Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Resistance 2008 PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 213

_______________________________________________________________________________________________________________

FAKULTET FOR INGENIØRVITENSKAP OG TEKNOLOGI – NTNU TRONDHEIM


NORGES TEKNISK-NATURVITENSKAPELIGE UNIVERSITET

FACULTY OF ENGINEERING SCIENCE AND TECHNOLOGY – NTNU TRONDHEIM NORWEGIAN


UNIVERSITY OF SCIENCE AND TECHNOLOGY
_______________________________________________________________________________________________________________

UK-2008–80/IV

TMR4220 Naval Hydrodynamics

SHIP RESISTANCE

Knut Minsaas and Sverre Steen

DEPARTMENT OF MARINE TECHNOLOGY


Revised March 2008

MARINTEKNISK SENTER INSTITUTT FOR MARIN TEKNIKK


MARINE TECHNOLOGY CENTRE DEPARTMENT OF MARINE TECHNOLOGY
TRONDHEIM, NORWAY
TABLE OF CONTENTS

1 INTRODUCTION ....................................................................................................................................... 3

2 RESISTANCE AND RESISTANCE COMPONENTS ............................................................................ 7


2.1 INTRODUCTION ..................................................................................................................................... 7
2.2 MODELLING AND CALCULATION OF FRICTIONAL RESISTANCE ............................................................. 8
2.3 VISCOUS PRESSURE RESISTANCE ........................................................................................................ 23
2.3.1 Transom stern resistance .............................................................................................................. 25
2.4 VISCOUS RESISTANCE ......................................................................................................................... 27
2.4.1 Form factors.................................................................................................................................. 27
2.4.2 Form factor according to Hughes, Prohaska and Inui.................................................................. 31
2.4.3 Finding the form factor of appended models ................................................................................ 35
2.4.4 Empirical formulas for form factor ............................................................................................... 36
2.5 RESISTANCE DUE TO HULL ROUGHNESS ............................................................................................. 43
2.6 WAVE RESISTANCE............................................................................................................................. 50
2.6.1 The Kelvin Wave Pattern............................................................................................................... 52
2.6.2 Some considerations about wave resistance. ................................................................................ 54
2.6.3 Wave Pattern and FN ..................................................................................................................... 59
2.6.4 Wave Breaking Resistance ............................................................................................................ 64
2.7 AIR RESISTANCE ................................................................................................................................. 66
2.8 RESISTANCE DUE TO A TRANSOM STERN ............................................................................................ 69
2.9 APPENDAGE RESISTANCE ................................................................................................................... 71
2.9.1 Resistane of tunnel thruster openings............................................................................................ 73
2.9.2 Scaling of appendage resistance from model tests ........................................................................ 74
2.10 RESIDUAL AND TOTAL RESISTANCE ................................................................................................... 76
3 RESISTANCE PREDICTION ................................................................................................................. 79
3.1 INTRODUCTION ................................................................................................................................... 79
3.2 PREDICTING RESISTANCE FROM MODEL TESTS .................................................................................... 80
3.2.1 Example of Estimation of Ship Resistance based on Model Tests................................................. 84
3.3 EMPIRICAL RESISTANCE PREDICTION METHODS .................................................................................. 86
3.3.1 Calculation according to Hollenbach ........................................................................................... 88
3.3.2 Resistance calculated according to Holtrop's Method (1988) ...................................................... 94
3.4 NUMERICAL RESISTANCE PREDICTION METHODS ................................................................................ 99
3.4.1 Disturbance of the Free Surface due to Sinks and Sources......................................................... 100
3.4.2 Thin ship theory........................................................................................................................... 105
3.4.3 Estimation of Wave making Resistance by Panel Methods ......................................................... 107
3.5 A COMPARISON BETWEEN RESISTANCE CALCULATED BY THEORY, MODEL TESTS AND REGRESSION
ANALYSES ...................................................................................................................................................... 110
4 CORRELATION FACTORS FROM TRIAL ANALYSES ................................................................ 115
4.1 TRIAL ANALYSES.............................................................................................................................. 119
5 RESISTANCE IN SHALLOW WATER............................................................................................... 123
5.1 A CLOSER STUDY OF WAVE PATTERNS NEAR A SHIP IN SHALLOW WATER AT DIFFERENT VELOCITIES
129
6 DIRECT MEASUREMENT OF RESISTANCE COMPONENTS .................................................... 137
6.1.1 Measurement of the Wave Pattern and Estimation of the Wave Resistance................................ 145
7 METHODS FOR RESISTANCE REDUCTION.................................................................................. 149
7.1 INTRODUCTION ................................................................................................................................. 149
7.2 HOW TO REDUCE THE WAVE PATTERN RESISTANCE OF A SHIP?....................................................... 150
7.3 APPLICATION OF BULBS FOR REDUCTION OF WAVE RESISTANCE .................................................... 157
7.4 DEVELOPMENT OF ROUGHNESS AND FOULING IN SERVICE ............................................................... 164

1
8 MULTIHULLS ........................................................................................................................................ 169
8.1 STAGGERED HULLS .......................................................................................................................... 169
8.2 SINGLE HULL WITH OUTRIGGERS ..................................................................................................... 171
9 RESISTANCE OF CATAMARANS ..................................................................................................... 175
9.1 INCREASE IN RESISTANCE DUE TO ROUGHNESS ................................................................................ 177
9.2 RESISTANCE DUE TO THE TRANSOM ................................................................................................. 177
9.3 DETERMINATION OF THE FORM FACTOR ........................................................................................... 178
9.4 DECOMPOSING OF RESISTANCE IN PRACTICAL MODEL TESTING ...................................................... 179
9.5 CORRELATION FACTOR ..................................................................................................................... 179
9.6 PUBLISHED RESISTANCE DATA FOR CATAMARANS .......................................................................... 180
10 SWATH .................................................................................................................................................... 185
10.1 COMPONENTS OF RESISTANCE .......................................................................................................... 189
11 GENERAL SYMBOLS AND DEFINITIONS ...................................................................................... 199

12 REFERENCES ........................................................................................................................................ 205

13 INDEX ...................................................................................................................................................... 209

2
1 INTRODUCTION

This condensed representation of teaching material should not to be comprehended as a


complete textbook, but as a supplement to the lectures.

One of the main ideas in the lectures has been to show that analytical, experimental and
numerical methods are complementary and not alternatives in practical work.

Those who are looking for a deeper theoretical treatment of the material, may therefore not be
satisfied. On the contrary, others may find the material too theoretical.

Many formulas and much material has been included to make it possible to use the
supplement as a kind of "handbook".

It is important to recognize: The formulas are there to be understood, and not only to be
memorised.

The following sources, have frequently been applied in this supplement, and would be worth a
study:

Baba, E "Study on Separation of Ship Resistance Components", Mitsubishi Technical Report


No.59.

Baba, E. Mitsubishi Heavy Industries. "An Application of Wave Pattern Analysis to Ship
Form Improvement", Journal of Society of Naval Architects of Japan"

Bertram, V,"CFD for Ship and Offshore Design ",1999.

Havelock, T.H. "Wave Resistance Theory and its Application to Ship Problems", SNAME
Transactions, Vol.59.

Hoerner, S "Fluid-Dynamic Drag".1965.

Hoerner/Borst (1985), “Fluid Dynamic Lift”, Hoerner Fluid Dynmics Albuquerque.

Inui,T, Tokyo University "From Bulbous Bow to Free-Surface Shock Wave-Trend of Twenty
Years Research on Ship Waves at the Tokyo University Tank", 3rd George Weinblum
Memorial Lecture.

Inui, Kajitani, H and Miyata,H.”Wave making Resistance of Ships", SNAME


Transactions,Vol.70.

Inui, Tokyo Univerity,"Advances in Calculation of Wave-Making Resistance of Ships" The


Society of Naval Architects of Japan 60 th Anniversary Series.

Lammeren, Troost, Koning, Delft University of Technology "Resistance Propulsion and


Steering of Ships", The Technical Publishing Company H.Stam-Haarleem-Holland.

3
Lap, van Manen, NSMB, Delft University of Technology, "Fundamentals of Ship Resistance
and Propulsion".

Principles of Naval Architecture, Volume II, "Resistance, Propulsion and Vibration" Jersey
City NJ1988, The Society of Naval Architects and Marine Engineers"

Rawson-Tupper, Brunel University "Basic Ship Theory" Volume 2, Longman

Robb, AM, University of Glasgow "Theory of Naval Architecture" London Edinburgh

Sedov, L.I “Ploskie Zadachi Gidrodinamiki I Aerodnamiki, Gosudarstvennoe Izdatel’stvo


Techniko-Teoreticheskoi Literatury, Moscow-Leningrad,1950”.

Schlichting/Truckenbrodt, Technische Hochschule Braunschweig/Technische Hochschule


München."Aerodünamik des Flugzeuges.Grundlagen aus der Strömungsmechanik
Aerodynamik des Tragflügels" Springer -Verlag.

Sharma. Technische Hochschule Aachen "Jahrbuch der Schiffbautechnischen Gesellschaft


1962"

White, F.M."Fluid Mechanics"Mc.Graw-Hill

Wigley, C."The Theory of the Bulbous Bow and its Practical Application", Trans. NECI,
Vol.52

Wigley, C."Calculated and Measured Wave making Resistance for a Series of Forms defined
Algebraically, the Prismatic Coefficient and Angle of Entrance being varied Independently",
Trans.INA, Vol.84.

Zierep, J: Technische Hochschole-Karlsruhe, "Grundzüge der Strömungslehre."Berlin


Heidelberg, Springer Verlag

1.1 Dimensions, Units, Density for Water and Air

During the calculation of hydrodynamic forces, it is important to keep order on notions,


dimensions, symbols and units. In this chapter and in chapter 11 the most frequent notions and
symbols within resistance and propulsion been summarised.

The dimensions have been built around three basic units in the international SI (International
System of Units) system.

These are:

L = length, M = mass and T = time.

4
DIMENSION(F) DIMENSION(M)
DESIGNATION/ Based on force Based on mass UNITS
SYMBOL
F, L, T M, L, T
Length = L L L Meter, m
Force = F F L Newton, N
M 2
T

Mass = M F 2 M Kilogram, kg
T
L

Time = T T T Second, s
Velocity = V L L m/s
T T

Acceleration L L m/s2
2 2
T T

Pressure = p F M Pascal, Pa=N/m2


L
2
LT 2

Work = W FL 2 Joule, J=Ws=Nm


Energy = E M L2
T
Power = P L 2 Watt, W=N m/s
Energy flow = E F M L3
T T
Mass density = ρ 2
M kg/m3
F T4 3
L L

Mass flow = m T M kg/s


F
L T

Dyn. viscosity = µ T M Ns
F 2
Pas = 2
L LT m

Kinematic viscosity m2 /s

5
=ν L
2
L
2

T T

The notions weight (w) and specific weight (γ) follow from:

w = M · g and γ = ρ · g

where

g = acceleration of gravity = 9.81 m/s2 .

In ship technical calculations standard values are applied for density, salt water and fresh
water.

Fresh water, has the following density:

o
C 0 5 10 15 20 25 30
ρ (kg/m3) 999.8 999.9 999.6 999.0 998.1 996.9 995.6

The salinity varies from location to location. Normally the salinity varies between 3.4 and 3.5
%, with a mean value of 3.47 %.

With the assumption that the salinity is 3.5 %, the density is:

o
C 0 5 10 15 20 25 30
ρ (kg/m3) 1028.1 1027.7 1026.8 1025.9 1024.7 1023.3 1021.7

Air, has the following density at po=1 bar

o
C 0 10 20 40
ρ (kg/m3 ) 1.275 1.250 1.210 1.120

Model tests are normally done at oC = 15, and speed prognoses mainly done for oC = 15.

6
2 RESISTANCE AND RESISTANCE COMPONENTS

2.1 Introduction

The resistance of a body moving steadily through a fluid is the component of the total net
force in the direction opposite of the motion. The forces acting on the body can be divided
into pressure force – acting in a direction normal to the surface of the body, and shear force –
acting tangentially to the surface and in the direction of local relative fluid motion. The shear
force is caused by friction between the fluid and the body. The resistance due to this shear
force is termed viscous resistance.

On a very thin flat plate moving longitudinally in the fluid, the only resistance that can act is
the resistance due to friction. This makes it possible to measure the frictional resistance
separately from the pressure resistance. The frictional resistance of a ship is defined as the
resistance of a flat plate of the same speed, area and length as that of the ship. Thus, the
difference between viscous and frictional resistance is that viscous resistance includes three-
dimensional flow effects, while frictional resistance do not include such effects.

If the fluid is inviscid, there is no friction force, but there might still be a net pressure force.
The pressure resistance in inviscid fluid is due to wave making resistance. (Remember that a
body deeply immersed in inviscid flow has no resistance, according to d’Alemberts principle).
For a streamlined body at the free surface, like a ship hull, the pressure resistance is mainly
due to wave making. Thus, in ship resistance theory it is a common simplification to divide
the total resistance into wave making resistance and viscous resistance, and to assume that
these are independent of each other. As we shall see in this compendium, this is a rather
coarse simplification.

On a blunt body, like for instance a sphere, moving deeply submerged in viscous fluid, there
is a large pressure resistance. This is due to flow separation, which is caused by viscous flow
effects. Thus, this resistance is pressure resistance due to viscous flow effects, called viscous
pressure resistance. A ship will also have this kind of resistance, for instance on a submerged
transom stern, on appendages, and on the aft body of full ships.

In the following sections, we will discuss the various resistance components of ships, based
on the basic division into components as outlined above, but further subdivided into the
following components:

• Frictional resistance
• Viscous resistance
• Resistance due to hull roughness
• Transom stern resistance
• Wave Resistance
• Wave breaking resistance
• Air resistance

7
2.2 Modelling and Calculation of Frictional Resistance

Mainly from Sclichting / Truckenbrodt, Wieghart and White

The main point in this section is to show


• how the frictional resistance component can be determined if type of flow, velocity
and "characteristic" length are known

As mentioned, frictional resistance is the resistance due to friction on a flat plate aligned with
the flow. In this section we will show how the frictional resistance can be calculated, and list
the most commonly used empirical formulas for frictional resistance – the friction lines.

The shear force in the fluid is expressed as:


du
τ ( y) = µ (see Figure 2.1) (2.1)
dy

with the following expression for the velocity gradient in the boundary layer:
du V
=k (2.2)
dy L

where
L = Characteristic length (for example ship length)
V = Full velocity outside the boundary layer (for example ship velocity)
u = Velocity in the boundary layer at a given y value (distance from the wall)
k = Constant
µ = Cinematic viscosity of the fluid

The frictional force in the flow direction for a characteristic area then becomes:
V
FFR = k ⋅ µ ⋅ ⋅A (2.3)
L

Figure 2.1 Development of boundary layer on a flat plate

8
The ratio between frictional force and inertia force is:

V
k ⋅µ ⋅
⋅A
FFR L constant ⋅ν constant
= = = (2.4)
ρ
Fd C ⋅ ⋅ 2 ⋅ A L ⋅V RN
D V
2
where
V ⋅L
RN = (2.5)
ν
is the Reynolds number.

If the Reynolds number is equal for two models with equal geometry but with different scales,
this means that the ratio between frictional force and inertia force is identical for the models.
This is in principle a requirement for testing of scale models, but as we shall see, the
requirement is usually not fulfilled in testing of ships and marine structures. That is why
calculation of frictional resistance is important, also when resistance is found by model
testing.

Viscous flow can be either turbulent or laminar, and that difference has a major impact on the
resistance. The difference between laminar and turbulent flow is best illustrated by studying
Figure 2.2 from the tests of Reynolds. He released black ink into the flow at different
velocities, as illustrated. He also defined the parameter:
D
RND = VM ⋅ (2.6)
ν
where
D = Diameter of the tube
VM = Mean velocity in the tube

When RND <2000, the jet of ink is thin and regular as shown in Figure 2.2 (a). The flow is
identified as laminar when the particles are flowing parallel in the direction of the flow. For
RND between 2000 and 13000, the colour is spread as in Figure 2.1 (b). The flow is
recognised as turbulent after having passed a region of transition, where the ink does not
entirely fill the tube.

With a high-speed camera, a different picture than in b is obtained. This is shown in c. The
ink is distributed in small vortices and we can see that we have got small transverse motions
in the fluid.

For increasing velocity ( RN ) the point of transition is moving forwards. As shown in Figure
2.1, the same pattern as for the flow along a flat plate is obtained. After a laminar region
behind the leading edge, a transition region is observed, before the flow finally becomes
turbulent.

9
Figure 2.2 Reynolds pipe-flow experiment

The ratio between frictional force and the inertia force, is now expressed as:
FFR
CF = (2.7)
ρ
⋅V 2 ⋅ S
2
where the characteristic area S is the wetted surface.

10
For turbulent flow in ship hydrodynamics, the following empirical relation between C F and
R N is often applied:

0.075
CF = (2.8)
[log RN − 2]
2

This gives the following C F values, which may be compared to values in Figure 2.3:

log(R N ) 5.5 6 6.5 7 7.5 8 8.5 9


3
10 C F 6.12 4.69 3.70 3.00 2.48 2.08 1.78 1.53

For laminar flow, Blasius's equation is the most frequently used formula:
1.327
CF = (Curve 4 in Figure 2.3) (2.9)
RN

On the same figure, different C F curves applied before introduction of the ITTC-curve have
been traced.

Figure 2.3 Frictional resistance of laminar and turbulent flow over smooth flat plates

11
In the transition region between laminar and turbulent flow, the resistance has often been
expressed as:
0.075 A
CF = − (2.10)
[log RN − 2]
2
RN

where A is dependent of the turbulence level in the main flow which is defined as:

1 u '2 + v '2 + w '2


Tu = (2.11)
V 3

u’ = deviation from the mean velocity in the x direction


v’ = deviation from the mean velocity in the y direction
w’ = deviation from the mean velocity in the z direction
V = velocity in the direction of the main stream

The meaning of u '2 is "mean square" of the frequency. The mean is taken over a sufficient
long time in the x direction.

If the Reynolds numbers for the flow up to the point of transition is:

R NX = V ⋅ X TR (2.12)
ν
where X TR is the distance of the point of transition from the leading edge, the following A
values are obtained:

R NX 3⋅105 5⋅105 106 3⋅106


A 1050 1700 3300 8700

As has been observed, the flow may be either laminar or turbulent, or partly turbulent
depending on R N and the turbulence level in the fluid. This is the case both in full-scale and
in model- scale.

For a ship model, fully turbulent flow is required in order to get a flow that is similar to the
full scale flow. This is obtained by provoking the flow in such a way that it becomes turbulent
in spite of the Reynolds number being lower than the usual limit for transition to turbulence.
On ship models this is achieved by placing turbulence stimulators close to the bow.
Frequently used turbulence stimulators are studs, thread, or sand strip. The thread shall have
a diameter of about 1-1.5 mm, and being fastened to the hull surface normal to the flow, as
shown on Figure 2.5. It is only recommended for relatively slow models, due to the risk of
ventilation along the tread at higher speeds. The studs are 3 mm high with a diameter of 1.5
to 2 mm. They are placed in a vertical row with an internal distance of 10 mm between the
studs. By comparing Figure 2.4 with Figure 2.5 one can see how the resistance of a stream
lined body changes with application of turbulence stimulation in the critical (transitional)
region, which is 105 < R N < 107 . The resistance, which is randomly varying for critical
Reynolds numbers without stimulation, now follows the turbulent curve.

12
Figure 2.4 Resistance of a body of revolution without turbulence stimulation

Figure 2.5 Resistance of a body of revolution with turbulence stimulation

Froude proposed to divide the total resistance in two parts, one due to wave generation, which
manifests itself as pressure on the hull, and one due to shearing forces along the hull. The
difficult problem is to separate them. Froude measured the frictional resistance in towing tank

13
tests with plates having lengths between 3 and 15m.

Later, Schoenherr, Tiedemann, Kempf, Grunow-Schultz and Hughes made similar tests.
According to Froude, a "residual" resistance, which mainly is composed of wave making
resistance, is now determined by subtracting the frictional resistance RFM from the total
resistance:

RRM = RTM − RFM (2.13)

Le Besnerais summarised Froude's results and expressed the frictional resistance for a plank
at 15 0C degrees as:
γ
RF = ψ ⋅ ⋅ S ⋅ V 1.825 (2.14)
1000
where
γ = Specific mass (kg/m^3)
V(m/s) = Velocity of the plank
0.258
ψ = 0.1392 +
2.68 + L
L = Model or ship length in m
As already mentioned, different alternative friction formulas have been proposed, like the
Prandtl-Schlichting's (1955) formula:

0.455
CF = (2.15)
⎡⎣log ( RN ) ⎤⎦
2.58

ITTC (1957) (International Towing Tank Conference) summarised and analysed different
results and proposed a correlation line without claiming that this was an exact expression for
the frictional resistance. It was called a correlation line because it was tuned to give the best
match between model and full scale tests when extrapolation from model to full scale was
done without the use of a form factor. Compared to a “true” flat plate friction line, the
ITTC’57 correlation line includes a form factor of (1+k)= 1.1194 (this is the ratio between the
ITTC’57 correlation line and the Hughes line, which was supposed to be a purely flat plate
friction line). In spite of this, the ITTC’57-line is widely used as a flat plate friction line, both
for extrapolation of ship resistance with the use of a form factor, and for other purposes where
a flat plate friction line would have been more appropriate.

The ITTC’57 correlation line is given in equation (2.8) as:

0.075
CF = (2.8)
[ log( RN ) − 2]
2

In the following table, CF is calculated for different RN according to the Prandtl and the ITTC
formulas given above.

14
Estimation of frictional resistance based on change in momentum.

The velocity in the boundary layer increases with the distance from the hull and finally
becomes identical with the ship speed. The velocity distribution is a function of the
longitudinal pressure distribution and the local Reynolds number.

If the pressure increases in the direction of the flow, the velocity in the boundary layer
decreases. Finally, the velocity may even become negative and the flow unable to follow the
contour of the hull. This effect is called separation. Separation also occurs at discontinuities
on surface of the hull, like the crossing between the transom and the hull surface. When
separation occurs, backflow usually occurs.

We have seen how it is distinguished between laminar and turbulent flow. If the flow is
laminar, the elements in the fluid are parallel with each other.

Turbulent flow is regarded as "disturbed" flow where the elements also have transverse
motions. This was illustrated in Figure 2.2 (b) and (c) as well as in Figure 2.6.

The velocity distributions in turbulent and laminar flow are radically different. The
distribution of turbulent flow is "full" as shown in Figure 2.7.

Figure 2.6 Transition from laminar to turbulent flow

15
Figure 2.7 Laminar and turbulent boundary layer velocity distributions

Rotation

There is a distinction between different types of motion in turbulent flow like rotational
motion, translatory motion and deformation. A single fluid particle may be exposed to all
these types of motion. The conception of rotation is demonstrated in Figure 2.8.

It is interesting to know if the flow is exposed to rotation or not because it may be shown that
Bernoulli's equation is valid only when the flow is free from rotation.

An element of fluid (ABCD) in the flow has rotation with velocity components along the
sides as shown in Figure 2.8.

The integral along the contour is:

dΓ = ∫ v ds
s (2.16)

Finally:
∂v ∂u ∂v ∂u
d Γ ABCD = dxdy − dydx = dy ⋅ dx ⋅ ( − ) (2.17)
∂x ∂y ∂x ∂y

∂v ∂u
The last link ξ = − expresses the vorticity in the flow and is bound to the circulation
∂x ∂y

16
around the element of vorticity ABCD through d Γ ABCD = ξ dA .

Figure 2.8 Flow rotation

Figure 2.9 is an example of the importance of vorticity. The shear flow between the particles
in a fluid is as we have seen:

du
τ =µ⋅ (2.1)
dy

where µ is the dynamic viscosity.

If the flow is turbulent, we may according to Douglas/Gasiorek/Swaffield (1983), due to the


random motion of the particles, modify the viscosity in such a way that the apparent shear
flow is expressed like:
du
τ = (µ + ε )⋅ (2.18)
dy

Where ε is the eddy viscosity .

Sometimes ε becomes much larger than µ. The effect of the transverse motions in the
turbulent main flow is as follows:

The flow is composed of the mainstream u with a fluctuating component in the direction of
the mainstream u' and a component normal to the mean flow direction v'. This gives a shear
stress opposing motion, which is given by the product of fluid density and the average product
of the normal velocity fluctuations over an incremental time period:

τ ε = − ρ (u + u ')v ' = − ρ u ' v ' (2.19)

17
Figure 2.9 Eddy formation in the turbulent boundary layer

The shear flow at the wall opposing motion over ∂x is equal to the change in momentum and
is:

δ

∂x ∫0
τ0 = ρ ⋅ ⋅ u ⋅ (V − u ) ⋅ dy (2.20)

18
If momentum thickness is defined as:

δ
u u
θ =∫ ⋅ (1 − ) ⋅ dy , (2.21)
0
V V

the local frictional coefficient may now be expressed as:

τ0 dθ
Cf = = 2 ⋅ (2.22)
2 ρ ⋅V
1 2
dx

For the flow along a flat plate, a boundary layer as illustrated in Figure 2.1 is obtained.
Ideally, the boundary layer thickness is equal to the distance from the surface of the body and
to the point where the flow is undisturbed by the body. It is difficult to determine the distance
on this basis, because the velocity approaches the undisturbed velocity asymptotically. Thus,
the thickness of the boundary layer is defined as the distance where the local velocity is 99 %
of the full velocity (u V = 0.99) .

The velocity in the turbulent boundary layer is with acceptable accuracy approximated by (see
Figure 2.7):

u y 1
= ( )n (2.23)
V δ
where
n = Constant dependent of RN
y = Distance from the surface
δ = Boundary layer thickness
The boundary layer may be divided into an external and an internal part. In the internal part,
the viscous molecular shear forces are the dominating, and in the external forces due to
vorticity are dominating. Between these two regions, there is a transition region.

Applying Cole's wake parameter for the accurate velocity profile of a two-dimensional
turbulent boundary layer, the following relation between local frictional coefficient and
V ⋅δ
Reynolds number RNδ = is obtained:
ν
1

C f = 0.02 ⋅ RNδ 6
(2.24)

The relation is based on experimental tests.

With the velocity formula in equation (2.23) applied for the velocity distribution in the
boundary layer, we get:

θ n
= =χ (2.25)
δ 3n + n 2 + 2

19
With the definitions and symbols applied above, the local frictional coefficient becomes:


1
d ( χδ ) dR
C f = 0.02 ⋅ RNδ 6
= 2⋅ = 2 ⋅ χ ⋅ Nδ (2.26)
dx dRN x

In addition, the following solution is finally obtained:


1
1 1
− ⎡ n ⎤7 −
C f = 0.02 ⋅ RNδ 6
= 0.0377 ⋅ ⎢ ⎥ ⋅ R 7
(2.27)
+ 2
+
Nx
⎣ 3n n 2 ⎦
Where
V ⋅δ
RNδ =
υ

V ⋅x
RNx =
υ
This is the local frictional coefficient. The coefficient for the entire length of flow becomes:

1
1
⎡ ⎤7
L
7 n −
CF = ∫ C f ( x) ⋅ dx = ⋅ C f ( L) = 0.044065 ⋅ ⎢ ⎥⎦ ⋅ RN
7
(2.28)
x =0
6 ⎣ 3n + n 2
+ 2

Applying the velocity distribution given above, the boundary layer thickness for flow along a
plate without pressure gradient in the direction of the flow is:

6
⎡ 3n + n 2 + 2 ⎤ 7 −
1
δ = 0.022034 ⋅ ⎢ ⎥ ⋅ RN
7
⋅x (2.29)
⎣ n ⎦

Where x is the flow length.

It is now possible to calculate CF ⋅103 according to different n values and Reynolds numbers
and to summarise the calculations as follows:
log( RN ) 6 7 8 9

Prandtl/Schlichting 4.46 3.00 2.12 1.57


ITTC 4.69 3.00 2.08 1.53
n=5 4.52 3.25 2.34 1.68
n=7 4.39 3.15 2.27 1.64
n=9 4.28 3.08 2.22 1.59
n=10 4.24 3.05 2.19 1.58

20
See also Figure 2.3. It is obvious that the ITTC formula, Schoenherr's formula and Prandtl -
Schlichting's (1955) formula agree well with each other, but it has not been said that any of
these formulas are the accurate expressions for 2 dim turbulent flow.If it is assumed that the
ITTC (1957) and the Prandtl -Schlichting's formula estimate CF accurately, n varies between
five and nine when RN varies between 106 and 109 . This means that the velocity distribution
is "fuller" in full-scale than in model scale.

As an example:

Assuming that the ship speed is 35 knots and that the hull is slender, the boundary layers in
model and in full-scale for a ship of 100 m's length are:

Model Ship
N 5 9
L (m) 4 100
V (m/s) 3.6008 18.004
RN 1.263 ⋅107 1.513 ⋅109
δ L 0.01835 0.00919
δ (m) 0.0734 0.919
θ L (momentum thickness) 0.00305 0.00092

θ (m) 0.0122 0.092

It is observed that the boundary layer thickness relative to the ship length is approximately
twice as large on the model as on the ship. This has consequences for the velocity “seen” by
the propeller – the wake. This is discussed in the propeller compendium.

The apparent thickness of the hull becomes larger at the stern due to addition of the
displacement thickness to the geometrical thickness. In calculations made according to the
potential theory, this may be taken into account.

So far, we have considered the friction and boundary layer development on a flat plate. The
boundary layer development depends also on the pressure gradient in the direction of flow.
When analysing the effect of pressure gradient on the boundary layer development, the
displacement thickness, defined as:

δ
u δ
δ = ∫ (1 − ) ⋅ dy =
*
= displacement − thickness (2.30)
0
V 1+ n

is used together with the "shape factor". The shape factor H is:

n+2 δ*
H= = (2.31)
n θ

21
Karman developed an integral equation for the relation between shear forces, pressure and
velocities along a body with pressure variation in the flow direction. He applied Prandtl's
simplification of the Navier-Stokes equation for a flow, which does not vary with time and
arrived at the following expression:

dθ θ dU τ Cf
+ (2 + H ) ⋅ = = (2.32)
dx V dx ρV 2
2

2τ w
where the local frictional coefficient C f = , must not be mixed up with the total
ρV 2
frictional coefficient CF.

For a flat plate without separation and a high Reynolds number n>7 in equation (2.31), which
means that H<1.4. With a positive pressure, gradient n becomes smaller than 7.

Tests have shown that separation occurs when H>2.4 in turbulent flow, and at 3.5 when the
flow is laminar.

22
2.3 Viscous Pressure Resistance

The pressure force normal to the surface of the hull, decomposed and integrated in the
direction the ship is moving, is called pressure resistance. The wave resistance is for a surface
ship the dominating part of this resistance.

To consider the pressure resistance that is not due to wave making, we consider a body deeply
submerged in the fluid. D’Alembert’s paradox says that a body deeply submerged in inviscid
fluid has no resistance. Thus, the pressure resistance not due to wave making is related to
viscous flow effects, so we call it viscous pressure resistance. The boundary layer created by
the frictional forces changes the local velocity, and therefore also the pressure, creating a
viscous pressure resistance. However, for high Reynolds number flows, as we have in marine
hydrodynamics, the boundary layers are generally thin, so this effect is not very important.
The main sources of viscous pressure resistance in marine hydrodynamics are:

• Flow separaration
• Generation of vortices and circulation
Of these two sources, flow separation is the most important, if there is actually flow
separation present in the flow.

In Figure 2.10 the flow separation on a cylinder in turbulent and laminar flow is shown. It is
seen that the separation point, and thereby the flow pattern is quite different in laminar and
turbulent flow. This has a strong effect on the drag, as can be seen in Figure 2.11. How
important the viscous pressure resistance might be compared to the remaining viscous
resistance is illustrated in Figure 2.12.

Laminar separation (low Rn) Turbulent separation (high Rn)

Figure 2.10 Separation pattern on a cylinder in laminar and turbulent flow (White,
2005)

23
Figure 2.11 Pressure distribution around a cylinder

Figure 2.12 The importance of streamlining and flow separation on the drag (White,
2005)

24
2.3.1 Transom stern resistance

For a normal ship with partly submerged transom stern, the flow at the transom is separated,
and the flow is dominated by chaotic backflow (see Figure 2.13). This flow is similar to the
flow behind a projectile. The separated flow creates a drag force, called base drag. Based on
experiments with projectiles in air, the following empirical formula has been made:
0.029
CDB = (2.33)
C FB

where CDB and CFB are drag coefficients based on the transom area. The relation between CFB
and the normal friction line based on wetted surface is:
S
CFB = CF ⋅ (2.34)
SB

Where SB is the transom area. It is more practical to formulate the transom stern drag
coefficient using wetted surface area. The expression becomes:

SB 3
( )
DB S
CDB = = 0.029 ⋅ (2.35)
ρ CF
⋅V ⋅ S
2

2
where
SB = area of the transom
S = wetted surface of the body up to the transom (without transom area included).
The base drag should not be included in the form factor because of its special CF dependence.
On ships with a large transom and with low FN s, this component may be considerable.

Figure 2.13 Base drag (Przirembel)

25
The suction pressure at the transom arises as a result of the external flow, which carries away
the dead water behind the transom (see Figure 2.13). This suction effect decreases the
pressure on the transom. The boundary layer reduces the suction depending of how thick the
"coat" around the dead water is. A thick boundary layer means that the "coat" becomes
effective and the suction reduced. This means that the transom stern resistance should
increase with decreasing frictional resistance coefficient CF, which means it increases with
increasing Reynolds number. This effect is clearly seen from equation (2.35). The dependence
of CDB of Reynolds number is thus very different from the dependence of CF on RN. Thus, the
transom stern drag should be separated from the viscous or frictional resistance when the total
resistance is calculated, or scaled from model to full scale.

If the wetted surface is applied instead of the frontal area, the resistance becomes:

DB ( S B S )3
CDB = = 0.029 ⋅ (2.36)
ρ 2 CF
V S
2
where
S B = base, transom or frontal area
S = wetted surface in front of the transom (the area of the transom not included)

As already mentioned this component can not be included in the form factor due to the special
CF dependence and should be scaled separately. On ships with large transoms and low FNs,
this component may reach considerable values.
As mentioned, the formula for base drag presented here is derived for a “projectile-shaped”
body in infinite fluid. When applying this formula to surface ships, it should be noted that it
only applies as long as the speed is low enough for the transom stern to be wetted – that the
backflow in the separated flow zone fills the separated zone with dead water. For higher
speed, the transom will be dry, and the transom stern resistance can then no longer be
calculated using equation (2.36). How to calculate the drag of a dry transom is explained in
section 2.8.

26
2.4 Viscous Resistance

The main point in this section is


• to show how the ship shape and fullness influences on the "frictional resistance". It is
also shown how this effect is determined and taken into consideration in speed
prognoses based on model tests.
• and to show that this part of the total resistance on full ships is an important part of
the total resistance which should be reduced to a minimum.

The viscous resistance is somewhat larger than the frictional resistance and is proportional to
this. Viscous resistance includes three-dimensional flow effects on the frictional forces. Due
to the volume of the ship, the velocity along the hull is larger than the velocity of the ship.
This increases the viscous resistance compared to the resistance of a flat plate with the same
speed, length and wetted surface as the ship. This increase is the difference between viscous
and frictional resistance.

2.4.1 Form factors

Form factors are commonly used to express the relation between viscous resistance, CV, and
frictional resistance CF:
CV = (1 + k ) ⋅ CF (2.37)

Here k is the form factor. Note that often the value of 1+k is quoted as the value of the form
factor. If the form factor value is larger than 1, it is most likely because the value of 1+k is
given. The form factor might also be divided in separate components, for instance a form
factor caused by velocity increase due to displacement effects, and a form factor related to
viscous pressure resistance.

For a better understanding of the expression "viscous resistance" of a ship it is worth first to
study two idealised cases:

The viscous resistance of a foil in two-dimensional flow is:

RT
CV = = 2 ⋅ (1 + k0 + k1 ) ⋅ CF (2.38)
ρ
⋅V 2 ⋅ Ap
2
where
k0 = contribution to the frictional resistance, due to increase in velocity caused by
displacement of the flow or by thickness.
k1 = due to pressure resistance near the following edge, which is proportional to CF
CF = frictional coefficient of a flat plate.
Ap= projected area

27
S = wetted surface

tmax
Analyses of tests with thin profiles ( < 0.15 ), shows that a good approximation is:
c

t max
CV = 2 ⋅ (1+ 2 ⋅ )⋅ C F (2.39)
c
where
tmax = maximum profile thickness
c = length of the chord line
If experimental test results from tests in turbulent flow are plotted as a function of the
thickness, it is observed that CD increases linearly with the thickness up to approximately 20-
25 % of the thickness as shown in Figure 2.14. If the thickness increases beyond this value,
CD increases with a higher exponent. A closer analysis show that the flow does not follow the
contour of the profile, but separates over the last part of the profile length.

From the experiments, it is found that the drag in this region increases with:

t max 4
CV = 2 ⋅ 60 ⋅ ( ) ⋅ CF (2.40)
c

The contribution to total CV shown in equation (2.40) is due to the fact that the viscous flow
has changed the pressure distribution along the profile. Thus, this term is viscous pressure
resistance.

With reference to the expression for CV given in equation (2.38), the form factors will be:

tmax
k0 = 2 ⋅
c

tmax 4
k1 = 60 ( )
c

28
Figure 2.14 Viscous resistance coefficient of foil sections as function of the thickness to
chord ratio (Hoerner, 1965)

If the viscous resistance from tests with a cigar shaped body is studied (Figure 2.15), it is
observed that a similar effect as on profiles is present.

Based on the tests in turbulent flow the drag is expressed as:

⎡ 3
⎤ R
CD = CF ⋅ ⎢1+1.5 ⋅ ( D max ) 2 ⎥ = 1 T2 (2.41)
⎣ L ⎦ 2V S

where
Dmax = maximum diameter
CF = frictional coefficient for a flat plate with the same length L and velocity as the
ship.
S = wetted surface
Dmax
If experimental values of CD are plotted as function of it increases as for the profiles
L

29
Dmax
when > 0.2. This is illustrated in Figure 2.15 and indicates the presence of viscous
L
pressure resistance in addition to resistance due to “super velocity”. Analyses of test results
indicate that this resistance is a function of the viscous resistance in the forebody and up to the
point of separation and is approximated with:

∆CD = 7 ⋅ CF ⋅ ( D max )
3
(2.42)
L
The total resistance is then:

⎡ 3

CD = CF ⋅ ⎢1+1.5 ⋅ ( D max ) 2 + 7 ⋅ ( D max )3 ⎥ (2.43)
⎣ L L ⎦

Figure 2.15 Viscous resistance coefficient of bodies of rotation, as function of the


diameter/length ratio (Hoerner, 1965)

30
2.4.2 Form factor according to Hughes, Prohaska and Inui.

It is found that the wave resistance is approximately proportional to V 4 for typical surface
ships.

In ship resistance theory, two assumptions are often made.

1. That the wave resistance coefficient varies as the forth power of the speed.
2. That there is no change caused by viscosity in the wave resistance

It is therefore justified to express the wave resistance as:

CW = m ⋅ FN 4 (2.44)

where m is a constant

The first assumption is only true at fairly low speeds, when the wave resistance is small. As a
function of speed the wave resistance is not even monotonic, and it is possible that a formula
as mentioned above should contain an expression of the wave resistance with an algebraic
equation instead of the simple assumption of proportion to the forth power of the speed.

If the approximation given above is accepted, the specific model resistance is:

CT = CW + (1 + k ) ⋅ CF (2.45)

4
Let CW = mFN , then:

CT F 4
= 1+ k + m ⋅ N (2.46)
CF CF

CT
Values of therefore plot on a straight line with slope m and intercepting 1+k on the
CF
ordinate axis. Prohaska plotted the results of about 200 model experiments in this manner. For
C
FN between 0.1 and 0.2, T values for a great majority did plot on straight lines, enabling the
CF
determination of 1+k. However, for full forms, the points plotted on concave curves,
indicating that either 1+k or m is speed dependent. Some examples of Prohaska plots are
shown in Figure 2.16. It is seen that one might experiment with other exponents than 4 of the
FN, and choose the exponent that gives the best straight line fit, within the range
approximately 2-7.

CT
For twin-screw models tested with appendages the - values plot on convex curves. This is
CF
explained by the presence of a certain separation resistance on the appendages. Assume that
this resistance is approximated by:

31
CS = a + b ⋅ C F

where a and b are constants, then

CT a F 4
= (1 + k + b) + ( 4 + m) ⋅ N
CF FN CF

But the (apparent) form factor and the (apparent) wave resistance have increased. As the term
a
4
decreases with speed, the convexity of the curve is explained.
FN

Figure 2.16 Determination of form factors according to Prohaska

A different explanation to the convexity may be a rapid increase in wave resistance above the
power formula. Consequently, the correct assessment of m in such a case, should be the value
corresponding to the tangent at zero or low speed. When determining the form factor using
Prohaska’s method, results for Froude numbers in the range 0.1<FN<0.2 should be used.

32
An alternative is to assume that there is no wave resistance left at FN <0.1 and to let:

CT
= f ( FN = 0.1) = 1 + k (2.47)
CF

This procedure requires tests at lower speeds than the Prohaska method, so that it is more
prone to laminar flow over parts of the model hull, in which case our formula for CF is no
longer valid. Also, if the ship has a transom stern and/or a bulb close to the free surface, then
the there will be other resistance components than viscous resistance present at FN <0.1.
If the ship has a submerged transom stern, this can be corrected for by modifying equation
(2.46) into:

CT − CBD F 4
= 1+ k + m ⋅ N (2.48)
CF CF

Other known resistance components can also be subtracted in the same manner as CBD. The
same principle can be applied to equation (2.47).

As mentioned above, the exponent n is in the region 2 - 7, depending on ship shape, trim and
draught. It is assumed that the chosen n-value is suitable when a linear relation is obtained.
The slope of the curve gives the m value. Then k is calculated as illustrated in Figure 2.16.

It is also possible to determine the form factor through so called geosim tests where a
minimum of two models of different scales but equal shape are tested in the towing tank.

The form factor is now determined from:

CT 2 − CT 1
1 + k = 1 + k0 + k1 = (2.49)
CF 2 − CF 1

where
CT2 = coefficient of total resistance for model nr 2
CT1 = coefficient of total resistance for model nr 1
CF2 = coefficient of frictional resistance for model nr 2
CF1 = coefficient of frictional resistance for model nr 1

The wave resistance follows from:


CW = CT2 - (1+ k )·CF2

At least two models are required. Therefore, this procedure is expensive and not used in
commercial testing. However, in special cases it may give valuable information about the
scaling of the resistance components.

Figure 2.17 shows a plot of CT vs. CF used to determine the form factor from a series of
geosim model tests. Figure 2.18 shows how the form factor is independent of Froude number
for low speeds, while at high speeds there is a significant form factor dependence. In the same
figure, also the wave resistance is shown as a function of FN.

33
Figure 2.17 Determination of form factor from geosim model tests

34
Figure 2.18 Form factor values determined for different speeds from geosim model tests.
(Baba)

2.4.3 Finding the form factor of appended models

As mentioned in the previous section, it might be difficult to determine the form factor of
models with appendages. This can be due to laminar flow and transitional flow on the
appendages at low speeds. Here we will present a method to find the form factor of the
appended model when towing test results of the model without appendages are available. A
further treatment of appendage resistance and scaling of appendage resistance is given in
section 2.9.
The form factor for the appended model is determined by equating the values of wave
resistance in the range of Froude Numbers between 0.1 or 0.2 and 0.4. In this range it is
assumed that the appendages do not modify the wave resistance, and that the residual
resistance equals the wave resistance, so that:

RRMap = RRMnh

Where

RRM is model wave resistance, equaling the model residual resistance


Subscript ap means appended model
Subscript nh means naked hull

35
Thus:

RTMnh = (1 + knh ) ⋅ CFM ⋅ 12 ⋅ ρ M ⋅ S Mnh ⋅ VM2 + RRMnh

RTMap = (1 + k ap ) ⋅ C FM ⋅ 12 ⋅ ρ M ⋅ S Map ⋅ VM2 + RRMap

RTMap = (1 + kap ) ⋅ CFM ⋅ 12 ⋅ ρ M ⋅ S Map ⋅ VM2 + RTMnh − (1 + knh ) ⋅ CFM ⋅ 12 ⋅ ρ M ⋅ S Mnh ⋅ VM2

S Mnh S S
CTMap = (1 + kap ) ⋅ CFM + CTMnh ⋅ − (1 + knh ) ⋅ CFM ⋅ nh = (1 + kap ) ⋅ CFM + CRnh ⋅ Mnh
S Map S ap S Map

S Mnh
CTMap − CRnh ⋅
S Map
1 + kap = (2.50)
CFM

To find the wave resistance coefficient of the naked hull, CRnh, the form factor of the naked
hull must be determined experimentally using the method of Prohaska, as explained in the
previous section. Thus,
CRnh = CTMnh − (1 + knh ) ⋅ CFM (2.51)

2.4.4 Empirical formulas for form factor

If we for some reason not are able to determine the form factor from model tests with the
actual ship, it is possible to apply results from regression analyses of model test results.
Holtrop(1978-1984) published results of regression analyses of the model test results database
of the towing tank in Wageningen. He published a complete method for predicting the
resistance of ships, and that method included a formula to predict the form factor:

B 1.06806 T 0.46106 L 0.121563 L3 0.36486


k = − 0.93 +0.487118 ⋅ C14 ( ) ⋅( ) ⋅( ⋅( ) ⋅ (1 − CP )
-0.604247
) (2.52)
L L LR ∇

where
LCB
LR = L ⋅ (1 − CP +0.06 ⋅ CP ⋅ ) = distance from the bow to the parallel mid ship
4CP − 1
region
Lcb = longitudinal position of the centre of buoyancy in front of 0.5 L in percentage of
L.

The coefficient C14 is a function of the shape of the stern post and is expressed as:

C14 =1+ 0. 011Cstevn

36
According to Holtrop the following values should be used:

Shape of sternpost Cstern


Pram or gondola -25
V-shaped sections -10
Normal frame shape 0
U-shaped frame shape 10

Such formulas can never represent anything else than the average of the ships used in the
analyses. The form factor given in equation (2.52) is a prediction of the form factor that
would be found in a model test, using Prohaska’s method. As discussed, this is not necessarily
the “true” form factor. The measured form factor, using Prohaska’s method might include
resistance components that are not proportional to FN1, and should thus not be included in the
form factor as it is used in standard powering prediction methods.

Since the experimentally determined form factor might contain resistance components that are
not proportional to FN1, and since there is significant uncertainty to the measured form factor
value, some ship model tanks prefer to apply a form factor given by the main dimensions of
the ship instead of form factor determined from analyses of the towing tests with the actual
ship.

For this form factor, MARINTEK applies:

k = k0 + k1 = 0.6 ⋅ Φ + 145 ⋅ Φ 3.5 (2.53)

as an approximation. This gives a viscous resistance, which is close to the resistance of a body
of revolution, see Figure 2.19.

In the expression for the form factor in equation (2.53) the fullness parameter Φ** has been
applied. It is a result of a transformation of the ship to an equivalent body of rotation.

CB
Φ= ⋅ B ⋅ (TA + TF ) (2.54)
LWL

where
CB = block coefficient of the ship
TA = draught at the aft perpendicular
TF = draught at the forward perpendicular
B = maximum breadth of the hull
LWL = length of the waterline

37
Figure 2.19 Empirical expressions for form factors, as applied at MARINTEK.

**
The volume of the ship is:

∇ = 2 CB L B T

For the mid ship section the "equivalent diameter" is introduced:

4
D= 2 BT
π
Many ships may have the same D and L but different degrees of fullness

The following parameter is therefore introduced:

38
D
Φ =CPCB
L
With CP values in the region 0.85 - 0.90:

CB
Φ ≈ ⋅ 2 ⋅ B ⋅T
L

is an expression for the fullness of the ship.

In Figure 2.20, the form factor for different ships has been plotted as function of
∇ 500 2
3
= ⋅φ
( L 10) CB

As expected when L/T approaches 0, k approaches k = 0 as solution for the flat plate. If L/T
approaches ∞ , k will also approach ∞ because the ship then becomes broader with separation
and vortex generation therefore.

If the difference between form factors determined according to Inui/ Prokaska’s method and
k0 is large, it is likely that the pressure resistance is too high and that it still is space for
reduction of resistance through improvement of the lines.

In Figure 2.19 an occasional selection of form factors from model tests ( determined at FN =
0.10) plotted as a function of Φ. Where these values are high, it is assumed that the form
factor contains "false" viscous resistance in the form of induced resistance (or vortex
resistance), the components being strongly dependent of the shape of the frames.

There seems to be a lower limit for the measured form factor, which is assumed to represent
"an absolute minimum". This form factor may be approximated by:

k = k0 + k1 = 0.6 ⋅ φ + 11808 ⋅ φ 5.15 (2.55)

Because the ship is not a body of revolution, different types of separation may occur, as
indicated in the upper part of Figure 2.21

The vorticity behind the hull contains energy. With the horizontal and vertical velocities in a
plane behind the hull, the induced resistance may be expressed as:

ρ
Di = ⋅ ∫∫ (Vy 2 + Vz 2 ) ⋅ dy ⋅ dx (2.56)
2

Where Vy and Vz are velocities induced by the vortices in the y and z directions.

Figure 2.21 shows results from measurements of induced resistance behind the model. These
measurements show that the resistance is in the range of:

CDi = 0.1⋅10−3 − 10−3

39
Figure 2.20 Form factor dependence of L/T and fullness. (Inui, 1957)

It is important to observe how it depends on the radius of the transition between bottom and
sides and that it is important to have soft transitions. On the models in the lowest part of
Figure 2.22, the strength of the vortex has been measured at 4 positions behind the model for
the different models. It looks as if the relative intensity of the vortices increases with the size
of the model like the resistance due to the transom, which increased with decreasing CF
values.

It looks as if there is a given region:

∆k1 = 11808 ⋅ φ 5.15 − 145 ⋅ φ 3.5

where it is unclear how much of the resistance is viscous, and how much is pressure
resistance.

When k1 > 11808φ 5.15 it is reason to assume that the pressure resistance becomes increasingly
dominating.

When the main dimensions have been fixed, changes in the shape of frames, after body and
fore body is the only way to reduce the resistance. Model tests and advanced CFD
(computational fluid dynamic) calculations are therefore required.

40
Figure 2.21 Induced resistance due to vorticity

41
Figure 2.22 Different types of vortex generation. (Lugt)

42
2.5 Resistance due to Hull Roughness

The main intention with this section is

• to show the importance of roughness on the frictional resistance, and that for full
scale ships the roughness replaces the Reynolds number as the variable that
determines the magnitude of the frictional resistance coefficient.

The roughness of a surface increases the frictional resistance. In order to investigate the
influence of roughness, Nikuradse covered different plates with sand roughness as indicated
in Figure 2.23. He determined the frictional resistance of flat plates with sand roughness of
different magnitude, characterised by the average roughness height ks. Based on his tests,
Prandtl-Schlichting expressed the resistance of rough plates with the formula:

L
C F = (1.89 +1.62 ⋅ log
-2.5
) (2.57)
ks

It is observed that in equation (2.57), CF is a function only on the non-dimensional roughness


height L/ks, not a function of the Renyolds number, as we are used to.

Figure 2.23 The increase of frictional resistance coefficient with roughness

43
Usually, the increase of frictional resistance coefficient due to roughness is denoted ∆CF. If
the relative increase of frictional resistance, expressed by the ratio:
CF + ∆CF
CF

is studied for different values of ks/L, results as shown in Figure 2.23 and Figure 2.24 are
obtained. For other types of roughness than the sand roughness, we may in praxis apply the
so-called "equivalent sand roughness" as shown in Figure 2.23, where the constant k in
diagram is used in stead of ks.

The roughness of the hull is not only a function of the type of coating that is used, but also a
function of the amount of rust, fractures in the coating, fouling and traces from machine work.
The typical roughness of a modern hull is different from the roughness shown in Figure 2.23.
It has softer wave peaks.

It is seen from Figure 2.23 that the total frictional resistance CF+∆CF becomes practically
independent of the Reynolds number above a certain critical Reynolds number, and that the
value of this critical Reyolds number decreases with increasing roughness. Furthermore,
CF+∆CF increases with the roughness height.

It is worth to observe that it for a given roughness and velocity, is possible to extrapolate the
increase in resistance due to roughness with the following formula:

⎡ log ( RNm ) - 2 ⎤
4

∆CFs ( RNs ) = ∆CFm ( RNm ) ⋅ ⎢ ⎥ (2.58)


⎣ log ( RNs ) - 2 ⎦

Figure 2.24 Increase of frictional resistance due to roughness

44
The increase due to roughness can be measured on castings of the surface in a wind tunnel, in
a cavitation tunnel or in the towing tank. Having the formula shown above in mind, the
measured increase due to roughness can be extrapolated to full scale. RNm in equation (2.58)
is not the RN of the ship model, but it is the RN which is valid for the casting.

Roughness allowance for ships is, as mentioned above, based on measurements of resistance
in a towing tank, wind tunnel or cavitation tunnel with castings of different hull surfaces.
Results from such measurements may be approximated by the formula:

∆CF = ⎡110 ⋅ ( H ⋅ V ) − 403⎤ ⋅ CF 2


0.21
(2.59)
⎣ ⎦
where
V = ship velocity in m/s
H (10-3 mm) = mean distance between the highest ” mountain top” and the deepest
"valley" in a "landscape" over a length of 50 mm. If ∆CF according to the formula is
<0, then ∆CF =0.
When predicting the resistance of a new ship, the roughness is normally not known, and a
roughness of H=150 is used as a standard value. Equation (2.59) is not really non-
dimensional, and is clearly a purely empirical formula. Equation (2.59) is the standard
roughness correction in use at MARINTEK.

0.003
Smooth
H=75 my
F+ ∆ C F

0.0025 H=150 my
H=300 my
Total frictional coefficient C

0.002

0.0015

0.001

0.0005

0
1.0E+07 1.0E+08 1.0E+09 1.0E+10
Reynolds number R N

Figure 2.25 Total frictional coefficient found by use of the ITTC'57 line and equation
(2.59)

45
In Figure 2.25, the total frictional coefficient, including roughness allowance according to
equation (2.59) is shown. It is seen that above a certain Reynolds number, the total frictional
resistance coefficient is no longer dependent on the Reynolds number, but only on the
roughness. It should also be noted that the choice of roughness allowance is more important
for the full scale resistance than the choice of friction line.

An alternative formulation for the roughness allowance, which was proposed by Townsin and
is currently recommended by ITTC for ship resistance predictions, is:

⎡⎛ H ⋅10−6 ⎞ 13 ⎤
−1
∆CF = 0.044 ⎢⎜ ⎟ − 10 ⋅ R 3⎥
+ 0.000125 (2.60)
⎢⎝ LWL ⎠ ⎥
N

⎣ ⎦

This roughness allowance is compared with the roughness allowance in use at MARINTEK in
Figure 2.26. Note that previously it was customary to include the roughness allowance in the
correlation allowance, resulting in a combined correction factor that could be called roughness
allowance or correlation allowance. This should be kept in mind when values of either
roughness allowance or correlation allowance are stated in reports or scientific papers.

0.003

Smooth

0.0025 Townsin
Total frictional coefficient C F+∆CF

MARINTEK

0.002

0.0015

0.001

0.0005

0
1.0E+07 1.0E+08 1.0E+09 1.0E+10
Reynolds number RN

Figure 2.26 Comparison of total frictional resistance including roughness, comparing


the formula in use at MARINTEK (equation (2.59)) and the formula proposed by
Townsin and recommended by ITTC (equation (2.60))

46
After some time in service, the roughness may under special conditions reach values being
many times as large as 150. It is therefore important to choose optimum docking intervals for
maintenance of the surface. This is further discussed in section 7.4.

The recent ban by IMO of TBT-type antifoulings has lead to a rapid development of
antifouling paint technology. Especially the development of so-called foul release paint
systems might have significant impact on the roughness, both on new ships and on the
development of roughness in service. The foul release paints keep marine growth off by being
so slippery that as long as the ship is in service at speeds above approximately 15 knots, the
growth will not be able to withstand the forces from the water flow. The surface of the foul
release paints have typically a lower roughness and a very different roughness texture. As
indicated from Figure 2.23, the type of roughness has significant impact on the roughness
allowance.

47
Figure 2.27 Wave making of ships modelled with sink-source distributions (Horn,
Lammeren, Troost)

48
Figure 2.28 Wave generation (Wigley)

49
2.6 Wave Resistance

(From Lunde, (1951), Havelock, (1951), Inui et al (1962), Baba, (1969 -1972) Wigley, (1935-
1942) and Bertram (1999).

The main intention in this section is

• to show that the wave resistance next to the frictional resistance is the most
important components of the total resistance and to show how it is identified.

The wave resistance is an important component of the total resistance. It is a function of the
ship shape, expressed by both main parameters and other hull shape parameters. In the
following, we have tried to show how wave resistance is generated and how it is possible to
change the resistance by changing the shape of the hull.

The volume displacement changes the velocity along the hull in a similar way as the thickness
changes the velocity on the surface of a foil. At the fore body of the hull, the water is forced
outwards. In the after body, it flows back towards the centre line of the ship. This can be
simulated mathematically by placing so-called sources in the fore body and sinks in the after
body as shown in Figure 2.27.

According to Bernoulli's equation, a pressure distribution along the hull as indicated in Figure
2.27 is created. The free surface is deformed and the deformation is a function of the
waterline shape. Waves are generated near bow, stern and "shoulders" as indicated in Figure
2.27 and Figure 2.28.

The change in surface level is easy to understand from the following relationship:

ρ
⋅q + ρ ⋅ g ⋅ z + p = constant
2
(2.61)
2
where
z = change in surface level
q = resulting velocity along the hull
p=pressure at the surface
When the flow is approaching the bow, it is retarded. According to Bernoulli's equation, this
should give an increase in pressure, but since the surface pressure only is equal to the
atmospheric pressure, this is compensated for by an increase in z, which is identical to an
increase in elevation of the free surface.

Near the middle of the ship, the velocity is at a maximum. Without the free surface, this
would correspond to a decrease in pressure. However, with atmospheric pressure at the free
surface, a reduction in z is required in order to keep the energy constant along the streamline.

This change of surface level is indicated in Figure 2.28 and Figure 2.33. The change follows
the ship and gives waves near the bow and near the stern. Waves are also generated near the

50
shoulders if these are "sharp" as shown in Figure 2.28. In reality, waves are formed along the
entire hull, but they are only “visible” where the changes in the curvature of the surface are
"sharp".

The waves consist of two types, namely waves due to the local disturbance following the ship
and the free waves. The waves due to the local disturbance are non-oscillatory and decrease
rapidly as they move further away from the ship. They are called the inner wave system, or
Bernoulli waves, since they are related to the pressure-velocity distribution around the hull.

The free wave pattern consists of oscillatory waves, which are not present ahead of the ship
but only in its rear and over an endless stretch of distance (if viscous dissipation is excluded).

The total amount of energy transferred to the water increases steadily with time. The waves
related to the local disturbance maintain a constant amount of energy concentrated in the
ship's proximity, which means that the wave related to wave making resistance is not caused
by the inner (Bernoulli) wave system, but only the free-wave pattern.

Figure 2.29 Picture of the wave pattern of two sailing boats (Wikipedia)

51
2.6.1 The Kelvin Wave Pattern

In Figure 2.29 the wave patterns of two sailing boats are shown. When seen from the boat,
these wave patterns appear to be steady, which means that the wave pattern has the same
speed as the boat. It is also seen that the wave pattern is composed of waves with different
angles. The angle between the wave direction and the direction of travel is here denoted θ, see
Figure 2.30.

Figure 2.30 Definition of wave direction

Havelock (1951) took note of the fact that at a given θ-value, a one dimensional, sinusoidal
wave was travelling at an angle θ away from the x axes with a propagation speed Vcosθ, see
Figure 2.31. He called the one-dimensional waves just mentioned "elementary waves" and
showed that the so called Kelvin wave group is produced as a result of the mutual interference
of several such waves in the range

π π
− <θ <
2 2
We cannot distinguish each elementary wave by visual observation, but we can see the Kelvin
wave group , which comes out as a final result of the mutual interference of the elementary
waves.

The Kelvin wave group, illustrated in Figure 2.31 has two wave systems; transverse waves
and diverging waves, with angles in the range:
Transverse waves: 0 < θ < 35°
Diverging waves: 35° < θ < 90°
The diverging waves have an angle of close to 90° at the centreline. Towards where the
transverse waves meet the diverging waves, the angle decreases towards 35°. The transverse
waves has an agle of 0° at the centreline, increasing to 35° where they meet the diverging
waves. Where the transverse and diverging waves meet, they both have an angle of 35°, but
are ¼ wave length out of phase relative to each other. The common wave front formed by the
diverging and transverse waves has an angle of 19°28’ relative to the direction of travel. This

52
is often called the Kelvin angle.

The phase velocity Vp of each wave component must be the same as the speed U of the ship,
since we know that the wave pattern is steady seen from the ship. Thus, the relation between
angle and wave length (or frequency) is given as:

ω
Vp = = U cos θ (2.62)
k

Thus, the larger the angle θ, the shorter the wave length. This is observed in the wave pattern,
where the transverse waves are longer than the diverging waves. Since the diverging waves
are short, they are often steep, and often the most visible part of the wave pattern. This
doesn’t mean that they are the most important for the wave resistance. Especially for
conventional ships, the transverse waves often give the most important contribution to the
resistance. The resistance of the two wave systems for a typical ship, determined using
numerical calculations, is shown in Figure 2.32.

Figure 2.31 The Kelvin wave group

53
Figure 2.32 Contribution to the wave resistance from transverse and divergent waves
(Kelvin)

2.6.2 Some considerations about wave resistance.

Mainly from Horn and Taylor

In accordance with the expression given above, it is assumed that the wave resistance is
proportional to the square of the amplitude. It is further assumed that the bow and the stern
are each forming a wave system disturbing the surface with the wave amplitude ζ . Since the
wave resistance increases with the square of the amplitudes, an important task is to relate
these amplitudes to the shape and the velocity:

Assume that the surface is disturbed by two points with a distance approximately a = β⋅L =
0.9 L from each other. These points are the origin of two wave systems, simple amplitudes
forming the resulting wave.

Lord Kelvin's wave pattern is in our discussion simplified by omitting the diverging waves.
We do also assume that the transverse waves from the stern and bow are trochoidal waves.

54
Trochoidal waves

In wave theory, we do normally use sinusoidal waves. The trochoidal wave is somewhat sharp in the crest and
λ
flat in the trough like a simple wave in a model tank. The wave is reasonable realistic for waves up to about
20
in height. Other characteristics of the trochoidal wave, such as velocity, period, and pressure change with depth,
are the same as for the simple harmonic wave. However, it fails to meet certain requirements of classical
hydrodynamics and cannot be derived from the velocity potential. The only reason why it is used here is that it
has a convenient form from the geometrical point of view and its resemblance with the tank wave.

The trochoidal wave is described by a point on the radius of a circle as the circle rolls along the underside of a
horizontal straight line. If R is the radius of the rolling circle and λ is the wave length from crest to crest

λ = 2π ⋅ R .
The parametric equations of the trochoid are:

x = R ⋅ ω t + r ⋅ sin ω t
z = R + r ⋅ cos ω t
The radii of the circles in which the particles move decrease exponentially with depth; as in the case of the
harmonic wave

r = r0 e kz

where r0 is the radius of a particle at the surface and k = = the wave number.
λ
z is measured to the centre of the circle in which the particle moves.

The total energy in the trochoidal wave over one wavelength is approximated by:

1
Etot = ⋅ ρ ⋅ g ⋅ λ ⋅ b ⋅ H 2
8
This is identical to the expression for the sinusoidal wave. As for the sinusoidal wave, the wavelength is:

2π 2
λ= ⋅c
g
where

c = the speed of the wave (or the moving pressure point)

55
It is assumed that the bow and stern waves are trochoidal waves, which belong to two separate
wave systems with equal wave length λ. If a is the distance between a wave crest belonging to
the first wave system and a wave crest of the second wave system, the wave height of the
resultant system is:

2π a
H 2 = H12 + H 2 2 + 2 ⋅ H1 ⋅ H 2 ⋅ cos (2.63)
λ

H1 = wave height of the first system


H2 = wave height of the second system
It is easy to see that the resultant wave height is at a maximum (H1 + H2) or a minimum (H1 -
H2), for the following values of a:

a = n⋅λ
1
a = (n − ) ⋅ λ
2

n is any whole number. Thus, the maximum resultant wave height occurs when a wave crest
of one system coincides with a crest of the other, and the minimum resultant wave height
when a crest coincides with a hollow (trough).

We have assumed that the bow-wave system begins with a crest, the stern-wave system with a
hollow. The distance between the crest of the bow-wave and the hollow of the stern-wave is
called the wave-making length. This length is half a wavelength smaller than the distance
between the crests of the bow and stern waves, hence:

λ
β ⋅L = a− (2.64)
2

2π ⋅ V 2
Substituting in equation (2.63), using λ = we get:
g

2πβ L
H 2 = H12 + H 2 2 − 2 H1 ⋅ H 2 ⋅ cos (2.65)
λ

Above it has been assumed that the waves have equal breadths. In the place where the stern
wave has this breadth, the bow-wave system is already much broader and correspondingly
lower. Its height H1 in the resulting wave system is decreased to:

k ⋅ H1 (k < 1)

The energy in the transverse wave system is proportional to the wavelength λ, wave breadth b
and to the square of the wave height H. Thus the total energy over b is:

Etot = C ′ ⋅ b ⋅ H 2 ⋅ λ (2.66)

56
where C ′ is a constant

Applying potential theory, a first approximation for the wave heights is:

H1 = A⋅ V2

H2 = B⋅ V2

where A and B are constants, depending upon the ship shape.

If the ship is moving a distance x, the following energy is supplied over the same length:

x
E = Etot ⋅ (2.67)
λ

This gives as a rough approximation for the resistance over a breadth b due to the transverse
waves:

Rw ⋅ x = E = C ′ ⋅ b ⋅ H 2 ⋅ x (2.68)

The breadth b is proportional to the wave length, which is proportional to the square of the
V
velocity V ( FN = ). This gives:
gL

Rw = C ′ ⋅ V 2 ⋅ H 2 (2.69)

Based on the considerations above, the wave resistance due to the transverse waves is
therefore roughly approximated by:

RW ⎛ β ⎞
CW = =C⋅ V 2
⋅ ⎜ A 2 + B 2 − 2 ⋅ k ⋅ A ⋅ B ⋅ c o s( ) ⎟ (2.70)
ρ ⎝ FN
2

⋅V 2 ⋅ S
2

C = constant depending upon the shape of the ship


S = wetted surface
k=constant due to reduced wave height of the bow wave behind the stern
If A and B were constants, the wave-making resistance would vary as V4, the specific
resistance as V2. Bow and stern, however, are by no means concentrated pressure points, and
consequently A and B vary in a rather pronounced degree with the form of the wave pattern
and thus with speed. Nevertheless, the general trend of the wave-making resistance curve
shows a certain similarity to a parabola of the fourth degree.

k is not a constant either, but increases rapidly with speed. Consequently, the same is true for
the interference phenomena. If this expression is presented as a function of FN, a similar trend
as shown in Figure 2.32 is obtained.

57
Figure 2.33 Generation of wave resistance, showing how the interference between the
bow and stern wave systems create humps and hollows in the wave resistance curve.

58
2.6.3 Wave Pattern and FN

It has been shown how FN is an important parameter. It is therefore interesting to study this
closer. For simplicity, it is assumed that the waves generated by the bow and the stern are two
dimensional, as assumed above.

On a normal ship, the distance between the crest of the bow wave and the hollow of the stern
wave is given almost entirely by the geometry of the ship, and is approximately of the order:

b = β ⋅L (2.71)

where
L = ship length in the waterline
β = 0. 9 to 1. 0
These waves have the same velocity as the ship and consequently identical wavelengths.

The relation between wave length (distance from wave crest to wave crest in the same wave)
λ and the ship velocity is:

λ⋅g
V = (2.72)

If the trough of the bow wave and the trough of the stern wave are corresponding, the distance
b (the wave-making length) is related to the velocity by:

b⋅ g
V= (2.73)
n ⋅π
where
n = 1, 3 ,5,-----
The wave crest from the bow wave and the trough from the stern wave correspond for the
following n values:

n = 2, 4, 6, ------
This is illustrated in Figure 2.33, and shows that it is possible to reduce the amplitude of the
resulting wave by phasing the bow and stern wave in a correct manner (letting the wave crest
from the bow wave cancel the trough from the stern wave). Manipulating the expression given
above, we get:

V β
FN= = (2.74)
g⋅L n ⋅π

According to this expression for FN, a minimum or a maximum in wave resistance is obtained.
Whether it is a maximum or a minimum depends on n. If n is an odd number, the wave

59
resistance has a maximum. If the bow wave gives a trough or a crest near the stern, the
following is obtained for β = 0.9:

N=1 FN =0.54 a trough and maximum wave resistance


N=2 FN = 0.38 a peak and minimum wave resistance
N=3 FN =0.31 a trough and maximum wave resistance
N=4 FN = 0.27 a peak and minimum wave resistance

Without interference between the waves, the resistance should have increased with the square
of the velocity. The interference effect adds to the "basic" resistance, gives humps and
hollows (troughs) in the resulting resistance curve, as indicated in Figure 2.33. According to
this relation, it is natural to apply the Froude number and to draw the resistance
dimensionless, as showed in Figure 2.32. The challenge for the designer is to avoid the
resistance peaks, where these are marked. In reality this is a question of choosing the
economical speed for a given ship length.

Already in 1932 Weinblum made a theoretical calculation of the wave resistance for a number
of ship shapes and compared the results with model test results for the same shapes.
Moreover, Graff, Kracht and Weinblum (1966) made model tests with a systematic series of
models where the breadth and draught are kept constant, while the prismatic coefficient is
varied. The results are shown in . It is observed that the residual resistance reaches its
maximum at FN=0.5 for all models. It is also observed that the wave resistance has a steep
increase when FN>0.25. and obtained similar relations as discussed above.

Examples of typical dimensions, speeds and Froude numbers for various types of ships are
shown in the table below.

60
Figure 2.34 Residual resistance from a systematic model test series (Graff, Kracht,
Weinblum, 1966)

For the designer it would now be natural to ask:


How does the distance between fore and after body influence on the final value of the
wave resistance?

May after body and forebody be optimised independently with regard to wave
resistance?

Which is contributing most to the wave resistance, after body or fore body?

In order to answer these questions, Baba (1969, 1972) made measurements and calculations of
wave resistance on ship models with simple generic shapes, with circular bow and stern, and
parallel mid-body. The results are shown in Figure 2.35. Note that the wave resistance
coefficient is expressed using B2 instead of wetted surface for non-dimensionalisation.

The length was varied while the breadth; after body and fore body remained unchanged. The
tests showed that the wave resistance for L/B > 5.5 was no longer influenced by the ship
length. This is roughly also the case for the elliptical cylinders and for bodies having seem-
circular bow and stern, see Figure 2.35 and Figure 2.36.

61
Figure 2.35 Wave resistance of models with semi-circular bow and stern, and different
lengths (Baba)

Finally, the length of the fore body was kept unchanged, while the after body was changed as
shown in Figure 2.37. It is seen that in this case, the change of aft body shape didn’t influence
the wave making resistance. The changes in the after body were marked and manifested
themselves in different form factor values. It should be noted that for other ship geometries
and other types of aft body changes, the wave making resistance might indeed change. The
most notable example of this might be the introduction of large overhung flat sterns on large
cruiseliners and containervessels.

These results illustrate three important points during generation of wave resistance:

• After and fore body may be optimised separately, respectively for viscous resistance
and wave making resistance.
• The wave resistance is strongly influenced by the shape of the fore body, while
changes in the after body have considerably influence on the viscous resistance.
L
• For a given ship the wave resistance does not change with the length if > 5.5
B

62
Figure 2.36 Influence of L/B-ratio on the wave resistance coefficient of an elliptical
cylinder (Baba)

Figure 2.37 Effect on the wave resistance of changing the aft body shape (Baba)

63
2.6.4 Wave Breaking Resistance
If the velocity near full bows is reduced over a region, wave elevation is also observed in the
region. If the elevation of the wave becomes large enough, the wave will break. In this case,
one part of the wave energy will not spread out as longitudinal and transverse waves.
However, a belt is generated around the ship as shown in Figure 2.39.

Breaking waves are also from time to time observed behind the transom stern as shown in
Figure 2.38. According to this, the wave resistance may now be divided into:

• Wave breaking resistance RWB


• Wave pattern resistance RWP

It is not simple to separate these two components. In a simplified manner, one may regard the
wave breaking resistance as a part of the wave resistance but it is linked to the energy required
to form the foam belt observed in the wake of ships with full bow. Outside this belt, there is a
wave system similar to the wave system of slender ships without wave breaking as indicated
in Figure 2.39. The energy in the external system is linked to what we normally call wave
pattern resistance.

Figure 2.39 show how large the different resistance components are, compared to the total
resistance for a typical tanker. The actual tanker was equipped with two different bows. It is
indicated that the bulb in ballast reduces the wave breaking resistance.

Figure 2.38 Resistance components when the stern wave is breaking (Baba)

64
Figure 2.39 Wave breaking resistance

65
2.7 Air Resistance

The main intention in this section is

• to show that the air resistance in some cases is an essential component which we
should try to keep at a minimum

Air resistance is resistance due to the flow around the superstructure of the ship. The
resistance is normally found using drag coefficients based on earlier wind tunnel tests for
similar ships. If the air resistance is very important, or the requirement for accuracy is
particularly high, the air resistance should be determined from wind tunnel tests. The tests
may even be made in a cavitation tunnel with a double model of the superstructure.

The air resistance is expressed as:

ρ air
RAA = Cairp ⋅ ⋅ V 2 ⋅ Ap (2.75)
2
where
Cairp = resistance coefficient for the superstructure (0.5 - 0.8).
AP = projected area of the super structure in the longitudinal direction
V = wind velocity

The air resistance may on high speed craft be as high as 10 % of the total resistance.

The expression given above is often transformed in such a way that the wetted surface S is
used instead of the projected surface Ap , as for the other coefficients used to express the ship
resistance:

ρ air A
C AA = ⋅ C airp ⋅ p (2.76)
ρ S

The following values are typical mean values for Cairp:

CAR FERRY 0.55


CONTAINER SHIP 0.55
DESTROYER 0.60
DIVING SHIP 0.60
FERRY 0.45
FISHING BOAT 0.70
FRAIGTHER 0.55-0.65

66
PASSENGER SHIP 0.40
RESEARCH VESSEL 0.40
SUPPLY SHIP 0.55
TANKER 0.75-0.90
TENDER 0.55

The air resistance is calculated applying the ship speed. However, the velocity of the wind is
included in the calculations. In analyses of trials, we have to consider the wind velocity some
times. The difference in estimation of air resistance and wind resistance is exemplified in the
following:

Assume that the ship has a projected cross section area of 100 m2. With Cairp = 0.65 and a
total efficiency of ηD = 0.70 the power requirement due to air resistance is:

VS(knots 15 25 30 35
PD(kW) 25 115 200 316
∆VS = 5 m/s 68 222 359 516
∆VS = 10 m/s 131 364 540 765
∆VS = 15 m/s 216 541 774 1062

If the ship first is moving with the wind, which has a velocity of 5 m/s and then is moving
against the wind, it is observed that the loss against the wind is not by the gain when the ship
is mowing with the wind. This is clear from the following table. The reason is evident that the
resistance due to the wind is proportional to the square of the velocity.

VS(knots 30 35 40 45
PD(kW) 200 316 471 672
PD with the wind 349 516 729 993
PD against the wind 91 165 270 413
PD mean 220 341 500 703

It is normal that model tests are made with a projected surface, which is smaller than on the
ship. Often the model is placed below the carriage where the air velocity becomes larger than
the velocity of the carriage. In such cases, the air velocity below the carriage Vair should be
measured. If that is done RAAm can be calculated according to:

2
1 ⎛V ⎞
RAAm = ⋅ ρ air ⋅ ⎜ air ⎟ ⋅ V 2 ⋅ APm ⋅ Cairp (2.77)
2 ⎝ V ⎠

The ratio between Vair and V may reach as high values as 1. 25. It is therefore important with
this correction before CR is calculated and important to model the superstructure correctly.
The best solution is to place the model well in front of the carriage to avoid the disturbances.

67
On hydrofoil craft, the hull may become an essential lift and a moment in addition to the air
resistance. To be able to estimate the required control forces it is necessary to know how these
forces influence on the craft. In such cases, the forces are often measured separately during a
towing test or in a wind tunnel. An alternative is CFD (computational fluid dynamic)
calculations of the air forces.

The air velocity is independent of the height z above the still water level. However, the wind
velocity varies significantly with height due to the boundary layer induced by viscosity. The
speed at a height z above the still water level can be related to the speed at 10 m height by the
relationhip:

1
⎛ z ⎞ 8.85
U ( z ) = U ( z = 10 ) ⋅ ⎜ ⎟ (2.78)
⎝ 10 ⎠

If we assume that the wind velocity is 10 m/s at z = 10m , the velocity distribution within the
boundary layer of the wind is:

z 0.5 1.5 2.5 3.5 4.5 5.5 6.5 7.5 8.5 9.5
V ( z) 7.1 8.1 8.6 8.9 9.1 9.3 9.5 9.7 9.8 9.9

The mean value of the wind velocity becomes: Vmean=9.00 m/s. The “effective” wind velocity
for different maximum heights of the superstructure becomes:

hmax (m) 10 20 30 40
Ve (m/s) 9.00 9.50 9.67 9.75

The resistance due to air and wind now becomes:


2
ρ ⎡ ⎛V ⎞ ⎤
Rair + wind = ⋅ C AA ⋅ S ⋅ ⎢V ± ⎜ e ⎟ ⋅ Vwind ⎥ (2.79)
2 ⎣ ⎝ 10 ⎠ ⎦
where

ρ air Ap
C AA = ⋅ Cairp ⋅ (2.80)
ρ S

68
2.8 Resistance due to a Transom Stern

The main intention with this section is

• to show how the resistance due to the transom is defined

For typical displacement ships with submerged transom sterns, model tests indicate that the
transom is wetted for FN < 0.20. If the FN becomes larger, the water separates at the crossing
between the transom, the sides and the bottom. In this case, the pressure against the transom is
equal to the atmospheric pressure. The hydrostatic pressure that acts at all other part of the
hull below the waterline is “lost”, and this loss of hydrostatic pressure is considered as a
resistance force. The resistance of the ventilated transom is then:

H
Rtransom = ρ ⋅ g ⋅ ∫ z ⋅ B( z ) ⋅ dz (2.81)
0

Where

H = maximum depth of the transom


B(z) = the width of the transom at different draughts

The resistance can be expressed on dimensionless form as:

H
ρ ⋅ g ⋅ ∫ z ⋅ B( z ) ⋅ dz H
g z H
2 ∫
Ctransom = 0
= 2⋅ ⋅ B( z ) ⋅ ⋅ dz = 2 ⋅ g ⋅ k ⋅ 2 (2.82)
1
2 ρ ⋅ S ⋅V 2 V 0 S V

Ctransom is then dependent of the constant k, which is dependent of the transom stern
shape.
It is observed that the transom resistance is scaled like the wave resistance and that it is
natural to regard it as a part of the residual resistance. This is reasonable since the hydrostatic
force in this case is linked to a disturbance of the surface.

When the transom is still wetted, the so-called base drag due to the suction at the transom is
registered. The resistance of a wetted transom stern is discussed in section 2.4. It can be
expressed as:

(S S)
32
R
CDB = 1 transom = 0.029 ⋅ transom (2.83)
2 ρV S
2
CF

where
Stransom = transom stern area
S = wetted surface of the "ship" in front of the transom

69
CF = frictional coefficient of the hull

It has been indicated that the transom becomes dry for FN>0.2. This is a typical speed, but it
depends also quite strongly on the depth of the transom, which is not reflected in this
criterion. A better criterion to predict when the transom becomes dry is to use the transom
depth Froude number:
V
FNT = (2.84)
g⋅H

The criterion is then that the transom becomes dry when FNT>2.5.

As is mentioned above it is not a point separate the resistance of a dry transom when scaling a
model test result. It is, however, quite important to include this component when calculating
the resistance without model testing. When calculating the resistance, a practical criterion for
when the transom is dry is when the resistance found from equation (2.82) is smaller than the
resistance found from equation (2.83).

70
2.9 Appendage Resistance

The main intention with this section is

• to discuss the resistance of various types of appendages; how it can be scaled and
how it can be calculated.

An appendage is an addition (or possibly a subtraction) from the underwater part of a hull. It
is not always clear what is considered an appendage and what is not considered as an
appendage. Typical appendages on displacement ships are:

• Propeller shafts and shaft brackets (on twin screw ships with open shafting)
• Stabiliser fins
• Motion damping foils
• Tunnel thruster openings
• Bilge keels
• Rudders
In relation to model testing, rudders are often considered not as an appendage, but included in
the naked hull resistance. This is due to tradition and because, at least on single screw vessels,
the rudder is mostly inside the boundary layer of the hull.

Bilge keels are usually not included in model testing of resistance. The resistance of the bilge
keels is added to the ship resistance by adding an amount of wetted surface equal to the
wetted surface of the bilge keels to the wetted surface of the ship. This means that the ship
hull Reynolds number is used to predict the frictional resistance of the bilge keels, and that no
drag for vortex generation or similar is included. This method then represent the absolute
lower limit of the bilge keel resistance.

Appendages have usually small volume and are usually well submerged, so that it is safe to
assume that they don’t influence the wave making resistance. The exception, which might be
important for high speed craft, is if they change the trim of the vessel. Thus, the appendage
resistance can be assumed to consist of viscous resistance and viscous pressure resistance.
Therefore, the appendage resistance should be separated from the residual resistance when
scaling resistance from model to full scale.

When calculating the appendage resistance, one must take into account the local Reynolds
number of the appendage - that means calculating the Reynolds number based on the flow
length over the appendage, not the hull length. Furthermore, the flow velocity might be
influenced by the presence of the ship hull – thus a local wake fraction should be considered.
Then, a form factor specific for the appendage should be applied. This means that a generic
formula for calculation of appendage resistance is:

RApp = 12 ρV 2 (1 − wApp ) CF App (1 + k App )


2
(2.85)

Where

71
wApp is the hull wake fraction at the position of the appendage
CF App is the frictional resistance coefficient for the Reynolds number based on
V (1 − wApp ) ⋅ LApp
appendage length and the local velocity: RN App =
ν
kApp is the form factor of the appendage – which depends on the shape of the
appendage
If the appendage is located close to the propeller disk, the wake measured in the propeller
plane might be used to estimate the wake fraction of the appendage. If it is located further
forward, the wake is often dominated by the viscous boundary layer, which can be estimated
using standard formulas for boundary layer thickness and velocity distribution on a flat plate.

The form factor depends on the shape of the appendage. For foil shaped appendages, like
rudders and stabilizer fins, the following formula applies:

1 + k = 1 + 2 ( t c ) + 60 ( t c )
4
(2.86)

Where t/c is the typical thickness/length-ratio of the appendage. The last term is viscous
pressure drag due to partial flow separation, while the second term is the classical form factor,
due to increase of velocity due to displacement. The formula is empirical, and taken from the
ultimate reference for empirical drag data of generic shapes; Hoerner “Fluid Dynamic Drag”.

For “torpedo-shaped” appendages (bodies of revolution), the following formula might be used
to calculate the form factor:

1 + k = 1 + 1.5 ( D L ) + 7 ( D L )
1.5 3
(2.87)

Where D/L is the diameter/length-ratio of the appendage. The last term is viscous pressure
drag due to partial flow separation, while the second term is the classical form factor, due to
increase of velocity due to displacement. The formula is empirical, and taken from Hoerner
“Fluid Dynamic Drag”.

When there is a sharp transition between bodies with different boundary layer thickness,
vortices are formed, and the formation of these vortices lead to resistance – often called
interference resistance. Such vortex formation is shown in Figure 2.40. A typical example of
relevance to appendages is a stabilizer fin mounted on a hull. The stabilizer fin will have a
much thinner boundary layer than the hull, so a “necklace vortex” is formed. The resistance
created by this vortex can be estimated by:

(
CD int = FF 17 ( t c ) − 0.05 ( t 2 S )
2
) (2.88)

where
FF = a factor between 0 and 1, depending on the fairing radius
t = maximum thickness of the appendage
t/c = typical thickness/length ratio of the appendage
We have assumed that FF = 0.2 would represent a normal fairing. If the fairing radius is close

72
to 0, which would probably be the case for a stabilizer fin, FF becomes close to 1.00. As can
be seen from equation (2.88), the interference resistance is considered independent of
Reynolds number. Thus, it need not be separated when calculating the residual resistance of a
model test.

Figure 2.40 Vortices generated at the intersection between different parts of a hull,
leading to interference resistance

2.9.1 Resistane of tunnel thruster openings

When the water flows along the hull and passes the tunnel thruster openings, the water in the
outer part of the tunnel will circulate as illustrated in Figure 2.41.

The circulation requires energy to be maintained. This energy is taken from the flow along the
hull and manifests itself as an increase in the resistance relative to the resistance of the naked
hull. This contribution to the total resistance is a function of boundary layer thickness and the
shaping of the junction.

If the following edge of the exit or entrance is too sharp, separation occurs and the resistance
increases. If the edge is drawn inward, the resistance decreases. If it is rounded the resistance
decreases further.

Tests in towing tanks shows that the resistance caused by tunnel openings increases the ship
resistance by:

R = k ⋅ D m ⋅V 2 (2.89)
where
k = constant
m=2-3

73
D = diameter of the tunnel openings
V = velocity of the ship

which of course shows that it is beneficial to keep the openings as small as possible. The best
way to reduce the resistance is to reduce the circulation in the tunnel for example by mounting
ribs normal to the flow as shown in Figure 2.41. The actual direction of the flow over the
opening is frequently determined from paint smear tests in model scale.

It has been mentioned that the total increase in resistance can be determined from model tests.
These tests are not always reliable, because the resistance of the tunnel is small compared to
the resistance of the ship and because the effect of the openings of a model can be the same as
from a turbulence stimulator. Nevertheless, based on such tests it has been possible to propose
the following empirical approximation for the resistance increase due to tunnel openings in
percent of the total model resistance:

dRm A 1
= 15 ⋅ t ⋅ (2.90)
Rm S Fn gLm

S = wetted surface of the ship


At = cross sectional area of the tunnel
If it is assumed that the resistance of the model is relatively larger than the resistance of the
ship due to different CF values in model and in full scale. The values above must therefore be
increased with approximately 30 % if they shall be valid for the ship.

Figure 2.41 Removal of vortex generation through application of flat bars (Brix)

2.9.2 Scaling of appendage resistance from model tests


Scaling of appendage resistance from model tests is one of the most difficult issues in
powering prediction based on model tests. There is not one agreed method, but several
alternative methods, and we will present the most commonly used methods.

The appendage resistance of the model is usually found by towing the model with and without

74
appendages, and taking the difference as the total appendage resistance.

RApp M = RTM − RTM naked (2.91)

Where RTM naked is the total resistance of the model without appendages (naked hull).
For high speed crafts, where the appendages might influence the trim, one should apply some
kind of correction to avoid having the resistance change due to trim as part of the measured
appendage resistance. To find the resistance of each appendage, which would be of interest in
the scaling process, one has to test with and without each appendage separately, and this is
much too time consuming for routine model testing. The uncertainty of the resistance test
results might also make the relative accuracy of the appendage resistance quite low.

Some practically applied appendage scaling methods are:

• To include the appendage resistance in CR – which is really not to scale it. This is a
very simple method, which is probably why it is fairly widely used. The scale effect
of typical appendages will with this method be implicitly included in the correlation
allowance (see chapter 4).
• To scale the appendage resistance using a scale fraction β: C App S = β ⋅ C App M , where
0.5< β <0.8. The value of β varies between towing tanks, and is based on experience.
The choice of β must be seen in relation to the value of the correlation allowance.
• To calculate the appendage resistance in model and full scale, according to equation
(2.85). This method doesn’t require that the model is tested with and without
appendages, which makes this method quicker to apply than the β-fraction method.
This method gives less scale effect – higher full scale resistance, than the β-fraction
method, but smaller full scale resistance than just including the appendage resistance
in CR. An alternative way of using this method, which can be attractive for small
models where only towing tests are performed, is to test without appendages, and add
the calculated appendage resistance in full scale.
• To scale the appendage resistance as frictional resistance, which means that

C FS
C AppS = C AppM ⋅ (2.92)
C FM

This is similar to using equation (2.85) with kApp=0 and wApp=0.

• To scale the appendage resistance as:

C FS + ∆C F 2.5
RAppS = RAppM ⋅ ⋅λ (2.93)
C FM

where λ is the scale ratio. Converted to resistance coefficients, equation (2.93)


becomes:

ρ S C FS + ∆C F −0.5
C AppS = C AppM ⋅ ⋅λ (2.94)
ρM C FM

75
2.10 Residual and Total Resistance

The intention with this section is

• to show how the residual resistance is composed.

The term residual resistance was introduced by William Froude as the term that was left when
the frictional resistance was subtracted from the total resistance measured in a model test. His
idea was that when the frictional resistance was subtracted, the residual was mainly composed
of wave resistance, which could easily be scaled when the Froude numbers of the model and
ship was equal. His idea is still the fundament of ship model testing. However, the method is
usually refined by subtracting not only the frictional resistance of a flat plate with the same
area as the ship (or model), but subtracting also the form drag, air drag, transom stern drag
and so forth. The principle is to subtract all non-Froude scaled resistance components, so that
the residual can be Froude scaled without introducing scaling errors. As our discussion of
resistance components should have shown you, it is not easy to determine the non-Froude
scaled resistance components accurately and completely, so the residual is still not just wave
resistance, but composed of a variety of contributions, some of which aren’t really Froude
scaled.

It is also important to be aware that there is not a universally accepted definition of residual
resistance. Which components that are subtracted from the total to compute the residual
resistance varies a lot between model testing institutions. The only thing that all methods have
in common is that the frictional resistance is subtracted. The International Towing Tank
Conference (ITTC – http://ittc.sname.org/ ) maintains recommended procedures for model
testing and scaling, but these are not followed unanimously by the members. In addition, the
definition of residual resistance might be different for some special types of ships, for instance
SWATH ships (see chapter 10).

A very detailed decomposition of the resistance of a standard displacement ship is shown in


Figure 2.42.

76
Figure 2.42 Typical resistance components of a normal displacement ship (Harvald)

A formula for determination of residual resistance from model test for standard ships,
compatible with the standard procedure in use at MARINTEK, and in line with the previous
discussions in this chapter is:

CR = CTM − (1 + k ) CFM − C AppM − CBDM − C AAM (2.95)

Where

RTM
CTM = total resistance coefficient for model, CTM =
2 ρVM S M
1 2

0.075
CFM = frictional resistance coefficient for model, CFM =
( log10 RNM − 2 )
2

CAppM = appendage resistance coefficient for model


CBDM = transom stern resistance coefficient for model
CAAM = air resistance coefficient for model
The inclusion and expressions for the three last terms in equation (2.95) depends on the type
of ship, whether it has significant appendages or not, whether it has a transom stern, and
whether the air resistance has any significance. If the speed is so high that the transom is
completely dry, the transom stern resistance should not be subtracted when calculating the
residual resistance, since the drag of a dry transom is part of the wave resistance.

77
Since the residual resistance is assumed to mainly consist of wave resistance, it is assumed
free of scale effects in a Froude scaled experiment. Also the form factor is assumed free of
scale effects. The total resistance coefficient is then calculated as:

CTS = CR + (1 + k )( CFS + ∆CF ) + C AppS + CBDS + C AAS + C A (2.96)

Where CA is the correlation allowance, which is only added if CR is found from model tests.
The correlation allowance is treated in more detail later. The formula for CTS depends on the
formula for CR – the full scale counterparts of those subtracted to calculate CR must be added
when calculating CTS. Thus, equation (2.96) is valid only when equation (2.95) is used to
calculate CR.
Full scale resistance is calculated as:

RTS = 12 ρVS2 S S CTS

78
3 RESISTANCE PREDICTION

3.1 Introduction

Accurate prediction of the ship resistance in full scale is important for the prediction of the
propulsion power, and for calculation of the propeller thrust, ,which again is of crucial
importance when selecting the right propeller. Accurate prediction of the ship resistance is
difficult, due to the fairly complex three-dimensional flow pattern around the ship.

Resistance calculation methods might be divided in three different categories, as shown in


Figure 3.1. The first category – purely empirical methods – are methods which are based on
practical experience. This is typically how the power requirement for pleasure boats is
determined, but also some ship owners trust this way of determining the required power of
their new ship. This category is not further covered in this book.

Another category that is only briefly covered is numerical solutions of Navier-Stokes


equations. In this case, we do not necessarily need to divide the resistance in viscous and
wave resistance, although this is quite commonly done.

The main emphasis in this chapter is on the category of methods using Froude’s hypothesis.
These methods can be further subdivided in:

• Empirical methods to determine CR and form factor


• Model testing to find CR and form factor
• Numerical calculations to determine the wave resistance
We shall discuss each of these types of methods in the following sections.

Purely ”Froude’s Numerical


empirical hypothesis” solution of N-S
equations

”guesswork” Empirical Model tests to Numerical CFD


based on calculation find k and CR calculation of RANS and
resistance of of k and CR wave resistance similar methods
similar ships to find CW (⇒CR)

Figure 3.1 Classification of resistance prediction methods

79
3.2 Predicting resistance from model tests

The main intention with this section is

• to show how the resistance of the ship model resistance is measured, and to show
how the ship resistance is estimated based on the tests

The resistance of a ship is normally estimated on basis of model tests made in tanks with a
length of 200 - 300 meters and with a breadth of 10 -12 meters. The depth varies from 4 to 6.
The model length varies between 3 and 8 meter. Figure 3.2 shows a cross-section of a towing
tank with the carriage and the model connected to the carriage, while Figure 3.3 shows how
the model is fastened to the carriage, and how the trim and resistance are measured.

Figure 3.2 Cross section of a towing tank, with carriage and model.

A/D Filter Amplifier

"peg" for fixing the model


Towing carriage during acceleration
Measurement of:
Model resistance RTM Force transducer
Model speed VM (Dynamometer)
Sinkage fore and aft
Ship model
Flexible connection between model and transducer

Trim posts
Figure 3.3 Test set-up for towing tests

80
During the tests, the model resistance is measured as a function of the velocity. The model is
hull is an accurate scale model of the ship hull, but usually excluding small details, such as
cooling water intakes.

It has been shown above how the ratio between gravity force and inertia force has to
be in balance on ship and model, in order to let wave making and wave resistance be
identical, and that this requirement is obtained if the Froude number is the same for
ship and model.

Ls
If the speed of the ship is VS, the length of the ship LS and the scale λ = , the model
Lm
velocity is:

Vs
Vm = (3.1)
λ

Normally, only a single model is used for the testing of a given ship. However, some times
models of different scale or length are used. This may give results as indicated in Figure 2.17.
Tests of this type are called "geosim tests".

Before the measured model resistance is applied to determine the residual resistance and
finally the ship resistance, the total model resistance RTm must first be made dimensionless:

RTm
CTm = (3.2)
ρ
⋅V ⋅ S m
2
m
2
When CTm is determined, the frictional resistance of the model is calculated:
0.075
CFm =
[log RNm − 2]
2

applying
Vm ⋅ Lm
RNm =
ν
The next step would be to determine the form factor k by plotting:

C Tm - C AAm - C DBm - C APm


CFm

n
F
as a function of N as already shown in Figure 2.16, but this would require model testing at
CF
low speeds (which is time consuming and inaccurate) and which gives too high form factor. A
form factor, which is too high, gives a predicted resistance, which is too low. This is because
too much model resistance is scaled as viscous resistance. Therefore, the form factor k is as
shown often determined from the main dimensions of the ship.

Finally the air resistance CAAm is determined and the residual resistance of the model is

81
estimated as:

CR ( FN ) = CTm − (1 + k0 ) ⋅ CFm ( RN ) − C AAm − CDBm (3.3)

It is assumed that CR (FN ) is identical for ship and model.

It is also possible to measure the wave resistance directly by measuring the wave profile when
the model passes as indicated in Figure 6.10.

If the residual resistance is known, the total resistance coefficient for the ship is:

CTs ( FN ) = (1 + k0 ) ⋅ (CF + ∆CF ) + CR ( FN ) + C AAm + C AP + CDBm (3.4)

In full-scale, the resistance is corrected applying an empirical factor CA. This factor contains
other components than resistance. It is determined through analyses of trials.

The procedure for determination of CR (FN) is for each FN or for each velocity as sketched
below:

CB
Φ To be calculated ⋅ B ⋅ (TA + TF )
LWL
Measured during the VS
VM (m/s)
tests λ
Measured during the
RTm (N) RTm
tests
RTm
CTm To be calculated ρ
⋅ V 2m⋅S m
2

k To be calculated k = 0.6 ⋅ Φ +145 ⋅ Φ 3.5

V ⋅ Lm
RNm To be calculated
ν
0.075
CFm To be calculated
[ log RN − 2]
2

ρ air A Pm
CAAm To be calculated C AAm = C AP ⋅ ⋅
ρ vann S m
CR (FN) To be calculated CTm − (1 + k ) ⋅ CFm − C AAm

In this example, it has been assumed that the ship is without transom stern. Therefore CDB is
not applied.

We have further, not applied any resistance due to appendages. This resistance is for larger
ships small and often included in CA.

82
The ship resistance is now estimated as in the following table:

V(m/s) Given V = 0.5144VS(Knots)

⋅ B ⋅ (TA + TF )
CB
Φ Calculated Φ=
LWL

K Calculated k = 0.6 ⋅ Φ + 145 ⋅ Φ 3.5

BT Calculated B, T

L, ∇
1
L ∇ 3 Calculated

V ⋅L
RNs Calculated
ν
0.075
CFs Calculated
[ log RN − 2]
2

Calculated as L
CR(FN)
shown above Cp ,FN ,

1/3

∆CF and ∆CF=[110·HV 0.21- 404] ·CFs2


CAAs Calculated ρ luft A ps
C AAs = C AP ⋅ ⋅
ρ vann S s

CTs Calculated (1+k)·(CFs+∆CF)+CR+CAAs

ρ
CTs ( FN , RN ) Vs 2 S s
Rs(FN,RN )
Calculated
(N) 2

83
3.2.1 Example of Estimation of Ship Resistance based on Model Tests

Imagine that a tanker has been model tested. Model and ship had the following main
dimensions:

DIM/COEFFICIENT MODEL SHIP


LWL = length of waterline (m) 6.022 239.19
B = breadth (m) 1.057 42
T = draught (m) 0.307 12.20
CB = block coefficient 0.834 0.834
∇ = displacement (m3) 1.623 101664
S = wetted surface(m2) 9.060 14291
CAam 0.0288 × 10-3 0.0288 × 10-3

The form factor follows from k = 0.6 ⋅ Φ + 145 ⋅ Φ 3.5 which gives: 1 + k = 1.1327

Model results:

Vs Vm RTm
FN CTm × 103 CFm × 103 CVm × 103 CR × 103
knots m/s N/9.81
12.5 1.020 0.133 1.863 3.875 3.350 3.790 0.056
13 1.061 0.138 2.004 3.852 3.326 3.767 0.056
13.5 1.102 0.143 2.151 3.834 3.303 3.741 0.064
14 1.143 0.149 2.304 3.820 3.281 3.716 0.075
14.5 1.184 0.154 2.465 3.809 3.260 3.693 0.087
15 1.224 0.159 2.634 3.804 3.240 3.670 0.088
15.5 1.265 0.165 2.813 3.805 3.221 3.640 0.106
16 1.306 0.170 3.007 3.817 3.203 3.628 0.160
16.5 1.347 0.175 3.222 3.846 3.185 3.608 0.209
17 1.388 0.181 3.468 3.900 3.168 3.588 0.283
17.5 1.428 0.186 3.760 3.989 3.152 3.570 0.390
18 1.469 0.191 4.117 4.129 3.136 3.552 0.548

The residual resistance follows from:

CR = CTm - CVm - CAAm

84
Example: Speed prognoses

The prognoses are made based on the following hull roughness, air resistance and correlation
factor:

H = 150
CAA s= 0. 0288× 10-3
CA = - 0. 228 × 10-3 (correlation factor, a function of ship dimensions and model
testing technique).
The total resistance follows from:

CTs=CVs+CR+CAAS+CA

Vs RTs RNs × 10-9 CR ×103 CFs × 103 CVs × 103 ∆CF ×103 CTs ×103
Knots kN

12.5 527 1.295 0.056 1.483 1.838 0.140 1.694


13 572 1.346 0.056 1.476 1.838 0.147 1.695
13.5 620 1.398 0.064 1.469 1.838 0.154 1.695
14 670 1.450 0.075 1.462 1.838 0.160 1.703
14.5 724 1.502 0.087 1.456 1.838 0.166 1.726
15 782 1.554 0.088 1.450 1.838 0.172 1.727
15.5 846 1.605 0.106 1.445 1.838 0.178 1.745
16 917 1.657 0.160 1.439 1.838 0.183 1.800
16.5 1001 1.709 0.209 1.434 1.838 0.189 1.848
17 1103 1.761 0.280 1.429 1.838 0.194 1.919
17.5 1233 1.812 0.390 1.424 1.838 0.198 2.029
18 1404 1.864 0.548 1.419 1.838 0.203 2.187

The resistance is calculated fromCTs:

ρ
R = CTs ⋅ ⋅ V0 2 ⋅ S
2

for the actual velocities.

85
3.3 Empirical resistance prediction methods

The main point in this section is

• to warn against uncritical use of programs for resistance calculation based on


analyses of occasional databanks or hull series found in the literature

Most ship model tanks and large shipbuilders have extensive data for model tests and full
scale analyses. This material is a good basis for extrapolation and interpolation but can hardly
be used for optimisation of the resistance for given main dimensions.

Nevertheless, essential reductions are obtained, but because of systematic model testing of the
actual ship and not because of interpolation in databases.

Theoretical calculation of resistance is still an academic exercise. The shipbuilder must


therefore rely on model testing, even if it is easy to point out limitations in model testing
techniques.

The two most important parameters for estimation of resistance are the fullness and FN . The
optimum combination of ship length, velocity and fullness is given in the following table:

FN 0.15 0.20 0.25 0.30 0.35


CB (upper) 0.85 0.76 0.67 0.58 0.49
CB (lower) 0.79 0.70 0.61 0.52 0.43

Figure 3.4 Todd's methodical experiment of models of single-screw merchant ships


(Todd, 1963)

In Figure 3.4 (Todd' s 60 series) (1963) the total resistance for ships with differences in

86
fullness is given. If the figure is studied, it is observed that there is a strong increase in
resistance above a certain speed for all the ships. The service speed is therefore usually placed
well below this critical speed.

Below, the main characteristics of a selection of some hull series are given:

SHIPTYPES SOURCE L/B L/∇1/3 CP CB


Trawlers Ridley, Nevitt 56 – 3.85 - 0.55 - 0.70
63 5.23
Trawler Lackenby 0.625-0.810
NPL Coasters Dawson31-60 4.44 - 0.625-0.81
8.00
SSPA Tankers Edstrand 53-56 0.725-0.800
SSPA Cargo Edstrand-Freimanis 6.54 - 0.525-0.750
8.14
SSPA Coaster Warholm 53 4.5 -7.5 0.600-0.750
BMT Moon-Lackenby 66 0.600-0.800
60 (ATTC) Todd 5.5 - 4.79- 6.84 0.614 - 0.600-0.800
8.25 0.805
Taylor S. S. Taylor 0.70 - 0.48 - 0.86
8.75
Fullform-low L/B Keil75 2.73 - 3.73 - 0.77 0.75
4.74 5.13
Data Sheets SNAME 150 ship
Cylindrical bow Muntjewerf 6.5 0.800-0.850
Rounded bilge ( Marwood, Bailey 0.397-0.693
NPL) 69

Strong requirement is present for simple programs, which can be used to calculate the
resistance. Such programs have for a long time been in use throughout the world, and are
based on analyses of previous performed model tests. They are developed for use at an early
stage in the design process, where only a rough estimate of the resistance is required. It must
be mentioned that such programs can not be used, if accurate results are required. Most of
these methods have one thing in common, namely that they are based on "old fashioned" hull
shapes that are not in use any more. Due to the weight that has been laid on energy saving hull
shapes the last 30 years, these methods will therefore estimate a too high resistance. Besides,
many relevant hull parameters have often not been included in the analysis. For that reason,
the scatter may become large.

Such methods must therefore be applied with care and resistance prediction and hull design
left to or made in co-operation with a ship model tank, which has access to updated resistance
data.

87
In the table below, a few examples, indicate what may be expected, if different methods
available in the literature are applied. Deviations between calculated and measured model
resistance have been compared. The mean values of the deviations and the standard deviation
for different methods are:

Method single-screw twin-screw


mean value std deviation mean value std deviation
Holtrop/Mennen +2.7% 13.4% +8.4% 17.9%
Guldhammer/Harvald +4.8% 15.2% +12.1% 23.0%
Lap-Keller +2.9% 13.4% +16.2% 19.7%
Series-60 +2.4% 13.4% +17.7% 22.4%
Oortmerssen +5.7% 14.8% +6.8% 20.2%
Danckwardt Trawler -4.3% 17.9% +17.9% 31.5%

It is clear that these high standard deviations are unacceptable. A main problem is that none of
the methods considers the bulb in a satisfactory way. Most ships of today are equipped with
bulbs.

An other problem, is that a number of small details in the shape of the frames, that at first may
look insignificant are not considered, like transitions between shipsides and bottom. The lines
in the after body are probably the most important factor.

3.3.1 Calculation according to Hollenbach

The intention with this section is

• to show how it is possible to make a rough estimation of the resistance based on


analyses of previous model tests.

Probably the most recently published empirical method for resistance prediction for
conventional ships is Hollebach (1998). He published a regression analysis of ships tested in
the Vienna Ship Model Basin, which is one of the oldest ship model basins in the world. He
only considered models which have been tested later than 1975. As indicated below in
comparisons with other methods, it is seen that the improvement in the prognoses is
remarkable, but still the deviations are large.

single-screw single- screw twin-screw


design draft ballast draft design draft
Mean standard Mean standard mean standard
deviation deviation deviation
Holtrop-Mennen - 0.5 % 12.8 % 6.3 % 16.1 % 5.8 % 18.4 %
Guldhammer 0.8 % 11.0 % 10.5 % 17.9 % 11.2 % 19.2 %
Lap-Keller - 0.5 % 12.9 % 27.9 % 32.9 % 14 % 23.4 %
Series-60 - 1.0 % 11.6 % 37.3 % 42.7 % 15.2 % 23.3 %
Hollenbach - 1.0 % 9.4 % - 0.2 % 11.2 % 3.5 % 13.3 %

88
The standard deviation for Hollenbach’s method is lower especially for the ballast condition
and for twin-screw ships.

Below, Hollenbach's method is presented in order to serve as a basis for simple programming.
An important parameter in Hollenbach's method is the length of the ship defined as in Figure
3.5.

It is worth to observe that the definition of the length LOS is different for ballast and full load.

Froude's number is an important parameter in the calculations. Different lengths LFn are
applied. The length is defined as follows:
Lfn
L OS / L < 1 L OS
1.0 > LOS / L < 1.1 L + 2 / 3 * ( LOS - L )
1.1 < L OS / L 1.0667 * L

In the calculations the Froude number is defined as:

V
Fn =
gL fn

Hollenbach defines the coefficient of residual resistance in a special way:

RR
CR =
ρ
⋅V0 2 ⋅ T ⋅ B
2

Figure 3.5 Definition of ship lengths in Hollenbach's method

89
The total resistance must consequently be expressed like:

ρ
RTm = ⋅ V0 2 ⋅ [CFm ⋅ S + CR ⋅ B ⋅ T ]
2

CR is normally expressed with a "standard value" CR,Standard in the following way:

a1 a2 a3 a4 a6
⎛T ⎞ ⎛ B⎞ ⎛ L ⎞ ⎛ L ⎞ ⎛ D ⎞
CR Hollenbach = CR ,Standard ⋅ CR , Fnkrit ⋅ k L ⋅ ⎜ ⎟ ⋅ ⎜ ⎟ ⋅ ⎜ os ⎟ ⋅ ⎜ wl ⎟ ⋅ ⎜ P ⎟
⎝ B ⎠ ⎝ L ⎠ ⎝ Lwl ⎠ ⎝ L ⎠ ⎝ TA ⎠
a5
⎡ TA − TF ⎤
⎢⎣1 + L ⎥⎦ ⋅ (1 + N Rud ) ⋅ (1 + N Brac ) ⋅ (1 + N boss ) ⋅ (1 + NThr )
a7 a8 a9 a10

where
TA = draught at the aft perpendicular
TF = draught at the forward perpendicular
DP = propeller diameter
Nrud = number of rudders
Nbrac = number of brackets
Nboss = number of bosses
NThr = number of side thrusters

The standard value for CR is:


CR,Standard = b11 + b12 ⋅ Fn + b13 ⋅ Fn 2 + CB ⋅ ( b21 + b22 ⋅ Fn + b23 ⋅ Fn 2 ) + CB 2 ⋅ ( b31 + b32 ⋅ Fn + b33 ⋅ Fn 2 )

where

⎡ ⎛ Fn ⎞ ⎤
c1

CR , Fnkrit = max ⎢1.0, ⎜ ⎟⎟ ⎥


⎢ ⎜F
⎝ n , krit ⎠ ⎥
⎣ ⎦
Fn , krit = d1 + d 2 ⋅ CB + d3 ⋅ CB 2

k L = e1 ⋅ Le 2
The formulas are used in the following regions:

Fn ,min = min ( f1 , f1 + f 2 ⋅ ( f 3 − CB ) )
Fn ,max = g1 + g 2 ⋅ CB + g3 ⋅ CB 2
Values of the various coefficients are found in a table, named Table II, that is found at the end
of this section.

In the calculations the definitions "mean resistance", "max resistance" and "min resistance"
are applied. The max and min values are limits for 5 % of the statistical material, which
deviates from the mean value. Minimum resistance is an expression for what is possible to
obtain through extensive calculations and model tests. The max values are for ships with "bad

90
shapes".

The maximum resistance is expressed as:

RT max = h1 ⋅ RTmean

where h1 is taken from the table.

1− t
For the hull efficiency ηH = , if the ship is a single screw ship, Hollenbach proposes to
1− w
use the formula:

−0.58 0.1727
⎛R ⎞ ⎛B⎞ ⎛ DP 2 ⎞
η H ,model = 0.948 ⋅ C B 0.3977
⋅ ⎜ T,mean ⎟ ⋅⎜ ⎟ ⋅⎜ ⎟
⎝ RT ⎠ ⎝T ⎠ ⎝ B ⋅T ⎠

In ballast draft and for single screw:

0.2991 −0.3.2806 −0.2317


⎛L⎞ ⎛L ⎞ ⎛D ⎞
η H ,model =1.055 ⋅ C B
1.0099
⋅⎜ ⎟ ⋅ ⎜ wl ⎟ ⋅⎜ P ⎟
⎝B⎠ ⎝ L ⎠ ⎝ T ⎠

If the model is twin screw, the following formula is recommended:

0.0285
⎛ DP 2 ⎞
η H ,model = C ⋅ CB 0.1202
⋅⎜ ⎟ ,
⎝ B⋅T ⎠

where the constant C has the following values:

Two shafts and rudder C = 1.125


Two "skegs" og two rudders C = 1.224
Two shafts with brackets and one C = 1.086
rudder
Two "built-in" shafts and one rudder C = 1.096

The residual resistance was in Hollenbach's (1998) analyses determined by assuming that the
form factor = 0 for all ships. This assumption is too rough to be used in the full scale
prognoses. It is therefore reason to believe that a conversion of Hollenbach's (1998) data in
such a way that the data can be applied with a form factor may improve the situation.

Hollenbach's (1998) CR data can be converted in the following way:

B ⋅T
CTm = CFm + CRHollenbach ⋅ = CFm ⋅ (1 + k ) + CR
S

which gives:

91
B ⋅T
CR = CRHollenbach ⋅ − k ⋅ CFm
S

CFm follows from the actual Froude's number:

6 ⋅ FN ⋅ 6 g
RNm = ⋅106
1.1395

0.075
CFm =
[ log( RNm ) − 2]
2

With form factor the resistance is:

CTs = ( CFs + ∆CF ) ⋅ (1 + k0 ) + CR + C A

CTs = total resistance


k = form factor
CF = frictional resistance
∆CF = added frictional resistance due to roughness
B ⋅T
CR = residual resistance defined as CR = CRHollenbach ⋅ − k ⋅ CFm
S
CA = correlation factor from analyses of trial results (see separate section)

It is proposed to apply a form factor between:

k = 0.6 ⋅ φ + 145 ⋅ φ 3.5

and

k = 0.6 ⋅ φ + 11808 ⋅ φ 5.15

where

CB
Φ= ⋅ B ⋅ (TA + TF )
L

This form factor was derived through analyses of different model test results. With the
information given above, it is now easy to make a program for estimation of the resistance. In
such a program, we may include the polynoms, which are given for the propellers in the B
series. Alternatively, form factors according to Holtrop may be used. (See 2.8)

In case, we just follow Hollenbach's (1998) procedure and definition of CR, this may be done
in the following way:

92
B ⋅T
CTs = (CFs + ∆CF ) + ⋅ CRHollenbach + C A
S

RR
CRHollenbach = residual resistance coefficient, defined as:
ρ
⋅ V0 2 ⋅ B ⋅ T
2

Figure 3.6 Values of the coefficients used in Hollenbach's method for resistance
calculation

93
3.3.2 Resistance calculated according to Holtrop's Method (1988)

The main point in this section is

• to present a method for calculation of ship resistance based on analyses of model


tests made in the ship model tank in Wageningen up to 1988.

The method is based on a wide but to some extent "old fashioned" selection of ships. The
weak point as in Hollenbach’s method, is the representation of the bulb. However, it is a result
of the most extensive statistic analyses of model test results ever made.

The method was originaly published by Holtrop and Mennen (1978). The wave resistance was
given by a simplification of a formula originally developed by Havelock (1951) where it is
assumed that two disturbances moving in the water at a distance λL from each other gives the
wave resistance:

⎛ ⎞
−2
− mFN
Rw = ρ ⋅ g ⋅∇ ⋅ ⎜ k1 ⋅ e 9 + k2 ⋅ e− mFN + k3 ⋅ e − mFN ⋅ cos ( λ ⋅ FN −2 ) ⎟
−2 −2

⎜ ⎟
⎝ ⎠

It is assumed that the two disturbances have an infinite breadth. The first term within the
brackets is a correction for the influence of the diverging waves.

The interference between the transverse waves is taken care of by the cosine link. The
constants k1, k2, k3, and m are functions of the hull shape.

Holtrop transformed Havelock's formula and applied the following expression in his

analyses:

Rw = k4 ⋅ k5 ⋅ k6 ⋅∇ ⋅ ρ ⋅ g ⋅ exp ⎡⎣ k7 ⋅ FN d + k8 ⋅ cos(λ ⋅ FN −2 ) ⎤⎦

where k4, k5, etc are constants.


λ is approximated by:

L
λ = 1.446 ⋅ C p − 0.03 ⋅
B

In the analyses it is distinguished between three velocity regions:

A: FN < 0.4

B: FN = 0.4-0.5

C: FN > 0.5

In region A where FN < 0.4 :

94
RWA = c1 ⋅ c2 ⋅ c5 ⋅∇ ⋅ ρ ⋅ g ⋅ exp ⎡⎣ m1 ⋅ FN + m4 ⋅ cos(λ ⋅ FN ) ⎤⎦
d −2

1.07961
c1 ⎛T ⎞
⋅ ( 90 − iE )
−1.37565
2223105 ⋅ c7 3.78613 ⋅ ⎜ ⎟
⎝B⎠
0.33333
c7 ⎛B⎞
0.229577 ⋅ ⎜ ⎟
⎝L⎠
when
B
< 0.11
L
c7 B
L
when
B
0.11 < < 0.25
L
c7 L
0.5 − 0.0625 ⋅
B
when
B
> 0.25
L
1
m1
L ∇ B 3
0.0140407 ⋅ − 1.75254 ⋅ − 4.79323 ⋅ − c16
T L L
c16 8.07981 ⋅ CP − 13.8673 ⋅ CP + 6.984388 ⋅ CP 3
2

when
CP < 0.8
c16 1.73014 − 0.7067 ⋅ CP
when
CP > 0.8

In region C where FN > 0.5 :

RWC = c17 ⋅ c2 ⋅ c5 ⋅∇ ⋅ ρ ⋅ g ⋅ exp ⎡⎣ m3 ⋅ FN d + m4 ⋅ cos(λ ⋅ FN −2 ) ⎤⎦

c17 ∇ L
6919.3 ⋅ C -1.3346 ⋅( ⋅(
2.00977 1.40692
M 3
) -2)
L B
m3 B 0.326869 T 0.605375
-7.2035 ⋅ ( ) ⋅( )
L B
m4
c15 ⋅ 0.4 ⋅ exp(−0.034 ⋅ FN −3.29 )

95
c2 exp(−1.89 ⋅ c3 )

c5 AT
1 − 0.8 ⋅
B ⋅ T ⋅ CM

0.56 ⋅ ABT 1.5


c3
⎣ (
⎡ B ⋅ T ⋅ 0.31 ⋅ ABT + TF − hb ⎤
⎦ )
3
c15
-1.69385 for L < 512

c15 L
-8 3
− 1.69385 + ∇
1/3
for 512 < L < 1726.91
2.36 ∇
0
c15
when
L3
> 1726.91

D -0.9

λ L
1.446 ⋅ CP − 0.03 for
B
λ 1.446 ⋅ CP − 0.36 for L/B>12

In the equations, Holtrop applied the following symbols:

L Waterline length
B Breadth
TF draught at the forward perpendicular
TA draught at the aft perpendicular

∇ displacement (m3)

ABT transverse bulb area


AT transom area
HB distance to the centre of gravity of the bulb area
above the keel
CM mid ship area coefficient
CP prismatic coefficient

96
CWP waterline area coefficient
iE half of the angle of inlet in degrees
LCB longitudinal position of the centre of buoyancy in
front of 0.5L in % of L
LR Length of run(see separate formula)

If iE in c1 is unknown, the following formula is used:

⎡ L ∆ ⎤
iE = 1+ 89 ⋅ exp ⎢ −( )0.80856 ⋅ (1 − CWP )0.30484 ⋅ (1 − CP − 0.0225 ⋅ LCB )0.6367 ⋅ ( L R )0.34574 ⋅ (100 3 )0.16302 ⎥
⎣ B B L ⎦

where the «length of run» is:

LCB
LR = L ⋅ (1 − CP + 0.006 ⋅ CP ⋅ )
4CP − 1

If 0.4 < FN < 0.55 the following approximation for the wave resistance is applied:

RW ( FN = 0.55) − RW ( FN = 0.4 )
RW = RW ( FN = 0.4 ) + ( FN − 0.4) ⋅
1.5

A difficult and unclear point, also in this method, is as just mentioned the influence of the
bulb on the wave resistance.

The resistance due to the bulb is expressed by:

ρ⋅g
RB = 0.11 ⋅ exp ( −3 ⋅ PB −2 ) ⋅ F 3 NI ⋅ ABT 1.5 ⋅
1 + F 2 NI

where

ABT
PB = 0.56 ⋅
TF − 1.5 ⋅ hB

and the FN based on immersion is:

V
FNI =
ρ ⋅ (TF − hB − 0.25) ⋅ A BT + 0.15 ⋅ V 2

This resistance is added to RW. The influence of the bulb on Rw is taken care of by the factor
c2. RB is to consider as wave resistance, when the bulb is close to the surface.

97
Further, there is a contribution from the transom on the total resistance:

ρ
RTR = ⋅ ρ ⋅ V 2 ⋅ c6
2
where
c6 = 0.2 ⋅ (1 − 0.2 ⋅ FNT ) for FNT < 5

and c6 = 0 for FNT = > 5


V
FNT =
2 ⋅ g ⋅ AT
B + B ⋅ CWP

It is assumed that the wetted surfaces used in the formulas are calculated according to the
formula:
⎡ B ⎤ A
S = L ⋅ (2 ⋅ T + B) ⋅ CM ⋅ ⎢0.453 + 0.4425 ⋅ CB − 0.2862 ⋅ CM − 0.00346 ⋅ + 0.3696 ⋅ CWP ⎥ + 2.38 ⋅ BT
⎣ T ⎦ CB
If the speed prognoses are made according to this method and the results are compared to the
trial results, the standard deviation becomes higher than the standard deviation from a
comparison between trial results and prognoses based on model tests.

The final total resistance is:


RT = RW + R B + R F + RTR + R A
where
ρ
RF = ⋅ S ⋅ V 2 ⋅ (CF + ∆CF ) ⋅ (1 + k1 )
2
and

RA = 12 ρ ⋅ S ⋅V 2 ⋅ C A =correlation factor

The formfactor k1 is given by:


1.06806 0.46106 0.121563 0.36486
⎛B⎞ ⎛T ⎞ ⎛ L ⎞ ⎛ L3 ⎞
⋅ (1 − CP )
−0.604247
1 + k1 = 0.93 + 0.487118 ⋅ c14 ⋅ ⎜ ⎟ ⋅⎜ ⎟ ⋅⎜ ⎟ ⋅⎜ ⎟
⎝L⎠ ⎝L⎠ ⎝ LR ⎠ ⎝∇⎠
The coefficient c14 = 1 + 0.011⋅ Cstern accounts for the stern shape. It depends on the stern shape
coefficient Cstern .Typical values for Cstern are:

Type of afterbody Cstern


Pram with gondola: -25
V-shaped section shape: -10
Normal section shape: 0
U-shaped section shape: +10

98
This formfactor does not contain any correction for the appendage resistance. The correlation
factor CA is based on analyses of trial results and is mainly a function of ship length and
fullness of the hull .

3.4 Numerical resistance prediction methods

Here we will discuss numerical methods to predict the wave resistance of ships. The
dominating principle for this type of calculations is potential flow theory.

It is also possible to calculate the wave resistance by numerical solutions of the Euler
equations and of course the Navier-Stokes equations. These two methods will not be further
discussed – we will concentrate on potential flow theory.

Potential flow methods can be categorised in different ways. One way is how the hull is
represented by sinks and sources:

• Slender ship theory


• Thin ship theory
• 3-D panel methods
In slender ship theory, the geometry is represented by a single row of sinks and sources. This
means that only the longitudinal distribution of buoyancy is represented. These methods are
no longer in practical use for resistance prediction.

In thin ship theory, the geometry is represented by a distribution of sinks and sources on the
centreplane of the vessel, so that the distribution of buoyancy both longitudinally and
vertically is modelled. Input to the calculation is thus a “half-beams” distributed in a vertical
grid. Thin ship theory has proven to be useful for systematic optimisation studies of multihull
configurations, and when a quick estimate of wave resistance is needed. It is further covered
in section 3.4.2.

In 3-D panel methods, sinks and sources are distributed over the real surface of the submerged
part of the ship hull. The ship hull is discretised into panels, or elements. If the source strength
is constant over the panel, it is called a first order panel method. If the source strength is
allowed to vary, for instance linearly or using a spline, it is called a higher-order method.
Higher order methods require less panels to obtain the same accuracy as a first-order method,
but they are more complicated to program and require more computational resources per
panel than first order methods. Another important classification is between linear and non-
linear methods. A linear method panels the hull up to the undisturbed free surface, so the
shape above the undisturbed free surface is not taken into account. In a linear method,
panelling of the free surface can be avoided by use of a Green’s function. A non-linear
method needs panels up to the real waterline at the speed in question. It also needs
panelisation of the free surface. Panelisation of the free surface means many more panels are
needed, so it becomes more computationally demanding. Panelisation up to the real waterline
at speed means that iteration is required, since the real waterline at speed is unknown and
must be found from the calculation. Since iteration is required, the re-panelisation required for
each step in the iteration should be automatic, something that is fairly complicated to program

99
in a robust manner. Non-linear methods are definitely more accurate than linear methods.
Many different computer codes using 3-D panel methods for wave resistance calculations
exist, but most of them are academic codes, or codes developed and used by single research
institutes. One of the few commercially available wave resistance calculation programs is
Shipflow, made by Flowtech in Sweden http://www.flowtech.se/

In the following, the theoretical foundation for potential flow methods for wave resistance
calculations is outlined.

3.4.1 Disturbance of the Free Surface due to Sinks and Sources


Any ship-shaped form, moving on the surface of a liquid, can be represented by a suitable
distribution of sources and sinks on the surface of the hull or along it's middle plane. It is
therefore necessary to consider at first the velocity potential of one such source moving at any
distance below the free surface of the fluid. Consider a simple source of strength m, moving
parallel to the axis of x with the velocity c. Assume that the source has suddenly been formed
at the point (0, 0, -f) in the fluid at the time t = 0 and is active during the time interval dτ.

This is an impulse motion, but on the free surface only the constant atmospheric pressure and
no impulse pressure is present. Consequently, the velocity potential during the short time
interval dτ in a point P is given by the difference between the contribution from a source
situated at the point (0, 0, -f).

If the free surface of the liquid had been a rigid surface, the image system of the source would
have been a source of strength m situated at the image point (0,0,f). However, the free surface
is not a rigid surface. Therefore, it seems natural to assume that we must have some sort of
image system above the free surface. Let us therefore assume that this image system can be
built up of a certain distribution of sources. First, the potential of the source itself and image is
considered:

m m

r1 r2

where the distances from the sources to a given point is:


2
r1 = x 2 + y 2 + ( z + f ) 2
2
r2 = x 2 + ( z − f ) 2

This impulse motion leads to an immediate surface elevation, which when the source dies out,
will be spreading out on the surface as a point disturbance or spherical waves under the
influence of the gravity force, like when a stone is falling vertically on a surface at rest.

The velocity potential of this source at the time t was according to Havelock given by an
expression which was the potential for a wave motion, where the wave height decreases and
the wavelength increases as the spheres spread out.

A pressure point moving on the surface may be substituted by a continuous process where the
motion of the fluid is created by an impulses on the surface at the points x1, x2, x3 at the
moments t1, t2, t3, where x1<x2<x3 and t1<t2<t3. Each of these impulses will create a series of
ring waves, which due to the interference will create a wave system built up of longitudinal

100
and diverging components. If more and more impulses are created and the time interval
between the impulses are reduced, we will have the same motion of the fluid as when a single
pressure impulse is in rectilinear motion on the surface. Each impulse has a given velocity
potential and a given motion in the fluid. The velocity potential at the time t is then expressed
by the sum of the velocity potential for each of the impulses at the time t.

**
The velocity potential φ has the property that that the velocity of the flow in any direction is the partial derivative
of φ with respect to that direction. Thus for a uniform stream of velocity U, the velocity is defined by the
expression:

∂φ
= −U ,
∂x

or

φ = −Ux

and

∂φ ∂φ
= =0
∂y ∂z

This method which was applied by Kelvin, may as indicated above be applied in our case. We do therefore
assume that the motion which is created when a source is moving below the surface, may be build up from an
infinite number of momentanesly working sources along the path. The motion due to the source is then obtained
by summing all these steps up to a given time t.

If it is assumed that the motion of the source starts with the velocity c and that this velocity later is constant, it is
possible to determine the velocity potential for a constant and rectilinear motion and then make an integration up
to t. Finally, after a lengthy calculation we get the complicated expression:

π

m m 4mk0 2 cos(k x ⋅ cos θ )
φ= − − ⋅ ∫ sec2 θ ⋅ dθ ∫ ⋅ cos(k y ⋅ sin θ ) × e k ⋅( z − f ) ⋅ dk
r1 r2 π 0 0
k − k0 ⋅ sec θ
2

π
2
−4mk0 ⋅ ∫ sin(k0 ⋅ x ⋅ sec θ ) ⋅ cos( k0 ⋅ y ⋅ sec 2 θ ⋅ sin θ )e k0 ⋅( z − f )⋅sec θ ⋅ sec 2 θ ⋅ dθ
2

where

g
k0 = =wave number
c2
The first and second link represents the velocity of the source in an infinite parallel flow and the velocity
potential of an equally sink in an infinite flow. The two last links represent the mirror of the source in the free
surface.

101
Waves generated by Advancing Sinks and Sources

When an isolated group of waves are moving with almost the same wavelength, the velocity of advance is
smaller than the velocity of each single wave forming the group. If a given wave is studied, it is observed that it
advances through the group as it slowly dies out as it approaches the leading edge of the group. The previous
position in the group is replaced by other waves all moving forward from the following part.

The simplest way to study a group described as above is to study the composition of two wave systems:

Imagine that two sinusoidal waves with the same maximum wave heights are progressing along the positive
direction of the x-axis with different advance velocities and wavelengths, and that the wave heights are:

ξ1 = a ⋅ sin(k0 ⋅ x − ωt )

ξ 2 = a ⋅ sin [ (k0 + δ k0 ) ⋅ x − (ω + δω ) ⋅ t ]

where

a = the amplitude of the waves

g
k0 = = wave number
c2

ω
c= = celerity of the wave
k0


ω= = wave period
T
The wavelength is given by


λ=
k0

Because δk0 and δω are small, we may as a good approximation express the resulting wave as:

ξ = ξ1 + ξ 2 = A ⋅ sin(k0 x − ωt )

where
1
A = 2a ⋅ cos ( xδ k0 − tδω )
2

It is from this evident that the amplitude of the advancing sinusoidal wave varies slowly with x and t because δk0
and δω are small units. For a given t value several x values are possible, where the amplitude A is close to 0 or
close to 2a. On the surface, a series of groups separated with almost calm water, like when a stone is thrown in
ω
the water, is observed. Each wave has a celerity equal to U = , while the maximum value of A and therefore
k0
∂ω
the group centre has the velocity U g = . These two velocities can be used to describe the wave pattern
∂k0

102
generated by a ship or a body moving on the surface. Below we have tried to do this by applying an analogy with
sub and supersonic flow in aerodynamics.

The general Free Surface Condition


The relation between the pressure p, the velocity potential φ of the disturbing flow and the
vertical location in the fluid z, in the presence of gravity and a uniform stream, - c is obtained
by Bernoulli' s equation:

⎡ (c + φ z ) 2 + φ y 2 + φ z 2 ⎤ P c2
p
+ ⎣ ⎦ + g ( z + ξ ) = + + gz (3.5)
ρ 2 ρ 2

where

ζ= wave elevation
P is the atmospheric pressure and the reference datum is z = 0

On the free surface, the dynamic condition is that dP = p. If the squares of the perturbation
terms are neglected, the expression for the wave elevation becomes:

1 ⎡ ∂φ ∂φ ⎤ 1 ⎡ ∂ ∂ ⎤
ξ= ⋅ c − = ⋅ c − ⋅φ (3.6)
g ⎢⎣ ∂x ∂t ⎥⎦ g ⎢⎣ ∂x ∂t ⎥⎦

Introduction of the local particle velocities and the kinetic condition (that the velocity of the
free surface along its local normal must be equal to the component of the fluid velocity along
that normal) gives:

∂x ∂φ ∂y ∂φ ∂z ∂φ
= −c + −c; = ; = (3.7)
∂t ∂x ∂t ∂y ∂t ∂z

The general free surface condition is finally obtained and expressed as:

∂ ∂ ∂φ
( − c ) ⋅φ + g ⋅ =0 (3.8)
∂t ∂x ∂z


If the problem is stationary or ⇒ 0 , the boundary condition becomes:
∂t

∂ 2φ g ∂φ
+ ⋅ =0 (3.9)
∂x 2 c 2 ∂z

c
If the Froude number FN = is introduced, it is clear that this parameter is important in
gL
the problem.

If x and z are expressed in non-dimensional form with the unit of length L and c the unit
velocity, the boundary condition now becomes

103
∂ 2Φ ∂Φ
− K0 ⋅ L ⋅ = 0, ( Z = 0) (3.10)
∂X 2
∂Z
where
φ
Φ=
VL
x
X=
L
z
Z=
L
and
L⋅g 1
K0 ⋅ L = 2
= 2
c FN

If FN becomes very small, the first term in the equation for the boundary condition is
eliminated and the following approximation of the boundary condition is made:

∂Φ w
= = 0, ( Z = 0), (3.11)
∂Z c

In the high Froude number range, the free surface changes to the free flow surface where no
action of gravity is considered. Here the component of velocity in the direction of advance is
constant.

This means that in the low Froude number range, the free surface z = 0 acts as if it was a rigid
wall. For the purpose of obtaining the ship geometry from an assumed singularity distribution,
this rigid-wall effect is represented by a mirror image of the assumed singularity distribution.
We may as well study the problem where the geometry of the ship is given and the singularity
distribution must be obtained.

If the complete boundary condition given above is applied instead of its approximation, the
effect of the free surface can be included.

There are many ways to represent the hull in this problem. The singularities can be placed in
with different positions like curved surface distribution, vertical plane distribution or
horizontal plane distributions.

104
3.4.2 Thin ship theory
Michell's linearization in the treatment of the ship-surface condition, which often has been
used, must be mentioned. He related the ship geometry and the corresponding source
distribution directly by:

∂y
m( x, 0, z ) = −2 (3.12)
∂x
where
y = y(x, z) = equation for the ship surface.
As shown by equation (3.12), the source distribution is a function of longitudinal and sources
are distributed only on the centreplane of the ship

Figure 3.7 shows some examples where Mitchell's theory has been applied and comparisons
with model test results have been made. It is seen that fairly good correspondence is obtained
for these hulls, which are fairly slender, low block hulls. Today, a widely used
implementation by Larry Lazauskas of Mitchell’s theory, called Michlet, is available for free
at http://www.cyberiad.net . It is especially well suited for estimation of resistance of
multihulls and very slender monohulls, but has also been used with some success on
conventional monohulls.

105
Figure 3.7 Comparison of measured resistance with results of calculations using
Michells thin ship theory (Weinblum)

106
3.4.3 Estimation of Wave making Resistance by Panel Methods

From: V. Bertram, H. Söding, Jensen, F.M.White, J.P.Breslin, and Poul Andersen

In potential flow and at low speed if the flow is irrotational and a velocity potential exists, it is recalled that:

V = ∇φ
or
∂φ
u=
∂x
∂φ
v=
∂y
∂φ
v=
∂z

The continuity equation ∇ ⋅ V = 0 reduces to Laplace's equation for φ:

∂ 2φ ∂ 2φ ∂ 2φ
∇ 2φ = + + =0
∂x 2 ∂y 2 ∂z 2

The momentum equation

dV
ρ⋅ = ρ ⋅ g − ∇p + µ ⋅∇ 2 ⋅ V
dt
reduces to Bernoulli's equation:

∂φ p 1 2
+ + V + g ⋅ z = const
∂t ρ 2

where V = ∇φ

Typical boundary conditions are known free- stream conditions

∂φ ∂φ ∂φ
Outer boundaries: , ,
∂x ∂y ∂z

• Boundary condition: No velocities normal to the boundary at the body surface:

∂φ
= 0 where n is perpendicular to the body
∂n
When there is a free surface in the problem with a free surface pressure pa, the Bernoulli equation supplies a
relation at the surface between V and the elevation z of the surface. For steady flow.e, g.

2
Free surface: V = ∇φ = const − 2 ⋅ g ⋅ zsurf
2

107
The analyses of Laplace' s equation is very well developed and there are many analytical techniques, including
numerical finite elements and numerical boundary elements. Having found φ ( x, y, t ) from such an analysis, we
then compute V by direct differentiation after which we compute p from the Bernoulli equation.

Boundary Conditions used in the Estimation of Ship Resistance

If we study the transom stern condition, we generally assume that the flow separates and the transom stern is dry.
Atmospheric pressure is then enforced at the edge of the transom stern (z = zT). The inherent assumption is that
the water flows only in longitudinal direction.

φ x + U 2 − 2 gzT = 0

The kinematic condition is fulfilled if the water does not penetrate the water surface z=ζ

∇φ ⋅∇ς = φz

• The dynamic condition is fulfilled if there is atmospheric pressure at the water surface z=ζ

1 p 1
(∇φ ) 2 + g ⋅ z + = U 2
2 ρ 2

• Radiation condition: Waves created by the ship do not propagate ahead. This condition is not valid in
shallow water cases when the flow becomes unsteady and solitary waves are pulsed ahead.

• Decay condition: Far away from the ship, the flow is undisturbed :

lim x2 + y2 + z2 →∞ ∇φ = (−U , 0, 0)

• Open boundary condition: Waves generated by the ship pass unreflected any artificial boundary of the
computational domain

• Equilibrium: The ship is in equilibrium if trim and sinkage are changed in such a way that the dynamical
vertical force and the trim moment are counter-acted

• Bottom condition (shallow-water case) : No water flows through the sea bottom.

• Kutta Condition (for Catamarans/Swaths): At the stern/end of the strut the flow separates

The decay condition substitutes the open - boundary condition if the boundary of the computational domain lies
at infinity.

Potential flow methods are used with special techniques to ensure that the flow separates. However, the point of
separation has to be determined externally. For geometries with sharp after bodies, this is quite simple.

Conditions to be fulfilled in the Calculations

Because the ships are not thin as required in the thin ship theory, panel methods are used instead of the classical
methods that use Kelvin or Havelock sources which fulfils automatically a crude approximation of the dynamic
and kinematic free-surface conditions. Kelvin sources are complicated and require great care in their numerical
evaluation. Rankine sources on the other hand are quite simple.

The potential of a Rankine point source is a factor divided by the distance between the point source and the
considered point in the fluid domain. The factor is called the source strength. The derivative of the potential in
arbitrary spatial direction gives the velocity in that direction and the operation is simple to perform.

108
We recall that a point source with co-ordinates (xq, yq, zq) the distance between source point and field point is
given by

r = ( x − xq ) 2 + ( y − yq ) 2 + ( z − zq ) 2

The potential is given by

−σ
φ=
4π r
The velocity vector is obtained through differentiation with respect to x, y and z:

⎧φ x ⎫ ⎧ x − xq ⎫
⎪ ⎪ 1 ⎪ ⎪
v = ∇φ ⎨φ y ⎬ = σ 3 ⎨
y − yq ⎬
⎪ ⎪ 2π r ⎪ ⎪
⎩φ z ⎭ ⎩ z − zq ⎭

Panel methods divide surfaces into a finite number of "panels" and corresponding number of "collocation
points". A panel distributes Rankine sources with a given distribution over the surface. By a proper adjustment
of the source strengths, the boundary condition is fulfilled at these points with good approximation.

Panel methods are supposed to fulfil:

• Laplace's equation
• The decay condition
• Bottom condition with mirror images of the panels at the bottom for shallow water cases
• Side wall condition through repeated use of mirror images at vertical canals wall

Once a potential has been determined, the forces can be determined by direct pressure integration on the wetted
hull. The forces are corrected by the hydrostatic forces at rest. The force in the x direction corresponding to the
wave resistance is:

F = ∫ p ⋅ n ⋅ dS − ∫ pS ⋅ n ⋅ dS
S S0

The none dimensional wave resistance coefficient is:

F
CW =
ρ
⋅V 2 ⋅ S
2
In these equations:

S = wetted surface

S0 = wetted surface at zero speed

P = hull pressure determined from Bernoulli's equation

pS = hull pressure at rest

n1 = inward normal on the hull

In the same way, moment and sinkage may be calculated. In an iterative process corrections or adjustments are
made and the calculations repeated for each step. This yields the final resistance of the ship.

109
Those who would like to read more about this topic, are recommended to read:

"Practical Ship Hydrodynamics" by Volker Bertram. ISBN 07506-48511 Butterworth and Heinemann 2000 and
"CFD for Ship and Offshore Design" 31 WEGEMT SCHOOL 1999

3.5 A Comparison between Resistance calculated by Theory, Model Tests and


Regression Analyses

The main point in this section is

• to discuss limitations in estimation of resistance of conventional ships applying the


methods described above

• to show that model testing probably is the most accurate method to determine the
resistance of the ship

It has been shown how different statistical procedures can be applied for estimation of ship
resistance, and that these methods will give very different results.

It is of special interest to see how model test results agree with simple theoretical calculations
and calculations based on regression analyses, when the resistance of a given ship is
determined or optimised for different loading conditions.

As an example, consider a ship (from: "Hovedprosjekt 2000 by Sønju, Olafsrud, Rype and
Soma from HIO") with the following characteristics:

LKVL = 174. 559 m


B = 29 m
CB = 0.682
Displacement: 31449.543 m3
Modelscale: 1:75

tested in the towing tank at Tyholt.

The model, which is shown in Figure 3.8, has 3 different bow designs. These bows can easily
be interchanged during the test program.

The test program was as follows:


• Towing tests first without bow, then with a large and a small bow. Tests at different
waterlines and trim angles.
• Estimation of the total resistance with Hollenbach's (1998) program for the same load
conditions as above.
• Estimation of the wave resistance according to Michelet's "thin ship theory".
The results of the comparison of resistance from model test and the two different calculations
are shown in Figure 3.9, Figure 3.10, and Figure 3.11

110
Figure 3.8 Model with exchangeable bulb

Figure 3.9 Model test results with three different bulbs.

111
Figure 3.10 Comparison of model tests and calculations. DWL, small bulb.

Figure 3.11 Comparison of model tests and calculations. WL2, small bulb.

Based on this, we may draw the following conclusions:

1. The model results show that the bulb has large influence on the resistance depending
on trim and draught. This is clear from the figures for ballast and full load.

2. If the test results are compared with theoretical calculations according to Michlet,
bad and unacceptable correlation is obtained as shown in Figure 3.10 and Figure
3.11.

112
3. The tests demonstrated that one should be careful and not rely too much on
procedures like Hollenbach's in the optimisation of bulbs.

"Thin ship theory" has, as we already know strong limitations for full ships. It is therefore
interesting to see what may be obtained with more advanced methods:

From time to time workshops are arranged, where experts discuss and compare their results.
Figure 3.12 shows results from a workshop arranged in 1995. Applying pure theory, the
model resistance was calculated for a tanker model as shown in Figure 3.12 and later
compared with results from model tests. As observed, the correlation between theory and
experiment is not the best. In this example, viscous resistance is the most important part of the
model resistance.

Figure 3.12 Comparison of residual resistance from model test and calculated by
different institutions. (Proceedings of CFD workshop, Tokyo, 1995)

113
On more slender ships like Ro-Ro ships, the wave - resistance will be more dominating than
on fuller ships and a more accurate calculation of the wave resistance is required.

Currently, the wave resistance is usually calculated with more accurate methods than the "thin
ship theory". An example is shown in Figure 3.13. Non- linear theory with two dim Laplace's
equation and a three dim surface condition is applied for calculation of wave or residual
resistance for a Ro-Ro ship. The calculations, which have been made for two conditions, are
compared with model results. It is observed that the correlation between calculations and test
results in this case is good.

Conclusion:
In the following a summary of accuracy in calculation of resistance with different methods is
made:
• When the predictions are based on regression analyses of a material with a limited
number of hull parameters and with a too restricted selection of ships, the speed
prognoses will be inaccurate.
• If the calculations are based on analyses of model test results where the models are of
the same type as the actual ship, prognoses that are more accurate are obtained.
• Theoretical calculations of the wave resistance with modern methods can give reliable
results.
• It is reason to believe that only prognoses based on model tests can give fairly reliable
results in calculation of the total resistance. This is specially the case, when the shape
of the hull is special.
Finally, it should be noted that all large and commercial tanks have large databases also from
testing of models with modern hull shapes. When data of this type is applied in the programs,
optimum hull shapes are determined with better accuracy than indicated for methods
published in the literature.

Figure 3.13 Comparison of experimental and calculation results of residual resistance


for a fast Ro-Pax ship.

114
4 CORRELATION FACTORS FROM TRIAL ANALYSES

The main point in this section is

• to show that it is necessary to apply empirical correlation factors for scaling of model
resistance in spite of the fact that the scaling methods may be advanced. These
factors varies mainly with the details of the scaling method.

The following notions of resistance components are frequent in use:

Rw Component connected to generation of gravitation waves.


Rwb Resistance due to wave breaking.
Rwp Resistance due to wave making, measured from elevation of the surface some
distance from the model.
Rp Pressure resistance due to integration of the pressure along the hull in the direction
of motion.
Rv Viscous resistance due to integration of the shear forces along the hull in the
direction of motion. To this, the part of the pressure force, which is proportional to
CF, is added.
Rpv The part of the pressure resistance which is proportional to CF
Rs Resistance due to spray
RAA Air resistance
RBD Transom stern resistance

It is not simple to separate the different components from each other and to scale them, as
should be clear from the text so far. As an example, the pressure resistance sometimes
manifests itself by vorticity. This resistance is sometimes strongly RN dependent. Other times
it is independent of RN.

Generally, the resistance could alternatively have been expressed like:

CT = CF + a ⋅ CF + b ⋅ CF −0.5 + CDi + CWB + CW 0 ⋅ e − a ⋅ CF + C AA + C App + CBD (4.1)

where

CT = total resistance coefficient


CAA = air resistance
CApp = appendix resistance

115
Other definitions are:

CF Frictional resistance for a flat plate with the same length and speed as the
ship
a ⋅ CF Amount of pressure resistance proportional to CF
b ⋅ CF −0.5 Pressure resistance created in regions of the after body behind forced
separation(base drag)
CDi = d·CF-0.5 Induced resistance due to vortices formed at the shoulders of the after body
CWB Resistance due to wave breaking
CWP Resistance due to wave making (without viscous influence)
CWP ⋅ e − a ⋅ CF Resistance due to wave making (with viscous influence)

This procedure makes the wave making resistance also dependent of CF and not only on FN.
The constants are functions of the shape parameters of the ship. In other words, there is reason
to question the form factor concept.

However, as long as the detection of the components is so complicated and the interpretation
of the scaling laws so unclear, we have to accept simpler scaling laws. A normally accepted
definition of the model resistance is therefore:

CTm = (1 + k )CFm + CRm + C AAm + C Appm (4.2)

where

1+ k = form factor = f (L, B, T, CB etc)


CRm = CWB + CDi + CWP e − a ⋅ CFm
For the ship, the total resistance is:

CTs = ( 1+ k )( CFs + ∆ CF ) + CRm + CAAs + CApps+ CA (4.3)

where

CA = empirical correlation factor.


We have then in reality avoided to scale some of the resistance components separately, and
instead included them in CR. This will have impact on the correlation factor CA:

C A = CFs ⋅ ( d 0 − k ) + CWP ⋅ e − a ⋅ ( CFs − CFm ) + b ⋅ CFs + c ⋅ ( CFs −0.5 − CFm −0.5 ) (4.4)

Only few of the constants are identified. Therefore CA must be established on a pure empirical
basis where only model and trial results have been analysed.

In the following, it is shown how CA is determined from trial results, and how it is used in,
speed prognoses based on model tests.

During the towing tests the resistance RTm is measured at different velocities. During the

116
propulsion tests, the propeller must generate a thrust to balance the resistance:

This thrust is expressed as follows:

Rm
Tm = (4.5)
1− t

where the thrust deduction is:

Rm
t = 1− (4.6)
Tm

This is an expression for the fact that the propeller changes the boundary layer and the
pressure distribution in the after body. It is demonstrated that the working propeller increases
the resistance of the ship.

Normal t values are between 0.05 and 0.30 depending on propeller loading and after body
shape.

During the open water tests, the propeller is normally tested without the rudder. Some times
the rudder is regarded as an integrated part of propulsion system and is placed behind the
propeller during the tests.

During the open water tests the relation between thrust, torque, number of revolutions and
velocity is determined. The velocity is represented by the advance ratio:

Vm
Jm = (4.7)
n ⋅ Dm

where
Vm = advance velocity of the propeller
n = propeller revolutions
Dm = propeller diameter
Thrust and torque is expressed on dimensionless form as:

Tm
KTm = (4.8)
ρ ⋅ n 2 ⋅ Dm 4

and
Qm
K Qm = (4.9)
ρ ⋅ n 2 ⋅ Dm 5

Because Vm , Tm , Qm and n are measured during the propulsion tests, KTm and KQm can be
calculated. Assuming « thrust identity » we may at a given model velocity use the measured
KTm value and enter the open water diagram for the model propeller and take out the
corresponding J0, KT0 and KQ0 values. We will now usually get a slightly different KQ value

117
than measured during the propulsion tests. The ratio is defined as the relative rotative
efficiency ηR. This term is an expression for the fact that a propeller working in a wake field
behind the model and in front of the rudder has a different efficiency than in open water.

KQ
ηR = (4.10)
K Q0

The normal range for ηR is 0.97 to 1.05.

To J0 there is a imaginary velocity:

V0 = J 0 ⋅ nD (4.11)

This velocity is connected to the model velocity Vm in the following way:

V0 = (1 − wm ) ⋅ Vm (4.12)

where the effective wake fraction is defined as:

V0
wm = 1 − (4.13)
Vm

This is an expression for the fact that the propeller behind the ship "sees" a velocity, which is
smaller than the ship velocity. This is due to the boundary layer and the potential flow.
Normal values for w is from 0.10 to 0.40.

Due to the presence of bilge keels, rudder shafts, brackets, domes, openings for tunnels etc,
we get added resistance, termed appendage resistance (see section 2.9). It is difficult to
estimate this type of resistance because each element will to a large extent influence on the
resistance of the other elements. Calculation of the resistance and the effect of the influence of
each small element are complicated and necessarily not very accurate. Only when a single
element is very large, a separate resistance calculation of the element is recommended.

It is therefore normal praxis to tow the model with and without appendices. The difference in
resistance ∆R gives:

∆R
C App = (4.14)
1
ρ ⋅Vm 2 ⋅ S App
2

S App = 2 ⋅ ∑ ln ⋅ hn = area of the wetted surface of all appendages.

hn = equivalent height of a single appendage.


ln = equivalent length of appendage.

It is now possible to define an equivalent length of SAP and a corresponding RNAP

118
S App
LApp =
2

V ⋅ LApp
RNApp =
υ

A different procedure is to calculate the appendage resistance in model and in full scale and to
apply the ratio between the calculated resistances for scaling of the appendage resistance,
determined by model tests.

However, it is not easy to determine how much of CApp which is frictional resistance and how
much which is pressure resistance . An alternative to the method sketched above, is to let
CApp be part of CR and avoid to scale CApp separately. In this case, any scale effect will be
included in the correlation factor CA. See also section 2.9.

4.1 Trial Analyses

During the trials VS, RPM and KQS are measured. The power or the torque may be measured in
different ways. The most reliable way is to apply strain gauges glued to the shaft. From the
model results

K QS
K QOS = (4.15)
ηR

is calculated.

From the propeller diagram applying thrust identity JS and KTS is derived, which makes it
possible to calculate equivalent propeller velocity and propeller thrust:

RPM
VOS = J S ⋅ ⋅ DS (4.16)
60

and the wake:

VOS
ws = 1 − (4.17)
VS

This gives the scale effect of the wake:

∆w = wm − ws (4.18)

The propeller thrust follows from:

2
⎛ RPM ⎞
Ts = KTs ⋅ ρ s ⋅ ⎜ ⎟ ⋅ Ds
4
(4.19)
⎝ 60 ⎠

119
The ship resistance is:

Rs = Ts ⋅ (1 − t ) (4.20)

with the resistance coefficient:

RTs
CT = (4.21)
1
⋅ ρ ⋅ (1 − ws ) ⋅ Vs 2 ⋅ S
2

From the speed prognoses, the following coefficient for the total resistance is obtained:

CTo = (1+ k )(CFs + ∆ CF ) + CR + CAP + CAA + CA (4.22)

If it is assumed that CT = CTo , a value for CA is found.

Ideally CA should have been zero, but due to different ways of treating scale effects and
different ways of doing the tests the final value varies from laboratory to laboratory and will
often be different from 0. Because of this, different ship model tanks have their own values
for CA and ∆w from analyses of trials. Examples are shown in Figure 4.1 and Figure 4.2.

When speed prognoses are made on basis on model tests, one proceeds as described above.
CA and ∆ w for the actual ship is taken from analyses of trials for ships, which are similar to
the actual ship. CT0 and CT are calculated and finally the propeller thrust Ts.

Then wS = wm - ∆ w and VS are calculated and applying the propeller diagram, the RPM and
the power for different velocities are determined.

In speed prognoses, it is important to use an open water diagram corrected for scale effects. If
scale effects are present, the RPM will not be correct. The trial analyses will also give
incorrect CA and ∆w values. In the propeller compendium it is shown how corrections for
propeller scale effects are made.

Figure 4.1 Correlation factors

120
Figure 4.2 Wake scale effect, found from correlation studies.

121
122
5 RESISTANCE IN SHALLOW WATER

Mainly from: Stokes, Lamb, Havelock, Wigley, Horn, Schlichting, Van Lammeren, Clay/ Medwin and Wiegel

The main point in this section is

• that the resistance may increase suddenly if the depth is suddenly reduced and that
the increase in resistance is connected to a strong change in trim and immersion.

• that certain combinations of speed and depth should be avoided to keep resistance,
change in trim and draught at a minimum.

Up to now, resistance has only been discussed if limitations like depth, channel walls or the
presence of other ships are absent.

The problem of wave making in shallow water is of increasing importance in practice. With
the general increase of ship speed, water ways which could formerly be considered very deep,
must now be transferred to the category of restricted depth. For many a fast ship the North
Sea with a depth of 40 meters has become a shallow waterway, whose influence on resistance
and consequently on speed can clearly be observed.

If the depth is limited, the resistance may for given combinations of velocity and depth
change radically, as shown in Figure 5.1. The same is valid for trim and depth.

Figure 5.1 Effect of shallow water on wave resistance (Havelock, 1908)

123
The resistance reaches a maximum at a depth Froude number of around 1.0. The depth Froude
number is based on depth instead of ship length:

V
FNh = (5.1)
g ⋅h

where
h = depth.
In Figure 5.2 the change in residual resistance with FN and h/L is shown for a particular ship.
Other ships have a similar trend. If the correct h/L value is used in FN, it is seen that the peaks
are close to FNh.=1.0. It should also be noticed that at some velocities, the shallow water effect
reduces the resistance.

Figure 5.2 Residual resistance for a ship in deep and shallow water (Graff, Kracht and
Weinblum)

It is important to have the shallow water effect in mind in waters like the North Sea and the
Baltic. FNh.=1.0 gives the following critical velocities:

h (m) 5 10 15 20 25 40
Vs (knots) 13.62 19.25 23.58 27.23 30.44 38.51

This example shows that many resistance peaks in the actual region are possible for example
for large and small catamarans.

This increased resistance is related to changes in the wave pattern, as shown in Figure 5.3.
The most important parameter in the study of wave resistance in shallow water is as
mentioned the depth Froude number FNh (see equation (5.1)).

124
Figure 5.3 Wave patterns caused by a point disturbance of pressure advancing over
water of finite depth (Havelock, 1908)

125
Referring to Figure 5.3, it is normal to distinguish between three regions:
1. For unlimited depth when FNh < 0.7 both diverging and transverse waves are generated.
Were these two sets of waves intersect, a large superimposed wave is noticed. The
angle formed by a line through the intersections (cusp locus) and the centreline θ
depends upon FNh, being 19.48 degrees.
2. Above Fnh = 0.7 the angle which originally was 19.47 degrees, changes successively
and reaches finally at Fnh = 1.0 the angle 90 degrees. The wave follows the craft and is
not lacking behind.
3. When Fnh > 1 the angle is reduced again and only diverging waves remain. This is the
supercritical region, contrary to the sub critical region for Fnh < 1.
The conditions are analogous to the conditions for sound waves. The same equations are valid
except that the Mach number V/a where a = speed of sound, is replaced with the Froude
number FNh.

Havelock (1951) was one of the first who investigated the wave resistance of a ship with
varying depth. He got results as shown in Figure 5.1. It is observed that a peak is present at h
/ L = 0.75 and at FN = 0.86. That corresponds to Fnh = 1. It is observed that the wave
resistance, for this case, is smaller than with unlimited depth.

The safest way to determine the change in resistance is by model testing. However, it is also
possible to determine the change due to shallow water by theoretical calculations.

Example of determination of resistance at limited depth:

If the wave resistance for deep water is known, it is possible to estimate the resistance for the
same craft on limited depth.
According to the "thin-ship theory" when Fnh >1 , the wave resistance can be expressed as:

ρ ⋅U 2 1
⋅ ∫ ⎡⎣ S ∗∗ ( x ) ⎤⎦ ⋅ dx
2
Dwh = (5.2)
2h ⋅ Fnh 2 − 1 0

where
S**(x) = the derivative longitudinally of cross section areas.

Hofman and Kozarski (1999) estimated the wave resistance for a Wigley hull where the
draught was T/L = 0.05. The results from the calculations at different FN and h/L ratios are
plotted in Figure 5.4. The results are presented in curves showing the ratios between wave
resistance on restricted depth and wave resistance on unrestricted depth. The calculations
show that it is important to avoid the dark fields in the left part of the diagrams where the
resistance may increases strongly.

As an example, the depth where the wave resistance is twice as large as for unrestricted depth
is searched for. The diagram shows that this requirement is fulfilled when h/L = 0.23 and
V
FL = = 0.435 . For different ship lengths the following depths and velocities are
gL
obtained:

126
L (m) 25 75 100 150
h ( m ) critical 5.75 17.25 23.00 34.50
V (m/s ) critical 13.24 22.94 26.49 32.44

If 0.1 < h/L < 0.5 the borders for the critical region and the maximum value for the resistance
Rw (h)
ratio rmax = can be represented by the polynomial:
Rw (h = ∞)

h L h L h L
FL = a1 + a2 ( ) + a3( ) + a4 ( ) + a5 ( ) + a6 ( ) + a7( ) (5.3)
L h L h L h

where the constants follow from:

a1 a2 a3 a4 a5 a6 a7
FL(low) 0.21757 0.04229 0.04143 0.17549 -0.00521 0.27237 0.00017
FL(kr) 0.68693 -0.03774 -0.09583 0.04807 0.01033 0.27060 -0.00045
FL(upper) 0.55786 0.36675 -0.04949 0.12859 0.00636 0.01695 -0.00024
rmax 49.396 -91.456 -11.268 -63.489 1.0062 0 0

All that is needed in order to apply the diagram in Figure 5.4 is the ship velocity V and the
length L together with the depth h. Investigations indicate that the diagram may also be
applied for other T/L values and ship shapes than those used in the calculations.

The diagram in Figure 5.4 is divided in three parts:

1. Sub-critical region, where most of the conventional ships operate

2. Super-critical region where the resistance may become smaller than in the sub-critical
region

3. The critical region, where resistance and change in trim and sinkage are at a maximum.

The diagram is useful at an early stage in the design of ships for rivers and lakes and
operation of vessels in shallow water..

The diagram shows clearly that the shallow water effect:


• can be neglected in the sub critical region
• may even be beneficial in the supercritical region
• is undesirable in the critical region
Because of this, (if it were possible) the hull should be designed for the sub critical or the
supercritical region. If forced to operate in the critical region, the ship should operate so far at
the right as possible. The left side of the diagram should under all circumstances be avoided.

127
Figure 5.4 Increase in resistance due to shallow water (Hofman and Kozarski)

128
5.1 A closer Study of Wave Patterns near a Ship in Shallow Water at different
Velocities

The main point in this section is

• try to explain why the wave pattern near a ship in shallow water has an appearance
which depends on the Froude's number

• to go more thoroughly into the meaning of critical, supercritical and sub critical
waves

To better understand the wave pattern generation near a ship in shallow water, it is useful to
apply the definitions group and phase velocity, where group velocity is the velocity of the
energy in the wave system.

From the elementary wave theory, we recall the following relations:

The wave number, the number of waves per unit distance along the x-axis is:


k= (5.4)
λ

where λ is the wavelength

If the depth is finite, it follows from the boundary condition at the bottom and the free surface
condition, that the phase velocity becomes:

gλ 2π h
U ph = ⋅ tanh( ) (5.5)
2π λ

If kh 1 this tends to the phase velocity for deep water:

g λ∞
U p∞ = (5.6)


If h 1 we enter a regime where the phase velocity depends only on the depth.
λ

In water with finite depth h, the group velocity is:


h
1 λ
U gh =( + ) ⋅ U ph (5.7)
2 sinh( 4π h)
λ

129

In deep water ( h 1) the group velocity reduces to:
λ

1 gλ 1
U g∞ = = U p∞ (5.8)
2 2π 2


In shallow water ( h 1) the group velocity becomes:
λ

U gh = U ph

Velocities according to the formulas above are calculated and presented in a table as functions
of:
h 1
= (5.9)
λ∞ Fnh ⋅ 2π
2

h 1
= 0.000 0.100 0.200 0.300 0.400 0.500
λ∞ Fnh ⋅ 2π
2

U ph
0.000 0.700 0.900 0.950 0.975 1.000
U p∞
U gh
0.000 0.590 0.595 0.550 0.525 0.500
U p∞
U gh
1.000 0.843 0.661 0.579 0.539 0.500
U ph

Above, it is observed that the group velocity in deep water is half the phase velocity. This is
illustrated in Figure 5.5, where a sinusoidal wave is followed at regular intervals. The wave is
moving from left to right after it has been "switched on" and then suddenly "switched off" in a
wave tank. The line to the right, marks the front of the wave system but also the energy or the
group velocity. The line to the to the left marks the point where the wave ceases to exist.

130
Figure 5.5 Illustration of group velocity (Newman, 1977)

Similar conditions exist for ship waves:

Imagine that the waves are generated by advancing sink or sources moving with constant
speed U. The corresponding group velocity of the waves is:

U
Ug =
2

We have also shown the wave pattern for critical, sub critical and super critical velocity
without explaining why the wave patterns have their specific shape. In order to understand
this we may regard our disturbance as fixed at its origin, with a uniform current U flowing
past. In this frame of reference the waves, which move with the ship, are stationary, so ω = 0
and

g
U ⋅ cos χ = =Up
k

Where χ is the angle with respect to the x axis of the progressive wave due to the pressure
point. In other words:

A system of plane progressive waves moving at an oblique angle χ with respect to the x-axis
will appear steady state to an observer moving along the x-axis with velocity U p sec χ

131
Figure 5.6 Wave crest generated by a moving pressure point (Stokes "Water waves")

Imagine a point of disturbance starting at A as indicated in Figure 5.6. After a given time the
disturbance has moved to point B. The disturbance sets up transverse and diverging waves as
shown in Figure 5.3 where the waves with their special pattern move with the point.

As mentioned in section 2.6 the transverse and diverging waves intersects on a "ray" or line at
an angle θ with the direction of advance – the so-called Kelvin Angle. In deep water this
angle is 19°28. The crest tangent of the transverse wave has an angle χ to the same direction
as shown in Figure 5.6. The "rays" are straight lines radiating from the origin of the
disturbance. The waves must travel to the left to balance the flow U to the right. Therefore the
angle becomes:

U gy
tan θ = (5.10)
U − U gx

where
U gx = U g ⋅ sin χ
U gy = U g ⋅ cos χ

and finally:
1
sin χ ⋅ cos χ
tan θ = 2 (5.11)
1
1 − cos 2 χ
2

χ 20 30 32.50 35.27 37.50 40 45


θ 16.05 19.11 19.37 19.47 19.41 19.21 18.43

Thus, the significant waves, moving in a direction χ , are situated along the "ray" or the radial
line:

132
y
= tan θ = tan19 28
x

From the table above it is clear that the maximum value of θ occurs when χ is 35.27 . The
waves are confined to a sector symmetrical about the negative x-axis, with included semi
angles of 19°28. Within this sector, for a given point (x, y) and corresponding ratio y/x, two
solutions of the equation exist; hence two waves with distinct angles χ move in different
directions. No waves exist outside the sector or the wedge.

If we return to Figure 5.6 and accept that there are significant waves as described above and
this time also accept that wave velocity which is gh , we may have different ratios between
group velocity and phase velocity ranging from 0.5 to 1.0. This gives:

Vg
2 1
sin θ = =
Vg 2
Vp − −1
2 Fnh

If Fnh < 1

θ values according to this formula are presented in Figure 5.3 and the following table. In
water of finite depth, the most noticeable feature is that as depth h reduces or the speed
U increases, increasing the wavelength, the wedge angle increases until at
U
Fnh = =1
g ⋅h

it reaches 90 . This is another form of Froude number which says that the ship speed U has
reached the critical speed gh which is the maximum possible speed of gravity waves, and
where the group velocity Ug = U, the face velocity.
At lower speeds, the wave energy is falling behind the ship. Viewed from the ship the wave
crests have zero crest speed. Looking straight back, the transverse waves are travelling at
speed U towards the ship and being swept back at speed U by the current. Their energy is
travelling at speed Ug<U towards the ship but is also being swept back at speed U by the
current, and so falls back, allowing a train of waves to form. As U reaches gh so Ug reaches
U and the transverse wave energy builds up to a maximum near the ship giving an almost
two-dimensional wave but not leaving anything downstream. If U now increases further,
transverse waves cannot keep up with the ship, since they would be beyond their fastest
possible speed, and so there are no transverse waves. There is still a diverging wave pattern
and as U continues to increase the wedge angle decreases again. The wedge angle is easily
calculated from the Doppler relation:

U gy
tan θ =
U − U gx

133
We now assume that U cos χ = gh and U g = gh , which gives:

gh ⋅ sin χ ⋅ cos χ sin χ


tan θ = =
U − gh ⋅ cos χ 2 U
− cos χ
gh

Introducing
U
Fnh =
gh

we have:
1
cos χ =
Fnh

and finally:
1
tan θ =
2
( Fnh − 1)

Thus for a given Fnh there is only one value of tanθ and only one line or ray on which
1
θ = π − χ . The wave crest is actually along this line or ray. With a restricted water depth h,
2
the speed of progressive waves U can never exceed the speed Us of the wave of translation.
This speed is independent of the wavelength and is completely defined by the water depth,
which corresponds exactly with practical observation.

In terms of analogy with compressible gas flow, the wave crests are Mach lines and Fnh is the
Mach number, see Figure 5.7.

An equation for shallow water waves that satisfies the condition on the free surface and
Laplace' s equation is the following linearized equation:

∂ 2φ (0) ∂ 2φ (0) 1 ∂ 2φ (0)


+ − ⋅ =0
∂x0 2 ∂z0 2 gh ∂t 2


If the motion is regarded as a steady-state motion in a moving reference frame, is replaced
∂t
U∂
by - and the linear equation above takes the form
∂x

∂ 2φ (0) ∂ 2φ (0)
(1 − Fnh 2 ) ⋅ + =0
∂x0 2 ∂z0 2

where Fnh is the Froude number based on the depth h.

134
In the solution of this equation there is a solution for the sub critical flow Fh<1 and one for the
supercritical case Fh>1. Physically this corresponds respectively to the aerodynamic
conditions of subsonic and supersonic flow if the Froude number is replaced by the Mach
number U/a because this equation is also the equation governing the linearized steady state
two dimensional flow in aerodynamics, with compressibility effects represented by the speed
of sound which is a.

Imagine now that this analogy is applied to explain what happens in shallow water and it is
referred to (a) in Figure 5.7 where a point representing the vehicle is radiating a wave
spreading with the speed of sound. After a certain time step dt, the wave front has got the
radius a dt. In meantime, the point with velocity U has travelled over a distance U dt. Since U
< a, a picture as on (a) in Figure 5.7 is obtained. We are still in the sub critical region where
the V shape depends on U/a.

Figure 5.7 Sound waves (Mach)

Figure 5.7 (c) an example of supercritical flow is shown. If we apply the relation given above
between θ and Fnh, we get the following values for θ and Fnh, which are also given in Figure
5.3:

135
Fnh θ (degrees)
0.38 19.47
0.42 19.47
0.50 19.48
0.55 19.50
0.60 19.62
0.70 20.30
0.82 23.70
0.92 39.32
0.96 59.45
0.99 78
1.00 90
1.005 84
1.41 45
1.73 35
2.00 30
3.00 19.47

In a towing tank, the depth is approximately 5 to 6 meters. The maximum speed of the models
is approximately 10 m/s. A consequence of this is that speed prognoses based on results from
tests in a towing tank should be made with carefulness.

From Figure 5.1, Figure 5.2 and Figure 5.4 and the few examples below it is evident that:
• Fast ships with lengths up to 40 meters operate at a FN number close to 0.90 and at depths
between h/L= 0.10 and 0.30, where the influence of shallow water is marked. However, if
the model tests shall be relevant for deep water the test should be made at h L ≥ 2.0
which means h ≥ 6 − 8m . In other words: The model tests are not always 100%
representative for the service condition.
• A fast ship with a length of 150 meters and a speed of 50 knots operates at FN = 0.67. At
this speed the critical depth is 67.4 which gives h/L = 0.45. This shows that a larger an fast
ship often may operate near the critical speed in full scale. In model scale the depth in a
normal tank is about 5.5m, which gives h/L=0.33. If we study Figure 5.1, it is indicated
that tests in the towing tank are valied for shallow water. However if it is required that the
model tests shall be valid for "deep water", this can not be fulfilled.
• A frigate has a length of 100 meters and does often operate at h/L=0.23 and FN=0.435. A
normal model length would be 6 meters (h/L =0.92). As seen from Figure 5.2 and Figure
5.4, the wave resistance may be doubled, due to the shallow water effect.

In cases where both Fnh and FN >1.25, the transverse waves are not present any more. Figure
2.32 indicates that the wave resistance in this case is due to divergent waves and decreases
with FN. The resistance due to the transom stern becomes more important.

136
6 DIRECT MEASUREMENT OF RESISTANCE
COMPONENTS

the main point in this section is

• to show how it is possible to measure the wave resistance, viscous resistance and
resistance due to wave breaking separately, using special measurement techniques.
The waves generated by the ship are distributed in the direction of motion (transverse waves)
and out to the sides (divergent waves) (see Figure 2.31). At low and moderate speed most of
the wave energy is in the waves following the ship (transverse waves). If the speed is high,
most of the energy is in the divergent waves.

It is possible to measure velocities and static pressure at different distances from the surface
and out to the sides behind the ship. The measurements can be made with a pitot rake placed
at a distance equal to 0.5 L behind the model, as shown in Figure 6.3 and below. During tests
with a conventional model, measurements are normally made in a depth down to 1 m and in a
width of about 3 m.

In such measurements, one control plane is placed in front of the model and one control plane
is placed behind the model. The planes follow the model as indicated in Figure 6.1.

Figure 6.1 Definitions for wave resistance measurement

The total resistance can, as already discussed, be expressed by the change in pressure and
momentum between two control planes, one far in front and one behind the model:

RT = ∫∫ (( p + ρ ⋅ u ⋅ (u − V )) ⋅ dy ⋅ dz − ∫∫ (( p + ρ ⋅ u ⋅ (u − V )) ⋅ dy ⋅ dz (6.1)
A B

where A is the control plane far in front of the model while B is the control plane behind the
model.

137
If the total resistance is known, and the wave resistance is wanted, the viscous resistance is
subtracted from the total resistance. The viscous resistance follows from measurements of the
total pressure between the control planes. This is done applying Bernoulli's equation, as
indicated in Figure 6.2.

Figure 6.2 Measurement of total head

The definition total head has been applied and is defined as:
ρ
H = p+ρ⋅g⋅z+ ⋅ ⎡⎣ (u − V ) 2 + v 2 + w2 ⎤⎦ (6.2)
2

where u, v and w are velocities induced by the hull in the x, y and z directions and V is the
velocity of the ship.

A has a position so far in front of the ship that these velocities are zero or close to zero.

The total head far in front of the ship is then:

ρ
H0 = ρ g z + V2
2

which gives the following total resistance:

ρ
RT = ∫∫ ( H 0 − H ) ⋅ dy ⋅ dz + ρ ⋅ g ⋅ ∫∫ z ⋅ dy ⋅ dz +
2 ∫∫
((v 2
+ w2 − u 2 ) ⋅ dy ⋅ dz (6.3)
B B B

The total resistance is according to this, expressed by the velocities induced in the control
plane behind the hull. The second link is transformed in such a way that the total resistance
becomes:

ρ ρ
RT = ∫∫ ( H 0 − H ) ⋅ dy ⋅ dz + ⋅ g ⋅ ∫ ζ 2 ⋅ dy + ⋅ ∫∫ (v 2 + w2 − u 2 ) ⋅ dy ⋅ dz (6.4)
B
2 B
2 B

138
where ζ is the wave height behind the boat in control plane B at the breadth y
It is assumed that the flow on the outside of the viscous wake is free from rotation, and that H
therefore is constant along the streamline (and that Bernoulli's equation can bee applied).

Inside the wake, the flow has rotation. For such cases it is assumed that the width of the wake
1 1

is proportional to x 2 . Maximum wake (maximum back flow velocity) is proportional to x 2 .

With these assumptions:

∫∫ ( H
B
0 − H ) ⋅ dy ⋅ dz =ρ ⋅ V ⋅ ∫∫ uτ ⋅ dy ⋅ dz
B
(6.5)

in the wake where:

τ
uτ = = friction − velocity and a measure of formation of turbulence and
ρ
transmission of momentum due to turbulent fluctuations in the wake
τ = shear stress
The viscous resistance is expressed as:

RV = ∫∫ ( H 0 − H ) ⋅ dy ⋅ dz (6.6)
B

and the wave resistance as:

ρ ρ
RW = ⋅ g ⋅ ∫ ζ 2 ⋅ dy + ⋅ ∫∫ (v 2 + w2 − u 2 ) ⋅ dy ⋅ dz (6.7)
2 B
2 B

H and Ho are measured in different ways. It is for example normal to measure H directly as
shown in Figure 6.2, where the total head is measured directly applying a Prandtl tube.

The wave pattern resistance can be measured by measuring the wave elevation in a
longitudinal cut along the side of the ship, as sketched in Figure 6.10.

It is also possible to determine the viscous resistance by measuring the velocities and static
pressure at different distances from the surface and in the breadth behind the model. The
measurements are made with pitot rakes at approximately 0.5 L behind the model as shown in
Figure 6.3.

139
Figure 6.3 Measurement of viscous resistance using a rake of pitot tubes. The vertical
rake of tubes must be traversed across the width of the wake.

As shown before, it is possible to express the viscous resistance with the total pressure behind
and in front of the model as:

RV = ∫∫ ( H 0 − H ) ⋅ dy ⋅ dz (6.8)
B

where H and Ho can be measured in different ways but it is normal to measure H directly
without measuring pressure and velocity separately.

The viscous resistance is registered as a change in pressure and momentum between an


undisturbed point in front of the model and a point in the control plane behind the model. The
plane is placed so far behind the model that the static pressure no longer is influenced by the
local flow.

Figure 6.4 shows as an example results from such measurements for a slender model with a
sharp forebody and for a full tanker model. It is observed that two types of curves are present.
The tanker model has a top in the middle but also one top on each side. These tops on the
sides are not due to the viscous resistance. They contain the energy in the breaking wave, as
indicated in Figure 6.5.

140
Figure 6.4 Results of wake traverse measurements of two different ship types (Baba)

When the wave is breaking, the energy is transformed in such a way that it may be registered
as false viscous resistance as shown in Figure 6.5.

Two components of the viscous resistance are therefore identified:

Rv = Rv 0 + Rv1 = ∫ ( H 0 − H ) ⋅ ds + ∫ ( H 0 − H ) ⋅ ds (6.9)
A B

The first integral is valid for the central part, while the second integral is valid for the region,
which is influenced by the wave breaking.

Figure 6.5 Division of wake zone due to viscous resistance and due to wave breaking

141
The last component may become large but is not proportional to CF. It is only dependent on
FN. In resistance calculations, this component must therefore be added to Cw if Cw has been
measured by wave cuts. Figure 6.6 and Figure 6.7 show that good correlation is obtained
between resistance from towing of the model and resistance obtained by summing up the
components measured separately.

Figure 6.6 Resistance components of a slender ship

Figure 6.7 Resistance components of tankers in ballast.

At MARINTEK, several tankers were tested in a research program on energy saving in


shipping. Different fore and after bodies were tested and the wave resistance Rw and the
components Rvo+Rv1 measured. Two of the hulls are shown in Figure 6.8 and Figure 6.9,
together with the wake measurement results and derived resistance. It is possible on both
models to recognise the shape of the waterlines or the frames in the after body. It is observed
that Rw+Rvo+Rv1 corresponds well with measured total resistance.

142
6

CT
5.5
Resistance coefficients *1000

CW
5

CV CV
4.5

3.5

CF
3
0.14 0.15 0.16 0.17 0.18 0.19 0.2 0.21 0.22
F N [-]

Figure 6.8 Frame shape, wake measurement results, and calculated resistance for ship
model M32 (Marintek)

143
4.8
CT
4.6

4.4
Resistance coefficients *1000

CW
4.2

4
CV CV
3.8

3.6

3.4

3.2
CF
3
0.14 0.15 0.16 0.17 0.18 0.19 0.2 0.21 0.22
F N [-]

Figure 6.9 Frame shape, wake measurement results, and calculated resistance for ship
model M49 (Marintek)

144
6.1.1 Measurement of the Wave Pattern and Estimation of the Wave Resistance.

We have seen that the wave resistance increases with the square of the wave height, and seen
that it is possible to measure the wave resistance by measuring the wave pattern of the ship or
the model. In the following, it will be shown that this is also the case if the flow is three dim.
Figure 6.10 shows arrangements for such measurements in three- dim flow.

Figure 6.10 Longitudinal cut measurement of wave pattern resistance

According to Havelock (1953), a three- dim wave system created by a ship may be expressed
as:
π
2
ξ ( x. y ) = ∫π S (θ ) ⋅ sin ⎡⎣ K

0 ⋅ sec 2 θ ⋅ ( x − x0 ⋅ cos θ + y ⋅ sin θ ) ⎤⎦ ⋅dθ
2
π
2
+ ∫ C (θ ) ⋅ cos ⎡⎣ K 0 ⋅ sec 2 θ ⋅ ( x − x0 × cos θ + y ⋅ sin θ ) ⎤⎦ ⋅ dθ (6.10)
−π
2

where
ξ ( x, y ) = wave height at a given point (x, y)
θ = angle between the wave direction and the longitudinal axis of the ship. See Figure
2.30.

145
S (θ ), C (θ ) = amplitude functions (or constants) determined from analysis of a
measured wave system and dependent on the shape of the waterlines of the ship.
g
K0 =
V2
g = acceleration of gravity

It is also possible to calculate these functions theoretically for a given ship shape.

Assume two planes one in front of ship and one behind the ship perpendicular to the direction
of the ship's course. The plane A is far ahead of the ship. The plane B is far in the rear. At a
given time, the wave system is crossing the plane B. After a time step ∆ t the wave system has
moved to the plane A.

The increase in energy in the fluid between the two planes is equal to an increase in wave
energy corresponding to the increase in the area of the free-wave pattern. By introducing the
distance ∆x = V ⋅ ∆t between the two planes, Havelock by a lengthy calculation obtained an
expression for the time average of this energy as:

π
2
cos3 θ
E ⋅ ∆x = π ⋅ ρ ⋅ V 3 ⋅ ∆t ∫ A2 (θ ) ⋅ ⋅ dθ (6.11)
−π 1 + sin 2 θ
2

where
A2 (θ ) = S 2 (θ ) + C 2 (θ )
is the square of the wave amplitude.

This increase of energy is attributable to the work W1 performed by the ship as it advances
against wave-making resistance, RW:

W1 ⋅ ∆t = RW ⋅ ∆x = RW ⋅ V ⋅ ∆t

The second source, W2 is the work performed through vertical plane B, by the fluid through
the rear plane. W2 is the energy transfer due to the wave motion. The values of W2 and energy
E changes with time according to the phase of the wave. The time average is expressed as
follows:

π 2
cos5 (θ )
W 2 ⋅ ∆t =
2
⋅ ρ ⋅ V 3 ⋅ ∆t ⋅ ∫
−π
A2 (θ ) ⋅
1 + sin 2 (θ )
⋅ dθ (6.12)
2

From the law of conservation of energy:

E ⋅∆x = W1 ⋅ ∆t + W 2 ⋅ ∆t

the wave resistance is expressed in the simple form:

146
π

W π
2 2
RW = E − = ⋅ ρ ⋅ V 2 ⋅ ∫ ⎡⎣ S 2 (θ ) + C 2 (θ ) ⎤⎦ ⋅ cos3 θ ⋅ dθ (6.13)
V 2 0

It is observed that the wave making resistance is obtained by an integration of the quantity
A2(θ) weighed by the quantity cos3θ. This shows that the wave making resistance increases
little when θ varies. It is also clear that mainly the amplitude of the transverse wave
determines the wave making resistance. As the diverging wave, especially at a large θ-value,
has a small wavelength, it is more visible than the transverse wave when ship's waves are
observed from above downward vertically. This is because of the fact that the diverging wave
has a steeper slope than the transverse wave. This sometimes tends to give an erroneous
impression about the relative importance of these two waves with respect to wave-making
resistance.

It is now possible to connect the visible ship's wave directly to the wave-making resistance
and to compare theoretical computations with experimentally observed ship's waves.

147
148
7 METHODS FOR RESISTANCE REDUCTION

7.1 Introduction

Also when discussing resistance reduction, it is appropriate to divide the resistance in viscous
and wave resistance.

Reduction of viscous resistance can be done by:

• Reduction of wetted surface


• Reduction of frictional resistance
• Reduction of viscous pressure resistance
Reduction of wetted surface by shaping the hull is usually not something that one should
consider, as little can be obtained in this way. For a ship of a given size, it is more important
to determine the shape for minimisation of wave resistance than for minimisation of wetted
surface. Another way of reducing the wetted surface is to introduce air cushions and air
pockets. This is traditionally used only for high speed vessels such as SES Air cushion
vehicles. It has also been proposed to create air pockets under the flat bottom of for instance
tankers, but few (if any) such projects has been realised.

Reduction of frictional resistance is mainly done by minisation of the roughness.


Minimisation of roughness on a new ship is a matter of high quality workmanship (flush
welding seams and good paintwork), and selection anti-fouling paint system. Minisation of
roughness on sailing ships is important for ship operators, and optimisation studies of docking
intervals and paint systems have been performed repeatedly. Some information on this is
given in section 7.4.

Reduction of viscous pressure resistance is a question of good ship design. To minimise the
viscous pressure resistance, one should take care to:

• Align bilge keels and propeller shaft brackets with the local flow direction
• Take care to avoid flow separation in the after body.
• Avoid sharp corners and flow across knuckles
• Minimise the submerged transom stern area
The three first mentioned items are typically things one should look for in a model test, and
where a streamline paint test is the right tool. Increasingly, numerical flow calculations can
also be used for orientation of bilge keels and propeller shaft brackets. To avoid flow
separation in the afterbody, the hull shape can’t be too full in the stern.

Reduction of wave resistance is often where most can be achieved. At an early stage in the
design, a lot can be gained by proper selection of main dimensions. One should remember that
the wave resistance has a major hump around FN=0.5, so this FN-range should be avoided if
possible. If it is not possible to avoid this FN-range, the hull must be very slender, so that in

149
practice a multihull configuration must be used.

After the main dimensions are fixed, one is left with optimisation of the details of the hull
shape. This is treated in a general (and fairly theoretical manner) in section 7.2. Often, the
most important detail to optimise is the bulb design. This is discussed from a more practical
viewpoint in section 7.3.

7.2 How to reduce the Wave Pattern Resistance of a Ship?

By following the outline given in chapter 6, it is possible to connect the visible ship's wave
directly to the wave-making resistance. This is convenient when making corrections of the
hull form and when comparing the theoretical computations with experimentally observed
ship's waves.

The expression for the wave resistance presented above indicates that the wave resistance
could be reduced considerably by adding a new system to the original system. This system
must produce the amplitude functions S(θ) and C(θ) equal to the original ones, but with
opposite signs. All elementary waves of the ship and the new system must be in inverse
phase. If the amplitude function of the main hull is denoted by A (θ) and the function of the
new system by B (θ), we must have

π
A(θ ) = B(θ ), ( θ = 0 ≈ ) (7.1)
2

If this condition is satisfied in the rear of the ship, the free wave pattern of the bow-wave
system of the main hull and the wave system of the bulb (new system) are superposed, we get
the following wave height:

π
2
ζ F + D ( x, y ) ∼ ∫ [A
−π
F (θ ) − BF (θ ) ] sin
(7.2)
2

⎡ K 0 sec2 θ ⋅ ( x − x0 cosθ + y sin θ ) ⎤ dθ


⎣ ⎦

The requirements for B(θ) are satisfied for a doublet representing a bulb. The resistance of the
resultant wave system produces a wave-making resistance as expressed by the following
equation:

π 2

π 2
RW , F + D =
2
⋅ ρ ⋅V 2 ⋅ ∫π [ A

F (θ ) − BF (θ ) ] ⋅ cos3 θ ⋅ dθ (7.3)
2

It is a well-known fact in primary hydrodynamics that a certain three-dimensional doublet can


replace a sphere placed in a uniform flow with its axis directed against the flow. This concept
has also been accepted in the bulb treatment without any question.

It can be shown that that the free-wave pattern caused by a travelling point-doublet is a sine-
component wave with a negative sign, which means that this free wave of a doublet starts

150
with a trough. By matching AF and BF correctly, the total resistance may be minimised as
exemplified in Figure 7.1.

Inui investigated a given ship, first with and then without bulbs. The wave making resistance
for the ship is shown in Figure 7.1. Figure 7.2 shows the amplitude functions AF of the main
hull and BF of the bulb wave and the integrand functions on the right side of the equation for
the resistance. Figure 7.3 shows the wave profiles. It is observed that the bulb has reduced the
wave pattern as well as the resistance considerably.

Figure 7.1 Wave making resistance for a ship with different bulb configurations (Inui)

151
Figure 7.2 Amplitude functions versus Froude number (Inui)

Figure 7.3 Measured and calculated wave profiles, with and without bulb (Inui)

152
Baba (1973) followed ideas by Hogner (1938) and Everest (1970) and used a "thin ship"
which was added to a given basic hull form so as to reduce the total wave - making resistance.

The pattern of the "thin ship" was calculated theoretically and superposed linearly on the
measured wave- pattern of the basic hull.

Everest had pointed out that analytical studies of measured wave data indicated regions on the
hull where modifications are effective in reducing wave - making resistance. On the bow of a
high-speed liner form, he calculated the effect of the addition of small displacement volume
on wave - making resistance of a basic hull. This is also the principle in Baba's work. Both
Everest and Baba apply Hogner's so called "influence lines".

We have seen that it is possible to express the wave pattern far behind the ship by:

π π
2 2
ξ ( x. y ) = ∫ S (θ ) ⋅ sin [ x, y ) ] ⋅ dθ + ∫ C (θ ) ⋅ cos [ x, y)] ⋅ dθ (7.4)
−π −π
2 2

and to calculate the wave resistance by:

π 2
RW = ⋅ ρ ⋅ V 2 ⋅ ∫ ⎡⎣ S 2 (θ ) + C 2 (θ ) ⎤⎦ ⋅ cos3 θ ⋅ dθ
2 0

where

ξ ( x, y ) = wave height at a given point (x, y)


θ = angle between the wave direction and the longitudinal axis of the ship. See Figure
2.30.
S (θ ), C (θ ) = amplitude functions (or constants) determined from analysis of a
measured wave system and dependent on the shape of the waterlines of the ship.
If the wave amplitude functions a (θ) and b (θ) for the thin ship is introduced, this results in a
modification of the wave pattern (wave heights), which gives the following modification of
the wave resistance:

π 2
⋅ ρ ⋅ V 2 ⋅ ∫ ⎡{S (θ ) + a (θ )} + {C (θ ) + b(θ )} ⎤ cos3 θ ⋅ dθ
2 2
RW = (7.5)
2 ⎣ ⎦
0

The total resistance of the "new" ship may now be expressed as a sum of different terms,
namely the resistance of the basic ship, the resistance of the thin ship itself and the resistance
due to the interaction between the two hulls.

Hogner expressed the wave making resistance only by the basic hull form coefficient and the
interaction term between them:

153
1
B 1
CW = CW 0 + 4 ⋅ Cm ( ) ⋅ 2 ⋅ ∫ A(ξ ) ⋅ G (ξ ) ⋅ dξ (7.6)
L FN 0

where

A( x)
A(ξ ) =

AΘ = midship − area
S (θ )
S=
L
B = beam
L = shiplength
Cm = midship − area − coefficient
RW 0
CW 0 =
ρ
⋅ V 2 ⋅ L2
2

The function G (ξ) is called "influence function" and is calculated from the measured wave
data (See Figure 7.4).

One of the constraints for the thin ship is that the volume of the total ship must be constant.
This gives the following requirement for the area distribution of the thin ship:

∫ A(ξ )dξ = 0
0

With the influence function, the thin ship is added in such a way, that the sectional area curve
becomes the optimum one.

Baba (1972) tried to improve the resistance of a given ship (1129) as shown in Figure 7.4 and
Figure 7.5. By changing the sectional area of the basic ship, he arrived at two modified
models with much lower resistance than the basic version. The figures mentioned above show
the resistance curves, wave patterns, amplitude spectra and influence factor.

154
Figure 7.4 Modification of basic lines using wave pattern analysis (Baba, 1972)

155
Figure 7.5 Amplitude spectra and influence functions for shape optimisation (Baba,
1972)

156
7.3 Application of Bulbs for Reduction of Wave Resistance

The main intention with this section is

• to show how it is possible to reduce wave resistance by application of bulbs.

In the previous section it has been discussed in fairly general and mathematical term how the
wave resistance can be reduced by modification of the hull shape. As is also discussed in the
previous section, the most efficient way to reduce the wave resistance of an existing and
conventional ship is to equip the ship with a bulb. It is possible to neutralise the effect of the
wave peak generated near the bow by the wave system set up by the bulb itself.

Figure 7.3 shows an example of how the level of elevation of the free surface has been
changed in fore and aft ship and how the wave system of the bow has reduced the total wave
elevation and wave resistance.

T. Tsutsumi studied the residual resistance of a slender ship with bulbs of different size as
indicated in Figure 7.7, where the residual resistance for different F N is drawn as a function
of the ratio between bulb area and mid ship area. In this case, minimum resistance was
obtained when the ratio was about 15%.

Yasukawa started with a slender and modified Wigley hull and first changed the longitudinal
distribution of cross section area without applying bulb. He then reduced the wave resistance
by making the fore body more slender, but obtained no marked reduction before he finally
applied a bulb as in Figure 7.8. The figure shows that each ship has an optimal bulb size.

Application of bulbous bows has a long history. Before the turn of last century, model tests
showed that bulbs could reduce the resistance. In full-scale, bulbs where first applied in USA
in 1912. Later European shipbuilders followed up in the thirties and from 1950, the bulb
became increasingly common. Today the bulb is a normal and integrated part of the ship. The
question is not "to use a bow or not", but "what type of bow shall we use".

Application of bows have been carefully investigated theoretically and experimentally by Inui
(1962 and 1980), Wigley (1935 - 1942), Sharma (1967), Kinoshita, Kracht (1964 - 1985),
Weinblum (1964) among others. Figure 7.6 examplifies shaping of bulbs, according to the
three categories classified by Kracht:

∆-type Drop shaped bulb, where the centre of gravity is


positioned in the lower part of the bulb.
O – type Bulb with an oval cross-section.
∇-type Bulb where the centre is positioned in the upper
part of the bulb.

157
Figure 7.6 Different types of bulbs

Today a ∇ shaped bulb, reducing slamming in a seaway, and with a prolonged waterline in
ballast, is preferred.

Gains in the region 5 - 20 % has been obtained with bulbs between FN = 0. 17 - 0. 70. For FN
's higher than 0. 25 the main task of the bulb is to push the bow wave forward. At small FN,
the effect of pushing the volume forward is to make the shoulders "softer". A bulb may
therefore be useful also for full and slow ships.

In a seaway, the bulb reduces the pitch motions, but if the wavelength becomes larger than
1.35 - 1.55 times the ship length, it may be opposite. In this region, the motions are already
small. The increase in resistance in a seaway is often larger with bulb than without bulb.

A sharp transition between bottom and ship side results in formation of vortices in the region
near the "shoulders". Vortices are also formed in the waterline from the bow and afterwards.
It is known that this formation increases the resistance. With a bulb, these vortices are
efficiently suppressed and the resistance thereby reduced.

Åke Williams (1981) published results from SSPA with different bow shapes on a tanker as
shown in Figure 7.9. Compared to a ship having sharp bow, the power requirements were as
follows:

Bow type Vertical cylinder Elliptical bow Bulb Sharp bow


Full load 95 % 90 % 97 % 100 %
Ballast 95 % 90 % 92 % 100 %

He did also show how the propulsive power changed on a Ro-Ro ship when applying different
bulb sizes, the bulb as shown in Figure 7.10. Comparing with the "basis" ship, which had no
bulb, lower power was required at light load or in ballast with small bow. A large bow gave
lower power at full load and at light load, but higher power in ballast.

He did also vary the height of the bow as shown in Figure 7.11. This was done both for a ship
with typical U shaped frames in the fore body and for a ship with typical V shaped frames in
the fore body with the following results:

158
Bow type No bulb Low bulb High bulb
Full load 100 % U 95 % U 97 % V
Light load 100 % U 90 % U 82 % V
Ballast 100 % U 90 % U 105 % V

Figure 7.7 Effect of bulb size on residual resistance (Tsutsumi )

159
Figure 7.8 Optimisation of wave resistance (Yasukawa)

160
Figure 7.9 Hull shapes of tankers (Williams, 1981)

161
Figure 7.10 Hull shapes of Ro-Ro ships (Williams, 1981)

162
Figure 7.11 Hul shapes of container and Ro-Ro ships, with varying bulb height.
(Williams, 1981)

163
7.4 Development of Roughness and Fouling in Service

The main point in this section is to exemplify how fouling and roughness increases when
a ship is in service and how the resistance is influenced.

A high number of studies of the development of roughness over time and its influence on the
resistance are available. Both theoretical and experimental methods have been applied in the
studies. In some cases, resistance and roughness have been measured in laboratories on
castings of ship hull surfaces.

The roughness increases over time. The growth rate is dependent on the number of
"recoatings", type of coatings and the frequency of damages. If the hull is sand blasted and
recoated, the roughness and the increase in roughness are reduced considerably. On modern
surfaces, the shape of the roughness is quite as important as the amplitude.

Most ship hulls are covered in coatings which are reactivated in service. First, the plating is
covered with a "primer" then with an "anti fouling" which "gives off" poison from the surface.
Figure 7.12 illustrates the effect of anti fouling. Finally most of the poison is washed away
and a stage where it has no effect is reached.

If the upper layer is removed by brushing, there may still be enough poison available in the
lower layers to regain the original effect. The brushing may in some cases even reduce the
roughness.

An other type of coating is the self-polishing anti fouling where binder and poison form a
unit. The poison is dissolved in the water and due to the friction and the turbulence, the water
will polish the surface, and a fall in roughness over time is experienced.

Figure 7.12 Self-polishing anti-fouling

164
Figure 7.13 Reduction in roughness due to self-polishing effect of self-polishing anti-
fouling.

Figure 7.13 shows the time history of a typical self-polishing surface. A reduction after each
docking is indicated due to the self-polishing effect. It is also clear that there is an increase in
the permanent roughness due to damages and corrosion of the steel plates. It is therefore
important to know the direct reason for the roughness increase.

The reasons for the increase may be many. However, the main source is damage due to
mechanical devices like tugs, fenders and gnawing of wires. These damages are observed
mainly in the forepart and at the bottom of the ship. It is therefore important to know where
on the ship the increase in roughness has the strongest effect on the resistance.

If the ratio between CF and CF + CW is studied for different relations, a result as sketched in
Figure 7.14 is obtained where the ratio has been plotted as a function of the prismatic
coefficient. It is clear that a slender ship with low speed suffers more from increase in
roughness than a full high-speed ship.

One may, for a large tanker over a 6 years period, expect an increase in power of
approximately 10 % due to hull roughness. For a LPG ship, the corresponding increase will be
approximately 5 %.

It is difficult to know if the speed loss is caused by changes in the hull roughness or is due to
changes in the roughness of the propeller. It is also difficult to understand the relation
between fouling and frictional resistance. One thing is clear: it helps to shrub and brush hull
and propeller during docking.

165
Figure 7.14 Ratio between frictional and total resistance, as function of ship size and
prismatic coefficient.

The first question, which arises, is: "What is an optimum docking interval?"

It is difficult to establish criterions because intensity of fouling is variable, dependent of time


of the year, and global region. It is therefore interesting to do continuous measurements of
changes in the power - velocity relations and in the power- RPM relations.

For the analyses of such material, a reliable mathematical model of the relation between ship
speed, power, RPM, displacement and trim is needed. Two of the main parameters are
roughness of propeller and hull. Other parameters are temperature and salinity of the water.
Very few ship-owners have models of the type described above, in spite of the fact that it is
easy to put up such models.

Many attempts have been made to be able to trace the effect of growth in fouling during
service. Figure 7.15 shows examples from such observations over time for a 22000-tdw
tanker. During docking the hull was sandblasted and covered by a conventional anti fouling.
The roughness before and after docking was 600 and 250 my. (0. 60 and 0. 25 mm). The
docking resulted in an increase in speed of 0.8 knots. After 8 - 9 months, another speed loss
could be registered and after 12 months, the velocity was the same as before the docking.

166
Figure 7.15 Speed loss due to hull fouling of a 2200 TDW tanker.

Figure 7.16 shows the time history of fouling on a 350000-tdw tanker. After the docking, the
ship was sandblasted and self-polishing coating was added. The roughness was assessed to be
of the order of 100 my. after the docking. In addition, the propeller blades were polished
during the docking. After this the ship speed stayed constant for 11 months.

On travel, it was suddenly discovered that the ship had a marked speed loss in the order of
0.5 - 0.7 knots. The propeller was immediately polished, and the speed increased with 0.5
knots. After 21 months, the ship lost speed again. The speed loss reached as high values as 1.
6 - 2. 0 knots. The hull was inspected and well developed fouling was registered on the
bottom and along the sides.

Figure 7.16 Speed loss de to fouling of a 350 000 TDW tanker.

Finally, an example in Figure 7.17 from US Navy must be given some attention. After some

167
time in service, the ship had to increase the power by 50 % to keep the speed. This
corresponds to an increase in ∆CF of approximately 10 -3, which is very much. The hull was
then gradually cleaned and the fouling washed away on different places until the entire hull
was clean again. It is suggestive that the propeller was responsible for 32 % of the total power
increase.

If the contribution from the dome is subtracted the propeller is responsible for 50 % of the
power increase. This corresponds to an equivalent sand roughness on the propeller of 1000 -
1500 my.

These examples show how important surveillance and maintenance are.

Figure 7.17 Effect of fouling of different parts of a naval combatant.

168
8 MULTIHULLS

The main point in this section is

• to show how it is possible to reduce the wave resistance of slender and high-speed
displacement ships applying hulls in addition to the main hull, and to keep the total
resistance at an acceptable level

A possible application of multihull ships is in cases where slender ships have small resistance
but bad stability. This will often be the case for displacement vessels designed to operate at
FN>0.45. The main hull may then be supported by two or several outriggers To place two
hulls in asymmetric relation to each other as shown in Figure 8.1 is an other alternative.

With such solutions the waves of the additional hulls interferes with the waves generated by
the main hull. This interference can be both positive and negative, and it is important to study
the interaction to optimise the position of the outrigger hulls in such a way that the profit
already gained by application of the slender main hull is attractive in spite of the resistance of
the additional hulls.

8.1 Staggered Hulls

The principle for this solution is shown in Figure 8.1. As observed, the wave system of the
first ship is levelled by the wave system from the following hull, which has the same shape as
the first hull. A wave peak from the first ship fills in this way a trough from the following ship
and no resulting wave is obtained.

Figure 8.1 Wave intereference of catamaran with staggered hulls (Søding, 1997)

169
Starting with a conventional catamaran, one of the hulls is moved half a ship length in the
longitudinal direction. Resistance results as shown in Figure 8.2 and Figure 8.3 are then
obtained. The figures show model test results from tests at HSVA by H. Söding (1997) at the
university of Hamburg. The models were towed with and without longitudinal displacement
(stagger) between the hulls. Not only the resistance, but also pitch and heave were reduced
with this solution as shown in Figure 8.3.

Figure 8.2 Comparison of total and wave resistance of a catamaran with (broken line)
and without (continuous line) staggered hulls. (Søding, 1997)

Figure 8.3 Comparison of heave and pitch of a catamaran with (broken line) and
without (continuous line) staggered hulls. (Søding, 1997)

170
8.2 Single Hull with Outriggers

In the last years very slender ships with high speed has been taken into use. A problem with
these slender ships is insufficient stability. In order to improve the stability, the ships have
been equipped with two or four outriggers. It is important to optimise the position of the
outriggers to minimise the resistance penalty of the outriggers.

This solution is built upon the principle that the wave system of the main hull is influenced by
the wave system of the outriggers.

This type of ship is especially actual for use when the speeds are between 30 and 40 knots,
and we have ship lengths in the region 150 - 250 meter. This gives a FN close to 0.5, which is
the FN where the wave resistance is at a maximum, as has been demonstrated earlier (see for
instance Figure 2.34).

One such example is shown in Figure 8.5. Applying thin ship theory according to Mitchell
(see section 3.4.2), Suzuki and Ikehata (1993) calculated the wave resistance for a trimaran
built of so called mathematical hulls, where the main hull and the side hulls have identical
shapes.

For this trimaran and this main hull the connection between axial and sideways position was
optimised by looking for the y0 position, which for a given x0 position gives the smallest wave
resistance. For positions as shown in Figure 8.4 towing tests with a model, which was free, to
trim and heave were performed. During the tests resistance, sinking and trim was measured
for two design conditions, FN = 0.4 and FN = 0.5. Figure 8.4 shows comparison between
model test results and calculations.

• It is clear from Figure 8.4 that the wave resistance for a trimaran always is larger than for
the main hull alone but that there is room for optimisation.

• The tests indicate that it is more profitable to place the side hulls near the after body than
near the fore body, in spite of the fact that according to theory the positions should give
identical results.

Finally sinking and trim were measured. It was shown that the configuration with the side hull
near the after body gave the smallest trim. This may explain that also the wave resistance is
smallest for this case.

It is fairly easy to do such calculations, using the free software Michlet, available at:
http://www.cyberiad.net/michlet

171
Figure 8.4 Results of model tests (CR) and calculations (CW) for different sidehull
positions, including main hull alone (MH-0). TR-0 is sidehull alone. The definitions of
the sidehull positions are shown in the upper part of the figure

172
Figure 8.5 Optimisation of outrigger position, performed with thin ship theory

173
174
9 RESISTANCE OF CATAMARANS

The main intention in this section is

• to show how the resistance of conventional catamarans may be calculated

• to underline that frictional resistance and resistance due to the transom stern at
moderate speeds are two important resistance components

Catamarans have the following main components of resistance:


1. Viscous resistance.
2. Wave pattern resistance.
3. Air resistance.
4. Resistance due to the transom stern.
5. Different types of pressure resistance not mentioned above.
The wave pattern resistance is mainly resistance due to divergent waves.

Strictly, the resistance of the model should be expressed as:

CTm = γ CW ( Fn ) + CFm (1 + k β ) + CAA (9.1)


where
CTm =resistance coefficient for the total resistance
CW (Fn ) = resistance coefficient for the wave pattern resistance without the presence of
other hulls. The wave resistance is mainly due to divergent waves.
γ = correction of the wave resistance due to the presence of the other hull
k = form factor without presence of the twin hull.
β = correction of form factor due to the presence of the other hull
0.075
CFm =
[ log RNm − 2]
2

Vm ⋅ Lm
RNm =
υ
υ =1.1395*.10 -6
Lm = waterline length
Vm = model velocity
CAA = resistance coefficient for air resistance

In full scale:

175
CTs = γ CW ( Fn ) + (CFs + ∆CF )(1+ k β ) + CAA (9.2)

where

0.075
CFs =
[ log RNS − 2]
2

Vs Ls
R Ns =
υ
υ =1.1895 * 10 -6
LS = skip length in the water line
∆CF = increase in viscous resistance due to roughness
The air resistance is:

1 2
RAA = ⋅ V0 ⋅ Ap ⋅ Cairp (9.3)
2
where
Ap = projected area normal to the direction of the motion
The air resistance coefficient with wetted surface is then:

RAA ρ air A
C AA = = ⋅ Cairp ⋅ p (9.4)
1
⋅ ρ ⋅V0 2 ⋅ S ρ S
2
where
S = total wetted surface of the hull
For a good superstructure: Cair = 0.45

176
9.1 Increase in Resistance due to Roughness

The increase of friction due to roughness is discussed in some detail in section 2.5.
Based on measurements the increase in resistance is for practical use expressed as:

∆CF = ⎡110 ⋅ ( H ⋅ V ) − 403⎤ ⋅ CF 


0.21 2
(9.5)
⎣ ⎦
where
H = roughness in microns (my) measured over a wave length of 50 mm
V = ship velocity in m/s.
CFs = frictional resistance coefficient as defined above.

A normal value for H for high speed crafts like catamarans is 50 my -75 my for smooth hulls.
With fouling H easily reaches values as high as 200 my.

9.2 Resistance due to the Transom

It is assumed that the water " slips" or separates at the edge between transom and hull in such
a way that the entire transom is exposed to a pressure equal the atmospheric pressure.

This assumption is not valid when FN is small. When FN is small, the typical hollow behind
the transom disappears and the flow will partly cover the transom. The resistance will then be
of the same nature as " base drag " (see section 2.3.1). To decide if the transom will be dry for
a certain speed, the following “rule of thumb” can be used:
V
Transom is dry if: FNt = > 2.5 (9.6)
g ⋅ Ttransom

The resistance of a dry transom is also discussed in section 2.8. When the water "slips" a force
opposite the direction of movement is generated, which is:

Rtransom = ρ ⋅ g ⋅ ∫ ( h − z ) ⋅ B ( z ) ⋅ dz (9.7)

where
h = maximum draught at the transom
z = distance from the deepest point on the transom
B (z) = breadth of the transom at z

This component is often called hydrostatic resistance and might be of importance for
moderate FN and large transoms, as is often found on waterjet propelled craft. At very high
FN s when the ship is planing, this component becomes small compared to the hydrodynamic
component, which is due to the hydrodynamic lift of the craft.

177
9.3 Determination of the Form factor

It has been shown how the wave - resistance and the trim for conventional ships reach their
maximum at FN = 0.5. Most catamarans operate at FN > 0.5. If FN increases above 0.5, the
wave resistance decreases and finally if the FN is high enough the total resistance is composed
of resistance due to dynamic lift, frictional resistance and air resistance.

The form factor may be determined as for conventional ships when FN<0.5, but for this type
of ships the conventional procedure is inaccurate and the form factor not applicable for the
highest speeds, partly because the resistance of the transom stern will not vanish or vary with
FN4 at low speeds.

As an alternative the form factor may be therefore be determined by increasing FN until the
wave resistance becomes vanishing, like in Inui/Prohaska’s method by plotting:

CTm − C AA CW ( FN ) FN n
= (1 + k ) + m ⋅ = (1 + k ) + m ⋅ (9.8)
CFm CFm CFm

and by changing n until a linear relation is obtained between

CTm − C AA F
n
and N
CFm CFm

The intersection with the y-axes gives the form factor k. The wave or pressure resistance is
now expressed as:

Cw = m ⋅ FN
n

where
n
F
m = a constant derived from the plot when the N dependence is linear (normally for
C Fm
n values between 4 and 6 for FN<0.5 but near n=-1.25 for FN>1.0).

Figure 9.1 shows examples of such plots for a model tested at 4 draughts. This way of using
the procedure deviates from the way, which is used for conventional ships. The form factor is
now determined at the highest possible model velocity.

178
Figure 9.1 Form factor of a high speed catamaran determined from high speed towing
tests.

9.4 Decomposing of Resistance in Practical Model Testing

It is standard practise to express the resistance in a simplified way as:


CTm = CFm+ CR ( FN ) + CAA (9.9)
and
CTs = (CFs+∆ CF ) + CR ( FN )+ CAA + CA (9.10)

where

CR ( Fn ) = residual resistance coefficient


CA = correlation factor
0.075
CF =
( log RN − 2 )
2

9.5 Correlation Factor

If the resistance, as an alternative, is expressed applying the form factor, and the two
expressions for CTS are compared, the following expression for CA is obtained:
CA = k ·β ·( CFs - CFm ) (9.11)

Model tests indicate that k-values as high as 0.5 are possible, and that kβ varies between 0.35
and 0.85 depending on distance between the hulls and the fullness of the hulls.

179
These values have been determined at very low FN s without considering the fact that the craft
has a wetted transom when FN < 0.2 with "base drag", in other words that it is a suction or a
sub pressure on the transom until the water slips the transom.

If the stern is wetted, this resistance is:

3
R 0.029 ⎡ Sb ⎤ 2
CDB = 1 base2 = ⋅ (9.12)
2 ρV S CF ⎢⎣ S ⎥⎦

where

CF = frictional resistance coefficient for the ship (ITTC)


CDB = resistance coefficient for the resistance due to the transom stern referred to the
wetted surface
Sb = area of the transom
S = wetted surface without transom stern

If this resistance is subtracted from the total resistance before the form factor is determined
for small FN, form factors around 0. 25 are obtained for conventional catamarans.

For conventional catamarans, normal CA values are from - 0.0002 to - 0.0003. Analyses of the
trial results give similar indications.

9.6 Published Resistance Data for Catamarans

In RINA 1996 "Resistance Experiments on a Systematic Series of High Speed Displacement


Catamaran Forms: Variation of Length-Displacement Ratio and Breadth-Draught Ratio" Mol
land, Wellicombe and Couser have published data for CT ( Fn ), CR ( Fn ) for different hull
shapes and hull parameters. Some of these results are presented in Figure 9.3.

S is now the distance between the centres of the hulls and not the wetted surface. B means
maximum width of each single hull and finally T means draught.

The models were tested both as single hulls and as catamarans and had the following main
data:

1
Model L L/B B/T CB CP CM L ∇ 3

1 1.6 8.0 2.5 0.397 0.693 0.565 7.39


2 1.6 12.8 1.5 0.397 0.693 0.565 8.51

The form factor (1+ k) for the single hull and 1+ k β for the catamaran) was:
Model Single hull S/L=0.2 S/L=0.3 S/L=0.4 S/L=0.5
1 1.30 1.41 1.39 1.48 1.44
2 1.28 1.44 1.43 1.44 1.47

180
The form factor has, as mentioned before, high values. In Fig. 90, it is observed that the
residual resistance for the catamaran becomes equal to the resistance of the single hull, when
the distance is large enough. Normally S / L will be between 0.2 and 0.3.

At MARINTEK, different hulls have been tested in various commercial ship development
projects during the last years. The results from these tests up until about 1990 are shown in
terms of residual resistance coefficient in Figure 9.4. Observe that "Displ" in this specific case
has the meaning of tons and not m3. A regression line for wetted surface for these craft is
shown in Figure 9.5. The wetted surface can then be approximated by:


S ( m 2 ) = 1.75 ⋅ T ⋅ L + (9.13)
T

∇ = displacement for each hull in m3.


T = draught in m
Much of the resistance data, which has been published, are valid for models that have been
allowed to trim in a natural way. Generally the craft trims astern and the trim reaches a
maximum at FN = 0.5. The trim varies from 1 to 3 degrees depending on the ship dimensions.
Above FN=0.5 the trim angle will then be constant with increasing FN. However all ships have
an optimal trim at a given speed. This trim is often realised artificially with a trim board or in
other ways. Figure 9.2 gives an example of a trim optimisation for a craft running at 35 knots
and shows how the trim and the resistance vary with the FN.

Figure 9.2 Effect of trim on model resistance

181
Figure 9.3 Experimentally determined residual resistance coefficient for two different
series of catamaran hulls (Couser and Molland)

182
Figure 9.4 Regression diagram for high speed displacement catamarans, relating
residual resistance to Froude number and slenderness ratio. The diagram is made based
on commercial model tests performed at MARINTEK up to about 1993

183
Figure 9.5 Wetted surface values for high speed displacement catamarans tested at
MARINTEK.

184
10 SWATH

The content of this section is to a large extent taken from Jerry L. Gore's article “SWATH
SHIPS" in Naval Engineers Journal Modern Ships and Craft, February 1985.

The main intention in this section is to

• present an unconventional vehicle with special qualities and to show how the
resistance is calculated for this type of ships

SWATH means: "Small Water plane Area Twin Hull". In Figure 10.1, a typical SWATH is
shown. The advantage with the SWATH is that it has much better sea keeping qualities than a
conventional craft with the same length. The reason is the small water plane area of the
SWATH.

Reduced motions minimise the speed reduction in a seaway. The drawback is increased
immersion.

The improved sea keeping ability of the SWATH has been demonstrated several times. An
excellent comparison was made with several craft and conventional ships on trial (side by
side) under identical sea conditions. See Figure 10.2. The sea keeping ability for a SWATH of
220 tons compared to the qualities of a conventional ship of 100 tons and 3100 tons.

Figure 10.1 Typical SWATH ship

185
Figure 10.2 Comparison of motion of SSP Kaimalino and USCG monohulls

186
The excellent sea keeping ability of the SWATH is explained by reduced water plane area
which reduces wave-induced forces and moments.

Pitch and heave periods are expressed as:

m ⋅ r5 2 + A55
Tstamp = 2π ⋅ (10.1)
ρ ⋅ g ⋅∇ ⋅ GM L

m + A33
Thiv = 2π ⋅ (10.2)
ρ ⋅ g ⋅ Awl

where
m = mass of the ship
∇ = volume of the ship
A33 = added mass in heave
A55 = added mass in pitch
Awl = water plane area
r5 = radius of inertia in pitch
GML = longitudinal metacentre height.
It is observed that a small waterline area and small metacentre height give large values of the
natural periods in heave and pitch. The frequency of response is very different from the actual
wave periods. Combined with small excitations this leads to small motions in a seaway.

When the ship is travelling, the hull has an unsymmetrical pressure distribution in the
longitudinal direction with a corresponding moment turning the hull with the nose up or
down. This moment (Munk's moment) is given by:

M munk = k ⋅ A33 ⋅ V 2 (10.3)

where k is a constant

The added mass in heave is approximately proportional to ∆. It is therefore allowed to make


the following expression:

M munk = k ⋅ ∆ ⋅ V 2 (10.4)

The righting moment is:

M T = ∆ ⋅ GM L ⋅ sin θ (10.5)

where
∆ = displacement
θ = pitch angle

187
Because the waterplane area dA is small, GML is also small. Therefore MMunk will as the
speed increases soon become larger than the rightening moment MT and the craft becomes
unstable and needs righting surfaces in the form of fins and flaps.

It would have been possible to improve the situation by changing the distance between the
centre of gravity and the centre of buoyancy, but this leads to an increase in calm water
resistance.

Figure 10.3 shows how a SWATH behaves in heave at different wave lengths. It is observed
that the wave, which gives resonance at 0 velocity, creates no problems at 20 knots. Figure
10.4 shows the influence of the control planes on the pitch motions.

Figure 10.3 Response Amplitude Operator (RAO) in heave for a SWATH

188
Figure 10.4 Effect of pitch control on measured single amplitude motion of SWATH
Seagull at 24 knots in sea state 5

10.1 Components of Resistance

A SWATH has the following resistance components in calm water:


1. Viscous resistance.
2. Wave resistance.
3. Spray resistance from the struts.
4. Resistance in the junction between strut and submerged body
5. Air resistance.

Viscous Resistance of the Struts.

If there is one strut on each hull, the total viscous resistance of the struts is expressed by:

⎛ t ⎞ ρ 2
Rvst = 4CFst ⋅ ⎜1 + k ⋅ ⎟ ⋅ ⋅ V0 ⋅ Lst ⋅ H st (10.6)
⎝ Lst ⎠ 2
where
Hst =distance from the upper side of the submerged body to the surface
Lst = length of the strut:
0.075
CFst = + ∆CF
[log RNst − 2]
2

k = form factor ( k = 2 )

189
t = maximum thickness of the strut

Viscous Resistance of the Submerged Body.

The resistance of the body of revolution is expressed as:

⎡ ⎛ Dmax ⎞
1.5
⎛ Dmax ⎞ ⎤ ρ 2
3

RVuv = 2 ⋅ CFuv ⋅ ⎢1 + 1.5 ⋅ ⎜ ⎟ + 7⋅⎜ ⎟ ⎥ ⋅ ⋅ V0 ⋅ Suv (10.7)


⎢⎣ ⎝ Luv ⎠ ⎝ Luv ⎠ ⎥⎦ 2

where
Luv = length of the submerged body
Suv = wetted surface of the submerged body
Dmax = [4Amax / π] 0.5 = equivalent maximum diameter for the body
Amax = maximum cross section
0.075
CFuv = + ∆ CF
[ log RNuv − 2]
2

V0 Luv
RNuv =
υ

Junction between Strut and Submerged Body.


In the junction between a strut and a flat submerged body, two boundary layers meet and a
vortex or "horseshoe vortex", is formed. The energy in the vortex is equivalent to a resistance,
which based on experimental results is approximated by:

Rnech t 0.0003
CDnech = = 0.75 ⋅ − (10.8)
1 Lst ( t ) 2
ρV0 2t 2
2 Lst

where
t = maximum thickness of the strut.
This drag is valid for a junction without fillet. A fillet in the junction reduces this drag.

If the junction between two struts is “sharp”, (the radius in the junction is very small), the
following formula can be applied:

2
⎛ t ⎞
CDnech = 17 ⋅ ⎜ ⎟ − 0.05 (10.9)
⎝ Lst ⎠
where:
t1 + t2
t=
2
If the junction is softened fitted with a "soft" junction with a fillet or a large radius in the

190
junction, the drag is reduced to:

⎡ ⎛ t ⎞2 ⎤
CDnech = k ⋅ ⎢17 ⋅ ⎜ ⎟ − 0.05⎥ (10.10)
⎢⎣ ⎝ Lst ⎠ ⎥⎦

where
k (r ) = constant which is a function of the fillet radius (0.15-0.30).
The formulas indicate that a "soft" junction with a fillet or a large radius in the junction,
reduce the drag:

Unfitness with these formulas is the effect of the radius in the junction. A very small radius in
the junction gives high resistance.

t
For different values of , the drag coefficient becomes:
Lst

t
0.1000 0.1500 0.2000
Lst
⎛ t ⎞ 0.0003
CDnech = 0.75 ⋅ ⎜ ⎟ − 2
⎝ Lst ⎠ ⎛ t ⎞ 0.0450 0.0992 0.1425
⎜ ⎟
⎝ Lst ⎠
2
⎛ t ⎞
CDnech = 17 ⋅ ⎜ ⎟ − 0.05 0.1200 0.3325 0.6300
⎝ Lst ⎠
⎡ ⎛ ⎛ t ⎞2 ⎞⎤
CDnech = ⎢ 0.30 ⋅ ⎜17 ⋅ ⎜ ⎟ − 0.05 ⎟ ⎥ 0.0360 0.0998 0.1890
⎢ ⎜ ⎝ Lst ⎠ ⎟⎥
⎣ ⎝ ⎠⎦

Spray Resistance due to Struts.


Each of the struts generates spray in addition to waves. The energy in the spray is related to
the "spray resistance" which based on model tests is given by:

Rspray
CDsprut = = 0.24 (10.11)
ρ
⋅V0 ⋅ t
2 2

2
where
t = thickness of the strut
As long as FN < 0.5 the strut will only generate one wave system. If the velocity becomes
larger, the wave brakes and builds up. It then and spreads out to the sides in a spray. The
ρV 2 L
extent of the spray is a function of FN , RN and Weber's number ( We = ).
ϒ

191
Weber's number is an expression for the surface stress, like on drops, in capillary flow and for
small waves.)

Rspray t
If the test results of CDX = are plotted as function of , we get:
ρ x
⋅V 2 ⋅ x 2
2

2
⎛t⎞ t
CDX = 0.24 ⋅ ⎜ ⎟ up to = 0.4 (10.12)
⎝ x⎠ x

If the struts become thicker, the resistance may be expressed as:

CDx= 0.12 (10.13)

x = distance from the leading edge to the position of maximum thickness


t = maximum thickness
The transition is abrupt and often caused by ventilation of the region behind x = t / 0.4. If the
drag coefficient is expressed with t instead of x, this gives:

CDt = 0.24

which is valid for FN = 3 where FN is based on the chord length

Residual Resistance and Trim Moment.

Finally, we have the wave resistance from the struts, the submerged body and the interference
between these:

In "The HSVA Systematic Swath Model Series" Peter Schenzle (1995) has published results
from systematic tests with Swath ships. For different configurations of trim, immersion, hull
distance, strut thickness and diameters the resistance and trim moments were measured at
different Froude numbers. The results are shown in Figure 10.5 to Figure 10.10.

The following parameters where applied:

L = hull length = length of strut


CPH = prismatic coefficient
DH / L = relative hull diameter
ts / DH = relative strut thickness
Bs / L = relative hull distance
D / DH = relative hull immersion
DH = [4Amax / π] 0.5 = equivalent maximum diameter for the body
Amax = maximum cross section of the hull

192
Instead of CR and CM for residual resistance and moment, Schenzle applied:

RR
εR = (10.14)
ρ ⋅ g ⋅∇

My
εM = (10.15)
ρ ⋅ g ⋅∇ ⋅ L

where the volume of the entire submerged, body was approximated by:

⎡ t ⎛ ⎛ d ⎞ ⎞⎤
∇ = L ⋅ DH 2 ⎢0.3 + 1.17CPH + s ⋅ ⎜⎜ 0.95 + 1.675 ⎜ − 1⎟ ⎟⎟ ⎥ (10.16)
⎢⎣ DH ⎝ ⎝ DH ⎠ ⎠ ⎥⎦

The wetted surface of the entire body was approximated by:

⎡ ⎛ t ⎞ d ⎤
S = L ⋅ DH ⋅ ⎢ 2 + ⎜ 1.12 + 0.90 ⋅ (1 − s ) 2 ⎟ CPH + 4 ⎥ (10.17)
⎢⎣ ⎝ DH ⎠ DH ⎥

A negative moment means that the craft is trimming forward. These results clearly
demonstrate the nead for trim control.

The residual resistance in the diagrams was estimated by application of the ITTC curve,
which was used for the viscous resistance.

In estimation of ship resistance, these diagrams for residual resistance can be applied. The
total resistance is expressed as:

ρ
Rtots = ε R ( FN ) ⋅ ρ ⋅ g ⋅∇ + ⋅ V0 2 ⋅ S ⋅ [CFS + ∆CF + C A ] (10.18)
2
where
CA = correlation factor
S = total wetted area of struts and body of rotation
In the calculation of CFS, the length of the strut is applied as " the length of the hull."

When the method described above is applied, one should apply the same type of hull as used
by Schenzle.

If other types are actual, a method based on the equations proposed initially is recommended.
These equations include most of the components except wave resistance. Here it is possible to
apply the results from Schenzle's tests and to estimate a " wave " or "residual component". If
that is the case, the FN of the model and the total resistance is determined according to the
concept shown above.

Then the residual resistance is determined from the expression:

193
ρ ρ
R ( FN ) = Rtotm ( FN ) − ⋅ V0 2 ⋅ [ Sroot ⋅ CDrot + CDstrut ⋅ S strut ] − ⋅ V0 2 ⋅ t 2 ⋅ ⎡⎣CDnech + CDspray ⎤⎦ (10.19)
2 2

This gives the residual resistance coefficient for struts and hull:

R ( FN )
CR ( FN ) =
ρ
⋅Vm 2 ⋅ Suvm
2

It is observed that Suvm and not S are applied in this expression.

The total resistance for the ship is now:

⎡ ⎡ D D ⎤⎤ ρ
Rtot = ⎢CR ( FN ) + 2 ⋅ CFuvs ⋅ ⎢1 + 1.5 ⋅ ( max )1.5 + 7 ⋅ ( max )3 ⎥ ⎥ ⋅ ⋅ V0 2 ⋅ Suvs + Rvst + Rnech + Rspray
⎣ ⎣ Luvs Luvs ⎦ ⎦ 2

where
⎛ t ⎞ 1
Rvst = 4CFS ⋅ ⎜1 + 2 ⎟ ⋅ ρ ⋅ V0 2 ⋅ Lst ⋅ H st
⎝ Lst ⎠ 2

We do now have two different ways to determine the resistance. The last method gives larger
freedom in the choice of hull parameters, but both methods contain questionable assumptions.

Finally, the wake w and the thrust deduction t must be determined. Tests have shown that the
wake determined by thrust identity, can be approximated by w = 0. 12 - 0. 14 if the propeller
diameter is 80 % larger than the hull diameter.

An useful approximation for the thrust deduction is:

t = w * η0 (10.20)

194
Figure 10.5 Residuary resistance and trim moment. Effect of lower hull diameter DH/L
variation.

Figure 10.6 Residuary resistance and trim moment. Effect of strut width ts/DH variation.

195
Figure 10.7 Residuary resistance and trim moment. Effect of demihull spacing Bs/L
variation.

Figure 10.8 Residuary resistance and trim moment. Effect of lower hull submergence
d/DH variation.

196
Figure 10.9 Residuary resistance and trim moment. Effect of lower hull fullness CPH
variation.

Figure 10.10 Residuary resistance and trim moment. Effect of trim ∆T/L variation.

197
198
11 GENERAL SYMBOLS AND DEFINITIONS

Area/Lengths:

SYMBOL DEFINITION UNIT


DESIGNATION
c = cord length profile length M
z = vertical position distance between two levels in the M
fluid
h = depth vertical distance from the free M
surface and to a given point in the
fluid
D = diameter normally the propeller diameter M
L = characteristic length normally the ship length but also M
ship length in general
R = radius normally propeller radius M

S = wetted surface surface in contact with water m2


A = area wetted surface but also projected m2
area

λ = Ls/Lm = scale ratio between length of prototype 1


and model length

199
Velocities:

SYMBOL DEFINITION UNIT


DESIGNATION
V skip velocity in general m/s
U0 main velocity m/s
U velocity in general m/s
C, u ,v, w velocity in the boundary layer m/s
resulting velocity
q m/s
q 2 = u x2 + v y2 + wz2
n number of revolutions m/s
Knots and
VS ship velocity
m/s
V ship velocity m/s
VA advance velocity of the propeller m/s
propeller induced axial velocity
UA m/s
behind the propeller
propeller induced tangential
UT m/s
velocity behind the propeller

200
Force/mass:

SYMBOL DEFINITION UNIT


DESIGNATION
M = mass Kg
g = acceleration due to m/s2
gravity
R = resistance force opposite to the N
direction of motion
RW = wave resistance resistance due to wave N
generation
D = resistance, drag force on the body opposite N
to the direction of motion
FFR = friction force N
Fd = dynamic force ρ N
V 2A
2

τ = surface tension/shear τ = µ du/dy N/m2


force

Pressure:

SYMBOL DEFINITION UNIT


DESIGNATION
∆p = difference in pressure pressure difference Pascal, Pa = N/m2
between two points
pa atmospheric pressure Pascal, Pa = N/m2
p pressure Pascal, Pa = N/m2

201
Coefficients:

SYMBOL DEFINITION UNIT


DESIGNATION
CD = drag or resistance D -
coefficient 1
ρV A 2
2

CL = liftcoeffisient L -
1
2 ρV A 2

CP = pressure coefficient ∆p -
1
2 ρV 2

CT = resistance coefficient , R -
total resistance
2 ρV S
1 2

CF = friction coefficient F FR -
2 ρV
1 2

Characteristic numbers:
SYMBOL
DEFINITION UNIT
DESIGNATION
RN = Reynolds number VL/ν -
V
FN = Froude's number -
gL
V
J = advance ratio -
nD
dy Ns
µ = dynamic viscosity µ =τ 2
du m
µ
ν= Kinematics viscosity ν= m2/s
ρ

Energy:
SYMBOL DEFINITION UNIT
DESIGNATION
EH = potential energy ρgz Joule, J=Ws=Nm
EV = dynamic energy 1
2 ρV 2 Joule, J=Ws=Nm
EP = pressure energy P Joule, J=Ws=Nm
Etot = total energy E H + EV + E P Joule, J=Ws=Nm

202
Resistance:

SB area of transom
S area of wetted surface
L ship length
FN Froude's number
∆C F aditional resistance due to roughness
RN Reynolds number
V0 ship velocity
R resistance in general
RT
CT = coefficient for the total resistance
1
2 ρV 2 S
RTm coefficient for the total resistance of the
CTm = 1
2ρV 2 S model
R
CW = 1 W 2 wave resistance coefficient
2 ρV S

R
CF = 1 F 2 coefficient for frictional resistance
2 ρV S

R
C App = 1 App2 coefficient for appendage resistance
2 ρV S

D viscous resistance coefficient for a body of


CDV = 1 V 2
2 ρV S
rotation
Cvortex coefficient for vortex resistance
R
CV = 1 V 2 coefficient of viscous resistance
2 ρV S

R
CDB = 1 BD2 coefficient of base drag
2 ρV S

R
C AA = 1 AA2 coefficient of air resistance
2 ρV S

R
CR = 1 R 2 coefficient of residual resistance
2 ρV S

RT total resistance
RW wave(pattern) resistance
RF frictional resistance
RV viscous resistance of the ship
DV viscous resistance of a body of rotation
RDB resistance due to the transom
RApp resistance due appendages(brackets, domes,
propeller shafts etc )

203
R AA air resistance
k0 formfactor due to increase in local velocity
k1 formfactor due to pressure resistance
k = k1 + k 0 formfactor in general
TA draught at the aft perpendicular
TF draught at forward perpendicular
T draught in general( normally amidships)
CB block coefficient
LVL length in waterline
H hull roughness in microns
longitudinal position of centre of buoyancy(
LCB ( or lcb ) also used to denote the distance of CB abaft
amidships( Lpp/2))

204
12 REFERENCES

Abbott, I.H," Theory of Wing Sections New York." Mc Graw - Hill 1949

Baba.E. (1972), " An Application of Wave Pattern Analysis to Ship Form Improvement "
Journal of Society of Naval Architects of Japan"

Baba.E.(1969), " Study on Separation of Ship Resistance Components," Mitsubishi Technical


Report No.59.

Barratt,M.J.1965. “The wave drag of a Hovercraft” J. Fluid Mech.(1965),vol.22,part1,p.39-47

Blume, P. and Kracht, A.M. "Prediction of the Behaviour and Propulsive Performance of
Ships with Bulbous Bow in Waves." Transactions of the Society of Naval Architects and
Marine Engineers. 93. 1985.

Clayton,B.R and Bishop,R.E.D ” Mechanics of Marine Vehicles”. Spon London(1982).

Dawson, J and Bowden, B. S. " The Prediction of the Performance of Single Screw Ships on
Measured Mile Trials”. NPL. Ship Report, 168. National Physical Laboratory, Teddington
1972.

Dawson, J og Bowden, B. S. " Performance Prediction Factors for Twin Screw Ships." NPL.
Ship Report, 172. National Physical Laboratory, Teddington 1973.

Doctors, L.J and Sharma, S.D, “The wave resistance of an air cushion vehicle in steady and
accelerated motion”, Journal of Ship Research, 1972,pp.248-260.

Douglas,J.F, Gasiorek,J.M, Swaffield,J.M .1983 "Fluid Mechanics"ISBN027300462x.Pitman

Doust, D. J.(1962 - 63), " Optimised Trawler Forms," Trans. NECI, Vol.79

Doust, D.J.(1963), " Statistical Analysis of Resistance and Propulsion Data ". National
Physical Laboratory, Ship Report 42, Feltham, England

Eggers, K., Sharma, S., and Ward, 1967. "An Assessment of some Experimental Methods for
determining the Wave making Characteristics of a Ship Form." Society of Naval Architects
and Marine Engineers Transactions 75: 112-157.

Gadd, G. 1976. " A Method of computing the Flow and Surface Wave Pattern around full
Forms." Royal Institution of Naval Architects Transactions 118:207-216

Gertler, M. (1954), A Re - Analysis of the Original Test Data for the Taylor Standard Series
TMB Report 806 DTRC

Graff, W., Kracht, A. and Weinblum, G.P. (1964) " Some Extensions of DW Taylor Standard
Series". SNAME Transactions Vol.72.

205
Gore Jerry L. " SWATH SHIPS", Naval Engineers Journal Modern Ships and Craft. February
1985.

Guldhammer, H. E. and Harvald, S.A.(1974), " Ship Resistance and Principal Dimensions (
Revised)," Akademisk Forlag, Copenhagen.

Havelock,T.H.1932. “The theory of wave resistance”: Proc.Roy.Soc.A,138,339-48.

Havelock, T.H. (1951), “Wave Resistance Theory and its Application to Ship Problems”,
SNAME Transactions, Vol.59.

Hofman and Korzarski, Shipbuilding Progress 1999.

Hollenbach Klaus Uwe "Estimating Resistance and Propulsion for Single - Screw and Twin -
Screw Ships" Schiffstechnik Bd.45-1998 / Ship Technology Research Vol.45-1998.

Hollenbach Klaus Uwe "Beitrag zur Abschätzung von Widerstand und Propulsion von Ein
und Zweischraubenschiffen im Vorentwurf. IfS-Rep.588, Univ.Hamburg."

Holtrop, J. and Mennen, G. G.J (1978), "A Statistical Power Prediction Method," ISP,Vol.25

Holtrop. J (1984) "A Statistical Re-Analyses of Resistance and Propulsion Data "International
Shipbuilding Progress.

Hoerner, S.F. (1965 ), Fluid Dynamic Drag,"Hoerner Fluid Dynamics, Bricktown, New
Jersey.

Hoerner/Borst (1985) ”Fluid-Dynamic Lift”, Hoerner Fluid Dynamics Albuquerque.N.M.

Inui, T., Kajitani, H.and Miyata, H. (1962), "Wave making Resistance of Ships," SNAME
Transactions, Vol. 70.

Inui, T. (1980), "From Bulbous Bow to Free-Surface Shock Wave-Trend of Twenty Years
Research on Ship Waves at the Tokyo University Tank," 3rd George Weinblum Memorial
Lecture.

Isherwood, R.M. " Wind Resistance of Merchant Ships". Transactions of the Royal Institution
of Naval Architects, 115, 1973

Karlsson, R. I. "The Effect of Irregular Surface Roughness on the Frictional Resistance of


Ships". 1979 SSPA Report

Keil, U. und Schenzle, p. ( 1975), " Modellversuche mit Extrem Breiten Schiffsformen-
Widerstadsversuche mit Extrem Breiten Schiffsformen (Cp = 0.77 ), " Institut für Schiffbau
der Universität Hamburg, Bericht Nr. 333.

Lamb,H.1926. On wave resistance. ProcRo.Soc.A,111,14-25.

206
v. Lammeren, W. P. A, Troost, L., and Konig, J. G. 1948. " Resistance, Propulsion and
Steering of Ships". Haarlem, Netherlands: H. Stam.

Lee, C.M, and M. Martin. "Determination of the Size of Stabilising Fins for Small Water
plane Area Twin-Hull Ships, DTNSRDC Report No. 4495, November 1974.

In, W.C and Day, W.G. " The Still-Water Resistance and Propulsion Characteristics of Small
Water plane Area Twin-Hull Ships". AIAA / SNAME Advanced Marine Vehicles
Conference, San Diego, California, February 1974

Molland, Wellicomb and Couser, "Resistance Experiments on a Systematic Series of High


Speed Displacement Catamaran Forms: Variation of Length-Displacement Ratio and Breadth-
Draught Ratio" RINA 1996.

Schenzle, Peter "The HSVA Systematic Swath Model Series". Fast 95 Lübeck 1995

Schlichting and Truckenbrodt " Aerodynamik des Flugzeuges" del II, Berlin, Heidelberg.

Sharma, S.D "Zur Problematik des Schiffswiderstandes in zähigkeits und wellenbedingte


Anteile" Jahrbuch der Schiffbautechnischen Gesellshaft 59.Band 1965

Suzuki, Ikehata " Fundamental Study on Optimum Position and Outriggers of Trimaran from
View point of Wave Making Resistance" Fast 93, 1993.

Söding, H " Drastic Resistance Reductions in Catamarans by Staggered Hulls" Fast 1997,
Sydney Australia.

Taniguchi, K. " Model - Ship Correlation Method in the Mitsubishi Experimental Tank"
Mitsubishi Technical Bulletin December 1963.

Tatinclaux J.C.”On the Wave Resistance of Surface-Effect Ships” Transactions of SNAME


1975

Todd, F. H. (1963), " Series 60-Methodical Experiments with Models of Single - Screw
Merchant Ships, DTMB Report No. 1712 DTRC.

Townsin, R. L., Byrne, D., Svensen, T.E. and Milne, A. " Estimating the Technical and
Economic Penalties of Hull and Propeller Roughness." Transactions of the Society of Naval
Architects and Marine Engineers, 89 1981

Zierep, J.(1997)."Grundzüge der Strömungslehre" Berlin Heidelberg, Springer Verlag.

Wigley, C. (1942)."Calculated and Measured Wave making Resistance for a Series of Forms
defined Algebraically, the Prismatic Coefficient and Angle of Entrance being varied
Independently, " Trans. INA, Vol. 84

Wigley, C. (1935-36). "The Theory of the Bulbous Bow and its Practical Application, "Trans.
NECI, Vol.52

207
White, Reports of the Performance Committee and the Resistance Committee, Proceedings of
the 11-21th International Towing Tank Conference (ITTC).

Williams Åke " Kan nya skrovformar hjelpa oss hålla nere bunkerskostnaderna?". SSPA
Allmaen Rapport 1981.

208
13 INDEX

3
3-D panel method
linear................................................................................................................................................................. 99
non-linear ......................................................................................................................................................... 99
3-D panel methods ................................................................................................................................................ 99
A
Air resistance ................................................................................. 66, 67, 68, 81, 85, 115, 175, 176, 178, 189, 203
appendage resistance
scaling .............................................................................................................................................................. 74
Appendage resistance............................................................................................................................................ 71
Appendages..............................................................................................................................................31, 82, 118
B
Baba .....................................................................................................................................3, 50, 61, 153, 154, 205
base drag ............................................................................................................................................................... 25
beta-factor ............................................................................................................................................................. 75
bilge keels ............................................................................................................................................................. 71
Blasius................................................................................................................................................................... 11
boundary layer ................................................................................................................................................ 15, 19
Boundary layer........................................................................................... 8, 19, 20, 21, 26, 73, 117, 118, 190, 200
boundary layer thickness....................................................................................................................................... 19
Bow wave.................................................................................................................................................59, 60, 158
Brackets....................................................................................................................................................90, 91, 118
Bulb..................................................................................................................................................................... 157
Bulbous bow ....................................................................................................................................................... 157
Bulbs ......................................................................................................................................88, 113, 151, 157, 158
C
Catamaran ............................................................................................................................................180, 181, 207
Cole's wake parameter .......................................................................................................................................... 19
Correlation ...................................................................................................................85, 92, 98, 99, 179, 193, 207
correlation allowance .................................................................................................................................46, 75, 78
Correlation factor .........................................................................................................................................115, 116
Critical speed .........................................................................................................................................87, 133, 136
D
d’Alembert’s paradox ........................................................................................................................................... 23
d’Alemberts principle ............................................................................................................................................. 7
Depth Froude number ......................................................................................................................................... 124
displacement thickness.......................................................................................................................................... 21
Doust................................................................................................................................................................... 205
Drag ........................................................................................................... 3, 25, 116, 177, 180, 192, 201, 203, 206
Dynamic lift ........................................................................................................................................................ 178
E
eddy viscosity........................................................................................................................................................ 17
Efficiency.......................................................................................................................................................67, 118
Euler equations...................................................................................................................................................... 99
F
Field .................................................................................................................................................................... 109
flow separation...................................................................................................................................................... 15
Flowtech.............................................................................................................................................................. 100

209
form factor ............................................................................................................................................................ 27
Form factor........................25, 26, 28, 31, 32, 33, 36, 37, 39, 62, 81, 84, 91, 92, 116, 175, 178, 179, 180, 181, 189
foul release paint systems...................................................................................................................................... 47
Friction.................................................................................................................................................164, 201, 203
friction line.............................................................................................................................................................. 8
frictional resistance ................................................................................................................................................. 7
Froude number ...............................................................................................89, 103, 104, 126, 133, 134, 135, 192
G
Gadd.................................................................................................................................................................... 205
Geometric.............................................................................................................................................................. 55
geosim tests........................................................................................................................................................... 33
Gertler ................................................................................................................................................................. 205
Gore .............................................................................................................................................................185, 206
Graff...............................................................................................................................................................60, 205
Group velocity .............................................................................................................................129, 130, 131, 133
Guldhammer ..................................................................................................................................................88, 206
H
Harvald...........................................................................................................................................................88, 206
Havelock ................................................................................ 3, 50, 52, 94, 100, 108, 123, 126, 145, 146, 153, 206
Heave ...................................................................................................................................................171, 187, 188
High speed craft .................................................................................................................................................... 66
Hoerner ............................................................................................................................................................... 206
Hollenbach ...................................................................................................................88, 89, 91, 92, 110, 113, 206
Holtrop .............................................................................................................................36, 37, 88, 92, 94, 96, 206
Hull efficiency ...................................................................................................................................................... 91
Hull roughness .........................................................................................................................................43, 85, 165
Hull series ....................................................................................................................................................... 86, 87
I
Ikehata..........................................................................................................................................................171, 207
IMO....................................................................................................................................................................... 47
Induced resistance ..........................................................................................................................................39, 116
interference resistance........................................................................................................................................... 72
International Towing Tank Conference ...................................................................................................14, 76, 208
Inui .........................................................................................................................3, 31, 39, 50, 151, 157, 178, 206
Isherwood............................................................................................................................................................ 206
ITTC.............................................................................................................................11, 14, 46, 76, 180, 193, 208
K
Karlsson .............................................................................................................................................................. 206
Kelvin.............................................................................................................................................................52, 108
Kelvin angle .......................................................................................................................................................... 53
Kelvin wave group ................................................................................................................................................ 52
Kozarski .............................................................................................................................................................. 126
Kracht.....................................................................................................................................................60, 157, 205
L
Laminar ............................................................................................................................................................. 9, 15
laminar flow .......................................................................................................................................................... 11
Lammeren ................................................................................................................................................3, 123, 207
Lee ...................................................................................................................................................................... 207
local frictional coefficient ..................................................................................................................................... 19
M
MARINTEK ......................................................................................................................................................... 45
Mennen ....................................................................................................................................................88, 94, 206
Michell .........................................................................................................................................................105, 153
Molland ............................................................................................................................................................... 207

210
momentum thickness............................................................................................................................................. 19
Multihull ............................................................................................................................................................. 169
N
Navier-Stokes equations ....................................................................................................................................... 99
O
Outriggers ............................................................................................................................................169, 171, 207
P
Potential flow theory............................................................................................................................................. 99
Prandtl-Schlichting................................................................................................................................................ 14
Pressure distribution.........................................................................................................................28, 50, 117, 187
Pressure resistance ................................................................................. 23, 27, 30, 39, 40, 115, 116, 119, 175, 178
R
Relative rotative efficiency ................................................................................................................................. 118
Residual resistance......................................... 69, 76, 81, 82, 84, 89, 91, 92, 93, 124, 157, 179, 181, 193, 194, 203
Reynolds number ...........................................................................................................................9, 12, 19, 20, 203
Roughness .................................................................. 43, 44, 92, 164, 165, 166, 167, 168, 176, 177, 203, 206, 207
Roughness allowance............................................................................................................................................ 45
rudder .................................................................................................................................................................... 71
S
Schenzle .......................................................................................................................................192, 193, 206, 207
Series.......................................................................................................... 3, 4, 86, 88, 92, 102, 180, 192, 205, 207
Shallow water....................................................................... 108, 109, 123, 124, 126, 127, 129, 130, 134, 135, 136
shape factor ........................................................................................................................................................... 21
Sharma .............................................................................................................................................4, 157, 205, 207
Shipflow.............................................................................................................................................................. 100
slender ship theory ................................................................................................................................................ 99
Speed loss.............................................................................................................................................165, 166, 167
Spray resistance ...........................................................................................................................................189, 191
SSPA..............................................................................................................................................87, 158, 206, 208
stabiliser fins ......................................................................................................................................................... 71
Strut..............................................................................................................................108, 189, 190, 191, 192, 193
SWATH ...............................................................................................................................185, 187, 188, 189, 206
Søding ................................................................................................................................................................. 207
T
Taniguchi ............................................................................................................................................................ 207
TBT-type antifoulings........................................................................................................................................... 47
Thickness .................................................................................................................................20, 28, 190, 191, 192
thin ship theory ..................................................................................................................................................... 99
Thrust deduction ..........................................................................................................................................117, 194
Todd .........................................................................................................................................................86, 87, 207
Townsin............................................................................................................................................................... 207
Transom ....................................................................... 25, 26, 40, 64, 69, 82, 96, 98, 108, 136, 175, 177, 180, 203
Transom resistance................................................................................................................................................ 69
transom stern......................................................................................................................................................... 25
transom stern resistance ........................................................................................................................................ 25
Transom stern resistance ....................................................................................................................................... 69
Trim ............................................................................................... 33, 108, 110, 123, 166, 171, 178, 181, 192, 193
Trim moment ...............................................................................................................................................108, 192
trochoidal waves ................................................................................................................................................... 55
Troost .............................................................................................................................................................. 3, 207
Tsutsumi.............................................................................................................................................................. 157
Tunnel openings.................................................................................................................................................... 73
tunnel thruster ....................................................................................................................................................... 71
Tunnel thruster ...................................................................................................................................................... 73

211
Turbulent......................................................................................................................................................9, 11, 15
turbulent flow........................................................................................................................................................ 16
V
viscous pressure resistance...................................................................................................................................... 7
Viscous pressure resistance................................................................................................................................... 23
viscous resistance.............................................................................................................................................. 7, 27
Viscous resistance .................... 29, 30, 37, 39, 62, 81, 113, 115, 137, 138, 139, 140, 141, 175, 176, 189, 193, 203
W
Wake ......................................................................................................................................64, 118, 119, 139, 194
effective.......................................................................................................................................................... 118
Wave generation ................................................................................................................................................. 201
Wave height ...................................................................................................................56, 102, 139, 145, 150, 153
Wave making resistance.............................................................................................14, 51, 62, 116, 147, 151, 153
Wave pattern resistance .................................................................................................................................64, 175
Wave resistance......23, 31, 32, 33, 50, 54, 57, 59, 60, 61, 62, 64, 69, 81, 82, 94, 97, 109, 110, 124, 126, 136, 138,
139, 142, 145, 146, 150, 153, 157, 169, 171, 175, 178, 189, 192, 193, 201, 203
Weinblum...................................................................................................................................3, 60, 157, 205, 206
Wetted area ......................................................................................................................................................... 193
Wetted surface ................................................... 10, 26, 28, 29, 66, 84, 98, 109, 176, 180, 181, 190, 193, 199, 203
White........................................................................................................................................................4, 107, 208
Wigley................................................................................................................................4, 50, 123, 126, 157, 207
Williams.......................................................................................................................................................158, 208
Y
Yasukawa............................................................................................................................................................ 157

212

You might also like