moves between two walls in a direction perpendic-
ular to the direction of motion of the ship. The angle of the plate is varying
48 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 2.25: Square rigged sail

with its position, so that the water is pressed towards the rear during the
motion of the plane. As with fish propulsion this is mainly of theoretical
importance.

The same is true for magneto-hydrodynamic propulsion . In this case a


strong magnetic field accelerates the flow in a magnetic duct. In principle
no moving parts are required for this type of propulsion. The high electrical
resistance of sea water makes the efficiency extremely low, however. It still
has to be developed for other fluids with more suitable properties.
January 4, 1994, Pro pulsors 49

Figure 2.26: Modern sailing ship (Wind Spirit)

Figure 2.27: Modern square rigged sailing ship


50 G.Kuiper, Resistance and Propulsion, January 4, 1994

a;a

ro ....., 4 r.

- - - "

Figure 2.28: Flettner rotor propulsion


Chaptei' 3
Intermezzo: Resistance of
Simple Bodies
,
Purpose: Brief introduction to non-dimensional formulations and hydro-
dynamic concepts relevant for ship resistance.

To understand the physics of the flow around a ship it is useful first


to look to the flow around a very simple body such as a flat plate in flow
direction.

3.1 Non-dimensional Coefficients.


Each flat plate has its own resistance RT . There is a need to compare the
resistance of plats of various lengths at various velocities. Experiments
show that the resistance of a plate is proportional to the square of the
velocity and proportional to the area S[m2] of the plate(heightH x lengthL).
When the resistance of the plate is measured in different fluids it appears
that the resistance is proportional with the density p[kg Im3] of the fluid.
The resistance RT[N] of a plate of arbitrary dimensions at an arbitrary
velocity V[mIsec] can therefore be expressed by a single number containing
these proportionalities. This number is the drag coefficient Cd:

Cd 1/2pv2s (3.1)

The resistance RT in this equation is the total resistance. It is called the


total resistance because various components of the total resistance will be
distinguished later. The dimension of the drag coefficient is found from the
dimensions of its components to be 1. That means that the drag coefficient

51
52 G.Kuiper, Resistance and Propulsion, January 4, 1994

is a real coefficient, because it is non-dimensional. Each plate, whatever its


size or velocity, has the same drag coefficient in the sarne circumstances.
An important clause is under the same circumstances. As will be shown
later this means that the flow has to be similar in all cases, which is true
when there are no other pa- rameters for the drag than the size and the speed.

.020

.010
.009
.004
.002
.006
co .00.3
e 0.242
.004 LCIG,

.003 . /;:r F
CF L3213 TURBULENT (SC NOE P44E RR )
I. LAMAR (BLASIUS)
.002

.001
10
107 to'
UA
REYNOLDS NUMBER

Figure 3.1: Drag coefficients of a flat plate

3.2 Drag of a Flat Plate.


When the drag coefficient of a flat plate is measured at various velocities
the dots in Fig. 3.1 are found. In this diagram the velocity in the abscissa
is replaced by the Reynolds number ULlv, which will be discussed later. It
is noted at first glance that the drag coefficient is not a constant! So there
are other parameters involved in the drag of a cylinder. Systematic tests
showed that at high velocities the drag coefficients of plates with the same
product V x L, where V is the velocity of the fluid and L is the length of
the plate, are the same. When the temperature is varied the viscosity of
the fluid is changed and systematic tests showed that the drag coefficient of
all cylinders will collapse on one line when plotted on the abscissa V X LiV
where v is the kinematic viscosity of the fluid[m2/sec]. This parameter is
called the Reynolds number:

VL
Rn = (3.2)
V
January 4, 1994, Simple bodies 53

The drag coefficient Cd in Fig. 3.1 has therefore been plotted against the
Reynolds number.

The Reynolds number is again non-dimensional, so that in Fig. 3.1 all


parameters are ex-pressed non-dimensionally. That means that the drag
coefficient is a fundion of the Reynolds number only and that Fig. 3.1 is
valid for all possible flat plates aligned with the flow. This property is the
purpose of expressing the parameters non-dimensionally.
In principle the dots should therefore form one single curve, which is not
exactly the case. So there are still other phenomena which influence the
resistance of a flat plate. The spread of the dots is caused by phenomena
in the boundary layer.

3.3 Boundary Layer Flow.


A boundary lay-er exists because the fluid particles at the wall of the flat
plate stick to the plate (no slip condition). At some distance from the plate
the free stream velocity V occurs. The region where the velocity varies
from zero to the outside velocity is called the boundary layer. This is a
region where strong velocity gradients occur. Due to these strong velocity
gradients the viscosity of the fluid has a large influence in the boundary
layer. At Reynolds numbers above 1000 this region is thin compared to the
length of the plate. The pressure from the outside of the boundary layer to
the wall in a thin boundary layer can be considered as constant, so that the
pressure in the outer flow is equal to the wall pressure. For calculations of
the outer flow the thin boundary layer can then be neglected.

In the boundary layer the velocity approaches the free stream velocity
asymptotically. The thickness 6' of the boundary layer is defined as the
distance from the wall where the velocity is 99 percent of V. The velocity
gradient at the wall determines the friction force between the fluid.
The shape of the velocity distribution in the boundary layer can be char-
acterized by various quantities. When the boundary layer is replaced by a
layer with uniform velocity V outside the boundary layer, with the condi-
tion that the same fluid moves through the layer, the displacement thickness
81 is found. This can be expressed by

V(51 = f ( V v)dy (3.3)


o

where y is the distance to the wall and v(y) is the local velocity in the bound-
ary layer. When the boundary layer is replaced by a layer with velocity V
54 G.Kuiper, Resistance and Propulsion, January 4, 1994

having the same momentum, the momentum thickness 9 is found:


1 15
9= v2dy
-V2 o

The ratio between the momentum thickness and the displacement thickness
is called the shape factor H of the boundary layer.

3.3.1 Laminar and Turbulent Flows.


The boundary layer can have different conditions. In a laminar boundary
layer the particles in the boundary layer are gliding smoothly along each
other, so that no motions perpendicular to the flow occur. In a turbulent
boundary layer the smooth motions disappear and violent motions perpen-
dicular to the direction of the motion occur. The turbulent motions of the
fluid particles cause an exchange of energy between the layers in the bound-
ary layer and as a result the velocity distribution in the boundary layer is
different, as shown in Fig. 3.2.

LAMINAR TURBULENT

Figure 3.2: Velocity distribution in laminar and turbulent boundary layers

From the velocity gradient at the wall in this Figure it follows that a tur-
bulent boundary layer results in a higher friction force than a laminar one.
The scatter in the dots in Fig. 3.1 is caused by the transition from laminar
to turbulent boundary layer flow. The line at lower Reynolds numbers is
the drag coefficient when the boundary layer is fully laminar. At a Reynolds
number around 105 transition to turbulence occurs. This transition begins
at the downstream edge of the plate and moves towards the leading edge
of the plate with increasing Reynolds number. At a Reynolds number of
106 transition occurs immediately at the leading edge and the boundary
layer on the plate is fully turbulent. The line in Fig. 3.1 at higher Reynolds
January 4, 1994, Simple bodies 55

numbers is the drag coefficient for fully turbulent flat plate boundary layers.

In Fig. 3.1 the Reynolds number is expressed based on the length L of


the plate. The location of transition depends on a local Reynolds number
R, based on the distance x from the leading edge of the plate 1 . Under
ideal conditions transition takes place at a fixed value of R,. When this
would be the case the dots in Fig. 3.1 would still form a single line. However,
transition is very sensitive to disturbances such as vibrations of the plate,
turbulence in the incoming flow, surface irregularities etc. This causes the
scatter of the dots in Fig. 3.1.
Because of the higher friction at the wall the boundary layer thickness
of a turbulent boundary layer increases more rapidly in flow direction than
that of a laminar one, as is illustrated in Fig. 3.3

8
TRANSITION

LAM I NAR T URBULENT

Figure 3.3: Development of boundary layer thickness

3.3.2 Effects of the Pressure Gradient.


In the case of a flat plate there is a constant pressure along the boundary
layer. This is not the case when the plate is at an angle to the flow or when
the plate has a thickness distribution. The effect of a pressure gradient on
the development of a boundary layer is very strong. A favorable pressure
gradient occurs when the pressure decreases in flow direction. Such a pres-
sure gradient reduces the growth of the boundary layer thickness, both for
laminar and turbulent boundary layers. It also delays transition to tur-
bulence of a laminar boundary layer. A favorable pressure gradient also
increases the velocity gradient at the wall and delays separation, a phenom-
enon which will be described below. Inversely an adverse pressure gradient
1It should be noted that transition is a very complicated process, which does not
occur at one location and in one moment. The description given here is a very strong
simplification of what really happens.
56 G.Kuiper, Resistance and Propulsion, January 4, 1994

causes a strong increase of the boundary layer thickness and stimulates


transition and separation.

3.4 Drag of a Two-dimensional Cylinder.


To illustrate some other floW phenomena a cylinder will now be used instead
of a flat plate. The drag coefficient of a cylinder on the basis of Reynolds
number is shown in Fig. 3.4 The Reynolds number of the cylinder is based

...
80
80 i li I 1 Ili
0 k nons!
40
CD
. 0.1
20 0.3
1.0 Ifeassfri
10
...., .. 10 8).
Z9 WieselsOarger
- 410
4 0 WO
300.0
2 -- Theoryllvelo lamb
-

/ .......44.4444-....4.00.0.4404.,,,,,
0.3
0.8
Oh

02
tee
0.1
1 L.
10- 10 10i 102 103 104
4 6 8105 2 8 8We

113,2,

Figure 3.4: Drag coefficient of a cylinder

on its diameter D(cyl) and the drag coefficient is the drag coefficient per
unit length
RT
d=
112pV2D(cyl)
The pressure p along the cylinder has been non-dimensionalized by the
stagnation pressure 1/2pV2. The stagnation pressure occurs where the ve-
locity on the cylinder is zero. The non-dimensional pressure coefficient is
expressed as the pressure difference between the local pressure p and the
undisturbed pressure pco.

At low Reynolds numbers the drag coefficient is high and it decreases


gradually to one at ./77, = 1000. At a Reynolds number between 1000 and
2 x 105 the drag coefficient is constant with a value of approx. 1. The flow
January 4, 1994, Simple bodies 57

pattern in this range of Reynolds numbers is given in Fig. 3.5a for subcriti-
cal flow. The flow separates at a position close to the location of minimum
pressure. At a Reynolds number between 2 x 105 and 5 x 105 the drag

SUBCRMCAL FLOW

SUPERCRMCAL FLOW

Figure 3.5: Flow Pattern around a cylinder

coefficient drops drastically to a value of about 0.3. This is due to the fact
that separation is delayed , which causes a much smaller wake, as shown in
Fig. 3.5 for supercritical flow.

The wake behind the cylinder is die to separation. Separation occurs


when the velocity gradient perpendicular to the wall becomes zero, as illus-
trated in Fig. 3.6. As a result the friction becomes zero and downstream
of the separation a region with back flow occurs. The streamline along the
wall separates from the wall and becomes the boundary of the separated
flow region.

When the drag coefficient is around 1 the boundary layer on the cylinder
is laminar before it separates close to the location of maximum thickness.
58 G.Kuiper, Resistance and Propulsion, January 4, 1994

PO p2>pl p3>p2

SEPARATION

Figure 3.6: Velocity profiles in the boundary layer around separation

The result of separation on a cylinder is that the pressure at the down-


stream side of the cylinder remains lower than on the upstream side, which
causes an additional resistance. The pressure distribution over a cylinder
at low Reynolds numbrs is shown in Fig. 3.7 by the dotted line (laminar).
The location of laminar separation is independent of the Reynolds number
and consequently the drag coefficient remains constant. This condition is
called the subcritical condition.
For comparison the line indicated by potential theory 2 is the pressure
distribution without separation in inviscid flow. In that case the pressure at
the back of the cylinder recovers to the stagnation pressure and the pressure
distribution is fully symmetrical. As a result the resistance is zero.

At a Reynolds number of about half a million transition of the boundary


layer from laminar to turbulent flow occurs at or upstream of the location
of separation. When the boundary layer becomes turbulent before separa-
tion occurs the flow pattern changes also because the turbulent boundary
layer separates much later than the laminar one. As illustrated in Fig. 3.5b
this strongly decreases the width of the wake and increases the base pres-
sure, which decreases the drag. This is reflected in the base pressure on the
cylinder, as shown in Fig 3.7 (turbulent). This is called the supercritical
condition. (Note that there is a critical condition in Fig. 3.7 at an interme-
diate Reynolds number, where the base pressure is even higher than in the
supercritical condition. This complication will be ignored here.)
The reduction of the drag coefficient of a cylinder at a Reynolds number of
around 2 x 105 is therefore due to the transition from laminar separation to
turbulent separation.

2The meaning of potential theory is described in chapter 10


January 6, 1994, Simple bodies 59

FO" ENTIAL THEORY

locf

i
_

o
_

CRITICAL

x7.1
TURB
x---:
__.....
ENT
x
t -
ii
.. LAMI AR
.... ....x..\.. \''' ...
V\: v i %

\
x
Ix
1

\ .."

2 /. I 1
x\ .

x
x

3o 30 60 90 120 150 180 210 240 270 300 330 360

Figure 3.7: Pressure distribution on a cylinder, from Achenbach.

3.5 Drag Components.


The pressure over the flat plate is constant and the resistance is only due
to frictional forces. The force on the cylinder can be decomposed into
pressure forces perpendicular to the cylinder (the pressure from Fig. 3.7)
and friction forces parallel to the cylinder surface. The integration of the
drag component of the pressure forces is the pressure drag or form drag.
The integration of the drag coefficient of the friction forces is called the
frictional drag. These two drag components are not independent. In the
case of the cylinder they are strongly interdependent, because the frictional
resistance determines the location of separation and this location determines
the pressure drag. On more streamlined bodies like ship hulls the location
of (turbulent) separation is less dependent on the Reynolds number.

3.6 Additional References.


Drag coefficients of a sphere and of a flat plate perpendicular to the flow
can be found in Schlichting [42]. Experimental drag coefficients of a range
of shapes can be found in Hoerner [13].
60 G.Kuiper, Resistance and Propulsion, January 4, 1994

3.7 Additional Data.


Since the flat plate boundary layer is used under many circumstances some data of the
laminar and turbulent boundary layer at zero pressure gradient will be given below.
Laminar flow: The thickness of the laminar boundary layer 5 is

= (3.4)

So the boundary layer thickness increases with N/X. In non-dimensional notation the
boundary layer thickness can be written as Rn5 = 5 OR.7z. The displacement thickness
of the laminar boundary layer can be written as:
0.34
6* =

. The local friction coefficient is proportional to the slope of the velocity distribution
perpendicular to the wall. This local friction coefficient is
0.664
Cf = (3.5)

The local friction coefficient per unit surface is defined as

Cf =
D9V2

where F is the local friction force on a unit surface.

This translates into a drag coefficient of a flat plate of


1.328
= (3.6)

where the Reynolds number is based on the length of the plate.

Turbulent flow: An approximation of the velocity distribution in the turbulent


boundary layer derived from pipe flow is the so-called 1/7th power /aw. In that case the
velocity distribution in the boundary layer is always
T7_ T, (3.7)

The boundary layer thickness can then be written as


= 0.37xRz-4 (3.8)
This means that the boundary layer thickness increases with x.4 instead of with ,N,/ in
laminar flow. The turbulent boundary layer will therefore be thicker than the laminar
one.
The displacement thickness can easily be derived using the 1/7th power law to be 81 =
The local friction coefficient can be written as
C1 = 0.0576(Rn4r+ (3.9)
and the drag coefficient based on a plate of length 1 and unit width as

Cd = 0.072(Rn1)-1 (3.10)
January 4, 1994, Simple bodies 61

3.8 Summary.
The resistance of a body can be expressed in non-dimensional terms as the
resistance coefficient. The resistance coefficient or drag coefficient of a sub-
merged body is dep- endent on the Reynolds number only. The drag coefficient
of a flat plate depefids on the transition from laminar to turbulent boundary
layer flow. The loge. ation of transition depends on the local Reynolds number
and on the pressure gradient. The drag coefficient e.g. of a cylinder varies
with the Reynolds number depending on the type of separation which takes
place. Under the subcritical condition laminar separation occurs. In the
supercritical condition turbulent separation occurs. The total resistance of
a body can be divided in form drag and frictional drag.
Chapter 4
Resistance, Wake and Wake
Distribution
Objective: A description of the relation between hull form, resistance and
wake distribution

The resistance of the ship is caused by the flow around the hull and this
flow around the hull is also reflected in the wake of the ship. The wake is the
velocity distribution behind the ship hull. The wake is important because
its magnitude is related with the ship resistance and the wake distribution
is important because it is the inflow distribution of the propeller. When this
distribution is very non-uniform the propeller will cavitate more extensively
and more violently.

4.1 Resistance and Wake.


The resistance of a body can be related with the wake behind the body. To
illustrate this we use the laws of conservation of mass and momentum.
Consider a body,as in Fig. 4.1. The shape of the body is irrelevant. For
sake of simplicity the body is assumed to be two-dimensional, so the flow
is identical in every plane parallel to the drawing. A control volume is
defined with plane A (width 2a)upstream of the body, where the velocity
is V everywhere. Downstream a plane B (width 2b)is chosen, where the
velocity is u(y). The outer boundaries are streamlines, which means that
no fluid passes through these boundaries. The outer boundaries are taken
at such a distance from the body that the velocity near the boundary of
section B is equal to V. This makes it possible to assume a fluid pressure
Po everywhere over the control volume. (Note that this is not evident in the

62
January 4, 1994, Wake 63

Streaml me

PO

Figure 4.1: Control volume around a body

region where the -velocity u is smaller than V. It is an assumption, often


made for convienience!) This assumption means that there is no pressure
force acting on the control volume.
Assuming incompressible fluid the law of conservation of mass requires that
b
2a V= udy (4.1)
b
This relation will be used later.

The drag force RT on the body is related with the loss of momentum
over the control volume. Momentum is entering the control volume through
plane A. The volume per unit time entering plane A is 2aV. (Note that
this is per unit length perpendicular to the drawing, the dimension of the
volume is thus kg Isecm instead of kg Isec) Its momentum is p2aU2. In
plane B momentum is leaving the control volume. At an arbitrary position y
relative to the centerline a flow volume udy passes plane B. The momentum
leaving plane B can therefore be written as

Mout p fb u2dy

Since the drag force is the only force present and since there is no resultant
pressure force the drag is equal to the loss of momentum over the control
volume. So:
R = p2aV2 p u2dy
64 G.Kuiper, Resistance and Propulsion, January 4, 1994

Using eq. 4.1 this can be written as

R=
p b
u(V u)dy (4.2)

Since the integrand is zero outside the wake region (because U u = 0) the
choice of b is not important. So eq. 4.2 can be used over the wake region
only, where u U.1
A further simplification can be obtained when it is assumed that the wake
velocity u is not too much different from the incoming velocity U. In that
case (V u)2 0 and thus it can be derived that

u(V u) = V(V u) (V u)2 V(V u)

So eq. 4.2 can be written as

R = pV f!b(V u)dy (4.3)

So in this linearized case the resistance is directly related with the velocity
deficit behind the body. The velocity deficit is called the wake. Expressed
as a fraction of the undisturbed velocity it is the wake fraction .

4.2 Flow along a Ship Hull.


The flow along the ship will remain attached when the hull is well designed.
A boundary layer will develop from the bow to the stern. In the bow region
there is a favorable pressure gradient and the boundary layer will remain
thin. A typical difference_ with e.g. a flat plate is that the flow, and thus
the boundary layer, is not two-dimensional. This gives rise to cross flow.
In the stern region there is a strong adverse pressure gradient and the
boundary layer will become thick. The boundary layer in the stern region
will become so thick that its thickness is no longer small compared to the
ship length or breadth. Some form of separation may occur there.

In the following some three-dimensional effects will be described and


their effect on the resistance and the wake distribution will be described.

4.3 Cross Flow.


Consider a boundary layer on the ship hull. There is not only a pressure
gradient in the flow direction, but also in a direction perpendicular to the
1Zie diktaat SWO I, pag. 2-5
January 4, 1994, Wake 65

flow. As a result the velocity vectors in the boundary layer will not remain
in one plane, but will change direction towards the low pressure region when
approaching the wall. This is shown in Fig. 4.2. 2

Figure 4.2: Velocity vectors in a 3D boundary layer

The streamlines outside the boundary layer will therefore have another
direction than the streamlines at the wall. This fact is to be remembered
when paint is used at the surface of a model hull to find the direction of the
flow around the ship, e.g. for the application of fins or stabilizers.

VISCOUS REGION SURFACE OF


IN EXTERNAL SEPARATION (BUBBLE
STREAM

LINE OF SEPARATION

IMITING STREAMLINES

SURFACE OF SOLID BODY

Figure 4.3: Transverse separation

2A11 figures in the remainder of this chapter: courtesy of M.Hoekstra


66 G.Kuiper, Resistance and Propulsion, January 4, 1994

4.4 Separation.
Separation in a three dimensional space (3D) occurs in two different man-
ners.

The first is similar to 2D separation , where the flow velocity decreases


until zero and becomes ne-gative. This situation is shown in Fig. 4.3. The
separation line runs in transverse direction relative to the local flow and
downstream of the separation line a "dead water" region occurs. This type
of separation is called two-dimensional separation or transverse separation.
On ship hulls this type of separation has to be avoided, because it increases
the drag, just as in 2D flow on a cylinder. Regions where such unwanted
separation can occur are specifically the regions in front of or above the
working propeller, as shown in Fig. 4.4.

Figure 4.4: Transverse separation on a ships hull

In 3D the flow lines can also converge because the body becomes smaller.
In that case the fluid moves away from the surface simply because of the
law of continuity (Fig. 4.5). The flow lines in such a region will exhibit
January 4, 1994, wake 67

Figure 4.5: Thickening of stream tube in converging flow

a separation line in streamwise direction, as shown in Fig. 4.6. The flow

SURFACE OF
SEPARATION

VISCOUS REGION
STREAMLINE IN
EXTERNAL STREANI

Lai;gab:al:I. lII

LIMITING STREAMLINES

SURFACE OF SOLID BODY

Figure 4.6: Streamwise separation

at the separation line has a component both in streamwise and in normal


direction. The outgoing flow has the tendency to "roll-up" into a vortex.
The vorticity thus shed is lost energy and is felt as extra resistance.

Such separation lines cannot be avoided on ship hulls and the design
of a good ship hull is mainly the control of these separation lines, so that
the wake behind the ship remains small. The control of the vortices is also
important because it is a means to make the propeller inflow more uniform.
Two examples of 3D separation on the bow of a ship are shown in Fig. 4.7
for two different ships. In the first case the separated vortex remains at-
tached to the hull, in the second case the separation line rolls up and forms
a vortex, which in this case is a bilge vortex. Sometimes more than one
68 G.Kuiper, Resistance and Propulsion, January 4, 1994

SEPARATION LINE
ATTACHMENT LINE
LIMITING STREAMLINES

SEPARATION LINE

Figure 4.7: Longitudinal separation at bow

vortex with different signs are generated at various positions on the hull.

Separation at the bow, as shown in Fig. 4.7, is suppressed by proper bow


design. Such vortices are, however, stronger and nearly inevitable in the
afterbody region, because the adverse pressure gradient along the afterbody
stimulates boundary layer growth and separation. Typical are the vortices
originating from the bilge near the end of the parallel middle body, the
(bilge vortices).

4.5 The Wake behind Simple Ship like Bod-


ies.
Some simpler forms will be helpful for a good understanding of the wake
generation.
Consider a simple hull form with U-shaped frames , as in Fig. 4.8. The
water will be pressed sideways in this case, causing higher flow velocities at
the sides than under the keel. The pressure at the sides of the ship will be
lower than at the bottom. The flow will go from bottom to side and a bilge
vortex will form which rotates clockwise. This vortex shows up at the aft
January 4, 1994, Wake 69

cwL

5 4 3 2 1 BODY PLAN

CWL

, , .., /, N. \
, , / , ...- _ TUFT GRID

, , , , i // / /
-.... \
-"" 1 k
AT A.P.

/ / / / / / \ ./\ 4
t\
1
,
/ , / / f -1/4 .....

, . / / / 1/4 bt '1/4 '1/4 ..... - ,... .. I s ...

,,/ It% /

Figure 4.8: Wake behind a U-shaped hull form

perpendicular as shown.

Another simplified extreme form is the Pram-type hull shown in Fig. 4.9.
The water will be forced down and the lowest pressure occurs here at the
bottom, so that a vortex with counterclockwise direction is generated. The
vortex rolls up under its own induction and shows up in the near wake as
shown.

In these simple cases separation may be suppressed considerably by com-


bining U-shaped and Pram-type hull form, as shown in Fig. 4.10. In this
case the amount of energy which is left in the wake wil be minimal. No
separation occurs and the wake will be completely due to the velocity dis-
tribution in the boundary layer along the hull. This body will have the
lowest resistance because a minimum of energy has been left in the wake.
The main component of the resistance will be frictional resistance. The
form resistance will be small.
70 G.Kuiper, Resistance and Propulsion, January 4, 1994

CWL

BODY PLAN

CWL
S \ \ I i

k
g 4 / I
4

\
N.

/ TUF T GRID
AT A. P.

/ / I
/ / /

Figure 4.9: Wake behind a pram-type hull form

The effect of 3D separation on the resistance is large. In Fig 4.11 the


bilge radius is systematically reduced from model A to D, which means a
reduction of the strength of the bilge vortex. The corresponding resistance
curves show that the effect on resistance is considerable.

4.6 Horse-Shoe Vortices.


A different type of separation should be mentioned:horse shoe vortices.
These are formed when a strut or fin attaches to a surface while both sur-
faces have a boundary layer. In this case also 3D separation takes place,
but in a special way. A vortex developes around the front of the strut and
because of its form this is called a horseshoe vortex. An example of it is
given in Fig. 4.12. These vortices are also transported with the flow and,
when arriving in the wake, will further complicate the wake structure.
January 4, 1994, Wake 71

CWL

ROOT PLAN

CWL

///rttt
/ I

///////// / / I
/

1
I TUFT GRID
AT A.P.

/ / / / i1
/ ///1/1
/ /
1111,11111
o
r / / 1

Figure 4.10: Wake behind an optimized simple hull

4.7 Visualisation of the Flow around the


Hull.
The flow around the hull can be visualised using paint tests. Paint is applied
along frames of the hull, approximately perpendicular to the flow direction.
During a run at a certain speed the paint forms streaks in the flow direction.
Examples of such tests are shown in Figs 4.13 and 4.14 It should be kept
in mind that the direction of the paint streaks is the direction of the flow
at the wall. When a strong crossflow is present in the boundary layer the
direction of the streaks can be quite different from the flow direction outside
of the boundary layer. Paint tests are useful for the detection of regions of
separation, because that occurs in the boundary layer. Paint tests are also
used to determine e.g. the proper location of bilge keels, because in those
regions the cross-flow is generally low. It is very dangerous to determine
e.g. the strut orientation in the afterbody with paint tests, because of the
crossflow there.

In case of cross flow the use of tufts is better. The tufts are flexible
72 G.Kuiper, Resistance and Propulsion, January 4, 1994

RESISTANCE

MODEL A A
EN
Anil
t Inid0/
r 1"
o
_AEI
MODEL 8

-.. -

MODEL C

SHIP SPEED

MODEL ID

Figure 4.11: Effect of variation of bilge vortex on resistance

wires, mounted on top of a needle perpendicular to the hull. The wire


positions are photographed during a run with the model. The direction of
the wires indicate the direction of the outer flow. It is even possible to mount
wires at more than one position on the needle, so that the crossflow can be
visualised. The needles may disturb the flow by increasing the boundary
layer thickness, but that risk cannot be avoided.

4.8 Ship Wake.


The wake behind a ship generally has a complicated structure because it is
the result of the retardation of the flow in the ship's boundary layer and of
many separated flows around the hull. The wake behind a ship is generally
only measured in the propeller plane, which may be only a fraction of the
January 4, 1994, Wake 73

ip
1
SURFACE OF SEPARATION
AIPSEP.:RAT/N
STREAMLINE

4///
1A, LINES OF SEPARATION

UNE OF ATTACHMENT 401, SURFACE FLOVi PATTERN

Figure 4.12: Horseshoe vortex around a strut

Figure 4.13: Paint test on afterbody of a fast ship

total wake. The wake in the propeller plane without the propeller action is
called the nominal wake.
74 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 4.14: Paint test on the bow

4.8.1 Representation of the wake


The representation of the wake in the propeller plane is done by representing
the axial,tangential and radial velocity components separately. An exam-

....,

rill .1.039

illKI,
..011
0.672

o
o 45 90 135 18d
POSITION ANGLE

Figure 4.15: Example of axial wake distribution

pie of the axial wake distribution as a simple diagram is given in Fig 4.15.
The axial velocities are expressed as a fraction of the ship speed. A rather
complicated wake peak is present in this figure in the top position of 180
degrees. The radial and tangential components of the wake can be plotted
in a similar way.

Another way to plot both axial and tangential components of the wake
January 4, 1994, Wake 75

170 /so 190


160 200

100

80

Figure 4.16: Example of Plotting the Wake

is shown in Fig. 4.16. The axial wake distribution is given as contourplots,


in which the low speed regions reflect the core of the vortices in the wake.
The tangential flow velocities are plotted as a vector diagram, in which the
same vortical structures should be visible.

4.8.2 Relation between hull form and wake distrib-


ution
The relation between the shape of the hull and the wake structure is compli-
cated. In Fig. 4.17 three variations of the same afterbody are shown. The
wake structure in the propeller plane is shown in Fig. 4.18. In general a hull
shape with small curvatures will shed few separated vortices and will gen-
erate a smoother wake. This is the case with a V-shaped hull form, which
shape will have the lowest resistance. However, the non-uniformity of the
flow in the propeller plane is large and a large portion of the wake passes
outside the propeller plane, which decreases the total efficiency, as will be
discussed later. A U-shaped hull form has more longitudinal separation and
therefore has a more uniform propeller wake, since the boundary layer from
the ship is rolled-up into the propeller plane. This U-shaped hull form will
have a higher resistance, but the interaction with the propeller may offset
this, as will be discussed later. A further increase of the uniformity of the
76 G.Kuiper, Resistance and Propulsion, January 4, 1994

V STERN
V - STERN
SULSOUS STERN

Figure 4.17: Variations of the hull form

(25--7----f 050
030
-4,rzar.
, :LPN
ti,e4C3464
atellga
714.111:1.41korlititi;
uotgo.0
\ WOO1/4" 0.20

V- STERN BULBOUS STERN

Figure 4.18: Effect of hull form variation on the axial wake

axial wakefield can be obtained using a bulbous stern, where local separa-
tion lines from the bulb will roll-up an make the wake more uniform.

The determination of the ship's resistance and wake is generally done


by model experiments. Model tests do not always show clearly why some
forms are better than others and calculation techniques become available
nowadays to calculate certain aspects of the flow around the ship's hull.
January 4, 1994, Wake 77

Both experimental and calculation techniques are necessary to design an


optimum hull form.

4.8.3 Wake_ Fraction


As discussed in cha-pter 3 the velocity deficit behind the ship is a measure
for the resistance. It should be kept in mind that this is the case when the
wake is measured over a such a region that the velocity at its boundaries
is equal to the ship speed. This area is much larger than the area of the
propeller plane, in which plane the wake data of a ship are defined.

The velocity deficit in the propeller plane (without the propeller present)
can be integrated over the propeller plane. This results in an entrance
velocity V, in the wake.
127r fro
Ve = Ffir urdO clr

This velocity is- the average entrance velocity in the propeller plane when
the propeller is absent. When the propeller is absent this wake is called the
nominal wake. It is made non-dimensional with the ship speed V, as

Vs
Wn = (4.4)
Vs
This is the definition of the nominal wake fraction and it is the non-
dimensional form of the velocity deficit V, V, in the propeller plane. The
wake fraction determines the relation between the entrance velocity at the
propeller and the ship speed by the relation

= Vs(1 wri) (4.5)


In Fig. 4.19 some axial wake distributions are shown for a number of
stern shapes. The nominal wake fractions are also given.

The wake distribution is responsible for unsteadiness in the propeller


loading during one revolution. In this course only the average wake will be
used for the propeller inflow and unsteady effects will be neglected.

An effect of the definition of the nominal wake over the propeller plane
only instead of over the whole region of the velocity deficit behind the ship
is that the nominal wake fraction does not necessarily correspond with the
resistance. In case of a pram hull form, with a very flat afterbody, the
velocity deficit corresponding with the resistance is distributed over the
breadth of the hull and only a small fraction of the velocity deficit is found
78 G.Kuiper, Resistance and Propulsion, January 4, 1994

1/2
o
AoP AoP A

C L_
0.80
60

0.40
0,20

\ 010 080
0,10

070

0,50

0.30

0.10. 0,10

wn = 0,399 f. wn = 0,406 wa = 0,371


8L 8 81..

Figure 4.19: Examples of wake of various stern shapes

in the propeller plane. In that case a very small nominal wake fraction will
be found, although the resistance may be high due to. e.g. strong bilge
vortices which pass outside the propeller plane.

4.9 Design Considerations.


The frictional resistance is dominated by the wetted area of the ship. This
cannot be influence very much by the shape of the hull. So the ship hull with
minimum resistance will be the hull with the lowest form resistance. This
can be obtained by avoiding three-dimensional separation. In general this
means minimum curvature of the frames. V-shaped frames will therefore
January 4, 1994, Wake 79

have the lowest resistance.


As mentioned the resistance of the hull is reflected in the velocity deficit in
the wake. When the velocity deficit behind the hull can be concentrated
in the propeller plane, the propeller will accelerate the fluid again. This
reduces the losses due to hull resistance. U-frames and bulb sterns are used
to generate vortices in such a way that the velocity deficit behind the ship
is rolled up in the propeller plane, although this increases the resistance.
Chapter 5
Wave Resistance
Objective: Description of the wave system generated by a surface ship
and determination of interference of the various wave systems.

When a ship moves through an undisturbed free surface it generates


waves. These waves contain kinetic and potential energy which has to be
genereated by the ship propulsion system. In terms of forces the waves
result in a drag, which is called the wave resistance. To understand the
character of wave resistance some knowledge of basic properties of surface
waves is necessary. Unless mentioned otherwise it will be assumed that the
water is deep.

5.1 Surface Waves.

IA' "I) riZEC r1C) N V,,


c:2
-----C7--------E) k l*
0 0 Q C' l.
O 0 Q CO O
8 o Q o. 8
8 c 9 e

Figure 5.1: Motions of fluid particles in a wave

In a surface wave the fluid particles describe orbital motions. As shown


in Fig. 5.1 the radius of the orbital motion decreases with increasing depth.

80
January 4, 1994, Wave Resistance 81

When the wave height is small compared to the length of the wave and the
water depth is large compared to the wave length these orbital motions are
circles. The circles indicate the path of the fluid particles over time. The
particles move in clockwise direction (angular velocity w)and the wave crest
moves from left fo right (with a velocity viv). After a time T = 2r/w the
particles have comi5leted a full circle and the wave crest has moved over one
wavelength A. In-the crest the 'particles move with the wave direction, in
the trough the velocity is backwards. The average position of the particles
over time remains unchanged.. The resulting wave height is a sine function
and these waves are therefore called sinuoidal-wa ves . The wave form is a
balance between the gravity force and the centrifugal forces of the orbiting
particles and these type of waves is therefore called gravity waves .
These gravity waves have some specific properties.

5.2 Properties of Surface Waves.


5.2.1 The Dispersion Relation.
At the free surface the combination of gravity forces and centrifugal forces is
perpendicular to the surface. This means that a relation between the orbital
velocity w and the wave velocity vu, exists. This leads to the important
dispersion relation

(5.1)
An arbitrary wave in one direction can be decomposed into various si-
nusoidal wave components. These wave components have different wave-
lengths. The various components of the wave will travel with different
velocities and after some time the various wave components are therefore
found in different locations. This is the dispersive effect of the waves.

5.2.2 Energy in a Wave.


The energy in the wave consists of potential and kinetic energy. In the wave
crest the potential energy is highest. Averaged over one period of the wave
the energy density per unit area of a plane wave can be calculated as
1
Eu, = pgh,2 (5.2)
8

where h,, is the wave height from crest to trough. Note that the unit of E
is Nlm = NmIm2.
82 G.Kuiper, Resistance and Propulsion, January 4, 1994

5.2.3 The Group Velocity.


Consider a two-dimensional wave of one frequency, generated at the end of
a deep towing tank by a flap-type wave maker moving at a frequency w. in
practice wave generators generate not only the oscillation frequency of the
wave maker, but also higher frequencies. The wave velocity of the generated
wave is defined as A/T = t. In combination with eq. 5.1 this gives

2r g
w=
A

and the wavelength follows directly from its frequency.. The wave velocity
v is the velocity of the wave crests. However, at t seconds after the wave
maker has started the front wave in the tank is not at a distance of t x
from the wave maker, but only halfway that distance. This is because a
wave front moves with half the velocity of the wave crest. This can ea.sily
be observed at a wave front, where it seems as if the waves disappear when
reaching the front. The velocity of the wave front is called the group velocity.
It is the velocity with which the wave energy is transported. As will be
shown below this property is important for the wave system behind a ship.

5.3 The Kelvin Wave System.


The foregoing considerations can be used to describe the wave system
around a ship hull. To simplify matters the ship hull is considered as one
point, moving at a speed V.

In Fig. 5.2 the ship is- at position A at an arbitrary time t. A wave


system at an arbitrary direction 9 with the course of the ship moves with
the ship when the crest velocity is V/ cos O. After a time At the ship is in
position B. The wave crest at an angle O is still emitted. However, the wave
generated at time to at A in not in position C, because it is a wave front
which moves with the group velocity.. It is only in D, halfway of AC.

This mechanism can be generalized, because the location of C for waves


in an arbitrary direction is a circle through ABC, drawn as circle I in
Fig. 5.3. The location of the wave front in arbitrary direction after a time
At is a circle with half this radius,drawn as circle II. The wave fronts ra-
diated by the moving pressure point will therefore be restricted to a wedge
tangent to this smaller circle. It is a matter of geometry to derive that the
top angle of the wedge is 39 degrees. The waves with a crest tangent to the
lj
January 4, 1994, Wave Resistance 83

Figure 5.2: Single wave generated by a pressure point

Figure 5.3: Boundary of radiated wave system

wedge, in point D, have an wave direction of 55 degrees.

This leads to the wave pattern, known as the Kelvin wave pattern2 , as
shown in Fig. 5.4. The waves, radiated in the direction of motion are waves
with the longest wavelength. They form a transverse wave system behind
2After Lord Kelvin or Sir William Thomson, a British mathematician (1824- 1907)
84 G.Kuiper, Resistance and Propulsion, January 4, 1994

TRANSVERSE
WAVE CRESTS 1928'

DIVERGING WAVE CRESTS

Figure 5.4: The Kelvin Wave System

the ship, approximately perpendicular to the direction of motion. At the


same time a diverging wave system occurs, originating from waves radiated
sideways, which have a lower crest speed (see Fig. 5.2) and thus a shorter
wavelength. The crest of a diverging wave is hollow, as shown in Fig. 5.4.
The frontal envelope of these diverging waves makes the distinct angle of
19.5 degrees with the path of the ship and the diverging waves near the
boundary have a direction of 55 degrees.

5.3.1 The Froude Number.


When a body with length L1 generates a certain wave system, the same
pattern can be generated by a geometrically similar body with length L21
when the ratio of the generated waves to the length of the body is the same.
Since the radiated wavelength is proportional with V2/g, this means that
the wave system of the two bodies is similar when V2 I(gL) is the same.
The square of this ratio is written as
V
Fn = (5.3)
\fir,
and is called the Froude number,, after R.E. Froude who has first used
it. Two wave systems are similar when the Froude number is the same.
This is important for model testing of ship hulls.

5.3.2 Resistance due to a Kelvin Wave System.


The wave resistance of a pressure point moving at speed V can be found
from an integration of the wave energy passing through a control plane at
January 4, 1994, Wave Resistance 85

some distance behind the pressure point perpendicular to its path. This
wave energy flux is equal to the resistance R times the velocity V. The
resistance can be written in a relatively simple form as

R= 17{-pV2 1712 h(9)2 cos' OdO (5.4)


8 -/r/2

From this formulation it can be seen that the waves with 9 close to
90 degrees, which are the shorter waves in the diverging wave system, con-
tribute less to the resistance than the longer waves in the transverse system.

The Kelvin pressure point has no length. A body with a certain length
scale however prefers to generate waves with a wavelength of its own length.
This length can be compared with the wavelength of the longest wave
= 2T-(/2)/g. When the body length L is small relative to the maxi-
mum wavelength the waves are primarily radiated in the direction with a
large angle O. The ratio L/A is inversely proportional with the square of
the Froude nuMber, so this occurs at high Froude numbers. Inversely when
L/A is large the Froude number is low and the radiated waves tend to be
dominated by the longer waves, of which the crest has a small angle with
the path, the transverse waves. So at low Froude numbers the transverse
wave system dominates, at high Froude numbers the divergent wave system
dominates.

5.3.3 The Wave System of a Ship.


A ship hull can be considered as a system of pressure points, each generating
their own Kelvin -wave system. When the wave height is small the effects
of the various pressure points can be superimposed. The total wave system
of a ship has characteristics as in Fig. 5.5. Two major wave systems can
be distinguished, one at the bow and one near the stern. The wave system
of the bow generates the wave profile along the hull. The dominance of
the transverse wave system at low Froude numbers is illustrated in Fig. 5.6.
The dominant diverging wave system at high Froude numbers is illustrated
in Fig. 5.7.
From eq. 5.2 it follows that the wave energy per unit surface area is
proportional with the square of the wave height h. The energy contained
in surface waves over a distance x behind the ship will be proportional with
xxbxh,v2, where b is the breadth of the wave system behind the ship. This
is the energy necessary to overcome the wave resistance Ru, over the same
distance x, so
R.x =
86 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 5.5: Wave system of a ship

Figure 5.6: Transverse wave system along the hull

Assuming that b is proportional to the wave length A, the wave resistance


is proportional to Alz,,,2 . Since A is proportional with the wave velocity, tit
(eq. 5.1), which is proportional to the ship speedV,, the wave resistance will
be proportional with V32 x h. The wave height hu, is related with the pres-
sures, which according to Bernoulli are proportional with V32. So ultimately
January 4, 1994, Wave Resistance 87

"":1*.

Figure 5.7: Wave pattern at high speed

the wave resistance of the ship will be proportional with V96. This extremely
high power is the reason that ships travel at relatively low speedg compared
with e.g. cars. At high speeds the wave resistance becomes prohibitive.

The wave energy of the wave systems generated by a ship can be mea-
sured by measuring the contours of the waves passing through a control
plane aside of and behind the ship. This method is called a wake scan . It
requires a complete formulation of the radiated waves, however, to analyse
a wake scan and to determine the wave resistance. The analysis is compli-
cated and time-consuming. The method is therefore not commonly used.
The determination of the wave resistance of a ship hull by experiments will
be described in chapter 7.

The wave resistance is made non-dimensional by

CW 1/2pV2S (5.5)

Here S is an arbitrary surface. The area S is taken as the wetted surface of


the ship. Since the wave resistance varies approximately with V6 the wave
resistance coefficient varies with V4 or, in non-dimensional terms, with Frl.
88 G.Kuiper, Resistance and Propulsion, January 4, 1994

5.4 Wave Interference.


As can be seen in Fig. 5.5 the wave system of a ship is composed of different
Kelvin wave systems. Grossly simplified the wave system of the ship can be
considered as a wave system at the bow and one at the stern. The trans-
verse waves generated at the-se two locations will interfere. When the crest
of the bow wave coincides with the crest of the stern wave, the two wave
systems will reinforce each other. (The two wave systems are in phase).
When the two wave systems are exactly out of phase and of equal strength
the wave system behind the ship will disappear and the wave resistance will
be zero. This illustrates that the interference of the different wave systems
generated by the ship hull is important for the wave resistance.

Since the wave length depends on the ship speed the wave height behind
the ship varies with the speed. Consequently also the wave resistance of the
ship as a function of the speed has a wavy character (oscillates), with "hol-
lows" and "humps", as shown in Fig. 5.8. The humps in Fig. 5.8 seem rather

2oone

Hum!)

.S
15000
...
a.
.-.
..7.
Op
CC
a

10000

t
I

Horn !
Hollow
5000

Hollow

o I

10 15 20 2 30
Stn0 soeed ,r. knots

Figure 5.8: Wave Resistance of a Ship Model

shallow because the wave resistance increases rapidly with increasing speed.
The wave interference becomes more evident when the wave resistance is
plotted non-dimensionally as the wave resistance coefficient C versus the
Froude number, as in Fig. 5.9.
January 4, 1994, Wave Resistance 89

MODEL DIMENSIONS
471.8S' O 10 ',-
PRISMATIC COEFFICIENT 0.636
MODEL SYMMETRICAL
ABOUT AMIDSHIPS

0IS 61.1 - 1{7 o c.6 o


Et;

Figure 5.9: Example of Wave Resistance

5.4.1 A Two-dimensional Simplified Hull Form.


The location of humps and hollows can be approximated using a simpli-
fied hull form, which exaggerates the wave systems generated by a normal
ship hull. Such a body was used by Wigley (1930) for calculations of the
wave system. The hull is two-dimensional and consists of a parallel middle
body and two similar wedges, as shown in Fig. 5.10. Four wave systems
are present in an extreme form: the bow wave, the shoulder waves at fore-
and afterbody, and the stern wave. The transverse wave component of each
system and its location relative to the hull is also shown in Fig. 5.10 . Note
that the bow and stern waves begin with a crest, the shoulder waves begin
with a trough. This is understandable when it is realized that a shoulder
causes a low pressure region along the hull. The fluid particles tend to
move away from the curved wall due to the centrifugal forces, causing a
wave trough. On the other hand a bow has a stagnation point, where the
pressures are high and a wave crest is formed.

The superposition of the wave systems in Fig. 5.10 has a clear resem-
blance with the measured wave system, showing that the error due to su-
perimposing is small.
90 G.Kuiper, Resistance and Propulsion, January 4, 1994

bow for ward shoulder after shoulder --stern

symmetrical surface disturbance


(primary wave Tysteni)

bow wave system

forward shoulder wave system

after shoulder wave system

stern wave system


X

total wave system


N. (measured)

total wave system(calculated)


'o-'

Figure 5.10: Wave systems generated by a Wigley Hull

5.5 Economical Speed.


A merchant ship will be designed in such a way that it operates in a hollow
of the resistance curve at the design speed. In practice the design problem
is opposite, the desired economical speed is known and the ship has to be
designed in such a way that it speed is in a hollow. Such a speed is called
an economical speed . This means in the first place a proper choice of the
ship length.
January 4, 1994, Wave Resistance 91

In the past the two dominant wave systems discerned were the bow wave
and the aft shoulder wave. These two wave systems have an opposite sign
because the bow _wave starts with a crest and the aft shoulder wave with
a trough. Favorable interference occurs when the distance between both
wave systems is equ- al to k times the wave length A, where k is an integer.
In that case a trough of the bow wave will be at the crest of the stern wave.
The wave length of the transverse wave system can be found from eq. 5.1
to be A = 271-V2/g. When the distance between both wave systems is called
the wave making length L , it is required that

L(economical) = k2R-V2 (5.6)

The problem is, of course, to determine the wave making length, because
especially the shoulder waves are not a single wave system but a combination
of many systems distributed over the length of the hull. From Fig. 5.10)
the wave making_length can be assumed to be related with 2a + I . The
prismatic coefficient Cp of a Wigley hull is VI. So the wave making length
can be written as

Lw = CpLwi
This relation was used by Baker and Kent in 1913 (!) [2]. They con-
cluded from experiments that the wave making length could be approxi-
mated by
A
Lw = CpLw/ +
Using the latter wave making length in eq. 5.6, the requirement for
minimum wave resistance can be written as:

V2 2Cp
(5.7)
g Li(econornical) 7-(4k 1)
Note that the left hand of this equation is the square of the Froude
number, as defined in eq. 5.3.
The assumption that the bow and aft shoulder wave are dominant is true
for old hull forms, where a cruiser stern often had a distinct aft shoulder.
The modern hull shapes, especially those with a transom stern, have gentler
curved waterlines in the afterbody and the bow and forward shoulder wave
systems have become more important.

For slender ships the curve of sectional areas can be used as a measure
of the distribution of the wave systems. For full ships the shape of the
92 G.Kuiper, Resistance and Propulsion, January 4, 1994

waterline is more important.

The entrance angle of the waterline has been considered as a measure for
the strength of the bow wave system. However, at present the interference
between the faired bulb and the forward shoulder makes it possible to use
a higher entrance angle (4rid consequently a smoother forward shoulder)
in combination with a proper bulbous bow design. No rules of thumb are
applicable here. The approach from Baker and Kent as described above
is more to gain qualitative understanding than a rule for hull design. Op-
timization is only possible by experiments or by Dawson calculations, as
described in chapter 13.

5.6 Hull Speed.


When the wave length of the transverse waves behind the ship is equal to
the ship length length a certain limit is reached. There is a wave crest at the
bow and at the stern. Frofn eq 5.1 it is found that when A = L and vi = V,
the Froude number is equal to 0.4. When the length becomes smaller or
the speed higher the ship will trim considerably because it is at the rear
side of its bow wave. The resistance increases dramatically in that case and
for normal displacement ships the available power will not be enough to
attain this condition. The Froude number at which this occurs is about 0.5
and the velocity at which this Froude number is attained is called the hull
speed . Such a situation can be approached in case of overpowered yachts
or tugs.(see Fig 5.11). The hull speed is the speed with the maximum wave
resistance, as is also shown in Fig. 5.9.

5.6.1 High Speed Ships.


High speed ships such as planing vessels will have enough power to overcome
the hull speed. For these ships the speed with the maximum wave resistance
coefficient is also called the hump speed . At Froude numbers above 0.5
the wave resistance coefficient decreases again, as illustrated in Fig. 5.9.
Above the hump speed the diverging waves dominate and these waves have
a smaller contribution to the resistance, as follows from eq. 5.4. Also the
extreme trim decreases again. The resistance at the hump speed is often
decisive for the power to be installed in high speed ships and for the propeller
design. Although at high Froude numbers the diverging waves dominate,
the transverse wave system is still there and the angle of the diverging waves
still have an angle of 19.5 degrees with the ship, as shown in Fig. 5.7. This
is contrary to the situation at restricted depth, as will be discussed later.
January 4, 1994, Wave Resistance 93

atitiNPA

"zz
4i"t4t,
4

Figure 5.11: Wave system close to the hull speed

5.7 Bulbous Bows.


The application of a bulbous bow is equivalent to the generation of a sep-
arate wave system. If the bulbous bow is at the correct position relative
to the bow it will generate a through at the location of the bow, which
partially cancels the bow wave. Of course this is a conceptual simplification
because the wave systems are not single systems and the wave height is con-
siderable, so they cannot simply be added. Since a favorable interference
of two wave systems occurs only at one speed, the bulb is best for only a
limited speed range. At other speeds the interference effects can even be
opposite and an increase in wave resistance may occur.

5.8 Shallow Water Depth.


With increasing depth the wave height decreases rapidly. For a sinusoidal
wave with small amplitude as described above the decrease is

h(y) = hwe-21rYIA (5.8)


94 G.Kuiper, Resistance and Propulsion, January .4, 1994

When y/A = 0.3 the reduction factor of the wave height is already 0.15. The
energy of the waves at this depth is, according to eq. 5.2, only 2.5 percent
of the wave at the surface. For practical purposes there is deep water when
the water depth is greater than one third to one half of the wavelength, so
h/A <3. When the water depth is less the water is considered shallow and
other effects become important.
In shallow water the circular fluid motions of Fig. 5.1 become elliptical
with the largest chord in horizontal direction. At the bottom the motions of
the fluid particles reduce to periodic rectilinear motions along the bottom.
The most important effect of restricted water depth is that it imposes an
upper limit on the wave velocity. Instead of eq. 5.1 the velocity of a wave
does not depend on its wavelength but on the water depth:

gh
V= (5.9)
27r

As long as the ship goes much slower than the wave velocity in restricted
water the wave system will be the same as in unrestricted water. When the
ship approaches the maximum wave velocity in restricted water, the radi-
ated waves will have the same speed as the ship. A phenomenon similar
to a shock wave in compressible flow will occur. All radiated waves will
have the same speed as the ship. As a result the crest of the wave system
will become perpendicular to the path of the ship. The trim will increase
strongly. This condition is independent of the length of the ship, but the
critical speed in shallow water depends on the Froude number based on the
water depth VA/FE (from eq. 5.9.

Above the critical speed a transverse wave system cannot exist because
the waves cannot stay with the ship. The angle of the diverging wave system
will decrease with increasing speed, similar as a Mach cone in supersonic
speed (Fig. 5.12).
January 4, 1994, Wave Resistance 95

r-

Figure 5.12: Wave pattern generated at supercritical speed


Chapter 6
Intermezzo: Scaling Rules
Purpose: Introduction of scaling laws and the relation with non-dimensional
parameters

For model tests some special rules called scaling laws have to be obeyed.
The problem is the same when models at different scales are compared,
because each size of model with the same hull-form has a different resistance.
As has been seen before the resistance can be made non-dimensional as

CT = (6.1)
V32 S

where R=resistance in N, p= specific mass of water in kg1m3,V3 =Ship


speed in mlsec and S= the wetted surface of the ship in m2.

To represent the frictional resistance coefficient as a function of the ship


or model speed in one single curve, as for a simple body, the velocity V, has
to be expressed non-dimensionally too, as the Reynolds number fin:

(6.2)

Similarly the wave resistance coefficient reduces to one function for all
sizes of ships when plotted as a function of the Froude number Fn:

(6.3)

Until now these parameters have been determined based on experimental


evidence. They can also be derived by dimension analysis.

96
January 4, 1994, Scaling Rules 97

6.1 Dimension Analysis.


A general method to derive non-dimensional parameters by dimension analy-
sis is by using the the so-called H-theorem. 1 This theorem states:

The number of non-dimensional parameters is equal to


the number of parameters involved minus the number of
dimensions involved.

The parameters are the physical quantities like length, mass, viscosity, com-
pressibility etc. As an example the parameters viscosity, mass and gravity
are now considered, because these have been used in the definition of the
Froude and Reynolds number.

The resistance R can be written as


R = f (v, V, L, g)
in which

v=kinematic viscosity in 7.722 I sec

V=velocity in m/sec
R=resistance force in N
L=length scale, e.g. the length of the ship in m.
g=acceleration due to gravity, 9.8 m/sec2
There are five parameters involved in this equation. The number of di-
mensions in the foregoing parameters is 3 (N,m,sec). The PI theorem states
that there are now 5-3=2 non-dimensional parameters ri which determine
the problem.

The two unknown parameters are found as follows. Any non-dimensional


parameter II can be written as a product of the parameters involved:

= RavbgcvdLe
This product has to be non-dimensional, so the sum of the exponents
has to be zero for each dimension involved. The dimensions of the parame-
ters R,V,g,v and L in 11 are:

1See SWO 1
98 G.Kuiper, Resistance and Propulsion, January 4, 1994

m 0 1 1 21
sec 0 -1 -2 -1 0
N 1 0 0 0 0

When H is to be non-dimensional the sum of each dimension has to be zero.


This leads to the following equations:
For the dimensions m:

b+c+2d+e=0
For the dimension sec.:

b 2c d=0
For the dimension N:

a=0
This gives three equations with five unknowns. We can therefore express
e.g. a,b and e in c and d. When that is done the result is:
b= (2c d)
a=0
e= c d
The non-dimensional parameter II can now be written as:
Rov(2c+d)gcvd.u_d

This can be rewritten as


g v
H (L172)c(Vdd
In the first factor the Froude number can be recognized, in the second
factor the Reynolds number.

In the foregoing the resistance was considered to be a function of four


parameters only. In model testing other phenomena such as cavitation, also
can play a role and these parameters also have to be properly scaled. In
case of cavitation e.g. the deviation of the pressure from the vapor pres-
sure is an additional parameter. In propulsion tests additional parameters
are the length scale of the propeller and the number of propeller revolutions.

In the following table a number of parameters which can play a role are
listed, together with the non-dimensional scaling rule which is added by the
January 4, 1994, Scaling Rules 99

parameter nomenclature dimension non-dim. param. name

Ship length L 771

_
kg
Fluid mass p m3

Velocity V .1mec

Pressure N _P___
P 727 pV2 Euler number

Viscosity v in 2 Rn VL
Reynolds number
sec v

Gravity m
g sec2 Fn -- \riV Froude number

Propeller diameter D 772


L
75 geometric similari

Propeller revolutions n sec-1 j v.


nD Advance ratio

Cavitation N a= PO -Pv
Po pe, m2 1/2042 Cavitation numbe

Surface tension s N We =
m PV2L
3
Weber number

Compressibility c Tn
sec M =E C
Mach number
2
Diffusion d m
sec
VD
d Peclet Number

Nuclei diameter d m cin.


L geometric similari

Gasconcentration C lig_ c
in3 P

Table 6.1: Scaling Laws

inclusion of such a parameter in the scaling problem. The basic dimensions


are length, time and mass. Note that the time in the table is hidden in the
parameter velocity. This table is only a selection of all possible parameters.
In unsteady problems and in cases with heat transfer other non-dimensional
numbers appear.
100 G.Kuiper, Resistance and Propulsion, January 4, 1994

6.2 Physical Meaning of Non-dimensional


Parameters.
The non-dimensional para:meters represent the ratio between two different
physical phenomena. E.g. the Reynolds number has something to do with
the viscous forces in the fluid. The Froude number with the gravitational
forces in the fluid. The third force component of the fluid forces are the
inertia forces, which cause the pressure variations when the velocity or its
direction is changed. It can easily be understood that these forces should
be kept in the same proportion when the scale is altered. From the physical
definitions it follows that:

the inertia forces are proportional with mass times acceleration. For
mass we can write p x P. Acceleration is V/t, in which t is the time.
This results in pl3VIt. The ratio Ilt is the velocity V, so inertia forces
have the dimension pV212

the frictional forces are proportional with the velocity gradient VII
times the area /2 times the dynamic viscosity p, so these forces have
the dimension pV1.

the gravity forces are proportional with mass times gravity pl3g.

Here p is the dynamic viscosity in Nseclm2. The dynamic viscosity is


related with the kinematic viscosity v by v =
It can be seen that the Reynolds number is the ratio between the inertia
forces and the viscous forc-es in the fluid:

pV2/2 VI
pV1 v

Similarly the ratio between inertia forces and gravitational forces gives
the Froude number:
p v 2 12 v2
=
pg13 gl
The physical meaning of the non-dimensional parameters makes clear
what the magnitude of such a parameter means. A high Reynolds number
mean.s that the dynamic forces dominate and that viscous forces can be
neglected. A high Froude number means that gravitational forces play a
minor role. The physical meaning of the non-dimensional parameters is also
important for scaling, as will be explained below.
January 4, 1994, Scaling Rules 101

6.3 Scaling Rules.


When two models of different size are compared, the comparison should be
made in similar condsitions. This means that the various forces involved
should be kept in- the same proportion, just as with geometrical scaling,
where the dimensi6ns in three different directions are kept in proportion.
The situations are- similar when the non-dimensional parameters are kept
the same. The non-dimensional parameters therefore act as scaling laws.
Each non-dimensional parameter leads to a scaling law. Maintaining the
Froude number at model scale e.g. means that
Vs V,
VgLs= VgL,
where the index m is for the model, the index s is for the ship. This gives
a prescription for the model test conditions:

V, = V,
Here a is the scale ratio. In such a way the model test conditions
can be derived from the full scale conditions using the scaling rules. The
maintenance of the scaling laws guarantees similarity of the flow pattern at
model and full scale.

6.4 Scale Effects.


It is not always possible to maintain the scaling rules. This is e.g. the case
for the Reynolds number, as will be shown in chapter 7. The deviation
of the scaling law results in dissimilarities between model and full scale.
These dissimilarities are called scale effects. The physical interpretation of
the non-dimensional parameters can be used to predict the effect of dissim-
ilarities or scale effects. A deviation in the Reynolds number e.g. will result
in differences in the regions where viscosity is important, so in the bound-
ary layer. A deviation of the Froude number will result in a deviation in
similarity of the wave system. Scale effects are a major source of problems
in model testing.
102 G.Kuiper, Resistance and Propulsion, January 4, 1994
Chapter 7
Resistance Prediction using
model tests
Objective: Prediction of ship resistance from model test results.

7.1 Elements of Ship Resistance.


The resistance of a ship is the force required to tow the ship at a specified
speed. Note that this is a condition without a propulsor.

Three components can be distinguished in the total resistance. The first


component is the frictional resistance. The frictional or viscous resistance
of a ship is due to the tangential forces along the hull. The pressure resis-
tance of form resistance is due to the pressure forces on the hull. These
.

definitions are valid when no waves are generated at the free surface. At
Froude numbers above 0.1 the waves cannot be neglected any more. The
major effect of the waves is on the pressures on the hull, although indirectly
the boundary layer, and thus the frictional resistance, is also affected by
the waves.

It should be mentioned that William Froude distinguished only the fric-


tional resistance and the residuary resistance . The latter component com-
prised the wave resistance as well as the form resistance.

The distinction between the resistance components is important for re-


sistance tests at model scale. A resistance test is a test where a model hull
is towed in a towing tank or model basin. During these tests the resistance

103
104 G.Kuiper, Resistance and Propulsion, January 4, 1994

force as a function of the model speed is measured. The question now is


how to predict the resistance at full scale from the model test results. This
process is called the extrapolation of model test results to full scale. The
simplest rule is to determine non-dimensional coefficients from the model
tests, such as a resistance- coefficient, and to apply that coefficient to full
scale. However, the deviatfon from the full scale Reynolds number makes
this approach incorrect ana in the nineteenth century the results of model
tests were useless because no relation between model test results and the
performance of (sailing) ships at full scale could be found.

Model tests became useful only after the different components of the
resistance were distinguished, because these components have to be extrap-
olated to full scale in different ways. This is because the different resistance
components are governed by different scaling laws.

7.2 Scaling Laws for Model Tests.


Assume that the resistance of a ship depends only on the viscosity of the
fluid and on the waves generated at the free surface. This assumption
excludes e.g. the effect of surface roughness, of wave breaking (surface ten-
sion). The non-dimensional parameters controlling the resistance are then
the Froude number and the Reynolds number. In chapter 6 it was shown
that the non-dimensional resistance coefficient at any scale remained the
same for model and ship when those numbers were kept the same at model
and full scale.

When the Froude number is maintained the following relation exists


between model parameters (index m) and the same parameters of the ship
(index s).

TI T42
(7.1)
gLyn= gL,
When the Reynolds number is maintained this means that

VL Vs.L,
(7.2)

Writing V, from eq. 7.2 as

V, = (V3L3)1L (7.3)
and substituting this in eq. 7.1 results in
January 4, 1994, Resistance Tests 105

Ls3
.L3 1

This means that both scaling laws can only be maintained when the test
is done at the same scale as that of the ship. In other words:

It is impossible to maintain both the Reynolds number


and the Froude number simultaneously in model tests.

In practice a choice has to be made. At least one of the scaling laws


has to be abandoned. When the Reynolds number is maintained the model
speed is found from eq. 7.3 to be V, x a where a is the scale ratio Ls1
The required model speed is very high indeed and in most towing tanks this
is unattainable.1 At these high speeds the waves in the towing tank would
be extreme because the Froude number is not maintained. No method is
available to correct the model tests results for the deviations of the Froude
number, so in praCtice the Froude number is maintained and the effects of
the Reynolds number deviation at model scale requires corrections. These
corrections are the main complications in the extrapolation of model data
to full scale predictions.

So in resistance and propulsion tests at model scale the Froude number


is maintained. The model velocity thus becomes:
V3
Vm =

The Reynolds number of the model is much lower than that at full
scale,since
Rns= a 1'5 Ram
So the Reynolds number at model scale is too low. Since the Reynolds
number is the ratio between the inertia forces and the viscous forces this
means that at model scale the viscous forces are too large. The viscous
forces are dominating in the boundary layer along the hull and at model
scale the boundary layer is too thick. Phenomena which are controlled by
the Reynolds number, such as transition from laminar to turbulent flow
and flow separation are different between model and full scale. The main
problem in extrapolating model test results to full scale is in the assessment
of the scale effects which occur due to the improper Reynolds number at
model scale.
lOnly in some cavitation tunnels, where no free surface exists, this goal can be
achieved
106 G.Kuiper, Resistance and Propulsion, January 4, 1994

7.3 Froudes Hypothesis.


It has been the merit of William Froude to distinguish the several compo-
nents of the hull resistance and to relate these components to the scaling
laws. He distinguished between the wave resistance and the residuary resis-
tance, which is the sum of form and frictional resistance. Then he made a
drastic simplification, which has worked remarkably well. Froudes hypoth-
esis was that

The components of the resistance are independent of


each other.

Having made this basic assumption it follows that the viscous resistance
depends on the Reynolds number only and that the residuary resistance
depends on the Froude number only. As can be seen in case of a cylinder
this is a gross simplification, because even without free surface, the relative
contribution of the form -resistance and the viscous resistance at various
Reynolds numbers differ considerably. However, when there are no large
separated regions in the afterbody, Froude's hypothesis is more accurate.
For ship hulls large separated regions are undesirable, so for a good hullform
Froudes hypothesis can be used with reasonable accuracy. This restriction
should be kept in mind, however, when large separated areas do occur, e.g
when offshore constructions are towed or when relatively blunt shapes such
as tunnels and thruster shafts are investigated.

7.4 Determination of Resistance Components.


The determination of the various resistance components of a ships hull can
be illustrated with the results of resistance tests with a series of models
at various scales (The "Simon Bolivar" family of models). For each of the
models a resistance test was carried out over a certain speed range. The
total resistance in non-dimensional form is shown in Fig. 7.1. Just as in the
case of simple bodies the resistance force of a ship is made non-dimensional
as

RT
CT .ipv32s (7.4)

where R.resistance in N, p= specific mass of water in kg/m3, Y,=Ship


speed in m/sec and S. the wetted surface of the ship in m3. The precise
determination of the wetted surface will be discussed later. When all scaling
laws would be fulfilled the resistance curves of all models would coincide.
January 4, 1994, Resistance Tests 107

\
9000
LAP-TROOST EXTRAPOLATOR

(X -144

IN!
9000

IlL 1 \\
. .84

II
7000

<X.= so

V
ji
6000
- =38
,,,
o 0(.25
i-- -
Ci 1-I"
cX. 21

5000
1
IL A.,,,,.
0(.180(=15

SCHOENHERR MEAN LINE 10, 1


4 000
741,139
NIIIIIIIIIIkglr i

-...,...

3000 -
5.0 5 S--...log Ret 6.0 65 7,0

Figure 7.1: Resistance Coefficients of the "Simon Bolivar" Model Family

Because the Reynolds number is not maintained this is not the case. The
dotted line connects tests at the same Froude number. At other Froude
numbers similar lines can be drawn, with in the limit the line for F 0.
This line is also shown in Fig. 7.1.

Fig. 7.1 shows that the resistance coefficients of various model sizes col-
lapse into one line for one Froude number. The explanation is that a part
of the total resistance coefficient is due to wave resistance and this wave
resistance coefficient is the same for all models at the same Froude number.
108 G.Kuiper, Resistance and Propulsion, January 4, 1994

The wave resistance coefficient is the vertical distance of the total resis-
tance above the line for Fn = 0. Consequently the resistance coefficient
below F, > 0 is the residuary resistance coefficient, which is due to form
resistance and frictional resistance and which coefficient depends on the
Reynolds number.

7.4.1 Determination of the Frictional Resistance.


To determine the frictional resistance coefficient Froude assumed that the
residuary resistance coefficient was related to the resistance coefficient of a
flat plate with the same length and wetted surface as the model hull. So
he did numerous experiments to determine the resistance coefficient of flat
plates as a function of Reynolds number. Froude himself did not arrive
at one single line on the basis of Reynolds number due to laminar flow
and edge effects in his measurements. His results depend not only on the

ill
Reynolds number, but also on the length of the plate, as shown in Fig. 7.2.
Friction lines on the ba-sis-of Reynolds number were developed later, both

0,006
1
0.005

0.004 is N,
\\ Froude
= 0.25m

1\....\ic. .
.0.50m
.400110). . 1.00m
0.003 ..../B .2.50m
.7,0ro-- :vsT,
...,...k .0. =25,0m

11
.50.0m
-4%. 1111 Irel II = 100m

0,002 .
,4..,, :omm

Schoenherr

J \
109 106 107 108 108 101
Ree

Figure 7.2: Frictional resistance coefficients according to Froude

theoretically from boundary layer theory and experimentally. These plate


lines are for turbulent boundary layer flow from the leading edge on and
extend until Reynolds numbers as occur at full scale. An example is the
Schoenherr mean line ,which is also shown in Fig. 7.2 as well as in Fig. 7.1.
January 4, 1994, Resistance Tests 109

Another is the ITTC plate line. 2 These plate lines have relatively simple
formulations. The Schoenherr mean line is formulated as:
0.242
= log(R7, x C f) (7.5)
\IGtf

The ITTC 1957 line is defined as:


0.75
Cf (7.6)
(logio 2)2
As a matter of fact it is not important if a flat plate with a certain wet-
ted area has a resistance coefficient according to the mentioned lines. These
lines, especially the ITTC plate line, have been adjusted in such a way that
the best correlation with full scale results was obtained in extrapolations as
will be described below.

Having assumed that the wave resistance and the residual resistance
are independent, the next assumption made is that the frictional resistance
coefficient of the ship is equal to the resistance coefficient of one of the plate
lines.

7.4.2 Determination of the Form Resistance.


From the frictional resistance coefficient, as determined with a plate line,
the form resistance or pressure resistance can be determined experimentally
as the difference between the line Fr, = 0 in Fig. 7.1 and the plate line. In
Fig. 7.1 the resistance coefficients of many geosims are available. In case
of a resistance test from one model only one resistance curve is available at
the model Reynolds numbers. An additional assumption is then required
about the relation between the frictional resistance and the form resistance
to derive the form resistance at full scale from that at model scale.

Several hypotheses have been made. Froude assumed that the form re-
sistance was independent of the Reynolds number, as shown in Fig. 7.3.
He therefore took it as a part of the residuary resistance. Lap took the
Fr, = 0 line to be the plate line after being horizontally shifted over a con-
stant value, based on an analogy with pipe flow. This is shown in Fig. 7.4.
Hughes took the form drag as proportional to the viscous resistance, mul-
tiplying the viscous resistance coefficient Cf with a constant factor k, as
shown in Fig. 7.5. The Reynolds dependent component of the resistance
thus becomes (1 + k)Cf. The factor 1 -I- k is called the form factor. One
of these assumptions has to be used to derive the form resistance at the full
2International Towing Tank Conference
110 G.Kuiper, Resistance and Propulsion, January 4, 1994

[
FROUOC CXTRAPOLAMN HENN

. - .

EMMA=
111M-444=11
111111111- -

Figure 7.3: Definition of form resistance according to Froude

scale Reynolds number from that at model scale. The method of Hughes
or the form factor method is part of the ITTC 1978 extrapolation method
and is most widely used nowadays.

The determination of the form factor of the model from the point at
the lowest Froude number of the resistance test is often difficult, because
in that condition the speed of the model is low and the forces are small.
Consequently the measuring errors are relatively large. It is possible to use
resistance data when the wave resistance is not zero by using the knowledge
that the wave resistance coefficient without interference effects is propor-
tional to F:, as inentioned in chapter 5. The total resistance coefficient can
than be written as
Ct = (1+ k)Cf cF4
Or
CT Fn,4
= (1 IL- k) c
f T

When this ratio acf is plotted versus F7`,' the curve is straight at Fn =
and the constant c can be determined using data over a larger Froude range.
(This approach is sometimes called Prohaska's method). The difference
between the direct approach and the regression method is illustrated in
Figs. 7.9 and 7.10 of the example below.
January 4, 1994, Resistance Tests 111

LAP-TROOST EXTRAPOLATION METHOD

A- - -

Will j
.
.

allf-
. .

--- -

plate friction coefficients


- - - - - 4 c a I.
P.IIIIIIIIIIIIIIII'""
mar:Ans

- g R.1

Figure 7.4: Definition of form resistance according to Lap-Troost

-.. HUGHES EXTRAPOLATION NE1I400


--..

e !!
plea friction coefficients

1
I

1
14. -""siligieft
l

,
It 11 It
.AMINI
.IIIIIII

I iI ir

Figure 7.5: Definition of form resistance according to Hughes

Typical values of the form factor 1 k are given in Table 7.1


112 G.Kuiper, Resistance and Propulsion, January 4, 1994

CB 1+k
< 0.7 1.10-1.15
0.7-0.8 1.15-1.20
> 0.8 1.20-1.30

Table 7.1: Typical values of the form factor

The magnitude of the form factor is primarily dependent on the shape


of the afterbody, although Holtrop uses also a dependency of the Froude
number in his most recent regression model (see chapter 8).

7.4.3 Determination of the Wave Resistance.


The wave resistance coefficient of the model C, over the speed range of
the model is now found_by subtracting the frictional and the form resistance
coefficients from the measUred total resistance coefficient.

7.5 Extrapolation of Resistance Tests.


Having determined the components of the total resistance of the model, the
measured data of a resistance test at model scale can be extrapolated to
full scale. The resistance of the model is generally measured from a very
low speed up to the design speed. The design speed at model scale is found
from the full scale speed using the full scale Froude number and maintaining
it at model scale (eq. 7.1). The total resistance is made non-dimensional in
the usual way as
R,
Ctm =
1/2pVm 2 S,

The wetted surface is taken as the frame length from keel to waterline,
integrated over the length (and multiplied by two, to account for both sides
of the ship). Note that the length is taken along the centerline and not
along the waterlines.

During the model test the Froude number was maintained. This means
that the wave resistance coefficient at model and full scale are the same.
The total resistance coefficient of the ship can therefore be found from

Ct, = (1 + k)Cfs Cw + (7.7)


January 4, 1994, Resistance Tests 113

The form factor k and the wave resistance coefficient are directly found
from the model test. The frictional resistance coefficient at full scale can
be read from the plate line using the full scale Reynolds number.

The additional resistance coefficient ca is a new element. This coefficient


is a correlation coefficient based on experiences at full scale. It accounts
both for extrapolation errors (due to the various assumptions made) and
for effects at full scale like the occurrence of a relatively rougher surface at
full scale than at model scale. Holtrop [14] gives a simple relation for ca as

Ca = 0.006(L1 + 100)-0.16 0.00205 (7.8)

Although only the length is used as a parameter, similarly as in the case


of Froudes plate lines, the additional resistance coefficient now decreases
with increasing length and this can be attributed at least partly to roughness
effects. The effect of surface roughness in general requires special attention.

7.5.1 Froude's Extrapolation Method.


The method in which a form factor is used is also called a three-dimensional
extrapolation method. As mentioned before William Froude distinguished
only two resistance components, the frictional and the residual resistance.
Froude also used different plate lines as an extrapolator. Further the ex-
trapolation method is the same. Eq. 7.7 becomes:

Cts = C fs Cres Ca

The correlation factor ca is of course different from the correlation factor


used for the three-dimensional extrapolation method. The Froude method
is still used incidentally, so it is only mentioned here.

7.6 Effects of Surface Roughness.


The effect of surface roughness on the plate line depends on the roughness
height k. Assume that the roughness height is the distance between the
lowest and the highest points on a very regular rough surface. Physically the
effect of roughness on the local friction coefficient depends on the thickness
of the boundary layer. When the roughness height is small relative to the
boundary layer thickness the surface is hydrodynamically srnooth and the
roughness has no influence. When the roughness height is of the order of
the boundary layer thickness the local friction coefficient depends on the
roughness height only and becomes independent of the boundary layer and
thus of the Reynolds number. The effect on the plate friction line is shown
114 G.Kuiper, Resistance and Propulsion, January 4, 1994

in Fig. 7.6, which diagram has been derived from tests with roughened
pipes. The smooth line is the regular plate line, the other lines depart from
that line and become gradually horizontal with increasing Reynolds number.
When the local friction line is horizontal the magnitude of the resistance
coefficient depends on the ratio 11k,1 being the length of the plate.
The ratio between the rou- ghness height and the boundary layer thickness

0,014
NIIMI NII .1111ft. .
----- 1 T

0.010

t
NMAIMM
'M
,, Niuwibs.-..
' 1fV111&Min.h
... _2.'-.................-......,..
11111 .*

1.103

-wiLWW111111\\IMIIMMIBMWMFM11.1
--'Wl-Mmii~bnikWiMg=AMEWIMA
-

-- .0. ill`-',MI I CM \lima VIII I1Z : ..b-Al


- . 1-%
0.005

0.004 . W -
' -''''111111111IMNIftralki
--Th....
\IA I lag 6. -
"%%-i,MIIMM\I
. - - ,m., . _, 1 1 1 mil =I 1 m i lommimm ....

.. 111111WIMITEN
04
2.10.
..I.
II. _-
5.10

Sal M
1.105
0.003
-. \2AO
- .."1---*-1L_. 1M110-
4111-tp,PA --..'14
mi..=: -.01111MI 1.106
0,002

---4744
tasiklft

0.001
55 75 80
Ig Rot --5. 8.5 9,0 95

Figure 7.6: Resistance coefficients for flat plates with sand roughness ac-
cording to Prandtl-Schlichting

can be expressed as a roughness Reynolds number R k :


Vxk
=
Lines of equal roughness Reynolds number are plotted also in Fig. 7.6.

At full scale Reynolds numbers the hull of a ship is nearly always fully
rough and the resistance coefficient is independent of the Reynolds number.
In terms of an additional resistance coefficient this means an increasing
value of Ca with increasing Reynolds number, or with increasing length.
This is reflected in Holtrops equation 7.8. Bowden et al. [3] relate the Ca
to the roughness of the full scale hull by

103ca = 105(lc3/L)333 0.64 (7.9)


for L > 400m.
January 4, 1994, Resistance Tests 115

When such a formula is used the roughness height has to be known. For
the extrapolation of model tests a value k, = 150p is often used.

7.6.1 Equivalent Sand Roughness.


The roughness height is often determined indirectly by its effect on the
resistance instead of by its geometrical properties. This is done by measur-
ing the local friction coefficient of a surface with unknown roughness. The
measured friction coefficient is then compared to the friction coefficients
of plates with carefully distributed sandroughness of a uniform grain size.
Many of such tests with sandroughness were carried out by Nikuradse (1932
). The roughness size belonging to the measured friction coefficient then
gives the equivalent sand roughness height.
It should be recognized that the empirical values of ca are a correction
on the extrapolated values, without identifying the physical cause. This
correction is obtained from correlations between predictions from model
tests and full scale experiences. 3 In predictions from model tests the ex-
trapolator line (plate line) , the method to calculate form resistance and
the additional resistance coefficient ca are always combined. The value of
Ca depends on the choice of the extrapolator and of the way of determin-
ing the form resistance and on the facility in which the tests are carried
out. Although roughness effects are part of the value of ca, it is in fact
a correlation coefficient, containing all the errors made in the process of
extrapolation from model to full scale.

7.7 Appendage Drag.


A separate resistance component is the appendage drag. Model resistance
tests are generally carried out with the appendages, including the rudder.
This has been a matter of debate, because the rudder is sometimes also con-
sidered as part of the propulsor. There is ample reason for that, because the
effect of the rudder on the propeller is much larger than on the hull. His-
torically, however, the rudder is taken as an appendage of the hull, however.

Appendices with a considerable wetted surface, such as a rudder or a


bilge keel, contribute significantly to the frictional resistance. They in-
crease the wetted surface. Appendices which cause considerable vorticity,
such as open shafts, contribute mainly to the form resistance. Since both
3Note that the resistance at full scale is very difficult to measure. The ship resistance
is always found by inference from propulsion data.
116 G.Kuiper, Resistance and Propulsion, January 4, 1994

components of the resistance are extrapolated as functions of Reynolds it


is not necessary to separate the appendices from the hull. Resistance tests
are therefore carried out with appendices. If only a bare hull resistance is
available the appendage drag has to be estimated from drag coefficients of
simple bodies.

7.8 Effective Power.


The power required to tow a ship is found from

PE = VsRT (7.10)
Here PE is the Effective Power in kW and RT is the total resistance
without propeller.

7.9 Effects of Laminar Flow.


The plate line or extrapolator used to calculate the frictional resistance is
related with the turbulent line in Fig 3.1 on page 52. It is important to have
a sufficiently high Reynolds number at the model scale to ensure turbulent
boundary layer flow all over the model. Laminar flow regions will lead to
much lower frictional resistance or/and to earlier flow separation, causing an
unrealistic form resistance. Since the Froude number has to be maintained
at model scale the only parameter to increase the Reynolds number is a
large model size. This is limited by the proportions of the towing tank,
because the cross section of the model has to remain below about 1/50th of
the cross section of the towing tank to avoid significant wall effects on model
resistance. As a result the model Reynolds number is often in the range
where both laminar and turbulent flow may occur. Turbulent flow over the
bow is ensured at model scale by using turbulence stimulators This is a
.

strip over one frame near the bow. The strip consists of sandgrains glued
to the model (Fig. 7.7, or of studs: a row of cylinders of approx. 2mm high
and at a distance of approx. 1 cm (Fig. 7.8. Apart from the stimulation of
the boundary layer into turbulence such a strip also has its own resistance
[30], which is generally not accounted for.
On smaller models at low speeds more turbulence strips may be necessary
on the afterbody to avoid "relaminarization" of the boundary layer.

7.10 Wake Scale Effects.


In particular when the boundary layer at the model is fully turbulent, the
frictional resistance coefficient of the model will be much higher than that
January 4, 1994, Resistance Tests 117

Figure 7.7: Boundary layer tripping by sandstrips

Figure 7.8: Boundary layer tripping by studs

of the ship. This means that the model has a relatively high resistance
compared to full scale. As was shown in chapter 3 the wake is related to
the resistance. So the wake fraction of the model will be higher than that
of the ship. This is important for the propeller inflow and will be discussed
in chapter 20.
118 G.Kuiper, Resistance and Propulsion, January 4, 1994

Length between perpendiculars Lpp 180


Length on waterline 185.06
Breadth moulded 32.24
Draft moulded on F.P. TF 9.05
Draft moulded on A.P. TA 9.63
Displacement v-olume moulded DDD 32466
Displacement-weight in salt water 33278
Wetted surface without appendages si 6762
Wetted surface with appendages S2 6828

Table 7.2: Data of container ship.

V, V, R,
m/sec knots N
0.989 9.01 26.32
1.207 11.00 39.74
1.427 13.01 55.39
1.591 14.50 69.09
1.756 16.01 81.85
1.920 17.51 95.23
2.084 19.00 111.14
2.249 20.50 129.74
2.414 22.01 155.60
2.523 23.00 179.02

Table 7.3: Results of resistance test

7.11 Exam.ple of Resistance Extrapolation.


An example of the extrapolation of the resistance from model tests is given below. The
example is for a containership with the characteristics as given in Table 7.2.

The turbulence stimulators were studs. The scale ratio was 22, so the model length
was 8.18 m. The design speed of the ship was 20 knots. The corresponding model speed
can be found from the Froude number equivalence FT, = Fris, so that V = V3 //X.
For a design speed of 20 knots (1 knot = 1 mile/hour = 1854 m/hr):
,)(1\ 1854 1
= 2.196m/sec.
Vin = (``" 3600 --\M
The temperature of the tank water was 12.9 degrees Celcius.

A resistance test has been carried out for a speed range of the ship of 9 to 23 knots.
The results of the resistance test are given in Table 7.3

Since the average model speed over the run of the carriage may differ somewhat
January 6, 1994, Resistance Tests 119

Vs Ctm Rram Cfm Csm/Cfm -b-rs,m Fnm4/Cfm


knots x107
9.01 .00382 .689 .00320 1.192 .109 .044
11.00 :00387 .841 .00309 1.250 .133 .101
13.01 .0Q386 .994 .00300 1.284 .157 .203
14.50 .00387 1.11 .00295 1.313 .175 .319
16.01 .00376 1.22 .00290 1.299 .193 .482
17.51 .00366 1.34 .00285 1.283 .211 .700
19.00 .00363 1.45 .00281 1.289 .229 .985
20.50 .00364 1.57 .00278 1.309 .248 1.352
22.01 .00379 1.68 .00275 1.379 .266 1.816
23.00 .00399 1.76 .00273 1.463 .278 2.183

Table 7.4: Calculation of residual resistance coefficients

slight deviations from the chosen speed values are possible.

The test results_are now plotted in non-dimensional form as resistance coefficients


versus Reynolds number. The resistance coefficient is calculated as Ct = 112/1:v7s.. The
tank water is fresh water and its density is 1000 kg/m3. The wetted area of the model
with appendages is 6868/222 = 14.1074m2.For 20.5 knots the result is:
129.74
(20.5knots) = = 0.00364
1/2 x 1000 x 2.2492 x 14.1074
The kinematic viscosity of water at 12.9 degrees is read from Table 22.1 to be
1.20493 x 10-6. (Linear interpolation in the table is suficiently accurate). As refer-
ence length the waterline length of the model (8.412m) is used. For 20.5 knots the
Reynolds number is:
2.249 x
Rn(20.5knots) = 8.412= 0.157 x 108
1.20493 x 10-8
To determine the wave resistance the Froude number is calculated. With a gravita-
tional acceleration g = 9.81m/sec the Froude number at 20.5 knots is:
2.249
Fn(20.5knots) = = 0.248
V9.81 x 8.412
The results of the extrapolation are worked out in Table 7.4 from the data in Ta-
ble 7.3.

The form factor can now be calculated from an extrapolation of Ctrn/ C in, towards Fn =
or from an extrapolation of Fn4/Cfm towards Fnm = 0. Both graphs are given in Figs. 7.9
and 7.10. It can be seen from Fig. 7.9 that it is very difficult to extrapolatetowards
Fr, = 0 because the wave resistance approaches the friction line asymptotically. This is
easier in Fig. 7.10. The result of the graphical extrapolation toward zero Froude number
is that (1 k) = 1.14. Using this form factor k the wave resistance coefficient can be
calculated:
Cw, = Ctm - (1+ k)Cfm
The result is shown in graphical form in Fig. 7.11.
120 G.Kuiper, Resistance and Propulsion, January 6, 1994

1.4

1.3

Ct /Cf

1.2

1.1

10
o 0.1 0.2
Fn

Figure 7.9: Determination of the form resistance on the basis of F.

1.4

1.3

Ct /Cf

1.2

1.1

lo I I 1 I 1

o 0.1 0.2 0.3 0.4 0.5


Fn4 /Cfm

Figure 7.10: Determination of the form factor with Prohaska's method

From this Figure it can be seen that there is a hump in the wave resistance at 14
knots. The design speed is correctly in a hollow. The wave resistance at design speed
is only about 10 percent of the total resistance at a Froude number of 0.25. At higher
speeds the wave resistance increases sharply.

The Froude number at model and full scale are the same, so the wave resistance
coefficient of the model is also that of the ship. The frictional resistance coefficient of
the ship can be found from the ship data. The ship speed is 20.5 knots or 10.558 m/sec.
The standard temperature at which full scale data are calculated is 15 degrees Celcius,
at which the kinematic viscosity u = 1.1883 x 10-6. This value is read from Table 22.1.
The Reynolds number at 20.5 knots is now
10.558 x 185.06
Rn3 = 1.038 x 109
1.883 x 10-'
January 4, 1994, Resistance Tests 121

0.0070

0.0060

0.0050

CTM
0.0040

0.0030
CFM

0.0020

0.0010

0.0000
6.0 10.0 14.0 18.0 22.0 26.0

SHIP SPEED Vs IN KNOTS

Figure 7.11: Resistance Test Results.

The resistance coefficient at that Reynolds number is 0.00144. (Table 19.2. The wave
resistance coefficient is found from Table 7.4 to be 0.00364 (0.00278 x 1.14) = 0.00047.
The total resistance coefficient at that speed is than found from Ct, = (1 + k)Cf, +
Cwm + Ca. With a correlation allowance Ca = .00038 the total resistance coefficient of
the ship is found to be

Cts = 1.14 x 0.00144 + 0.00047+ 0.00038 = 0.00249

The results of the resistance extrapolation for the speed range of 17 to 22 knots is
now given in Table 7.5.

The resistance of the ship can be calculated from the total resistance coefficient Ct,.
The standard condition for full scale is salt water with density p = 1025kg/m3. At 20.5
knots this gives:

R, = 0.00249 x 0.5 x 1025 x 10.5582 x 6828 = 971kN


122 G.Kuiper, Resistance and Propulsion, January 4, 1994

V, Cu, C1, , C1, R3 PE


knots x105 x105 x105 kN kW
17.0 42 147 248 665 5814
17.5 . 41 147 247 700 6300
18.0 41 146 246 737 6824
18.5 41 146 245 777 7391
19.0 41 145 245 819 8002
19.5 42 145 245 865 8673
20.0 44 144 247 915 9419
20.5 47 144 249 971 10239
21.0 51 144 253 1034 11171
21.5 58 143 259 1109 12266
22.0 66 143 267 1196 13537

Table 7.5: Extrapolation of resistance test results.

The effective power PE = Vs X R, has also been calculated. At 20.5 knots this is

PE = 10.558 x 971000 = 10239kW


Chapter 8
Resistance Prediction using
Statistical or Systematic Data
Objective: Making an estimate of the resistance in the preliminary de-
sign stage.

Design guidance can be obtained from the results of previous model


tests. Some general guidelines can also be deducted from hydrodynamic
considerations.

8.1 General Considerations for Hull Design.


The first choice to be made is the blockcoefficient. This is strongly related
with the Froude number because at higher foude numbers the wave resis-
tance becomes dominant and a low blockcoefficient is required. A statistical
review of the relation between the blockcoefficient Cb and the Froude num-
ber of some 200 ships is given in Fig. 8.1 from Townsin. In this figure also a
regression formula is given. Due to the small scale of Cb the regression curve
seems reasonably accurate, but at e.g. a Froude number of 0.25, which is
typical for containerships, the blockcoefficient varies between 0.55 and 0.7,
so a considerable variation is possible.

Typical Froude numbers and the position of humps and hollows are
shown in Fig. 8.2.

The next general consideration is about the longitudinal distribution of


the displacement, expressed by the curve of sectional areas. A parameter

123
124 G.Kuiper, Resistance and Propulsion, January 4, 1994

0.4._ 100F;, e23 -4-Tan(8C5-5.6)

02_

Cs-0-7 44tTanl 25(-23-F,


0.1 1
I I I I I 1

.5 .6 -7 43 -9
C3
Figure 8.1: Relation between blockcoefficient and Froude number

of this curve is the position of the center of buoyancy relative to the aft
perpendicular A-13 . As mentioned in chapter 1 the shape of the forebody
January 4, 1994, Statistical Resistance Predictions 125

WAVE

RIGATES

FISHING BOATS
TUGS
-

WAVE PL ANING
FE RRIES
CONTAINERS

TANI

0.1 0.2 0.3 0.4 0.5 0.6 0.7 01.6 0.9 1.0 1.1
Fa

Figure $.2:_ Typical Foude numbers for various types of ships

is important for the wave resistance. In the afterbody separation of the


flow is the greatest risk. At increasing speed (and thus Froude number)
the risk of separation does not increase significantly. As a result A-13 will
shift from greater than L/2 (forward of midships) to less than 1/2 (aft of
midships) with increasing Froude number. Slender ships will therefore have
their longitudinal center of buoyancy a few percent of the length aft, full
ships a few percent forward of midships.

The third general consideration is on the vertical distribution of dis-


placement, which is the choice between U and V-frames. When U-frames
are applied the displacement is further from the surface and the wave resis-
tance will be less. On the other hand application of U-frames easily leads
to three dimensional separation in the afterbody, which increases the form
resistance. In general V-shaped ships will have less resistance at moderate
Froude numbers. The wake of these ships, however, is not favourable for
propulsion, as will be discussed later.

More quantitative estimates of the resistance of a given ship can be


derived from the results of previous model tests. There are two methods to
represent experience

Results of systematically varied hull forms.

Analysis of existing data by regression analysis.


126 G.Kuiper, Resistance and Propulsion, January 4, 1994

These two approaches will be treated shortly.

8.2 Systematic Series.


In order to obtain a systematic series of hull forms a parent model has to
be chosen. This hull form is systematically altered so that only one or
a restricted number of parameters are varied. Examples are the "Taylor-
Gertler" series [9] (1907-1954) or the "Series 60" [44] (1963), . 1
In case of systematically varied ship forms one is in fact interpolating
between available ships from the series. Note that extrapolation outside the
range in which the parameters are varied is very dangerous! Also the series
has a distinct shape of the parent form and strictly speaking the resistance
data are valid only for the forms in the series. Before such methods are used
it is recommended to read the publication for a proper judgement of the
possibilities and the restrictions of the series. Since in most cases the hull
forms are rather out of date, the hull form is generally modified in present
designs. This makes that the prediction is only an approximation for use
in the preliminary design stage.

An example of the results is Taylor's series. From 1907 until 1914 ad-
miral Taylor measured the resistance of models which were systematic vari-
ations of a parent form, the cruiser "Leviathan". Taylor removed the ram
bow, moved the center of buoyancy to midships and used a 3% bulbous
bow. The lines of the parent form of the Taylor series are given in Fig. 8.3.
Experiments were carried out for five values of L 1 A113,2 two values of B/T

`,....
07
06
04
02
_

Figure 8.3: Lines of the parent form of Taylor's series

'For a review and appraisal of formdata and resistance see Gallin,College MT8 On-
twerpen van schepen I, par.1.2.2.
2where A is the displacement weight
January 4, 1994, Statistical Resistance Predictions 127

and eight values of the prismatic coefficient G. The B/T values used were
actually 2.25, 2.92 and 3.75. Initially the values of B/T=2.92 were not
published.In total 80 models were towed!
Gertler reanalysed his data in 1954 [9]. He corrected the datof Taylor for
tankwater temperature, laminar boundary layer flow and tank blockage.
The tests of B/T=2.92 were recalculated to B/T=3.0. Gertler estimated
the frictional resistance from the Schoenherr mean line. The remaining
resistance , which contains both the form drag and the wave drag, was
combined as residuary resistance, just as William Froude had done (see
chapter 7). Gertler gave the residuary resistance in 117 diagrams as:

CT= f(BIT,VIIL1,Cp,AILL)

The midship coefficient of the series was always 0.925 and the position of
the center of buoyancy was always at midships.

The diagrams of Taylor are still used. An example of Taylor's original


diagrams is given in Fig 8.4. In these diagrams the residual resistance
is found as RID in which D is the displacement in tons. The residuary
resistance is based on the original Froude extrapolator lines. 3
An example of Gertlers representation is shown in Fig. 8.5. The men-
tioned series is very old and the hull shapes are therefore quite different
from those used today.

A well known series is the "Todd-60" series. The parent lines were de-
veloped in 1948 for a single screw merchant ship. Separate parent forms
were developed for separate block coefficients (0.60,0.65,0.70,0.75 and 0.8).
For each block coefficient an optimum location of the center of buoyancy
was determined and the total resistance was given for that condition. The
results have been published by Todd in 1963 [44].

More recently other series have been published, such as the Guldham-
mer series [10], published in 1965 and 1969. Most of those series are based
on a restricted number of variations.

The measurement of a series with sufficient variation of parameters is


very costly, however. The application of regression analysis on existing
databases has therefore been developed more recently.

3Take care of the units in these diagrams.


128 G.Kuiper, Resistance and Propulsion, January 4, 1994

2.25 = 0,80 Vln knots L. in It , Am lbs


I ( 11
210

farialpis
lirdiallia
200

tI
Fr r 11117113111.E.M1 la NW= 2,1 2 26 2.1 A2 6.0 7 e4 o 10

r, 11111111.1111 MINIIIMINIIM3111M631111111111MLIMIll i
So
11111111101111111111111111LIIMINIMIIIILIIIMM111111111 HJ
lill1111 NIIIMIISIIIIIIIIIIRRINICIPLIMIR111111 1011
NI
0.10 0,59 0,60 0,621 070 075 0.60 0.

Figure 8.4: Example of Taylors diagrams for residuary resistance

8.3 Regression of Available Data.


When a large number of models have been tested interpolations can be
made to estimate the resistance of an arbitrary hullform. To make inter-
polation possible specific parameters have to be defined to characterize the
hull form. The choice of the proper parameters determines the accuracy of
the interpolation, and thus of the prediction for an arbitrary hullform. Also
the range of the parameters of the hullforms in the database restricts the
possibilities of interpolation, because again extrapolation is generally very
dangerous. The database for regression analysis is generally restricted to a
certain class of ships, which restriction is comparable to the choice of the
parent hull form in systematic variations. However, in case of regression no
single typical hullform can be identified.

An example of an early analysis of a large number of data is the analysis


of Lap [23] for single screw ships. Lap defined the viscous resistance coef-
ficient as the "Schoenherr"-line. For the residuary resistance he used the
following parameters:

prismatic coefficient Cp or 0
January 4, 1994, Statistical Resistance Predictions 129

/
V
FROUDE NUMBER r---
VgL
016 018 020 022 024

3 0 a 10-3 A

! c =0.70 I

ll
1 P
A
1

A
VALUES OF ViL3
701110-3
6.0 a 10-3

l
50110-3
40 a10
41411
3.0 :10-3
2.0 a10-3 --5

..-1 41 ''\
rf- -09-5x701
1
I

o
1 "G.
Ut08
SPEED-LENGTH RATIO
0 755

v/-
09 I0

Figure 8.5: Example of Gertler's diagrams for residuary resistance

Froude number, expressed as VsINICpLwi

position of centre of buoyancy LC B in % of Livi

He found that the value of B/T could be varied over quite a range without
affecting the resistance. The average BIT value in Laps database was 2.4.

The residuary resistance coefficient was defined as

Cr = -
1
p1/.,2 Am
Here Am is the midship area coefficient. Note that the "Froude number"
as defined above contains both the regular Froude number Vs/Vir and the
prismatic coefficient q5 and is not non-dimensional. It can also be written
as J1/410)

Based on the position of the centre of buoyancy a and the prismatic


coefficient 0 the hullforms were divided into five groups, as given in Fig. 8.6.
130 G.Kuiper, Resistance and Propulsion, January 4, 1994

42 7

.1

o
drea
te

o
a WrIP?OlV
AV Ilir
PIPV
-2
1

0,70 0.75 0.80

Figure 8.6: Relations between position of LCB and prismatic coefficient by


Lap

For each group the residuary resistance coefficient Cr is given in graph-


ical form. One of these diagrams is given in Fig. 8.7.
The resistance in the tank condition of the ship becomes

1
R = (C1
Crs Ca)-2PVs2S

For the calculation of the total resistance the wetted area S should be
known. Lap gives the approximation

S = (3.4V'/3 + )V113

where V is the volume displacement in m3


January 4, 1994, Statistical Resistance Predictions 131

60

0
"%PAC 2 4 )
Al 111 1 I/
io WM 17 A V IfAr"
o ../MAPIFA
r '
F FP- Apr
_./). _ __.!......---.--"_ _______ *

___-----
0

o
1,1 1.2 1.3
(Vrr-

Figure 8.7: Example of a diagram of residuary resistance from Lap

8.4 Design of Curve of Sectional Areas by


Lap.
Based on the same database Lap determined the curve of sectional areas
based on the choice of the LCB. He relates the total prismatic coefficient
and the LCB with the prismatic coefficients in fore- and afterbody, as shown
in Fig. 8.8.
Based on the prismatic coefficient of the fore- or afterbody Lap also
gives the curve of sectional areas over the length of the ship, as shown
e.g.in Fig. 8.9
Lap also gives a graphical relation between the prismatic coefficient and
the angle of entrance of the waterline. These figures can be used for pre-
liminary hull design.
As mentioned before this regression is based on a database as available
in 1954. These diagrams should therefore be used as a rough indication for
preliminary design and for resistance prediction.
132 G.Kuiper, Resistance and Propulsion, January 4, 1994

, s

EWA
vs

0.70
.6"UP "C"\ AI
I AllAY
R
6O

CAS
:M
iatirl7A A.
.../111;
CAO

070

CAO OAS 0,7S 010


0.

Figure 8.8: Relation between prismatic coefficient of the whole ship and of
fore-and afterbody

8.5 The method of Holtrop and Mennen.


Based on a more recent database of Marin Holtrop and Mennen used re-
gression analysis in 1978 and later. [14], [15]. Since they had computer
power ayailable they were no longer restricted to a very simple formula-
tion. Several powers of parameters and cross-products could also be used.
In case of Ref. [15] the database consisted of 334 models. The database is
continuously updated and recent programs are based on a database of 1988.

A short review of the parameters used for the resistance prediction will
be given below.

The total resistance coefficient is divided into components:

CT = (1 + k)Cf + Cu, + Ct, + Cb + Ca


Here the coefficients Cf,Cu, and Ca are the resistance coefficients due to
January 4, 1994, Statistical Resistance Predictions 133

0.75

0.70

s.

065

060

10 20 30 40 50
So ca mast el per cona 61

Figure 8.9: sectional area of forebody according to Lap

friction, waves and the additional coefficient as discussed in section 7. Cb is


the additional pressure resistance coefficient due to a bulbous bow near the
water surface and Ce,. is the additional resistance due to transom immersion.

The form factor (1 + k) is given as a function of the following non-


dimensional parameters:

1 + k = f(LIB,LIT,LCB,VIL2,C)
The wave resistance is given as

= f(F,CM,VIL3,LIB,BIT,ABTIBT,ATIBT,T1,14,Cp)
where:
V = Displacement in m3,
Cm = midshipcoefficient,
L. waterline length,
B. moulded breadth,
AT =transom area in m2,
hb= vertical position of centre of transverse bulb area above keel in 7n,
ABT= transverse area of bulbous bow in m2,
117, = Froude number,
134 G.Kuiper, Resistance and Propulsion, January 4, 1994

Tf= forward draft in in,


Cp= prismatic coefficient.

For the wave resistance- coefficient in [15] three formula for three Froude
number ranges are given. -

The wetted area is also given in a regression formula as

S = f(L,B,T,Cm,Cb7Ctupl ABT)

in which Cb is the block coefficient and Cp is the waterline coefficient.

Holtrop and Mennen also give estimates of the appendage resistance


as a form factor k2 similar as the form factor for the hull. The appendix
resistance is defined as -.

= 0.5PV2SCIPP(1 k2)C f

where Cf is read from the extrapolator line at the proper Reynolds num-
ber. k2-values for a range of appendages are given. The k2 factors allow for
a certain form resistance of the appendices, while most methods take only
frictional resistance into account.

Additionally Holtrop and Mennen also give an estimate for the resis-
tance of a bow thruster tunnel and for the correlation coefficient Ca. For
design purposes a regression formula for the angle of entrance of the water-
line is given.

In a more recent adaptation of the method in 1988 some influence of


the Froude number on the form factor (1 + k) was tentatively established.
Moreover, it appeared more accurate to express the wave resistance not by
two or three formula for certain speed ranges, but to give an expression for
the wave resistance coefficient for discrete values of the Froude number.

The formula's are complex and suitable for a computer program. Since
more parameters are used for the calculation of the resistance, more has
to be known in the preliminary design stage. When a certain parameter
is not yet known, it is best to analyse the influence of variations of that
parameter.
January 4, 1994, Statistical Resistance Predictions 135

8.6 Example of Resistance Prediction.


An example of a resistance calculation is given below. The ship is the same containership
as used in chapter 7. The input of the program (called DESP) is given in Fig. 8.10.

RESULTS RESISTANCE CALCULATIONS

BLOCK COEFF (ON LWL) (-) 0.5826


PRISM COEFF (ON LWL) (-) 0.5994
LCB -(PER CENT LPP FWD OF 1/2 LWL) -1.59
WETTED SURFACE AREA HULL (STAT) (M2) 6559.46
WETTED SURFACE AREA TOTAL(STAT) (112) 6724.46
HALF ANGLE OF ENTRANCE (STAT) (DEGR) 14.58
CORRELATION ALLOWANCE (-) 0.000260
1+K0-1 HULL (-) 1.172
1+K0 HULL AND APPENDAGES (-) 1.180

Y FN FORM RFR1C RAU RAPP RNAYE RBUL8 RTRANS RAER R PE YM RM Y/ CE


KT FACT EN KN KN KM KN EN KN KN KW M/S N D1/6 FROUDE
1 0.012 1.167 1.9 0.2 0.1 0.0 0.0 0.0 0.1 2.7 1 0.110 0.48 0.18 391
2 0.024 1.167 6.9 0.9 0.3 0.0 0.2 0.0 0.5 10.0 10 0.219 1.67 0.35 455
3 0.036 1.167 14.8 2.1 0.6 0.1 0.6 0.0 1.2 21.7 33 0.329 3.49 0.53 491
4 0.048 1.167 25.2 3.8 1.0 0.3 1.3 0.0 2.1 37.7 77 0.439 5.91 0.71 515
50.060 1.168 38.3 5.9 1.4 0.7 2.3 0.0 3.3 58.0 149 0.548 8.90 0.88 533
6 0.072 1.168 53.8 8.5 2.0 1.3 3.6 0.0 4.7 82.5 254 0.658 12.45 1.06 549
7 0.085 1.168 71.8 11.6 2.7 2.0 5.0 0.0 6.4 111.0 399 0.768 16.53 1.23 562
8 0.097 1.168 92.2 15.2 3.5 2.7 6.5 0.0 8.3 143.2 589 0.877 21.10 1.41 575
9 0.109 1.169 114.9 19.2 4.3 3.1 7.9 0.0 10.5 178.5 826 0.987 26.12 1.59 588
10 0.121 1.170 140.0 23.7 5.3 3.1 9.2 0.0 13.0 216.9 1115 1.097 31.57 1.76 602
11 0.133 1.171 167.4 28.7 6.3 4.1 10.1 0.0 15.8 259.6 1469 1.206 37.56 1.94 612
12 0.145 1.172 197.1 34.2 7.4 6.8 10.8 0.0 18.8 307.3 1897 1.316 44.15 2.12 619
13 0.157 1.173 229.0 40.1 8.6 10.3 11.2 0.0 22.0 359.0 2400 1.426 51.25 2.29 625
14 0.169 1.174 263.2 46.5 9.9 14.7 11.3 0.0 25.5 414.6 2986 1.536 58.83 2.47 631
15 0.181 1.174 299.6 53.4 11.3 20.7 11.1 0.0 29.3 474.9 3664 1.645 66.96 2.64 635
16 0.193 1.173 338.1 60.7 12.8 29.9 10.7 0.0 33.3 541.2 4454 1.755 75.74 2.82 637
17 0.205 1.171 378.9 68.5 14.3 44.3 10.1 0.0 37.6 615.5 5382 1.865 85.34 3.00 636
18 0.217 1.169 421.9 76.8 15.9 63.9 9.5 0.0 42.2 697.8 6462 1.974 95.76 3.17 633
19 0.229 1.166 467.0 85.6 17.6 91.6 8.8 0.0 47.0 790.9 7731 2.084 107.25 3.35 627
20 0.242 1.162 514.2 94.9 19.4 132.0 8.1 0.0 52.1 899.6 9256 2.194 120.23 3.53 615
21 0.254 1.158 563.6 104.6 21.3 192.5 7.4 0.0 57.4 1030.9 11137 2.303 135.34 3.70 596
220.266 1.153 615.1 114.8 23.2 282.7 6.9 0.0 63.0 1194.7 13520 2.413 153.48 3.88 568
23 0.278 1.148 668.7 125.5 25.2 404.5 6.4 0.0 68.9 1392.6 16477 2.523 174.80 4.06 537
240.290 1.143 724.5 136.6 27.3 554.8 6.1 0.0 75.0 1621.4 20019 2.632 148.99 4.23 506

Figure 8.10: Input for a resistance calculation of a containership

A number of statistically based choices have been made in the input. The appendix
resistance factor k2 is taken as 0.5. The statistical resistance allowance Ca has been taken
as 0.00005. The aperture coefficient has been taken -5, which is the average between nor-
mal and V-shaped frames.
136 G.Kuiper, Resistance and Propulsion, January 4, 1994

The result of the statisical calculation is given in Fig. 8.11

DATE:I991- 1- 7 TIME: 10H 46MIN CALC NO= 58

INPUT DATA DESPM (RELEASE 5 FEBRUARY 1990) BY: J.HOLTROP


COMMENT: -
Example Ro-Ro shlp
CODE NUMBER (8. 10 OR 11) 10
WHEN LWL IS 0: RESISTANCE DATA IN INPUT TABLE!
LENGTH ON WL (M) 185.060
LENGTH PP (M) 180.000
BREADTH (M) 72.240
DRAUGHT FWD (M) 9.050
DRAUGHT AFT (M) 9.630
VOLUME OF DISPLACEMENT (M3) 32466.0
CM (-) 0.9720
CWP (0 IS ALLOWED) (-) 0.0000
LCB (XLPP FWRD OF 1/2 LPP) -3.00
HALF ENTR ANGLE (0 IS ALLOWED) (DEGR) 0.00
APPENDAGE AREA (M2) 165.00
APPENDAGE FACTOR 1+K2 (-) 1.50
BULB TRANSVERSE AREA (M2)
CENTROID OF BULB ABOVE KEEL (M) 4.600
TRANSOM AREA/MIDSHIP SECTION (-) 0.000
ALLOWANCE FOR HULL FORM ON CA (-) 0.00005
AIR EXPOSED TRANSV AREA (10 + 11)(M2) 1000.0
C-STERN (-20=PRAM.-10=V.0=NORMAL9+10=U) -5.
C-APERTURE (0=CONV.10=EXPOSED SHAFT(S)) 0.
SPEED (KNOTS) 22.00
SCALE RATIO (IF 0 THEN NO CALC) (-) 22.0000
MEASURED TOTAL WETTED SURF. AREA (M2) 0.00
WIDTH OF TOWING TANK (M) 0.00
DEPTH OF TOWING TANK (M) 0.00
TEMP.OF TOWING TANK WATER (DEGREES C) 15.00
PITCH-DIAMETER RATIO MODEL PROP (-) 0.000
CHORD LENGTH 0.75R/DIAMETER (-) 0.000
MEASURED MODEL WAKE FRACTION (-) 0.000
ADDITIONAL ALLOWANCE ON CA IN SCALING 0.00000
LOWEST PROP TIPS ABOVE KEEL PLANE(M) 0.000
NO. OF PROPELLERS (-)
OPEN(0) OR DUCTED PROP(19A OR 37)(-) 0
NO.OF BLADES (-) 4
TOTAL AVAILABLE SHAFT POWER (KW) 16755.
MIN PROP DIAM (IF 0: NO DESIGN) (M) 6.500
MAX PROP DIAMETER OR MIN RPM (M)/(RPM) 122.000
MAX RPM (RPM) 122.000
ADDITION BAR(OPEN) OR BAR(DUCTED)(-) 0.050
THE FOLLOWINS DATA ARE NOT USED IF ETAR=0
REL-ROTATIVE EFFICIENCY (-) 0.000
THRUST DEDUCTION FRACTION (-) 0.000
WAKE FRACT. AT MIN PROP DIAMETER (-) 0.000
WAKE FRACT. AT MAX PROP DIAMETER (-) 0.000
V(KNOTS) R(KN)

Figure 8.11: Results of resistance calculation

In the top of the figure some results are given, which were calculated directly from
the input (e.g. the blockcoefficient) or which were calculated based on statistical data in
the program (the wetted surface, the half angle of entrance, etc.)
January 4, 1994, Statistical Resistance Predictions 137

Vs PE statistical PE model test perc. error


17.0 5814 5382 +8
19.0 8002 7731 +3.5
21.0 11171 11137 +0.3
-
22.0 13537 13520 +0.01
Table 8.1: Comparison of statistical and model test prediction

The result of the statistical calculation is compared with the extrapolated result from
the model test in Table 8.1.
The prediction is accurate for the design speed, but at other speeds considerable
differences occur. Deviations of some 5 percent may be considered as common.

8.7 Resistance of Small Vessels.


For small vessels the wave resistance is generally dominant. In 1971 van
Oortmerssen [34] used regression analysis for the resistance and propul-
sion data of small conventional ships like tugs and trawlers. Based on the
Havelock equation for a pressure disturbance travelling below a surface, he
used terms of the form eFn-2 for the wave resistance. The coefficients were
polynomials depending on LIB,BIT,ie, Cm, Cp and LCB, derived from a
database of 93 ships. Here ie is the angle of entrance of the load waterline.

Since each ship had data at many speeds the total database comprised
970 points. The viscous resistance was derived from the ITTC57 extrapola-
tor. For the form factor an expression based on CB, LIB and BIT is given
with a statistical coefficient. The total resistance is given by v.Oortmerssen
as resistance per ton displacement R/A.

Many other series have been investigated. For a review of some series
of high speed displacement ships see van Oossanen(1980)[36].
138 G.Kuiper, Resistance and Propulsion, January 4, 1994
Chapter 9
Intermezzo:Equations of
Motion
Objective: To show the origin of the equations of motion. The derivation
of the formula's is not important. The purpose is to show that the equations
of motion are th equivalent of Newtons second law, to show the assump-
tions used to arrive at the equations (Hooke's law, Newtonian fluid). Also
the concept of rotation is introduced.

The equations of motion describe the relation between forces and mo-
tions in the fluid. The equations are rewritten forms of Newtons' second
law:

F(t) = mi (9.1)
where F=force in Newtons, m=mass in kg and x=position in in.
The equations of motion therefore describe a relation, not a situation!
The situation, local velocities and pressures (forces), can only be found
from a combination of the equations of motion and boundary conditions.
The boundary conditions are situations in pressure or velocity at a certain
position at a certain time. When the boundary conditions are sufficient, the
velocities and pressures at other times and positions can then be calculated
using the equations of motion.

Exarnple A simple and well known example is the application of Newtons'


law directly to a body with mass m. The equation of motion is eq. 9.1 and
the boundary conditions are e.g.

139
140 G.Kuiper, Resistance and Propulsion, Januar-y 4, 1994

x(0) =
F(t) = f(t)
This makes it possible to calculate the velocity of the body at an arbi-
trary time t from:
--It
1= f(e)dti C (9.2)
TTE

From the boundary conditions it is found that C = 0. In this case the


equation of motion is integrated.
When the velocity (t) is given as a boundary condition instead of f(t)
the forces can be derived directly from the equation of motion by using the
time-derivative of the given velocity.

This example is meant to illustrate the character of the equations of


motion in a fluid. In the following the equations of motion will be derived
without using vector notation, so that knowledge of it is not necessary. The
derivation will also be in one direction only and for incompressible flow. The
purpose of this derivation is to give basic insight in the equations of motion
and to show the concepts of stress-strain and of rotation-deformation. A
basic insight in the equations of motion is also necessary to understand
simplifications which will be made later on, in order to make calculations
feasible.
Before deriving these equations of motion another basic equation has to
be formulated.

9.1 The Continuity Equation.


The continuity equation expresses that no mass is lost when there is no flow
through a control surface. When a cube of fluid is observed at a certain time
its shape may change over time, its density may change, but the amount
of mass is unchanged regardless the deformations or compression. For an
incompressible fluid this means that not only the mass, but also the volume
of such a cube remains unchanged. We will assume incompressible fluid
here, because the line of reasoning remains the same as in compressible
flow. Water can nearly always be considered as incompressible.
The formulation of the continuity equation is an example of the use of
an arbitrary control volume which is fixed in time and position.
Consider a control volume as shown in Fig. 9.1 with sides dx,dy and dz.
The velocity of the fluid entering the control volume from the left side
ADEF is now called u. (Note that this is the average velocity over the
plane, not a local velocity. This is useful because afterwards the size of
January 4, 1994, Equations of Motion 141

Figure 9.1: Control volume fixed in time and position

the control volume will be reduced infinitely, so that this average velocity
approaches the local velocity in point A.) The velocity leaving the control
volume at the right side BCGH is then u + edx. Partial derivatives are
used here although only the derivative in x-direction is relevant.
The total volume change dV in a time-step dt is now:

dV = (audx)dydzdt (avdy)dxdzdt (awdz)dxdydt


ax ay az
In incompressible flow this volume change must be zero for any volume
and for any time-step, so that

au au aw
(9.3)
4T;+-f-Ta7z-=

This is the continuity equation for incompressible flow. This equation can
be used in the formulation of the equations of motion.

9.2 The Equations of Motion.


For the formulation of the equations of motion we consider the forces on
a cube of fluid. These forces are decomposed in three directions, so that
142 G.Kuiper, Resistance and Propulsion, January 4, 1994

at each side of the cube there is one pressure force (indicated a) and two
friction forces (indicated 7).

ciz
G+

Figure 9.2: Forces on a cross section of a cube of fluid

In Fig. 9.2 a cross section of a cube is shown with at the bottom the
pressure force a and the fiction force 7 in x- direction. The related forces
at the opposite side of the cube are a+ pidz and 7 + t dz. These forces are
always in opposite directions and the resulting forces on the fluid particle
can therefore be written as't dz and tz-dz only.
In Fig. 9.3 the resulting forces acting on a cube of fluid are indicated.
Note that in this case this is not a control volume, but a material cube of
fluid at a certain time. The first index indicates the plane at which the
force is acting, the second index is the direction of the force. So r a
friction force on a plane perpendicular to the x-axis in the direction of the
z-axis.
We can now formulate the resulting force in each direction on the fluid
particle. For sake of simplicity this will only be done in the x-direction.
The resulting force in x-direction is
ao-
= (--- dx)dydz (--
dy)dxdz -F dz)dxdy
ax ay az
This force can now simply be used in eq. 9.1, resulting in
January 4, 1994, Equations of Motion 143

Figure 9.3: Resulting forces in x direction

du
= pdxdydzdt (9.4)

where dxdydz is the volume of the element. The result of this substitu-
tion is

du ao-, Ty 1zx
(9.5)
dt ax ay az
In eq. 9.5 the relation between the fluid motion u and the stresses in the
fluid is given. The stresses in the fluid are in turn related to the deformations
and compressions in the fluid. For a further formulation of the equations of
motion the relation between the deformations (strain) in the fluid and the
stresses on a fluid element is required.

9.2.1 Rotation and Deformation.


The stresses in the fluid will now be related to the deformations of the fluid
particle. The cube, considered before, will be deformed by the forces on
its sides. For simplicity only two dimensions will be considered, so a cross
section of the fluid particle in the x-z plane as shown in Fig. 9.4.
144 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 9.4: Deformation of a fluid particle

The originally rectangular shape of the fluid particle under consideration


will be deformed after a time Lt to an arbitrary cross section as given in
Fig. 9.4. After the time-step:

B(1,0) is in B' (I, el)


D(0,1) is in D' (t-1 , 1)

C (1,1) is in C"(I , I+ l)

Two elements can now be distinguished in the deformation:

the rotation of the diagonal AC, which is g). This is called


the rotation of the element.
the deformation of the element, which is written as 1'9u+ g). This
is called the dilatation of the element.

The dilatation of the element causes a strain in the element.

9.2.2 Relation between Stresses and Strain.


The relation between the dilatation of the element and the frictional stresses
is written as

au
Tyx = ,Ty.) (9.6)
and
January 4, 1994, Equations of Motion 145

7-zz +)
/Su aw,
az ax
Note that the frictional stress in the x-direction contains also the velocity
gradients in the y-direction. The parameter z is the dynamic viscosity of
the fluid.
Also when no dilatation occurs a deformation is possible, as shown in
Fig. 9.5

Figure 9.5: Deformation due to pressure gradient

In this case the pressure gradient in z-direction causes a deformation.


In such a case the following relation yields:

aax
ax = ( + i2)
a
ax 2 Su
ax
(9.7)
Here p is the mean static pressure. This relation is not evident. It is valid
for an isotropic fluid and the stress-strain relation uses the same dynamic
viscosity as the relation between viscous stresses and deformations. The
fluids for which this relation is valid are called Newtonian fluids. Water is
such a fluid. An example of a non-newtonian fluid is blood.
Using the foregoing the equations of motion can be formulated.

9.2.3 Navier-Stokes Equations.


Eq. 9.5 can be rewritten using the foregoing as:

du Op 5 Su a au Su
= (9.8)
146 G.Kuiper, Resistance and Propulsion, January 4, 1994

a au aw

ap 82u a2v 02u 52u 02w


+
ax ax2 axay aZ2 aza)x
ap 52u a2u a2u, a (au ay aw,
a+likax2+ uy2 az2)+Paxkax +-Vaz'
x
The total derivative of the velocity u can be written in its partial deriv-
atives as(-1.
dt = at
-ku11.4+
ax v ay 8z so that the formulation of the equation
of motion in the x-direction yields:

auau
uT31--
au
= (9.9)
ap a2u a2u a2u
ax+11(ax2+ ay2 az2)
O au ay aw,
+ (
ax sax + ay+ az)
This equation is also called the Navier-Stokes equation, because the
French mathematician Navier and the English mathematician Stokes devel-
oped this equation independently in the middle of the nineteenth century-.
In incompressible flow this equation simplifies because eq. 9.3 can be
used, resulting in

, au au au au
Pt-57 + u-6T 2,-(93 tv-az-} = (9.10)
ap 52u 52u 02u
- -5; -EY( uz2)

The equations in the other two directions can simply be found by changing
the x,y and z and u,v,w variables and indices.

9.3 A Simple Example.


The Navier-Stokes equations are too difficult to solve in general. As a simple
illustration the flow between two parallel walls can be considered (Fig. 9.6).
In this case there is only motion in one direction, and only the equation of
motion in that direction has to be solved. Moreover, the flow is steady and
does not change in time.
When the flow is in x-direction the Navier-Stokes equation in x direction
has to be considered. Since the flow is steady the time-derivatives are zero.
January 4, 1994, Equations of Motion 147

\N
(y)

N,N

Figure 9.6: The flow between two parallel walls

Also the x-derivatives of the velocity are zero since the flow profile remains
the same at all x-stations. There is only a pressure gradient in x-
direction. The pressure gradient is negative since the pressure decreases in
x-direction. Because the flow is two-dimensional there are no derivatives in
z-direction and the v- and w-components of the flow are zero. This simplifies
the N.S. equation to:

ap a2u
(9.11)
ax = uy2
The solution of equation 9.11 is now straightforward:

au y ap
C1 (9.12)
ay ax +

=
2/./ ax
+Cy+C 1 2 (9.13)

This is the integrated equation of motion. The constants can be deter-


mined using boundary conditions. These conditions are:

V=

u(b) = u(b)=
where b and b are the locations of the wall in y-direction. With these
boundary conditions it has been assured that the flow at the wall is parallel
to the wall and the condition of viscous flow has also been imposed: the
velocity of the fluid close to the wall is zero.
The solution is C1 = 0 and C2 = . This leads to the solution:
148 G.Kuiper, Resistance and Propulsion, January 4, 1994

1 ap y2)
(9.14)
ax\
The velocity distribution between the walls is thus parabolic.

9.4 The Euler Equations.


Because the Navier Stokes equations are very difficult to solve, both analyt-
ically and numerically, simplifications have been applied. The first simpli-
fication is the neglect of viscosity. This removes all terms with p from the
Navier Stokes equations. The result is called the Euler-equations . From
the Euler equations some specific results can be derived . Consider the
stationary Euler-equation in x-direction:
au ap
ayy p ax
This equation can be -rewritten in terms of rotation au
ay t-1 as

ap lau2 au
--pax = ax v ay
au2 , au av , ay

la au ay
(9.15)
-,T;(v2 + v2) + v(-6 .T;)
This form can also be integrated in x-direction:

1
+ C = (u2 + y2)
2
v(
ay
)dx
ax
(9.16)

Eq. 9.16 is called the Euler equation . These equations of motion describe
the flow in an inviscid, rotational flow. The euler equations can be used in
flow regions where vortices, generated elsewhere, are present, but where no
new vorticity is generated. The path and behavior of the vorticity can be
calculated by Euler solvers.

9.5 The Bernoulli Equation.


The Euler equations can be simplified further.There are two possibilities to
do this:
January 4, 1994, Equations of Motion 149

along a streamline, where f v(t u )ds =Oz

in irrotational flow where Lu


ay ar =
In both cases the integration leads to the Bernoulli-equation

1 T72
p ,
v =C (9.17)
2p
When gravity is the only external force the constant can be written as
pgh + C, so that the Bernoulli equation takes the well known form

p + pV2 + pgh = C (9.18)


2

Note that in an inviscid and irrotational flow the Bernoulli equation


is valid anywhere in the flow (it has the same constant everywhere in the
flow) . In an inviscid but rotational flow the Bernoulli equation has the
same constant aldng a streamline. Every streamline has its own constant,
however.

9.6 Summary.
The equations of motion in the flow are a translation of the second law of
Newton:Force = mass X acceleration When this law is applied to a particle
in the flow the Navier stokes equations evolve. In deriving the Navier Stokes
equations use has been made of the law of Hooke, which relates the defor-
mation of the fluid with the required stress. Also the fluid is characterized
by a single viscosity in all directions and for all deformations. The class of
fluids with these characteristics are called Newtonian fluids.

The Navier Stokes equations can be simplified by neglecting the viscosity.


This results in the Euler equations. A further simplification can be made by
neglecting the rotation of the flow. In that case the equations of motion can
be integrated and the Bernoulli equation evolves:

p + 1/2pV2 C onstant
150 G.Kuiper, Resistance and Propulsion, January 4, 1994

t
Chapter- 10
Intermezzo:Potential flow
Objective: To show the use of singularities in potential flow calculations.

The equatins- of motion relate the pressures in the fluid with the flow
velocities.In an irrotational, inviscid flow the equation of motion can be
written as the Bernoulli equation, eq. 9.17. The continuity equation is given
in eq. 9.3. These two equations are the constitutive equations for an irro-
tational, inviscid flow. That means that they are valid anywhere in the flow.

A potential flow is now defined as a flow field which can be described by


a scalar function, generally indicated as (I)(x, y, z), which is the potential.
The velocities in the flow in any direction are the partial derivatives in that
direction of the potential:

u= ao
ax
ao
v=
ay

w. ao
az
The rotation of a potential flow e.g. in the x y plane can be written in
terms of the potential
au av a ao a8
ayxOyx ax ay
and this is always zero for a function which is twice differentiable. So
a potential flow is always irrotational

151
152 G.Kuiper, Resistance and Propulsion, January 4, 1994

In such a flow the relation between the pressures and the velocities can
be described by the Bernoulli equation.

When the continuity equation is written in terms of the potential this


results in
a2c, a2c.
(10.1)
ax2 + ay' + az2 =

This equation is called the Laplace equation . In order to describe a flow


field it is therefore necessary to find a solution of the Laplace equation.

The strength of the potential flow description is that it is possible to


define a number of elementary solutions of the Laplace equation. Because
the solutions are superposable, an arbitrary flow field can be expressed in
terms of superpositions of these elementary solutions.

10.1 Singularities in Potential Flow.


10.1.1 Uniform Flow.
The simplest solution of the Laplace equation is

(I) = Ux

Its second derivative in x-direction is zero and the derivatives in y and


z directions are also zero. So this potential satisfies the Laplace equation.
The meaning of this potential is a uniform steady flow in x-direction, since
the only velocity component is u = U.

10.1.2 Source.
Another solution is the potential of a source. In two dimensional flow its
potential is

el) = --11n(r) (10.2)


2r
r is the distance from the source with strength Q. It requires some algebra
to show that this potential satisfies the Laplace equation.

The velocities in radial direction can be found from the derivative of the
potential function to r, so
January 4, 1994, Potential Flow 153

o.
Vr =
2rr
There is only a velocity in radial direction because ve = 0.

Figure 10.1: velocity field of a potential source

This is called-a, source because the amount of fluid passing through a


circle at radius r is always a. The flow field of a single source is shown in
Fig. 10.1. When the sign of the source is negative it is called a sink.

In three dimensions the potential of a source is


a-
(I) = (10.3)
47r
The velocities in radial direction are found from

Vr ao
== 4rr2
ar
o.

Here the amount of fluid passing through a sphere around the source per
unit time is always Q.

10.1.3 Vortex.
The potential of a two-dimensional vortex is

r
o = --u (10.4)
2r

The velocities are now in tangential direction

ve =
2rr
154 G.Kuiper, Resistance and Propulsion, January 4, 1994

There is no velocity in radial direction since the potential function is


independent of r. Here F is the vortex strength. The integral of the velocity
around a contour in the flow is called the circulation of the vortex, defined
as

F= veds

The flow field of a two-dimensional vortex is shown in Fig. 10.2.

Figure 10.2: velocity field of a potential vortex

The two-dimensional vortex can be considered as a vortex element per-


pendicular to the plane.
In three dimensions the same formulation of a vortex can be used. The
vortex becomes a vortex element, and the velocity induced by such a vortex
element is in a plane perpendicular to the vortex element, as shown in
Fig. 10.3. The velocity induced by the vortex element dl with strength F
in an arbitrary point P can be written as
F sin O
dv = ds (10.5)
471-r2

For an infinitely long straight vortex in 3 D the induced velocity is

v= (10.6)
2ra

where a is the distance perpendicular to the vortex line. This is a well


known solution in electricity, known as the law of Biot and Savart
January 4, 1994, Potential Flow 155

Figure 10.3: Velocities induced by a line vortex element

An isolated vortex element cannot exist in a potential flow field, because


in such a flow field no vorticity is generated. So a vortex element has to
be part of a co-ntinuous vortex or it splits into two vortex elements with
different directions, but with the same total strength. A vortex can only
end on a solid wall, which is the boundary of a potential flow field.

10.1.4 Dipole.
An elementary solution of the Laplace equation which is also often used is
a combination of a source and a sink (a negative source). The source and
sink are infinitely close together, but the product of source/sink strength
and distance remains finite. This product is the dipole or doublet strength
iI
Consider a source of strength Q and a sink of strength -Q at a distance
1 from each other. The potential in an arbitrary point P is
Q 1 1

47r 7-1 r2 )
The factor between brackets can be written as
r2 r1
rir2
When l + 0 the product r2r1 r2 and the distance r2 r1 lcos0, where
0 is the angle between the axis of the dipole and the arbitrary point P.
When Q1 --* g the potential of the dipole can be written as

=
pcos(0)
(I) (10.7)
47rr2
156 G.Kuiper, Resistance and Propulsion, January 4, 1994

The velocity field belonging to a dipole with its axis in the x-direction is
shown in Fig. 10.4.

Figure 10.4: velocity field of a potential doublet

The velocity field indicates that dipoles can be written as vortices and
vice-versa. In fact it can be shown that a vortex can be written as a deriv-
ative of the dipole. So a velocity field containing dipoles is equivalent with
a velocity field with vortices.

Note that a dipole has a direction. This direction is the axis of the
position of source and sink. The potential of a. dipole is in fact the derivative
of the potential of a source, so
a
(dipole) = -(-9-720(source) (10.8)

This relation will be used in panel methods (see below).


The velocity of a source or vortex is infinite at the location of the source
or vortex itself. The mentioned potentials are therefore called singularities.
Note that a single singularity describes the flow field everywhere in the flow.

10.2 A Simple Example of Potential Flow.


The use of singularities to describe a flow field is now illustrated by the
combination of a uniform flow with a single dipole. A uniform flow in x-
direction has a potential function (I) = Ux. When a dipole in x direction is
put in a uniform flow the potentials can be added, so:

0(dipole) = Ux (10.9)
271-r
January 4, 1994, Potential Flow 157

The velocity in radial direction is then found from


ao N
v,. = = U cos 0(1 + ) (10.10)
2rUr2
For r > co this reduces to v,. = U cos 0, which is the undisturbed uniform
flow in x-direction.
There is another special location, where
N
r2=
2rU
At that location the radial velocity is exactly zero, so there is only tan-
gential velocity. When this location is considered as the boundary between
inner and outer flow it turns out that the combination of a dipole and a
uniform flow represents the flow around a cylinder!

Figure 10.5: flow around a dipole in uniform flow

Note that the tangential velocity along the cylinder is not zero, as it
would be in real fluid. This is because the solution is in potential flow,
which is inviscid.

The tangential velocity along the cylinder is also be derived from the
potential function:

v ==U.94)

At the location of the cylinder r2 =


sin 9(1
UR2
N
27rU 7
so

ve = 2U sin 6'
158 G.Kuiper, Resistance and Propulsion, January 4, 1994

So there is a sinusoidal distribution of the velocity on the sphere. The


maximum velocity at O = 7r/2 is twice the incoming velocity U. Because this
is in a potential flow (irrotational) Bernoulli's law can be applied to find the
pressure from the velocities. The inviscid pressure distribution can thus be
found. At the top and bottom position (0 = 7r/2) the velocity is maximum
and the pressure is consequently minimum, the minimum pressure is found
from:

Po +112pU2 prnim + 1/2p[2q2


Or

Pmin
1/2pU2
PO
=3
where Po is the pressure in the undisturbed flow. The value -3 is called
the pressure coefficient.

This example shows tliat a single dipole can describe the flow around a
cylinder. Combinations of singularities in the flow can describe almost any
flow patterns. When the flow around an arbitrary body has to be calculated
the potential flow theory is used to locate singularities in the flow and to
determine the strength of these singularities in such a way that the body
contour is exactly a streamline.

Because the pressure distribution in our example of the cylinder is sym-


metrical in x-direction there is no resulting dragforce on the cylinder. This
is a general feature of bodies in a potential flow (Paradox of d'Alembert).

The resistance of a body in potential flow is zero.

This is due to the absence of the boundary layer, which causes not only the
absence of a friction force along the surface, but also the absence of pressure
drag.

10.3 Forces on a Vortex.


It is easy to add a circulation around the cylinder by adding a vortex with
strength r in the center of the cylinder. The potential of the flow is similar
to eq. 10.9 with the potential of the vortex added. The radial velocity
January 4, 1994, Potential Flow 159

in eq. 10.10 does not change, because the radial derivative of the vortex
potential is zero. The tangential velocity component changes from eq. 10.11
into:

_ ve = U sin 0(1
N r (10.12)
2rUR2) + 2rr
The velocity distribution now becomes asymmetrical. At the top of the
cylinder with O = r the velocity is
N
= U(1 2rUR2) + 2rr
At the bottom of the cylinder the velocity is
N
vo=_, = +U(1
2rUR2) 2rr
The pressure distribution, which can again be found by application of
Bernoulli's law, also is asymmetrical. Integration of the pressure over the
cylinder gives after some algebra the resulting force pUr in the direction
perpendicular to the flow.

So a vortex does give a force, perpendicular to the flow. The relation


between the side force L on a vortex and the incoming flow U is called
the law of Kutta-Joukowsky:

L = pUr

Vortices are so important because they can represent a lift force in the
fluid. This will be used for the description of airfoil wings or propeller
blades. An example of the application of lift on a single vortex is the Flettner
rotor , shown in Fig. 2.28. The rotating cylinders generate a circulation
around the cylinder due to frictional forces in the boundary layer close to
the wall. In wind the result is a side force perpendicular to the wind.

10.4 Panel Methods.


Without proof it is stated here that any potential flow field may be written
as a surface integral over the boundary S of a source distribution and a
dipole distribution on S. (See e.g. [19]). The dipole distribution has its
direction normal to S. Note that the boundary surface S is the complete
boundary of the flow, as shown in Fig. 10.6. When the outer surface goes
to infinity the surface S can be considered as the surface of the body and the
cut from the body to infinity. When this cut has no thickness only dipoles
160 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 10.6: Boundary S of a potential flow field

Figure 10.7: Panelling of an arbitrary body

are present there. When there is no lift on the body a source distribution
over the surface of the body remains.

For a short indication of panel methods it is sufficient to consider a


source distribution on a body. This situation is different from the example
of a cilinder, because in that example the dipole was in the center of the
cylinder and not on the surface.

The boundary condition on the body is that the normal velocity on the
body is zero, so
04)
=O (10.13)
an
This condition is valid everywhere on the surface of the body.

The panel method is a discretization of this boundary problem. The


surface of the body is approximated by flat elements. At each element a
January 4, 1994, Potential Flow 161

constant source strength of unknown strength is assumed. So the source dis-


tribution on the body is also discretised. The boundary condition eq. 10.13
is now applied on one point of the element. This point is the control point,
where the boundary layer is fulfilled.

Consider the velocity, induced by an arbitrary source panel in one control


point Pi, as sketched in Fig. 10.7. The potential in P1 of an arbitrary panel
is
4) = UPanei as
fpanel zirr
where dS is a part of the panel. This integral is the influence function of
a panel on a control point. The main variable is the distance r between
the panel and the panel length and position S. The potential in Pi is the
sum of the influence functions of all panels and the normal velocity is the
normal derivative of this sum.
A singularity occurs in the influence function when Pi is located in the
panel which is tonsidered, because in Pi the radius r is zero. This leads to
a singular integral over the panel with Pi. Fortunately this integral has a
finite solution and can be analytically formulated.

When r is relatively large (say a few times the panel size), the integra-
tion can be omitted because the panel distribution can be considered as a
single source with strength a x S. The majority of the panels can thus be
treated as discrete sources.

The normal velocity induced by all panels in the control point Pi can
thus be formulated as the sum of the influence functions of all panels. Ap-
plication of the boundary condition of zero normal velocity results in a set
of N linear equations, where N is the number of panels. A matrix of N x N
results and this matrix can be solved.
There are different methods to formulate the problem, but the basic as-
sumptions are the same. Refinements are possible by using curved panels
instead of flat ones and by using linear or quadratic formulations for the
source distribution on the panels. For a short review see [12]. An introduc-
tion in the formulation of the problem is given e.g. in [19].

10.4.1 The Lifting Problem.


In the foregoing description it was assumed that there was no lift on the
body. When lift is present two important additions have to be made.
162 G.Kuiper, Resistance and Propulsion, January 4, 1994

The first one is that there are not only singularities on the body, but also
in the wake, which is expressed by the cut S in Fig. 10.6. Consequently the
wake has to be panelled too. This can be a problem because the position of
the wake is not always known. In case of a flat wing it can be approximated
by positioning it in the plane of the wing, which is only true when the dipole
strength or vortex strength is small. This wake model therefore introduces
again some linearization. On a propeller the problem is more complicated
because the wake has a complicated shape. This shape can be determined
only by the condition that the wake is in the direction of the (induced)
flow. The induced flow is again dependent on the dipole strength, and this
dilemma is characteristic for the propeller problem. Simplified wake models
will be treated in the sections on propeller design and analysis.

On the body and in the wake a dipole distribution is added to the


problem. Every panel has not only a (uniform) source strength u, but
also a (uniform) dipole strength of unknown magnitude. The influence
functions of the panels_ become more complicated as a result. The general
shape of the influence fun-ction is given in eq. 10.14.

1 a
(Dpi = + (10.14)
anelr arpdS
In this equation the factor has been absorbed in the strength of the sin-
gularities.

The boundary condition remains the same, but an additional bound-


ary condition is necessary because there are additional unknown dipole
strengths in the influence functions. A solution can be found by prescribing
as an additional condition that the flow at the aft stagnation point remains
finite. It is easier to consider a wing, where the flow velocity at the trailing
edge has to remain finite. This condition is called the Kutta condition and
it determines the strength of the dipoles or vortices in the wake. It adds
one equation to the matrix of influence functions.

A common problem of a panel method with lift is the flow around an


airplane, as shown in Fig. 10.8. The lift on the wings is represented by
dipoles over the wings and the wake. The fuselage is covered with sources
only.

Although the formulation is simple in principle, the proper panelling is


often not so easy. When the panels are chosen wrongly or inconsistently,
the matrix of influence functions becomes ill-conditioned and the solution
is no longer reliable.
January 4, 1994, Potential Flow 163

LIFTING STRIP
OF PANELS
N
EIOUNO
VORTICITY
TRAIUNG EDGE
N- LINES SEGMENT

N.
-.TRAILING
TRAILING EDGE VORTEZ WAKE

Figure 10.8: Panels over an airplane

10.5 Summary.
The velocities in an incompressible, irrotational flow can be described by the
Laplace equation and the Bernoulli equation. The latter equation relates the
pressures with the velocities:

p + 1/2pV2 = C
These flows are called potential flow fields.

Potential flow fields can be described with singularities. These are po-
tentials which are singular(infinite) in the core. Typical singularities are
vortices and sources. A body in the flow can be described by a closed stream-
line, which is generated by a singularity distribution in the flow. In a po-
tential flow no force is exerted on a body without circulation.(Paradox of
d'Alembert). When circulation is present the force is perpendicular to the
flow and its magnitude is

= pVF
where r is the circulation around the body contour.
164 G.Kuiper, Resistance and Propulsion, January 4, 1994

Single vortices have the property that they cannot begin or end in the
flow. They are either closed or are connected to a solid wall.

The velocity induced by a vortex in the flow can be found from its po-
tential. For a straight line vortex element this velocity is
given by the law of Biot and Savart, eq. 10.5.

The flow around an arbitrary three dimensional body can be described


with a surface distribution of singularities. When no circulation is present
a surface distribution of sources (and sinks) is enough. When circulation is
present also vortices or dipoles have to be distributed over the surface and
over the wake.
Chapter 11
Intermezzo: Boundary Layers
Objective: A basic approach of boundary layer theory and the definition
of a turbulence model

Consider the flow over a two-dimensional body in the x-y plane. The
mean flow is in the x- direction.The flow can be considered inviscid, except
in a region close to the body, where a strong velocity gradient is present. In
this boundary layer region the equations of motions are the Navier-Stokes
equations. When the flow remains attached the viscous region is thin rel-
ative to the chord of the profile and this thin layer is called the boundary
layer.

A two-dimensional body is considered, so the two-dimensional Navier-


Stokes equations apply. These equations are:

u(
au au1= ap +
+ v79-i
,a2u 82u,
--pOx iiax2+ 0y2)
ay ay 1 ap , 82y 82y
zt + y--( = p ay aX2 ay 2
Ox

+ -=o
au ay

In these equations the dynamic viscosity tt has been replaced by the


kinematic viscosity v, which is found from u =

165
166 G.Kuiper, Resistance and Propulsion, January 4, 1994

11.1 The non-dimensional Navier Stokes Equa-


tions.
For further use these equations are first non-dimensionalized by replacing

u by u' = u IU
x by x' = xll
y by = Yll
P by =
where U is the incoming undisturbed flow velocity and l is a relevant length
of the body. For one term this is done as follows:

u Ou
as
u (8-6,(u
u a?) (11.2)

u'(98::( U12)

This can be done for all terms in eq. 11.3, which leads to the non-
dimensional equivalent of the Navier-Stokes equation:

au lap u 82u 52u


au=
u-X - F v-5Y
- -pax+ U1(ax + -ay )
(11.3)
av av 1 ap a2v,
u--5-;+v, . ----F(--2-+
p ay
u a2v
Ul ax ay
) (11.4)
au av
0
Ox ay
In eq. 11.5 all primes are deleted, but all velocities and distances are now
non- dimensional.The quantity vlUl is the inverse of the Reynolds number.

11.2 The Boundary Layer Equation.


Terms which are small when the boundary layer is thin will now be ne-
glected. The boundary layer with thickness S is thin when < < 1. Because
the profile is assumed to be thin also the vertical velocity y will be small.
This leads to the following orders of magnitude of the non-dimensional
quantities:
u = 0(1)
x = 0(1)
January 4, 1994, Potential Flow 167

y = 0(8)

y = 0(6)

This makes it possible to identify terms of order 6. In x-direction eq. 11.5


is
u-
au
ax
au
y ay =
lOp
p ax
1 a2u
ax2
02u \
ay2)
The terms of this equation have the following order of magnitude:
,1 cl 1 Op 1 , ,
1 1
1- + 0- =
1 6 p ax
+ + -)
R, 12 62
The last term between the brackets can only be of order 1 when RT, =
OW, which is the case for large Reynolds numbers. Neglecting the terms
of order 6 the boundary layer equation is found:

au au lOp 1 t02u
(11.5)
u az ay = as+ ay2 )

It is important to remember that the Reynolds number in this case has


to be of the order 62 So when the boundary layer thickness is of the order
of 1 percent of the length of the body, the Reynolds number based on the
length of the body has to be of the order of 104. Still these equations are
only valid for laminar boundary layers.

In y-direction a similar procedure can be followed. The Navier-Stokes


equation in that direction is

ap lOp 1O2v 32v


az v ay P:5 1:?7,8x2+ ay
which has terms with the following order of magnitude
6 8 1
+ 6-6 = --pderppx + 82(1S + (1)
(52

All terms except the pressure term are now of order 6 or smaller, so the
equation of motion in y-direction reduces to
168 G.Kuiper, Resistance and Propulsion, January 4, 1994

This means that the pressure in the boundary layer is constant over the
boundary layer thickness. This is an important conclusion because now the
pressure outside the boundary layer can be identified with the wall pressure.

Outside the boundary layer y = O. Eq. 11.5 reduces for high Reynolds
numbers to
u --
au
ax
ap
ax
This can be integrated in x-direction, which leads to
1
p + 2pU2 =C

So outside the boundary layer the Bernoulli equation is valid again. The
fluid there can be considered as potential flow when y = O.

11.2.1 Scaling the Thickness of the Boundary Layer.


In the foregoing derivation of the boundary layer equations the assump-
tion was made that R a 1/N/75'. Inversely this means that S' a VT.
(The prime means the non-dimensional boundary layer thickness !) When
a model is now towed at scale ratio 25 at the full scale Froude number v ,
vgl
the ratio between Rns and R, is 125, where the index in stands for model
and s for ship. This means that

=
Smi

So the boundary layer at the model is relatively more than ten times
thicker than at the ship. Relatively, because the boundary layer is taken
relative to the size of the ship. This causes the scale effects on the wake
distribution and on the resistance.

11.3 Solutions of the Boundary Layer Equa-


tions: Blasius.
The boundary layer equation eq. 11.5 can be integrated when it is assumed
that the velocity profile at every x-position has a similar shape:

(11.6)
January .4, 1994, Potential Flow 169

Such a solution is called a similarity solution. The specific friction at the


wall is found from
,u,
Tw Ilk) y=0 (11.7)

Blasius assumed a simple similarity solution by taking the velocity distrib-


ution in the boundd,ry layer as

u(Y)
U
= C(Yr 6.

This made it possible to integrate the boundary layer equation in combina-


tion with the no slip condition at the wall. The resistance D of a flat plate
with length l and width b can be found from the integration of the specific
friction coefficient T over the length of the plate at both sides:

D r(s)ds
2bo
The solution of Blisius for a flat plate is
1.328
cf = (11.8)
R,
where
Cf =--
1 T2 r
2 P.'

11.4 Turbulence.
The foregoing boundary layer equations are valid for laminar flow. The
viscosity used in the equations is the kinematic viscosity, a property of the
fluid. Simultaneously a high Reynolds number is assumed, so the Blasius
solution is only valid for R>104 and until the boundary layer becomes tur-
bulent at R. 105

Turbulence is the occurrence of random motions additional to the mean


velocity. So the instantaneous velocity of a fluid particle can be described
by the mean velocity and a turbulent component:

u = T./ ui with 17 =
= i7 + y' with -/-)7 =
p= p' with /7 =
To illustrate the effect of turbulence the Navier-Stokes equation in x-
direction (eq. 11.1) is considered:
170 G.Kuiper, Resistance and Propulsion, January 4, 1994

au au ap a2u 82u
P(ti tW)
Substituting the instantaneous velocity in this equation and taking the
time average results in a new formulation, the term p can be written as

= P + P' =
The latter is true because the time-average of the turbulent pressure com-
ponent is zero. So this pressure term is not changed by turbulence. This
is true for all right hand terms, but not for the left hand terms. E.g. the
term uti can be written as

uau1 a2u
ax
- 23x2
(11.9)
1 a(ri2 + u'2)
2 ax
,22 ,

2 ax2 2 ax2
The turbulent equation of motion in x-direction becomes in this way:
au au or, 52,
ax
'17)
ay
=
ax + '1(ax2 ay2 )
1 .9%72 au'vi
2.p(-8 + ay )
The extra terms require further attention, because they describe the
effect of turbulence. The term asat will be small relative to the derivative
aui
ay vi in y direction, because the variations of the velocities normal to the
wall (in y-direction) are much larger than those in flow direction. To assess
the effects of turbulence only the second termat-4L'-''-
ay
is sufficient. Similarly
the termaS 2will be small relative to ay2 These term can therefore also be
neglected.
When it is assumed that the term -1pu'v' can be written as

it possible to rewrite eq. 11.11 as:


6,, ar, a2t,
p(riax+ T))
ay
= ax
+ IL
ay2
- AT ay2
January 4, 1994, Potential Flow 171

When written in this form is is clear that the turbulence can be taken
as an additional viscosity, also called the eddy viscosity. This is of course
.

due to the assumption of eq. 11.11. This assumption is called the turbu-
lence model, which in this case relates the turbulent viscosity in a linear
fashion with the velocity gradient in the boundary layer. Other turbulent
models have been formulated. No final turbulence model that can describe
the phenomena in real flow has been formulated yet.

The eddy viscosity is generally much greater than the dynamic viscosity,
so A, is greater than ft The determination of A is therefore important. A
well-known model is that of Prandtl, who assumed that

A, = p12
where l is the average distance of the turbulent motion in the boundary
layer. The model is therefore called the mixing length model. The problem
is of course to determine this length. In practice there is not such a single
length scale, but this model gives a conceptual physical background to the
turbulence model-of eq. 11.11.
172 G.Kuiper, Resistance and Propulsion, January 4, 1994
Chapter 12
Flow Calculations without
Waves
Objective: Review of the use of calculations for resistance and flow. No
extensive mathematical formulations are used, the purpose is to be able to
use the available programs intelligently.

A complete description of the flow around an arbitrary body is given


by the Navier-Stokes equations, provided that at high Reynolds numbers a
solution can be found for the formulation of the eddy viscosity, the ficticious
viscosity which is caused by the turbulence in the flow. The flow around
a ship hull can be found in priciple by applying the no slip condition at
the wall and by applying the appropriate boundary conditions at the free
surface. The free surface conditions are twofold. The first is the kinematical
free surface condition : the fluid particles at the free surface have to remain
at that surface. The second is the dynamical
free surface condition : the pressure at the free surface is always equal
to the atmospheric pressure. In principle this problem is well formulated
and using a volume discretization of the fluid it can be solved.

The solution of this problem is still not really feasible. One reason is
that the viscous phenomena require a very fine grid because the scale at
which energy is dissipated is very small. On the other hand the scale of the
waves at the free surface is large. In principle a large flow region with a
very fine volume grid has to be used for the solution and this is still beyond
the present computing capacity. Another reason is that the free surface is
a part of the solution. It is not known beforehand where the free surface
boundary condition has to be applied.

173
174 G.Kuiper, Resistance and Propulsion, January 4, 1994

Programs using either volume discretizations or surface panelling there-


fore contain essential simplifications. Only when the implications of these
simplifications are understood a proper use can be made of Computational
Fluid Dynamics (CFD).

The first and basic simplification is that the regions of viscous flow and
the regions of potential flow with a free surface are separated. This is an as-
sumption comparable with the Froude hypothesis, which separates viscous
and residuary resistance.

By far the most important calculation methods are those which calcu-
late the inviscid outer flow. The presence of a thin boundary layer makes
it possible to regard the flow outside the boundary layer as a potential flow
and the pressure at the wall is equal to the pressure at the outside of the thin
boundary layer. In that case the flow around a ship hull can be regarded
as a potential flow, because the viscosity can be neglected and as a result
no rotation is generated. The uniform inflow in front of the ship is also
irrotational, so the flow- aro. und the ship can be described by the Laplace
equation. The boundary conditions are the free surface conditions and tan-
gential flow along the outside of the boundary layer. Since the boundary
layer is considered thin this can be approximated by tangential flow along
the hull.

The viscous region can in principle be calculated using boundary layer


equations, because the boundary is considered to be thin. Efforts have
been made to use two-dimensional boundary layer equations along stream-
lines. The boundary layer is considered two-dimensional when no cross
flow perpendicular to the streamline occurs. The streamlines and the pres-
sure distribution along the streamlines have to be found from the potential
calculations of the outer flow. The results of such calculations give a distri-
bution of local friction coefficients along the streamline. Integration of the
longitudinal component over the hull gives the
resist ance.
These calculation methods have not been very successful. On an average
ship hull the effects of cross-flow and separation are too important to be
ignored, and the presence of a thick boundary layer in the stern region has
too much influence on the resistance. cross-flow are too important to be
ignored. Three-dimensional boundary layer calculations are complicated
and they blow up in regions of separation, where vortices leave the hull.
In those regions the basic assumptions of boundary layer flow are violated.
Until now no reliable calculation methods are available to calculate the re-
sistance. Efforts to do this are made using the full Navier-Stokes solutions,
as will be discussed later in this chapter.
G.Kuiper(MT512) January 4, 1994, Flow Calculations I page: 175

The separation of the viscous and the inviscid regions neglects the inter-
action between the viscous flow region and the potential flow region. The
main region where problems occur is the wake, where the flow is highly ro-
tational and viscosity cannot be neglected. Still it is a part of the potential
flow region. But also elsewhere on the hull where any type of separation
occurs, such as at the bilges in the forebody (see Fig. 4.7) the assumptions
are violated. Also when the boundary layer becomes very thick, as in some
regions in the afterbody, the solution of the potential theory will be inac-
curate. Keeping in mind these restictions the potential flow calculations
can be used intelligently to optimize hull forms or to calculate flow patterns
which are difficult to measure at model scale.

12.1 Potential Flow Calculations.


12.1.1 Panel Methods without Free Surface.
The simplest category of calculations of the flow around the ships hull are
the potential flow calculations without a free surface. The undisturbed free
surface is assumed to be unchanged by the flow around the hull. In practice
this is realised by mirroring the ship hull at the undisturbed free surface.
This so-called double hull is used for flow calculations in an unbounded
fluid. Because the undisturbed waterline is a plane of symmetry, the veloc-
ities normal to that plane will always be zero. A drawback is, of course,
that twice the amount of panels have to be used to cover the whole double
body.

When no lift is present the singularity distribution used is a source dis-


tribution. The sources are distributed on the discretized hull surface. The
elements of the hull surface are called panels. As boundary conditions the
condition of tangential flow at the ships hull is used. This condition is ap-
plied at the center of the panels, the so-called control points.
The simplest panel shape is a quadrilateral flat plane , on which a uniform
distribution of sources with constant strength is placed. This implies that
the distribution of the sources is a step-function instead of a continuous
distribution. This approach, first applied by Hess and Smith [11], approx-
imated the shape of a ships double hull with a large number ( N) of such
flat panels. On each panel a uniform source/sink distribution is placed with
an unknown source strength Q. The potential function (I)(x, y, z) can then
be written as the sum of the potential functions of all elements. The ve-
locity at an arbitrary control point can be expressed by influence functions
of the panels. The influence function of the own panel is singular, but the
176 G.Kuiper, Resistance and Propulsion, January 4, 1994

singularity can be integrated over the panel.So after some mathematics the
velocity in an arbitrary control point can be expressed as a linear function
of the N unknown source strengths Q.
Next the boundary condition of tangential flow is applied at each control
point. This results in N boundary conditions for the N panels. This system
can be solved, resulting in the strength Qi at every panel. The velocities at
every position at the ships hull and around the ship can then be found from
the derivative of the potential function (I)(x, y, z) in the desired direction.
Note that the panel size limits the accuracy of the derivative, because the
derivative has to be determined from the difference between two panels.
The pressures can be derived from the velocities using Bernoulli's law.

It is important to realize the consequences of the simplifications made.


First it is potential flow, so the fluid is inviscid. That means that necessar-
ily the resistance is zero (Paradox of d'Alembert) and that flow separation
does not occur. Just as in the case of a cylinder the pressure recovery in the
afterbody is complete. The water surface is also undisturbed, so the wave
resistance is also zero.

How can the results of such calculations be used? They can be used to
assess qualitatively the relative merit of various alternatives and therefore
to optimize the hull before model tests or further calculations are carried
out. They can also be used to indicate improvements, because the calcu-
lations provide data such as pressure and velocity distributions, which are
not measured. Examples are given below.

12.1.2 Assessment of Various Bulb Designs.


In Fig. 12.1 three forebodies are shown of a full bulk carrier. [18] For
these designs panel calculations were carried out, The results are shown
in Fig. 12.2. The dotted lines in this Figure are lines of equal pressure
coefficient. Since the pressure coefficient Cp is defined as 1P/2,2, a low
(highly negative) value of Cp means a low pressure on the hull.
A bulbous bow is used to minimize wave resistance, but excessive flow
separation has to be avoided too, because that increases the residuary re-
sistance again. The risk of flow separation is largest when the pressure
gradients are large. In this case the pressure gradients are largest when the
minimum pressure is low. From the calculations this minimum pressure oc-
curs at frame 19 near the bottom. The location does not vary significantly
between the alternatives. The forebody with the smallest risk of flow sepa-
ration is the one with the smallest perturbation of the flow, so the one with
the highest minimum pressure coefficient. Ship P1 has a minimum pressure
G.Kuiper(MT512) January 4, 1994, Flow Cakulations I page: 177

SHIP P1

SHIP P2

Figure 12.1: Two bulb configurations for a bulk carrier

coefficient of -0.35, which is lower than that of ship P2. This indicates that
ship P1 has a potentially higher residuary resistance.

The results from model tests are shown in Tabel 12.3. Ship P1 has in-
deed a significantly higher power requirement. So although the resistance is
not calculated, the relative merit of these designs could be estimated from
a comparison of the pressure gradient in the forebody.

The assessment of the pressure distribution from potential calculations


requires considerable experience and comparison with experimental data.
The most important application of this type of calculations is in combination
with experimental data, instead of as a replacement of experiments. Suppose
that ship P2 was the initial design. After model experiments is would have
been found to have a high resistance compared to e.g. statistical data or
to other similar ships. The experimental results give no indication of the
cause, nor any indication about the remedy. From the pressure calculations
178 G.Kuiper, Resistance and Propulsion, January 4, 1994

SHIP P1

SHIP P2
Figure 12.2: Pressure distributions on two different forebodies

it can be concluded that the fullness of the lowest waterlines should be


reduced, because the pressure gradient there is very steep. The pressure
distribution on an alternative design can also be calculated, to see if the
change has the desired impact, such as in ship Pl. When this is not the
case further improvements can be made. After such an optimisation the
new design can be tested again experimentally to determine quantitatively
the effects on the resistance.

12.1.3 Knuckles and Bulge Keels.


G.Kuiper(MT512) January 4, 1994, Flow Calculations I page: 179

Speed Ship P1 Ship P2


in
knots PD PD

12 100 84.6
13 100 86.5
14 100 88.3
15 100 85.8
16 100 85.7
17 100 90.3

Figure 12.3: Speed-power relations of three forebody configurations

The velocities along the hull can be calculated both in magnitude and
in direction. This can be plotted as a vetor diagram. This simulates a
"tuft test" in a towing tank, where small wires are observed to find the flow
direction and especially to find regions of flow separation. The calculations
will never give flow separation, but can only indicate the risk of separation.
On the other hand, the calculations give the magnitude of the velocities,
and thus the pressure distribution, which is not found from a tuft test.

In Fig. 12.4 an example is given of the calculated flow direction on a


hull with a knuckle. The calculations indicate that the flow along the upper
knuckle is tangential to the knuckle, but flow separation can be expected
at the lower two knuckles. These results can again lead to a change in the
design before model tests are carried out.

Similarly the calculation of flow lines along the hull makes it possible to
make an educated guess about the correct position of bilge keels or stabilis-
ers. It should be kept in mind that these calculations are applicable when
the influence of waves is small and when the boundary layer remains thin.

12.1.4 Assessment of the Afterbody.


The boundary layer in the forebody is relatively thin, and the calculated
velocities and pressures in the forebody are generally accurate, provided no
separation takes place. This is more questionable in the afterbody. In the
180 G.Kuiper, Resistance and Propulsion, January 4, 1994

.--

/1

'
. ,
// >,

I
\\ /

co
o
o

Figure 12.4: Calculated flow lines along a barge

afterbody the neglect of viscosity is more drastic and the results of potential
calculations are very much qualitative and should be treated with greater
care. As an example of an assessment of an afterbody configuration the
optimisation of a shaft bossing is shown. Fig. 12.5 shows two alternative
designs of a bossing. The question is at what angle the bossing should be
placed to have minimal interference with the flow.
The dotted lines in this Figure are lines of equal pressure coefficients
on the hull. The differences are small, so the position of the bossing is not
important for the flow over the hull. The lines of equal pressure
coefficient on the inside and outside of the bossing, as given in Fig. 12.6,
give a better indication to distinguish between the alternatives. The pres-
sure coefficient at the inner side of the bossing of ship J is much higher
than that on the outer side, indicating that the bossing should be directed
outwards. The pressures on both sides of the bossing of ship L are more in
equilibrium, and also the pressure gradients are smaller. The calculations
indicate that the bossings on ship L are the most favourable.

The required power for the three configurations were measured at model
scale. The relative merit from the model tests is given in Table 12.7. In this
G.Kuiper(MT512) January 4, 1994, Flow Calculations I page: 181

0=AP.

NO INCLINATION

SHIP .1

0= AP.

SKEG INCLINATION INWARD OVER THE TOP

SHIP L

Figure 12.5: Two alternative positions of a shaft bossing

table the power PE is the power required to tow the ship without propulsor,
as discussed in chapter 7.The power PD is the delivered power in the self-
propelled full scale situation, as will be discussed later. Ship L has indeed
the lowest power requirement at all speeds. The relative merit as deduced
from the panel calculations was therefore confirmed by model tests. The
interpretation of such calculation results requires caution and experience,
however.

If the bossings are in the proper position can be found from the wake
field. . In Fig 12.8 the measured axial nominal
wakes are shown. These
results show that the flow at the outer side of the bossing of ship J showed
separation or at least a very thick boundary layer. The wake of ship L
182 G.Kuiper, Resistance and Propulsion, January .4, 1994

_ACY21.
C. 0.07?.......z.yVG0 040 0.020
...040

OUTIEI
0 040
0060 -.060
00 11020 .020

A.?.

SHIP J

'Loot

A P.

SHIP L
Figure 12.6: Pressure distribution on bossings of two containerships
G.Kuiper(MT512) January 4, 1994, Flow Calculations I page: 183

Ship

SPeed
in PE PD int./. PE PD inw*
knots

20 100 100 92.8 90.6


21 100 100 93.1 91.2
22 100 100 93.2 91.8
23 100 100 93.4 91.3

Figure 12.7: Speed-power relation for two afterbodies of a containership

SHIP J
.. 10
... , .... -
....
------. s
...
/
..'....."''' V

/
\ /
/ 03
. r 1
1 12
t nt
Ct TA
....

... ..../ i
a 56 CLAAtk
r/p - n 14 44 c.k.
o
o 00 iao 270 360

SHIP L

e
10
,
,

, 12
\ ,
0.5

%
f
. .....0
ini .....,-.....;,...
n 7ft
Cn.A-R14.4_
0 '515.

72- a i t .4.:- C(4-


o
o 90 160 270 360

Figure 12.8: Measured axial wake distributions of containership


184 G.Kuiper, Resistance and Propulsion, January 4, 1994

confirmed that the bossing was approximately aligned with the flow.

12.2 Navier-Stokes Solutions.


The most direct solution of the flow problem is of course a full solution
of the Navier-Stokes equations. This is still too difficult and some simpli-
fications have to be made. Hoekstra simplified the numerical scheme for
the calculation of the N.S.-equations in such a way that he could solve the
N.S. equations numerically going from stem to stern. (Parabolisation of the
equation). In this way effects near the stern, such as separation, have no
effect upstream, which they may have in real flow. This can in principle be
solved iteratively in Hoekstra's method, but the time(and thus cost-)factor
often prohibits this. Hoekstra also neglects the waves and uses a double
model for his calculations.
The solution of Hoekstra is viscous and in that case the technique of
surface distributions is n.o longer applicable. The flow has to be divided into
three-dimensional elements troughout the flow, which of course increases
the calculating costs considerably. These types of calculations are typically
suited for supercomputers.
The solution of Hoeksta allows flow separation, because the flow is vis-
cous. The roll-up of the separated flow can be calculated and this allows
the calculation of the nominal wake distribution in the propeller plane. As
an example the flow along the ship given in Fig. 12.9 is given at various sta-
tions upstream of the propeller plane.(The propeller plane is at 2x/L=1.0,
where L is the ships length)

WOE
L AMA
,1 \6,._'""'% "IilIAMPANYMIN
L.N.._11Wil44mail
..WILSONE

%_ rlifMill
Ftia6 .irmiwza)ArAll 11611/11M
..=1 /AM, al.--.. 1

Figure 12.9: Hullform for calculation of viscous flow


G.Kuiper(MT512) January 4, 1994, Flow Calculations I page: 185

The interesting region for the calculation is the sharply convex region
above the stern bulb. The boundary layer from the forebody becomes very
thick in that region, so that a real boundary lauer does no longer exist.
The relatively thin viscous region at 2x/L=0.7 is shown in Fig. 12.10. The
contours of equal axial velocity indicate the axial velocity as a fraction of
the ship speed. The vector diagram indicates the transverse velocity at the
cross-section.

J tiffi0
Figure 12.10: Calculated flow pattern at 2x/L=0.7

At 2x/L=0.9 the viscous region is considerably thicker, as shown in


Fig. 12.11
The roll-up of the separated flow can be seen on the vector-diagram of
the transverse flow. This roll-up is completed in the propeller plane, as
shown in Fig. 12.12.
In the contours of axial velocity this results in the strongly curved con-
tours in the upper part of the propeller plane.
The effect of the propeller can be incorporated in these calculations
when the propeller is represented as an "actuator disk", that is a simple
uniformly distributed force over the propeller disk.
186 G.Kuiper, Resistance and Propulsion, January 4, 1994

Figure 12.11: Calculated flow pattern at 2x/L=0.9

t'----::::".
\
* \\ : '
\ \\,,\
%\\ \\\- \\\
.\\
...
,""

-.1,J1
\--
iiI\v\x\O
\\\\ \\\ \
Ill
\\\\\N

\\\ \\\ , ,
, \
/I 1111\
pil ,1110 \\ \\\ \ ,
\\N
iiii %%01\ \
\\
, t I I
'
\\III\ 1 I i

iitttttit\\\\\\\\
t

Figure 12.12: Calculated nominal wake


Chapter 13
Flow Calculations with a Free
Surface
Objective: Review of the use of calculations for wave resistance.

When a free surface exists the free surface conditions have to be sati-
fied. To illustrate the formulation of such conditions and to illustrate the
linearization of the problem the linearized free surface conditions will be
derived in two-dimensional form.

13.1 The Linearized Free Surface Condition.


Consider a simple two-dimensional wave system which is stationary behind
a ship hull which moves at a velocity Vs, as shown in Fig. 13.1. Since this
is a stationary situation (in a ship mounted system of axes) the kinematic
boundary condition is that the fluid velocity at the surface is tangential to
the surface:

077,v, vV!,as

A potential flow is assumed. The potential is taken as the potential


of the undisturbed inflow V,x and a perturbation potential 0, so that
= Vsx +
The velocities vr and vy can then be expressed as
ao
v, = + -a- -;

187
188 G.Kuiper, Resistance and Propulsion, January 4, 1994

Vs

Figure 13.1: Kinematic Free Surface Boundary Condition

ao
VY =

The kinematic boundary condition in terms of the potential 0 becomes

017 Vay
ax (1/3 pl)

This condition can be linearized when the perturbation velocities are


assumed to be small relative to the undisturbed velocity Vs. That means
that the last term can be written as
ao
o
(V8 ) = V3(1 -F ) = V3(1 -F 0( epsilon))
(9:c

where e is small. Note that also the g is of order e. Neglecting terms of


order 62 means that the linearized kinematic free surface boundary condition
becomes

1 ao
(13.1)
ax = V3 ay
This means that in the linearized kinematic boundary condition the hor-
izontal perturbation motions are neglected.

The dynamic boundary condition at the free surface is the requirement


that the pressure at the surface is equal to the atmospheric pressure. This
can be formulated using the Bernoulli equation
G.Kuiper(MT512) January 4, 1994, Flow Calculations II page: 189

1 0(1.2 (34:12

Pa = P 75. ) PgY
where pa is the atmospheric pressure. At the surface this reduces to
1 &V (91,2
/9( - ) Pg71 =
Written in terms of the perturbation potential 0 this is
1 80 2 802
2P{(u 1+ P9.17 =
Both a. and ay are of the order 6. Neglecting terms of order 62 gives
the linearized dynamical boundary condition
1 , g (13.2)
2 's ax
Elimination of the wave height i from both the kinematic and the dy-
namical boundary-condition makes it possible to formulate the free bound-
ary condition without the (unknown) waveheight 77. Differentiation of eq. 13.2
to x gives

a77 1 a2 q5
=
ax g 3x2
In combination with eq. 13.3 this results in
vs2 520 50
g ax2
ay
The variables can be made non-dimensional by a length L. So

Y=L
, x
x=
L
V
V' =
L
Omitting the primes gives the non-dimensional free linearized free sur-
face condition
vs2 520 50
(13.3)
gL ax2+ Oy =
The coefficient gL
11- is the square of the Froude number. As expected the
Froude number enters the free boundary conditions.
190 G.Kuiper, Resistance and Propulsion, January 4, 1994

The linearized free surface condition eq. 13.3, which is also called the
Kelvin condition 1 does not contain the wave height 77. It can therefore
be applied before the wave height is known. This is a consequence of the
linearization. However, it still has to be applied at the unknown wave sur-
face. It can be shown, however, by expanding the wave height into a Taylor
series in 77 that the boundary condition can also be applied at y = 0. The
error is than of 0(62). This makes it possible to eliminate the wave height
completely from the formulation of the problem.

13.2 Kelvin Sources.


The sources mentioned until now were solutions of the Laplace equation,
with the property that their influence disappeared at infinity. These type
of sources are also called "Rankine sources" , because it is possible to for-
mulate other types of solutions of the Laplace equations. For example a
basic solution exists which satisfies both the Laplace equation and the lin-
earized free surface condition. The properties of this solution are similar
to a regular source: fluid is issuing from the source in all directions; but
the formulation is much more complex. It will not be given here. These
sources are called "Kelvin sources" of "Havelock sources". Using a distri-
bution of these sources instead of regular sources the free surface problem
can be solved by applying the hull boundary condition. By using Kelvin
sources the free surface condition is than satisfied implicitly. When the po-
tential has been calculated, the wave height can be found from the linearized
dynamical boundary condition eq. 13.2.

13.3 Applications of the Kelvin Sources.


13.3.1 The Michell Theory.
A classical application of Kelvin sources was introduced by Michell (1898).
The undisturbed flow is the uniform inflow with the undisturbed free sur-
face. The disturbances due to the presences of the ship are described by
Kelvin sources. The basic assumption is the assumption that the breadth
of the ship is small relative to the length and to the draft. The approach
is therefore called the thin ship theory. The sources can consequently be
positioned at the ship centerline.
The power of this approach is that the problem can be solved analytically,
1Lord Kelvin or Sir William Thomson (1824-1907) was a British mathematical and
physicist who formulated the properties of vortices in potential flow
G.Kuiper(MT512) January 4, 1994, Flow Calculations H page: 191

which was a prerequisite in the pre-computer era. The solution of the prob-
lem had the form of eq. 13.4

AM= fi(0) ab(x'Y) f2(x,y,O)dxdy (13.4)

A is the wave amplitUde. The angle O is the wave direction. (For coordinates
see chapter 5). b(x, y) is the half beam of the hull, the x,y plane being the
centerplane of the hull. This integral makes it possible to calculate the
radiated wave height in all directions. Since the radiated waves moves
with the ship the wave velocity is found from V, = Vslcos(0) and the
corresponding wave length is found from the dispersion relation eq. 5.1. So
using eq. 13.4 the wave energy radiated in an arbitrary direction can be
found (see eq. 5.2. The integral of the radiated energy over the full circle is
the wave resistance of the ship. The result is Michell's integral for the wave
resistance [33]

2
D = -2L4
rvs2 f2ir sec 0 [.1 exp( "g sec' 0)(y ix cos 0) dxdy] de (13.5)
0 ax Vs2

This integral is not easy to solve analytically. A numerical approach is


better suited nowadays. In that case panels with a constant Kelvin source
strength are placed at the ship centerline. The boundary condition of tan-
gential flow can be reduced for a thin ship to

vY
ab=
ax V,

where y is in the transverse direction and x is in the longitudinal direction


of the ship. b is half the breadth of the local waterline. In the case of a
thin ship the vertical velocities along the hull are thus neglected.

Application of the boundary condition in the control points of the panels


results in a set of equations. The solution gives the strength of the Kelvin
sources at each panel.
When the strength of the Kelvin sources is known, the potential at each
location can be calculated.Differentiation with respect to x and y gives the
local velocities. Since it is a potential flow the Bernoulli equation can be
used to translate the velocities into pressures. Integration of the longitudi-
nal components of these pressures over the hull gives the wave resistance.
At the undisturbed surface z = 0 (with z as the vertical axis) the linearized
dynamical free surface condition eq. 13.2 can again be used to determine
the wave height. In practice only the wave height at the centerplane is
calculated, which gives the wave contour along the hull.
192 G.Kuiper, Resistance and Propulsion, January 4, 1994

This calculation is carried out with the model in a fixed immersion, the po-
sition without speed. Integration of the pressures in vertical direction gives
a force distribution over the hull in vertical direction. This force distribu-
tion can be used to calculate trim and rise. This method is very approxi-
mate. Further iterations, with a corrected wave resistance calculation in a
trimmed position, did not increase the accuracy of the predictions, however.

An example of a calculation of the wave resistance of a hull with LI B =


12 and BIT = 2.5 is given in Fig. 13.2.
A special element is the effect of a transom stern, which is quite com-
mon. In that case the hull is modelled as an open stern. The breadth at the
stern has a finite value. The pressure over the transom stern is taken as the
atmospheric pressure. The resistance"due to this transom stern is responsi-
ble for about one third of the residual resistance coefficient in Fig. 13.2.

3.00 -

8 2.50 -
o
,- e ..
2.00 -
e ...,
. . .
tp
o MIL

0
vi 1.50
ti)
CC

1.00 - Measured
471

cc 0.50 - Calculated

0.00
0.15 0.2 0.3 0.35 0.4 0.45 0.5 0.6 0.7 0.8 0.9 1

Fraude number

Figure 13.2: Wave Resistance of a Slender Hull Calculated with Thin Ship
Theory

This picture shows a general trend in the results of thin ship theory:
overestimation of the interference effects of the waves. The humps and
hollows are more
pronounced than found experimentally. The position of the hump is
properly predicted. The result can also be very sensitive to the number of
panels and their distribution over the centerplane. The calculation results
deteriorate rapidly with increasing BIT ratio. The calculated trim and rise
shows considerable discrepancies with experiments.
G.Kuiper(MT512) January 4, 1994, Flow Calculations II page: 193

13.4 Kelvin Sources for Catamaran Hulls,


an Example.
An application in which the thin ship approach is more suitable is in case
of catamarans, where the breadth of each hull is indeed small relative to
the length. ( The ipproach is used in Marin's program CATRES.) In that
case the Kelvin sources are positioned at the centerplanes of each hull, as
shown in Fig. 13.3.

Figure 13.3: Position of Sources on a Catamaran

The fact that only sources are used is a simplification, because it as-
sumes that the flow around each hull is symmetrical. However, the induced
velocity of one hull at the location of the other induces also an asymmet-
rical flow around the other hull, which can only be described by dipoles
or vortices on the centerplanes. The result is a sideforce on the hull and
a corresponding induced resistance due to trailing vorticity. This effect is
neglected.

The strength of the Kelvin sources at the centerplane is calculated using


the boundary condition at both hulls. This implies that hull interference is
194 G.Kuiper, Resistance and Propulsion, January 4, 1994

taken into account, because the velocities induced by both hulls are taken
into account in the calculation of the velocity in the control points.

The resistance is calculated by integration of the pressures in longitudi-


nal direction. Since the problem is symmetrical integration over one hull is
sufficient. The wave resistance is found by doubling the wave resistance of
one hull.

An example of the calculated residual resistance using this approach is


given in Fig. 13.4 The LIB ratio of each hull was 8.64, the BIT ratio was

12.00

co, 10.00
Q

8.00
a)

4.00 -
:0 Measured

cc 2.00 - - Calculated

0.00
0.4 0.45 0.5 0.6 0.7 0.8 0.9 1

Froude Number

Figure 13.4: Residual Resistance of a Catamaran

1.85. The hull spacing was 32 percent of the length. A good indication of
the residual resistance coefficient can be found in this way, especially below
Fri = 0.5. Above that speed the effects of sprayrails and trim wedges can
cause serious errors in the prediction.

13.5 Dawson's Method.


The "Thin Ship Theory" considers the ship hull as a perturbation of the
undisturbed water. The linearized free surface condition does the same with
the waves, it is assumed that the waves are small disturbances to the smooth
water surface. These are severe restrictions for practical applications. Daw-
son (1977) designed a practical method to overcome these restrictions. He
used the "Hess and Smith" solution of the double body without free surface
G.Kuiper(MT512) January 4, 1994, Flow Calculations II page: 195

as the base flow. The source strength of the panels of the double body
are thus used as a base potential. The difference between this double body
potential and the potential around the hull including a free surface is then
assumed to be small. This difference is thus used as a perturbation of the
double body potential.

This makes it possible to redefine the free surface conditions. When the
base potential as found from the double hull is used for the half body (the
ship hull until the waterline without its image) the vertical velocities at
the undisturbed free surface are no longer zero, because they are no longer
canceled by the velocities induced by the image hull. Using the dynamical
boundary condition eq. 13.2 these velocities can be translated into a wave
height. This is the "double body waveheight", on which the perturbations
are superimposed. In principle the free surface boundary conditions should
now be applied at the wavy surface generated from the base potential. Daw-
son formulated a free surface condition in terms of the double body potential
and the perturbation potential, which could be applied at the undisturbed
free surface. 2.

Having defined the free surface condition, a panel distribution over an


area around the ship at the free surface can be formed. At these free sur-
face panels a uniform distribution of regular Rankine sources is placed. Care
should be taken of the radiation of wave energy at the boundaries of this
area. In combination with the panels of the hull the total region at which
boundary conditions should be applied is covered. At the free surface the
panels have a source strength equal to the perturbation potential, at the
hull the panels have a source strength equal to the perturbation potential
plus the (known) double hull potential. At the free surface the newly formu-
lated boundary condition is applied, at the hull the condition of tangential
flow is applied. This set of equations can be solved.

This technique circumvents the restrictions of the thin ship theory. It


can also be applied to full hullforms, as long as the viscosity does not play
a significant role. It still has its limitations in the waveheight, because the
disturbance potential is assumed to be small. Application to moderately
high waves appears to be possible however.

2The transition from the double hull waves to the undisturbed surface has some
complications, as discussed by Raven [39]
196 G.Kuiper, Resistance and Propulsion, January 4, 1994

13.5.1 Applications of Dawson's Method.


An illustration of a hull form with a high wave resistance is shown in
Figs. 13.5. This wave pattern has been calculated with the Marin program
"Dawson" . It shows a very strong interaction of all wave systems involved.
After optimization using the Dawson program the wave resistance could be
drastically reduced, as shown in Fig. 13.6. Especially the interaction be-

j='''
. .
.......... - -4414116,
10Sile.
'
...................................
..
WAVE HEIGHT MAGNIFICATION FACTOR 5.0

Figure 13.5: Dawson Calculation of Wave Pattern before Improvement of


the Hull

tween the bow wave system and the forward shoulder wave system has been
improved. This optimization meant a complete redesign of the hull shape,
as the drawn frames in Fig. 13.7 show. Actually only the main dimensions
and the displacement were maintained. The redesigned hull form has a
softer forward shoulder with a more slender forebody, moving the center of
buoyancy backwards. The bulb is more pronounced . The midship coef-
ficient has been increased, resulting in a lower prismatic coefficient . The
afterbody has become very full afterbody, which may increase the frictional
resistance. This should be verified using other calculation methods or by
experiments.
Although the three-dimensional graphs of Figs. 13.5 and 13.6 are nice to
see, it is often sufficient to study the wave pattern along the hull, as shown
in Fig. 13.8 for the same case. Here the drawn line is from the improved
G.Kuiper(MT512) January 4, 1994, Flow Calculations II page: 197

Figure 13.6: Dawson Calculation of Wave Pattern after Improvement of the


Hull

hull form. Note that the waves contours at the stern are not too much
different. These contours are somewhat distorted by the (violation of) the
linearization of the free surface condition. Viscosity will also reduce the
height of these waves. The most important gain for the wave resistance is
in the forebody.

In practice it is never allowed to redesign a ship hull so drastically as in


the foregoing example. For reasons outside hydrodynamics it is often neces-
sary to maintain the length of the parallel middlebody (containerships), to
maintain the position of the center of buoyancy (trim) and to maintain the
deck area and midship section. An example of a more realistic improvement
in the wave pattern is shown in Fig. 13.9. The original design is the dotted
line. In this case it is important to reduce the interference between bow
wave and bow shoulder wave. The wavelength of the shoulder wave is indi-
cated as Ao in this Figure. The bow wave should be at a position where the
shoulder wave has a minimum. This can be obtained by moving the bow
wave forward, resulting in the drawn wave profile. This profile has a lower
wave amplitude along the hull. Although these realistic improvements seem
small, the decrease of the wave resistance is considerable.
198 G.Kuiper, Resistance and Propulsion, January 4, 1994

,..4rAT 2o

Figure 13.7: Hull Forms before (dotted) and after (drawn) Optimization of
the Wave Resistance

These examples show that the bulb cannot be considered as a separate


part of the hull. A Dawson type calculation is not done to design a bulb,
but to improve the shape of the hull including the bow form.

13.6 General Considerations to Assess Pro-


grams.
Many calculation methods are made available to customers commercially.
The calculations are usually done by specialized institutions, because the
generation of the grid for the calculations and the performance of calcula-
tions require some measure of expertise.
Simpler programs such as the "Michell" programs, are generally run by
G.Kuiper(MT512) January 4, 1994, Flow Calculations II page: 199

77.7.7..7Mfel,...1211Sr.

Figure 13.8: Wave Contours along the Hull before (drawn) and after (dot-
ted) Optimization of the Wave Resistance

STERN

WAVE HEIGHT MAGNIFICA1 ION FACTOR 120.

Figure 13.9: Wave Contours along the Hull before (dotted) and after
(drawn) Optimization of the Wave Resistance

the user himself. The user is confronted with programs which are "black
boxes" for him. To judge the possibilities of a calculation method it is
important to realize the simplifications involved. This means that it is
important to know:

which equations of motion are used.

where are the boundary conditions applied

which quantities have been assumed small (linearized)


200 G.Kuiper, Resistance and Propulsion, January 4, 1994

which quantities have been assumed two-dimensional


which quantities have been neglected (physical parameters)
which corrections have been applied to overcome previous simplifica-
tions -

which empirical coefficients or data are used

The simplifications are often hidden, because corrections are applied.


As an example a simple two-dimensional boundary layer calculation can be
added to the "Hess and Smith" calculation to calculate the viscous resis-
tance or even the wake pattern. These elements should be mentioned in the
documentation of the program. Examples of the documentation of some
previously mentioned calculation methods are attached to this chapter.
Chapter 14
Axial Momentum Theory
Objective:A description of the actuator disk model for the calculation of
the induced velocity and the ideal efficiency. Determination of optimum
loading distribution and effects of rotation.

The propeller induces velocities in the flow around it. The determina-
tion of the induced velocities in the propeller sections is a main problem
in propeller design theory. The simplest model to estimate the induced ve-
locities is the axial momentum theory, in which viscosity is neglected, the
number of blades is assumed to be infinite and the rotation induced by the
propeller is also neglected. The only action of the propeller is to excert a
uniformly distributed axial force on the fluid.

14.1 Axial Momentum Theory.


The fluid is considered to be inviscid and incompressible. The propeller
is described as an actuator disk, which is a disk with diameter D which
causes a pressure jump Ap over the propeller disk. Because the thrust is
distributed uniformly the pressure jump at the propeller disk will be the
same at every position of the disk. The flow pattern is axially symmetric
and a cross section is shown in Fig. 14.1.

Nomenclature: The inflow has a velocity Ve at a pressure po, which


is the pressure in the undisturbed flow. The diameter of the streamtube
far upstream is Do and the area Ao is therefore 0.25rDg.The streamtube
contracts downstream of the propeller in the slipstream until the diameter
remains constant. This diameter is called D2 and the area A2. The veloc-

201
202 G.Kuiper, Resistance and Propulsion, January 4, 1994

Ve

Po

NI
PROPELLER
DIAMETER

Vs +2Va Vs + va

Po

SECTION A2

SECTION A,

STREAMTUDE

SECTION Ao

Figure 14.1: The axial actuator disk model

ity is axial there and the pressure is Po again, with an increased velocity
vaa. (Note that in Figure 14.1 the velocity increase is 2va, or twice the
increase at the propeller. This still has to be proven.) At the propeller disk
the pressure rises from p upstream to p Ap downstream of the actuator
disk. The thrust is then ApAi, where A1 = 0.257rD2. The velocity at the
actuator disk is continuous because the water is incompressible and has a
value Ve va. It is assumed that the flow in the slipstream has no rotation,
so only axial induced velocities occur.

A relation between the axial induced velocities va and vea with the pro-
peller thrust T can now be formulated using the conservation laws for mass
and momentum.

The first relation to be used is the conservation of mass or the continuity


equation:
VeA0 = (Ve + va)Ai = (Ve va.)A2

From these two relations the diameters Ao and A2 can be written in terms
of the propeller diameter D:
January 4, 1994, Momentum Theory 203

D02 . [ve v+ Va 1,02 (14.1)

D22 = [ Ve ]D2
Ve + Vaa

The law of conservation of momentum equates the force excerted on


the fluid with the net outflow of momentum . The control volume is the
streamtube from Ao to A2. The mass per unit time through Ao is pV,A0 and
the momentum inflow is pl/e2A0 Similarly the momentum outflow through
A2 can be written and the conservation of momentum requires that 1:

pVe2A0 p(Ve vaa)2A2 T=


Using eqs. 14.1 this can be written in terms of the propeller diameter D
as:

T n2 / T7
pk,
\,
-r va)vaa (14.2)
4
The thrust can also be written as T ApS and the pressures and veloc-
ities at the propeller disk are related with those upstream and downstream
by Bernoulli's law, which can be applied over those regions where no force
is applied on the fluid, that is upstream and downstream of the propeller
separately.

11
Upstream of the propeller:

Po
1,
2PVe` = P + 2P( Ve + v. )2
1

Downstream of the propeller:

Po + 2P(Ve + Vaa)2 = P AP va)2


2P(Ve
Subtracting these two equations gives:
1
Ap = P(2VeVaa Vaa2) (14.3)
2
and a second formulation for the propeller thrust is:

]n == 1)2p(Ve lvaa)vaa (14.4)


4 2

'This assumes that no net pressure force is present on the outside of the control
volume. It can be shown that this is true when the fluid outside the control volume is
large. A proof is given e.g. in [7].
204 G.Kuiper, Resistance and Propulsion, January 4, 1994

From eq. 14.2 and eq. 14.4 it is found that vaa = 2va.

The axial induced velocity at the propeller is half the


axial induced velocity the slipstream

The relation between the propeller thrust and the axial induced velocity
is now:

T = D2 p(Ve va)2va (14.5)


4
The propeller thrust is made non-dimensional with the propeller area
and the inflow velocity V,

CT -=
'D21pV2
-4-2 e

where CT is a thrust coefficient indicating the propeller loading. 2 Eq. 14.5


becomes

CT =
V

Ve
)
Va

Ve

or inversely

Va_ 1 _1
(14.6)
V: 2 2V

14.1.1 Efficiency.
The induced velocity in the slipstream represents energy supplied to the
flow behind the propeller. This is due to the fact that the fluid "gives way"
when a thrust is excerted to it. The loss of energy is reflected in an efficiency
which is lower than one. To formulate the efficiency we use a reference
system in the undisturbed water instead of connected to the moving disk.
In that reference system the propeller disk moves with a velocity Ve and
excerts a force T. The power delivered by the propeller is therefore TVE. In
the slipstream an velocity 2va is present. With the mass flow expressed as
the mass flowing throught the propeller disk, which is equal to that flowing
through the slipstream, this represents an energy of:
2This thrust coefficient CT is different from the thrust coefficient KT based on rpm.
To distinguish between both CT will be called the loading coefficient and KT the thrust
coefficient
January 4, 1994, Momentum Theory 205

E108t = p(Ve + va)(7iD2)(2va)2 (14.7)

The efficiency of the propeller can be written as


TV,
-=
TV, + Elost
Using eq.. 14.5 this can be rewritten as

VeP(Ve va)f irDi22v.


qo =
Vep(Ve + va)14,0122va + 2va2piDi2(Ve + va)
or:

170(14.8) 1

1 -F tige.

The induced veloctiy va can be written in terms of the propeller loading


using eq. 14.6 as:

2
710 = (14.9)
1 + 1/1 + CT

This represents the maximum efficiency which is theoretically possible


in an inviscid flow with a propeller which does not introduce any rotation
in the slipstream. It is therefore called the ideal efficiency. Its value is a
function of the propeller loading and this relation is graphically shown in
Fig. 14.2. When an efficiency is claimed which is larger than the ideal ef-
ficiency, basic laws of nature have been violated or a simple error has been
made.

In this actuator disk model the thrust Ap is considered uniformly dis-


tributed over the propeller disk. This is an assumption. Below it is shown
that this situation of uniform propeller loading gives the maximum effi-
ciency, so the actuator disk calculations really give the ideal efficiency in
uniform inflow.

14.2 Optimum Radial Loading Distribution.


In the preceding section we have assumed a constant radial loading distri-
bution Ap on the actuator disk. The momentum theory can also give an
indication of the optimum radial loading distribution when an annular disk
206 G.Kuiper, Resistance and Propulsion, January 4, 1994

100

090

0130

070

ni
060

150

040

a30
025 a5 2 16

CT

Figure 14.2: Ideal efficiency as a function of propeller loading

with radius r to r + dr is considered. The energy lost in the slipstream is


similar as in eq. 14.7, but now for an annular ring with area 2irrdr:

dE(r) 1 p(Ve va(r))(2va(r))2dS (14.10)


2
Similarly the thrust of an annular element is:

dT(r) = p(Ve va(r))2v3(r)dS (14.11)


Note that here the total pressure force on an annular element is again consid-
ered to be zero, thus neglecting any possible interaction between the elements.

From these two equations for an annular disk the loss of energy in the
slipstream can thus be written as:

dE = vadT (14.12)
The aim is now to minimise the energy loss for a given propeller thrust.
An increase of the trust AT at an arbitrary radius ra gives an increase of
the lost energy AEa = va(a)AT. To keep the total thrust constant a similar
decrease of the thrust AT has to be applied at another radius rb. This gives
a decrease of the lost energy AEbva(b)AT. When va(a) is smaller than va(b)
the total efficiency is increased. An optimum is therefore obtained when va
is constant over the radius. In that case the radial distribution of the thrust
at the propeller disk is also constant, so this is the actuator disk model as
January 4, 1994, Momentum Theory 207

used above.

The optimum radial loading distribution gives a uniform


axial velocity distribution in the slipstream
_

This property will be used later when the hydrodynamic pitch of the vor-
tices in the wake is chosen.
Chapter 15
The Propeller Geometry
Objective: Description of the propeller geometry and the names and de-
finitions used to describe it. Once the geometry is understood, the figures
and the definitions in the text should be sufficient for further use.

15.1 General Outline.


A sketch of a propeller is given in Fig. 15.1.
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 ro-
tation is as viewed from behind, that is towards the shaft. In normal for-
ward operation a right handed propeller rotates in clockwise direction when
viewed from behind . The propeller in Fig. 15.1 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 of the tip is the propeller
radius. The propeller diameter is, of course, twice the 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 moves forward the propeller inflow is at its back.) Because in forward
speed the back side has a low average pressure and the face side has a high
average pressure (this pressure difference generates the thrust), the face is
also called the pressure side and the back the suction side.

208
January 5, 1994, The Propeller Geometry 209

Di recti on of
rotati on

Trailing edge- 7-----


Leading edge
Fi 1 1 et area

Hub

Shaft

Back
Face

Figure 15.1: Sketch of a propeller

The propeller hub is of course rotationally symmetrical 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.

15.2 Blade Sections.


Consider an arbitrary propeller, as drawn in Fig. 15.2.
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 Fig. 15.3.


The side which meets the flow is the leading edge of the profile. The
trailing edge is generally sharp. A sharp trailing 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.
210 G.Kuiper, Resistance and Propulsion, January 5, 1994

P/4

Direction of
rotation

Figure 15.2: Cylindrical cross section of a propeller blade

Y
Suction sid

Camber Thickness

Tail Nose

Pressure side

Figure 15.3: Geometry of a propeller blade section


January 5, 1994, The Propeller Geometry 211

The straight line between the leading and the trailing edge of the profile
is the chordline 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 Fig. 15.3. 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 t(x) of the profile (see Fig. 15.3).
The line through the middle of the thickness over the chord is the camber
line of a profile. The vertical distance between the camber line and the
nose-tail line is the camber 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 between 0 and 1, or
as percentages of the chordlength . When the same camber and/or thick-
ness distribution is used for the blade sections at all radii, as is often the
case, the blade sections can simply be described by this distribution and
the radial distribution of maximum thickness and maximum camber. The
maximum thickness and maximum camber are often given as percentages
of the chord length.
Variations in a given camber and thickness distributions are often made by
varying the chordwise position of the maximum. These positions are gener-
ally expressed in percentages of the chordlength, measured from the leading
edge. (A section is described 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).
212 G.Kuiper, Resistance and Propulsion, January 5, 1994

15.2.1 NACA Definition of Thickness and Camber.


An alternative definition of thickness and camber has been given by the
NACA organisation in the United States (now NASA). This organisation
has carried out a massive amount of experiments on profile series [1] since
the thirties of this century. In the NACA definition the thickness is mea-
sured perpendicular to the camber line. In the NACA-method the nose of a
profile is no longer the extreme of the contour, but the intersection of the
camber line with the contour. Since the camber line is connected with the
thickness distribution, the camber line in the NACA method always has to
be determined iteratively.

Figure 15.4: Two ways of combining camber and thickness

For a given profile geometry the Naca definition results into a different
camber line than the definition using the maximum length. As a result the
nose point will also be different (as will be the tail location in case of a blunt
tail). Inversely, the construction of the geometry of a profile from a given
camber line and thickness distribution results in a different geometry. The
difference in profile contour when both definitions of the camber are used is
especially apparent near the nose. This is illustrated in Fig. 15.4. For thin
profiles the differences between both definitions of camber and thickness are
small. Since propeller blade sections are generally thin it is generally not
mentioned which method is used to define camber and thickness.
January 5, 1994, The Propeller Geometry 213

15.2.2 Root and Tip.


Two important points on the propeller can now be defined more precisely.
The first is the tip of the propeller blade. This is the location where the
chord length of a cylindrical section becomes zero. This occurs at radius r,
which is the outer radius R of the propeller.
The second location is the root section. This is more difficult to define.
The root section of a propeller is the intersection of the blade with the hub.
However, the hub is often not a circular cylinder, but has a conical or even
more complex shape. Therefore, a (design) hub diameter is defined as some
average hub diameter in the root region. The blade section in the cylinder
at the design hub radius is the blade root section. This section is found
when the blade is extended virtually into the hub. In this definition of the
blade root section the fillets are neglected.
The plane perpendicular to the shaft through the mischord of the root
section is the propeller plane. It is used as a pine in the coordinate system
in which the propeller geometry is defined.

15.3 Pitch and Pitch Angle.


The cylindrical cross section of a propeller blade, as shown in Fig. 15.2, is
now developed into a plane. (Fig. 15.5). In this figure the x-axis is the
projection of the center of the propeller shaft. In this developed plane a
number of parameters can be defined.

The chordline 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. 1. The propeller pitch P is defined as the increase in axial direction of
the pitch line over one full revolution 27rr. The dimension of the pitch is a
length.

The pitch angle (1) is the angle between the pitch line and a plane per-
pendicular to the propeller shaft.

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 pro-
peller drawing (Fig. 15.9), as will be discussed later. A significant radius,
which is often used as representative for the propeller, is the radius at
r R = 0.7. If a pitch value is given in the case of a variable pitch distribu-
tion it is usually the pitch at 0.71t.
1When the blade sections have a flat face (pressure side), the pitch line is sometimes
defined as the line trough the section face instead of the nose-tail line.
214 G.Kuiper, Resistance and Propulsion, January 5, 1994

\-\
Propel 1 er r erence 1 in

Pro eller plane

Rake

Skew Generator 1 ine


i nduced
rake

Blade reference line

Figure 15.5: Expanded cylindrical cross section of a propeller

15.4 Propeller Plane and Propeller Refer-


ence Line.
The pitch and pitch angle could be defined without the definition of a co-
ordinate system. A coordinate system in which the geometry of a propeller
is expressed is always chosen as a cylindrical coordinate system (x, r, 0),
fixed to the propeller, with the positive x-axis as the shaft center line in
the direction of the propeller back. (see Fig. 15.1).The origin of the co-
ordinate system is chosen on the shaft center line in such a way that the
plane through the origin and perpendicular to the x-axis, called the pro-
peller plane. , goes through the middle of the chord of the root section. The
coordinate O = 0 in the propeller plane is the radial through the shaft center
and the midchord of the root. This line is called the propeller reference line.

In Fig. 15.5 the intersection of the propeller plane with the expanded
cylinder at an arbitrary radius is given. The x/-axis is the intersection of
the plane x, r, O = 0 with the expanded cylinder. The intersection of the
cylinder with the propeller plane gives another line, which is perpendicular
to the x'-axis. Both form the coordinate system in the plane of the ex-
panded cylinder.
January 5, 1994, The Propeller Geometry 215

15.5 Rake.
Having defined the coordinate system some other parameters can be defined
in Fig. 15.5. 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), thus 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 gener-
ator line.

Only in case of a linear rake distribution from root to tip the generator
line is a straight line in the plane z, r, 8 = O. 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 Fig. 15.6.

Propeller plane

Rake

h ft centreline

Hub

Figure 15.6: Longitudinal cross section of a propeller with rake

The axial displacement of the blade sections has little effect on the pro-
peller performance. It increases the wetted surface of the blades somewhat
216 G.Kuiper, Resistance and Propulsion, January 5, 1994

and thus decreases the efficiency slightly. Because the blade thickness is
measured in axial direction rake decreases the thickness of the blades when
measured perpendicular to the blade surface. This may become important
in cases of extreme rake.

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. Rake may also be used in
the casting process to prevent gas inclusions.

15.6 Skew.
The midchord of the blade section in Fig. 15.5 does not coincide with 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 Fig. 15.5and the distance between blade ref-
erence point and the generator line is called the skew. . When skew is in the
negative direction of O 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 . A propeller without skew has a generator line which
coincides with the blade reference line.

Unlike the rake, the skew distribution is never linear. Because the skew
is defined along the pitch line the skew distributions used can be better
shown in other projections than Fig. 15.5, as will be discussed in the next
section.

15.7 Blade Contours and Areas.


The projection of the blade contour on the propeller plane gives the pro-
jected blade contour. An example is given in Fig. 15.7a as the drawn contour.
The propeller reference line and the generator line coincide in this projec-
tion as the vertical axis. Consequently rake is not visible in the projected
blade contour. The blade sections in this projection are segments of a circle.

Apart from the projected blade contour the developed blade contour can
be defined. The blade sections in the cylinder of Fig. 15.2 are rotated
218 G.Kuiper, Resistance and Propulsion, January 5, 1994

blades and the area of the propeller plane Ao (Ao = 0.2571-D2, where D is
the propeller diameter). Two blade area ratios are used: the projected blade
area ratio Ap I Ao and the expanded blade area ratio Ae/A0. The latter ratio
is physically most significant and when no further indication is given this
blade area ratio is meant.

Because the skew is measured along the pitch line, the skew distribu-
tion can be plotted directly in the expanded blade contour, as shown in
Fig. 15.8. The skew varies over the radius. To indicate the amount of skew
as a property af the whole propeller the skew angle is used. It is the an-
gle between the blade reference line and the line from the shaft center to
the tip. This angle can be defined in the expanded contour. In the ITTC
nomenclature, however, the skew angle is rather inconsistently defined in
the plane of rotation, and the skew angle has to be drawn in the projected
contour, as is done in Fig. 15.8.

Skew angle extent -

Propeller reference line


Generator line

Blade reference line

Hub radius

Root

Projected contour Expanded contour

Figure 15.8: Projected and expanded contour of a propeller with skew

The skew at inner radii is generally forward skew, at outer radii back-
ward skew is applied, as shown in Fig. 15.8. Such a skew distribution is
January 5, 1994, The Propeller Geometry 219

called balanced skew. This is done to reduce blade spindle torque and to
avoid excessive stresses in the blade root, which would occur due to cen-
trifugal forces if the skew was not balanced. As an indication of the shape
of the propeller blade the skew angle alone can be misleading. In that case
the skew angle extent is defined in the projected contour. A drawback of the
skew angle extent is that minor changes in the skew distribution at inner
radii, which have little impact on the propeller performance, can greatly
influence the skew angle extent.

15.8 Warped Propellers.


Because of its definition skew also generates skew induced rake, which causes
the blade sections to move out of the propeller plane. For the same struc-
tural reasons as with balanced skew a balanced rake can also be required.
This is done using, negative rake together with positive skew, such that the
negative rake compensates the skew induced rake. In that way the blade
reference line can be kept in the propeller plane. Such a propeller is called
a warped propeller.

15.9 The Propeller Drawing.


A propeller drawing contains the elements discussed before. An example is
given in Fig. 15.9.
At the left the longitudinal projection on the plane x = 0, r, O = 0 is
given. It gives the contour of the hub and a projection of the blade with the
generator line in the plane of the drawing. The propeller in Fig. 15.9 has
no skew, so the tip of the contour is at the outermost radius. With skew
this is no longer so. In this drawing also the the envelope of the blade over
a revolution ( the clearance curve or sweep) is sometimes given. The tip
contour is given separately in detail. This tip contour is not a projection
but an axial cross section at the location of the blade reference line.

Next to the longitudinal cross section the pitch distribution is given


graphically. The distance to the generator line of the projected contour is
the pitch.

The middle figure gives the projected blade contour. It is characterized


by the circle segments of the blade sections. Above this contour are the
details of the leading and trailing edges of a propeller section. The trailing
edge is not smoothly rounded, but has a knuckle at some distance from the
220 G.Kuiper, Resistance and Propulsion, January 5, 1994

Anti
singing
2g9e

LW IINErr:"-
smiumnomvAmair-
=EMS.

=1 111MW,
=1MM M1%1W

Figure 15.9: Example of a propeller drawing

trailing edge. This is the anti-singing edge , which serves to fix the separa-
tion of the flow from the blades. If there is no anti-singing edge vibrations
of the blade (singing) may occur. The vibrations are caused by excitation of
vortices leaving the trailing edge periodically, often at an audible frequency.

The figure at the right hand side of Fig. 15.9 is the expanded blade con-
tour. The nose-tail line is a straight line here and the shape of the blade
sections is given relative to this line. In this drawing the location of max-
imum thickness is also drawn. In Fig. 15.9 this location differs from the
midchord line at inner radii only.

To illustrate the effects of camber and skew on the propeller drawing


these parameters are varied in Fig. 15.10. The interpretation of these dia-
grams should be clear now.

15.10 Description of a Propeller.


The propeller geometry is generally defined by the following data:

1. Number of blades.
January 5, 1994, The Propeller Geometry 221

!KM
wow.
M1111111101111MIN
111119EM
WNW
Iimat2.07. Illrwat

Without skew and rake

With rake

,....---- =11.........,._
rwwm,...-.....-
anlaga.

imal...wiiik
rmm.
Mill.w- mi.miiiii_..
MIMI
MEN .
EllE1.111/MMIIIITMEINI111
MIR 'IMMIMMMINNER
WM!
MBE"
7.......
-,..,,,.
1111111111
1.1.....sii
poom......,
.-- - ''' -- '4

With skew

FASi MI
MMIS'IN
MIMM1 %-- MIIINI,,
,
1111170211-1111111111
INI
t-- INM11
ii=r.-/. IIMINMMIM
----.._
IM
'W.-

IIIIIIWIIIIINMIIVAIIIIIIIEMIMILM
IIMMIIIIIVIIMMINIMMILM111
RIMMIIINIMMIll -IIW =wassi
111111111=111111 -I1W
MMINUIMMI2017 ,
IIP=s:
0--

With skew and rake

Figure 15.10: Effects of skew and rake in a propeller drawing


222 G.Kuiper, Resistance and Propulsion, January 5, 1994

Diameter.
Radial distribution of rake.
Radial distribution of pitch.
Radial distribution oi skew.
Radial distribution of chord length.
Type of camber and thickness distribution. (locating also the posi-
tions of maximum camber and thickness)
Radial distribution of maximum thickness.
Radial distribution of maximum camber.
details of anti-singing edge.
hub shape
root fillets.

Instead of the non-dimensional camber and thickness distributions and


the radial distributions of the maximum camber and thickness the section
contours (pressure and suction side coordinates) may also be given. In that
case the definition of the nose region requires much attention and at least
50 points over the chord are required for a proper definition.

The precise definition of the root fillets is still lacking in practice. The
shape is often indicated by a radial cross sections of the blade. More pre-
cise definitions are required, especially when the blades are being milled
numerically.

15.11 Controllable Pitch Propellers.


A special type of propeller is the controllable pitch propeller,, often abbre-
viated as CPP or CP-Propeller. The blades of such a propeller can rotate
about an axis perpendicular to the shaft centre line, the spindle axis. The
spindle axis is the propeller reference line of Fig. 15.5. In exceptional cases
the spindle axis does not pass through the centre line of the shaft.

Because the hub is a complex mechanism in this case, the blades are
manufactured separately and mounted to the hub. The blades end in a
circular disk, which is called the palm. This disk is bolted to another cir-
cular disk in the hub, the carrier, which can be rotated mechanically or
January 5, 1994, The Propeller Geometry 223

hydraulically.

The variation of the propeller pitch poses additional constraints on the


hub geometry, because the blades should fair smoothly into the hub at all
blade positions. This requires that the palm has the shape of a sphere. This
is often not so and -then there is a clearance between blade and hub.

In case of CP-propellers the propeller reference line is not always chosen


through the midchord of the root section, but at another position which is
favorable from a manufacturing point of view. The spindle axis can then
be used as the propeller reference line, but it should be kept in mind that
such a shift of then reference line can strongly change the values of skew
and rake while the actual blade geometry is unchanged.
Chapter 16
Systematic Propeller Series
Objective: Propeller design in uniform flow using systematic test results

The performance of a propeller is characterized by its open water perfor-


mance as represented in the Kt Kg J diagram. This diagram will be
explained first.

16.1 Open Water Diagram.


As shown in the description of the propeller geometry the blade sections
have a certain pitch. When the propeller moves forward over that distance
during one revolution the chordline of the blade sections is in line with the
flow along the blades. When the forward displacement over one revolution is
smaller the propeller will develop more thrust. An important parameter for
the thrust of the propeller is therefore the axial displacement per revolution,
or the ratio V/n. This can be expressed in non-dimensional terms as the
advance ratio J

T
nD

where D is the propeller diameter and n is the rotation rate (sec') and V,
is the undisturbed velocity upstream of the propeller.

When the axial distance covered per revolution is smaller than the pitch
the difference is called the slip of the propeller. In non dimensional terms
the slip is expressed as PIDJ. The ratio PID is the pitch ratio. The slip
can also be expressed as a percentage of the pitch ratio. The slip is used
in older literature, but has been replaced by the advance ratio in modern

224
January 5, 1994, Systematic Propeller Series 225

literature.

In the foregoing the advance velocity has been made non-dimensional


with the velocity nD (proportional with the tip speed). The thrust and
torque can be made non-dimensional with the same speed and with the
propeller diameter-D. The result is the thrust coefficient IfT

-KT = pn2D4

and for the torque the torque coefficient KQ

A-Q
pn2D5

where T is the-thrust in Newton, Q is the torque in Nm. These are the non-
dimensional parameters in which the propeller performance is expressed.

These non-dimensional parameters can also be derived from dimensional


analysis. Two additional parameters are introduced now: the propeller di-
ameter D and the rotation rate of the propeller n. As has been mentioned
in chapter 6 the introduction of a propeller with diameter D introduces
a non-dimensional parameter DIL, which means that in model tests the
propeller should be geometrically scaled. The introduction of the propeller
revolutions n introduces the advance ratio J.

The propeller performance in uniform flow has the characteristics as


given in Fig. 16.1 1. In addition to the thrust and torque coefficients the
propulsive efficiency of the propeller is shown. The efficiency is the ratio
between the delivered power by the torque and the effective power of the
thrust. The power delivered to the propeller is the delivered power PD:

PD 27rQn

The effective power PE is the power delivered by the propeller thrust

PE = TV,

'Note that the torque coefficient is multiplied by 10 to separate it from the thrust
coefficient
226 G.Kuiper, Resistance and Propulsion, January 5, 1994

0.7
ETA

0.1

0.3

0.2

0.1

0.0 0.0
0.2 0.1 0.6 0.8 1.0
AOVANCE COEFF IC I ENT J

Figure 16.1: Open water diagram of a propeller

The velocity V, is the entrance velocity. It is the velocity of the propeller


relative to the undisturbed water. With propeller mounted coordinates this
means that the entrance velocity is the undisturbed velocity far upstream
of the propeller. The velocity in the propeller disk will be higher due to the
induced velocities, both in axial and tangential direction. The axial induced
velocities were calculated in the actuator disk model of chapter 14.

The efficiency of the propeller is the ratio between delivered and effective
power:

PE
770 =
PD
January 5, 1994, Systematic Propeller Series 227

This efficiency is defined without interference between propeller and


hull and with uniform inflow. These conditions are met when a propeller
is mounted on the front of a sting. It is therefore called the open water
efficiency.. 770 can be written in terms of thrust and torque coefficients as:

KtJ (16.1)
qo =
27rit,q
The efficiency is commonly plotted in the open water diagram as shown
in Fig. 16.1. Note that at a certain advance ratio the thrust coefficient
becomes zero. Then by definition the efficiency is also zero. This condition
will be close to the zero slip condition and thus depends on the propeller
pitch ratio.

16.2 The Quality Index.


At zero advanCe r-atio the efficiency is also zero. This condition is called the
bollard condition . The thrust coefficient is a measure of the thrust delivered
at zero speed. This can be an important design parameter (e.g. for tugs)
but the efficiency loses its meaning in that condition. In that condition the
quality index is sometimes used.
The efficiency indicates the quality of the energy conversion. It does not
say very much about the quality of the propeller itself, because a heavily
loaded propeller will always have a lower efficiency than a lightly loaded one.
A measure for the quality of a propeller is therefore the ratio between the
ideal axial efficiency, as derived in chapter 14 in eq. 14.9, and the efficiency
as defined above. This ratio can be written as
110 TVe(0.5 0.5V1 + CT) KT j + J2 + -8 KT)
27rmn 47rKQ V 7r

This "quality index" does not go to zero in the bollard condition, when
the advance ratio J is zero. Instead it is
K1.5
Quality Index (J=0) = T
VT7-1-KQ

16.3 Systematic Propeller Series.


The open water diagram gives a characteristic of the powering performance
of a propeller. Systematic series of propeller models have been tested to
form a basis for propeller design. The starting point of a series is its parent
form. The extent and applicability of the series depends on the parameters
228 G.Kuiper, Resistance and Propulsion, January 5, 1994

which are varied and on the range of the variations. There are several se-
ries, but one of the most extensive and widely used series is the Wageningen
B-series. The basic form of the B-series is simple and it has a good per-
formance. The extent of the series is large: some 210 propellers have been
tested.

The basic characteristics of the B-series are shown in Fig. 16.2.

Pitch distribution
1.0R 100%

0.045
B4-40
4 ' B4-55
_
' 84-70 i B4-85 .B4-100'

1.0R Pitch distribution

0 8R
.....,...---
_

0 7R \
0 : \ _____ _____
I
0
0
I
:
, :

R
t-
XII
X

W I
m v.
, _..........
___-_

I P11.1SE2 I i M Iii X.
.....
EEFFW
___

C" - 11"1!
x
_- - - -
85-45 B5-60 1 B5-75 B5-105/
tD --
, 0.040

Figure 16.2: General plan of the B-4.40, B4.55 and B-4.70 propellers

As shown the B-series propellers have


a constant radial pitch distribution at outer radii
very little skew
15 degrees backward rake angle (linear rake distribution)
a blade contour with fairly wide blade tips
circular back blade sections of NSMB-design
2NSMB:
Netherlands Ship Model Basin, nowadays MARIN: Maritime Research Institute
Netherlands
January 2, 1994, .Systenta,tie Propeller Series 229

The followiag parameters have been varied

the expanded blade atea, ratio AE/Ao from 0.30 to 1.2


the number of blades Z from 2 to 7
the pitch ratio D from 0.5 to 1.4

The propellers are indicated by their blade munber and blade area ra-
tio_ Propeller B-4_85 e.g.. Has four blades ;qTfd a bia.de area ratio of 0.85.
From each propelle,r an open water diagram NVaS I-rte.-as:urea. . Until now 210
propeller models have beert tested_ The re.suits are given in. open water di-
agrams per series of one blade number and area ratio. An example is given
in Figs. 16.3

The open water tests of the B-series were done at various rpm, so at
a variety of model Reynolds numbers, The B-series diagrams ha.ve been
corrected to a Reynolds number3 of 2 x 106 along the lines of the IT'TC57
method., as will be discussed in chapter 20. The correction is only

16.4 Propeller Hull Interaction.


The propeller -works behind the ship hull. Before a propeller can be designed
from open water diagrams it is necessary to estimate the interaction between
hull and. propeller. In this preliminary design stage this vy-ill be done in a
very simple way -using the wake fraction a_nci the thrust decluction factor.
The velocity V, as used in the open water diagrams? is the velocity far
upstrearn of the propeller. Behind the hull without propeller the velocity
at the propeller disk is called the nominal wake_ This wake is -Used as the
velocity- V. This assumes first that the nominal wake fraction is the same
in the propeller plane (w.here it is measured or calculated) and in a plane
several propeller diameters upstream of the propeller plane. It assume.s next
that the wake, is uniform over the propeller disk, bec-ause the inflow is used
in open water diagrams. where the irlow is uniform. ft finally a.ssnmes
that the nominal -wake distribution in this upstream plane is not afreeted
by the propeller aciion.

r-For the propeller inflow the nominal wake fraction is


used: Ve Vr,;(1 tu,)

Tb.e propeller has an effect on. the -hull, however- The propeller increases
the resistance of the ship bull by increasin!-4 the velocity along the hull
3A Reynolds number based on the chord len,g-th and inflow -velocity- at 0-76R.
230 G.Kuiper, Resistance and Propulsion, January 5, 199.4

1.6

NI
::::::
1 Yoe
mImi
It.=1 ma
..... IMO

am
a
Imam. !ROMEO asmaMiewsea 22:22:22:22..
E:
amt.
imam.

..
... ......
..................... .

.....
IN

..................

=:::
Moo..
=2EEEEE ttritr:1271
1. 3 EINittl.:PIIHR1111111111.5PREPn:21:7411155r:
Oa ..... ............... 11MerraiMa
111.. 111.011.1=1. One

1111 sow.

11
= =2
reala I... itsay
.................. .:::::::::.
. se ....... IN .........

"'" .. e
......ma...ma
......................................................................
.... ..... am

...................
-,
...... SONO
...Ea, a..- mama
..
:::,===m;
-::::.-:::::::...ttn.==;==:::.;:::::.:::::::::::..zraLr.:..=
..::::..- --v. :::.==::::1:::::::-.-::::::-:- ....
to - :===:: -.. . ------::::::::::=::::::::::::::::::::::::::::::. L -::::-::::::::::::::::: ..........
=7.

=
,,,..... a".11,
..4,
+- ,=1
----_____.....=..... --4 ,::::::::::::=.....::::. +-

-,rnI,
i=,,. ..:IBEL:r:iiiiii:naiL":":**` =:-.::-..r."-:::::=:::::::=4--r.,:::,:: ::;;;::::;;::;::::: ............

.,, ...
Q9
,alta 01. Mailaaama ......... as .=..

a .,
10KQ

:Nmaegm
aimm, .1=122=2:2="..."- ..
Tag

'
rMe

go. ...........=::=: !Ma,.


"W... .. ..... tt...ttr.ta
,:;.:00::............ -ylt,
..""r....
MM.
Ga.*
11:..111

07Mag..11.11111MIttle =0=g===....f.14.1110e
.:= 41===.1111rWIKIIKAVIIIVIN
=.1.awaaamego.easaaltswa............M.."

. .s...-....
.
.........11111...-1

MIMINM!..11
. ................... fia
,..!111111101111..Mal=01111iMaYaalndINO.,../Mi
1.--" 2:-..:===
ar11
."...
..... a no
.1.111 . ---- ---------
=:..=:`,........::6"""'m.
MOISMINIIMMYMINE.5.8.01MaYlame=181111

...._...=-=......=71.......-....-,Nr,O1111,...... =1:11.
.........."-=:...i..-=Fro::::.22::;a:It.::::;tittaIi-2::. 11.....2:2:2........t:Mr
....
11 ..... ....::
........
NOV: ,111/ 41.1.7,e0.0,11111M,-.11........ : .-.:.
.......................s. ..........--......,.....a...m..: .:-...n. - -=::::::=1----- ........ ::-
1110114.101.a.M: Mall Sall., .11 001.11:=
.. ..
Vim
.............
1

0 5 .. IMI,-,11,3 la,.../
_.........2===...-zzasr-..- am=1:::::::116,2 22223:22:=2........k....."............
=mill,...e, -,.-.--e
___11=111...
....
.v....7=........
11,
..... 2:
in.._ 111 MM./. ,,allp70On/:...--/ 1=t;=:21221.7....
=1:=:=.'."."..."1...:::::::P"... :
:22...""1"."......2%.% Elk,. :0"...i
=11/ albs /..."2""'"'""1,1:. ""'"'i"te".",`"":7""'":""*.r"=
.....-...-.............................................................,..k........,.....--..............................:::
.._,....p..4........z.r.m......r...........=.1... ,....-....."=.....=1,.........,04:: :.........nIMMOD ......16...:=V::...=....3.=..... .....
MOIR

.:::r::::::......:::
....._==,..;wrzsa:ft.4e12.-.4.,..4mgrwrar2
als....... ..=.,,,,.,,ame.416,,,,airr..,140..
IniallIMM 1 111"111
I =14::====::::::=1;3:
aa 111.11-aa
owe!. alalamimaal
Om marnata
ra
a.**11 MO
E. s ...01
Me WEI. MilamelOalnle

K
0 4 .....r...............7.......:,
10...
_1111,,MMINIIIIIManallalin.
EIDOMOIre
WASS, 111./... &MM. 11.1111M0/00. Wage 111111. IFO.
!41104:1;falLZ.,1DaeraarAibiZnte=a1=111:71"1.==.1:1=1=4.41.tntrar"....1-'-"
Ar, pr., How 11,1011MMINI,11=epoeMlaiRaeMIDEaly,a...11.,,./alMea
=1:1............-11::,.......:::::.74,..1.....a.........._........_.1.................. OS
%.
MO ........ = .. mamas.
11Fal.=1.
rj.N....=::..ar=:t..=s=:::.:::=nd.j,..........
8SOO
liammea Mill."
a
was= mom
"'"'::
taa
=....z..-...............:-
Iry
114.11

,...........^..,11/471.71-
.........,...............z.=......................_..1...,..===1....=....1"=.1.1,.;=====:,
.....,..
,...........:....1._........ "rams - Mae name. 1
.i.................r........nram..r.14%.: -,..a....z.... =2
MM.RSA 1.1111
IN ...... 2 .. -
:.=== .42= .. 1 ..... :::::=:: SONO ::
7 _7 _.......,...,......, ....0 Ray Mt ,11.../.11911.,MON11!, ...,.
..,.....rz.t1;4:..:..tII:e:dLr4:_t=.'%.._I..al=:=11=t,.I._;:nrz:r._.;:..::.....;:;:z: .... ..................... MINIMS Mil.
iIIIIINIWITY11.::4====11.6. ::.1Iltaa
.0,. .-ma "mow., ,..1111111111
=:...7.74,0!:....1.1.-:14=....M.Z.7...===ezwi=E:....7:::3.,,,,:+ii ii,,,n,.... 1:.,.=:-...................................1.11m4.1
_.........=,:r_..:1.:a::.frarin r-..:,: Ma;::041=a=
C.Ima,S11,111.1.1...
,...redieritr,,11LINNION....,8117.11.00
02 7............,,,
Ai -74:044,:ate;:4 41,:: ======.,--:...r.
igs es==
..... !Sill MI: SA01

elm , NM AV..., 9.11111. WSO/a.., ..... OM.; nme.:.a.


Mary =6.1 =Warr :."'"'"*"*":111..,..:::::
,.....,.....yrez.,../.1.....-....._-......-..................m................4....6.77I...!.....1:11=14!1"=1.:== z.....r.....MS:r.r.r+.1.....,.......
.e.' -rx-
_
--;:.=.........-4
....._. ::,
o 1 =://40-=""......V.:nr..-Z.M26-..-===ZTr.=:`-.:Z_ZnIT:IMINC.T.T.ZICIFI:-..;...:::::ht'7:1:Z7."'"
=.......___s_.....,...=.........,........=..,.....7.1.......".."In=razi...,;Li....,,1 -
,,,.._........,m..m. -....p 4=2= a......m.".1,:niauss==
,.._._........,..... ......................
..-".'122allti
= ..... :: ..-::....p. LI - -- = Err:
Ma; 1....... , e-,........7:::1;1:1.0..zwira
..............16.

Y.. . ...
01 Q2 03 RE7FE!hi----=le.z=91--aniezsiirogirAitittehz..7416291iiiiihrgienir-
04 05 06 07 08 09
ig 811r.
10
walla MM., IMINVINO
Oa.
11 12 13 14 15 16
J
El 4-70
1973

Figure 16.3: Open Water Diagrams of B-4.70 Propeller Series

(generally a small effect) and by decreasing the pressure around the stern.
The thrust to be developed by the propeller should thus be greater than
the resistance without propeller at the design speed, because the thrust has
to be equal to the increased resistance. The increase of the resistance due
to the propeller action is expressed as the thrust deduction factor t:
January 5, 1994, Systematic Propeller Series 231

t=
TR

T is the thrust to maintain a certain design speed and R is the resis-


tance without proReller at that speed, as found e.g. from the resistance test.

The assumption that the nominal wake fraction can be used to deter-
mine the propeller inflow is not consistent with the increase of the resistance
due to the propeller, as expressed by the thrust deduction factor. To be
more accurate the effective wake fraction should be used, but this refine-
ment is left for later.

In the preliminary design stage some estimates have to be made of the


thrust deduction factor and of the nominal wake fraction. These estimates
can be made using statistical data or even comparable ships.

16.5 Propeller Design Requirements.


The charts of the B-series can be used for a preliminary design of a pro-
peller. To be able to design a propeller the requirements should be defined.

The most open situation is when a ship is designed for a given design
speed. Using the wake fraction the propeller entrance velocity is known.
Also the resistance of the hull without propeller should be determined, e.g.
from statistical calculations or from a resistance test. Using the estimated
thrust deduction factor t the required thrust can be found from

T= (1 t)

The thrust deduction can be estimated using statistical formula's. A


common procedure is to measure it at model scale using a stock propeller
with an approximate diameter and with the required loading at the design
speed.

The propeller designer now has the freedom to choose the number of
blades, the blade area ratio, the rotation rate, the diameter etc. The cri-
terium is optimum efficiency and the result is a propeller and the required
shaft power for the design speed.

The situation is not always as open as in the foregoing case. A lim-


ited number of engines is available and especially the rotation rate of these
232 G.Kuiper, Resistance and Propulsion, January 5, 1994

engines is prescribed. The engine has a maximum power at which it can


operate continuously ( the Maximum Continuous Rate or MCR) and the
design condition of the engine is often chosen at 80 percent MCR. At that
operating condition the engine delivers its power at a certain rate of rota-
tion. That means that the _propeller designer has to design the propeller to
absorb that power at that required rotation rate.

Another restriction which occurs frequently is a restriction on the al-


lowable diameter. To maintain sufficient tip clearance while staying at or
above the baseline of the hull with the propeller often requires a limit to
the allowable diameter. In practice this maximum diameter is often smaller
than the optimum one, so the designer has to design a propeller with re-
stricted diameter.

Another possibility is that the ship exists and the engine has already
been chosen or even installed. Maximum propeller efficiency in this case
means maximum ship- speed. In such a case the available power and the
rotation rate is prescribed.

Several combinations of such restrictions and prescriptions are possible.


The open water diagrams can be used to meet these requirements, but first
some other considerations are necessary.

16.6 Choice of Number of Blades and Blade


Area Ratio.
Before the open water diagrams of the systematic series can be used to
fulfill the design requirements the parameters which are chosen on other
considerations than efficiency have to be defined. These parameters are the
number of blades Z and the expanded blade area ratio AE/Ao

The number of blades is chosen in relation to possible vibrations. An


8 cylinder engine and a four bladed propeller may suffer from resonance
frequencies because the blade frequency and the engine frequencies have
common harmonics . In that case the vibrations will become excessive,
resulting in damage.
The structure of the wake is also important for the choice of the number of
blades. When the wake has strong second and fourth harmonics, which oc-
4Harmonics are distinct frequencies in a periodical signal. The frequency content of
a periodical signal can be found by a Fourrier transformation, which describes the signal
as a sum of sine wave with distinct frequencies
January 5, 1994, Systematic Propeller Series 233

curs when there is a wake peak both in top position and in bottom position
of the blades, an even bladed propeller is at a disadvantage with respect to
shaft vibrations.
For Navy ships the number of blades is often chosen as high as possible to
reduce the danger of tip vortex cavitation.

The blade area ratio is chosen such that cavitation is avoided as much as
possible. Empirical formulas have been developed to choose the area-ratio.
An old and very simple formula is that of Taylor

AE/Ao = 1.067 0.229P/D

An empirical chart which is still frequently used is the chart of Burrill


(1943),as given in Fig. 16.4. The line based on the experience of MARIN is
also given in that chart. (Note that in this chart e = p, and AEIA = SI Fp)
From calculations with the lifting line theory using a 25 percent margin

/
0,4 1

0.3 / we\5411%2.-

..r/
/ -----;

,.....------
0.20
.. e.`\

I
0.15

of.

0.10

0.09

0.08
E
=10111111111
/
e<",
11111111E111
1111 Upper limit heavily - loaded propellers (Burrill)
IPMEMUpper limit merchant prop all type sections (Burril0
Upper limit merchant prop. aerofoil sections (Burril0
407 Mil -- -- Upper limit Wageningen tank

e
0.06
--- Upper limit Lerbs 4a= 7173 (1(i.3/2 Ps/q) - )
...,,'
-.
0.05
0.05 0.07 010 0.15 02 fl2 0.3 04 05 06 07 08 09 1.0 1.5 20 2.5 3.0

1/2 V20,622+ n. 0.70 )7

Figure 16.4: Empirical Chart for the Blade Area Ratio (Burrill 1943).

against pressure side bubble cavitation Fig. 16.5 was derived [28]. The
calculation method will be described in chapter 19. (Note that in Fig. 16.5
FalF = = pg and HID = PID. A handsome indication gives the
234 G.Kuiper, Resistance and Propulsion, January 5, 1994

r. 0,30

420 11111 .
z 5- /F .0,35
.Fa

0,20
_

0,15 EVII . z .4 - FaiF


1 .0.35 11 0,15
. z 5 - Fa f F. 0.50
- z .3-Fa/F. 0,35
ri4 fi ll.
o,io
ptil. "Pzf;.%1;ifl,as 1

was , Fa .. 0,10

MAKIN
.
Upper limit heavily-loaded propellers (Burri 0
0,08 zC. Fa F.0.65 --Lower limit merchant prop. aerofoil sections (Elurrill
0,08
H/13 .0,5

1111.
1

2 WO *V
3 HA3 .0.3 Data according to the
0,06
4 WO 1.1 circulation theory 0,06
z.2- Fa/F.0,65 5 14/13 1,3
3,05
1 1 1 1 1
i
0,05
U01 0,15 2 0,25 03 04 0,5 0.6 0,7 0,8 0,9 1.0
1

1.5 2,0 2,5 30

Po-e-0,8R7
u.0 D 6:11 + 25 Vo

Figure 16.5: Chart for the Required Blade Area Ratio from Lifting Line
Calculations.

formula of Keller
(1.3 + 0.3Z)T
AEI Ao = +k
(Po P)D2
in which P0 is the undisturbed pressure at the propeller shaft. This for-
mula has a more physical basis because it takes the blade loading and the
pressure into account. The propeller revolutions are not taken into account,
although these may also be important. The constant k is zero for fast naval
vessels and 0.20 for highly powered full ships.

The blade area ratio and the minimum pressure on the propeller can-
not be described accurately by a single curve or set of lines. Nowadays a
calculation, such as with the lifting line theory, has to be made to be more
accurate. These rather old charts and formula's are therefore only useful
for preliminary designs.

16.7 Propeller Design using B-Series Charts


After all these considerations the charts of the B-series can now be used to
design a propeller with maximum efficiency within the design restrictions.

The basic procedure is to calculate either the J-value, the KT value or


January 5, 1994, Systematic Propeller Series 235

the KQ-value and to read the efficiency from the series diagrams. However,
generally the data to calculate either parameter are not available. In that
case an estimate of the unknown data is made and this estimate is varied.
An example makes can explain this.

Assume that the. design speed is known and the rate of rotation is pre-
scribed. From the design speed the entrance velocity is derived using the
nominal wake fraction. Two variables of the advance coefficient are then
known: Ve and n. Only the diameter is unknown. In that case a diameter is
chosen arbitrarily. This makes it possible to calculate J. In the appropriate
diagram ( The number of blades and the blade area ratio have been chosen
as mentioned above) the corresponding efficiency from the open water dia-
grams such as in Fig. 16.3 can be read for all available pitch ratio's. The
pitch ratio with the maximum efficiency is chosen. When another diameter
is chosen another maximum efficiency is thus found. The diameter can be
varied until a maximum in the efficiency as a function of the diameter is
obtained.

This iteration technique can also be applied when more than one variable
is unknown. In that case a matrix of variations has to be calculated. The
most common case is when the rotation rate n and the diameter D are both
unknown. The correct iteration is to chose an initial diameter and an initial
rotation rate. The entrance velocity is known, so the advance ratio J can
be calculated and the maximum efficiency can be read from the diagrams.
The rotation rate is varied first to find an optimum rotation rate. Then a
second diameter is chosen and the whole procedure is repeated. Note that
the optimization of the efficiency with a fixed rotation rate n generally gives
another optimum than the optimization with a fixed diameter D. Both are
sub-optima.

This procedure is elaborate and well suited for computer programs. The
diagrams such as in Fig. 16.3 have been approximated by polynomials for
that purpose. There is a manual shortcut, however, which makes it possible
to use the diagrams by hand or to speed up computer calculations.

16.8 Elimination of Variables.


In all cases the entrance velocity of the propeller is assumed to be known.
When this is not the case it has to be estimated and optimized later by
variations as mentioned above. When one other parameter is unknown it is
possible to eliminate this parameter from the diagrams instead of optimizing
it by variations.
236 G.Kuiper, Resistance and Propulsion, January 5, 1994

16.8.1 Known Power and Diameter.


A common situation is e.g. that the available power PD and the rotation
rate at which this power is developed is known. The unknown variable is
the propeller diameter. In that case the diameter can be eliminated from
the diagrams by plotting KQIJ5 versus J instead of KQ versus J. This
value can be calculated because

KQ Q n5D5 Qn3 Ppn2


J5 pn2D5 Ve5 pV5 2irp1/5
The parameter KQIJ5 has been called Bp 5. Instead of plotting it against
J it was plotted against 1/J, which parameter was called S. The plots were
therefore called Bp- 6 diagrams. An example is shown in Fig. 16.6. The

1.4

i1 wirmirmay lltfitrimrretrovimmoriontrawoomatFa
IntrAttitm da41.4t4AinAlkirmammar4Atasrioraer Amino

\
t3

1.2 IPAveltrevOINVAWDAWASFACROART10.2RPERNEMPIE41
A NA sA411.4A.t.411.4s.terliAtte 411ss.karain DINEMIEGEON
ZW\AVARRATALVMDATI" VII AMEIMEN ter licso"40
1.1 int% ommlAiistekttArMitilekikArAtatimnittlE&Nab,..EMEAM!
NV V NV irr Pp,' rIfiErf WAN V % Pi VO V W. OWENS 022 VA
*M. AitAthiAtiAaitoraMtAkeikelm .W.4.4CORT0.11214 4-

"WM
1.0

P/ D
arSVOTWOrrekvit
WY A W.4 A Al_.....O.C.A. 4 .1...4
6 LA. 1.4_ VIIRFIMFAWIftr
VAASM..41.1Weo..AA
smoN
, ,......Neartirrovregartyryommresoordrow.NA.ier-vronvi
, .04_, 1,4 4, i_ 1 41, 11. .1, 4 .... .. _A A. Alb. .A_A41;401..A , 4 Atio
os milikk.44:gr_ A AblatattiketWiSWIRASIMAr. ',Ii*eallekriC:41M ,d8o ...

MrAIDAVAIISETOWEAPpARRIWENEWOR"Or4WfV1;021.: :,`,"
17 /0256:0044:44:rzer:Ordaiareidirardiardai 4"42.1MMet4lafi'Ar7...e
Mrargeitat ,,z-111-.YeAff%00y,"470-ppip,46.itlFeatimsrgr";ppwvr -
0.6 TAIRTENett.,:_Z-":"---/.4;,-.44A.VA 4.200,AMP.A.g.10M-4-%? - .
WAIMPIF ArAP4.--7---1:71r-i-22:07vivorros P-r - w.p.k- A7- ....-
__AdreArdrozolzeolorzi----i--;----------7-------57-4--:-4;----4-2.-514::-._Aio.*--r0.!--fgioo.,<,
0.3 .04 05 06 07 0.8 0.9 1.0
_re-_-di-___ _.;!_lig
1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0 2.1 2.2
B
P

Figure 16.6: Example of a Bu-8 diagram.

efficiency 7/ is indicated by lines of equal efficiency. Since in the case under


consideration the Bp value is known, the optimum efficiency and the corre-
sponding pitch ratio can be read directly. The line connecting all maximum
'In older diagrams the Bp value was not defined in metric units!
January 5, 1994, Systematic Propeller Series 237

efficiencies is also shown in Fig. 16.6. It is the line connecting the locations
where the efficiency curve is vertical.

For each propeller series (number of blades and blade area ratio) such a
diagram can be made. These diagrams have been used widely for manual
calculations of the-optimum diameter and pitch in the mentioned case of
known power and rpm.

16.8.2 Known Power and Rotation Rate.


When the power PD and the rotation rate n is known, a similar diagram
can be made using the combination KQ/J3.

IfQ PD
J3 27rpD2V3

In this parameter the rotation rate is eliminated. To separate it from


Bp this parameter has been called Bp2. The diagram can be plotted in a
similar shape as in Fig. 16.6 as a B2-(5 diagram.

The optimum line, as in Fig. 16.6 can be used without the whole dia-
gram. For a quick estimate of the optimum diameter when power and rpm
are known Fig. 16.7 can be used. Similarly when the power and diameter
are known Fig. 16.8 can be used. These figures illustrate that in both cases
different optima are obtained.

16.8.3 Known Thrust and Diameter.


When the ship speed is not fixed a resistance curve is often available. That
means that at every ship speed the resistance is known. Using a thrust
deduction factor the required thrust at each speed is known. In that case
a ship speed is chosen and the thrust is known. When also the diameter is
known the rotation rate can again be eliminated using

KT T
J2 pV2D2

Instead of making another diagram for this case, the use of the open
water diagrams will be illustrated.
238 G.Kuiper, Resistance and Propulsion, January 5, 1994

0.7

0.6

no 1.4 5.0

0.5

1.2 4.0

0.4

3.0

0.3

0.6 2.0

P/D

0.6- 1.0

0.4 t
0.5 1.0 1.5 2.0 2.5
1/4
J.
5/4
IcQ

Figure 16.7: Optimum Lines for given Power and Diameter.

16.8.4 Known Thrust and Rotation Rate.


To be complete the last case, when the diameter is known with given thrust,
is also mentioned. The apparent combination to eliminate the diameter is

KT Trd
J4 pV4

In this case also the open water diagrams can be used, as will be illus-
trated below.
January 5, 1994, Systematic Propeller Series 239

no

5.0

4.0

3.0

2.0

LO

o
0.25 0.50 0.75 1.00 1.25
K1/4.3-3/4

Figure 16.8: Optimum Lines for given Power and Rotation Rate.

16.9 Optimization using the Open Water


Diagrams.
As has been shown it is possible to make diagrams for every possible combi-
nation of known and unknown variables. In practice it is more convenient to
use one diagram for all cases. The most suitable diagram is the open water
diagram. Elimination of variables can still useful, as will be illustrated in
the case of known thrust and rotation rate.

Assume that from the thrust, the entrance velocity and the rotation rate
240 G.Kuiper, Resistance and Propulsion, January 5, 1994

the value KT I J2 can be calculated to be

KT
J4 = 1.1387
This relation can be plotted in the relevant diagram, e.g. in the open
water diagrams of the B4.85 propellers (A four bladed propeller with 0.85
blade area ratio), as shown in Fig. 16.9.

Figure 16.9: Determination of Optimum Diameter from B4-85 Series

At each crossing of this curve with a KT curve of the open water di-
agrams (the open dots) the corresponding efficiency can be read (the full
dots). The optimum efficiency is found to be 0.61 at J = 0.62 and a pitch
ratio of 0.9. Since the entrance velocity and the rotation rate are known
January 5, 1994, Systematic Propeller Series 241

the optimum diameter can be calculated from the optimum value of J.

Note that the optimum condition is not the maximum efficiency of pro-
peller B4-85(P/D=0.9). This is because of the required combination of
thrust, rotation rate and entrance velocity. A different choice of these pa-
rameters will give-a different optimum, as will be shown in the example
below.
In Fig. 16.9 the torque coefficients have been omitted. These are also
known, however, so the required power in the optimum condition can also
be calculated.

16.10 Example.
The task is to design a four-bladed propeller for a Ro-Ro-ship. The ship involved is the
same ship as used in chapter 7 and the propulsion data will be derived later in chapter 20.
The blade area ratio is chosen first. Using Keller's approximation the diameter has
to be known. A first estimate of the propeller diameter is 0.711, where To is the draft at
the aft perpendiculai. This results in a diameter D = 0.7 x 9.63 = 6.74 (see Table 7.2
on page 118).

The resistance of the ship at 22 knots was found from the resistance test to be
1196 kN (Table 7.5). Assume that the thrust deduction t has been determined from a
propulsion test to be 0.211. (A propulsion test will be discussed in chapter 20). The
required thrust for a speed of 22 knots is therefore:
1196
T= = 1516kN
(1 t) (1 0.211)
The required area ratio according to Keller is:
(1.3 + 0.3 x 4)1516000
A.e/A0 = = 0.85
(105 2300)45.43
Here the vapor pressure Pt, is taken at 2300 Pa. The value of k in Kellers formula is
chosen as zero, because of the very slender afterbody. The design diagram for the B4-85
series will thus be used, as shown in Fig. 16.10.

Assume that the wake fraction can be measured with a pitot tube and that it has
been found to be 0.251. So at 22 knots the incoming velocity of the propeller Ve is 16.48
knots or 8.49 m/ sec.

The thrust, diameter and entrance velocity are now known and the rotation rate has
to be optimized. The combination of parameters in which the rotation rate is eliminated
is KT/J2:
1516000
Kt = J2 = 0.486J2
1025 x 8.492 x 6.52
This curve has been plotted in Fig. 16.10. The optimum efficiency can be derived
in the same way as in Fig. 16.9. The optimum efficiency is 9 = 0.6 at an advance ratio
of 0.65 and a pitch ratio of 1.0. As can be seen in Fig. 16.10 the optimum is very flat
January 5, 1994, Systematic Propeller Series 243

The advance ratio in the optimum condition is 0.65. The optimum number of revo-
lutions is found from:
8.49
ns = 2.01rps
0.65 x 6.5

Note that again the optimum condition is not at the maximum efficiency of Propeller
B4-85(P/D=1.0).

The optimization as carried out above can be continued by keeping the optimum
propeller revolutions at 2 rps. The optimum diameter for this rps can be found by
eliminating D from Kt and J. The result is

Tn21516000 x 22 x J4 = 1.1387J4
K t = pv.e4 4 =
1025 x 8.494
This curve has been used in the example of Fig. 16.9. This resulted in an optimum
efficiency of 0.61 at J = 0.62 and P/D=0.9. So the optimum diameter for 2 rps is
8.49
Dopt = = 6.85m
2 x 0.62
So by increasing the diameter from 6.5 to 6.85 meters the open water efficiency is
increased from 0.6 to 0.61. The optimum is therefore reasonably flat and it is probably
more important to choose the propeller diameter in proper relation with the wake to
optimize the hull efficiency, as will be discussed in chapter 20. The larger diameter is not
always feasible while maintaining the required clearances between propeller and hull, as
prescribed by the insurance companies.

The blade area ratio and number of blades can be varied to further optimize the
efficiency. This is a time consuming work and the open-water curves have therefore been
represented in polynomials, so that computerized optimization can be carried out.

16.11 Four Quadrant Measurements.


The open water curves of propellers are restricted to the forward advance
operation, that is in the condition of forward speed of the ship and forward
thrust of the propeller. E.g. for stopping conditions and during maneuvers
it is important to know the characteristics of a propeller in other conditions.
For that purpose the "four quadrants" of a series of B-series propellers has
been measured. An example is given in Fig. 16.11.

The four quadrant results are given as a function of hydrodynamic pitch


angle arctan --r--0.714rnD (see Fig. 16.12. The value of varies over 360
degrees. From 0 until approx. 40 degrees the curves are the regular open
water diagrams. The range 0=90 until 180 degrees corresponds with neg-
ative revolutions and forward speed, as would occur when the propeller is
244 G.Kuiper, Resistance and Propulsion, January 5, 1994
16

c'T

-
12
Iiiiri,\\ 8 4-70
15'6.05
as
08

08
SS 10
12
//if \ -
/ ,--
, \\
-...,
,......,
/...., , 4
. rir wv 16,.
o.
i
\.,...,.. ,\,,
\) ,7 ,
o
,\-\
s \
.........._
,v,,,
/
.
4^,
.4/
,/ \ \x
,/.
____,,, I1,,\

08
11111 ,N

\ \./
\ i r.
_......_,

,.. ,..,,"/
/ /
i

1
\ /a \ i

--- .

\
1

% '

-20

-24 e
80 120 160W 200 240 280 320 360

Figure 16.11: Four Quadrant Measurement Results of B4-70 Propellers

reversed during a stopping manoever. The range fi = 180 270 degrees cor-
responds with negative revolutions and negative ship speed. In the velocity
diagram Fig. 16.12 this means that the resulting velocity V,. is coming from
the direction of the trailing edge. In the fourth quadrant fi = 270 360 de-
grees the velocity is negative and the revolutions positive. So the propeller
gives forward thrust while the ship speed is still negative. Such a condition
might occur when manoevring in a harbor.

The thrust in the four quadrant diagrams has been expressed as the
thust coefficient based on the resulting velocity Vr:
January 5, 1994, Systematic Propeller Series 245

2 th QUADRANT 1 th QUADRANT

X Trn D

3 th QUADRANT 4 th QUADRANT

Figure -16.12: Four Quadrants of Propeller Operation

Ct =
1/2p(K2 (0.77rnD)2)iD2

Similarly the torque coefficient has been expressed as

c, =
1/2p(Va2 (0.77rnD)2)D3

From the four quadrant diagrams some conclusions can be drawn about
the propeller characteristics in manoevring conditions. The maximum neg-
ative thrust is obtained at ,8 = 80 degrees. So the revolutions are still
positive in that condition. When the revolutions are reversed when the ve-
locity is still considerable is in the range of 160 degrees and the negative
thrust may even disappear. This has to do with stall of the blade sections,
as will be discussed in chapter 17.The same occurs at around 340 degrees.

The four quadrant diagrams also give information about the rpm during
a stopping manoever, which is generally controlled by the maximum torque
available. The maximum thrust during a stopping manoever can thus be
determined. From this the propeller strength can be checked. Determina-
tion of the length required to stop the ship can be found from an integration
of a range of transient conditions in a quasi static way.
246 G.Kuiper, Resistance and Propulsion, January 5, 1994

16.12 Propeller Design using the Optimized


Data.
When the geometrical parameters such as number of blades, blade area ra-
tio, and pitch/diameter ratio have been determined the propeller geometry
is known, because the geometry of that propeller from the series hags to be
taken. In practice the geometry of the series is often adapted for reasons
of e.g. cavitation. Common adaptations are the choice of a different blade
contour and the choice of different blade sections. When other blade sec-
tions are used the maximum camber of the alternative sections has to be
equal to that of the B-series, otherwise the derived optimum pitch has to be
corrected for the difference. This can be explained after the characteristics
of blade sections have been discussed.

16.13 Other Series.


The B-series has been -developed by Marin. The same institute has also
developed systematic series of other propellers. A notable example is the
series of propellers in an accelerating nozzle, the Ka series, as developed
by Oosterveld [37]. A recent series is a series of controllable pitch propellers
in a nozzle. All information on systematic series available at Marin has
been summarized in [22]. The data have been summarized in polynomi-
als to make them accessible for computer programs. The latter booklet
contains a floppy disk with all data. Examples of how to use the series
information is also given in this booklet.

As mentioned before the basis of a series is some parent type of propeller.


Each series has its parent propeller. A series with a high blade area ratio,
suitable for high speed ships, is the Gawn series [8]. Other institutions have
made their own series, such as the Swedish Towing Tank SSPA [27] and
others [29].
Chapter 17
Profile Characteristics
Purpose: Introduction to the lift and drag properties of profiles, the behav-
ior of the pressure distribution and loading distribution with varying angle
of attack and the geometry of profile series

A characteristic feature of a profile is that it generates lift. This is


caused by the fact that the flow is forced to leave the profile smoothly at
the tail. This in turn is an effect of viscosity, which prevents the velocities
from becoming extremely large at the tail. The condition of smooth depar-
ture of the flow at the profile tail is called the Kutta- condition. It can be
formulated as tangential flow at the tail, or as equal pressure at both sides
of the tail.
When it is assumed that the flow remains attached to the profile and that
the boundary layer is thin relative to the profile thickness, the flow field
around the profile can be considered as a potential flow. Outside the bound-
ary layer the viscosity can be neglected and no rotation is present in the
undisturbed inflow, so the two conditions for potential flow are met. The
Kutta condition maintains the viscous flow pattern around the profile in po-
tential theory and this condition has to be applied explicitly in a potential
flow.

17.1 The Pressure Distribution.


Consider a profile moving at a certain angle of attack through the fluid,
as defined in chapter 15. Taking the coordinate system in the nose of
the profile, there is an inflow velocity V at the angle of attack a. As a
result of the inflow there is a certain pressure distribution over the surface
of the profile. When the pressure Po in the undisturbed flow is changed,

247
248 G.Kuiper, Resistance and Propulsion, January 5, 1994

the pressures over the surface of the profile will also change with the same
amount. So the pressure distribution is made independent of the pressure
Po by subtracting Po from the local pressure at the profile. This pressure
distribution p po is made non-dimensional with the total head 0.5pV2 of
the undisturbed inflow. The pressure distribution is thus expressed in the
pressure coe cient Cp as -

Cp
_P0.5pv2
Po

In the flow the Bernoulli Law is therefore valid, so the pressure distribution
at the surface of the profile can be related to the local velocity y by

p + 0.5pv2 = Po + 0.5pV2
The influence of the immersion h is neglected here, because the profile
is assumed to be nearly horizontal and at a small angle of attack. In non-
dimensional terms the Bernoulli equation can simply be written as
\2
Cp = 1 (17.1)
V)

There are two locations where the local velocity is zero. These locations
are the forward and rear stagnation points . In these stagnation points the
local pressure is equal to Po + 0.5pV2 and the pressure coefficient is +1. 1
An example of a pressure distribution of a profile at a small angle of attack
is shown in Fig. 17.1. The pressure distribution is plotted as a function of
the position on the projected chord. The projection is necessary to have the
inflow along the x-axis. The chord position is made non-dimensional with
the chord length c, so it goes from 0 to 1. Note that the pressure coefficient
is plotted as the negative pressure coefficient. At the suction side (the up-
per side) the pressure coefficient is negative over most of the chord, which
means that the local velocity is greater than the undisturbed velocity V. At
the lower side the pressure coefficient is closer to zero. At the sharp tail the
stagnation point will be at x cI2, at the rounded nose the stagnation
point may be at some distance from x = c12.

The relation between the chord length c and its projection cprol on the
direction of the velocity can be written as
y2
Cproj = c cos cc = c(1+ 0()
'The sign of the pressure coefficient is also frequently reversed, resulting in a stagna-
tion pressure of -1.
January 5, 1994, Profile Characteristics 249

UPPER SIDE

Figure 17.1: Example of the Pressure Distribution on a Profile

where y is the difference in height between nose and tail. When y I c is


considered small (< 1) the second order term can be neglected and the
projected chord is equal to the actual chord length. In this approximation
only terms linear in the small value y I c are taken into account and the higher
order terms are neglected. This is called linearization of the problem. It
is extensively used in the formulation of the potential flow around profiles,
which will not be treated here. The linearized thin profile theory is used to
calculate the velocities and pressures around thin profiles at an arbitrary
angle of attack. Some results are important and will be given below.

17.2 The Loading Distribution.


The pressure distribution as given in Fig. 17.1 is important for cavitation
and for the development of the boundary layer (chapter 18. The pressure
difference between suction side and pressure side is the loading of the profile.
So from the pressure distribution follows a loading distribution . The integral
of the loading distribution over the chord is the lift force L on the profile.
In non-dimensional terms the integration of the pressure coefficient gives
the lift coefficient
250 G.Kuiper, Resistance and Propulsion, January 5, 1994

CL
0.5pV2S

where S is an area (span times chord length). For a profile the lift coefficient
per unit span is also used, where S is replaced by the chord length c. Note
that the loading distribution is a force distribution which is perpendicular
to the x-axis. This is important because in potential flow there is no drag,
so the lift force is perpendicular to the flow. The difference between the lift
force perpendicular to the chord and perpendicular to the undisturbed flow
is negligible in linearized theory, because of the small angle of attack, but
the direction of the lift force is important for the calculation of thrust and
torque in propeller design. (chapter 19).

An important result of the linearized thin profile theory is that the load-
ing distribution is directly proportional to the camber distribution. This
property is the reason of the definition of the camber. Note that in the
derivation of this relation the problem has been linearized, so that it is only
valid for small angles of attack and thin profiles in a potential flow.

When the pressure at the suction side is called p- and the pressure at
the pressure side is called p+, the relation between the pressure and the
camber f and thickness t is as follows:

0.5(p+ p-)(x) cx t(x)


0.5(p+ -I- p-)(x) cx f(x)
or in words, the thickness distribution is proportional to the mean pressure
, the camber distribution is proportional to the pressure difference. This

separation is important because only the camber distribution is responsible


for the forces on the foil. In the calculation of the performance of propellers
(thrust and torque) the blade thickness can therefore be neglected.

17.2.1 The Lift Curve.


When the lift of a profile is measured and plotted in a non-dimensional way
as the lift coefficient C1, the result is a curve with the characteristics as
shown in Fig. 17.2
In the linearized thin profile theory the lift curve is exactly linear. The
lift curve crosses the x-axis at the zero-lzft angle . This angle is determined
by the camber of the profile. An increase in camber is increases the zero lift
angle. The slope of the lift curve is 2r for thin profiles. (Note that a is
January 5, 1994, Profile Characteristics 251

cl
Thrust breakdown
Flow separation

a0

Figure 17.2: Lift curve of a profile

in radians.) The thickness of the profiles and viscous effects cause a small
deviation of the slope from 27r, but in practice this deviation is never more
than 10% . The lift coefficient of a thin profile can therefor be written as

= 27r(a ao) (17.2)

At higher angles of attack the lift curve deviates from the straight line
and at a certain angle the lift curve drops sharply. This is the angle where
the flow separates from the profile and it is called the stall angle. For air-
foils this is a very important angle, but it will not often be reached with
propeller sections, because strong cavitation will occur much earlier.

17.3 The Zero Lift Angle.


The zero lift angle depends on the camber distribution. For a symmetrical
parabolic camber distribution as shown in Fig. 17.3 the zero lift angle can
be calculated using linearized lifting surface theory to be
252 G.Kuiper, Resistance and Propulsion, January 5, 1994

2fmax
o

Figure 17.3: Determination of the Zero Lift Angle of a Symmetrical Cam-


berline

This is a practical approximation of the zero lift angle of arbitrary sec-


tions.

17.4 The Leading Edge Suction Peak.


Consider a symmetrical profile (without camber). At zero angle of attack
the minimum pressure is located in the midchord region. An example of
such a pressure distribution at the suction side is given in Fig. 17.4 for
0 degrees angle of attack. The minimum pressure depends mainly on the
maximum thickness of the profile. At a small angle of attack (1) the pressure
will decrease in the leading edge region, while the minimum pressure is
only marginally lower than at zero angle of attack. At a certain angle of
attack (2) the pressure is approximately constant in the leading edge region.
At higher angles of attack (4) a leading edge suction peak occurs and the
minimum pressure decreases strongly with increasing angles of attack (5).
This is important for cavitation, as will be discussed in chapter 18

17.4.1 The Ideal Angle of Attack.


A special angle of attack is the angle at which the pressure around the
leading edge is symmetrical. In this case the flow direction at the leading
January 5, 1994, Profile Characteristics 253

-Cp

+1

Figure 17.4: Pressure Distributions at Various Angles of Attack

edge is exactly in the direction of the camberline. When the profile is thin
(only the camberline) this is the angle at which the flow is shockfree at the
leading edge. At other angles of attack the velocity becomes infinite at the
leading edge. For a symmetrical profile this angle is of course zero degrees.
For a cambered profile this angle is not trivial, because the streamlines
close to the profile will be deflected by the lift of the profile, as illustrated
in Fig. 17.5. The ideal angle of attack depends on the camber distribution.
A special case is when the camber distribution is symmetrical with respect
to the midchord position. In that case the ideal angle of attack is again
zero degrees, as illustrated in Fig. 17.6.

The ideal angle of attack is important because it is the angle of attack


where the pressure close to the leading edge is maximum. It is also the
angle which is the middle between the angles where a leading edge suction
peak occurs at the suction and the pressure side respectively.
Propeller blade sections are generally designed to operate at the ideal angle
254 G.Kuiper, Resistance and Propulsion, January 5, 1994

CAMBER LINE

Figure 17.5: Streamlines at Ideal Angle of Attack

CAMBER LINE

Figure 17.6: Streamlines of Symmetrical Camberline at Ideal Angle of At-


tack.

of attack.

17.4.2 Profile Drag.


The sectional drag is represented as the drag coefficient Cd. The drag D is
added to the lift as shown in Fig. 17.7 and is in the flow direction. The drag
force is due to the friction along the surface of the profile and has its main
direction in the chord direction (D'). Decomposition in the flow direction
introduces a small decrease of the lift force due to profile drag. In linearized
theory this decrease is neglected, as is the difference in direction between
D and D' and only the component D is used.
The drag force is expressed in non-dimensional terms as

Cd
pV2S
where S is the area of the profile, just as used for the lift coefficient. A
typical drag curve is plotted in Fig. 17.8. The drag coefficient is plotted
January 5, 1994, Profile Characteristics 255

Figure 17.7: Direction of Lift and Drag Forces

versus the lift coefficient. The drag coefficient depends on the Reynolds
number and the lower drag curves in Fig. 17.8 are drag curves for Reynolds
numbers from 3 to 9 million. 2 In this range of Reynolds numbers the
boundary layer is still laminar over part of the chord. This reduces the
drag coefficient considerably. The extent of laminar flow is especially large
at the ideal angle of attack (which is zero degrees in this case) and this
is reflected in the bucket of the drag coefficient at zero lift coefficient. An
increase in Reynolds number decreases the extent of laminar flow, which
results in an increase of the drag coefficient. When the laminar flow is
eliminated by roughness at the leading edge the drag coefficient increases
significantly and the bucket at zero angle of attack has disappeared.

The lift drag ratio can be considered as the performance of the profile.
A profile is a force multiplier and is very effective. E.g. at a lift coefficient
of 0.2 (a common lift coefficient for propeller blade sections) the drag coef-
ficient in Fig. 17.8 is only 0.01 for the roughened profile. So the drag force
is multiplied by a factor of 20! The lift drag ratio of profiles with laminar
boundary layers can be as high as 40.

The lift coefficient of propeller blade sections is restricted to approx. 0.3


in practice, because of cavitation. A minimum drag coefficient of 0.008 for
high Reynolds numbers (full scale) or 0.004 for Reynolds numbers as used

2Reynolds number based on chord length and undisturbed inflow velocity.


256 G.Kuiper, Resistance and Propulsion, January 5, 1994

.024

ig 020

.4b

.0/6

t .0/2
*.
V) .008

.004 o 3.
o
o
6.0
0
Figure 17.8 Typical drag curve of a profile (NACA 66-006).

at model scale can be used as a first approximation. Hoerner [13] gives an


approximation of the minimum drag coefficient as CD= Cf(i+ 2!), where
Cf is the friction coefficient of a flat plate. For Cf twice the extrapolator
line can e.g. be used, or the Blasius friction line when laminar flow is dom-
inant.

In case of profiles with a finite span the frictional drag is only a fraction
of the total drag, which is dominated by the induced drag. The induced
drag is due to a change in direction of the lift, which causes that the lift has
a component in the direction of the undisturbed flow, as will be discussed
in chapter 19. An accurate determination of the viscous drag is therefore
often unnecessary.

17.5 Profile Series.


As mentioned the NACA has defined and tested a large number of profiles.
The data have been summarized by Abbott and Doenhof in [1]. Part of
January 5, 1994, Profile Characteristics 257

these profiles are defined as a thickness and camber distribution, which of


course have to be combined in the NACA-way. In practice the other way of
combining them (perpendicular to the chord) is sufficiently accurate, except
for the calculation of the pressure distribution at the nose.)

17.5.1 Thickness Distributions.


NACA has used several series of thickness distibutions. A thickness distri-
bution which is often used for propellers is NACA-1 series. The 1-series has
a thickness distribution which is given in [1] in tabular form for different
values of maximum thickness. It is often indicated by two digits, the first
being the series (1), the second being the position of the minimum pres-
sure at zero lift in tenth of the chord. A commonly used distribution is
the 16-series, which has the minimum pressure at 60 percent of the chord,
measured from the leading edge.
Another series of thickness distributions is the 66-series, which differ from
the 16 series mainly by a thinner tail region.
The profile 16 and 66 thickness distributions are further indicated by their
maximum thickness yy (in percent of the chord) as e.g. NACA66-0yy. The
thickness distribution of the NACA66-006 profile (with 6 percent maximum
thickness) is given in Fig. 17.9. The Table also gives the velocity distrib-
ution y over the chord in terms of y/V. The squares of this velocity ratio
are related to the pressure coefficient 1 Cp, as follows from eq. 17.1. The
ratio ya /V in the table is the increase of the local velocity due to a unit
increase of the lift coefficient. This can be used to approximate the velocity
(and pressure) distribution at angles of attack.
In Figure 17.9 a graph of the pressure distribution (in terms of (y/V)2) at
zero angle of attack is shown, together with the pressure distributions at
a lift coefficient where the leading edge suction peak begins. As shown in
Fig. 17.9 the leading edge suction peak begins at a lift coefficient of 0.01.
From eq. 17.2 the angle of attack at this lift coefficient is found to be 0.0016
radians or 0.09 degrees. So this (thin) profile has a very small margin against
variations in angle of attack. This angle increases with increasing thickness.

Small variations of the maximum thickness can be obtained by multi-


plication of the tables such as in Fig. 17.9.

17.5.2 Camber Distributions.


An important series of camber distributions is the series with constant lift
from the leading edge on until a percentage of the chord. Behind that loca-
tion the lift goes to zero at the trailing edge in a linear way. The camberlines
are indicated as "NACA a=0.x", where x is ten times the fraction of the
258 G.Kuiper, Resistance and Propulsion, January 5, 1994

16

0/ Upper surface
12
o
Cl L ower surface
(H2
.8

N4C4 68-006

.4

0 .2 .4 6 8 1.0
xIC

X Y
(v/V)2 v/ V tiva/V
(per cent c) (per cent c)
- 0 0 0 0 4.941
- 0.5 0.461 1.052 1.026 2.500
- 0.75 0.554 1.057 1.028 2.020
- 1.25 0.693 1.062 1.031 1.500
2.5 0.918 1.071 1.035 0.967
5.0 1.257 1.086 1.042 0.695
7.5 1.524 1.098 1.048 0.554
10 1.752 1.107 1.052 0.474
15 2.119 1.119 1.058 0.379
20 2.401 1.128 1.062 0.320
25 2.618 1.133 1.064 0.278
30 2.782 1.138 1.067 0.245
35 2.899 1.142 1.069 0.219
40 . 2.971 1.145 1.070 0.197
45 3.000 1.148 1.071 0.178
50 2.985 1.151 1.073 0.161
55 2.925 1.153 1.074 0.145
60 2.815 1.155 1.075 0.130
65 2.611 1.154 1.074 0.116
70 2.316 1.118 1.057 0.102
75 1.953 1.081 1.040 0.089
80 1.543 1.040 1.020 0.075
85 1.107 0.996 0.998 0.061
90 0.665 0.948 0.974 0.047
95 0.262 0.890 0.943 0.030
100 0 0.822 0.907 0

L.E. radius: 0.223 per cent c

Figure 17.9: NACA 66-006 Basic Thickness Distribution


January 5, 1994, Profile Characteristics 259

chord where the lift starts to decrease. A camberline which is frequently


used is the NACA a=0.8 mean line, as given in Fig. 17.10. The table gives

2.0
re

1.0

o
ACA a 08 (modified)
mean ilne
2

o
o .2 .4 .6 e 10
x/c

ti, = 1.0 a; = 1.400 c.,,/, = 0.219


x Y<
dy,/cis PR AV/V = P R/4
(per cent c) (per cent c)
0 0
0.5 0.281 0.47539
0.75 0.396 0.44004
1.25 0.603 0.39531
2.5 1.055 0.33404 1.092 0.273
..

5.0 1.803 0.27149


7.5 2.432 0.23378
10 2.981 0.20618
15 3.903 0.16546
20 4.651 0.13452
1.096 0.274
25 5.257 0.10873
30 5.742 0.08595
35 6.120 0.06498
40 6.394 0.04507 1.100 0.275
45 6.571 0.02559
50 6.651 0.00607
55 6.631 - 0.01404 1.104 0.276
60 6.508 - 0.03537
65 6.274 - 0.05887 1.108 0.277
70 5.913 - 0.08610 1.108 0.277
75 5.401 - 0.12058 1.112 0.278
80 4.673 - 0.18034 1.112 0.278
85 3.607 - 0.23430 0.840 0.210
90 2.452 - 0.24521 0.588 0.147
95 1.226 - 0.24521 0.368 0.092
100 0 - 0.24521 0 0

Figure 17.10: NACA a=0.8 Mean Line


260 G.Kuiper, Resistance and Propulsion, January 5, 1994

the coordinates and the slope of the mean line. The slope can be used to
combine the mean line with a thickness distribution in the NACA way. The
parameter ./3,. is the local load coefficient, which is the difference between
the velocity at the upper surface p+ and that of the lower surface p- in
non-dimensional terms: PR = (p- p+)/0.5pV2. This load distribution is
given at the ideal angle of &tack ai = 1.4 deg.,which is given in the head
of the table. The lift coefficient at the ideal angle of attack cii = 1.0.

The pressure difference of a profile with zero thickness (a camber line) is


caused by a pressure decrease at the upper side and a pressure increase at
the lower side relative to the ambient pressure po From thin profile theory
it follows that the increase and the decrease have the same magnitude, so
the mean pressure at the camberline is equal to po The velocity increment
Av/V is the velocity increase at the upper side or the velocity decrease at
the lower sie of the camberline. It can be shown with Bernoulli's law that
Av = PRI4. The velocities at upper and lower side can be used to find the
local pressures and velocities on a combination of a thickness distribution
and a camberline.

Integration of the local load coefficient PR gives the lift coefficient. The
moment of the load distribution around the quarter chord point x = cI4
can also be calculated. The moment coefficient (per unit span) is in non-
dimensional form
Cm =
0.5pV2c2
From the moment and the lift coefficient the chordwise location of the re-
sultant lift can be calculated. This location is called the aerodynamic center

When the maximum camber is different from the maximum camber in


Table 17.10 all values have to be multiplied with the same factor as the
multiplication factor of the maximum camber.

17.5.3 Derivation of the Local Pressure of a Profile.


The velocities in the Tables were derived from potential theory. The veloc-
ities due to different causes may therefore be added. The velocity around
a profile with a certain camber- and thickness distribution at an angle of
attack can be considered as the sum of the velocities due to the thickness
distribution, the velocities due to the angle of attack and the velocities due
to the camber distribution. Assume the profile NACA 66-006 a=0.8, which
is a combination of the NACA 66-006 thickness distribution (Fig. 17.9) and
January 5, 1994, Profile Characteristics 261

the NACA a=0.8 camberline (Fig. 17.10) , at an angle of attack of 2 de-


grees. The velocity v/V at 10 percent of the chord on the suction side is
then:
the velocity due to thickness: 1.052 (table 17.9.
the velocity due to angle of attack. This angle is 2 degrees or 0.035
radians. The change in lift coefficient is thus 2r0.035 = 0.22. From
Ava/V in Table 17.9 the resulting velocity is 0.474 x 0.22 = 0.104.
the velocity due to camber. From Table 17.10 this is found to be
0.273.

The local velocity is thus 1.052 + 0.104 + 0.273 = 1.429 and the pressure
coefficient Cp = 1.04. At the pressure side the effects of angle of attack
and camber are opposite, so the velocity on the pressure side is 1.052
0.104 0.273 = 0.675.

17.5.4 Considerations to Choose or Design a Profile.


The foregoing examples are taken from [1]. This book contains many profile
series. Many propeller manufacturers have designed their own profiles or
have made modifications on the NACA profiles to serve a certain purpose.
A frequently used modification is the NACA66(modified) profile. This is
a modification of the NACA 66 profile, made by the "David Taylor Model
Basin" in Washington DC to make it more suitable for propellers.
Since it is often used the coordinates are given in Table 17.1. The main
modification is a thicker trailing edge, which makes the propeller blades
less vulnerable in backing conditions. Profiles for nozzles have more severe
restrictions from a manufacturing point of view, since these are made of
welded plates with webs. The pressure side of these profiles is often straight.
An example is the profile used in the Nozzle 19a from the Wageningen Ka
series, as given in Fig. 17.11.
The calculation of the pressure distribution over the nozzle and the
design of profiles with prescribed pressure distributions is very well possible
nowadays. These methods will not be treated here, but in more advanced
applications the use of profile series is replaced by profiles designed for a
specific purpose.
262 G.Kuiper, Resistance and Propulsion, January 5, 1994

chord position thickness


0 0
0.005 0.0665
0.0075 0.0812
0.0125 0.1044
0.025 0.1469
0.05 0.2066
0.075 0.2525
0.1 0.2907
0.15 0.3521
0.2 0.400
0.25 0.4363
0.3 0.4673
0.35 0.4832
0.4 0.4952
0.45 0.5
0.5 0.4962
0.55 0.4848
0.6 0.4653
0.65 0.4383
0.7 0.4035
0.75 0.3612
0.8 0.3110
0.85 0.2532
0.9 0.1877
0.95 0.1143
0.975 0.0748
1.0 0.0333

Table 17.1: Geometry of the NACA 66 Modified Thickness Distribution


January 5, 1994, Profile Characteristics 263

line
straight
cylindrical
c ear-
ance C

Figure 17.11: Profile of a Nozzle.


264 G.Kuiper, Resistance and Propulsion, January 5, 1994
Chapter 18
Cavitation

Objective: Introduction of cavitation and its detrimental effects

Cavitation is the phenomenon that water changes its phase into vapor
in flow regions with very low pressures. These low pressures are caused
by local high velocities. Cavitation is different from boiling because the
process of vaporization in cavitation is nearly isothermal, while in boiling
the vaporization is fed by heat transfer. The formation of vapor requires
some heat, but in cavitating flow this amount is so small that only a very
thin region around the cavity has a lower temperature than the mean tem-
perature.
The process of beginning of cavitation is called cavitation inception Pure
.

water can withstand very low pressures without cavitation inception. A


prerequisite for inception is the occurrence of weak spots in the flow, which
break the bond between the water molecules. These weak spots are gener-
ally tiny gas bubbles called nuclei . The presence of nuclei in water depends
on the circumstances. In sea water ample nuclei of all sizes are present, and
the inception pressure will be equal to the vapor pressure. At model scale
a lack of nuclei is common and the inception pressure will be lower than
the vapor pressure. This is a major cause of scale effects at model scale,
which means that the model conditions differ from those at full scale. There
are many causes of scale effects, which make it difficult to simulate cavita-
tion properly at model scale. In this chapter these scale effects will not be
treated. The picture given below of cavitation and its effects is therefore
very schematic. As always reality is infinitely more complex and this is very
strong in case of cavitation.

265
266 G.Kuiper, Resistance and Propulsion, January 5, 1994

18.1 The Cavitation Number.


The situation at full scale is that cavitation inception occurs when the local
pressure is equal to the vapor pressure. The local pressure is expressed in
non-dimensional terms as the pressure coefficient Cp. Similarly the cavita-
tion number is expressed non-dimensionally as
Po Pv
= (18.1)
0.5pV 2
where po is the undisturbed pressure in the flow, A, is the vapor pressure
and V is the undisturbed velocity of the fluid. The condition that cavitation
occurs when the local pressure is equal to the vapor pressure, means that
e.g. a profile or a propeller will begin to cavitate when the lowest pressure
is at the vapor pressure. This is expressed in non- dimensional terms as

o- = Cp(min)

The cavitation number or cavitation index o- is non-dimensional. This


means that it is the parameter which has to be maintained when model
tests are carried out. The cavitation index determines the pressure in the
test section of a cavitation tunnel.

18.2 Types of Cavitation.


The main parameter controlling this appearance is the pressure gradient.
However, cavitation has many different appearances and the judgment of
the effects is complicated. Some types of cavitation will be mentioned first in
a schematic way. After that the effects of the various types of cavitation are
discussed. Again it should be emphasized that the following classification
of cavitation is very schematic.

18.2.1 Bubble Cavitation.


When the fluid elements in the flow experience a gradual decrease in pres-
sure, cavitation will occur in the fluid. The nuclei will be the cause of
isolated cavities. These cavities move with the flow. This type of cavitation
is called bubble cavitation . It occurs when the minimum pressure on a blade
section is in the midchord region and when this minimum pressure is lower
than the vapor pressure. An example is given in Fig. 18.1. This type of
cavitation occurs when the blade sections are relatively thick and operate
at a small angle of attack. Near the root of controllable pitch propellers,
where the chord length is restricted because the blades have to pass each
January 5, 1994, Cavitation 267

other while the strength requires thick blade sections bubble cavitation is
sometimes difficult to avoid.

., .
-,..
--.4.,
--..v...,,...,.:-
7%

.
.
V
m"

Vet+
,

Figure 18.1: Bubble Cavitation on a Propeller Blade.

18.2.2 Sheet Cavitation.


When the pressure distribution has a strong adverse pressure gradient the
flow will separate from the body and a region of cavitation occurs. This
happens typically when a leading edge suction peak is present on a profile
while the minimum pressure is lower than the vapor pressure. In such a case
sheet cavitation occurs. Sheet cavitation is attached to the foil and the flow
moves around the sheet. The pressure in the cavity is approximately equal
to the vapor pressure. An example of such a sheet cavity is given in Fig. 18.2.

On commercial propellers the sheet cavity gradually merges with the tip
vortex. The rear of the cavity is smooth in such cases, as in Fig. 18.2. When
the tip of the blades is unloaded, as is often the case with Navy propellers,
268 G.Kuiper, Resistance and Propulsion, January 5, 1994

"`" Ara,.
1/4

Figure 18.2: Sheet Cavitation on a Propeller Blade

the length of the sheet cavity decreases towards the tip. The rear of the
cavity becomes cloudy in that case, as is illustrated in Fig. 18.3.

18.2.3 Root Cavitation.


At the blade root a type of cavitation can be present with a typical shape,
as is illustrated in Fig. 18.1. The root cavity has the shape of a wedge. The
top of the wedge can be at the leading edge, but it can also start on the
blade itself. Root cavitation is related with the horse shoe vortex which is
present at the blade root.

18.2.4 Tip Vortex Cavitation.


At the tip and hub of a propeller blade strong vortices leave the blade or the
hub. The pressure in the core of these vortices is low and when this pressure
is lower than the vapor pressure vortex cavitation occurs. An example is
given in Fig. 18.4. The vortex is generally very stable, so that the end is
far downstream in the flow. When the vortex passes a strong wake peak
it may break up and cause a complicated type of cloud cavitation (Fig. 18.5).
January 5, 1994, Cavitation 269

N0111.

r so,
41Ww.e.".14;'

-,etr
-

Figure 18.3: Sheet Cavitation with Unloaded Tip.

18.2.5 Propeller Hull Vortex Cavitation.


A special form of vortex cavitation occurs when a strong wake peak inter-
acts with the propeller in such a way that the tip vortex connects with the
hull. In that case a propeller hull vortex occurs (PHV-cavitation), as shown
in Fig. 18.6. This type of cavitation causes damage to the plating and an
extremely high noise level.

18.2.6 Unsteady Sheet Cavitation.


A propeller blade operates in a wake. The angles of attack of the blade
sections will therefore vary during one revolution and consequently the cav-
itation extent will vary with the blade position. In most cases there are
blade positions outside the wake peak where no cavitation will occur, or
where even pressure side cavitation will occur at the pressure side of the
blades. In the wake peak cavitation will occur at the suction side. The
growth and collapse of cavitation causes a break-up of the sheet cavitation
270 G.Kuiper, Resistance and Propulsion, January 5, 1994

Figure 18.4: Tip Vortex Cavitation.

Figure 18.5: Tip Vortex Breaking up in a Wake.

into a cloud of vortices and bubbles. Such a type of cavitation is called


cloud cavitation. An example is given in Fig 18.7. The development of
January 5, 1994, Cavitation 271

-..-.... I
-

Figure 18.6: Propeller Hull Vortex Cavitation.

cloud cavitation occurs when during the development of the sheet cavity a
part of the cavity separates from the main cavity and collapses separately
while moving with the fluid. An example of the development of such a de-
tached cavity is shown in Fig. 18.8.

18.2.7 The Mechanism of the Development of Cloud


Cavitation
As mentioned above cloud cavitation occurs at the rear edge of a steady
sheet or as a result of unsteady behavior of the sheet. The mechanism which
controls the development of cloud cavitation is not clear, but some aspects
are important.

The first aspect is the occurrence of a re-entrant jet at the rear end of
the cavity. In Fig. 18.9a a typical cross-section of a sheet cavity is shown.
The contour of a cavity has very little friction, so the flow can be considered
as inviscid. The pressure in the cavity is equal to the vapor pressure P.
The streamline just outside the cavity will approach the profile surface at a
272 G.Kuiper, Resistance and Propulsion, January 5, 1994

,
-
4'44

.4411.

.1 r
je
-011r
a 4

'frd
4

Figure 18.7: Cloud Cavitation at Full Scale.

large angle and the pressure at the surface will be much higher than p. In
case of a cavity closure perpendicular to the profile contour the pressure in
the fluid at the rear of the cavity will even be equal to the stagnation pres-
sure Po + 0.5pV2. This condition cannot exist in static equilibrium. At the
rear of the cavity a jet develops into the cavity, as shown in Fig. 18.9b. At
some moment the re-entrant jet hits the cavity surface and a complex situ-
ation occurs, where a part of the cavity becomes separated from the sheet
(Fig. 18.9c). It will move with the flow and collapse when arriving in a
region with higher pressures. The collapse is very complicated, because the
shape of the separated vapor region is far from spherical or two-dimensional.
Instead the vapor separates into parts and vortical structures are often ob-
served. This complex system of vapor and fluid is called cloud cavitation.

In two dimensional flow the process of separation can become very vi-
olent, when the re-entrant jet hits the front of the cavity and a large part
of the sheet separates. This is illustrated in Fig. 18.10. The view is on the
suction side of a profile in a narrow tunnel.
January 5, 1994, Cavitation 273

4;00':41!'t
- .46,
-

Figure 18.8: Development of Detached Cavitation on the Propeller of a


Containership (Full Scale).

In three-dimensional flow the re-entrant jet is deflected outward when


the cavity length is increasing with the radius. Such a situation is present
in Fig. 18.2. The result is a smooth rear end of the cavity without cloud
cavitation. When the cavity length becomes more or less constant, as in
Fig. 18.3 outside r=0.7R, the re-entrant jet mechanism generates cloud cav-
itation. The separated region is much smaller than the whole sheet, so a
cloudy edge occurs. It is not yet clear if cloud cavitation such as in Figs 18.7
and 18.8 can be explained by the behavior of a quasi-steady re-entrant jet
mechanism or if dynamic phenomena play a separate rle.

Cavitation can have four possible detrimental effects: erosion, radiated


noise, vibrations and loss of thrust.
274 G.Kuiper, Resistance and Propulsion, January 5, 1994

Figure 18.9: Development of Cloud Cavitation through a Re-entrant Jet.

%IF

Figure 18.10: Cloud Cavitation on a Profile.

18.3 Noise and Erosion.


All types of cavitation generate noise. Bubble cavitation is generally con-
sidered to be erosive. Cloud cavitation is considered very erosive. From
experience pressure side cavitation is also erosive. In propeller design it is
January 5, 1994, Cavitation 275

therefore tried to avoid these types of cavitation.

First the mechanism of erosion and noise generation will be discussed.


Generation of vapor from the fluid is a very rapid process. This means that
a vapor bubble, which moves into a lower pressure, will expand rapidly while
the pressure inside remains very close to the vapor pressure. When such
a cavity arrives in a region with a pressure higher than the vapor pressure
the same occurs: the bubble decreases in size without the pressure inside
becoming higher. When the bubble becomes very small the surface tension
also becomes large and this accelerates the collapse. The cavity therefore
collapses violently. This is the source of noise. When this occurs close to or
on the surface, the surface may be damaged. This damage is called erosion.
(Erosion is mechanical damage while corrosion is chemical damage to the
material)

18.3.1 The Implosion of a Single Bubble Cavity.


There are two mechanism of surface erosion. Consider a small bubble which
collapses close to the wall. In the final stage the bubble does not remain
symmetrical but deforms, as sketched in Fig. 18.11. In the center of the
bubble a jet is formed in the direction of the wall. The velocity in this
microjet is very high and the pressure when hitting the wall can be several
thousands of bars. This equals a hit with a very small hammer on the wall.
The result of cavitation erosion is a pitted surface.

SPI0Cle - WAVE 5

Figure 18.11: Collapse of a Cavity near a Wall.


276 G.Kuiper, Resistance and Propulsion, January 5, 1994

During the collapse of the cavity the velocity of the cavity wall becomes
extremely high, far higher than the velocity of sound in the fluid. Although
the fluid is generally considered incompressible, this is no longer the case in
the final stage of cavitation collapse. Around the collapsing bubbles shock
waves are formed, which also hit the wall. This mechanism explains why
imploding cavities can still damage the surface, even when they implode at
some distance from the wall.

When there is no gas in the cavity, the cavity will simply disappear after
collapse, this, however, is never the case. For inception of a bubble cavity
a small gas bubble (nucleus) is already required. During the expansion of
this nucleus gas is collected in the 'cavity by diffusion (Cavitation is an ef-
fective means of de-aerating the water). At the end of the collapse a small
amount of gas at very high pressure remains (The pressure is so high that
the gas can radiate light). This gas expands again and the bubble cavity
rebounds as numerous small bubbles. These bubbles act again as cavities
and collapse again. In this way the collapse of a single bubble cavity can
produce a multitude of pits and a very complex noise spectrum.

When a cloud of bubbles collapse simultaneously, the collapse can be


more violent than the collapse of single bubbles in the cloud. This is because
the pressure distribution in the cloud during collapse produces a higher
mean pressure in the center of the cloud, thus intensifying the collapse of
bubbles in that region. This explains why the collective collapse of bubbles,
as occurs in cloud cavitation, is so erosive.

18.3.2 Noise Radiation.


The shock-waves emitted from the cavity collapse result in radiation of a
high level of noise. The noise consists of " pressure spikes" due to the
collapse of isolated bubbles. Noise is generally decomposed into frequency
bands. A sharp spike has a constant energy in all frequencies. In the lower
frequencies (below say 5kHz) there is noise from all possible sources such
as the engine, the propeller without cavitation etc. Cavitation can have a
significant contribution in those frequencies. Cavitation has a specific con-
tribution in the higher frequencies, up to even 100KHz. An increase of the
noise level in those frequencies can be used to detect cavitation.

Radiated noise may be a problem for the habitability of ships. In that


case the inboard noise is a problem. A cavity type which generated very
strong inboard noise problems is the PHV-cavitation. Radiated noise is a
main problem for Navy ships. Their noise pattern is used to identify them.
Also the detection of submarines is seriously hampered by propeller noise
January 5, 1994, Cavitation 277

of the ship itself, while weapon systems like torpedo's are triggered by a
noise source, similarly as in air the projectiles are triggered by heat sources.
So for Navy ships there are many reasons to avoid noise. An important
characteristic for a Navy ship is therefore its inception speed, the speed at
which cavitation on the propeller begins.
subsectionVibrations.
Sheet cavitation can have a considerable volume. The dynamic behavior
of this large volume of vapor generates strong pressure fluctuations in the
water. The frequencies involved are the blade frequency (shaft frequency
times the number of blades) and multiples of the blade frequency. These
frequencies are lower than the noise frequencies. The pressure fluctuations
around the dynamic cavity have a wavelength which is large relative to the
distance to the hull. The pressure fluctuations are therefore independent
of the compressibility of the fluid. The pressure in the whole space under
consideration varies in phase with the pressure at the cavity surface. The
independence of the compressibility of the fluid distinguishes the cavitation
induced pressure fluctuations from cavitation induced noise.

The constant phase of the cavitation induced pressures makes that these
are effective in causing hull vibrations. This is different from the pressure
field from the passage of a blade without cavitation, which is felt at differ-
ent times at different places along the hull. This is sketched in Fig. 18.12,
where the distribution of the pressure disturbance at a certain time and
blade position is sketched. The pressure distribution due to cavitation
reaches its maximum and minimum everywhere on the hull at the same
time (in phase). The pressure distribution due to the passing blade with-
out cavitation is a pressure wave which moves over the hull surface with
the passing over the blade. Also the pressure amplitude of the cavitation
induced pressures decreases with 1/r (r is the distance to the cavity) while
the pressure amplitude of the non-cavitating blade decreases with 1/r2, so
much faster. When integrated over the hull the cavitation induced pres-
sures result in a much larger integrated force on the hull as the pressures
of the non-cavitating propeller, where the pressure peak and the trough are
opposite and the area at which the pressures act is smaller. 1.

The area over which the pressure fluctuations are integrated is impor-
tant. Especially in open stern container vessels and ships like that this the
hull area integrating the pressure fluctuations is large. Cavitation induced
are then very effective in causing hull vibrations. The flexible, open con-
struction of e.g. Ro-Ro- ships in the stern makes that the response of the
1The modelling of these pressure fluctuations as sources and dipoles are not treated
here
278 G.Kuiper, Resistance and Propulsion, January 5, 1994

// \\
N CAVITATION

/ \1
0
, I //
NON - CAVITATI NG

Figure 18.12: Pressure Distribution due to Cavitation and due to a Blade


Passage.

construction to these hull pressures is strong, resulting in vibrations. Un-


acceptable hull vibrations can be countered by redesigning the propeller or
by changing the response of the construction. The latter is only effective
when the vibrations are local. The most effective way to avoid vibrations
is to make the wake as uniform as possible.

18.3.3 Thrust Breakdown.


Partial cavitation on a profile will not affect its lift. On the contrary, a
small amount of cavitation may even increase the camber of the profile and
thus increase the lift. The effect of cavitation on the pressure distribution
at the suction side of a profile is shown in Fig. 18.13. At a cavitation index
above 1.25 there is no cavitation and there is a leading edge low pressure
peak with a minimum pressure coefficient of -1.18. At a cavitation index of
1.0 cavitation occurs (a < Cp(min)). Due to cavitation the minimum pres-
sure coefficient becomes equal to the cavitation index, which means that
the pressure at the cavity is equal to the vapor pressure. The pressure in
the leading edge pressure peak is thus increased. The cavitation extent,
however, is larger than the length of the minimum pressure peak, so the
pressure behind this peak is reduced. As a result the effect of cavitation on
the lift is minimal. This goes on until the cavity length is a considerable
fraction of the chord length (a = 0.5). When the mean pressure is further
January 5, 1994, Cavitation 279

reduced the cavitation causes a reduction of the lift. This reduction is grad-
ual, but fairly rapid.

s
(X = 3 e

sA = 0,0294
reA = 0.7500

,7*--
...................... . t .............0,5.. _,:,..._ _ 0,5
.......... ..... .
_ 0.3

tr x--
. . _ ..... 7 ..... .7 ..... .. - -_-___ .
p 0,3
_./_ 0 ,_
4

4,0-0.5 Is -05

-1,0

Figure 18.13: Pressure Distribution on a Profile at Various Cavitation Num-


bers.

On propellers different blade sections will suffer from a decrease of lift


at different conditions, so the effect of cavitation on the thrust will be more
gradual than on a single profile. When the cavitation becomes very exten-
sive at all radii the propeller thrust will disappear and the propeller suffers
from thrust breakdown indexthrust breakdown. On commercial propellers
this will rarely happen, because the propeller loading and the rotation rate
will be low enough. On highly loaded propellers and especially on propellers
which operate at high speed the thrust can be limited by cavitation.

The B-series propellers have been investigated at different cavitation


numbers, so that from the corrected open water diagrams the effect of cavi-
tation on thrust and torque can be found. An example is given in Fig. 18.14.
A measurable effect of cavitation on thrust is often encountered at Navy-
ships at full power or at tugs in towing condition, but also the performance
of fast ferries and fast containerships can be affected by cavitation.
280 G.Kuiper, Resistance and Propulsion, January 5, 1994

e 4-95
F' .10
I

10 =.
It k .4/4.1 ..A11116.---
:: ' r MIPIIMEM
4
NNWk4- A

NEIMI1
03 WPM
Ilbrfti
gift- ..411
,--imullmi
Ql
wilibh..
0 _
09 1.0 11 LS

Figure 18.14: Example of Open Water Diagram with Cavitation.

18.4 The Cavitation Bucket.


On profiles the minimum pressure coefficient determines the inception of
cavitation. The minimum pressure coefficient Cp(min) is generally plotted
as a function of the angle of attack as in Fig. 18.15
At small angles of attack the minimum pressure varies only slightly with
the angle of attack. The cornerpoints of the bucket indicate the angles of
attack where the leading edge suction peak begins.

The cavitation bucket' is important to judge the risk of cavitation on the


profile. The criterion for occurrence of cavitation is the minimum pressure
being equal to the vapor pressure, or when the cavitation index is equal
21n the past the plot was rotated over 90 degrees, so that the curve had the shape of
a bucket. This name has persisted ever since, so the curve is called a bucket
January 5, 1994, Cavitation 281

3 SHEET CAVITATION
SUCTION SIDE
a
2

BUBBLE
CAVIT.
NO CAVITATION

o
G- 2
Cp (min.)
SHEET CAVITATION
PRESSURE SIDE

Figure 18.15: Cavitation Bucket of a Profile.

to the minimum (negative) pressure coefficient. When the cavitation index


and the angle of attack is known the risk of cavitation can be read from the
bucket. The value a is plotted on the Cp axis of the cavitation bucket.
The range of angles of attack at that Cp-value within the bucket is the
range where the minimum pressure is higher than the vapor pressure. No
cavitation will occur there. Outside the bucket cavitation will occur. The
position outside the cavitation bucket gives an indication of the type of
cavitation. When the cavitation index is in the range of the horizontal
branches of the bucket sheet cavitation will occur, caused by the leading
edge suction peak. When the cavitation index is lower than the bottom of
the bucket bubble cavitation will occur.
The cavitation bucket belongs to the profile geometry. An increase in
maximum camber ( using the same camber distribution) results in a mainly
vertical shift (upwards) of the bucket. An increase in maximum thickness
(using the same thickness distribution) results in a wider bucket ( both
cornerpoints move in opposite directions). Because the minimum pressure
also decreases when the profile becomes thicker the bottom of the bucket
moves to the right at the same time.
A variation of the maximum camber is shown in Fig. 18.16. A variation
of the maximum thickness is shown in Fig. 18.17.

It should be mentioned that instead of the angle of attack a cavitation


bucket can also plotted against the lift coefficient. Because of the nearly
linear relation between the two this does not change the shape of the bucket.
282 G.Kuiper, Resistance and Propulsion, January 5, 1994

Arnew 08 c...80tC0re-
cp(C 77-/IcA-NYIV rro4 ,4 riwY k/

Figure 18.16: Cavitation Buckets for Varying Camber.

5
NACA 66 (TIM MOD NOSE & TAIL)
NACA a .1 CALIDERLINE

%f1:48ESRSA TO
EKidallipplal PPI,Nli
CL ..
CALIBER
1097 ( - .83 4R0 4.70) womOrdomm.mummemmill
d -41 - -.1
4
0 04 DECREES
MEMMEgg .01111 EM
WIAMENNEMad IIIMINmaiiiii
.. .4mErm mill- mg
BE iiII
iatki... mrommulOMM

loteil i 'Rift
am
Webbingliftwitimaffillin
mil
11441,06ME-N4M mu
III
11 I'
r i .,S2

"..." '

tlik
"TV'. --"..0111111
.5414

imumnssimar._ hENEMISBANft
,_ i
MomMA
3

1 "hmgsamawAsmmwmm7-
immulEmmEmms.lammp
-5

Figure 18.17: Cavitation Buckets for Varying Thickness (from Brockett).

The cavitation behavior of a propeller is represented in the inception


diagram as given in Fig. 18.18. The design condition of the propeller is
January 5, 1994, Cavitation 283

indicated with a cross. The inception lines of the various types of cavi-
tation are given on the basis of the propeller thrust coefficient. The sets
of curves for sheet and tip vortex cavitation form again a kind of bucket.
When the operational condition is in the bucket there is no cavitation of
that type. A typical curve of the propeller at various ship speeds is also
given in Fig. 18.18. As shown the thrust coefficient changes only slightly
with increasing ship speed, which indicated that both the thrust and the
resistance increase with ri2.

6
TIP VORTEX FACE TIP VORTEX BACK
(SCALED UP) (SCALED UP)

5
FACE SHEET

On

BACK SHEET
3

BACK BUBBLES

o
-0.1 0 0.1 0.2 0.3 0.4
K1

Figure 18.18: Inception Diagram of a Propeller.


284 G.Kuiper, Resistance and Propulsion, January 5, 1994

18.5 Cavitation Tests.


Cavitation is investigated at model scale using a cavitation tunneL This
is a closed loop in which the water is circulating. A cross section of a
large cavitation tunnel (the Grand Tunnel Hydrodynamique in Val de Reuil,
France) is given in Fig. 18.1-9. The velocity in most of the circuit is very low

large test section (LTS) Downstream tank (DI)

NEE ItiINIO
;:.,mi
'1111111UP
Resorber Pump

Figure 18.19: Cross Section of a Cavitation Tunnel.


and the pressure is high, so that turbulence can disappear and gas bubbles
from the cavitation in the test section can rise to the surface or dissolve.
Upstream of the test section a contraction accelerates the fluid, so that in
the test section the velocity is high and the pressure low. A propeller in the
test section can be driven by a shaft, mostly from downstream. A square
angle drive from the top of the tunnel is also applied.
In general a cavitation tunnel has no free surface. Consequently the
Froude number disappears as a scaling law and propellers can be tested
at higher rotation rates than according to the Froude scaling law. This
is important, because at low Reynolds numbers the boundary layer at the
propeller blades remains laminar over large areas, which results in scale ef-
fects on cavitation inception and on performance. To avoid these laminar
flow effects the leading edge of propeller blades is often roughened. This
has to be done with great care, to avoid that cavitation is controlled by the
amount of roughness.

The absence of a free surface does not completely remove effects of grav-
ity. There is still a variation in pressure over the height of the test section.
This results in a distribution of the cavitation index over the height of the
test section, and thus over the propeller disk. This distribution is only sim-
ilar to full scale when the Froude number is maintained. When e.g. the
rotation rate of the model propeller in the test section is taken twice the
January 5, 1994, Cavitation 285

rotation rate according to Froude, the cavitation index varies less over the
height of the propeller. This is illustrated in Fig. 18.20. Although this effect

/WATER SURFACE
o

(m)

to 2.0 2.5 3.0


an

X r MOOEL LENGTH SCALE FACTOR


n PROPELLER RATE OF ROTATION
Por STATIC PRESSURE
Pvr. VAPOUR PRESSURE
p MASS DENSITY OF WATER
D r PROPELLER DIAMETER

FULL SCALE OR MODEL SCALE VvITH nmodel n ship /-27 ( FROUDE )


6 02: MODEL SCALE WITH nmodel r 2 ^ship 151-

Figure 18.20: Variation of the Cavitation Index at two Rotation Rates.

is generally neglected, it can cause deviations in the cavitation behavior on


a blade in various positions.

An important aspect of cavitation testing is the effect of a wake. Most


tunnels are too small to accommodate the complete model, and even when
that is possible the wall effects will be large. Instead of using a complete
model, a scaled dummy model is often used in front of the propeller. This
dummy is often not a geometrically scaled part of the hull. The breadth
and the height may have different scale factors. The use of wires and san-
droughness on the dummy makes it possible to control the wake distribution
to some extent. When such a dummy is used the wake which is generated
in the propeller plane has to be measured and if necessary adjusted to make
it the same as the target wake distribution. The target wake distribution
is the nominal wake distribution from the towing tank, corrected for scale
286 G.Kuiper, Resistance and Propulsion, January 5, 1994

effects. Such a correction is also used for the extrapolation of the wake
fraction of a propulsion test, as treated in chapter 20.
The simplest technique to simulate a wake distribution is by screens in front
of the propeller. The interaction between propeller and hull is different how-
ever, from the interaction between propeller and screen, so this technique
requires much experience. Also the tangential velocities in the wake are not
properly simulated with screens.
Chapter 19
Lifting Line Propeller Design
Objective: The design of a propeller in a circumferential averaged wake
with an arbitrary radial loading distribution

In the foregoing the flow around the propeller blades has not been con-
sidered. In the actuator disk theory the propeller has been replaced by an
actuator disk, in the systematic series approach the global characteristics
of a propeller have been considered only. A more detailed treatment of the
flow around the propeller blades is given by the blade element theory. In
this approach the flow over the propeller is considered in a cylindrical cross-
section, similar as in Fig. 15.2 on page 210. The velocity components in
this plane as well as the blade section and the resulting forces on the blade
are given in Fig. 19.1. This is an important diagram and the elements will
be considered in detail.

The velocity components of the undisturbed flow (without the propeller)


are the axial velocity Ve and the rotational velocity x7rnD. The resultant
of these two components is the resulting velocity V.

The propeller induces additional velocities in the propeller plane. The


axial component of these velocities has been estimated in the actuator disk
model in chapter 14. The axial induced velocities va and the tangential
induced velocities vt together form the induced velocity vi. The resultant
of Vr and vi is the toal velocity 14, which is the actual inflow velocity of
the blade section. When the propeller loading is light or moderate and the
fluid is inviscid the induced velocity component vi is perpendicular to the
total velocity 14.
1The diagram is at an arbitrary radial cross section at radius r, so x = r/R. R is the
outer radius of the propeller, or half the diameter D.

287
288 G.Kuiper, Resistance and Propulsion, January 5, 1994

Figure 19.1: Velocity Diagram and Forces on a Blade Section.

The blade section has a pitch angle Op and is at an angle of attack a


with the total velocity Vt. The result of this is a lift force L. The lift force is
defined perpendicular to the incoming velocity V. Apart from the lift force
also a drag force D is generated by the blade section. Lift and drag can be
decomposed into thrust and torque.

The important simplification of this blade element approach is that the


January 5, 1994, Lifting Line Theory 289

radial velocities are neglected . As a result the blade section can be treated
as a two-dimensional profile at an angle of attack. The lift and drag gener-
ated by the blade section depend only on the angle of attack in this plane
and they are independent of the other radii. The problem is not fully two-
dimensional, however, because the induced velocity 7.7i depends not only
on the lift generaied at this radius, but also on the lift generated at all
other radii. The three- dimensional nature of the flow around the propeller
blades is thus reduced to the determination of the induced velocities. For
this determination the lifting line model is used.

19.1 Lifting Line Theory.


19.1.1 Two-dimensional Lifting Lines.
As mentioned in chapter 17, a profile in an inviscid fluid has no drag (para-
dox of d'Alembert), but generates a lift force perpendicular to the incoming
flow. The lift coefficient has the magnitude 27r( ao) (see eq. 17.2 on
page 251). Such a lift force in an incoming flow can be represented in po-
tential theory by a vortex. This was illustrated in the example in chapter 10
by the flow around a cylinder with circulation around it. In that case the
cylinder was represented by a dipole and the circulation was represented
by a vortex, both in the center of the cylinder. Now the dipole is omitted,
making the diameter of the cylinder zero. The vortex remains and the cir-
culation F is the strength of the vortex. In the thin airfoil theory it can
be shown that the vortex strength r is the actual circulation around the
profile. The lift generated by the vortex in a velocity V is pyr. So there is
a direct relation between the vortex strength and the angle of attack

r r(a ao)VS
where S is the chord times the span of the profile.

In this way the lift of the profile is represented by a two-dimensional


lifting line. Note that the drag has been neglected, lifting line theory is a
potential theory. The velocities in the flow outside the lifting line can be
determined with the law of Biot and Savart (eq. 10.6 on 154) to be F/27rx,
where x is the distance to the lifting line.

19.1.2 Lifting Lines in Three Dimensions.


Without proof it is stated here that a vortex in three dimensions cannot
end in a fluid. Similarly it cannot change in strength, because that would
mean that a part of the vorticity would disappear. This has consequences
290 G.Kuiper, Resistance and Propulsion, January 5, 1994

for a profile with a finite span. Consider a profile with a constant lift force
but with a finite span, as in Fig. 19.2. The profile is replaced by a vortex

Free vortex
Y

Figure 19.2: Bound and Trailing Vortices on a Profile with Uniform Span-
wise Loading.

of strength F. Since this vortex is fixed to the profile it is called the bound
vortex . The vortex cannot end in space and at the tips free vortices or trail-
ing vortices will leave the wing and move with the flow. When the loading
of the profile is light the path of the free vortices can be approximated by
the plane through the profile and the velocity vector. Far behind the wing
a starting vortex connects the two trailing vortices, so that these trailing
vortices do not end in the flow either. The starting vortex is considered to
be at infinity and therefore has no influence. It is therefore neglected .

The presence of the two trailing vortices has an important effect. The
trailing vortices induce a vertical velocity at the location of the lzfting line.
This is in contrast with the bound vortex elements, which do not induce a
velocity at other positions of the lifting line. When the lift is upwards, the
velocity induced by the trailing vortices is downwards. The induced velocity
component is therefore also called downwash . As a result of the downwash
the angle of attack of the inflow velocity is changed. In flow direction the
situation is as in Fig. 19.3. The lift L' is perpendicular to the total velocity
Vt, which is no longer in flow direction. As a result the lift has a component
in flow direction. This is a drag force, even while the fluid is invicid. This
drag force is called the induced drag . When the downwash is small relative
to the undisturbed flow velocity V the induced angle of attack ai is small
and the difference between L and L' is of 0(c4) and can be neglected. The
induced drag is of 0(cri) , however, and cannot be neglected. So
January 5, 1994, Lifting Line Theory 291

Figure 19.3: Inflow and Lift on a three-dimensional Profile.

The main effect of a three dimensional profile is the oc-


currence of the induced angle of attack and the induced
drag.

When a profile has a variable spanwise loading the situation is as in


Fig. 19.4. The strength of the bound vortex is r(s), where s is the position

Figure 19.4: Bound and Trailing Vortices on a Profile with Variable Span-
wise Loading.

on the span. Trailing vortices are shed from the bound vortex at all locations
and the strength of the trailing vortices is E. The downwash at an arbitrary
position of the span can be calculated by integrating the induced velocities
of all trailing vortex elements, as illustrated in Fig. 10.3 on page 155.
292 G.Kuiper, Resistance and Propulsion, January 5, 1994

19.1.3 Lifting Line Theory for a Propeller.


The lifting line model for a propeller is the same as for a three-dimensional
profile. In case of a ligthly loaded profile the downwash is small and the
position of the trailing vortices can be approximated by a plane through the
lifting line. This is more complicated in case of a propeller. The trailing
vortex will move with the flow, but the direction in the wake of a propeller
is not known a priory. As a first approximation the trailing vortices are
always positioned in the undisturbed flow. So the induced velocities in the
wake are neglected. The lifting line model of the tip vortex is then as given
in Fig. 19.5. The trailing vortex is a helicoidal line with a pitch angle equal

Figure 19.5: Bound and Trailing Vortex of a Propeller.

to the hydrodynamical pitch angle (see Fig. 19.1). The diameter and the
pitch of the trailing vortex remains constant.

This propeller wake model ignores two phenomena. The first is the
contraction of the wake, as was calculated e.g.with the actuator disk model.
The wake model is therefore only valid for ligthly loaded propellers. The
second is the roll-up of the vortices. This can be illustrated by the flow
behind a wing. Two adjacent trailing vortices A and B, both trailing from
the wing, have the same direction of rotation. As illustrated in Fig. 19.6 the
velocity of vortex A at the location of vortex B is downwards. The velocity
induced by vortex B at the location of vortex A is upwards. The result is
that the two vortices tend to wind around each other, which phenomenon
is called roll-up. Especially in the tip region, 'Where the trailing vortices are
January 5, 1994, Lifting Line Theory 293

strong, the helical plane in which the vortices are shed, tends to roll up into
a single tip vortex. The calculated position of the trailing vortices at some
distance behind a wing is shown in Fig. 19.6. Instead of being in the plane
of the wing, the trailing vortices are moved downwards (downwash) and the
plane is rolling up at its tips into the tip vortices of a wing.

0.1

Y 0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8I 3-As4-
0.9x 0 x)c 1.0

x - -x- -x - x-x -x-x -x x-x-x-x-x-x-x-x-x-x-x*x-x-xxx-xlook s


-01 -

Figure 19.6: Position of Trailing Vortices behind a Wing.

The wake model is important for the calculation of the induced veloc-
ities in the propeller plane. The induced axial and tangential velocities in
any location in the propeller plane can be calculated by integrating the
velocities, induced by the trailing vortices. This can be done numerically
using the law of Biot and Savart for each element of the trailing vortices.

The wake model can be made somewhat more realistic by taking the
pitch angle of the trailing vortices equal to the induced pitch angle Oi . This
creates a dilemma, however, because this angle depends on the induced ve-
locities and these induced velocities depend on the position of the trailing
vortices. The solution has to be obtained by iteration.

The lift at each blade section is also called the sectional loading. Note
that the radial loading distribution is not the thrust distribution over the
radius, because the loading is perpendicular to the total velocity Vt at each
radius (Fig. 19.1). The result of the lifting line theory for propeller is that

A relation is formulated between the radial loading dis-


tribution and the axial and tangential induced velocities
in the propeller plane

19.2 Optimum Radial Loading Distribution.


From the axial momentum theory a simple criterium for the optimum radial
loading distribution was found. The axial velocity should be independent
of the radius at and behind the propeller. In the blade element approach
294 G.Kuiper, Resistance and Propulsion, January 5, 1994

also tangential velocities are induced, as shown in Fig. 19.1. A sectional


efficiency of a blade element at radius r with a sectional thrust dT and a
moment dQ can be written as
dTV,
_ 77(r) =
dQ2irn

For a given total thrust the maximum efficiency is obtained when 77(r) is
constant over the radius. Otherwise thrust can be replaced from a radius
with a low local efficiency to a radius with a higher local efficiency, thus
increasing the local efficiency.

In terms of the lifting line approach the sectional thrust can be written
as
dT = pr(r)(27nr vt)
and the sectional torque can be written as

dQ = pf(r)r(Ve + va)
The sectional efficiency can thus be written as

(271-nr vt)Ve
=
(Ve va.)27rnr

The sectional efficiency can thus be written as

tan(13)
=
tan ()
and in case of an optimum efficiency this ratio has to be constant over the
radius. It leads to a certain radial distribution of the circulation r , which
has been named after Betz, who derived it in 1919.

19.3 Induction Factors.


Having established the position of the free vortices behind the propeller a
relation between the strenght of a trailing vortex and the induced velocity
components va and vt in the propeller plane can be derived using potential
theory. Because of the spiral shape of the trailing vortices this is a com-
plicated formulation. An analytical solution has been performed by Lerbs
(1951) [25], who presented the solution in diagrams. An analytical solution
has been formulated by Wrench in 1957 (for the formulations see [32]). The
precise formulation not important now, but it has the following form:
January 5, 1994, Lifting Line Theory 295

ua
= 1/2[TO /A(r,A,Z)ar f (ve,Ro,r')dr' (19.1)
Ve J rh ar
where R0 is the propeller outer radius and rh is the hub radius. The
integrand is the induced velocity at radius r due to the Z helicoidal trailing
vortices with strength R: at radius r'. The integral combines the induced
velocities from the trailing vortices at all radii. This reflects that the in-
duced velocities of the trailing vortices at all radii influence the induced
velocity at radius r. This is the three dimensional nature of the problem.

The function IA is the induction factor for the axial induced velocity.
The induction factor provides a relation between the radial loading distrib-
ution and the induced velocities at the propeller, provided the pitch of the
trailing vortices tan(A) is known.

The equation for the tangential induced velocity has the same form.
So the induction factors for axial and tangential induced velocity together
determine the induced velocity in Fig. 19.1.

19.4 Propeller Design using the Induction


Factors.
The design of a propeller is an implicit problem: the angle of attack of the
blade sections, and thus the lift, depends on the induced velocities, while
the induced velocities depend on the radial lift distribution. The problem
has to be solved by iteration.

19.4.1 Determination of the Inflow.


A lifting surface propeller design can start by prescribing a required pro-
peller thrust T and a radial loading distribution -L-P7 at the design entrance
velocity V, and propeller revolutions n. The entrance velocity Ve can be
a function of the radius, so that the tangential mean velocity of the wake
can be taken into account. The propellers thus designed are therefore also
called wake adapted propellers .
This starting point is a main difference with the use of series such as the
B-series, because there the radial loading distribution is fixed and the en-
trance velocity is the same at all radii because the open water diagrams are
made in uniform flow.
296 G.Kuiper, Resistance and Propulsion, January 5, 1994

The loading distribution is of course related with the trust by the relation
To
Lcos(3i) Dsin(A)dr = T (19.2)

The induced pitch angle is not known at this stage of the design, so as
a first approximation the hydrodynamic pitch angle has to be used. The
radial distribution of the loading L has to be chosen now. This distribution
is generally chosen for optimum efficiency, as has been discussed before, or
it can be chosen on the basis of the propeller requirements. Navy propellers
e.g. often have a strongly reduced tip loading to delay the inception of
tip vortex cavitation. An unloaded tip can be necessary to prevent hull
pressures due to cavitation. From the radial loading distribution L(r) the
radial distribution of the circulation can be found from
L(r)cos(0) D(r)sin(0)
r(r)
P177-

Here the velocity V, and the hydrodynamic pitch angle is used, both
without the induced velocities.

The drag D is also unknown at this stage, but as a first approximation


a drag coefficient of 0.0008 can be used. Together with V, instead if values
are set. The initial estimate of the chordlength is also used in the deter-
mination of the section geometry as described below. It is also possible to
neglect the drag force altogether in the first estimate. The viscous drag
force calculated from the first iteration can then be taken into account in
subsequent iterations.

From r(r) the induced velocities in axial and tangential direction va and
vt can now be calculated when the pitch angle of the trailing vortices is also
taken as the hydrodynamic pitch angle 0. This gives the induced velocities
and a first value for the induced pitch angle

With this first value of the process can be repeated from the begin-
ning. The calculation is repeated until Oi converges to a constant value.
The convergence of the calculation is very rapid and requires only a few
iterations.

19.4.2 Determination of the Blade Sections.


Having determined the direction and magnitude of the incoming velocities
at each radius the propeller geometry can be designed. The blade sections
at each radius have to be placed such that the required circulation and lift
January 5, 1994, Lzfting Line Theory 297

is generated. From the sectional lift coefficient all parameters except the
chord length are known, so the product

cLc =
0.5pVt2

is known. The cavitation index at which the blade section has to operate
can be calculated from the (minimum) immersion hin, of the blade section.
The cavitation index
Patm Pghmin
0.5pVt2
can be calculated since all variables in it are known.

For the determination of the camber ratio and the chordlength the cavi-
tation buckets of a series of profiles is used. From such a series the thickness
and camber distribution are known, the chordlength c and the camber ratio
fmax/c and the thickness ratio t lc have to be determined.
First the criteria for the choice of the blade sections are formulated. The
blade section should have a such a bucket that the design condition has
a certain margin against bubble cavitation and against pressure side cavi-
tation, as shown in Fig. 19.7. The design condition is the combination of

BUCET

CL
DESIGN CNDITION

JI

Figure 19.7: Combination of Design Conditon and Cavitation Bucket.

the lift coefficient


CL and the cavitation index a. The margin against bub-
ble cavitation is expressed as a percentage of the cavitation index in the
January 5, 1994, Lifting Line Theory 299

From a calculation at various radii the blade geometry, the pitch and
the blade contour is found. This completes the design of the propeller.
However, the simplifications of the lifting line approach are too rigorous
and some corrections, the lifting surface corrections, still have to be applied.
These will be treated later in this chapter.

19.4.3 Strength of the Propeller.


The strength of a propeller depends on the material and on the maximum
stresses in the material. The maximum stresses occur as compressive stress
on the suction side of the propeller blade. So the compressive maximum
stress of the material determines the strength of the propeller. For cunial
e.g. this is approx. 5.106 N M2 .

19.4.4 Stresses due to Loading.


The determination of the maximum stresses in the blades is complicated. A
finite element calculation can give an answer in this case. For approximate
answers the propeller blade is considered as a beam, as sketched in Fig. 19.8.
The thrust is assumed to attach at radius r1. The blade thrust Tb and the

Tb

Figure 19.8: Model of the Propeller Blade as a Beam.

blade torque Qb are the torque and thrust of the propeler, divided by the
number of blades. We now consider the blade section at the root of the blade
r2, where the maximum stresses normally occur. When it is assumed that
the blade section has its minimum moment of resistance in the direction of
the induced angle of attack ,(3, (see Fig. 19.1) the important moment Mz. is
300 G.Kuiper, Resistance and Propulsion, January 5, 1994

in that direction. So

= Qbsin(A)+Tb(ri r2)cos(0i) (19.3)

The relation between thrust and torque can be expressed by the propeller
efficiency 77p:
TV, TbVe
77P 2rQn 2rQbn
So
2r Q bnrip
Tb =
Ve

The blade torque can be expressed in terms of the power P = 2rQn =


2rZQbn delivered to the propeller. Using

Qb
2rnZ
equation 19.3 can be written as
P nD r2 sin(A) V, D 1 1

Mx = Zn V, (19.4)
D riP 1_77P 2r nDr1 r2 tg(3i)j
The maximum stresses in the blade section are found from
Mx
o- =

where W is the moment of resistance of the blade section. W can be written


as W = act2, where c is the chord length of the blade section at r2 and t is
the maximum thickness.

19.4.5 Stresses due to Centrifugal Forces.


Centrifugal forces cause a reduction of the compressive stress at the root.
A positive rake, however, makes that the centrifugal forces increase the
compressive stress again. Writing the mass of a propeller blade as Mb the
moment around Mr at the root is

Mz.(centr.) = M(wr)2[tg(e) l] 1(/3)

and the additional stress can be found from

cr(centr) =
,
111(centr)
cfct2
January 5, 1994, Lifting Line Theory 301

19.4.6 Approximate Methods.


The classification societies have issued calculation techniques based on this
type of formula's. Romsom has given calculation results from a beam theory
as described above in 1952 [41]. He used empirical values for moments due
to thrust and torque. He gives the maximum stress at two radii (0.2R and
0.6R) as
P 1
=
nZct2`jikL'2 tg(A)
The value tan(A) for a series propeller such as the B-series is not directly
known. It can be replaced by tan() = tan(f3)/i or tan(131) = JI(rnp).
The equation for the additional stresses due to centrifugal forces is written
as
a(centr)= n2 D2 [A. C K]
The values of C1, A and C are given by Romson in graphical form, both
for r2 = 0.2 and r2 = 0.6. The formula for C2 and K is given in formulas.
The Figures and formula of Romson are given below because of the original
publication is rather inaccessible. The diagrams below are given from van
Manen [28] (Note the units in these diagrams!).

Similar diagrams for the B-series propellers are given in Figs. 19.11 and
19.12, where the variable 8 is the inverse of the advance ratio J.

A considerable approximation of the beam approach is that the neutral


line in the blade sections is not the straight line at an angle Oi. The blades
are thin and the stresses are as in a thin shell, where the neutral line is
curved. So the value of a is a rough approximation, which is implicitly used
in the diagrams of Romson. Also the attachment location of the thrust 7.1
depends on the radial loading distribution. In the calculations of Romson
the optimum loading distribution has been used. The results from the beam
theory therefore have to be applied with empirical correction factors. The
correction factors are generally combined in the allowable stress, which is
taken some ten times the tensile strength of the material. This safety factor
also takes care of the variations in the stress due to the non-uniformity of
the wake and of the margin for fatigue.

19.4.7 Lifting Surface Corrections.


The simplification of the blade section from a camberline into one point is
a too drastic simplification of the situation. A correction of this relation
necessary. This correction originates from the fact that the induced velocity
vi is only an average over the chord. Due to the three dimensional character
302 G.Kuiper, Resistance and Propulsion, January 5, 1994

rip = 0,6

7 _
COMPRESSIVE STRESS.cenrraugal stress excluded

6= SHPmetr g C,lC2+_ A.L)


A 1,14

z n 0.085 s 'I.
EVA
WO
,
6 _ 112
where n = rpm / .6

5
:: turiCtIOrl of rake angie All A
Allpro.de
2: nurriber of Wades
C, = function of 110.6/0
4 1.08
c2=0.66 PI"/D

Ilidil Allall
S =thickness in cm
3 i =chord
Ac.: Ho., ,,,eran....4
TTO ,,
or.
ten_grh in m
1,06

2 ....dimor-4-
po-0,01p--mpl-m:da
II 1.04

,.._,,,,..- _......1.0v--
1

1,02

o
20
,- 4 6. B.
rake ante
10
in degrees
COMPRESSIVE STRESS DIJE TO CENTRIFUGAL FORCE
120 14 160 180
1,0r

Cc = e 02CA C -0.345)
where n- r P---
10M0

100
D= propeuerdiam in m
A= function of D/506 anct rake
angte
I C= function o( H0,6/0

60
_
40 0,9

0,7
_
0,5
05 06 07 08 09 1,0 11 12 1,3 1,4
H/0 r =0..

Figure 19.9: Stresses at 0.6R.

of the flow the induced velocity varies over the chord. This results in a
"curved" flow pattern along the chord, as shown in Fig. 19.13.
The curvature of the streamlines, resulting from the variation of the
induced velocity over the chord, reduces the camber of the profile. An "ef-
fective" camber can be found by straightening the flow and the camberline.
On this straight flow the two-dimensional relation between lift and angle of
attack can be applied.
January 5, 1994, Lifting Line Theory 303

i
r/R=0,2 1

COMPRESSIVE STRESS,centrifugal stress excluded


, A
SHPmetr 4 c, (c 2 4. 7 j
1 1,14

where
z n 0,0 8 5

n= rpm
521

1
oz'FrArrOpl
4414 1,12

=function of rake angle


O
Z =number of blades r AA 1,10
Ci=function of H02/0
OP''' pp
C2= 2,25 H0.2
s=thickness in cm r
I:chord length in m
Ai= H 92._
i0

fi z r
-40 Ii
IN .-
1
I/
1,08

,06
1Td

MI ,04

--------_%--11A ,0 2

25
Il ,00

20


150 11 O

c
8

4
100 _
09 1,1 1,3
Hip at 1113=0,2

Figure 19.10: Stresses at 0.2R.

To account for this effect the camber of the blade sections as found from
the buckets have to be corrected, because these sections were the effective
sections. This means that for the propeller geometry the camber has to be
increased to get the same lift in a curved flow. When the variation of the in-
duced velocities over the chord is not symmetrical to the midchord position
a correction of the angle of attack is also necessary. These corrections are
called lifting surface corrections because these are calculated using a lifting
surface calculation which, in contrast to the lifting line calculation, spreads
304 G.Kuiper, Resistance and Propulsion, January 5, 1994

rip= 0,6
COMPRESSIVE STRESS .centrifuga( stress excluded
d SHPmetr Ca (Cb +5.'20
7
sz I 1j4
z.n 0.085 1---
where n =rpm
6 - =function of rake angle 1.12
C1/4-o-.6711.3-1
z =number of blades
Ca=function of H0.6/0 .
00(
5
Cb: 19.4 1-10.6/0
A4 S} 8d-5 diagr 1,08
72P
szthIckness n cm
i = choro tenrh in 1,06
1 3

2 #dleill,1111.-.111
-morlAilMilliglill
f
1,04

_oil ...a .110%11111 1,02

o 1,00
6 8 10 12 14 16 18
rake angle ,n degrees

COMPRESSIVE STRESS, OUE TO CENTRIFUGAL FORCE


4 6,. n2.02 rA C -0.345)
where n =rpm/joo
0 = propellerdiam in m
A :ftion
unc 0/sas,and rake
- tion of Hog 0
1.1

0,9

O 0.7

0.5 0,6 0,7 08 09 1,0 1,1 1,2 1,3 1,4


--a... Ric) at riP = 0.6
Figure 19.11: Stresses at 0.6R for B-series Propellers.

the vorticity also over the chord. The correction due to the curvature of
the flow can be separated into a correction on the maximum camber and a
correction on the angle of attack. The camber correction k, is:

k, = f eom
Jeff
where f1 the corrected "effective " (two-dimensional) camber and
fgeo7n is the actual geometry of the blade section. The camber of the blade
January 5, 1994, Lzfting Line Theory 305

rIR =02
COMPRESSIVE STRE SS, centrifugal stress excluded
6= SHPmetr.E Ca (Cb51p)

A
z.n.0.085 5 2.1

where :n=rpm
function of rake angle
=

z= number of blades DA0.2A

pjoilli
Ca= f unction of H0.2/0
1
Cb= 66,7 Fl0/0
SI18p -6 r.
7, p clia9
s= thickness in cm

/All 1K11011.I
WW1"
i= chord length in
5 1
1,05

1,00
50 rake angle 10 in degrees 16
COMPRESSIVE STRESS DUE TO CENTRIFUGAL FORCE
c= n2 D2 (A.0 -0.58)
where n_ LP-Ln
100
kb0= propellerdiam in m

Ca
6 illimi
A=function of D/50,2and rake
libb.5=function of Ho2/0
1,0

--in .6

2
O 0,7
,2
0,9 1,1 1,3
1.4/0 at riiR =0,2

Figure 19.12: Stresses at 0.2R for B-series Propellers.

section, derived from the buckets, is the effective camber and this has to be
multiplied by Ke to arrive at the geometrical camber of the propeller blade
section.

When it is assumed that the blade section is at its ideal angle of attack,
this ideal angle of attack requires a lifting surface correction Ko

K = c'gern
aef f
306 G.Kuiper, Resistance and Propulsion, January 5, 1994

GEOMETRICAL CAMBER I EFFECTIVE CAMBER


DISTRIB ION i DISTRIBUTION

fjp.

Figure 19.13: Curved Streamline near a Blade Section.

where ceeff is again the "effective"( two dimensional) ideal angle of at-
tack and %eon-, is the corrected angle of attack for the propeller blade section.
Again the ideal angle of attack derived from the buckets was the "effective
angle and this angle has to be corrected to find the geometrical angle, from
which the pitch is derived.

The neglect of the blade thickness is correct in two-dimensional flow.


In three dimensional flow a finite thickness distribution leads to a small
correction of the angle of attack. This correction is taken into account by a
thickness correction factor. The relation between the "effective" ideal angle
of attack and the geometrical ideal angle of attack is then written as

aef f &age, ktBT F


where BTF is
tmaxl D 0. 003
BT F = + 0.003
1 r
with tmax as the maximum thickness at that radius.

Lifting surface corrections for the camber have been calculated already
in 1944 by Ludwieg and Ginzel . Systematic calculations using a lifting
surface program has been given for a series of propellers by Morgan et al in
1968 [31]. Morgan et al calculated the corrections for camber and angle of
attack for a series of conventional, non-skewed propellers with a blade area
ratio from 0.35 to 1.15 and an induced advance coefficient rtan(f3i) -=
from 0.4 to 2. These calculations were made for 4,5 and 6 bladed propellers.
The correction factors are given in tables.
January 5, 1994, Lifting Line Theory 307

To give an indication of the magnitude of the lifting surface corrections


some Figures from Morgan as shown here. An example of the camber
correction factors is given in Fig. 19.14. The correction factor for the ideal

3.0

t=s wk1=1.2

/./s
2.0
0.95

0.55

0.35
1.0
0.3 0.4 0.5 0.6 0.7 0.8 09

Figure 19.14: Camber Correction Factor for 5 blades and R-Ai = 1.2.

angle of attack is given by Morgan et al as

a(georn)
ka(Morgan)

Here ai3O is the ideal angle of attack for CL = 1 for the NACA a=0.8 mean
line. Since this angle is 1.4 deg., the angle of attack of the blade section can
be found from 1.41c,,CL degrees. An example of the lifting surface corrections
for the angle of attack is given in Fig. 19.15.
An example of the diagrams for the correction for blade thickness ict from
[31] is given in Fig. 19.16.
308 G.Kuiper, Resistance and Propulsion, January 5, 1994

irk1 Ar
1 .2 `/A = 0.75

2.0

ka

1.0
0.1 0.4 0.5 0.6 07 08 09

Figure 19.15: Ideal Angle of Attack Correction Factor for 5 Blades and
Blade Area Ratio of 0.75.

19.4.8 Viscous Forces.


The previous design method is completely based on potential theory and is
therefore inviscid. The effect of viscosity can simply be taken into account
by adding a drag force in the direction of Vt. The magnitude of the drag
force can be estimated by assuming a constant lift-drag ratio CIICd. Such a
lift drag ratio can be found from data of the two-dimensional profile series.
Approximate values of 15 to 25 can be used for the lift-drag ratio. Since
the angles of attack are always small in the outer radii also a constant drag
coefficient of 0.04 to 0.08 is sometimes used.

The drag force gives an increase in the propeller torque, but also a
decrease in the propeller thrust (see Fig. 19.1). In the calculation of the
circulation required for a prescribed thrust the calculated drag force can be
used for a second iteration if the initial estimate was not accurate enough.
January 5, 1994, Lifting Line Theory 309

yki =1.2 A
E/A = 0.75
o

1.0

Z:z4

0.3 0.4 0.5 0.6 0.7 08 09

Figure 19.16: Correction Factor for Induced Angle of Attack from Thickness
for r = 1.2 and Blade Area Ratio 0.75.
310 G.Kuiper, Resistance and Propulsion, January 5, 1994
Chapter 20
The Propulsion Test
Purpose: The prediction of the full scale performance from model test re-
sults

In the foregoing the hull and the propeller were treated separately. Now
both will be brought together in order to predict the propulsive behaviour
of the ship with propeller.

The propulsive properties of a ship with propeller are experimentally


determined in a propulsion test. In this test the ship model is driven by the
propeller at a certain ship speed. The model speed can be related to the
ship speed using the Froude number:
V, V,
,VFLT \fir,
So V = Vs/VEr, where a is the scale ratio. When no external forces are
present this condition is called the self propulsion point of model.

20.1 The Additional Towing Force.


As discussed in the resistance test the Reynolds number at model scale
is necessarily much lower than at full scale, resulting in a relatively higher
resistance at model scale. In the self propulsion condition at model scale the
propeller is therefore much heavier loaded at the same Froude number than
at full scale. In order to scale the propeller loading properly an additional
towing force on the model hull is necessary. The towing force can simply be
derived from the difference in total resistance coefficient of the model and
the ship:

311
312 G.Kuiper, Resistance and Propulsion, January 6, 1994

ACD = Cts Ct, = (1 + k) {C f(Rnm) C f(R,)} Ca (20.1)

The value of ACD depends on the extrapolator used. The extra towing
force, often indicated as the additional towing force of the model, can be
found from CD:

FD = AC D(1/2pV7,2 S) (20.2)
where S is the wetted surface of the hull as defined in the resistance test.

20.1.1 Self Propulsion Test with an Additional Tow-


ing Force.
One way to carry out a propulsion test is to estimate the rotation rate
required for a certain model speed. The additional towing force can be
calculated and is generally applied to the model hull by a weight, as shown
in Fig. 20.1. The model is tested with the propeller operating and the
additional towing force applied. The speed of the model is measured. When
the model speed deviates too much from the required ship speed, either
the additional towing force or the rotation rate of the propeller has to be
adjusted. This is only necessary at a few model speeds, the others can
be found accurately enough by interpolation. When the additional towing
force is applied the velocity of the model at a certain propeller rotation rate
is the self propulsion point of ship.

Figure 20.1: Application of Additional Towing Force.


January 6, 1994, The Propulsion Test 313

20.2 Overload Tests.


When very large and heavy models are used it can take a long time before
the model speed in the towing tank is constant. When the length of the
tank does not allow this, an overload test can be applied. In that case the
model is fixed to the carriage. The model is towed at a prescribed speed
with various rotation rates of the propeller. The towing force between the
carriage and the model is then measured. The self propulsion point of ship
is found from interpolation at the correct additional towing force FD. 1

When the extrapolator is not yet decided or may be changed afterwards


the overload test is at an advantage. The drawback is of course that it
takes more runs. A method to reduce the number of runs in an overload
test will be shown in the example in this chapter.

A propulsion test is carried out to determine the required power and the
obtained speed at a chosen propeller rpm. For that purpose the propeller
torque is also measured. For reasons of scaling, as will be discussed later, the
propeller thrust is also always measured. This applies both to the overload
test and to the free running propulsion test.

20.3 Scaling Laws.


The propulsion test is always carried out at the proper Froude number,
so that the wave pattern is properly scaled. When no viscous effects were
present the kinematic similarity between model and ship would ensure the
same advance ratio J at model and ship, so

Vs
= V,
n8D, nD,
Or

V, D.,
n= V, D,
ns

which results in n, = IA-ns.


In principle the propulsion test at model scale with the proper towing
force FD provides the non-dimensional coefficients KT and /fQ of the pro-
peller. These are non-dimensional and can therefore be used at full scale.
From these coefficients the forces and moments, and thus the required engine
power, at full scale can be found. However, the difference of the Reynolds
'In older publications the overload test is also called the continental method, the
other method being the British method.
314 G.Kuiper, Resistance and Propulsion, January 6, 1994

number between model and full scale makes this extrapolation considerably
more complicated.

20.3.1 Scale Effects.


There are still two major scale effects to be accounted for. The first scale
effect is that the propeller operates behind the model where, even when an
additional towing force FD is applied, the wake is still too large. The wake
is reflecting the resistance of the model, as shown in chapter 3. With the
(considerable) simplification of a uniform propeller inflow this is reflected
by a difference in wake fraction between model and ship.

The second deviation is due to the fact that the propeller operates at
model Reynolds number. The boundary layer at the propeller blades of the
model is also too thick. This requires corrections for the relation between
the rotation rate and torque and thrust. This is taken into account by
scaling the open water characteristics.

20.4 Propeller Hull Interaction.


To account for viscous scale effects the measured data of a propulsion test
are compared with the results of a resistance test at model scale and with
the open water characteristics of the propeller at model scale. From this
comparison the interaction effects between model and propeller are derived.
The viscous scale effects are subsequently taken into account by extrapo-
lating the interaction effects, the resistance and the open water character-
istics to full scale values. The extrapolation of the resistance test has been
described in chapter 7. The determination and the extrapolation of the
interaction effects and the extrapolation of the open water diagram will be
discussed next. The combination of the extrapolated full scale values will
give the prediction diagram for power and rotation rate as a function of
ship speed.

20.4.1 Thrust Deduction.


The propeller has an effect on the resistance of the hull. Consider a propul-
sion test with a towing force FD at a model velocity V, and propeller
rotation rate n. The torque Q and the thrust T are also measured. The
velocity V, is constant, so the total force in longitudinal direction on the
hull FD T is equal to the resistance R'of the hull. Since FD is known and
T is measured the model resistance R' can be calculated. The resistance R
at the same model speed can be found from a resistance test. A comparison
January 6, 1994, The Propulsion Test 315

will show that R' > R. This is due to the propeller action, which accelerates
the flow near the stern and thus lowers the pressure over the afterbody. The
difference in resistance due to the propeller is called the thrust deduction
and it is described as a thrust deduction factor t:

R' R
t=
R'

So the combination of the propulsion test with the resistance test gives
the thrust deduction factor, which expresses the resistance increase due to
the propeller action.

20.4.2 Taylor Wake Fraction.


Inversely the hull has an effect on the propeller inflow. It is not always
sufficient to measure the nominal wake fraction as mentioned in chapter 4,
because in behind condition the hull changes the velocities in front of the
propeller. The wake fraction in behind condition is now determined by
comparing the thrust or torque coefficients of the propulsion test with the
open water diagram of the same propeller. The advance ratio 4 at the
thrust coefficient from the propulsion test can be read from the open water
diagram of the propeller model. Since the propeller diameter and the ro-
tation rate at model scale is known, the inflow velocity Vet can be derived
and the wake fraction wt can be written as

Vs Vet
wt -=
vs

The wake fraction thus determined is the Taylor wake fraction , after
admiral David W. Taylor, who introduced it. It amounts to the use of the
propeller as a velocity measuring device with the open water curve as the
calibration.
The Taylor wake fraction is derived using the thrust coefficient of the propul-
sion test, or using thrust identity.. The Taylor wake fraction can in principle
also be found from the torque coefficient, so with torque identity, mainly
because of rotation of the wake. In practice there is a differences between
the wake fractions derived with thrust and torque identity. It is a matter of
definition and the thrust identity is generally used for the definition of the
Taylor wake fraction. The deviation between the torque Q' from the open
water diagram at thrust identity and the torque Q from the propulsion test
is called the relative rotative efficiency :
316 G.Kuiper, Resistance and Propulsion, January 6, 1994

Q'(openwater)
qr (20.3)
Q(prop.test)

Why this ratio is called an efficiency will become clear when the elements
of the efficiency are discussed.

It has been a point of controversy at which the rotation rate at which


the open water diagram has to be determined. In principle this should be
done at the same Reynolds number as the 'propeller behind the model in
the propulsion test, so at a Froude scaled rotation rate. In practice the
open water diagram is measured at a higher propeller Reynolds number, to
avoid laminar flow effects. In that case it is assumed that these effects do
not occur in behind condition.
Holtrop has proposed correctly to measure the open water characteris-
tics both at the model Reynolds number and at the maximum obtainable
Reynolds number. The difference can then be applied to the results of the
propulsion test. The open water diagram at high Reynolds number and
the corrected propulsion test results can then be used for extrapolation to
full scale. This is more elaborate, of course, because two open water curves
have to be measured.

Note that the velocity Vet from the Taylor wake fraction cannot directly
be measured in behind condition. It is not the water velocity just ahead
of the propeller in behind condition. The average velocity in the propeller
plane is the entrance velocity Ve plus the velocity induced by the propeller
action.

Note also that the Taylor wake fraction is not the nominal wake fraction
as measured with a pitot tube over the propeller disk, although the devi-
ations will not be large. The nominal wake fraction is measured over the
propeller disk and the Taylor wake fraction takes the inflow over a larger
diameter due to the contraction of the streamtube. The difference increases
therefore with increasing propeller loading and with a greater sensitivity of
the nominal wake to the propeller diameter. Also the presence of a rud-
der (generally considered to be part of the hull) behind the propeller in a
propulsion test has a significant effect on the thrust.

Note that the Taylor wake fraction refers to the wake over the propeller
disk and not over the total flow disturbance behind the ship, which will
extend over more than the ships breadth.
January 6, 1994, The Propulsion Test 317

Type of ship (1 w3)/(1 wm)


multiple screw ships 1.01
slender single screw ships 1.03-1.08
full single screw ships loaded 1.10-1.15
full single screw ships, ballast 1.15-1.30

Table 20.1: Typical values for wake scaling factors

20.5 Extrapolation of the Interaction Ef-


fects.
The thrust deduction factor is now considered to be independent of the
Reynolds number. This means physically that the thrust deduction is as-
sumed to be due to the pressure decrease in front of the propeller. The
thrust deduction is thus assumed to be an extra pressure resistance. The
increase in viscous resistance, caused by higher velocities over the hull, is
considered small.

The scaling of the Taylor wake fraction is more difficult, since no full
scale values can be measured. The extrapolation of the wake has to be used
in the extrapolation of the propulsion test and the prediction can be used to
compare with full scale results. This comparison can give statistical values,
but the extrapolation and thus the statistics contain much more than only
the wake scaling. A statistical formula for single screw ships as mentioned
in the ITTC 1957 report is

to, = (t + 0.04) + (w, + k)Cf, + Ca


t 0.04)(1 ( 20.4)
(1 + 1c)C f,
The wake fraction is thus related with the viscous resistance coefficient at
model and full scale and is changed proportionally. This is a reasonable
estimate for full ships with a heavily loaded propeller. For slender ships
this formula fails and a simple table, as given by Holtrop in Table 20.1,
gives an estimate. Each test facility has its own experience in extrapolating
the wake fraction from the model wake fraction.

20.6 Extrapolation of the Open Water Char-


acteristics.
The results of the open water test at model scale are different from the
full scale values due to viscous scale efFects. These scale effects have to be
318 G.Kuiper, Resistance and Propulsion, January 6, 1994

corrected to arrive at a prediction of the full scale propeller performance. 2


To extrapolate the open water diagram of the model to full scale the viscous
drag of the propeller is treated as the viscous drag of one blade section, the
equivalent blade section. Only the drag of the equivalent blade section is
considered because the Reynolds number is assumed to have no effect on
the lift of the equivalent blade section.

20.6.1 The Equivalent Blade Section.


As the equivalent blade section the section at 0.75R is generally taken. Cor-
rections to the drag of this blade section are related to corrections in thrust
and torque of the propeller. Lerbs(1951) [24] derived a relation between the
drag coefficient of the equivalent blade section and the propeller torque and
thrust using lifting line theory and some simplifications. Kuiper [22] gave a
simplified derivation of this relation, in which also the lift coefficient of the
equivalent blade section was related to the propeller torque and thrust. At
this point we will restrict ourselves to the relation the ITTC has adopted in
its ITTC57 extrapolation method. An empirical relation between a change
in the drag coefficient of the equivalent blade section and a change in the
propeller torque and thrust was given by the ITTC (Lindgren 1969), based
on correlation with only a few propellers:

Aift =- Kt, Kt, = ACD(equiv) x 0.28r5P (20.5)

AKQ = KqsKqm--= ACD(equiv) x 0.248 x (20.6)

Here c is the chord of the equivalent blade section, Z is the number of


blades, D is the propeller diameter and P is the propeller pitch. The drag
coefficient of a blade section is defined as

CD=
112pV2c1

where I is the span of the wing. The drag coefficient is also expressed as
the drag coefficient per unit length, so with 1 = 1.

The extrapolation of the open water characteristics is now reduced to the


determination of the change of the drag coefficient ACD of the equivalent
blade section.
2Just as with the ship resistance the open water characteristics at full scale are never
measured, but they are derived from propulsion test results.
January 6, 1994, The Propulsion Test 319

20.6.2 Extrapolation of the Drag Coefficient of the


Equivalent Blade Section.
An increase in Reynolds number from model to full scale will decreases the
CD-value of the equivalent profile. The decrease in drag coefficient can be
taken from the ITTC57 extrapolator using model and full scale Reynolds
numbers. The Reynolds number of the equivalent profile is based on V, at
the equivalent radius and on the chord length of the blade section at that
radius. The kinematic viscosity at model scale is taken at the temperature
of the tank water. The Reynolds number at full scale is always based on
the kinematic viscosity at a temperature of 15 degrees Celcius.

The correction according to the extrapolator line assumes that the bound-
ary layer at the model test is fully turbulent, since the ITTC line is a turbu-
lent plate line. Calculations of the drag coefficient of the equivalent blade
sections with Lerbs' method show however that at model Reynolds numbers
a considerable laminar effect is present, as shown in Fig. 20.2. A similar

0.02

Turbulent

-.Epormi
0
C .

Dann
(s/1 = 0.05)

0.01

Yokoo, C20]
Taniguchi, (21]
..4441114
SSPA, (22]
sSPA, (23)
van Manen, Victory ship, [10]
Laminar
Yokoo, [24] (s/1 = 0.06)
4.5 5.0 5.5 6.0
log Rn 0.75R

Figure 20.2: Drag Coefficients of Equivalent Blade Sections.

trend was found by van Oossanen in 1974 from similar calculations of the
B- series propellers (Fig. 20.3 [35]). The friction coefficient at full scale Cfs
is therefore taken from the ITTC57 extrapolator:

0.075
= (20.7)
(log R 2)2
320 G.Kuiper, Resistance and Propulsion, January 6, 1994

0.024 B3-80 B5-60


B3-65 A 85-75
a B3-50 85-105
a B3-35 B6-50
4- 84-40 B6-65
84-55 B6-80
0.020 B4-70 87-55
a B4-85 87-70
84-100 B7-85
a B5-45
V

0.016

0.012 2x ITTC 1957 turbulent flow line

CD

0.008 ++
4+

o 2x Blasius laminar flow line


0.004
o

1x10 2x105 3x105 4x105 5x105


Rn

Figure 20.3: Drag Coefficients of Equivalent Blade Sections of the B-Series


Propeller Models.

The total drag coefficient at model scale has been approximated from the
data of Fig. 20.2 as

cpm [1 211 [0.044 5 1

c
(20.8)

This line is shown in Fig. 20.2 as the ITTC line. Note that the effect of
blade thickness is accounted for by the factor I + 2tI c and that the factor 2
accounts for the fact that the wetted surface of a profile is twice the area IX C.
January 6, 1994, The Propulsion Test 321

At model scale the propeller is very smooth. At full scale the surface will
be rougher and the boundary layer is thinner, so the surface roughness at full
scale can increase the drag coefficient. An additional drag for the presence
of roughness at full scale is used in the ITTC method. As was shown in
Fig. 7.6 on page 114 the effects of small roughness elements become more
pronounced at higher Reynolds numbers and the drag coefficient becomes
independent of the Reynolds number. The ITTC57 roughness allowance is
therefore formulated as:
e -2.5
6Cf3 = (1.89 + 1.62 log - Cfs (20.9)

Here c is the chord at full scale of the 0.75R blade section. The roughness
height lc, is the equivalent roughness as defined by Nikuradse in 1942 (see
[42]). It is the roughness height which gives the same resistance coefficient
as a flat plate with a roughness height k, applied in a careful way. In prac-
tice the value of k, is taken as 150 microns (150 x 10-6m).

The drag coefficient of the equivalent blade section at full scale is now

CDs = 2 {1 + 2-t (Cfs + 6Cf8) (20.10)

The increase of the drag coefficient of the equivalent section is found from:

ACD Cpm CDs

The Reynolds number at the equivalent radius of 0.75R is based on the


undisturbed incoming velocity V,. and the chord at the equivalent radius.
So:

with
V,. = V(V(1 w))2 (27rnr)2 (20.11)
This is similar for model and full scale. The Reynolds number for the model
and the ship can be rewritten in non-dimensional terms as:

(1 wt)Vc1/1 + (77.)2
R, = (20.12)

in which x = of the equivalent radius. In practice the value 0.75 is always


used for the equivalent radius. The Reynolds number on the ship is found
from the same formula with ship values at 15 degrees Celcius.
322 G.Kuiper, Resistance and Propulsion, January 6, 1994

The above method is the ITTC-method. It is simpler to base the


Reynolds number on the rotational velocity only and to write eq. 20.11
as V,. = 2rnr, which gives a Reynolds number

nD2
= 0.75 r (20.13)
li

This makes little difference with the more accurate formulation because the
rotational velocity is far greater than the advance velocity at 0.75R. It also
has the advantage that the Reynolds number is independent of the advance
velocity Ve. The relation between model and full scale Reynolds number
can also simply be written as R,,, = Rnrn Vm

20.7 Extrapolation of the Propulsion Test


Results.
The extrapolation of the resistance test gives the resistance of the ship R,
at the ship speed Vs. The thrust deduction gives the propeller thrust

T, = R3I(1 t)

From the extrapolated wake fraction w, the entrance velocity of the pro-
peller can be found:
Ves = (1 ws)V3
With the full scale thrust and entrance velocity known the open water
diagram can be used, provided that the rotation rate n, is known. The
rotation rate can be eliminated, however, when the variable KT I J2 is used.
This variable can be calculated. A plot in the (extrapolated) open water
diagram gives the advance ratio J3. From J3 the rotation rate n, can be
found. At the derived J3, the value of 1-f, can be also be read from the
open water diagram, from which the required torque can be derived. The
required power in kW for that speed is found from

PD 2rQ3n3/1000

When this process is carried out at a range of ship speeds, the result is
the prediction of power and rotation rate at full scale versus the ship speed.
When an engine with a given power is installed the obtainable ship speed
can be read.
January 6, 1994, The Propulsion Test 323

20.8 Trial Condition and Service Condition.


The extrapolation of the resistance and propulsion test is based on an ideal
situation of perfectly calm water, no strong winds, no fouling on the hull
and the propeller. This condition is called the tank condition. Even in ideal
circumstances at sea there will be disturbances. This condition is called
the trial condition. In daily service the conditions will not be ideal and the
ship will experience an increased resistance due to fouling and due to the
weather. This condition is called the service condition. This will not only
lead to a decreased ship speed, but also to a different rotation rate, so that
e.g. a diesel engine cannot deliver its maximum power.

To avoid this situation of overloading the engine, a service allowance


is applied to the resistance of the ship. This correction is applied to the
(extrapolated) resistance curve. In the propulsion test this correction is
applied as a correction on the additional towing force. When an overload
test has been carried out the extrapolation for trial and service condition
can be done exactly as for the tank condition with a corrected FD.

When no overload test is present a trial or service allowance is applied


to the power instead of to the resistance. Assuming an unchanged total
efficiency the margins for the power can be equal to those applied to the
resistance. The power curve of the tank condition is increased by the mar-
gin. From the available shaft power of the engine the trial or service speed
is found. In that case the effect on the rotation rate of the propeller has to
be estimated. The rotation rate of the tank condition is then reduced by
e.g. 0.5 percent for each 10 percent allowance on the power. This corrected
rotation rate gives the rotation rate at the corrected ship speed.

The trial and service margin is often agreed upon in a contract. It is


generally expressed as a percentage of the tank power. So the resistance
correction has to be such that the required power increases with the con-
tractual margin. A value of 10 percent for the trial condition and 20 to 30
percent for the service condition is regular. The propeller rotation rate is
sometimes also lowered with 1 to 2 percent to allow for roughness of the
full scale propeller and for fouling of the blades.

20.9 Efficiencies.
The efficiency is the ratio of energy supplied to a system and the useful
energy delivered by the system. For the ship with propeller the energy
324 G.Kuiper, Resistance and Propulsion, January 6, 1994

output is RV,. The input is the engine power Qw. So the total efficiency is
RV,
= 27rnQ
This total efficiency can be divided into parts which are related to the
propeller performance without the hull and to the hull without the propeller.
The total efficiency is written as:

RV,
qt (20.14)
27rnQ
RV; TVE Qo
TV, 27rnQo Q
= 77hiloqr (20.15)
The total efficiency is thus divided into
The open water efficiency 770 _t_TVn
2irThis
riw o
is the efficiency of the
propeller alone in the mean inflow Ve. It can be derived from open
water diagrams of the propeller.
The hull efficiency rih = g. This efficiency can be expressed into the
Taylor wake fraction and the thrust deduction factor by writing

RV, T(1 t)V, 1 t


TV, TV3(1 w) 1 w
The thrust deduction is typically smaller than the wake fraction, so
the hull efficiency is larger than one!
The relative rotative efficiency 77r = 91 This efficiency reflects the
difference in torque in the wake and in open water at the same thrust.
The relative rotative efficiency is generally close to one.
The distinction of the open water efficiency and the hull efficiency is
important for the judgment of propeller designs. The traditional design
methods optimize the propeller alone. The interaction with the hull is
not taken into account. Still it is possible that a propeller with a lower
open water efficiency but a smaller diameter gives a higher total efficiency
because the interaction with the hull is better, which shows in a higher
hull efficiency. E.g. a higher wake fraction may improve the hull efficiency
when the total resistance remains the same. This reflects the requirement
of minimum energy loss in the wake. It is important that the energy loss
in the nominal wake is regained by the propeller. The propeller diameter
therefore has to be such that the wake is going through the propeller disk
as much as possible, as is illustrated in an idealized way in Fig. 20.4.
January 6, 199.4, The Propulsion Test 325

I fa

r-1 vs

WAKE

Li Hi

PROP.
IN.4tre
EXAM. I

k-1
I.

Vf.4. 2 Vr/
4

Figure 20.4: Relation between Optimum Propeller Diameter and Wake.

20.10 Variations on the Extrapolation Method.


In the previous method of extrapolation both the open water test results
and the resistance test results are required. To reduce the cost of testing,
efforts have been made to extrapolate the full scale prediction from the
propulsion test only. One method, which is applied successfully at Marin,
is the restricted overload test, as developed by Holtrop.

The test is an overload test, so the model is fixed to the carriage and the
model speed is fixed. The propeller revolutions are varied at each speed.
The test method is a restricted overload test because the variation of the
rpm is carried out at two or three model velocities only. Since the variations
of thrust, torque and towing force with rps are approximately linear and
constant, the self propulsion point of ship is calculated using these slopes at
other model speeds. In this way the results of the restricted overload test
is translated into the result of a propulsion test at the self propulsion point
of ship. This method is illustrated in the example in this chapter.

The results of the overload test can also be used to replace the resistance
test. Using the mentioned derivatives of thrust, torque and towing force to
the propeller revolutions, the conditions of zero thrust can also be found.
It can be assumed that in the condition of zero thrust of the propeller the
towing force is representative for the model resistance. the propeller at zero
326 G.Kuiper, Resistance and Propulsion, January 6, 1994

thrust may still affect the flow somewhat, because at zero thrust the lift at
all radii is not necessarily zero. A positive thrust at outer radii may be com-
pensated by a negative thrust at inner radii. Also there is still the rotation,
which the propeller induces by friction. In practice the resistance derived
from the restricted overload test differs a few percent from the results of
the resistance test.

With the restricted overload test results also the open water test can be
omitted. Instead of the open water curve the relation between the advance
ratio J,., -Z7z,
nmDm and the thrust and torque coefficient from the overload
points is used. In this relation the behind conditions are included, so that
no separate relative rotation efficiency is required. This method has an ad-
vantage because the relation between the open water test results and the
similar diagram from the overload test is not always constant over a large
range of loadings.

In the Marin method the advance ratio based on model speed is multi-
plied by the scaling ratio of the wake:

1 ws
and again the value (Kt + K)/ J2 as well as ifq is plotted against
this corrected advance ratio. The further procedure is the same as in the
extrapolation using the open water test. The extrapolation of the restricted
overload test is given in the example.

20.11 Example of Extrapolation of the Pro-


peller Open Water Diagram.
The open water diagram of Fig. 16.1, which happens to be the open water diagram of
the propeller of the containership in the example of chapter 7, has been measured at
model scale. Its characteristics are given in Table 20.2. It has been tested in open water
conditions at 12.9 degrees C and at 14.7 rps. The open water test results are given in
Table 20.3.
The Reynolds number of the equivalent blade section at 0.75R at model scale is calcu-
lated from eq. 20.13:

R,, = 0.757r 14.7 x 0.1241 x 0.2955 = 1.05 x 106


1.205 x 10-6
The drag coefficient at model scale is found from eq. 20.8:

0.044 5 1
CD,, = 2 [1 + 2 x 0.0305]
{10.082 10331] = 0.0082
January 7, 1994, The Propulsion Test 327

diameter D 6.5 m
Pitch ratio at 0.75r P075 0.907
Expanded blade area ratio AE/Ao 0.726
Number of blades Z 4
thickness ratio at 0.75R (t/c)0.75 0.0305
chord diameter ratio at 0.75R c0 75/D 0.42
model scale ratio A 22
revolution rate n, 2.079 sec-1

Table 20.2: Example of Full Scale Propeller Data

J Kto Kqo 170


0 .4959 .06646 0
0.1 .4514 .06106 .118
0.2 .4051 .0562 .232
0.3 .3574 .05010 .341
0.4 .3082 .04443 .442
0.5 .2578 .03858 .532
0.6 .2062 .03249 .606
0.7 .1573 .02610 .656
0.8 .1004 .01938 .660
0.9 .0464 .01226 .542
0.985 .000 .00584 .000

Table 20.3: Results of open water test

When the open water test was carried out at Froude scaled rpm the Reynolds number
at full scale is then found to be

= 1.05 x 106 x 221 51 1.07 x 108


1.1.28:53

The open water test is normally carried out at higher rpm than follows from Froude scal-
ing. In that case the full scale Reynolds number can be calculated also from eq. 20.13.

The friction coefficient from the ITTC57 extrapolator is found from eq. 20.7 or from
Table 22.3:
0.075
Cf, = = 0.0021
(8.029 - 2)2
The full scale drag coefficient without roughness will be
2(1 + 2 x 0.03050) x .0021 = .0045
This is indeed considerably lower than the model scale value of 0.0082. For the rough-
ness effect an equivalent roughness of 30 microns is chosen, so that the additional drag
coefficient is found from eq. 20.9:

2.73
- 2.5
bCf, = (1.89+ 1.62 log - 0.0021 = 0.0011
io x 10-6
328 G.Kuiper, Resistance and Propulsion, January 6, 1994

The friction coefficient of the full scale equivalent profile with the effect of roughness is
now 0.0021+0.0011 = 0.0032. The drag coefficient including the profile thickness is now
found from eq. 20.10:

Cps = 2 [1 + 2 x 0.0305] (0.0032) = 0.068

After all there is a decrease in drag coefficient from model to full scale of:

ACD = 0.0082 - 0.0068 = 0.0014


This results into an increase in thrust coefficient according to eq. 20.5:

AKt = -0.0014 x 0.28 x 0.907(0.42 x 4) = -0.00060


Similarly the increase in torque coefficient is found from eq. 20.6:

LRQ = 0.0014 x 0.248 x 0.42 x 4 = 0.00059

The model propeller operates at J = 0.633. The measured thrust coefficient at that
advance ratio is found from Fig. 16.1 to be 0.190. The torque coefficient is 0.0306. At
full scale the thrust and torque coefficients become 0.1901 and 0.0299 respectively at an
advance ratio of 0.626.
When the efficiency is calculated in both cases from eq. 16.1 the results are:

?Rim = 0.626

= 0.641
So the efficiency at model scale is about 1.5 percent lower than at full scale.

20.12 Example of the Extrapolation of the


Propulsion Test.
The propulsion test of the same Ro-Ro-ship with the data in Tables 7.2 and 20.2 will
now be used as an example for the extrapolation of a propulsion test.

The propulsion data from the measurement are given in Table 20.4.
This propulsion test was a limited overload test. At a speed of 18.5 and 21.5 knots
three propeller revolutions were tested. At the other speeds only one measurement was
done. The rpm at these model speeds were chosen on the basis of a prediction program
in such a way that the self propulsion point of ship was approximated.

The additional towing force can be calculated using eqs. 20.1 and 20.2. From the
extrapolation of the resistance test the friction coefficients for model (Table 7.4) and full
scale (Table 7.5) were already calculated. The additional resistance coefficient Ca is kept
at 0.0038. In the following the calculations for a speed of 21.5 knots will be illustrated.

The value of ACD at this speed is found from eq. 20.1:

ACD = 1.14(00275 - .00143) - 0.00038 = .00113


January 6, 1994, The Propulsion Test 329

Vs V, n F T Q
knots mlsec sec-1 N N Nm
17.02 1.867 7.250 33.36 74.85 3.419
18.51 2.030 5.580 99.72 11.33 .819
18.51 2.030 7.980 34.70 89.85 4.060
18.52 2.031 6.780 70.44 47.15 2.314
20.01 2.195 8.720 36.56 108.69 4.905
21.51 2.359 6.730 131.96 22.18 1.397
21.51 2.359 8.170 90.24 74.2 3.563
21.52 2.360 9.620 38.89 137.19 6.127
23.01 2.524 10.600 42.7 173.1 7.786

Table 20.4: Results of propulsion test

V, Vm FD n, K Kg
17 1.865 29.67 7.369 .1911 .02921
18 1.974 32.71 7.766 .1874 .02876
19 2.084 35.88 8.189 .1854 .02851
20 2.194 39.17 8.652 .1854 .02851
20.5 2.248 40.86 8.900 .1861 .02860
21 2.303 42.58 9.160 .1874 .02875
21.5 2.358 44.33 9.443 .1893 .02898
22 2.413 46.11 9.747 .1918 .02929

Table 20.5: Model propulsion test results

The additional towing force at that model speed becomes:

FD = 0.5 x 1000 x (2.36)2 x 0.00113 x 14.107 = 44.3N


At 21.5 knots the self propulsion point of ship is found from the intersection of the
curves of F,T and Q versus n at F = FD, as shown in Fig. 20.5.
The intersection results in n=9.44 rps , T=128.71 N, Q=5.822 Nm.

From Fig. 20.5 it can be seen that torque, thrust and towing force are quite linear
with the rps in the overload test. The slopes of these curves are also used to correct the
values of n,T and Q at other speeds, where only one rps is measured. The corrections
are made so that F = FD

In this way the results of the overload test are reduced to the results at the self
propulsion point of ship. The results for the measured speed range, interpolated at
intervals of one knot, is given in Table 20.5. The thrust and torque are given in non-
dimensional form as Kt and Kg.
The results of Table 20.5 are the same as when a propulsion test was carried out
with the proper additional towing force on the self propelled model.
330 G.Kuiper, Resistance and Propulsion, January 6, 1994

160

140

120

/00
10C1

80

60

40

20

5 6 7 a 9 10 11

Figure 20.5: Interpolation of Model Data from an Overload Test at 21.5


knots.

20.12.1 Comparison with Resistance Test.


From the thrust coefficients the full scale thrust can simply be calculated. At 21.5 knots
n, is found to be 9.443/ = 2.013. The diameter is 6.5 m. The density of sea water
is 1025. So:

T, = 1025 x 2.0132 x 6.54 x 0.1893 = 1404kN

The full scale thrust can be compared with the extrapolated resistance
from Table 7.5 and the thrust deduction factor t can be calculated. At 21.5 knots
January 6, 1994, The Propulsion Test 331

Vs n, n, T, R, t
knots Hz rpm kN kN
17 1.571 94.3 863 665 .230
18 1.656 99.3 940 737 .216
19 1.746 104.7 1034 819 .208
20 1.845 110.7 1154 915 .207
20.5 1.897 113.8 1226 971 .208
21 1.953 117.2 1307 1034 .209
21.5 2.013 120.8 1404 1109 .210
22 2.078 124.7 1516 1196 .211

Table 20.6: Calculation of thrust deduction

t =i 1109 = 0.210
1404
In Table 20.6 the results of the thrust deduction are given.
The torque to be delivered for this thrust and the rpm still are not
correct, because of the scale effects in the wake and on the propeller. To account
for these effects the Taylor wake fraction is first calculated from a comparison with the
open water test.

The results of the open water test are given in Fig. 16.1 and Table 20.3.
From the thrust coefficient at the self propulsion point of ship the advance ratio is
read from the open water diagram and the corresponding torque K qo is found. For 21.5
knots the thrust coefficient is 0.02898. From the open water diagram the advance ratio
J = 0.623 is found and Kqo = 0.03045. The relative rotative coefficient is calculated to
be

0.03045 (1.1
71" = 0.02898 =
The Taylor wake fraction wry, is found from the advance velocity

Ve = .632 x 9.443 x 0.2955 = 1.7635m/sec


at a model speed of 2.358 m/sec. So the wake fraction is
1.7635
Wrn = 1 = 0.252
2.358
The results of the calculation of the Taylor wake fraction are given in Table 20.7.
The values of the Taylor wake fraction can be used to correct for wake scale effects.
From experience the ratio
1 w,
1 = 1.055
is chosen. So (1 w,) = 1.055(1 0.252) = 0.789. At a ship speed of 21.5 knots the
entrance velocity of the propeller is thus
= 0.789V8 = 8.728m/sec.
332 G.Kuiper, Resistance and Propulsion, January 6, 1994

V, V, nit, Kt K1 Kqe J wt
17 1.865 7.369 .1911 .02921 .03067 .629 .266
18 1.974 7.766 .1874 .02876 .03023 .636 .261
19 2.084 8.189 .1854 .02851 .02998 .640 .257
20 2.194 8.652 .1854 .02851 .02998 .640 .254
20.5 2.248 8.900 .1861 .02860 .03007 .638 .253
21 2.303 9.160 .1874 .02875 .03022 .636 .253
21.5 2.358 9.443 .1893 .02898 .03045 .632 .252
22 2.413 9.747 .1918 .02929 .03076 .628 .251

Table 20.7: Calculation of Taylor wake fraction

J (Kt + AKt)/ J2
0.2 10.142
0.3 3.978
0.4 1.930
0.5 1.034
0.6 0.574
0.7 0.315
0.8 0.158
0.9 0.058

Table 20.8: Values of Kt/J2 corrected for full scale

Now the value of KT/J2 can be calculated:


KT,T, 1404.000
= 0.426
= PVeiD2 = 1025 x 8.7282 x 6.52
Interpolation in the open water diagram for full scale gives the advance ratio J,. The
full scale open water diagram can be derived from Table 16.1 by adding AKT = 0.0006
to the model thrust coefficients and by subtracting 0.0059 from the torque coefficients.
The full scale open water data from Table 16.1 are given in Table 20.8.
Interpolation at K/ J2 = 0.4258 in Table 20.8 gives an advance ratio J, = 0.657.
Together with the entrance velocity of the ship propeller of 8.728 m/sec the proper
propeller revolutions are found from
8.728
n, = = 2.04
0.657 x 6.5
Expressed in rpm the revolutions at 21.5 knots are therefore 123 rpm.

The torque coefficient at J, = 0.657 is read from Table 16.1 to be 0.0289-0.00059 =


0.0283. This is the open water torque coefficient at full scale. The torque coefficient
in behind condition is found by dividing this open water torque by the relative rotative
efficiency 1.051. This results in a torque coefficient of 0.0269. The power absorption at
21.5 knots it thus found from:
P = Kqpn3D527r = 0.0269 x 1025 x 2.0563 x 6.55 x 27r = 17469k W.
January 6, 1994, The Propulsion Test 333

V3 n, P, T lif no nis 17d


knots rpm kW kN
17 96.3 8610 863 1.050 .624 1.041 .682
18 101.6 9930 940 1.051 .628 1.051 .694
19 107.1 11541 1034 1.052 .630 1.056 .700
20 113.2 13613 1154 1.052 .630 1.055 .699
20.5 116.4 14860 1226 1.052 .629 1.052 .696
21 119.8 16282 1307 1.051 .628 1.050 .693
21.5 123.4 17977 1404 1.051 .626 1.047 .689
22 127.3 19975 1516 1.050 .623 1.047 .685

Table 20.9: Performance Prediction of Ro-Ro Ship

14 R F(T = 0)/R
knots N
17.02 90.78 1.046
18.51 105.53 1.030
20.01 123.23 1.026
21.51 146.10 1.029
23.01 179.23 1.034

Table 20.10: Resistance from overload test

The power to be delivered by the engine is found by taking the shaft losses into account.
Assuming the shaft losses to be 1 percent of the delivered power, the power to be delivered
at the shaft P, = 17469/0.99 = 17646 kW.
The results of the performance prediction for the ship, based on model test results,
is given in Table 20.9. The results in Table 20.4 are derived using the Marin method for
the restricted propulsion test, so slight differences occur.

20.13 Extrapolation of the Example using


the Marin Method.
From the overload test results in Table 20.4 the model resistance can be derived by
interpolation at T = 0, as illustrated in Fig. 20.5. The result is given in Table 20.10.
The resistance results from the overload test are approx. 3 percent higher than from
the resistance test. It is possible to take this into account, but it is not strictly necessary.
When the higher resistance is extrapolated with the lower thrust deduction factor the
prediction for full scale thrust will not be much different. In the following we will assume
that the resistance found from T = 0 is reduced with 3 percent, so that the resistance
extrapolation and the thrust deduction remains the same as in Table 20.6.

From the overload test the relation between thrust coefficient, torque coefficient and
advance ratio based on model speed can be found. The result can be found in Table 20.11.
These data are plotted in Fig. 20.6 on the basis of J,,.
334 G.Kuiper, Resistance and Propulsion, January 6, 1994

Vs Jt, Kt Kg
17.02 0.872 .1869 .02889
18.51 1.231 .0477 .01168
18.51 0.861 .1852 .02832
18.52 1.014 .1346 .02236
20.01 0.852 .1876 .02865
21.51 1.186 .0634 .01370
21.51 0.977 .1459 .02371
21.52 0.830 .1945 .02941
23.01 0.806 .2022 .03078

Table 20.11: Open water data simulation from overload test

0.3

0.2

Kr

0.1

o
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
J,Jv

Figure 20.6: Thrust Coefficient in Behind Condition.

At 21.5 knots the revolutions at the self propulsion point of ship has been calculated
in Table 20.5 to be 9.443 rps. The advance ratio J, in that condition is
January 6, 199.4, The Propulsion Test 335

2.358
(21.5knots)= = 0.845
9.443 x .2955
At full scale the incoming velocity in the propeller is higher than at model scale by a
factor 1.055. This is simulated by multiplying the advance ratio in Fig. 20.6 with 1.055.
This gives a new relation Kt Ji,. The curve or line thus obtained is shifted vertically
by an amount AKt to acount for viscous effects on the propeller. This line gives the
relation between Ktbi (behind condition, ship) at full scale and L at full scale.

Similar as with the ITTC extrapolation method the full scale advance ratio Jt, is
not known because the propeller revolutions are unknown. The correct advance ratio is
found from the full scale thrust, which is known from the extrapolation of the resistance
test in combination with the thrust deduction. At 21.5 knots the required thrust has
been calculated in Table 20.6 to be 1404 kn. At 21.5 knots or 11.073 m/sec the ratio
1404 x 103
Kt/.13 = = 0.2644
1025 x 11.0732 x 6.52
In Fig. 20.6 the ratio Ktb8/42 is plotted from the curve Ktbs and intersected at
0.2644. This gives for the ship <It, = 0.825. The revolutions are found from
11.073
ns = = 2.065
.825 x 6.5
This is 124 rpm, which is within the accuracy of the readings the same as the 123 rpm
found from the ITTC extrapolation method.

The torque for the ship is found from the relation between Kt, and Kgm as found
from the overload test in Table 20.4. The relation is plotted in Fig. 20.7.
The full scale thrust coefficient at 21.5 knots can be calculated from
1404 x 103
1<-13 = = 0.18
1025 x 2.0652 x 6.54
The relation plotted in Fig. 20.7 is the relation at model scale. Therefore the corre-
sponding torque coefficient is found at
= 0.18 0.0006 = 0.1794. 3 This gives Kg, = 0.02763. With the correction
AKg = 0.00059 the full scale torque coefficient is found to be Kg, = 0.02704. The
required power at 21.5 knots is therefore

Pd = 0.02704 x 1025 x 2.0653 x 6.55 x 2ir = 17797kW


With 1 percent shaft losses the shaft power to be delivered by the engine is:

P, = 17797/0.99 = 17977

3This accuracy cannot be read from the figure. It has been found from higher order
interpolation in a program. This shows that the accuracy of the extrapolation can be
limited by the reading of the diagrams.
336 G.Kuiper, Resistance and Propulsion, January 6, 1994

0.3

0.2

Korn

Q1

o
o 0.1 az 0.3
Km

Figure 20.7: Relation between Thrust and Torque Coefficient from an Over-
load Test.
Chapter 21
Propulsion Calculations.
Objective: An indication of available statistical formulas for the propul-
sion test parameters.

Following the sequence of this course the calculation of the wake frac-
tion, the thrust deduction, the relative rotative efficiency and the wake
scaling relations from statistical data and from flow calculations should still
be treated. The data in this field are, however, limited. Especially the
flow calculations are still in their infancy. Both the wake fraction with
propeller and the thrust deduction can in principle be calculated from full
Navier-Stokes solutions of the flow around the hull, where the propeller is
modeled as an actuator disk, generating a thrust force and eventually also a
circulation in the flow. These calculations have not been made on a regular
basis. When attempts are made, these are focussed on the effect of the pro-
peller on the wake distribution, more than on the prediction of the correct
average. These calculations at high (full scale) Reynolds numbers are not
yet feasible, so the scale effects on the wake or the full scale wake fraction
cannot yet be calculated . These topics are more suited for an advanced
course. Moreover, in this course only the mean wake is considered. So in
the following we will restrict ourselved to the statistical approximation of
the interaction parameters.

21.1 Statistical Prediction of the Model Wake


Fraction.
A very old but handsome approximation of the wake fraction was given by
Taylor in 1917:
w = 0.5CB 0.05

337
338 G.Kuiper, Resistance and Propulsion, January 7, 1994

This is of course based on very old hull forms. The prediction of the wake
fraction from 200 models towed at the Netherlands Ship Model Basin has
been attempted by Harvald in a thesis in 1950. He gave separate diagrams
for single screw and twin screw ships, as reproduced in Figures 21.1 and
21.2. In these figures 6 is the blockcoefficient CB and Psi is the wake
fraction. The wake fraction used by Harvald was the Taylor wake fraction
wT, which is also called the effective wake fraction. .

Nominal wake. factor


single screw silica
Add!for

0,40

t 0 1.:: 4.:
1_1111f:S4P:
_

owll I 1
IIIIIIIIIIIIIIIIIII i .
0,30
*AO L gnu
4

020
&0
1

Q55 --0,60 0,65 0,70 0,75

V
055 -060 --w 065 070 075
Correction in th wake- factor for shoot ci Ihe tram. sectionsl
0,05
a
U -Slue*

V- Sha..
0.05

Correction in the wake-factor


14
for the propeller diameter
0.05

0.05

.10 c03 404 405 C105 Ct07


1.0.0...,

Figure 21.1: Harvald's Diagram for the Determination of the Wake Fraction
for Single Screw Ships.

Regression formula's based on model test results were given by Holtrop


in 1977 [14]. For single screw ships he arrived at:

0.17774B2 7.65122
w, 0.577076 L 0.404422Cp 42,
(L L.C)2 p
in which
=
L(CtC)
January 7, 1994, Propulsion Calculations, 339

0,20

0.15
1/B -z. 6.5
11.1

0,10

0,05

o
0.55 0.60 0.65
cS

Figure 21.2: Harvald's Diagram for the Determination of the Wake Fraction
for Twin Screw Ships.

where Ct and C are the total and the wave resistance coefficients. The
difference is thus the residual resistance coefficient. For other parameters
see the next section.
For twin screw ships Holtrop arrived at:

w, = 0.4141383C/2, 0.2125848Cp 5.768516/.0,2


The model wake fraction requires an extrapolation to full scale. In 1977
Holtrop gave for single screw ships:

12
w, W3 = 7.65122(Ctm Ct3)(Ctm Cts 2Cw) D2

and for twin screw ships the same formula, but with a coefficient 5.769
instead of 7.65122. These formula's can also be used in the extrapolation
340 G.Kuiper, Resistance and Propulsion, January 7, 1994

of the propulsion test results.

Later Holtrop based his regression formula's for the wake fraction on
analyses of full scale trials and thus predicted the full scale wake directly.

21.2 Statistical Prediction of the Full Scale


Wake Fraction.
Based on full scale trial results in combination with extrapolated open water
diagrams Holtrop has made a regression formula for the wake fraction at
full scale [16]:

LB S T
ws= f(B'T'LD' LCB
D Cp Cm Cb C t ern/ Cv 7

in which S = wetted area,


L, B,T = length,breadth and draft,
D=propeller diameter,
LCB=distance of the center of buoyancy to the midship section, in percent
of the waterline length. Positive when the position is forward of the midship
section.
Cp=prismatic coefficient,
Cm= midship coefficient,
Csterm=a coefficient indicating the shape of the stern,
C, = (1 + k)Cf + Ca, the residual resistance coefficient.

For fast single screw ships with an open stern a simpler prediction for-
mula was sufficient:

tv, = 0.3CB + 10CCB 0.1


For twin screw ships a simpler formula could be given:

ws = 0.3095CB 10C,,CB 0.23D/v"BY

21.3 Statistical Prediction of the Thrust De-


duction.
The thrust deduction is even more susceptible to the shape of the afterbody
and the rudder configuration than the wake fraction. The reliability of sta-
tistical formula's is therefore even less. A rough approximation is again
January 7, 1994, Propulsion Calculations, 341

from Taylor (1917): t = 0.6w.

In [16] Holtrop gives a regression formula for the thust deduction

t = 0.25014()0.28956(VT-7"ID)0.2624 /(1 Cp+ 0.0225LCB)m1762+0.0015Cstern


L
Again for fast single screw ships with an open stern the thrust deduction
could be approximated by a single number: t = 0.10. For twin screw ships
this became:

t = 0.325CB 0.1885D/ViBT

21.4 Statistical Prediction of the Relative


Rotative Efficiency.
Again Holtrop in 1984 gave his latest regression formula as:
AE
71r = 0.9922 0.05908 -I- 0.07424(Cp 0.0225LCB)
Ao

For fast single screw ships with an open stern the relative rotative effi-
ciency can be estimated as 0.98. For twin screw ships this is [15]:

hr = 0.9737 -F 0.111(Cp 0.0225LCB) 0.063243-


342 G.Kuiper, Resistance and Propulsion, January 7, 1994
Chapter 22
TABLES

T Pv
C(Acius /V1rn2
0 608.012
2 706.078
4 813.951
6 932
8 1069
10 1226
12 1402
14 1598
15 1706
16 1814
18 2059
20 2334
22 2638
24 2981
26 3364
28 3785
30 4236
32 4756
34 5315
36 5943
38 6619
40 7375

Table 22.1: Vapor pressure of Water.

343
344 G.Kuiper, Resistance and Propulsion, January 7, 1994

Temp. kinem. visc. kinem. visc.


deg. C. fresh water salt water
m2I sec x 106 m21 sec x 106
0 1.78667 1.82844
1 1.72701 1.76915
2 1.67040 1.71306
3 1.61665 1.65988
4 1.56557 1.60940
5 1.51698 1.56142
6 1.47070 1.51584
7 1.42667 1.47242
8 1.38471 1.43102
9 1.34463 1.39152
10 1.30641 1.35383
11 1.26988 1.31773
12 1.23495 1.28324
13 1.20159 1.25028
14 1.16964 1.21862
15 1.13902 1.18831
16 1.10966 1.15916
17 1.08155 1.13125
18 1.05456 1.10438
19 1.02865 1.07854
20 1.00374 1.05372
91 0.97984 1.02981
99 0.95682 1.00678
23 0.93471 0.98457
94 0.91340 0.96315
25 0.89292 0.94252
26 0.87313 0.92255
27 0.85409 0.90331
28 0.83572 0.88470
29 0.81798 0.86671
30 0.80091 0.84931
Table 22.2: Kinematic viscosities adopted by the ITTC in 1963
January 7, 1994, Tables 345

Cf X 108
1 x 105 8.333
2 6.882
3 6.203
4 5.780
5 5.482
6 5.254
7 5.073
8 4.923
9 4.797
1 x 106 4.688
2 4.054
3 3.741
4 3.541
5 3.397
6 3.285
7 3.195
8 3.120
9 3.056
1 x 107 3.000
9 2.669
4 2.390
6 9.246
8 2.162
1 x 108 2.083
2 1.889
4 1.721
6 1.632
8 1.574
1 x 109 1.531
2 1.407
4 1.298
6 1.240
8 1.201
1 x 1010 1.17x
Table 22.3: Friction coefficients according to the ITTC57extrapolator.
346 G.Kuiper, Resistance and Propulsion, January 7, 1994

Temp. density density


deg. C. fresh water salt water
kg/m3 kg Irri3
0 999.8 1028.0
1 999.8 1027.9
2 999.9 1027.8
3 999.9 1027.8
4 999.9 1027.7
5 999.9 1027.6
6 999.9 1027.4
7 999.8 1027.3
8 999.8 1027.1
9 999.7 1027.0
10 999.6 1026.9
11 999.5 1026.7
12 999.4 1026.6
13 999.3 1026.3
14 999.1 1026.1
15 999.0 1025.9
16 998.9 1025.7
17 998.7 1025.4
18 998.5 1025.2
19 998.3 1025.0
20 998.1 1024.7
21 997.9 1024.4
22 997.7 1024.1
23 997.4 1023.8
24 997.2 1023.5
25 996.9 1023.2
26 996.7 1022.9
27 996.4 1022.6
28 996.2 1022.3
29 995.9 1022.0
30 995.6 1021.7
Table 22.4: Densities as adopted by the ITTC in 1963.
Bibliography
ABBOTT, I.H., von DOENHOFF, A.E., "Theory of Wing Sections",
Dover Publ., 1958.

BAKER, G.S., Kent, L., "Effect of Form and Size on the Resistance
of Ships" , Trans. R.I.N.A., 1913

BOWDEN,B.S., DAVIDSON, N.J., Resistance Increments due to Hull


Roughness Associated with Form Factor Extrapolation Methods", Nat.
Physical Laboratory, Techn. Memorandum No.380, 1974.

BRANDT, H., "Modellversuche mit Schiffspropellern an der Wasser-


oberflache", Schiff und Hafen, 1973.

CHUNG, Y.K., LIM, J.S., A Review of the Kelvin Ship Wave Pattern,
Journal of Ship Research, Vol.35, 1991.

DOREY, A.L., "High Speed Small Craft", 45th Parsons Memorial Lec-
ture, Trans. R.I.N.A.,1989

DURAND, W.F., "Aerodynamic Theory", Berlin, 1935.

GAWN, R.W.L., "Results of Experiments on Model Screw Propellers


with Wide Blades", Trans. Inst. of Naval Architects, Vol.79, 1937.

GERTLER, M., "A Reanalysis of the Original Test Data for the Taylor
Standard Series" , David Taylor Model Basin, Report 806, 1954.

GULDHAMMER, H.E., "Formdata", Danish Technical Press, 1965


and 1969.

HESS, J.L., SMITH, A.M.O., "Calculation of Non-Lifting Potential


Flow about Arbitrary Three-Dimensional Bodies", Jounal of Ship Re-
search, Vol.8, pp.22-44, 1964.

HESS, J.L., "Panel Methods in Computational Fluid Dynamics", An-


nual Review of Fluid Mechanics, Vol.22,1990

347
348 G.Kuiper, Resistance and Propulsion, January 7, 1994

HOERNER, S.F., "Fluid Dynamic Drag" ,published by the au-


thor,1971.

HOLTROP, J.,"A Statistical Analysis of Performance Results", Int.


Shipbuilding Progress, Vol. 24, 1977.

HOLTROP, J., MENNEN, G.G.J., "An Approximate Power Prediction


Method" , Int. Shipbuilding Progress, Vol.89, 1982.

HOLTROP, J. "A Statistical Reanalysis of Resistance and Propulsion


Data ", Int. Shipbuilding Progress, Vol.31, 1984.

Proceedings of the International Towing Tank Conference, issued every


three years after a conference meeting in various places in the world.

JONG, K. de ,"On the Optimization and the Design of Ship Screw


Propellers with and without End Plates", Thesis Groningen University,
The Netherlands,1991.

JONK, A., VOSSNACK, E., "The Use of Non-viscous Flow Calcula-


tions in Hull Form Optimization" , MARIN Workshop on Develop-
ments in Hull Form Design, 1985.

J.KATZ, A.PLOTKIN, "Low-Speed Aerodynamics", McGrawhill, 1991.

Publ. Kelvin 1904 and 1913.


KNAPP, DAILY, HAMMITT "Cavitation",McGraw-Hi11,1970.
KUIPER, G., "The Wageningen Propeller Series", Marin Pub!. No.
92-001, Wageningen, 1992.

LAP, A.J.W., "Diagrams for determining the Resistance of Single


Screw Ships ", Int. Shipbuilding Progress, 1954.
LERBS, H.W., "On the Effects of Scale and Roughness on Free Run-
ning Propellers", ASME J. Basic Eng., 1951.
LERBS, H.W., "Moderately Loaded Propellers with a Finite Num-
ber of Blades and an Arbitrary Distribution of Circulation", Trans.
S.N.A.M.E., Vol.60, 1952.

LEWIS, E.V., "Principles of Naval Architecture, Second Revision", Soc.


of Naval Architects and Marine Engineers, USA, 1988.
LIGHTHILL, J., "Waves in Fluids", Cambridge University Press,
1978.
January 7, 1994, Bibliography, 349

LINDGREN, H., Bjrne, E., "The SSPA Standard Propeller Family


Open Water Characteristics", SSPA Publ. No. 60, G6teborg, 1967.
MANEN, J.D. van, "Fundamentals of Ship Resistance and Propulsion,
Part B, Propulsion",Marin Publication No. 132a, 1952.
MAU Propeller Series. Various Authors in Japanese. E.g. YAMAZAKI,
S., " Open Water Test Series with Propeller Models of Small Blade Area
Ratio", Trans. West-Japan Soc. of Naval Architects in Japan, 1982.
McCARTY, "The roles of Transition, Laminar Separation and Turbu-
lence Stimulatioin in the Analysis of Axisymmetric Body Drag", Symp.
on Naval Hydrodynamics, London, 1976.

MORGAN, Wm.B., SILOVIC, V., DENNY, S.B.,"Propeller Lifting


Surface Corrections", Trans. S.N.A.M.E., 1968.
MORGAN, Wm.B., WRENCH, J.W., "Some Computational Aspects
of Propeller Design", Methods in Computational Physics, Vol.4, Acad-
emia Press, New York, 1965.

NEWMAN, J.N., Marine Hydrodynamics, The MIT press, Cambridge


(Mass), 1978.

OORTMERSSEN, G.VAN, "A Power Prediction Method and its Ap-


plication to Small Ships", Int. Shipbuilding Progress, 1971.
OOSSANEN, P.van, "Calculation of Performance and Cavitation
Characterisitcs of Propellers Including Effects of Non-Unzform Flow",
Thesis Technical Univ. Delft, Marin Publ. No. 457, 1974.
OOSSANEN, P.VAN, "Resistance Prediction of Small High Speed Dis-
placement Vessels , Int. Shipbuilding Progress, 1980.

OOSTERVELD, M.W.C., "Wake Adapted Ducted Propellers", Thesis


Technical University Delft, 1970.
RAVEN, H.C., "Variations on a Theme by Dawson", 17 th Symposium
on Naval Hydrodynamics, The Hague, 1988.
ROMSON, J.A., "Propeller Strength Calculation", The Marine Engi-
neer, 1952.

SCHLICHTING, H., Boundary-Layer Theory, McGrawHill, 6th edi-


tion, 1968.

TAYLOR, D.W., "The Speed and Power of Ships", New York, 1910.
350 G.Kuiper, Resistance and Propulsion, January 7, 1994

TODD, F.H., "Series 60, Methodical Experiments with Models of


Single-screw Merchant Ships" , David Taylor Model Basin, R and D
report 17, 1963.

VOSSNACK, E., VOOGD, A., "Developments of Ships Afterbodies",


Second Lips Propeller Symposium, Drunen, Holland, 1973.
YOUNG, R.T.,"Cavitation", McGrawhill, 1989.
Index
a-symmetrical duct, 39 camber line, 211
active rudder, 38 canard, 24
actuator disk, 201 carrier of a propeller blade, 222
additional resistance, 113 catamaran, 26
additional towing force, 312 cavitation, 29, 265
advance ratio, 224 cavitation bucket, 280
aerodynamic center, 260 cavitation inception, 265
agouti, 35 Cavitation tunnel, 284
air cushion vehicle, 24 chine, 21
angle of attack, 211 chord of a profile, 211
anti singing edge, 220 cloud cavitation, 270
appendage resistance, 134 conservation of momentum, 203
B-series, 228 containership, 14
back of a propeller, 208 continuity equation, 140, 202
behind condition, 315 contra rotating propellers, 34
Bernoulli, 149, 151 controllable pitch propeller, 33, 222
bilge radius, 71 cross flow, 66
bilge vortex, 69 curve of sectional areas, 123, 131
Biot and Savart, 154
blade area ratio, 233 d'Alembert, paradox of, 158
blade area ratio expanded, 218 deadrise, 20
blade area ratio projected, 218 deadweight, 13
block coefficient, 14 dilatation, 144
bollard condition, 227 dimension analysis, 98
bossing, 31 dipole, 155
bound vorticity, 290 dispersion, 82
boundary layer, 53, 165 displacement, 13
bow, 14 displacement thickness, 53
brackets, 31 diverging wave system, 85
bubble cavitation, 266 downwash, 290
bulb resistance, 133 drag coefficient, 51, 318
bulbous bow, 17, 94 duct, 37
bulbous stern, 77 dynamic free surface condition, 188
dynamic positioning, 31, 38
camber distribution, 211 dynamic viscosity, 101

351
352 G.Kuiper, Resistance and Propulsion, January 7, 1994

dynamical free surface condition, induced velocity, 204


173, 187, 189 induction factor, 295
economical speed, 91 Keller, 234
eddy viscosity, 171 Kelvin source, 190
effective camber, 304 Kelvin wave pattern, 84
effective wake fraction, 338 kinematical free surface condition,
efficiency, 205 173, 188
entrance velocity, 78 knots, 14
equivalent blade section, 318 Kutta condition, 247
equivalent roughness height, 115, Kutta-Joukowski, 159
321
erosion, 29, 275 laminar flow, 54
Euler equations, 148 Laplace, 152
expanded blade area ratio, 233 lenght of center of buoyancy, 124
extrapolation, 104 length between perpendiculars, 14
length of the waterline, 14
face of a propeller, 208 lift coefficient, 250
fin, 40 lift drag ratio, 255
fish propulsion, 47 lifting surface corrections, 303
Flettner rotor, 47, 159 loading coefficient, 204
form drag, 59 loading distribution, 249
form factor, 109 longitudinal center of buoyancy, 16
form resistance, 103 Ludwieg and Ginzel, 306
four quadrant diagrams, 243
free surface condition linearized, 189 magneto dynamic propulsion, 48
Michell's integral, 191
Froude number, 85, 101
momentum thickness, 54
generator line , 215
Navier Stokes, 146, 184
gravity waves, 82
Newtonian fluid, 145
group velocity, 83
nominal wake, 74
Hess and Smith, 175 nose tail line, 211
hollow, 89 nuclei, 265
horse shoe vortex, 71 open stern, 31
hull efficiency, 324 open water diagram, 224
hull speed, 93 open water efficiency, 227, 324
hump, 89 optimum radial loading distribu-
hump speed, 93 tion, 207
hydrofoil, 21 overlapping propeller, 34
overload test, 313
ideal efficiency, 205
induced advance coefficient, 306 paddle wheel, 43
induced drag, 290 palm of a propeller blade, 222
January 7, 1994, Index, 353

parent model, 126 separation, 67


PHV-cavitation, 269 service condition, 323
pi-theorem, 98 SES, 25
pitch angle, 213 shape factor, 54
pitch of a propeller, 213 sheet cavitation, 267
planing, 19 shockfree, 253
potential, 151 similarity, 102
pram hull, 70 similarity solutions, 169
pressure coefficient, 248 singularity, 156
pressure drag, 59 sink, 153
pressure gradient, 55 sinusoidal waves, 82
prismatic coefficient, 16, 131 skew, 216
propeller plane, 214 skew angle, 218
propulsion test, 311 skew induced rake, 216
propulsive efficiency, 225 slipstream, 201
pumpjet, 43 slot, 38
source, 152
radial loading distribution, 293 spindle axis, 222
rake, 215 spray rail, 21
rake backward, 216 stagnation point, 248
ramjet, 47 stall angle, 251
Rankine Source, 190 stern, 14
re-entrant jet, 271 suction peak, 252
rebound of a cavity, 276 supercavitation, 35
reference line (propeller), 214 surface piercing propeller, 34
regression analysis, 128 swath, 26
relative rotative efficiency, 315, 324 sweep of a propeller, 219
residuary resistance, 103, 106, 127,
128 tanker, 17
resistance, 103 Taylor, 233
restricted overload test, 325 Taylor series, 126
retrofit, 40 Taylor wake fraction, 315
Reynolds number, 101 thickness distribution, 211
right handed propeller, 208 thin ship theory, 190
roll-up, 292 thrust coefficient, 225
rotation, 144 thrust deduction, 230, 315
roughness, 321 thrust identity, 315
thruster, 31
sail, 45 tipplates, 36
scale effects, 102, 265, 317 Todd series 60, 126
scaling laws, 102 total efficiency, 324
self propulsion point of model, 311 trail condition, 323
self propulsion point of ship, 312 trailing vorticity, 290
354 G.Kuiper, Resistance and Propulsion, January 7, 1994

transom, 21
transom resistance, 133
transverse waves, 84
trim wedge, 21
tuft test, 72
turbulence stimulator, 116
turbulent flow, 54
U-frame, 69, 70, 76, 125
V-frame, 76, 125
vane wheel, 37
vibration, 29
viscous resistance, 103
Voight schneider propeller, 41
vortex, 153
vortex cavitation, 268
wake, 232
wake adapted propellers, 295
wake fraction, 65, 78
wake scan, 88
wall effects, 116
warped propeller, 219
wave making length, 92
wave resistance, 134
wave resistance coefficient, 88
weiss fogh propulsion, 47
wetted area, 134
wetted surface, 112
zero lift angle, 250
Appendix A
DICTIONARY

ENGLISH NEDERLANDS

accelerating duct versnellende straalbuis


acceleration versnelling
active rudder actief roer
actuator disk versnellende schijf
added resistance toegevoegde weerstand
additional towing force toegevoegde sleepkracht
advance ratio oortgangsgraad
aerodynamic center aerodynamich centrum
air cushion vehicle luchtkussenvaartuig
air cushion vehicle (acv) luchtkussenvaartuig
airfoil vleugelprofiel
angle of attack invalshoek
anti singing edge anti zingrand
appendage resistance aanhangsel weerstand
appendages aanhangsels
a-symmetrical duct a-symmetrische straalbuis
back of a propeller zuigzijde van een schroef
behind condition conditie achter het schip
bilge kim
bilge radius kimstraal
bilge vortex kimwervel
blade area bladoppervlak
blade area ratio bladoppervlak verhouding
blade root bladwortel
blade section bladdoorsnede
blockcoefficient blokcoefficient
bollard condition paaltoestand

355
356 G.Kuiper, Resistance and Propulsion, January 7, 1994

bossing asuithouder
bound vortex gebonden wervel
boundary condition randconditie
boundary layer grenslaag
bow boeg
bow thruster boegschroef
brackets asuithouders
breadth breedte
bubble cavitation bellencavitatie
bulbous bow bulbsteven
camber welving
camberline welvingslijn
catamaran catamaran
cavitation cavitatie
cavitation desinence verdwijnen van cavitatie
cavitation inception ontstaan van cavitatie
cavitation tunnel cavitatietunnel
center of buoyancy drukkingspunt
chine kniklijn
chord of a profile koorde van een profiel
cloud cavitation wolkencavitatie
compressibility samendrukbaarheid
conservation of momentum behoud van impuls
continuity equation continuiteitsvergelijking
contra rotating propellers tegengesteld draaiende schroeven
controllable pitch propeller verstelbare schroef
cros flow dwa,rsstroming
curve of sectional areas kromme van spantoppervlakken
deadrise vlaktilling
deadweight laadvermogen
decelerating duct vertragende straalbuis
density dichtheid
developed blade area ontwikkeld bladoppervlak
developed blade area ontwikkeld bladoppervlak
diffusion diffusie
dilatation vervorming
dimension analysis dimensieanalyse
dipole dipool
dispersion dispersie
displacement waterverplaatsing
displacement thickness verdringingsdikte
diverging waves divergerende golven
January 7, 1994, Dictionary, 357

draft (draught) diepgang


drag weerstand
drag coefficient weerstandscoefficient
draught (draft) weerst and
duct straalbuis
dynamic positioning dynamisch positioneren
dynamic viscosity dynamische viscositeit
economical speed economische snelheid
eddy viscosity turbulente viscositeit
efficiency efficientie
entrance velocity intree snelheid
equation of motion bewegingsvergelijking
erosion erosie
erosion erosie
expanded blade area uitgeslagen bladoppervlak
expanded blade area uitgeslagen bladoppervlak
extrapolation extrapolatie
face of a propeller drukzijde van een schroef
fin vin
finite differences eindige differenties
finite elements eindige elementen
fish propulsion visvoortstuwing
form drag vormweerst and
form factor vormfactor
form resistance vormweerst and
formfactor vormfactor
frame spant
free surface condition vrije oppervlakte voorwaarde
free vortex vrije wervel
friction coefficient wrijvingscoefficient
generator line generator
gravity zwaartekracht
gravity waves zwaartekrachtsgolven
group velocity groepssnelheid
heave stampen
horse shoe vortex hoefijzerwervel
hub naaf
hull romp
hull efficiency hull efficiency
hull speed rompsnelheid
hump hump
hydrofoil draagvleugel
358 G.Kuiper, Resistance and Propulsion, January 7, 1994

induced geinduceerd
inertia traagheid
kinematic viscosity kinematische viscositeit
knots knopen
laminar laminair
leading edge intredende kant
length lengte
lift lift
lift coefficient liftcoefficient
lifting line theory draaglijn theorie
lifting surface draagvlak
lifting surface theory draagvlak theorie
loading coefficient belastingscoefficient
loading distribution belastingsverdeling
midship grootspant
momentum impuls
momentum thickness impulsverliesdikte
newtonian fluid newtonse vloeistof
nominal wake nominale volgstroom
nose tail line neus staart lijn
nuclei kernen
open water diagram open water diagram
overlapping propeller overlappende schroef
overload test overbelastingsproef
paddle wheel rad
parent model moedermodel
perpendicular loodlijn
pitch spoed
pitch angle spoedhoek
planing planerend
plate line plaatlijn
potential flow potentiaal stroming
pram hull praamvorm
pressure coefficient drukcoefficient
pressure drag drukweerstand
pressure gradient drukgradient
prismatic coefficient prismatische coefficient
projected blade area geprojecteerd bladoppervlak
propeller schroef
propeller plane schroefvlak
propeller shaft schroefas
propulsion test voortstuwingsproef
January 7, 1994, Dictionary, 359

pump jet waterstraalvoortstuwing


radial radiaal
rake rake
ramjet ramjet
reference line referentie lijn
regression analysis regressie analyse
relative rotative efficiency relative rotative efficiency
residual resistance restweerst and
residuary resistance restweerst and
resistance weerst and
resistance test weerstandsproef
retrofit aanpassing achteraf
re-entrant jet re-entrant jet
right handed propeller rechtse schroef
roll slingeren
roll-up oprolling
rotation rotatie
roughness ruwheid
sail zeil
scale effect schaaleffect
scaling laws schaalregels
separation loslating
service condition diensttoestand
shape factor vorm factor
sheet cavitation vliescavitatie
shockfree stootvrij
similarity gelijkvormigheid
singularity singulariteit
sink put
sinusoidal wave sinus(vormige) golf
skew skew
slipstream volgstroom
slot opening
source bron
spindle axis vers telas
spray rail spray rail
stagnation point stuwpunt
stall overtrekken
stall angle over trekhoek
stem voorsteven
stern achtersteven
suction peak lage druk piek
360 G.Kuiper, Resistance and Propulsion, January 7, 1994

supercavitation supercavitatie
surface piercing propellers surface piercing propellers
surface tension oppervlaktespanning
swath swath
sweep of a propeller omhullende van een schroef
tangential tangentieel
tanker tankschip
thickness dikte
thin ship theory slanke schip theorie
thrust stuwkracht
thrust coefficient stuwkracht coefficient
thrust deduction zoggetal
thruster dwarsbuis
tip clearance vrijslag
tipplates tipplaten
torque koppel
torque coefficient koppelcoefficient
towing carriage sleepwagen
towing tank sleeptank
trailing edge uittredende kant
trailing vortex vrije wervel
trailing vorticity afgaande wervels
transition omslag
transom transom
transverse waves transversale golven
trial condition proeftochtcondi tie
trim trim
tuft test tuft test
turbulent turbulent
vane wheel vane wheel
vapor damp
vibration trilling
viscous resistance wrij vingsweerst and
vortex wervel
vortex roll-up oprolling van wervels
wake volgstroom
wake adapted propeller volgstroomschroef
wake fraction volgstroomgetal
wall effects wandeffecten
warped warped
waterline waterlijn
wave breaking golfbreking
January 7, 1994, Dictionary, 361

wave making length golfmakende weerstand


wave pattern golfpatroon
wave resistance golfweers t an d
wetted area nat oppervlak
wetted surface nat oppervlak
zero lift angle nullift hoek
Appendix B
WOORDENLIJST

NEDERLANDS ENGLISH

aanhangsel weerstand appendage resistance


aanhangsels appendages
aanpassing achteraf retrofit
achtersteven stern
actief roer active rudder
aerodynamich centrum aerodynamic center
afgaande wervels trailing vorticity
anti zingrand anti singing edge
asuithouder bossing
asuithouders brackets
a-symmetrische straalbuis a-symmetrical duct
behoud van impuls conservation of momentum
belastingscoefficient loading coefficient
belastingsverdeling loading distribution
bellencavitatie bubble cavitation
bewegingsvergelijking equation of motion
bladdoorsnede blade section
bladoppervlak blade area
bladoppervlak verhouding blade area ratio
bladwortel blade root
blokcoefficient blockcoefficient
boeg bow
boegschroef bow thruster
breedte breadth
bron source
bulbsteven bulbous bow
catamaran catamaran

362
January 7, 1994, Woordenlijst, 363

cavitatie cavitation
cavitatietunnel cavitation tunnel
conditie achter het schip behind condition
continuiteitsvergelijking continuity equation
damp vapor
dichtheid density
diensttoestand service condition
diepgang draft (draught)
diffusie diffusion
dikte thickness
dimensieanalyse dimension analysis
dipool dipole
dispersie dispersion
divergerende golven diverging waves
draaglijn theorie lifting line theory
draagvlak lifting surface
draagvlak theorie lifting surface theory
draagvleugel hydrofoil
drukcoefficient pressure coefficient
drukgradient pressure gradient
drukkingspunt center of buoyancy
drukweerstand pressure drag
drukzijde van een schroef face of a propeller
dwarsbuis thruster
dwarsstroming cros flow
dynamisch positioneren dynamic positioning
dynamische viscositeit dynamic viscosity
economische snelheid economical speed
efficientie efficiency
eindige differenties finite differences
eindige elementen finite elements
erosie erosion
extrapolatie extrapolation
gebonden wervel bound vortex
geinduceerd induced
gelijkvormigheid similarity
generator generator line
geprojecteerd bladoppervlak projected blade area
golfbreking wave breaking
golfmakende weerstand wave making length
golfpatroon wave pattern
golfweerst and wave resistance
364 G.Kuiper, Resistance and Propulsion, January 7, 1994

grenslaag boundary layer


groepssnelheid group velocity
grootspant midship
hoefijzerwervel horse shoe vortex
hull efficiency hull efficiency
hump hump
impuls momentum
impulsverliesdikte momentum thickness
intredende kant leading edge
intree snelheid entrance velocity
invalshoek angle of attack
kernen nuclei
kim bilge
kimstraal bilge radius
kimwervel bilge vortex
kinematische viscositeit kinematic viscosity
kniklijn chine
knopen knots
koorde van een profiel chord of a profile
koppel torque
koppelcoefficient torque coefficient
kromme van spantoppervlakken curve of sectional areas
laadvermogen deadweight
lage druk piek suction peak
laminair laminar
lengte length
lift lift
lift coefficient lift coefficient
loodlijn perpendicular
loslating separation
luchtkussenvaartuig air cushion vehicle (acv)
luchtkussenvaartuig air cushion vehicle
moedermodel parent model
naaf hub
nat oppervlak wetted surface
nat oppervlak wetted area
neus sta..art lijn nose tail line
newtonse vloeistof newtonian fluid
nominale volgstroom nominal wake
nullifthoek zero lift angle
omhullende van een schroef sweep of a propeller
omslag transition
January 7, 1994, Woordenlijst, 365

ontstaan van cavitatie cavitation inception


ontwikkeld bladoppervlak developed blade area
ontwikkeld bladoppervlak developed blade area
open water diagram open water diagram
opening slot
oppervlaktespanning surface tension
oprolling roll-up
oprolling van wervels vortex roll-up
overbelastingsproef overload test
overlappende schroef overlapping propeller
overtrekhoek stall angle
overtrekken stall
paaltoestand bollard condition
plaatlijn plate line
planerend planing
potentiaal stroming potential flow
praamvorm pram hull
prismatische coefficient prismatic coefficient
proeftochtconditie trial condition
put sink
rad paddle wheel
radiaal radial
rake rake
ramjet ramjet
randconditie boundary condition
rechtse schroef right handed propeller
referentie lijn reference line
regressie analyse regression analysis
relative rotative efficiency relative rotative efficiency
restweerstand residual resistance
restweerst and residuary resistance
re-entrant jet re-entrant jet
romp hull
rompsnelheid hull speed
rotatie rotation
ruwheid roughness
samendrukbaarheid compressibility
schaaleffect scale effect
schaalregels scaling laws
schroef propeller
schroefas propeller shaft
schroefvlak propeller plane
366 G.Kuiper, Resistance and Propulsion, January 7, 1994

singulariteit singularity
sinus(vormige) golf sinusoidal wave
skew skew
slanke schip theorie thin ship theory
sleeptank towing tank
sleepwagen towing carriage
slingeren roll
spant frame
spoed pitch
spoedhoek pitch angle
spray rail spray rail
stampen heave
stootvrij shockfree
straalbuis duct
stuwkracht thrust
stuwkracht coefficient thrust coefficient
stuwpunt stagnation point
supercavitatie supercavitation
surface piercing propellers surface piercing propellers
swath swath
tangentieel tangential
tankschip tanker
tegengesteld draaiende schroeven contra rotating propellers
tipplaten tipplates
toegevoegde sleepkracht additional towing force
toegevoegde weerstand added resistance
traagheid inertia
transom transom
transversale golven transverse waves
trilling vibration
trim trim
tuft test tuft test
turbulent turbulent
turbulente viscositeit eddy viscosity
uitgeslagen bladoppervlak expanded blade area
uitgeslagen bladoppervlak expanded blade area
uittredende kant trailing edge
vane wheel vane wheel
verdringingsdikte displacement thickness
verdwijnen van cavitatie cavitation desinence
versnellende schijf actuator disk
versnellende straalbuis accelerating duct
January 7, 1994, Woordenlijst, 367

versnelling acceleration
verstelas spindle axis
verstelbare schroef controllable pitch propeller (cpp)
vertragende straalbuis decelerating duct
vervorming dilatation
vin fin
visvoortstuwing fish propulsion
vlaktilling deadrise
vleugelprofiel airfoil
vliescavitatie sheet cavitation
volgst room wake
volgstroom slipstream
volgstroomgetal wake fraction
volgstroomschroef wake adapted propeller
voorsteven stem
voortgangsgraad advance ratio
voortstuwingsproef propulsion test
vorm factor shape factor
vormfactor formfactor
vormfactor form factor
vormweerst and form resistance
vormweerst and form drag
vrije oppervlakte voorwaarde free surface condition
vrije wervel free vortex
vrije wervel trailing vortex
vrijslag tip clearance
wandeffecten wall effects
warped warped
waterlijn waterline
waterstraalvoortstuwing pump jet
waterverplaatsing displacement
weerstand drag
weerst and draught (draft)
weerstand resistance
weerstandscoefficient drag coefficient
weerstandsproef resistance test
welving camber
welvingslijn camberline
wervel vortex
wolkencavitatie cloud cavitation
wrijvingscoefficient friction coefficient
wrijvingsweerstand viscous resistance
368 G.Kuiper, Resistance and Propulsion, January 7, 1994

zeil sail
zogget al thrust deduction
zuigzijde van een schroef back of a propeller
zwaartekracht gravity
zwaartekrachtsgolven gravity waves

Vous aimerez peut-être aussi