You are on page 1of 106

N-

J i.:

ft'r:;:

a. f>:
/ /

(f ;
oA
A / '<.
fr

f:
n

M
o
CJ
i
2iSLt>' o v e f

SHIPS and WATER

y
0*

.VlV
.h\
V * nK

by

J.A.H Paffett RCNC CEng FRINA HonFNI FRSA

N.Cham. 623.812 P127


Autor: Paffett, J.A.H
Titulo: Ships and water : a short book about the physical in
315787
257436

Ex.1
SHIPS AND WATER

A short book about the physical


interactions which occur between
a ship and her liquid environment

by

James A.H. Paffett RCNC CEng FRINA HonFNI FRSA

U; O2'O/
1 1 s-
UliOTEC
*0 ^ ?r ^ ^
*2X)o

Published by The Nautical Institute


202 Lambeth Road, London SE1 7LQ England
Telephone (0171) 928 1351

First published 1990 \

Copyright The Nautical Institute 1990 I

All rights reserved. No part of this publication may be reproduced, stored in a retrieval
system, or transmitted in any form or by any means, electronic, mechanical, photocopying,
recording or otherwise, without the prior written permission of the publishers, except; for the
quotation of brief passages in reviews.

Typeset by Javafame Computer Services, Saffron Pane, Hall Road, Lavenham,

Sudbury, Suffolk CO 10 9QU

Photographs by Ambrose Greenway and Laurence Dunn

Printed in England by O'Sullivan Printing Corporation, Southall, Middlesex, UB2 5LF

ISBN 1 870077 06 7
!8t.

CONTENTS
Drawings
Illustrations
Acknowledgements
Foreword

page
Chapter One The Nature f Water 1
Introduction, Density, Viscosity, Surface Tension.

r Chapter Two Waves 13


t Ocean Waves, Wave Propagation, Waves in Shallow Water,
Waves and Small Craft, Speed and Power, Resistance Prediction,
I Over the Hump, Waves and Currents.

Chapter Three Proximity 37


Bernoulli, Squat, Stability, Bank Effects, Ship-ship Interaction,
Shoals and Steering, Virtual Mass.

Chapter Four Foils, Fins and Blades 47


Lift and Drag, Lifting Surfaces in Water: Cavitation,
Hydrofoils, Fins, Rudders, Propellers, The Tip Vortex,
CP Propellers, The Paddle-Wheel Effect.

Chapter Five Turning Corners 62


The Hydromechanics of Steering, Directional Stability,
The Pivoting Point.

Chapter Six Fairness and Flow 68


Fairness and Fairing, Flow: The Boundary Layer,
Flow around Bodies, Separation of Flow, A Case Study,
Transom Stems, Trailing Vortices, Why Fairness?

Chapter Seven Stability and Stabilisation 80


Weight and Buoyancy. GZ Curves, The Metacentric Height,
The Inclining Experiment, Ships in Waves, Following Seas,
Free Surface, Stabilisers.

88
Sources of Vibration, Damaged Propeller, Wake Effects,
Pressure Effects, Cavitation, Singing, Wave-induced Vibration,
Diagnosing Vibration Causes.

Index

iii
DRAWINGS
Fig. page

1 The boundary layer 5

2 Pressure in a bubble due to surface tension 10

3 Wave propagation in deep water 15

4 Wave generation by a ship 18

5 Effects of shoal water 21

6 Broaching 24

7 Prediction of resistance from model experiments 28


u
Q O/T1-.Q +iii oiiaiiuw
sViol
I m t r
wa.L\_-x
trvn-t-aw
UU
OQ

9 Bank effects 41

10 Ship-Tug interaction 42

11 Steering showing the effect of shoal water 45

12 Forces on a submerged body in flow 48

13 Rudders at different helm angles 53

14 Propeller blade shapes 56

15 The propeller tip vortex 57

16 Steering, the forces required 63

17 Force on a yawed ship 64

18 The pivot point 67

19 Form drag 71

20 The stability curve 81

21 Loss of stability on a wave crest 85

iv
ILLUSTRATIONS

M.V British Wye in a North Atlantic storm


Photograph by Captain J Thomson, Master

The boundary layer

Spray

A ship's wave pattern

A destroyer at speed

Waves breaking

A speed boat planing

A hydrofoil at speed

Damage due to sea conditions in the region


of the Agulhas current

Naval replenishment at sea

Propeller under test in a cavitation tunnel

A modern skewed propeller

A full-bodied ship's bow

A bulbous bow on a thin fast ship

Slab-sided construction
ACKNOWLED GEIMDE/NTS

The relationship between ship and water is a topic of eternal


fascination and not a little mystery. I've tried to cast just a little hesitant
light on some of the fringes of the subject; there's a lot more yet to be
explained, once we understand it all. I should like to thank all those
individuals, too numerous to be named, whose conversation and
correspondence has helped me in putting this work together. The
contributors to SEAWAYS correspondence columns in particular have
provided a constant stimulus; may they keep on writing!

vi
FOREWORD

Navigators need to be forward looking. They are concerned about their


destination and planning their voyage. They constantly adjust and update
their course in the light of information received.

Pilots have a more intimate and immediate series of problems to overcome


as tbiey handle ships safely in and out of port, using their experience,
almost instinctively to manoeuvre ships in confined waters.

The maritime tradition stretches back thousands of years but, in general,


seafaring is a young man's profession and the skills needed are constantly
being acquired by each new generation of trainee.

Awareness and anticipation, the knowledge and experience to plan ahead


wilfh confidence and safety and to be flexible in the light of changing
circumstances are part of seamanship and essential for those engaged in
sea transport Seatime and certification are the two traditional ways of
obtaining these qualifications but it is not always obvious what needs to
be learnt

Ships and Water is a exemplary book which sets out to explain what the
mariner instinctively feels in plain language. It demonstrates that the
interplay of a ship with the sea is by no means straightforward and that
what may appear as common sense when derived from living on land may
be a misleading guide to the navigator.

Having been a pilot for most of my working life in one of the world's
busiest estuarial ports, I have had to learn, tentatively and by experience,
what Mr Paffett so clearly explains in the text.

It would have been useful to have had this book when I was a young officer
and trainee pilot, Its fascination lies in the fact that those who have been
most seasoned with salt are the most likely to appreciate the subtlety of
the concepts being discussed. The plain language and unequivocal style of
the author makes this the most readable of books.

Mr Paffett was made an Honorary Fellow of The Nautical Institute for his
belief and commitment in involving the mariner in the operational design
of ships. This book demonstrates perfectly why this honour was so richly
deserved.

Commodore G.G.Greenfield RD* FNI RNR


President, The Nautical Institute 1987 - 1990

vii
CHAPTER ONE

SHIPS AND WATER

Introduction

We have water all about us. We drink it, we swim in it, we wash in it and
generally enjoy its benefits while giving little thought to its physical
properties.

However, when we come to design or operate a ship we find that her


performance, and indeed survival, depend critically upon certain
characteristics of water, which deserve in this context a closer
examination.
j

The properties of water which mainly concern the mariner and ship
designer are density, viscosity and surface tension. (There are others, such
as specific and latent heats, refractive index, vapour pressure and so on
whichl bear upon meteorology and hence affect life at sea; but in the
preseiit discussion we shall concentrate upon the first three named).

Density

The density of a substance is the mass per unit volume, usually stated in
grams per cubic centimetre, or - more conveniently for ships - in tonnes
per cubic metre; the figure is the same. The density of pure fresh water is
close to 1-0 tonnes per cubic metre, and of seawater around 1-025 t/m3.
Seawater density varies somewhat, particularly in coastal and estuarial
waters where river water dilutes the salinity, but the figure of 1-025 t/m3
is an average for the open ocean and is usually used in the design of
seagoing vessels.

Reference is sometimes made in technical publications to specific


gravity. This is a pure number, being defined as the ratio between the
density of the liquid considered and the density of pure water at the same
temperature. Thus, the SG of sea water might be given as 1-025. Modern
engineering practice is to use density in preference to specific gravity.

As with every other known liquid, the density of water increases with
a fall in temperature. Water, however, is unique in that it reaches a
maximum density at 4 degrees Celsius. On cooling below this the density
begins to decrease again, and the solid formed on freezing - namely, ice -
has a density even less than that of the unfrozen water, so that it floats on
the surface. If water behaved like almost eveiy other material known to the
physicist, ice would sink and navigation in high latitudes would be
complicated by a continually varying and indeterminate accretion of ice on
the seabed; charts would be practically useless and grounding a constant
danger. The relatively minor hazard from ice floes and bergs is probably an
acceptable price to pay for a predictable sea-bed.

The density of water provides the familiar phenomenon of buoyancy. If


we consider a single cubic metre of water lying somewhere in a wide
expanse of still ocean, we note that it is stationary; it is not accelerating
anywhere. Therefore the resultant of all the forces acting on the surface of
this cubic metre from the surrounding volume of ocean must be exactly
equal and opposite to the downward force exerted by gravity on the cubic
metre, namely 1-025 tonnes weight. If we rapidly remove our cubic metre
of water and replace it by a solid body of the same shape and volume, the
surrounding water will be unaffected by the change and will continue to
exert toe same upward force on the solid body. In conventional wording,
the solid body will experience a buoyancy force equal to the weight of the
displaced fluid - the principle established by Archimedes of Syracuse
around 200 BC, following a research programme reputedly carried out in
his bath.

The most basic requirement of naval architecture is to balance ship


weight against buoyancy force. The submerged volume of the ship up to
the designed draught can be calculated from the hull drawings;
multiplying this volume - the displacement - by the seawater density gives
the buoyancy force; if this equals the all-up weight of the loaded ship, all is
well and she floats at the right draught. It must be admitted that there are
cases on record where new ships have turned out heavy and have floated
deeper in the water than intended.

It is worth noting that the draught of a given ship depends upon the
density of the water in which she is floating. If a ship goes from salt water
into fresh, she sinks a little into the water until the immersed vojume is
increased by about 2-5% so as to restore the buoyancy force\to the
magnitude that it had in salt water. This can be a matter of some practical
importance in the case of a ship entering an estuary or river mouth with
only a small clearance under the keel; the sinkage due to the density
change could be enough to cause grounding. >

Density has to be measured with particular care when one is making


a "draught survey", i.e. measuring the ship's draughts accurately so ks to
obtain an exact figure for the displacement mass. (This is sometimes done
as a check on the weight of cargo loaded). From the draughts (forward,
amidships and aft) and the ship's hydrostatic curves one can deduce the
submerged volume; multiplication by the density gives the ship mass. It is
important that the actual water in which the ship is floating be checked for
density by means of an accurate hydrometer. A range of water samples
should be taken from both surface and keel depth, forward and aft, and a
mean value taken for density, since density can vary considerably from
point to point in some harbours.

2
The submariner has particular cause to take an interest in seawater
density. A submerged submarine has no free water-plane, so that if the
boat is to be able to hover at constant depth she must adjust matters so
that the all-up weight of the vessel is exactly equal to the buoyancy force.
The displacement volume is fixed by the construction of the vessel, so that
the submariner can only achieve the necessary balance by taking in or
discharging ballast water. If the vessel has headway some vertical force for
fine adjustment is available by operating the hydroplanes, but if she is
stationary the planes have no effect and the ballast has to be adjusted very
critically.

, Even when perfect balance has been achieved, the stopped submarine
is still in an unstable situation, arising from the compressibility of the
steel hull. If she goes deeper, the increasing hydrostatic pressure
compresses the steel hull, so decreasing its displacement volume. The
water itself is also compressed, so increasing in density. Unfortunately, the
loss in hull volume more than offsets the gain in water density with
increasing depth. Thus, if a stationary submarine sinks a small distance
she loses buoyancy and begins to accelerate down towards collapse depth;
if she'rises, she gains surplus buoyancy and accelerates towards the
surface. Successful hovering demands the ability to detect incipient
changes in depth and to make rapid small adjustments in ballasting to
check rise and fall.

Seawater density in the open ocean is fairly constant, but in coastal


waters fed by river outflows the density can vary considerably. A
submarine passing from salt to brackish water in a river mouth can begin
to sink rapidly. Sometimes stratified water is encountered, where a layer of
fresh water overlies salt, or warm water tops cold; here a submarine can
sit stably on the interface between the two layers. Car
though, as density interfaces can move up and down through considerable
distances in a kind of wave motion. Great ponderous waves, resembling
ocean surface waves but moving more slowly, have been observed in the
density interface where salt Mediterranean water slides out past Gibraltar
under the less dense Atlantic surface water.

The submariner has another reason for taking some interest in the
density of the water which surrounds him. Sound waves travel in straight
lines at constant speed in a medium of constant density. But the velocity
of sound varies with the density of the medium, and if this is non-uniform
the sound will travel in paths which are curved rather than straight.
Seawater density depends upon both salinity and temperature, both of
which can vary considerably with depth. One consequence is that the
beams of sound emitted by the hunting frigate's sonar transmitter may be
curved in a complex manner as they radiate from the ship; indeed they
may be refracted upwards to the surface again without reaching the depth
of the submarine. Thus it can pay both hunter and hunted to measure
and understand the ocean density distribution; the submariner can exploit

3
the refraction phenomenon to hide from the searching sonar, the frigate
officers need to know the limitations of their equipment in planning their
search pattern. Surface ships are sometimes equipped with
bathythermographs, instruments which enable them to measure the
temperature variation between keel and seabed, so that sound behaviour
in the ocean can be predicted.

Viscosity

A ship is a solid body, surrounded and supported by a liquid. Let us


consider those properties which distinguish a solid from a liquid. Common
experience tells us that a solid has shape, while a liquid has not. But this
shapelessness of liquids is not absolute; one has to expend energy to make
a volume of liquid change its shape, and the viscosity of the liquid can be
regarded as a measure of the mechanical work done in effecting this
change. One can think of the physical effort needed to scoop up a spoonful
of treacle, compared with that for a spoonful of water. We speak of treacle
as being more viscous than water.

We can look into this a little more scientifically. Stress is a measure of


the force per unit area sustained by a body, while strain is a measure of
the distortion caused by that stress. Stress is of two kinds: direct stress,
which changes dimensions - elongation or compression - and shear stress,
which changes shape. A cube subjected to direct stress will stretch or
squash, but its corners will remain right-angled. However, a cube
subjected to shear stress will distort into a lozenge shape, and the change
in the angle at the corner of the cube, eg from 90 to 89 degrees, is the
shear strain resulting from the stress. We can think of the cube under
shear stress distorting like a pack of square cards, each card sliding
through a small distance relative to its neighbour.

Now a solid cube subjected to a constant shear stress will take up a


strained shape and there stop, with its corner angle steady at (say) 89
degrees, We could call this a shear strain of one degree. But a similar cube
of liquid will distort and go on distorting as long as the shear stress is
applied. The shear strain can still be measured; it is no longer constant,
but can be shown to be increasing at a constant time rate, at so many
degrees per second/ The time-rate of change of shear strain, moreover, is
in general proportional to the applied shear stress, and the constant of
proportionality is the viscosity. We can think of the viscosity of the liquid
as the shear stress needed to cause a given change of shear strain per
second. The more viscous the liquid, the more stress we need to apply to
generate a given rate of distortion. This accords with our everyday
experience with treacle,and it also enables us to express this quantity
viscosity in figures, and even to calculate the forces required to distort a
liquid at a given rate.

4
This brings us back to practicalities. The mariner is engaged in
distorting a liquid, namely sea-water, at some cost in fuel bills. The
distortion takes place in the boundary layer, the layer of water adjacent to
the ship's hull plating. The skin of water immediately in contact with the
plating wets it and travels along with it at ship's speed, say 10 knots. A
few millimetres further out the water is sliding somewhat sternwards
relative to the ship, so that its speed over the ground is say 9 knots.
Further out still the water is moving at 8 knots, and so on. Eventually, a
metre br so out from the ship - depending on the ship's length and the
distance from the bow - there is no measurable movement parallel to the
ship. (Fig. 1)

Figure 1 The Boundary Layer

From this we can see that there is a continual sliding motion of layer
across layer; cubes are being bent to lozenges and there is a continual
change of stiear strain in the water. Because the water has a finite
viscosity it follows that there must exist a shear stress. Integrated over the
whole wetted surface of the ship, this results in a total horizontal force,
exerted in the sternwards direction. This is usually referred to as the skin
friction, or the viscous drag. It represents something like 80% or more of
the total drag force on most merchant ships, the remaining 20% or less
being made up of wavemaking forces, appendage drag and so on, which
are considered in a later chapter.

5
A ship moving through the water carries with it a frictional "boundary layer".
This grows in thickness from zero at the stem to perhaps a metre or more at
the stern. The boundary layer should not be confused with the froth visible on
the sea surface; this arises mainly from the breaking of the bow wave crest.

As the skin friction, or drag of viscous origin, accounts for some 80%
or more of the fuel bill, it is not surprising that great efforts have been
made to reduce it. Useful gains have come from the pursuit of smoothness;
quite small roughnesses on a ship's bottom increase the drag significantly,
hence the efforts made over the years to prevent the growth of weed and
shells on the plating. Even organic slime has been shown to put up the
drag. Smoothness and cleanness certainly pay; however, there is a well-
defined minimum below which the skin friction cannot be reduced^ Even
with the highly-polished mirror surface sought by racing-yacht
enthusiasts, there will still be a finite shearing rate in the boundary layer,
hence a viscous shear stress and a consequent drag force on the vessel.
The most that the ordinary merchant ship operator can do is to invest in
the best available anti-fouling bottom composition, and to ensure that dry-
docking is carried out without delay when growth finally makes it
necessary.

6
Efforts are made from time to time to reduce the frictional drag below
the basic minimum by interfering with the boundary layer structure. The
most popular idea, periodically reinvented, is to inject a layer of air
between the bottom plating and the water, in order to achieve "air
lubrication". Repeated attempts to do this, on both model and full scale,
have been made without much success. The method would work if the air
could be persuaded to form a continuous thin layer between water and
bottom, but in practice the air very quickly breaks up into a stream of
discrete bubbles which roll along embedded in the boundary layer, leaving
the bottom plating as wet and as exposed to shear stress as ever. The
writer recalls attending trials of a seagoing vessel where the speed
difference between "air-on" and "air-off was not measurable.

One attempt at air lubrication did lead to success of a sort. The


inventor pushed up the air flow to a prodigious amount, and obstructed its
escape by fabric skirts around the periphery of the craft. The air layer
formed one enormous trapped bubble a metre or so thick, and the result
was the Hovercraft. Frictional drag was undoubtedly eliminated (partially
so, in the case of sidewall craft) but at the cost of considerable air-blowing
horsepower.

Another method of modifying the boundary layer structure is by


means of the so-called long-chain molecules. There are certain substances,
organic polymers of high molecular weight, which have a string-like
molecular structure, and have the property of reducing the net frictional
force if they are injected into the water of the boundary layer. The actual
mechanism has not been clearly explained, but it appears that the long
molecules by their keying action modify the distribution of shear strain in
such a way that there is a small reduction in the shear stress exerted on
the solid boundary, although the viscosity of the water itself is not
reduced. The phenomenon has been used commercially in reducing the
pressure drop along pipe lines, such as fire hoses. The effect has been
demonstrated on ship models, and on the full scale in a minesweeper. The
trials showed, however, that the cost and weight of the chemicals used -
and the consequential pollution of the ocean - would be out of all
proportion to the marginal reduction in skin friction achieved.

Ship designers and operators will continue to seek ways of reducing


skin friction on ships, and no doubt there will be further ingenious
inventions/However, the viscosity of salt water, like its density, is one of
those quantities we have to live with. We will still have to pay the price,
one way or another, when we want to move water about.

Starfaee Tension

We have all met, at some early age, the parlour trick whereby a sewing
needle is set floating on a dish of water. The steel, we are told, is kept from

7
sinking by the surface tension of the water; and, indeed, close examination
shows that the surface of the water near to the needle is deflected
downwards, as though the needle were sitting on an elastic skin which
sags under its weight. When we go swimming, however, no supporting skin
saves us from a ducking. The phenomenon only seems observable on the
near-microscopic scale. What is this curious tension, and can it affect a
thing as large as a ship?

The molecules of any liquid attract one another. A molecule well


inside a body of liquid experiences attractions from the other molecules in
all directions; the forces cancel out and the internal molecule is
unaffected. A molecule on the surface, however, experiences forces only on
one side, tending to hold it back into the main body. The effect is that any
local protrusions are pulled inwards and dents are pulled outwards; the
surface behaves, in fact, as a skin under tension. The result is that any
isolated body of liquid will tend to take up a compact form where all
cohesive forces balance - namely, a sphere. Everyday experience confirms
that small particles of water form spherical drops.

If we consider a line drawn across the surface of a body of liquid, it


will be evident that the particles on one side of the line are attracted to
those on the other. If we cut the liquid along the line, we find that a finite
force is needed to separate the two halves. The force T per unit length of
line is measurable and is characteristic of the liquid; for clean water it is
75 micro-newtons per millimetre.

If our line has a length L, the force to separate the liquid is TL. If we
pull the edges apart through distance D we do mechanical work equal to
TLD, and we create a fresh surface of area LD. The energy to create unit
area of new surface is thus TLD/LD, namely T. Thus we have a second
way of regarding surface tension, namely as the energy which must be
expended in creating a new liquid surface ofunit area.

The direct forces exerted on a ship by the surface tension of the ocean
surface are minute and can be disregarded. The phenomenon cannot be
written off as irrelevant, however; surface tension begins to matter when
we come to consider seakeeping, and in particular spray generation. A
one-metre cube of water has a surface area of 6 square metres. However, if
we break this body up into one millimetre cubes the surface area will be
6,000,000,000 square millimetres or 6000 square metres, around 1112
acres. To create all this new area we have had to do significant work
against surface tension. The energy to do this is freely available ih a ship
at sea in rough weather, and it comes as no news to the mariner that a
tonne of seawater hitting the bows can easily be .dashed into one billion
spray droplets each of volume 1 cubic mm.

8
One billion spray droplets

If we try to reproduce the ship seakeeping behaviour with a scale


model in the laboratory, however, we find that the spray behaviour does
not "scale". We can reproduce the wavemaking behaviour and wave
response easily enough by setting the model speed at a value
proportionate to the square root of the linear scale - a 1/100 scale model
is driven at 1/10 ship speed. But at these model speeds there is not the
energy present to break the water into linear-scaled spray. The model
driven into waves throws up only a few blobs resembling footballs when
scaled up to ship size. The inference is that, while models can give us
valuable information on powering, steering, motions and wavemaking, they
are of little use in predicting spray generation on the full scale.

A practical instance arises in the case of the so-called "spray rails"


commonly fitted on the sides of small fast craft. These rails are intended to
deflect outwards and downwards the film of water which would otherwise
climb up the side plating and break into clouds of spray, obstructing the
view from the bridge. Spray rails properly sited and shaped can be highly
efficacious, but design has to rely mainly on full-scale experiments and
experience, because model work is unrepresentative for the reasons given
earlier.

9
Once water is broken up into small particles, the surface tension
forces tend to pull each particle into the familiar spherical drops which
reach us as rain or spray. The same forces act when we interchange the
air and the water; a small air bubble in water assumes the same spherical
shape under the action of the same surface tension forces. If a bubble is in
equilibrium the air or gas pressure inside it has to balance not only the
hydrostatic pressure of the surrounding liquid, but also the inward
pressure in its skin, much as the air pressure in a child's balloon balances
the tension in the rubber (Fig. 2). From the geometry of the sphere, we can
show that the internal pressure P to balance surface tension T is equal to
2T/R where R is the radius of the sphere. Since the surface tension T is a
follows that the bubble pressure P becomes larger as the
bubble radius becomes smaller. In the limit one needs an infinite pressure
to create a new bubble from

Surface tension leads


to spherical droplets

FORCE DUE TO FORCE DUE TO


INTERNAL PRESSURE SURFACE TENSION

=AREAX PRESSURE =CIRCUMFERENCE X S.T.


?rR2P =2rrR.T

FOR EQUILIBRIUM OF jtR2P= 2TTR.T


SEMI-BUBBLE: .'.P= 2 %

Figure 2 Pressure in a bubble due to surface tension '

This has some bearing upon the phenomenon of boiling. It may be recalled
from the school physics lessons that, while tap water could be boiled in a
glass flask over a Bunsen burner easily enough, pure de-aerated distilled

10
I .
L SiBLiU:

water only boiled with a succession of bumps and bangs sometimes severe
enough to break the flask. This happened because, in the absence of
nuclear air bubbles in the water, steam bubbles could only be created by
building up a vapour pressure large enough to balance T at infinitesimal
values of R. Once a bubble was born, the pressure needed to balance T
rapidly fell with increase in R and the internal steam pressure became
suddenly available to accelerate the bubble walls outwards - hence the
bumps. Dunking a piece of dry porous ceramic in the water provided a
stream' of air bubbles of finite diameter; these served as nuclei and
enabled boiling to continue quietly. In the domestic context, water from
the kitchen tap arrives full of air bubbles and usually boils without
trouble. However, one can sometimes get a bump or two from the kettle if
one tries to reboil water which has earlier suffered prolonged boiling and
been allowed to stand.

What possible relevance has boiling to the behaviour of water under a


ship? There is, in fact, a very close connection when we come to consider
the flow through propellers. On the suction surface of a working propeller
blade the water pressure can fall to a very low value, well below
atmospheric. It will be recalled, again from school physics, that boiling
temperature depends upon pressure, and that if pressure is reduced to
near vacuum water can be made to boil at room temperature. This can
happen on the propeller surface; if it does, the water boils and bubbles
containing water vapour are formed. The bubbles travel downstream until
they reach a region of higher pressure, where they collapse. In collapsing
they create shock waves in the water, which may be severe enough to
cause mechanical damage to the propeller blade, or to any solid surface on
which the bubbles impinge, such as the leading edge of the rudder. This is
the phenomenon of cavitation, which is considered in more detail in a later
chapter.

Cavitation is highly undesirable, on account of the mechanical


damage which it can cause, and also because of the noise which it radiates
through the water. The phenomenon can be demonstrated in the
laboratory, but accurate reproduction of ship-scale behaviour requires
attention to be paid to the boiling behaviour of the water discussed above.
Cavitation vapour-bubbles form easily enough in the sea, where the water
is well loaded with air bubbles and other minute particles which form
nuclei for bubble formation. The laboratory tank, filled with clean fresh
water, may or may not have representative nuclei; if they are absent,
cavitation may be delayed and unrepresentative of the full-scale propeller.
Laboratory staff must thus pay attention to the boiling behaviour of the
tank water, a property depending upon surface tension.

This treatment of surface tension has been somewhat simplified.


Closer consideration has to allow for the fact that the surface tension at an
interface depends upon the physical chemistry of both substances involved
~ not only of the liquid, but also of the air or other gas. If three substances

11
are involved, as with a bubble or drop in contact with a solid surface, the
behaviour of the bubble or drop is affected not only by the surface tension
of the liquid, but also by the properties of the solid surface, in particular
its "wettability". A drop of water falling on most surfaces, a steel deck for
instance, will quickly spread out and wet the surface, but there are some
substances which are "hydrophobic" or difficult to wet, and drops of water
falling on them will roll about like marbles. Certain plastics such as PTFE
and, to a lesser extent, polyethylene, are hydrophobic.

A permanently hydrophobic paint finish would be highly attractive for


ships operating in high latitudes, as a means of avoiding the icing-up of
upperworks by spray freezing on contact. Unfortunately, anti-wetting
characteristics are difficult to preserve, as such properties in a solid, like
surface tension in a liquid, are easily affected by very small additions of
foreign substances. A thin film of oil or grease makes a steel surface
difficult to wet; but a small addition of soap or detergent to the water acts
the other way and helps to make the surface wettable. Inadvertent
contaminants of solid surfaces, even when present only in minute
quantities, can affect wettability either way; no news to the washer-up or
deck scrubber.

A thin film of contaminant on a liquid surface can modify its surface


tension. This effect is clearly visible at sea, when an oil slick suppresses
those very fine ripples or "capillary waves" caused by a light wind. These
waves are discussed in a later chapter.

12
CHAPTER TWO

WAVES

Ocean Waves

Waves on the ocean surface are only too familiar. They slow down the ship,
stress jthe hull and nauseate the passengers. Yet the actual mechanism of
wave propagation is a matter of great complexity, the detailed description
of which still exercises the skill of mathematicians. We discuss here those
features of ocean waves which concern directly the designers and
operators of ships.

A body of calm water in repose, undisturbed by wind or ships, has a


perfectly flat smooth surface. (A purist here might point out that the
surface is not strictly flat, it is spherical like the Earth; but in the range of
ship dimensions we are considering here "flat" is an acceptable
approximation). If by some means a local patch of water is raised above, or
depressed below, the mean surface and released, the disturbed body of
water will move vertically until the surface is restored to the mean level.
The energy released by the upheaval will be radiated from the disturbance
in the form of waves which move away as expanding concentric circles. So
much is obvious to anyone who drops a pebble into a pond.

What may not be so obvious is that there are two quite distinct
agencies acting to restore the disturbed surface to the mean level: gravity
and surface tension/The action of gravity is more clearly seen with water
in a glass U-tube; if the levels are disturbed the water will rock back and
forth until the two arms settle at the same level again. The ocean can be
thought of as a series of linked U-tubes, the swing of each being
communicated to its neighbour as energy passes from tube to tube, gravity
providing the restoring force throughout

The action of surface tension is not quite so obvious. As explained


earlier, the surface layer of water behaves like a skin or membrane under
tension. If it is disturbed locally, eg by being pulled up into a protuberance
and released, the disturbance will travel outwards along the surface like a
kink in a rope held in tension. The restoring force here is provided by the
surface tension stress in the liquid, resolved normally to the mean surface
where the local surface is deflected away from the mean. Surface tension
waves could in theory travel across the surface of a body of water in outer
space, where gravity waves could not exist in the absence of gravity forces.

Gravity waves and surface tension waves obey quite different laws.
> The physics of the situation is such that surface tension waves on water
are only observable on the small scale, with wavelengths of 25 mm or less,
and amplitudes of a millimetre or so. These waves, sometimes referred to

13
as ripples or capillary waves, are of no practical concern to the ship
designer, but the mariner may sometimes have seen them at sea, notably
as "cat's paws" when a light puff of wind momentarily ruffles the surface of
an otherwise dead-calm sea. It is notable that such waves damp out very
quickly, disappearing almost instantly when the breeze passes. Their
dependence upon surface tension is evidenced by their suppression when
an oil slick or other surface contaminant modifies the surface tension of
the water.

It is the local absence of capillary ripples which gives the sea a


"greasy" look and sometimes first draws attention to local contamination.
In the past oil was sometimes discharged deliberately on rough waters in
the belief that it would calm then. It probably had little direct effect on the
large amplitude gravity waves and swell, to be discussed below, but the
suppression of ripples may have reduced the amount of spray and spume
being blown off the sea by the wind, giving a calming impression, and the
smoothing of the surface probably slowed the energy transfer from wind to
the shorter gravity waves, leading in time to some reduction in their
amplitude. The overall benefit would have been pretty marginal. Deliberate
oil discharge would not be advocated nowadays as a practical means of
calming troubled waters.

Gravity waves are the familiar ones, with wavelengths ranging from a
few centimetres to half a kilometre or more. The energy in waves comes in
the first instance from the wind. A wind commencing to blow over an
initially calm sea produces at first ripples, then short gravity waves, then
longer and higher waves, all travelling in the direction of the wind. If the
wind drops the wave trains continue travelling across the ocean. The
shorter waves damp out fairly quickly, energy being dissipated by a
viscous mechanism within the body of the wave, but the longer waves -
referred to as swell, once the wind has gone - lose their energy more
slowly, and may travel on for thousands of miles before finally spending
their energy by breaking on the shores of some distant land. The swell
surging up Cornish beaches probably drew its energy from the winds of
some Caribbean storm.

The energy in waves has fascinated generations of inventors, all


seeking to turn this seemingly wasteful outpouring of nature into useful
electricity. Various ingenious devices have been made to work, and no
doubt more will appear. But there is one unavoidable snag with any wave-
energy recovery device: sheer physical size, and hence capital cost. The
energy flux available is predictable at so many kilowatts per kilometre
frontage, averaged over the typical year. The actual figure depends on the
sea area considered, the western coasts of the UK being well served by the
heavy swells coming in from the Atlantic. But calculation will show that if
the energy recovered is to be large enough to be commercially interesting
to the generating industry, the device will have to span an ocean front of
many kilometres. This means a coastline with long expanses of

14
I8 L I u / E

environmentally objectionable concrete works; or, if the device is a floating


one, a continuous line of moorings in coastal waters, with the barges or
other floating mechanisms constituting a massive obstruction to
navigation, and an active hazard too every time a unit breaks adrift in a
storm. There is probably a future for wave-energy recovery devices in
generating electricity in small amounts for isolated consumers, eg for
powering a remote lighthouse or navigation buoy. But for feeding the
national grid in an industrial country like the UK, wave energy compares
unfavourably with the other "renewables", wind and tide.

Wave Propagation
Sound waves and radio waves travel at velocities which are constant for a
given medium, regardless of wavelength. Ocean waves are quite different;
the wave crest velocity depends upon the wavelength - the longer the
wave, the faster travel the crests. In deep water the crest velocity,
sometimes called the celerity, is proportional to the square root of the
wavelength. A wave 100 metres long travels at 24iu knots, a wave 400
metres long travels at 48^2 knots.

An observer trying to judge the celerity of ocean waves by watching a


single crest soon encounters a difficulty: no single crest seems to exist for
very long. If one watches an individual crest closely, one will see that it
grows in height to a maximum and then shrinks to nothing. While this is
happening the following crest grows to take its place, only to fade in turn
to yield to a third crest, and so on. In fact, if a complete train of waves is
watched - not often practical at sea, but quite possible in a laboratory
tank - it can be seen that the leading wave is constantly dying out, its
place being taken by the second, then the third; the train as a whole
moves forward at a speed which is less than that of the individual crests,
which advance through it from rearguard to vanguard. (Fig. 3)

WAVELENGTH
CREST VELOCITY

GROUP VELOCITYVaC

Figure 3 Wave propagation in deep water

15
The wave pattern contains four elements - the bow transverse system, the
bow divergent system, and the corresponding two systems at the stern.

16
iiSL

The mean speed of the whole train is called the group velocity, and it
can be shown that in deep water it is equal to half the celerity of the
individual crests. From mathematical hydrodynamics it can be shown that
the total energy of the wave train advances through the ocean at a speed
equal to the group velocity, not the crest celerity.

It follows from the above that a single isolated wave crest cannot
travel unchanged across an expanse of deep water; should such a wave
somehow be created, it will at once degenerate into a group or train of
waves, advancing at the group velocity. If one wants to propagate a wave
unchanged, one has to feed it constantly with energy, which will be
dissipated into a lengthening train of waves following on astern.

v All this has direct relevance to the behaviour of ships. A moving ship
is a disturbance which generates waves. Observation of any vessel under
way in deep calm water will show that she is accompanied by a pattern of
\i7Q
TT A7A
U.T v rrpc+o
^Oi^u luViinh mairfairi
l i t uv < t i l +V>pir
l a i v i i ciyp
\ i * prifl
CU-^unncitinn linphQnrfpH
UaIV/AAH I rplQ+nrp
a wAu.^ wfn Lv
the ship. It follows from what has just been said about energy flux and
group velocity that the ship's unchanging wave pattern can only be
maintained by feeding energy into it. The rate of energy flux, energy per
unit time, has the dimensions of power or force times velocity. It can be
expressed as the ship's velocity through the water multiplied by a force.
This force is the wave-making resistance of the ship. It can be reduced by
making the ship thinner, so fining the angle of entry at the bow; hence the
fine lines of fast displacement vessels such as destroyers.

The actual wave pattern generated by a ship is a matter of some


complexity, even beauty, as one can see from aerial photographs of ships
under way in calm water. The pattern contains four elements: the bow
transverse system, the bow divergent system, and the corresponding two
systems at the stern. The two transverse systems interact. If we consider a
ship starting from rest, while she is moving at slow speed the bow
transverse system, necessarily moving at the same speed, will have quite a
short wavelength, giving several crests evenly spaced along the ship's side
from bow to stern.

The bow system starts with a crest near the bow. The similar stern
system starts with a trough near the stern. As speed increases, so does the
wavelength of the two systems. Eventually the first following crest of the
bow system coincides with the initial trough of the stern system; the two
systems to some extent cancel one another out, so reducing the amplitude
of the combined transverse system following astern of the ship. This is a
favourable situation, making for reduced wavemaking resistance, and so it
is sought after by ship designers.

If we increase the speed further, we get to the situation where the first
trough of the bow systems coincides with the stern system trough. The two
systems reinforce one another, maximising the wave height in the
transverse system and so the wave drag (Fig. 4). This is the "main hump"
which naval architects usually tiy to avoid by judicious choice of hull
length and shape.

(a) THE TRANSVERSE WAVE SYSTEMS AT LOW SPEED


NOTE: SHIP SFEED=WAVE CELERITY WHICH
DETERMINES L

(b) THE WAVE SYSTEMS AT "HUMP" SPEEUTHE FIRST


TROUGH OF BOW SYSTEM COINCIDES WITH START OF
STERN SYSTEM

Figure 4 Wave generation by a ship

Hump dodging tends to drive designers working in this speed range


towards long thin ships, so increasing the spread between bow crest and
stern trough. Sometimes, though, the hump cannot be evaded, and many
destroyers at their top operating speed are condemned to working
somewhere near the hump - evidenced by the sometimes spectacular wave
pattern at 30 knots.

Above the main hump speed there are no further humps.


Wavemaking resistance does not disappear, but the nature of the flow
round the ship changes. Hydrodynamic (as distinct from hydrostatic) lift
forces become significant, and the vessel begins to plane, The long thin
shape is no longer advantageous, and indeed vessels which are very fast in
relation to their length - as are many power-driven small craft - may have
length-to-beam ratios as low as 3, compared with the destroyer's ratio of 9
or so, and the ordinary cargo ship's 6 or 7.

The above has been a grossly simplified account of wavemaking


resistance, Those who want to go into the topic in more detail will find that
designers have developed mathematical methods of great subtlety for
generating hull shapes which will minimise the wavemaking drag of the

18
ship's hull at her designed operating speed. One should note that the
optimum shape is a function of speed; the shape which will minimise
resistance, and hence fuel consumption, at one speed will not in general
be the most economical one at another speed. Tankers designed for 16
knots are sometimes operated at 8 knots to save fuel; but even more fuel
could be saved by using a hull designed from the outset to carry the same
cargo at 8 knots.

A destroyer at speed. Sometimes the "hump" cannot be evaded and many


such vessels at their top operating speed are condemned to working
somewhere near the 'hump*. The hull forms have necessarily to be fine to
minimise wave-making,hence these vessels may have a length/beam ratio as
high as 10. ..,',. .,...

Wave drag is an important consideration in warships, fast passenger


liners, ferries and many small craft. However, it is only of secondary
importance in most cargo-carrying merchant vessels of moderate speed. In
these ships hull drag due to viscous effects usually referred to as skin
friction - far exceeds the wavemaking resistance. Where friction is the
main factor/ it pays to keep the surface area as small as possible; this
favours the short fat ship ~ it will be remembered that the solid of least
surface area for a given internal volume is the sphere. In minimising
friction it is also important to keep the surface smooth, a point discussed
in the section on viscosity.

19
Waves in Shallow Water

In discussing waves so far we have been careful to qualify the water as


deep - that is, exceeding in depth several times the length of the longest
waves considered. Theoretical hydrodynamics, confirmed by observational
evidence, shows that wave behaviour is considerably modified as the water
becomes shallower; mariners sometimes say that the waves "feel the
bottom". The effect becomes significant at depths of half a wavelength or
less.

Consider a wave train initially travelling in deep water. The train now
enters shoaling water. The first effect is a reduction in the crest celerity.
Since the frequency - ie, the number of crests arriving at a fixed point in
unit time - remains unchanged, It follows that the waves must become
shorter.

The change in velocity produces a refraction effect, like that


experienced by a beam of light entering a block of glass. If the shoal water
is on, say, the left of an advancing wave front, the wave crests on the left
will lag behind those still travelling in deep water, and the wave front as a
whole will wheel leftwards until it is moving towards the shallowest area.
Thus, if we have a coastline running East and West, bounded by a
shelving beach facing South, and a wave system enters from the West, the
waves will be progressively bent round towards the North until the crests
lie almosi^ This is why the
holidaymaker on the sands always sees the breakers coming almost
straight in towards him, whichever way the waves are travelling further
out at sea.

Because of this refraction effect, isolated patches of shallow water in


mid-ocean can have a focussing effect rather like a lens. If we consider a
straight wave front approaching a shallow patch, the centre of the front
will be retarded over the patch, while the two wings will be swung round
towards one another. Beyond the patch the wings will converge, their
energy becoming concentrated into a narrow stretch of sea, leading to
higher, steeper and more confused waves. The Dogger Bank in the North
Sea has been blamed for focussing effects of this sort, with fishing vessels
being endangered by the high seas building up on the down-weather side
of the Bank. {Fig. 5)

Waves reaching a shelving beach normally dissipate all their energy in


turbulence and in stirring up the sand and shingle, and so are not
reflected. Waves reaching a vertical stone cliff or concrete sea-wall,
however, can be reflected with little loss of energy. The interaction between
the incoming and reflected wave trains can the produce interesting effects,
notably standing wave patterns where the water simply leaps up and
down, an effect sometimes referred to as clapotis. Such wave patterns can

20
be dangerous to small craft. Reflected wave patterns have been Implicated
in fishing vessel losses off the Norwegian coasts, where wave reflections
from cliffs are common.

WAVE FRONTS ADVANCING

CONCENTRATION OF
WAVE ENERGY
WAVE REFRACTION
OVER SHOALS

BEACH

' SHOALING TO ZERO DEPTH:


WAVES BECOME SHORTER,STEEPER
+ FINALLY BREAK

Figure 5 Effects of shoal water

If we follow a wave into extremely shallow water, we find that its


nature changes. The group velocity moves towards and merges with the
celerity; this means that in very shallow water an isolated wave front
becomes possible, and in fact single such waves do occur in nature and
are familiar as the bores which are triggered by tidal flows into estuaries.

Shallow water waves are known as waves of translation, since there is


a bodily movement of water forward inside the wave. This is evident where
waves wash up on to a shelving beach; each wave carries a volume of
water up on to the beach, whence it can be seen flowing back into the sea
in the interval before the arrival of the next wave, leaving items of flotsam

21
like plastic bottles at the top of the beach. By contrast, there is no bodily
transfer of water forwards in deep water waves, where individual water
particles simply move in circular orbits around unchanging mean centres.

Deep water waves are roughly symmetrical, with slopes of leading and
trailing faces seldom exceeding a steepness of 1 in 7. A wave travelling into
shallow water, however, loses its symmetry; the leading slope becomes
progressively steeper, eventually reaching and passing the vertical, and the
crest finally curls over forwards and breaks. The steep-fronted wave can be
particularly dangerous to small vessels, and there is little doubt that some
fishing vessels have been capsized by the impact of a vertical-fronted wave
striking flat upon the side of the superstructure.

Waves breaking on a beach.

The velocity of a wave of translation is independent of its wavelength -


if it is a solitary wave it has no length anyway - but it is proportional to the
square root of the water depth (it is in fact equal to the square root of
depth multiplied by "g", the acceleration due to gravity). A wave in 1 metre
of water moves at 6 knots, in 2 metres at 8 k n o t s , in 3 metres at 10
knots, in 10 metres at .l.Q1'* knots and so on.

22
In 1000 metres of water a wave of translation would travel at 193
knots. Such waves do actually occur, albeit infrequently. They are not
generated by the wind, but by earth tremors of volcanic origin. In the open
ocean they may only be a few centimetres high and they sweep past
unperceived by the mariner; but on approaching the coast they become
tall and steep, and can crash on the shore causing death and destruction.
The translation characteristic of the wave carries thousands of tons of
water beyond the normal high-tide mark; ships have been lifted and
deposited 200 metres inshore. Such waves are sometimes referred to as
tidal waves, misleadingly so as they have no connection with the tides. In
the*. Pacific basin they are known as Tsunamis. Earthquakes in South
America have been known to generate Tsunamis able to wreak damage on
the shores of Japan 10,000 miles away

Waves and Small Craft

The small craft sailor lives at close quarters with the waves of the ocean
and h%s every reason to respect them, especially the steep-fronted variety.
A wave which is no more than a minor irritant to a big-ship mariner may
be a major hazard to a yachtsman.

Given such a hazard, the mariner in a small powered vessel will


usually prefer to meet the seas head-on; he has more freeboard forward
and end-on waves cause less rolling. Sometimes, though - eg, when
running for shelter or entering harbour - he may be obliged to bring the
sea astern. Sailing vessels have less choice in the matter, and may be
driven to running before the wind. Since wind and waves generally move in
much the same direction, this usually means that the vessel is moving in
the same direction as the wave crests. This is the following-sea situation. It
has its own peculiar dangers, which arise from the hydrodynamics of wave
motion. -

It is quite obvious that the water in a wave is in motion, motion which


can sometimes be violent enough to endanger the vessel. But simply
watching a wave train gives little idea of the nature of this motion.
Experimental observation and theoretical analysis, however, show that
individual particles in a deep-water wave move in circular orbits about
mean centres which are fixed in space. At the crest of a wave the particles
are at the tops of their orbits and are moving forwards in the direction of
wave travel, while at the trough they are moving back in the opposite
direction. This alternating advance and retreat, superimposed on the rise
and fall, can be distinctly felt by a swimmer.

The interesting situation arises when the mean ship speed through
the water is close to the wave crest celerity. This will mean that the ship
"takes up station" upon a wave crest for some considerable period of time.
If the initial ship speed through the water is marginally less than that of

23
the wave, then as the crest passes the forward orbital motion of the water
particles in the crest will increase the ship's speed over the ground,
possibly up to the wave celerity, and the vessel will then be swept forward
by the crest, effectively becoming locked on to the wave.

The free gift of a little extra speed looks like a bonus, and so it might
be in the open ocean, but in coastal waters it may not be so welcome. As
the waters shoal the waves slow down and shorten, but at the same time
they get higher and steeper. This means that a craft which was initially
moving more slowly than the crests may find as the water shoals that the
waves cease to pass by from astern and begin to take station on her, with
danger of pooping or flooding from a wave breaking over the stern. A small
craft may at this stage become locked on to the wave and get carried
shorewards faster than she would wish, in the ultimate being pushed up
on to the beach by a steep wave in the manner of a surf rider.

To avoid this predicament the mariner can stream a drogue or sea-


anchor from a line over the stern; this is a drag-inducing device which acts
to keep the vessel square to the wave crests, and also enables the propeller
revolutions to be kept up, so keeping a good flow over the rudder to help
steering without at the same time increasing the speed of the craft through
the water.

There are further dangers from following seas. In vessels of certain


hull forms there can be a significant loss of transverse stability when there
is a wave crest around amidships and troughs at bow and stern. This
matters little if waves are passing quickly, but if ship and wave speed are
close the crest-amidships situation may persist long enough to allow an
unwelcome heel to develop. This phenomenon is discussed in more detail
in the chapter on stability below. I

The other main hazard, and one more familiar to small craft sailors, is
broaching. A craft is said to broach when she is brought broadside-on to
the waves in spite of all efforts by the helmsman to prevent this
happening. The broadside situation is one to be avoided, particularly in
steep-fronted waves, because of the consequent heavy rolling and danger
of capsize. (Fig. 6)
WAVE CELERITY
i 8 L 10 \

Broaching is particularly likely to happen in following seas, where


wave crests are passing - or being overtaken - relatively slowly. If the
vessel is not quite square to the wave crests, hydrodynamic forces and
moments arise which act to increase any obliquity, driving the vessel
further round until the seas are on the beam. The avoidance of this
condition depends on the ability of the rudder to generate enough force to
maintain control of the vessel's heading. In a sailing vessel this in turn
depends on the vessel's speed through the water. The dangerous moment
arises as the crest passes the stern; inside the wave crest the water
particles in their orbital motion have a velocity in the same direction as the
vessel, so reducing the flow over the rudder, possibly even to zero. The
rudder becomes ineffective and the vessel yaws round into the trough, to
take wind and wave full on the beam.
V
The power-driven vessel is in a somewhat better situation. With a
rudder immediately behind a turning propeller there is always some
rudder force available, though this is still reduced by the orbital action
and n|ay not be enough if power and/or rudder size are limited. If power is
available but propeller revolutions have to be kept low to keep ship speed
down for navigational reasons, a skilled coxswain may be able to stave off
an incipient broach by giving a short burst of engine power to restore the
desired heading. Alternatively, as mentioned above, speed can be cut by
streaming a drogue, allowing propeller revolutions to be kept high enough
to keep a good flow over the rudder.

Speed and Power


In designing a new ship, we need at an early stage an answer to two
questions: How many horse-power shall we need to drive the ship at the
specified speed? And how must we shape our underwater hull to reduce
this power to the minimum?

The effective power absorbed in driving a ship through the water is


equal to the speed multiplied by the total resistance. For a given speed, the
problem reduces to predicting the resistance or drag force for a range of
possible shapes, and then selecting the best shape - that is, the one which
has the least resistance while still satisfying other requirements such as
those for carrying capacity and stability.

This matter of resistance prediction and form optimisation is one of


great complexity. It has given rise to a voluminous literature, much of it
highly mathematical. We shall attempt here only a short outline of the
main fundamentals; readers who want to go into the matter in more detail
must refer to the standard works on naval architecture.

Consider first the nature of the resistance force. This is the force
exerted by a fluid upon a body moving through the fluid, acting in a
direction opposing the motion. To maintain the motion we shall need to

25
apply a propulsive force to the body which is equal and opposite to the
resistance force.

(Note that other forces may arise which are at right angles to the line
of advance of the body. These are known as lift forces. They are small in
the case of a conventional displacement vessel moving in a straight line at
a moderate speed and are usually neglected, but in high speed craft they
become large enough to lift the craft partially out of the water - the craft is
said to plane. We shall return to lift forces later; for the moment we are
concerned only with the resistance force acting in the line of motion).

Fluid forces enter the body in two distinct ways; by tangential drag
acting on the surface in a direction parallel to the surface, and by normal
pressures acting at right angles to the surface.

The tangential drag is in the nature of friction, and arises from the
viscosity of the fluid, as discussed in an earlier chapter. It depends upon
the area of the body surface exposed to the fluid - the "wetted surface" -
and the square of the velocity and a frictional coefficient/ The coefficient in
turn depends upon the viscosity and density of the fluid, the length of the
surface and its roughness. The effect of roughness is marked; the
coefficient has a minimum value for a perfectly smooth polished surface,
an ideal much sought after by racing yachtsmen.

Continued exposure to seawater leads to a gradual loss of


smoothness, and consequent increase of friction, due to the growth of
weed, barnacles and other living matter, and sometimes the spread of
pitting due to corrosion. Ship designers usually allow for the growth of
friction due to such fouling by assuming a daily increase in the skin
friction. The better the anti-fouling paint, the smaller will be the increase.
A common allowance is 0-25% per day out of dock - probably a pessimistic
figure for some of the best modern compositions.

Frictional coefficients can be derived from published formulae. The


wetted area of the hull can be measured from the ship's drawings. The
skin friction resistance can therefore be predicted approximately at an
early stage in the design, since it depends mainly on overall dimensipns,
and is not sensitive to small changes in shape.

Fluid forces arising from the distribution of normal pressures over the
hull caused by the motion are of several kinds. In a surface ship (as
distinct from a submerged submarine) the component of overwhelming
importance is the wave-making resistance. We have already seen that the
maintenance of a constant pattern of advancing wave crests, such as that
accompanying a ship, demands an expenditure of energy in the form of a
force moving in the direction of wave advance. This force, which needs to
be exerted by the ship's propeller or sails, is equal and opposite to the
wave-making resistance. Its magnitude depends veiy much on the hull

26
shape and increases sharply with speed. While the skin friction force on a
given hull goes up roughly as the square of the speed, the wave-making
resistance behaves in a more complicated manner.

As already explained, the wave-making drag depends on the relation


between the ship's length and the wavelength of the wave system of equal
speed. At very low speeds the waves are short and wave-making resistance
is small, much less than the skin friction. However, wave-making rises
rapidly, though not uniformly, with speed. When wavelength reaches
around twice ship length we have "hump" conditions, where wave drag is
going up not as speed squared but rather as speed raised to the fourth or
even fifth power. Above this speed the rate of increase falls away again,
hydrodynamic lift forces become significant and the craft enters the
planing regime.

It will be seen that the two major components of resistance vary with
speed in very different ways. At slow speeds there is little wave-making
and friction predominates; in a slow cargo ship it may represent 80% or
even 00% of the whole. At such speeds it pays the designer to minimise
the welted area. The ship's displacement or under-water volume is more or
less fixed by the specified cargo-carrying capacity; the designer therefore
seeks the form of least surface area for a given volume. The ultimate in
this direction, the sphere, is clearly not practical, but the designer will
approach it as closely as he can and will favour the short fat ship. In the
early days of sailing cargo ships length-to-beam ratios of 3 or so were
common, with near-circular sections.

In the faster ship, wave-making becomes more important and the


lines have to be fined down, particularly forward. Around the hump we
have the very slim destroyer forms with length-to-beam ratios of 9 or even
10. Once beyond the hump we begin to plane and can move again towards
shorter hulls. The lively motion experienced by planing craft in sea waves
encourages shortening as an aid to maintaining structural strength.

Wave-making is not the only kind of "pressure" resistance. Other


components exist, known variously as form drag, separation drag, eddy-
making resistance, appendage drag, air resistance (not negligible in a fast
vessel with large upperworks), augment of resistance due to propeller
interaction with the hull and so on. All are of great interest to the
hydrodynamicists, but in a well-designed hull these resistance
components normally bulk small beside the wave-making resistance and
are often lumped together with wave-making under the convenient over-all
title of "residuary resistance" - meaning, in effect, all resistance which is
not due to skin friction. (Fig. 7)

27
MODEL SHIP

MEASURED
RESISTANCE
OR MODEL

Figure 7 Prediction of ship resistance from model


experiments

The distinction between residuary and true wave-making resistance


must not be forgotten, however. In the particular case of the fast fully-
submerged submarine, where the surface wave-making is nil, the non-
wave component of pressure resistance can become important. And in
designing the stern lines of any vessel, fast or slow, the possibility of
separation oi flow has tu be bouic in mind. Separation is explained in
more detail in a later chapter; suffice it to say at this stage that separation
occurs if the flow of water approaching the stern breaks away from the
surface of the hull and throws off a trail of large eddies behind the ship,
causing a large increase in drag. This happens if the stern is too bluff; that
is, if the waterlines slope in too steeply towards the stern-post. (A rule of
thumb sometimes used is to limit the inward slope of waterlines to 20
degrees relative to the middle-line). Separation in a well-designed hull
should be negligible.

The separation phenomenon is independent of speed, so that the


slope limitation applies even in the slowest of vessels. Such vessels are not
great wave makers, and so can have quite bluff bows. We thus have forms
which are very full forward and relatively fine aft, leading to a centre of
buoyancy (ie, centre of gravity of the displaced water) which is well forward
of amidships. This feature can be seen in the forms of the sailing ships of
"wooden-walls" days, and in the bodies of fish. Similar considerations also
lead to the tear-drop shape of nuclear submarines; these are far from'slow,
but their deep submergence suppresses wave-making and makes a blunt
rounded bow acceptable, which is very convenient for fitting torpedo tubes
and sonars.

In fast surface ships we need to reduce the angle of entry at the bow
to minimise wave-making, while less change is required at the stern. The
centre of buoyancy thus moves aft, and will be found well abaft amidships

28
1 "-JLiSiJiLli - * i

in fast ships like destroyers. Merchant ship designers have various


empirical formulae based on experience telling them where best to locate
the centre of buoyancy relative to amidships for various speed-length
combinations.

Resistance Prediction
The prediction of residuary resistance in the design stage is a matter of
some complexity. Unlike skin friction, it cannot be derived from a simple
formula, and it is sensitive to hull shape. Up to the middle of the 19th
century various attempts had been made to predict resistance by
experiments with ship models, without much success. The break-through
came when W. Froude established the distinction between skin friction
and, residuary resistance. His resistance prediction routine, followed in
principle to the present day, was to tow a scale model of the ship in a tank
of water and to measure the total resistance. He then calculated the skin
friction force, using coefficients derived from experiments with flat planks
towed edge-on. The calculated skin friction force was then deducted from
the tl>tal model resistance; the remainder was the model residuary
resistance. This he scaled up to ship scale in the ratio of the displacement
weights of model and ship, giving ship residuary resistance. He finally
calculated the ship skin friction; this added to the residuary gave the ship
total resistance. The accuracy of the method was proved by actually towing
a full-sized ship at sea using a dynamometer in the tow-rope. The method
has been amply confirmed by experience over more than a century.

In this procedure it is of course essential to tow the model at the right


speed, a figure which is not at first sight obvious. Froude pointed out that
model conditions represented those at ship when not only the model shape
but also the wave pattern shape were geometrically similar to those on the
full scale. This required that the wavelength of the transverse wave pattern
be scaled down in the same ratio as the length of the model; thus, if the
model was built to 1/84 of the ship size, then the waves in the model tank
should have a wavelength equal to 1/64 of the waves set up by the ship.

We know that the celerity of the transverse wave crests is equal to the
speed of the ship, and that the wave celerity varies as the square root of
the wave length. The relation between model length and wave length in the
laboratory tank is to be the same as that between ship length and wave
length at sea. This is achieved if the model speed is scaled down in a ratio
equal to the square root of the linear scale. Thus, if we have a 1/64 scale
model of a 24 knot ship we must tow it at 24 knots divided by the square
root of 84 = 24/8 ~ 3 knots. This is called the corresponding speed/ When
a model is towed at corresponding speed, the ratio residuary
resistance/displacement weight is equal to the similar ratio in the ship.
Note that the ratio involves displacement weights, not volumes; allowance
must be made for water densities if, as is usually the case, the ship
operates in salt water while the model is tested in fresh.

29
The model procedure is quite expensive in that it involves substantial
capital facilities - long towing tanks, elaborate dynamometry and recording
instrumentation, model-making workshops and so on. Many attempts
have been made to cut out the model work by predicting wave-making
resistance through mathematical hydrodynamics, and computer programs
are becoming available for this. In addition much archival material is
available from tank models which have been tested in the past,
representing both actual ships and shapes of "methodical series" in which
various shape parameters have been varied in a systematic manner. For a
commercial ship of conventional design an acceptably accurate resistance
prediction is nowadays usually possible using available archive material
and programs, without commissioning new model work. Model work still
comes into its own, however, for novel types of ship, and for conventional
ships under unusual conditions - eg, in very shallow water, or with heavy
trim due to damage.

Over The Hrnnp

Few cargo ships operate "over the hump". Among the smaller craft,
however, speeds beyond the hump are increasingly common. Once the
hump is passed hydrodynamic lift forces become increasingly important.
As they grow the craft rises out of the sea so that the volume of water
displaced by the hull becomes significantly less than that displaced when
the craft is floating at rest. This is the planing regime. At very high speeds
the craft seems merely to skim the surface, only a small patch of the hull
bottom remaining in contact with the water. The reduction in wetted area
offers a beneficial decrease in friction.

Th^ can be thought of as a planing craft in which the


volumetric displacement is reduced to near-zero, while the lifting surface -
the foil itself - is fully immersed so that the positive lifting pressure on its
under-surface is reinforced by a negative or suction pressure on its upper
surface.

The hovercraft on cushion but not under way is a displacement


vessel, in that the air cushion forces a depression in the water, displacing
liquid of weight equal to the weight of the vessel - as required by
Archimedes. As the hovercraft gets under way and increases speed it
encounters the same hump as any displacement vessel; if it has the power
to go faster it enters the planing regime and the depression in the /water
gets smaller. The hovercraft enjoys the absence of friction - total absence
in the amphibious craft, partial in the sidewall type. It also experiences a
THtlicr sinootlidr.. : over small waves than most conventional planing
craft, due to the elasticity of the air cushion between the sea and solid
structure. In large waves the craft is forced down off the cushion and these
advantages are lost.

30
At very high speeds hydrodynamic lift forces become insignificant.

The hydrofoil craft at rest is supported by ordinary buoyancy forces, but


under way the displacement and hence the buoyancy shrinks virtually to zero
and the craft is supported by hydrodynamic forces on the foils.

31
It is sometimes said that the very fast craft, hydrofoils in particular,
do not generate significant waves. This is quite wrong. Any craft, whether
supported by a displacement hull, a skimming bottom, an air cushion or a
foil, exerts a downwards force equal to its own weight upon the water
beneath it. Once under way this force constitutes a travelling disturbance,
and it will generate a travelling wave pattern moving across the sea at its
own velocity. At a high speed this wave may be very long eg, a craft (of any
sort) moving at 40 knots in deep water is accompanied by a wave pattern
with 271 metres between crests. Such a wave may not be easy to see
because it will be quite low in relation to its length, but it may still contain
significant energy, and its maintenance will accordingly impose significant
wave-making drag on the craft generating it.

Waves generated by high-speed craft may have some practical


importance in harbours where larger vessels are secured alongside. A low
but long wave thrown off by a speeding craft may cause berthed ships to
heave and surge, straining at their mooring wires and possibly hazarding
cargo working operations.

Interesting situations can arise when fast craft operate in shallow


water. We have already seen that the nature of waves changes as we go
into shallow water, where the deep-water wave (celerity determined by
wavelength) is replaced by the wave of translation (celerity determined by
water depth). In water 10 metres deep, for instance, the wave of
translation has a celerity of 19^4 knots. It is not possible for any wave to
travel faster than this in 10 metres of water.

This has practical importance. Consider a fast craft travelling at, say,
20 knots in deep water, accompanied by the usual wave pattern. The craft
now enters shoaling water. As the wave pattern is modified by the shallows
the resistance will initially increase, and the craft will have to increase
power to maintain constant speed. When the depth falls to below 10
metres, however, the transverse part of the wave pattern can no logger
keep up and simply disappears, leaving only a modified divergent wave
pattern. The wave-making resistance is sharply cut, and the craft actually
travels faster for a given engine power than she would do in deep water.

It is said that the reversal of the effect of depth on resistance was


exploited by one of the early builders of destroyers who contrived to run
his measured mile speed trials over a course that was conveniently
shallow. Owners nowadays are alive to the depth effect, and require speed
trials to be conducted over a measured mile of acceptable depth; the
measured mile off the Isle of Arran has something like 200 metres of
depth.

Competitive rowing and sailing events are nowadays often conducted


on inland waters, lakes and rivers of uncertain depth. It seems likely that
at competition speeds depth effects could be quite significant in some of
these courses, and open to exploitation by competitors able and well-
informed enough to position their craft to take advantage of the most
favourable depths. Truly fair competition would seem to demand either
really deep water, or water dredged to uniform shallowness throughout the
racing area so as to avoid favouring or penalising individual competitors.
In comparing results obtained over different courses, allowance must be
made for the effects of depth.
s
>
In the early days of inland canal traffic, water depths were of the order
of a metre with a corresponding critical speed (ie, the celerity of a wave of
translation) of 6 knots. Canal barges carrying cargo operated at the speed
of a^single ambling nag and so were safely sub-critical. However, operators
of the occasional passenger barge discovered the critical speed
phenomenon at an early stage, and used teams of horses to drag their
light craft up to super-critical speeds, whereafter a single trotting horse
could maintain speed. Presumably the boost phase had to be repeated
after 'every lapse below critical for locks or passing traffic.

The generation of waves of translation in shallow water can have


another effect of practical impact. Consider again the fast craft steaming
along a deep river or canal at 20 knots. The waterway now shoals, or the
tide falls, so that the depth approaches 10 metres. Eventually the craft
speed coincides with the critical speed for the waterway. The transverse
wave pattern now takes the form of a single crest, travelling at the same
speed as the craft. As this is a translation wave and not a deep water
wave, the group velocity equals the celerity and so no energy is fed back
into a following wave train. As the craft travels, more and more energy is
fed into the wave, which builds up and up in height until it constitutes a
sort of artificial bore. Eventually something has to give; probably the boat
is unable to maintain speed or course and slows down. Meanwhile the
bore carries on up-stream, playing havoc among small craft moored along
the banks.

In practice the build-up conditions are unlikely to persist for long


because of variations in water depth, course changes and so on. But the
possibility still exists of dangerous and destructive waves being generated
where fast craft are allowed to operate at speed in confined and shallow
waters. This applies equally to planing craft, hovercraft and hydrofoils -
notwithstanding that the two last have sometimes been suggested for river
service on the grounds that "they don't make waves". They do.

Waves and Currents ;


We have seen that a surface wave has a celerity - that is, a wave-crest
velocity - that depends upon both wave-length and water depth. The depth
effect is small if the water is deeper than about half a wave-length.

33
A consequence of the length-dependence of wave celerity Is the so-
called dispersion effect, whereby long waves travel faster than short ones.
This means that the long waves generated by a remote storm disturbance
reach an observer before the short waves; a long smooth swell may be the
mariner's first warning of trouble afar off.

Dispersion results in "grouping". A long uniform train of waves is


unstable, and even if generated artificially in a laboratory tank such a
train will eventually break up into separate groups of waves. Uniform
trains never occur in nature.

We saw earlier that each group of waves advances together at a group


velocity which, in deep water, is equal to just half the celerity of individual
wave crests. The observer at sea can see each crest advancing through the
group from the rear, growing in height and then shrinking again as it
reaches the front. It follows from the observed behaviour, and can also be
demonstrated from the mathematics, that the mechanical energy
contained in the wave system advances across the ocean at the group
velocity and not at the speed of individual wave crests. This seemingly
academic fact has practical consequences which can directly affect the
mariner's safety.

Consider what happens when a set of waves advances into an area of


sea where there is a local current flowing in a direction opposite to the line
of advance of the waves. At first sight it might be thought that the only
effect of the current would be simply to reduce the wave speed over the
ground by one knot for each knot of current, but in fact things are more
complicated than that.

Since the frequency of the waves, ie the number of crests passing over
a fixed point on the sea bed per minute, must be the same outside and
inside the current area, it follows that the waves inside the current area
must be shorter. Because of the wave-length dependence discussed
earlier, it further follows that the wave celerity through the water, and
hence the group velocity and energy transport velocity relative to the
water, are all reduced.

Hie advance of the energy over the seabed is thus slowed down by two
separate and distinct actions; the dispersion effect slows down the flow
through the water, and the current itself slows down the movement over
the sea-bed. The effects are additive. The practical importance of this lies
in the fact that the slowing down of the energy flux means that the amount
of mechanical wave energy held in a given plan area of sea is increased,
and so the waves in the current-affected area get higher and steeper.

It can be shown from the theory that when the local current velocity
reaches the critical value of one-quarter of the open-ocean celerity of the
wave crests, the wave energy advance velocity over the ground is reduced

34
i

* J^' s i-I0 : f-

to zero. This results in the extraordinary situation where all the wave
energy arriving in the wave system is brought to a complete stop. Although
the waves continue to advance into the current area, their great energy
becomes trapped there, leading to a patch of violently confused steep seas.
Down-weather there is an area of calm water which the wave energy
cannot reach. (These effects form the basis of pneumatic breakwater
schemes, which use surface water currents induced by streams of air
bubbles to break up advancing wave fronts).

Can the energy blocking property of currents affect ocean-going


ships? Consider a wave long enough to inconvenience such a ship, say one
with; a wave-length around 100 metres. This wave will have a celerity of 24
knots, so that a local current of one-quarter of this, 8 knots, will be
enough to stop the enormous energy of the wave system dead in its tracks.

This cargo liner was lucky to survive spectacular damage to her fore end when
she "fell into a hole in the water" and shipped a huge breaking wave off the
coast of South Africa in the region of the Agulhas current.

35
There are not many places in the world where currents as fast as this
can meet long ocean swells head-on, but an area in the Indian Ocean off
the coast of South Africa is one of them. Here the Agulhas Current is
reputed to flow southwards at speeds of up to S1^ knots on occasion,
sometimes meeting long swells running northwards from remote Antarctic
storms. A number of ships have been sunk and damaged by heavy seas in
this area, and in the '70s a large cargo liner was lucky to survive
spectacular damage to her fore end when she "fell into a hole in the water"
and shipped a huge breaking wave when steaming southwards along the
line of the current. So the energy-blocking effect can indeed endanger
ocean-going ships.

There is a risk that the mariner can bring this hazard upon himself
unwittingly. If he meets a current flowing strongly in the direction in which
he wants to go, the mariner will naturally be tempted to route his ship
through the middle of it so as to gain a free boost in speed over the ground
and hence to save on his fuel bill. This is a reasonable tactic - provided
there are no long swells coming from ahead. If there are, the mariner may
well be taking his ship into dangerous waters. If he is in any doubt, and
has the sea-room, he would be better advised to keep clear of the current.

If we consider the smaller vessels, there are of course many more sea
areas where current effects can lead to danger. In coastal waters tidal
currents ana "races" can commonly reach several knots, leading to high
steep seas where the current flows against the waves - the "wind against
tide" situation. The small-craft mariner should be as wary of tidal currents
as of shoal water. Current and shoals together make a particularly potent
combination.

36
CHAPTER THREE

PROXIMITY

Bernoulli

We have already seen that the speed of a ship can be affected by the depth
of the water under the keel. Speed is not the only feature to be affected by
the proximity of the bottom; draught, trim and steering behaviour all
depart from the normal when the vessel is under way in close proximity to
the seabed, to horizontal boundaries such as dock walls or channel banks,
or tt> other vessels. These aberrations are referred to collectively as
proximity effects. They are of more than academic importance, as they can
- and do - cause collisions and groundings.

Proximity effects spring largely from the Bernoulli phenomenon, well


known to engineers. The 18th century Swiss philosopher Daniel Bernoulli
established the relationship between the pressure and the velocity in a
stream^ of moving fluid, and showed that the pressure in a horizontal
stream of water decreases as the velocity increases, and vice versa. One
can think of water flowing through a pipe; if the pipe is locally squeezed in
so as to produce a wasp waist, the water in the constriction will be moving
faster than that in the unsqueezed pipe, and the pressure in the
constriction will be less than that in the pipe clear of the constriction. (The
layman might expect the opposite, reasoning that a constriction could be
expected to put the pressure up. This shows that common-sense is a poor
guide in matters of hydrodynamics) .

Bernoulli's equations show that pressure changes due to fluid


movement depend upon the square of the velocity. This means that
Bernoulli pressure forces grow rapidly as speed increases; and equally that
they shrink quickly as speed drops. We shall see that the mariner has
good reason to note this point.

Squat ^

Consider a displacement vessel in calm water. While at rest the ship is


entirely supported by hydrostatic pressures acting on the bottom plating.
Once under way, however, the pressure distribution is altered because of
the velocity of the water relative to the ship.

Say the ship is advancing at constant speed V. The physical picture is


unchanged if we regard the ship as being at rest while the water moves
past it. Remote from the ship the water is moving at speed V, but because
of the disturbance of flow caused by the presence of the ship's hull t i e
water close to the plating will be moving at a range of velocities differing
from V. As a consequence of the Bernoulli effect, the pressure distribution

37
on the plating now differs somewhat from that existing with the ship at
rest, The total upward force on the hull changes; at moderate speeds there
is usually a small net decrease, so that the hull sinks slightly in the water
until the increase in buoyancy restores equilibrium. At very high speeds,
as in racing small craft, the hull may rise. In general the change in mean
draught is accompanied by a change in trim.

The changes in draught and trim are together referred to as squat. It


should be noted that squat occurs in deep water as well as in shallow.
Squat is not usually of much practical importance in deep water, but in
shallow water it becomes larger and potentially dangerous. (Fig. 8)

CROWDING OF FLOW LINES UNDER BOTTOM LEADS TO


INCREASED WATER VELOCITIES WHICH LEADS TO
REDUCED UPWARD PRESSURES AND THUS INCREASED
DRAUGHT

Figure 8 Squat in shallow water

The reason for the increase in shallow water can be visualised; as the
hull approaches the seabed, the gap through which the water must pass
narrows, the velocity goes up, the pressure on the bottom plating goes
down and the squat increases. The effect is sometimes described as
bottom suction.

The magnitude of the squat sinkage depends upon the undef-keel


clearance and square of the speed. It can be quite significant. A^ship
drawing, say, ten metres of water can safely lie at rest or move slowly in 11
metres of water, but if she is steaming at speed she may well sjjuat
through the odd metre and strike bottom. It does happen from time to
time.

A vessel initially on an even keel will usually trim by the bow' when
moving in shallow water; that is, the bow will sink more than the Stern. A
vessel initially trimming heavily by the stern will trim even more by the
stern in shallow water. A vessel having some intermediate value of initial
stern trim will sink uniformly without change of trim; the value of the
initial^ trim to achieve parallel sinkage will depend upon the hull shape. In
practice most merchant ships in the usual loaded condition will squat and
trim by the bow when under way in shallow water.

38
Figures have been published from which parallel sinkage and trim for
various hull forms, speeds and depths can be estimated. Reliance should
only be placed on those formulae or graphs which take explicit account of
initial under-keel clearance. Whether he has such data to hand or not, the
wise mariner will remember that ships squat in shallow water, that the
amount of squat depends upon speed squared, and that if the echo
sounder shows only a small clearance under the keel it will be as well to
keep th speed down, especially over a rocky bottom.

Stability

Apart from causing squat, under-bottom suction can influence transverse


stability and even induce an unexpected heel. Consider a wide flat-
bottomed craft under way with small under-keel clearance. She now heels
transiently to starboard due to, say, a gust of wind striking the
upperworks. The starboard bilge sinks even closer to the sea-bed and the
port feilge rises. The suction under the starboard side is consequently
increased and that on the port side diminished. The effect upon the ship is
of a helling moment to starboard, acting to increase the initial heel. If the
ship ha^s adequate metacentric height her roll stiffness will still bring her
back to the upright, but if she is "tender" the righting moment may be
inadequate and the suction moment may take charge, increasing the heel
until the starboard bilge touches bottom

If the bottom is smooth and soft this need not be fatal, in fact the ship
may continue moving ahead with the starboard bilge slithering along the
mud until deeper water is reached. A case is on record of a cargo vessel
operating in the shallow parts of the Gulf of Mexico which developed an
unexplained steady heel. The heel vanished when the ship slowed down,
returned again with increasing speed and finally disappeared when she
reached deeper water.

In shoal water the mariner should remember that all Bernoulli effects
vanish as velocity tends to zero ^ ^An unexpected heel or soggy slow roll
motion should be taken as a warning to reduce speed.

Bank Effects

Bernoulli forces arising from the ship's motion ahead through the water do
not only act vertically. They can be significant, sometimes dangerous and
sometimes helpful, in the horizontal plane too, notably in the group of
phenomena known as bank effects.

Oonsidcr now a vessel which is moving in a waterway which is deep


but confined on one side, say to starboard, by a rigid vertical wall or bank.
If the vessel approaches the bank the flow of water between ship and bank
becomes confined, and so speeded up in comparison with that on the open
port side. Because of Bernoulli the pressure force on the starboard side

39
falls below that on the port and the ship experiences a net force urging her
to starboard - towards the bank. On the face of it this looks a dangerous
situation, and so it can be. But it is not a simple one.

Complications arise because, as with squat, we have not only a bodily


force acting, but also a moment tending to turn the ship about a vertical
axis. This happens because the suction forces are not uniformly
distributed along the ship's length. Their resultant acts somewhat abaft
amidships. This is equivalent to a force acting at the centre of gravity plus
a turning moment. The moment acts in the direction to deflect the ship's
head away from the bank, to port in the present case. In practice this
turning moment over-rides the bodily suction towards the bank, and
swings the ship's head to port so that she starts moving out to sea away
from the bank. A ship being deflected away from a solid boundary in this
way is sometimes said to be "pushed off by the bow cushion".

The effect can be strikingly demonstrated with a model in a tank. A


model directed towards the tank wall at a small oblique angle will sheer off
and head away from the wall without making contact, as though deflected
by an invisible fender, the helm being locked amidships throughout. One
cannot know for certain, but it seems likely that this cushion phenomenon
may have saved many a ship handler over the years from making
expensive contacts with dock walls.

A problem arises when we wish to steam on a fixed course parallel


and close to the bank. It is necessary to overcome the outward-turning
moment, and so the rudder must be set so as to direct the bows towards
the bank, to starboard in the present case. The rudder so set exerts a force
on the ship's stern to port; this can be thought of as countering the net
force to starboard exerted on the main hull by the Bernoulli effects .v

We have an interesting situation when a ship steams down the middle


of a narrow canal, where there are similar banks to port and starboard.
The bank suctions and moments are equal and opposite and so cancel out.
If the ship strays towards one side, the cushion effect on that side
increases and returns her to the middle line without helm orders from the
mariner. In favourable conditions, curves in the canal can be negotiated
simply by leaving the helm amidships. This is one of the few instances
where hydrodynamics is helpful to the mariner. Needless to say, he should
not relax even if he is not giving many helm orders, because irregularities
in the banks may call for rapid intervention by him. y^

Consider now what happens when there is a gap in the right bank - a
dock entrance, say, or a river confluence. A ship steaming parallel with the
bank suddenly finds that the Bernoulli moment to port disappears. The
ship at once begins to swing to starboard. If the mariner is not prompt
with his helm, his vessel will come so far to starboard that the bow will
strike the right bank where it reappears after the gap. (Fig. 9)

40
TURNING

(a) BANK EFFECTS


STARBOARD HELM IS NEEDED TO MAINTAIN COURSE
SHOWN

Figure 9 Bank effects

Bank effects are present even when the banks are submerged and so
not visible. Hie banks may perhaps form the boundaries of a dredged
channel in the middle of a wide expanse of shallow water. A residue of
bank effect still persists when the bank is so deeply submerged that the
ship could steam safely over it without touching. The bank need not be
continuous. A potentially dangerous situation can arise when a ship
approaches a patch of shallows obliquely. Even though the shoal may be
below keel depth, a cushion effect can arise which is big enough to swing
the ship's head through an angle demanding correction by helm action.

A particularly unpleasant collision occurred in the River Plate estuary


in the 1960s, when a refrigerated cargo vessel approaching a patch of
shallows experienced a "sheer", or unexpected change of course, which
carried her into the path of an oncoming tanker. On that occasion
correcting helm was applied, but rendered ineffective by the engine being
put astern. An experienced mariner will be alive to the risk of sheers
caused by bottom irregularities in shoal water and will be prompt with
correcting helm - keeping his propeller turning ahead to give the rudder
"bite".

41
Ship-Ship Interaction

The solid boundary causing Bernoulli effects need not be a dock wall or
bank; it can be another ship. Certainly, two ships in proximity suffer
suction effects which act to draw the two hulls closer together. But, as
with bank effects, the bodily force is accompanied by a turning moment.
This moment varies in a complicated way which depends on the relative
positions of the two ships. The moment does not always act in the 'safe'
direction as in a canal.

Ship proximity situations take two forms, encountering and


overtaking. In the case of two ships on opposite courses encountering bow-
to-bow the proximity situation lasts only a short period of time; the
accelerations induced by the suction forces and moments do not last long
enough to result in significant velocities being built up before the ships are
past and clear of one another. In fact the bow cushion effect, which both
ships experience as the bows approach one another, acts to push both
vessels on diverging courses and so again makes for safety.

The potentially dangerous situations are those that can arise when
one ship overtakes another on a parallel course, or takes up station in line
abreast. Here the forces can last for many minutes and have significant
effects. If the vessels get too close together the interaction forces can
become too large to be countered by the rudder, and the hulls will be
drawn into contact.

This situation sometimes arises with tugs attending large vessels. A


tug steaming close to the fore end of a large ship underway, perhaps
preparing to pass a wire, will sometimes experience a turning moment
which in these circumstances will be directed towards the biggerlvessel.
This moment may be large enough to overwhelm the tug's opposing helm,
and the tug will be drawn into the larger vessel's bow and rolled over. (Fig.
10}

TURNING MOMENT M SUFFERED BY TUG NEAR BOW MAY


EXCEED THE CORRECTING MOMENT AVAILABLE FROM
TUG'SRUDDER

Figure 10 Ship-Hig Interaction

42
A tug captain experiencing this form of interaction should not try to
outpace or turn away from the larger ship, but should reduce propeller
revolutions so as to drop back to the quieter waters abreast the middle of
the ship. Once there he can use his helm to put some distance between
the two hulls. The larger ship should in any case reduce speed when a tug
is making fast or operating near the ship's bow.

As; with squat and bank effects, the forces and moments acting
between ships in proximity have been well studied by the
hydrodynamicists and there is information in the literature from which
they can be predicted. But for practical purposes, the best working rule
when overtaking is to leave the widest possible clearance between hulls,
and to be ready to use helm to correct sudden unexpected sheers,

Naval replenishment at sea operations. Interaction forces can be very


significant and call for skilled ship handling.

If prolonged steaming in company on parallel courses is unavoidable


for tactical reasons, as in Naval replenishment-at-sea operations, the
wisest approach is to take up station on the beam, amidships abreast

43
amidships but a long way off, and then to close the gap slowly with
minimal change of heading until the required spacing is reached. The
main requirement, both in taking up and leaving station, is to avoid the
partially-overlapping situation, bows of one ship near stern of the other,
when steering moments are at their worst.

Shoals and Steering


Returning now to the effects of shallow water, we consider the influence of
under-keel clearance upon steering behaviour. Shoal water changes the
flow of water around the hull and rudder in such a way that the steering
behaviour is modified. There is a marked effect upon tactical diameter.

When a ship's rudder is put over, two things happen; she starts to
move on a curved course, and her speed drops. After a while at constant
helm angle the speed reaches a steady value and the ship settles down on
to a circular course of constant diameter, referred to as the tactical
diameter. The tactical diameter for conventional cargo ships commonly has
a magnitude of about three ship lengths - in deep water.

An immediate effect of shoal water is to increase the tactical diameter.


The effect is marked; the diameter can double to, say, six ship lengths
while the clearance under the keel is still of the order of half the ship's
draught. At the same; time the fall in speed is reduced. Figures for a large
tanker show that in deep water under full helm the speed falls to half the
full speed for zero he draught the
speed only falls to 0*85 iftat fbr zero helm.

The combined effect of the larger diameter combined with the higher
speed under helm in shallow water means that the time rate of change of
the heading angle is not very different from that in deep water. This has its
dangers, because the mariner pri fe or hear the gyro
compass moving round at its usual rate for a given helm angle, without
realising that his ship is using much more sea room for her turn than she
would in deep water. (Fig. 11) i

Directional stability is affected too. This is a property of a ^hip


discussed in more detail in the chapter on Steering; it must suffice to say
here that a ship which is directionally stable will continue on a straight
and steady course in calm water as long as the rudder is held amidships.
A ship which is directionally unstable under the same circumstances will
steam in a large circle to port or starboard, either heading being -equally
likely. To maintain a mean straight course, the helmsman (or autopilot) in
a directionally unstable ship has to make a continual series of rudder
movements to check swings away from the mean heading, much as one
might ride a bicycle. Most, but by no means all, ships in service are
directionally stable.

44

i
8 1 3 . -

INITIAL SPEED=V

Figure 11 Steering showing the effect of shoal water

Hie initial effect of shoal water upon directional stability is usually to


decrease it. The effect varies somewhat with hull form, but it is commonly
found that a ship which is marginally stable in deep water may find herself
directionally unstable in shoals, and so may begin to pay off to port or
starboard without warning while the helm is still amidships. An
experienced helmsman may "sense the bottom" without looking at the
echo sounder when he notices the heading beginning to wander. If shallow
banks are encountered obliquely, as discussed in an earlier section, the
sheer arising from Bernoulli effects may be amplified by the concurrent
loss in directional stability.

As water depth decreases further we reach the situation where the


keel is very close to the sea-bed, perhaps within a metre or less. Now the
stability situation is completely reversed. Directional stability moves from
small and negative to large and positive; so large, in fact, that the greatest
possible rudder angle produces only a small rate of change of heading.
Helmsmen have described a ship so placed as being "stuck on tram-lines"
- almost locked on to a straight course. One has to hope that a ship so
situated does not have to manoeuvre sharply to avoid collision.

Virtual Mass

Virtual mass is a hydrodynamic property of a solid body moving in a fluid.


The concept appears at first sight to be somewhat academic, but we shall

45
see that it has some practical relevance. It is introduced in the present
chapter because it is sensitive to proximity effects.

Consider the energy of a body of mass M moving with velocity V.


Elementary physics shows us that this body possesses kinetic energy - ie,
energy by virtue of its motion - equal to ^MV2. This is the same as the
mechanical work done in accelerating the body from rest to velocity V.

If now we have a ship of mass M moving at speed V, the energy of the


ship itself is ^MV2. However, the total work done by the accelerating agent
in bringing the ship up to speed is more than this, because it has given
motion not only to the ship itself but also to quite a lot of water which is
moving in various ways in consequence of the ship's motion. Work has
been done in getting all this water moving too, so that the total energy of
the whole ship-and-water system is considerably greater than l/aMV2. It is
convenient to put it as equal to the energy of a hypothetical body of mass
Mi, namely ^MiV2. The practical effect is that, from the point of view of
anyone wanting to accelerate or decelerate it, the mass of the ship has
been increased from M to Mi. Mi is referred to as the virtual mass of the
ship. The difference Mi - M is the added mass.

For ordinary motion ahead in deep water Mi is of the order of 1-1


times M, the actual ratio depending on the hull form. In shallow water the
figure is increased very considerably, because of the change in the flow
pattern round the hull, The practical consequence of this is that a given
force achieves less acceleration or deceleration in shallow water than in
deep. Ringing down for full astern power may take off speed at so many
knots per minute in deep water, but the rate of fall-off will be less than
this in shallow, The ship may actually travel further over the ground
before coming to a stop, in spite of the fact that the drag force is greater in
shallow water. \

Added mass is much greater for sideways motion than for motion
ahead, and this again is also increased by shoal water. Admittedly, ships
seldom move sideways; but it does happen during berthing, when the, ship
must be decelerated to a stop before the jetty carries away. The combined
effect of sideways motion and shallow water can be such as to increase the
virtual mass, not by just a few percent but to many times the mass of the
ship alone. This means a great deal to the ship designer who has to
dimension the side framing to take the berthing loads, and to the civil
engineer who has to build his jetty to take the same forces when they come
out of the shore side of the fenders. The mariner bringing a loaded tanker
alongside at a dredged oil berth to the sound of scrunching fenders may
like to remember that the momentum of perhaps a million tonnes of water
needs to be checked, as well as that of ship and cargo.

46
CHAPTER FOUR

FOILS, FINS AND BLADES

Lift and Brag

Water flowing past a solid body exerts a force upon it. This force can be
resolved into two components, known as lift and drag. Drag acts in the
direction of the water flow, and lift at right angles to this direction.
Identical forces operate if the water Is at rest and the body moves through
it with the same relative velocity.

If the body Is symmetrical about the axis of flow - as, for instance, in
the qase of a running torpedo - then lift forces are absent and the body
suffers only a drag force.
7

111 the case of a surface ship, we have a body which - provided helm is
amidships - is symmetrical about a vertical plane, but is non-symmetrical
about lany horizontal plane. A lift force, which may act upwards or
downwards, will therefore in general act in the vertical direction. In
conventional displacement ships this causes the draught under way to
differ from that at rest, but the effect is small and can generally be
neglected in deep water. In very fast vessels, however, the lift force can
raise and trim the hull by a significant amount, and we ultimately reach
the racing motor boat which virtually skims the surface, supported largely
by the lift force acting on a small patch of bottom. In this case the
hydrostatic buoyancy force due to the submerged hull is small in
comparison with the lift.

In both slow and fast vessels, of course, drag forces (as distinct from
lift) will arise; these will be of the frictional and other types already
discussed in an earlier chapter. In the present discussion we shall
concentrate upon lift, as distinct from drag.

Return now to the fully submerged body. Let this initially be a blade
with a section in the form of a long thin oval. If the oval is symmetrical
about its longer axis and the water flow is along the direction of this axis,
then the flow pattern will be symmetrical about the axis and no lift forces
will arise.

We now introduce asymmetry. This can be done either by making the


body asymmetrical in section, or by tilting it relative to the flow, or both. In
each case the flow pattern will be rendered asymmetrical and lift forces
will arise. (Fig. 12)

47
(a) SYMMETRY DRAG ONLY

(b) ASYMMETRY BY SHAPE DRAG AND LIFT

J i LIFT

( c ) ASYMMETRY BY TILT DRAG AND LIFT

Figure 12 Forces on a submerged body in flow

We can form a physical explanation of the way in which these forces


arise by referring back to the principle of Bernoulli, discussed in an Earlier
chapter. If we take our oval blade and distort it by flattening one siirface
and humping up the other, so that it looks in section like a plano-convex
lens, we shall cause a flow pattern in which the stream lines are crowded
more closely together over the humped side than they are along the. flat
side. This means that the velocity over the hump is higher than that over
the flat. Alternatively, one can reason that the water path over the hump is
longer than that over the flat, so that the particles have to move faster to
get from leading to trailing edge over the hump than along the flat. /

We have already seen from Bernoulli that speeding-up is accompanied


by a fall in pressure proportional to the change in the square of the
velocity, The velocity increase over the hump means that the pressure here
is less than that on the flat. In effect, the hump suffers a suction, and the
body experiences a lift force in consequence.

48
We now return to the symmetrical oval blade, and introduce
asymmetry by tilting it through a small angle, such as 5 degrees, relative
to the flow. The flow pattern again becomes asymmetrical, with higher
velocities over the back of the blade (ie, the surface facing away from the
oncoming flow). It is not so easy in this case to form a simple physical
picture to explain why the velocities over the back should be higher; for
our present purpose it must suffice to state that there is ample
experimental evidence for this being so. The key lies in the need for the
flow to' separate smoothly from the blade at its sharp trailing edge, and not
at some point forward of this; while the oncoming flow can meet and split
into two at a "stagnation point" which can be somewhat away from the
precise leading edge, on the side facing the oncoming flow. Hie net result
is that the travel over the back is longer, leading to higher velocities and
suction on the back as before. There is again a force on the blade, which
canrbe resolved into a lift and a drag.

|n practice, the lift and the accompanying drag increase roughly in


proportion to the angle between inflow and bla.de plane of symmetry, up to
some limiting angle. Above this angle -the flow begins to break down and
become confused, no longer separating cleanly at the trailing edge.
Instead, turbulent eddies are thrown off and travel downstream. The lift
falls to zero while the drag increases sharply. The blade is said to have
stalled.

Lifting surfaces in water: Cavitation

In marine craft we use under-water surfaces to generate lift forces for a


variety of purposes; examples are hydrofoils, stabiliser fins, propeller
blades and rudders. We consider first the case of a lifting blade operating
in water without reference to its application. The first thing to note is an
important difference between a blade operating in water and a similar
surface working in air. This is the limitation imposed upon the water blade
by the phenomenon of cavitation.

We have already touched upon cavitation in discussing surface


tension in an earlier chapter. It was pointed out that water boils at a
temperature which depends upon the pressure, and that by reducing the
pressure to a low enough figure water can be made to boil at room
temperature. The pressure on the suction side of a surface generating lift
will be below that of the undisturbed water in the vicinity. If we increase
the lift force, by increasing the velocity or the angle of incidence or both,
the suction wiU increase, bringing the pressure down to atmospheric and
lower, eventually reaching a value at which the water begins to boil,
forming bubbles of vapour which travel downstream with the water flow.

As soon as the bubbles are clear of the low pressure area on the back
of the foil they are exposed to the full pressure of the surrounding
seawater and they collapse, each with an audible and potentially damaging

49
bang. If the collapse takes place in contact with a solid surface, the
repeated pressure pulses from bubble collapse can cause pitting and
erosion of the surface.

Photo courtesy Admiralty Research Establishment


A model propeller under test in the laboratory. This propeller is over-loaded, so
that cavitation occurs in the core of the tip vortex. The cavity is swept away
downstream. In a ship it would cause noise and possibly erosion of the \
rudder. (The downstream position of the propeller shaft is dictated by the \
laboratory equipment). \

Cavitation is in general an indication that the designer is asking the


surface to provide too much force per unit area. The lifting force possible
depends upon a range of factors, but a very rough rule of thftmb
sometimes used in preliminary design work suggests that each square
metre of lifting surface cannot be expected to give much more than 7
tonnes force of lift without cavitating. The precise figure will depend^n the
section shape of the foil and the depth of immersion. 5

Cavitation is thoroughly undesirable on three distinct counts. First,


the loss in efficiency; a cavitating surface loses lift without a compensating
loss in drag.Secondly, the noise generated in the water is a gift for any
hostile submarine or acoustic torpedo; it may also be audible inboard and
an annoyance to crew and passengers. Thirdly, the erosion caused by

50
bubble collapse can seriously damage propeller blades, shaft bracket
arms, rudder leading edges, etc. In an extreme case cavitation can erode
holes right through propeller blades.

Hydrofoils

We now come to practical applications of lifting surfaces. The hydrofoil


craft under way at operating speed is fully supported by the lift generated
by its foil members. The lift force always needs to be generated in one
direction, namely upwards, and so a non-symmetrical foil section can be
used as this minimises drag. The section shape used resembles the
aerofoil shapes seen in aircraft wings, but with an important difference. A
typical aerofoil shape has a large suction peak a short way aft from the
leading edge; this is acceptable in an aeroplane, but could cause local
cavitation if the same shape were used in water. For a hydrofoil, therefore,
the Shape is subtly adjusted to give a more uniform distribution of suction
over |he top surface, with no marked local peaks.

Mother possible consequence of the suction on the foil top is


ventilation; that is, the penetration of atmospheric air from above the sea
down into the low-pressure area atop the foil. If this happens, the flow
pattern will be disrupted and lift force will be lost. The craft may well come
"off the foils", dumping the main hull back in the water with immediate
loss in speed.

Air is particularly liable to work its way down from the sea surface
along the trailing edges of the legs which attach the foil surfaces to the
hull. To stop this happening the legs are sometimes fitted with fences,
local horizontal fins which act to detach any air bubbles which find their
way down from the surface, and throw them downstream before they reach
the main lifting foil.

T^ and ventilation are sometimes confused


with one another. They are in fact quite distinct in both cause and effect,
the only features in common being that they often occur in the same parts
of the ship, and they are both deleterious.

Fins,:.
Another lifting surface found in some displacement vessels is the stabiliser
fin. The operation of fins is discussed in a later chapter; suffice it here to
say that the fin in this application needs to be able to exert force equally in
either direction, up or down. It is therefore made with a symmetrical
section, and so has not quite such a good lift/drag ratio as that attainable
in a quasi-aerofoil section.

Like any other lifting surface in water, a stabiliser fin can suffer from
cavitation and ventilation. Cavitation is avoided by using an adequate area

51
of surface and a suitable section shape. Ventilation is seldom a problem in
conventional displacement ships where the fins are well immersed at turn
of bilge; in any case the intermittent operation of the fins as the ship rolls
gives little time for ventilation bubbles to build up.

Rudders
Practically every ship has a rudder, a lifting surface vital to the operation
of the vessel. Like the fin, the rudder has to have a symmetrical section
because it must operate equally to port and starboard. A rudder can in
theory cavitate, though this seldom causes trouble because a rudder
spends only a very short period of its life generating the full lift of which it
is capable.

Rudders more commonly suffer ventilation, with air travelling down


the rudder post or stock from the surface. This is particularly liable to
happen in small fast craft, where spoiler fins are sometimes fitted to keep
air from reaching the rudder and propeller region.

It will be recalled that lift force increases with angle of incidence up to


the stalling angle, where the flow pattern breaks down and lift falls away.
Hydrofoil surfaces and stabiliser fins are usually operated safely below the
stalling angle. Most designs of rudder will reach maximum lift at an
incidence of around 35 degrees, stalling thereafter, and steering gears
usually limit helm deflection to this angle.

It is, however, possible to modify the conventional rudder to increase


the stall angle considerably. This can be done by fitting a flap to the
trailing edge, much in the manner of an aileron on an aircraft wing. The
flap is caused to tilt in relation to the main rudder blade by a mecl|anical
linkage. By suitable shaping and angling of the flap, the rudder d^n be
persuaded to generate useful lift up to angles approaching 90 degrees. A
similar result can be obtained by fitting a vertical cylinder to the leading
edge of the rudder; when the rudder is deflected the cylinder is spun at
speed. This energises the flow over the rudder blade in such a way^that
stall is inhibited. (Fig. 13)

When one of these late-stall rudders is combined with a steering gear


providing helm angles up to 90 degrees, a remarkable improvement in
manoeuvrability is possible. The lift at high helm angles is accompanied by
a high drag; in a single-shaft ship this largely cancels out the prdpeller
thrust ahead, so that the ship gathers little or no headway while the stern
swings sideways. The effect is like that of a lateral thrust unit fitted at the
stem, with the important difference that the full power of main engines is
used to generate thrust, compared with the very limited output of the
usual thruster motor. A ship equipped with a high-angle rudder has a
tactical diameter at maximum helm angle of virtually zero.

52
! . BlSu-, . Z

LIFT

(a) CONVENTIONAL
RUDDER AT SMALL
^ H E L M ANGLE

Propellers

Each blade of a screw propeller is a lifting surface. It spends nearly all of


its working life exerting force in the same direction - ahead - and therefore
a non-symmetrical section is used. For many years blade sections were
commorily "ogival^, a shape bounded by a straight line on the face or
pressure side and a circular arc on the back or suction side. (Note that the
face of a propeller blade is turned towards the ship's stem and the back
towards M bow), More recent practice has been to use quasi-aerofoil
section shapes, modifiedas with hydrofoils to give a more or less uniform
distribution of suction on the back so as to defer the onset of cavitation.

53
Propellers are occasionally called upon to exert thrust astern. When
so working the lifting surface is operating "upside down", with pressure on
the more convex surface and suction on the flatter. Under these conditions
the lift/drag ratio is poorer than when thrusting ahead, and there will be
less thrust for a given power. Where a screw propeller is called upon to
thrust equally in either direction, as with a tunnel-mounted propeller in a
lateral thrust unit, a symmetrical blade section would be used to equalise
performance to port and starboard.

The actual geometric shape of a screw propeller is one of some


complexity. It approximates to a helicoid. A helicoidal surface is defined as
that swept out by a straight line which is square to an axis fixed in space,
and which rotates at constant angular velocity about that axis while at the
same time advancing along the axis at constant linear velocity. The two
velocities are so related that the line completes one complete revolution
while advancing through a distance P along the axis. P is referred to as the
pitch of the helicoid.

Any one blade of a conventional propeller will lie very close to a


helicoidal surface, and the pitch of the surface will be the geometrical
pitch of the propeller. The early ogival propellers were made so that the
blade faces lay exactly on the helicoid, and the pitch was determined with
precision by the geometry of the helicoid. Such a propeller could advance
through a solid body in which a matching helicoidal cavity had been cut,
in the same way as a bolt moving into a nut. This propeller would
necessarily advance through the solid by a distance equal to its geometric
pitch at each revolution. If a propeller of pitch P working in water advances
by a distance P for each revolution, it is said to be operating at zero slip,
like a screw in a nut. In practice, a ship's propeller advances by a distance
Q per revolution, where Q is somewhat less than P. Slip is said tq. occur.
Hie slip ratio, defined by (P-Q)/P, is sometimes computed by ship's officers
as a check on propulsive performance. An increase in slip indicates an
increase in hull drag, due perhaps to hull fouling. ^
i
Later propellers depart from the helicoid, as will be seen, so that there
may be a degree of uncertainty as to the precise meaning when values are
quoted for pitch. A concept used by naval architects is the 'analysis pitch';
this is the distance advanced per revolution by a propeller turning in water
while generating zero thrust. Care is needed too in discussing section
shapes. Although it has been stated that the face of an ogival propeller
blade is flat, it will be found that a straight-edge will not lie flat across the
face of such a blade. This arises because of the way in which sectiofi shape
is defined. If the blade is considered to be cut by a cylindrical surface co-
axial with the shaft, then the shape formed by the intersection of the
cylinder with the blade Itself will be the blade section shape. This lies on a
cylindrical surface in space, but in the drawings of a propeller the cylinder
is considered to be rolled out flat on to the plane of the paper. If an ogival
propeller blade is sectioned and rolled out in this way, the face will be seen

54
f , ^.v^r-

to be rendered by a flat straight line on the drawing; but the corresponding


line on the actual blade will follow the helicoidal surface and will be curved
in space.

Propeller blades are usually narrow at the base where they are
attached to the boss, but they are thick here to maintain strength. The
width of chord, measured from leading edge to trailing edge at constant
distance from axis, is a maximum around mid radius falling to zero at tip,
so thaHthe blade outline seen from astern is elliptical. These ellipses can
be quite narrow in low-powered ships, but if power is increased while
propeller diameter is limited by draught, cavitation considerations will
make it necessary to increase the blade areas so that the ellipses get
fattef. At very high powers the blades may resemble distorted discs, with
blade overlapping blade.

Photo courtesy Stone Manganese Marine Limited

A modern propeller with moderate skew.

55
The elliptical outline is often distorted by giving the blade a certain
amount of "sweep-back" or "skew"; that is, the mid-points of the chords
are placed not on a single radial line, but are displaced predominantly aft
along the helix by increasing amounts from root to tip. This gives a blade
of bent-back appearance, an arrangement adopted with the aim of
reducing vibration. The vibration can arise from non-uniformity in the
ship's frictional wake behind the ship's stern when under way, and in
particular from the intense wake "shadow" behind the stern-post of a
single-screw ship. Here the water is largely moving along with the ship, as
compared with the water further out which is sweeping past at near ship-
speed. The effect of the variation in the approach velocity is to expose the
propeller blade to a transient increase in inflow angle as the blade swings
through the shadow. This results in a sudden jllCrC3.S6 3X1 both lift and
drag forces on the blade, The force pulses are transmitted to the hull
through the shaft and cause vibration, if the blade is skewed or swept
back, the different parts of the blade pass the stern post at different
instants so that the force pulses are spread out in time and the intensity of
vibration is reduced. Skewed propellers are more expensive to produce,
and extreme skew may produce strength problems, but a moderate degree
of skew is common practice in single-screw merchant ships. Heavily
skewed propellers are sometimes used in submarines. (Fig. 14)

ELLIPTICAL RAKED

UNLOADED
TIP

Figure 14 Propeller blade shapes

T^ skew is reached when the blade turns right round on


itself and reaches the boss again, some distance downstream of its first
root. The blade thus takes the form of a loop of ribbon, following a
convoluted curve in space. The exaggerated sweep-back angle of such a
propeller makes it more or less immune from fouling by cordage.

56
Another form of distortion sometimes seen in merchant ship
propellers is rake. In a raked propeller the mean axis of each blade is not
square to the shaft axis. Sternwards rake is used to increase the distance
between the blade tips and the stern post, so lessening wake-shadow
effects. Sternwards rake increases the bending moments in the blade roots
as the moments due to hydrodynamic forces on the blades are augmented
by moments due to centrifugal forces. A propeller with sternwards rake
will thus need thicker blade roots than an equivalent unraked one in the
same material.

The JTip Vortex

Consider now what happens around the tip of any surface generating lift.
We have already seen that there is positive pressure on the face and
negative pressure on the back. This causes fluid on the face near the tip to
flow outwards towards the tip, and fluid on the back to flow inwards from
the tips. These motions are imposed on the normal chordwise How of the
fluid, ajid result in a rotary motion in the fluid flowing downstream from
the tip! This forms the tip-vortex, a phenomenon of some practical
importance.

Hydrodynamic analysis of vortex motion shows that velocities are at a


rUlUAilllUlll
->-l o V1" 11 rriU^/j^l
orvr^rnonVjinrf
UUVyllUlg U.1V- nAra OllU
orirl OU
n piCOOUl^l?
nfooeiiyat!liVi
Vioro nro U.L
V- OLV O "f<X
O liUlllilJlUiU.
rn l'ni TV! 11 TTt
For a foil operating in water the tip vortex core pressure can be well below
atmospheric, so inviting cavitation. This means that a heavily-loaded foil
may not only suffer from cavitation on its back; it may also throw off a

57
Tip-vortex cavitation is common in heavily loaded propellers. The
cavity sometimes extends in the form of a helix for a considerable distance
downstream from each propeller blade tip. In clear water the vortex cavity
may be visible to an observer leaning over the stern guard-rails, and
schnorkel-equipped swimmers can see (and hear) the tip cavitation in
passing motor boats. (Fig. 15)

Persistent tip vortices impinging on a ship's rudder can rapidly erode


the steel of the rudder by bubble-collapse erosion. Pits in the rudder or
skeg steelwork at the level of the propeller blade tips are a sure indication
of tip-vortex cavitation.

Where a propeller is heavily loaded, the effects of tip-vortex cavitation


can be avoided or mitigated by "unloading the tips". In the conventional
propeller the thrust forces have a considerable concentration near the
blade tips. By reducing the chord or the angle of incidence or both near
the tips, with compensating increases at smaller radii, the distribution of
force can be made more uniform along the blade radius while maintaining
the same total thrust. The result is that the trailing vortex is spread out
into the form of a sheet shed from the trailing edge instead of being
concentrated into a "rope" shed from the tip. This reduces core velocities
and suctions, and so delays cavitation. Propellers with unloaded tips can
sometimes be identified by their broad non-elliptical blades with a narrow
"bump" at each tip.

The angle of incidence may be varied with radius in order to unload


the tips, as above. Variation may also be introduced in order to match the
propeller to an inflow velocity field which changes with radius because of
the ship's frictional wake. Variation of this angle with radius means that
the blade will not lie on any single helicoidal surface, and the pitch as
defined by geometry will vary with radius. A propeller with such features
will have no single value of pitch, and is sometimes referred te> as a
"variable pitch" propeller. It must not be confused with the propeller with
blades that can be rotated mechanically about their radii to vary thriist at
constant shaft revolutions. Mechanically adjustable propellers, are
conventionally referred to as controllable pitch or CP propellers.

CP Propellers

A propeller in which the pitch could be changed while working would


confer great operational advantages. It would enable thrust to be^varied
without change in shaft revolutions, from full ahead to full astern,'and the
optimum pitch could be selected to give best fuel economy for a wide
variety of loading conditions, particularly useful in a tug or trawler.

True change in geometric pitch would require twistable rubber blades,


as going from one pitch to another would require different angular changes
at different radii. This is clearly not a practical proposition. An
58
approximation is therefore adopted: in the CP propeller, mechanical
arrangements are made to rotate each blade bodily about a radial axis.
This indeed enables thrust to be adjusted and reversed at constant shaft
speed, but it must be appreciated that the CP propeller blade only lies
along a helicoidal surface, and so has optimum performance, at one
particular angular setting. This is referred to as the design pitch setting.
When blades are rotated away from this setting, it follows from the
geometry of the situation that a given angular rotation will produce a
bigger thange in geometric pitch at the blade tip than at the blade root.

This departure from true helicoidal form reduces propeller efficiency.


At tlie setting remote from design pitch, probably full astern, we may even
have the blade roots and tips thrusting in opposite directions, a wasteful
situation. It follows that a CP propeller at full-astern setting will not drive
a ship as fast astern as a normal fixed-pitch propeller fed with the same
horsepower through a shaft turning astern. A CP propeller is thus not just
a simple substitute for a reversing gearbox in the shaft line.

Tl|e CP propeller loses out in comparison with a fixed-pitch propeller


in othefr ways as well. The bulky boss housing the operating mechanism
reduces propeller efficiency marginally, and the complex mechanism
increases weight, vulnerability and cost. The ship designer considering CP
propellers may find the advantages and disadvantages finely balanced. The
advantages seem to gain the day mainly in those ships with frequent need
for manoeuvring, eg ferries, or where there are widely varying conditions of
propeller loading, eg tugs and trawlers.

The Paddle-Wheel Effect


It may be convenient to mention here the so-called paddle-wheel effect.
This has nothing to do with actual paddle-wheels, but refers to a
phenomenon encountered in single-screw ships. When such a ship is
starting from rest, it is often found that the stern swings to one side or the
other as the screw begins to rotate and before the ship gathers headway. If
the propeller is right-handed, the stem moves to starboard if the propeller
is rotated in the ahead sense, and to port if in the astern sense. Skilled
ship handlers make use of this effect when manoeuvring in confined
waters.; ^v-

Thep has been explained from time to time by authorities


that ought to know better (including at one time the Admiralty Manual of
Seamanship) by saying that the water is denser at keel level than near the
surface, and so the blades encounter more resistance lower down and
push the ship in opposition to the motion of the lower blades.

This ac provides the ship handler with a convenient mnemonic


for reminding him which way his stem will swing in response to first
engine orders, but it is in fact wildly erroneous. Sea water admittedly

59
increases in density as one goes down, but the increase is only about
0-004% over 10 metres increase in depth. The density effect could thus
only account for a force of a few newtons at most in the largest ship, which
is far too small to account for the observed ship behaviour.

The true explanation lies rather in the asymmetry of the water flow
through the propeller disc, caused by the proximity of the ship's hull. The
astern case is easier to visualise. With a right-handed propeller turning
astern, the water stream at the top of the disc encounters the stern plating
with a strong forward and some portward velocity, exerting a portward
force on the hull. There is a corresponding starboardward velocity at the
bottom, but the lines here are finer and some of the flow anyway passes
from one side to the other under the keel. The portwards force
preponderates, giving the well-known kick to port.

The ahead case is more complex, and several different effects may
operate together. In a full-bodied ship, the presence of the hull - and of the
, free surface - makes the inflow into the disc very non-uniform. In first
turning from rest, the flow at the top may be so obstructed by hull and
free surface as to stall, and perhaps ventilate, the flow over the blades, so
reducing the athwartships force. The lower blades, working unstalled,
deliver the larger sideways force, hence the kick to starboard.

Conditions change as the ship gathers headway. Axial flow through


the propeller disc strengthens and stall ceases, so that the blades are
biting all the way round. The top blades, however, are working in the
stronger wake shadow, and so encounter higher angles of incidence as
they pass through 12 o'clock. This means that thrust and sideways force
pass through a maximum. There is less compensating side force from the
blade at 6 o'clock, so that the net mean force on the stern is now to port.

But when steaming ahead we also have to consider the sideways


forces exerted by the propeller race on the rudder. With helm ^dead
amidships, the efflux velocity from a right-handed propeller turning ahead
has a starboardward component at the top of the rudder and a portward
one at the bottom, leading to opposing forces on the rudder. Depending on
the shape and span of the rudder, the form of the wake shadow and; the
draught, the resultant may be to one side or the other, but is most
probably to starboard - so opposing the portwise force on the propeller. A
further complicating factor is the influence of the suction field just forward
of the propeller on the stern plating; this will probably result in a net force
to starboard.

It will be seen that the paddle-wheel effect when under way ahead is
the resultant of a wide range of factors, mainly connected with the design
of stern body and rudder, and it can be expected to vary from ship to ship,
and probably with draught in a given ship. It is also affected to some
degree by under-keel clearance. But in any case, once the ship is under

60
way ahead the rudder will be working and a very small constant helm
angle will be enough to cancel out any lateral force arising from paddle-
wheel effects. The helmsman may not be conscious at all of applying this
correction as he holds the ship on course.

Altogether, there are so many uncertainties about the paddle-wheel


effect that the wise mariner encountering a strange ship will avoid making
any assumptions about the effect, and will feel his way carefully until he
finds how the ship responds, particularly when starting from rest.

ei

5
I
%
%
rT\

61
CHAPTER FIVE

TURNING CORNERS

The Hydromechanics of Steering

Now that we have discussed the function of lifting surfaces, we can


consider an evolution in which the whole ship behaves bodily as a lifting
surface; namely, changing course.

We know that we can steer a ship with a rudder, and that the rudder
is usually at the stern. Why should it be there, and not somewhere else,
such as at the bow? The answer is by no means obvious. The steering of a
ship is, in fact, a process of some subtlety, and to examine this process we
need first to go back to some fundamental mechanics.

Newton tells us that a body moving through space will continue


indefinitely at constant speed in a straight line unless acted on by a force,
and that if such a force acts it is equal to the mass of the body multiplied
by the acceleration.

If our body is heading, say, due North and we wish to change course
through 90 degrees to starboard so as to head due East, we can bring it to
rest by applying a Southwards braking force for the necessary time, and
we can then send the body off on its new course by applying an
accelerating force in the Easterly direction. This would, however, be
wasteful in energy; a better and more practical procedure would be to
arrange for the body to follow a curved path linking the old course and the
new, say an arc of a circle, without slowing down.
I
While on the circular path, although the speed along the path is
constant, it can be shown from the geometry of the situation that the body
will have an acceleration directed towards the centre of the circle equal to
the speed squared divided by the radius of the circle. To achieve this,
Newton tells us that we must apply a force equal to this acceleration times
the body's mass. This force is "centripetal", that is, directed towards the
centre of curvature of the path. This direction is at right angles to; the
direction of advance at any instant.

If the body is a ship, we need not only to change the direction of


translation, but we need also to rotate the hull about a vertical axisf^o that
she finishes up moving bows first and not sideways. So, to finish up both
pointing and moving in the right direction, we need to apply: -

1. A thrust force up the line of shaft to counter the ordinary


resistance to motion ahead through the water, and

62
A couple, acting clockwise as seen from above, to rotate the hull
in azimuth against the resistance of the water, and

3. An athwartships force acting to starboard through the ship's


centre of gravity to provide the centripetal acceleration required
to follow the circular path.

We can show from basic mechanics that the couple and the force
through the CG can be replaced by a single force to starboard acting
through a point forward of the CG. (Fig. 16)

(a) TO FOLLOW CURVED PATH A B


THERE NEEDS TO BE A
CENTRIPETAL FORCE F
AND A COUPLE m

G

!

(b) THIS CAN BE CONSIDERED AS AN


EQUIVALENT SINGLE FORCE F
OFFSET Xm/F

Figure 16 Steering-the forces required

up is already provided by the propeller. But


how can we generate the sideways force? One way might be to lay out an
anchor in the sea out to starboard, bringing the cable in to a winch fitted
at the requisite distance forward of the CG, and heaving in on it with the
requisite force. A simple calculation will show that a VLCC on a moderate

63
turn would need a cable tension of the order of a thousand tonnes force,
so we clearly need to look for something more practical.

The solution lies in the hydrodynamics of lifting surfaces. We simply


cause the ship to yaw to starboard. It is convenient to think of the ship's
whole hull as a short wing or lifting surface, sticking downwards into the
water. Under normal steaming conditions there will be no yaw - that is,
the ship will be moving in the direction in which her bow is pointing.
Conditions being symmetrical, there will be no lift force, either to port or
starboard. But if the ship's head yaws to starboard, we at once get a lift
force exerted to starboard of the line of motion. [We also get an increase in
hull drag, so that the ship slows down unless the engines generate more
shaft revolutions in compensation).

Now this hydrodynamic lift force, which is generated by the complex


water flow around and under the hull, has an important property: it acts
through a point forward of amidships. By one of those fortunate
coincidences that are all too rare in engineering, this point is usually fairly
close to that required for the athwartships force discussed above.
Sometimes the points will actually coincide, but in most cases the lift point
is not quite far enough forward. We thus need to apply a supplementary
couple, both to initiate the yaw in the first place and also to maintain the
yaw rate during the turn. (Fig. 17)

OFFSET y DEPENDS ON SHIP FORM


IT IS COMMONLY SOMEWHAT LESS THAN
X IN FIG 16

Figure 17 Force on a yawed ship

know from basic mechanics that a couple plus a force can be


replaced by the same force displaced sideways. It is not in our power to
move the main lift force directly, but it is quite feasible to generate a small
athwartships auxiliary force at some point along the ship's length. If the
point selected coincides with the point of action of the main lift, the

64
turning action will be zero. We must push somewhere else. To generate a
yaw moment to starboard, we can exert a force either to starboard forward
of the main lift, or to port abaft it, the moment being equal to the auxiliary
force times its fore-and-aft separation from the main lift force. Since the
main force acts well forward of amidships, it follows that we get much
more moment for a given auxiliary force by putting our force-generating
device at the stern rather than at the bow. Hence the conventional rudder.

Because of the fortunate coincidence referred to above, the amount of


force demanded from the rudder is very modest; it is far smaller than the
total lift force generated by the yaw. The rudder is in fact one of man's
earliest amplifying mechanisms; for many years it enabled unaided human
brawn to control ship
masses of hundreds of tonnes. (The brawn has
hydraulics to help it nowadays).
sit will be seen from this reasoning that a bow rudder will work, but it
will Vbe relatively ineffective. A powered lateral-thrust unit, being
necessarily fitted some way back from the bow, and so closer still to the
main Centre of lift, will be even less effective. The LTU suffers from the
further drawback that it only generates full thrust when the ship is at rest
or nearly so; much of the thrust of an LTU is lost for hydrodynamic
reasons when the ship gains headway. So we can see on two separate
counts that an LTU has little use as a steering device at sea, whatever its
virtues for manoeuvring in harbour.

There are, of course, other and more practical reasons for fitting the
rudder aft. The slipstream of the propeller enables the rudder to give more
force per unit area than would be available from a bow rudder; there is
more room in the stem to house the steering gear; and a bow rudder is
more vulnerable to damage, But it should be remembered that the main
reason for putting the rudder at the stem is hydrodynamic in nature.

Directional Stability
We have seen that the steering behaviour of a ship depends upon the
relation between the two centres. In most ships the centre of lift caused by
a small yaw is somewhat abaft the centre of action of the athwarts hips
force required to maintain a curved course. This means that, if the rudder
is held amidships, any transient yaw - induced, perhaps, by wind or waves
- will be damped out and the ship will continue on a straight course. Such
a ship is said to be directionally stable. This is in general a desirable
quality, but one can have too much of it; the greater the directional
stability, the greater the rudder force needed to follow a path of given

.the--i-two---centres coincide, the ship with helm amidships will


continue on a straight line until disturbed. If a yaw is induced by external
factors, however, the ship will enter upon a curved path, and will continue

65
to follow it with the helm still amidships. Such a ship has neutral
directional stability.

If the centre of lift is forward of the other centre, the ship with helm
amidships will not continue on a straight course. She will swing on a
curved path, either to port or to starboard, and can only be held on to an
approximately straight course by repeated applications of correcting helm.
Such a ship is directionally unstable. This feature occurs mainly in ships
with very full lines in the after body, such as some tankers with high block
coefficients. Steering a directionally unstable ship demands constant
activity from the helmsman, or auto-pilot, even in calm water. Quite apart
from the load on the steering gear and the helmsman, instability is
undesirable because the continual helm activity, and consequential
yawing motion about the mean heading, increases resistance and fuel
consumption.

We have spoken of the two centres as though they were points fixed in
the ship. This is only approximately true; the centre of lift in particular
depends to some extent on yaw angle, so that a ship can be directionally
unstable at small helm angles and stable at larger angles. Such a ship
responds normally to large helm angles, but still needs to be "ridden" by
the helmsman, using small helm angles, to maintain an approximately
straight course without too much wander.

The centres also move with changes of draught and trim; stern trim
makes for greater stability. The under-keel clearance affects ship
behaviour as well; the changes are complex, but the overall result is that
in shallow water a given rudder angle produces a smaller yaw angle than
in deep water. The effects are two-fold: the ship follows a less curved path,
and the loss of speed due to turning is less. With an under-keel clearance
of 0-4 times the draught, the diameter of the circle steered may beitwice
the diameter for the same helm angle in deep water. The consequences of
this have already been discussed in the chapter on proximity effects. \

Pivoting Point : : :: ,
Mariners sometimes maintain that there is a point somewhere in a ship's
length about which she pivots or turns in some way. There may be some
risk of confusion between this point and the two centres referred to in the
discussion above. The pivoting point is not directly related to these
centres.^.-'.'. t.,.

In following a curved path in the sea a ship is indeed rotating about a


ccntxca tixxs centre is in the water several ship's lengths out to port or
starboard. At a given instant, any point in the ship's length is moving at
right angles to a radial line joining that point to the centre of curvature of
the path. One can think of the ship as part of a great wheel turning about
the centre.

66
It follows from the geometry that an observer at the stern of a ship
turning to starboard will be moving through space in a direction angled to
port of the ship's middle line, and one at the bow in a direction to
starboard. At some point along the ship's length the observer will find
himself moving parallel with the ship's middle line. This is the mariner's
"pivoting point". {Fig. 18)

Since the yaw angle in a turn necessarily carries the bow in towards
the centre of the turn, it follows from the geometry that the pivoting point
must be forward of amidships. In fact in some cases, such as fast planing
craft, the pivoting point can be so far forward as to lie outside the vessel
completely. On the other hand, in slow ships in shallow water the
reduction in yaw angle carries the pivoting point back towards amidships.
This effect may have to be allowed for by mariners handling ships in
shallow and confined waters.

67
CHAPTER SIX

FAIRNESS AND FLOW


Fairness and Fairing
Fair, fairness and fairing are words which tend to be bandied about by
naval architects. Like stability and some other words used in engineering,
the noun fairness and its associated verb and adjective have a range of
specialised meanings; one has to select the appropriate one from the
context.

In the shipyard, fairing a ship's structure is the process of assembling


the various parts - plates, angle bars and so on - so that the joints line up
accurately with one another ready for welding. A completed structure is
said to be fair if it does not exhibit irregularities from faulty assembly, or
excessive corrugation from welding shrinkage.

But in the design office, fairness relates to the shape of the ship as
designed rather than as actually built. The term is applied in particular to
the underwater shape, as set out in the lines plan. In this drawing the hull
shape is defined by a set of curves; the waterlines (sections of the hull
surface by a series of horizontal planes or waterplanes), bow-and-buttock
lines (sections by vertical planes parallel to the plane of symmetiy) and
displacement sections (sections by transverse vertical planes). Sometimes
diagonals are also shown (sections by oblique planes which intersect the
plane of symmetry in a horizontal line).

The naval architect requires all curves shown on the lines plan to be
fair. There is no precise and agreed mathematical definition for the^ term;
one of the classic works of reference described a fair curve simply as one
which is "pleasing to the eye" - the naval architect's eye, presumably For
many years ships' lines were based on little more than this purely
subjective criterion. Curves were drawn using bent wooden battenfe or
splines, held down on to the drawing board with lead weights. The fairest
curve was reckoned to be the one drawn with the stiffest batten, held down
with the smallest number of weights.

This method worked well enough when applied with eyeball and
pencil, and successful ships were designed for many generations in this
manner. However, programming a computer to generate lines curves 'calls
for criteria which are more precise and can be expressed in numerical or
analytical terms. Much mathematical ingenuity has gone into lines-
generating systems. The main aim has been to minimise the rate of change
of curvature along any given curve, while maintaining the displacement
and other hydrostatic particulars required by the ship design. Certainly,
modern computer fairing programs generate lines plans which are

68
pleasing, indeed beautiful, to the eye of the most critical naval architect.

But why should this quality of fairness, however defined, be regarded


as so highly desirable? Above water there is perhaps some aesthetic appeal
in a tastefully swept sheer line or elegantly flared bow, but most of the hull
surface is out of sight under the water and so appearance is irrelevant.
Here the main aim of fairness is the reduction of hull resistance. The
assumption that fairness necessarily reduces drag due to water flow is
natural, universal - and questionable, as we shall see. To discuss this
further we must examine the nature of water flow past a ship's hull. This
has already been touched upon in earlier chapters dealing with frictional
andfvave-making resistance; we must now go into the matter more closely.

Flovr: the Boundary Layer

Consider a uniform stream of water flowing at speed V past a long plank


aligned in the direction of water flow. As we have already seen in
discussing velocity, the water layer actually in contact with the plank will
be at rest, while layers further out from the surface will be moving at
velocities between zero and V. The volume of water where the velocity is
reduced perceptibly below V by the presence of the plank is called the
frictional boundary layer. It increases progressively in thickness from the
leading to the trailing edge of the plank. Downstream, clear of the trailing
edge of the plank, the boundary layer persists for some distance as a
stream of water moving at a mean speed less than V; this is the frictional
wake of the plank.

The flow in the boundary layer can take one of two forms, laminar and
turbulent, In laminar flow the successive layers move smoothly over one
another like shuffled playing cards, the layer velocities increasing
smoothly (but not linearly) as we go from the solid surface towards the
outside of the boundary layer. In turbulent flow, however, the sliding
motion breaks up into a mass of small eddies, so that the velocity
measured at any fixed point in the boundary layer would show a rapid
fluctuation in magnitude and direction. However, the mean velocity at the
point will still be directed sternwards, and will increase as before with
distance from the surface of the plank. The eddying motion is confined
within the thickness of the boundary layer.

In practice, a new boundary layer is usually laminar near the leading


edge, but becomes turbulent as the flow travels aft. In a ship the boundary
is only laminar for a metre or so aft from the stem; thereafter it is fully
turbulent, increasing in thickness until it is a metre or more thick
approaching the stern. In calm weather the frictional boundary layer is
clearly visible to an observer looking over a ship's side.

The boundary layer only comes into existence because the water is
viscous. If there were fluid with zero viscosity, it could flow past a solid
surface with no boundary layer; in the case of our plank the velocity would
be V right up to the surface of the plank. No such fluid exists and so the
situation cannot arise in practice. Nevertheless the flow of "ideal" inviscid
fluids is studied by hydrodynamicists because it can be modelled
mathematically with relative ease; viscous and turbulent flows are more
difficult to represent mathematically. And the theoretically-predicted
behaviour of ideal fluids can offer useful guidance to the behaviour of real
fluids - provided the effects of viscosity and the existence of boundary
layers can be allowed for.

Flow around Bodies


It is useful at this stage to introduce the concept of stream lines. A stream
line is the line describing the path in space followed by any given particle
in a state of steady flow. In the laboratory stream lines can be made
visible, eg by ejecting jets of dye into water flowing past a model. If stream
lines are parallel the flow velocity is constant in magnitude; if they
converge the velocity is increasing, ana the pressure (because of Bernoulli)
is decreasing. If the flow lines diverge, velocity is decreasing and pressure
increasing.

Let us now replace our flat plank by a solid body, say an ellipsoid of
revolution or cigar-shape, deeply immersed in the fluid with its major axis
in the direction of flow. If the fluid is inviscid there will be no boundary
layer. As the flow approaches the bow the flow lines diverge and the
pressure rises, so that the pressure due to flow exerts a sternwards force
on the body. Abreast amidships the lines are crowded together and the
pressure falls; the port and starboard suctions however cancel one another
out. At the stem the lines again converge, matching the pattern at the
bow: the fluid here exerts a forward force on the body. It can be,shown
from the mathematics that the bow and stern forces are equal knd so
cancel out; this is true whether the body is symmetrical fore-andlpft or
not. The body thus exerts no drag force on the body. This is the so-called
Paradox of D'Alembert, and it seems to defy common sense. \

In real life, of course, the fluid is always viscous and the body will
suffer a frictional drag of the sort already discussed under Viscosity in
Chapter 1. But viscosity leads to another type of drag which is quite
distinct from skin friction. This arises because of the effect of the
boundary layer upon the stream lines.

Return now to the ellipsoid body. Let it now be immersed in a real


fluid, like sea water. A boundary layer will at once form, thin at the bow
and getting thicker towards the stern. From the point of view of the fluid
particles outside the layer,the boundary layer of slow-moving particles
behaves like a slight enlargement or swelling of the solid body. In
consequence the stream lines towards the displaced slightly
outwards compared with those a of the body the

70
lines are still spread outwards by the boundary layer travelling
downstream as the frictional wake. As a result the divergence of the lines
at the stern is less marked than at the bow, the slowing down is less and
the local rise in pressure - the "recovery pressure" - is less. The
consequential forwards force on the stern is less than the sternwards force
on the bow, and there is a net drag force on the body, confirming everyday
experience. It is important to note that this force, sometimes called form
drag, is entirely distinct from the skin friction drag. Friction drag enters
the hull through shear stresses acting tangentially to the hull plating, and
form drag through direct stresses (ie, pressures) acting at right angles to
the plating. Form drag is sometimes described as arising from a deficiency
in the stem recovery pressures. (Fig. 19)

FORM DRAG ==RESULTANT OF BERNOULLI FORCES


=Fj-F 2 =NOT ZERO :
(NOTE: FORM DRAG IS DISTINCT FROM FRICTIONAL
DRAG)

Figure 19 Form Drag

In a well-designed ship's hull form drag is usually very small. In


predicting hull resistance from model experiments, form drag is usually
lumped together with wavemaking under the heading "residuary
resistance", and is so assumed to follow the same scaling laws as the
wavemaking resistance.

So far we have assumed that the boundary layers in our ellipsoidal


body and our ship behave in the same manner as in the plank, growing
slowly from stem to stem, and that the water outside the layer follows
round the form of the vessel smoothly just as an ideal fluid would follow a
mathematical representation. There is ample experimental confirmation

71
that this is what does happen, provided that the stern lines of the body or
vessel do not converge too rapidly towards the stern. If, however, the lines
are brought in too sharply, the whole flow pattern can become disrupted;
the flow is said to separate. In this context the term separation has a
special meaning.

Separation of Flow-
Let us consider in more detail the flow pattern around the stern of our
submerged body. Abreast amidships we have seen that the water is
moving aft relative to the vessel at a speed slightly above V. In the after
body, however, the flow lines diverge as the body becomes thinner. The
water speed outside the boundary layer drops because of the spreading of
the flow lines. Inside the boundary layer this drop in speed is
superimposed on the slowing down already caused by the viscous friction.
The two effects acting together can bring the water close to the hull to a
complete stop, and even cause some of it to begin flowing back towards the
bow. When this happens the boundary layer begins to thicken abruptly.
The flow lines outside the layer no longer follow the converging stern lines
of the solid body; instead they break off and flow directly downstream
parallel with the main flow of water remote from the body. The space
between the innermost layers of smooth-flowing water and the body
surface is occupied by a grossly enlarged boundary layer, which now
contains large eddies. The agitation persists downstream of the body,
where it forms the turbulent wake.

Flow breakdown of this nature is referred to as separation of flow. It


should be understood that the water itself does not necessarily separate
from the hull; the stern plating is still usually in contact with eddying and
turbulent water (an exception is the transom stern - see below). It is not
the water as such that separates, but rather that part of it which is|flowing
smoothly and uniformly. One can think of the innermost stream line; this
hugs the hull closely as it comes aft from the stem, separated only t>y the
relatively thin boundary layer. Somewhere abaft amidships, this line
breaks away or separates from the hull, leaving the intervening space full
of confused and eddying water.

Separation of flow can, and sometimes does, happen in ship^ and


boats. In displacement vessels of moderate speed separation is entirely
pernicious, as its main effect is to suppress the recovery pressures on the
stern and hence to increase the hull drag. In a badly separating hull this
increase can be very significant. In a sailing vessel steering can be'affected
if the rudder is blanketed in the confused water of the turbulent wake,
which has only a low mean velocity relative to the hull. In a powered
vessel, the rudder usually has the benefit of the propeller race, even if the
hull flow has separated; but even so separation can affect steering as can
be seen from the case study below.

72
Photo courtesy B.P.

We can see that a slow ship can have quite a bluff entry forward because wave
making is not a major concern. Some tankers have bulbous bows: these
function by reducing vortex formation around the forefoot.
73
Separation in a displacement ship is usually caused by too steep a
slope in the lines of the after body. A rule of thumb sometimes quoted is
that waterlines here should not have a slope greater than 20 relative to
the middle-line plane; slope in buttock and diagonal lines should also be
limited. This results in one of the classic ship-design conflicts; the ship
designer wants to place his engine-room as close to the stern as possible,
but if he puts it too close the beam required for the machinery pushes the
lines out so that the critical slope is exceeded, flow separates and the ship
fails to reach her designed speed.

It should be noted that separation is not sensitive to speed; it imposes


a degree of fineness in the stern lines in slow ships as well as in fast. We
have seen that a slow ship can have quite a bluff entry forward because
wave-making is not a major concern; hence the "cod's head and herring's
tail" form seen in the sailing ships of earlier years. Similar considerations
lead to the tear-drop shape of nuclear submarines; although these vessels
are far from slow, they avoid wave-making by operating remote from the
interface, and they use their blunt fore ends to house a variety of
transducers and weaponry.

In a hull form which has been steepened so that the flow is almost
but not quite separating, we may have flow conditions which are barely
stable, and where a minor disruption can cause the flow to separate. Here
it is particularly important for the plating to be smooth, with no
discontinuities in direction or curvature. So we have at last established
some connection between flow and fairness. Note that unfairness alone
does not cause separation, but it can trigger separation if other conditions
are favourable.

ACase Study \
\
\
We can illustrate the sometimes curious effects of separated flow by
describing the case of a certain small craft that had steering difficulties.
While the rudder was of conventional area and design, the boat's response
to helm was extremely sluggish, and the turning circle was large.
Investigation showed that the deck at the stern had been made elliptical in
plan for operational reasons. This elliptical form had been carried down
into the underwater hull, leading to waterlines which were unduly steep.
Model experiments showed that separation was occurring, and that the
point of separation was unstable, wandering forward and aft along the
ship's side. This was confirmed by observation at ship. Trials showed that
when the helm was put to starboard the yaw produced speeded5 up the
flow round the port quarter, suppressing the separation there; while the
slower flow round the starboard quarter aggravated the separation on that
side/ The streamline flow on the port quarter restored the recovery
pressures there; these, having a component to starboard, acted in
opposition to the rudder force. As a result, the vessel (in the Cox'n's words)
"steered like a pig".
-mkio-i e

The steering was much improved by fitting wedge-shaped "puddings"


on both quarters, effectively putting sharp outward kinks into the
waterlines. These kinks pinned down the separation to fixed points, so
that the recovery pressures were fixed at permanently low values each
side, at some cost in hull drag. (The owner disliked the appearance of the
puddings and unshipped them, preferring to live with his poor steering).
y
Transom Stems

In vessels which are fast in relation to their length, particularly small craft
where a large beam is dictated by stability and other considerations, a
completely separation-free conventional stern would result in an
unacpeptably long hull. In such cases the designer usually accepts that
some separation has to be lived with, and the problem resolves itself into
minimising its adverse effects. This means seeing that the separation
happens at a place chosen by the designer. His usual procedure is to cut
the stqpi off square and to leave the corners sharp - surely the grossest
kind oft unfairness. When the craft is under way, the flow separates quite
cleanly^where it is meant to, at the sharp corners. At low speeds the
transom drags along behind it the usual turbulent water, but as speed
increases the flow shoots aft from the bottom as well as the sides, leaving
the transom dry. The greater part of the turbulent wake disappears, but
this does not mean that the separation drag has gone too; the dry transom
has no water pressure on it at all so that recovery forces are virtually zero.

Transom-corner separation is completely stable, and there is no


interference with steering due to the separation wandering about. In
transom-sterned craft the rudder is usually below the transom, well
immersed in the stream-line flow, and possibly benefiting from the
propeller race as well.

To work properly a transom must have sharp square edges all round.
On occasion transoms have been given radiused corners, for aesthetic or
other reasons. The result is that separation is no longer clean; water is
drawn round the corners to fall into a confused tumbling mass pulled
along behind the transom at all speeds. As well as being unsightly, this
kind of flow augments the hull drag because of the suction around the
curved corners. This is a case where the extreme of unfairness - a sharp
corner of zero radius - is to be preferred to a supposedly fair curved
profile.

Trailing Vortices
, Separation is not always confined to the after end. A curious form of
separated flow sometimes occurs in the fore-body of full-bodied ships.
Where the entry is blunt, the flow lines near the bow plunge sharply
downwards at an angle of slope which may reach 45 in the region of the

75
forefoot. If the hull form here is not carefully designed the flow may
separate obliquely on each side, attaching again near the forward shoulder
where the flow lines bunch and the flow speeds up. The effect of the
localised detachment is to throw off trailing vortices. These stream astern,
rather like an aircraft's wingtip vortices or the tip vortices from propeller
blades.

The trailing vortices yield no surface disturbance and cannot be seen


on the full scale, but they can be demonstrated in the laboratory with
models. In an extreme case similar vortices can also be thrown off from the
shoulders in the after body; a ship so afflicted will have four separate
vortices trailing astern from the hull. These vortices exist quite separately
from the normal frictional wake and the propeller race.

The point about hull-generated vortices is that they contain energy,


and so represent an additional component of hull drag. The hull designer
must therefore try to lay out his hull lines, both forward and aft, in such a
way that no trailing vortices are induced. The drag-reducing virtues
claimed for bulbous bows in some slow full-bodied vessels probably lie in
the smooth profile offered to the downward-plunging flow near the forefoot,
so avoiding separation where the water follows round the turn of bilge to
get under the fiat bottom. (Bulbous bows in fine fast ships operate in quite
a different manner, by reducing wave formation at the bow) .

Trailing vortices may be a further embarrassment if they enter the


propeller disc; the irregular inflow to the propeller so produced may be a
contributory cause in some cases of vibration.

Why Fairness?

Whatever its attractions aesthetically, we have seen that fairnesk alone


does not guarantee satisfactory water flow around a hull. One can d^ign a
hull where all lines are ^pleasing to the eye" and satisfy fevery
mathematical criterion devised for fairness, but still find that separation
occurs in the after body. Likewise the most elegant, carefully faired ^fore-
body can still throw off trailing vortices. Clearly, fairness alone is not
enough. .::;. ^

On the other hand, fairness may be supererogatory. The turn of bilge


amidships in most merchant ships exhibits gross unfairness, being formed
from a quarter-circular cylinder joining fiat bottom to flat sidd The
sectional curve is a circular arc butting on to two straight lines? - with
infinite rate of change of curvature at the joints, offending against all
criteria of fairness. Again, the bilge keel itself represents the ultimate in
unfairness, a knife edge where curvature jumps from zero to near-infinite
and back.

76
The bulbous bow in a fast ship operates by reducing wave formation. The bulb
enables the buoyancy forward to be carried low down, so minimising the
waterline angle of entry.

77
These unfairnesses are entirely acceptable because the flow in
ordinary ahead motion is along them and not across them, The designer is
careful to position the bilge keel on the hull so that it lies exactly along the
expected line of flow throughout its length. In a slow ship the flow
direction is nearly enough fore-and-aft at constant depth along the parallel
body amidships, so that the bilge keel can be placed in a diagonal plane.
In a fast vessel such as a destroyer it cannot be assumed that the flow lies
along the diagonal plane, and it may be necessary to carry out model
experiments to plot the actual line of flow before designing the bilge keel.

Flow direction predictions are sometimes needed for other hull


features as well. For example, some fast twin-shaft lifeboats have their
propellers recessed into hollows tunneled into the hull, so that the
propellers and shafting are protected if the boat runs aground. The edges
of the tunnels make more or less sharp corners as seen in section, a
feature that would cause separation and trailing vortices if the flow were to
cross them obliquely. Care is therefore taken to align the tunnel edges in
the expected direction of flow when the vessel is under way at operational
speed.

The flow past a ship's hull can sometimes take unexpected forms. At
one time it was customary to assume that "flow in the fore body is along
diagonals and in the after body along buttock lines", but this was a crude
approximation. Where direction is important - as in designing the tunnels
referred to above - some objective method of flow prediction is needed. The
classical method is to use a physical ship model in a tank; flow direction is
indicated by tufts of thread anchored at various points around the hull, or
by streams of dye emitted from small holes in the model. Lately computer
methods have become available, saving the cost of model and tank work.
\
Some designers have exploited the tolerance of flow to aligned
unfairnesses, and have produced hull designs with knuckles or corners
parallel with the flow. This enables the steelwork to be simplified > to a
series of flats, so avoiding the need for frame bending and plate rolling.
The result is a cheaper, if less beautiful, ship. Other designers have
accepted curvature, but have restricted it to cylindrical and conical forms
which are easy to roll; any plate with curvature in two directions, such as
a spherical or ellipsoidal surface, is difficult and expensive to form.
It will be clear from the earlier discussions that these simplified forms
can be fully acceptable hydrodynamically provided full account is taken of
the flow directions around the hull, so that flow is not called upon to pass
across sharp convex corners. The designer working in GRP or FRP, of
course, has no call to use simplified forms; his material enables him to use
as much two-directional curvature as he likes. However, all designers need
to watch carefully areas of diverging flow lines, where asking the water to
diverge too sharply may result in separating flow and all the troubles that
stem from it.

78
Some ships have been built with angular comers or knuckles in the hull
plating. This enables steelwork to be simplified and makes for economy in
construction, but the knuckles must be carefully aligned with the water flow
to minimise drag.

All in all, fairness is a complicated business. There is much more to


designing a ship's lines than just making them "pleasing to the eye". One
must study the water and the ways in which it behaves, which are subtle
and sometimes unexpected. Perhaps the truly fair line is one which is
"pleasing to the water".

79
CHAPTER SEVEN

STABILITY AND STABILISATION

Weight and Buoyancy


Two ships steam in company on parallel courses in a beam sea. One rolls
through 5 out-to-out, the other through 10. Which ship is the more
stable?

The landsman, identifying stability with steadiness, would plump for


the 5 ship, and take passage in her for preference. The mariner, knowing
something about resonant rolling and related matters, realises that there
is more to this question than first meets the eye, and that it cannot be
answered without more information, The 10 ship could even be the safer
one.

In engineering the word stability has a specialised meaning; it is that


property of a system which causes it to return to its equilibrium condition
after suffering a transient disturbance therefrom. In the more specific case
of a ship floating upright in calm water, stability is the property which
returns her to the upright after she has been heeled by a transient heeling
moment.

A ship at any time is acted on by two major forces: her weight W,


acting vertically downwards through the centre of gravity G; and the
buoyancy force F, acting vertically upwards through the centre of
buoyancy B. For a ship floating in equilibrium in calm water, F is equal to
W and B lies in the same vertical line as G. The forces cancel out, and
there is no heeling moment. ^

If now the ship is heeled, G and B will no longer in general lie on the
same vertical line, and the forces will form a couple. The offset distance
between the lines of action of weight and buoyancy is called the righting
lever GZ. The moment of the couple acting on the ship is equal to GZ times
the weight of ship W. GZ is taken to be positive if the moment acts to
return the ship to the upright, negative if it acts to increase the he^l. A
positive GZ thus implies stability.

The quantity GZ varies with angle of heel. With increasing heel it


usually increases to a maximum, thereafter falling again to zero and
becoming negative at large angles. The span of heel angles over which GZ
is positive is referred to as the range of stability; the curve falls through
zero at the angle of vanishing stability. A ship forcibly inclined up to any
angle within the range of stability and then released will return to the
upright; if she is released beyond this angle she will capsize. A ship acted
on by a constant and persisting heeling moment, such as might be caused

80
by a shift of cargo, will capsize if the moment carries her beyond the angle
of maximum GZ. Note that this angle is smaller than the angle of
vanishing stability.

GZ curves

The stability of a ship cannot be expressed as a single figure; the only


complete statement of stability is a set of GZ values for all operating
displacements and trims, for all heel angles up to at least 90. Naval
architects compute such values of GZ at an early stage in the design
process. Hie information is commonly given in the form of "GZ Curves",
graphs of GZ plotted on a base of heel angle. Such curves assume a given

(a) CROSS SECTION OF A SHIP HEELED WHERE THE


BUOYANCY ACTS VERTICALLY UPWARDS AND

(b) A TYPICAL GZ CURVE

Figure 20 Stability curve


81
GZ curves have various characteristics, much mulled over by naval
architects, notably the initial slope, the maximum ordinate, and the area
below the curve (measured in metre-radians) up to various stated heel
angles, Maritime administrations commonly lay down stability standards
for certain categories of ships in terms of GZ characteristics; for example,
the IMO standards for passenger vessels specify minimum values for the
above quantities, and the UK Department of Transport standards for sail
training vessels stipulate minimum values for the range of stability as a
function of craft length.

Ship's officers should ideally have access to the GZ curves for their
vessel, and they should be able to correct them for G being at a different
height from that assumed when the curves were plotted. Thus, if G rises
through a distance h, the decrease in GZ value at heel is h.sin .

The Metacentric Height


Consider a ship in calm water heeled through a small angle . The vertical
through the heeled centre of buoyancy cuts the middle-line plane at a
point M, referred to as the metacentre. The distance GM is the metacentric
height. This is effectively constant for small values of G. The distance
between M and the upright centre of buoyancy B, designated BM, is
derived from the classic formula BM = I/V, where I is die moment of
inertia (strictly, the second moment of area) of the waterplane and V is the
volume of displacement, both easily calculated from the lines plans. (For
the derivation of this formula, see the early pages of almost any book on
Naval Architecture). Note that the quantity I depends upon waterplane
widths cubed: hence the strong influence of ship's beam on stability.

From the geometry we can see that GZ = GM. for small values of .
It follows that the slope of the GZ curve at the origin is equal to GM. The
value of GM can thus easily be derived from the GZ curves by drawing a
tangent to the curve at zero heel; this line cuts the ordinate at 57-3- (one
radian) at a height equal to GM, as measured on the GZ scale. This
tangent is sometimes shown already drawn in on sets of GZ curves. ;

The metacentric height GM is a measure of a ship's stiffness^ or


resistance to heeling moments, at small angles of heel. There was at one
time a tendency to regard GM as a prime criterion of stability, to the
exclusion of other stability characteristics; this could lead to design errors.
It is only too easy to design a ship or boat with a satisfyingly large GM, but
with a GZ which falls to zero at a dangerously small heel angle. The' classic
case usually quoted is the mid-Victorian turret warship HMS CAPTAIN,
which capsized under sail in the Bay of Biscay when the wind heeling
moment brought the lee rail under and GZ went negative.

82
A wise designer will keep his eye constantly on the whole GZ curve,
not only for the intact ship but also for the ship in possible damaged
conditions.

The Inclining Experiment

The ship's designer knows where the vessel's centre of gravity G ought to
be, but it is advisable in practice (indeed, obligatory in UK passenger
shipsKto conduct measurements at ship from time to time to find out
where G actually is. The metacentric height provided us with a convenient
procedure for doing this, by means of the so-called inclining experiment or
test^

The ship is subjected to a known heeling moment, applied by moving


ballast weights across the deck or transferring liquids between tanks, and
the resulting heeling angle is measured. Since the moment is equal to ship
weight times GZ, we can find GZ; since GZ = GM.sinG, we can deduce GM.
The position of M depends entirely on ship's geometry, and so does not
change in service. The height of M above keel KM for a range of draughts is
shown! on the ship's hydrostatic curves, having been calculated by the
designer. Knowing KM and GM, we can locate G.

Repeated inclinings throughout a ship's life will usually show that G


rises with time, mainly because of numerous small unrecorded additions
of weight in the higher parts of the ship. The designer of a new ship should
allow for some upward drift of G through the expected life of the ship.

Mariners will appreciate that every ship has a natural period of roll.
This depends upon the GM; the higher the GM, the shorter the period.
Indeed, an excessive GM can cause a rolling motion that is so quick and
lively that the ship is rendered uncomfortable and even dangerous to
passengers and crew; passenger vessels usually have a relatively low GM.

With a small vessel, the connection between period and GM provides


an alternative to the inclining experiment as a method for finding GM.
With the vessel in calm water, a roll is induced by craning a sided weight
quickly off the deck, and the resulting motion is timed by stop-watch over
several cycles as the motion dies away. GM is deduced from a formula of
the type: period = K divided by the square root of GM, where K is an
empirical constant depending on the type of vessel. The method is only
approximate, but it is cheap and quick, and it may on occasion serve for a
rough check on stability if there is any doubt about a ship's condition, eg
after damage . At sea, a slow and soggy roll motion may give early warning
of loss of stability, eg due to icing on the upperworks.

83
Ships in Waves

Stability as discussed so far is a calm-water concept. We can, however,


apply our stability theory to a ship in waves. A good approximation to roll
behaviour can be obtained by assuming the ship to be floating in calm
water, but acted upon by a cyclic moment caused by the departure of the
sea surface from the horizontal. Mathematical analysis shows that the
rolling amplitude depends not only on the wave slope but also on the
encounter period, rising to a maximum when the encounter period
coincides with the natural period of roll of the ship; this is the condition of
resonance.

Such resonance is a familiar phenomenon at sea. If resonant rolling is


troublesome, the short-term treatment is to change the frequency of wave
encounter by changing course or speed. The longer-term treatment,
sometimes adopted in passenger ships, is to design the ship with a
deliberately low GM so as to make the natural period of roll as long as
possible, well clear of any likely wave-encounter period. The designer's
problem with such ships is to maintain safety while retaining a low GM.
He can still get a fat GZ curve and a large vanishing angle if he has plenty
of freeboard to play with and watertight structure high up.

Following Sea

Consider now a ship in a following sea, where the wavelength is


comparable with the ship's length. The waves themselves, with crests
square to the ship's axis, will have no heeling action. If the ship is slow,
the waves will pass her by from stern to bow, but a fast ship (particularly
in shallow water) may find herself keeping station on a single wave for
some time. If the trough is amidships and the crests at bow and stern
there is not much to worry about, but if the crest is amidships therd could
possibly be some trouble. This is because a ship's sides are usually^more
or less vertical amidships, but flared towards the ends. The fall in water
levels at the ends reduces the waterplane widths there/ while the* rise
amidships produces no compensating increase. The quantity I,, the
waterplane moment of inertia, is therefore diminished, leading to a
reduction in BM and hence in GM. ;

Thus, a ship perched on a following wave can lose part of her effective
GM over a period of time. In these circumstances a moderate heeling
moment, as from the wind, can produce an unexpected and even alarming
angle of heel. This effect is particularly marked in forms with heavily flared
and overhanging sterns, where stability depends upon keeping the stern
wet. Ship's officers cannot do much about the hull form/ but they may be
able to avoid the wave-riding situation by adjusting course or speed. If the
waves pass the ship fairly quickly the periods of GM reduction still occur,
but are too short in duration to allow embarrassing heels to build up.

84
A related phenomenon is the so-called "parametric rolling". If a ship is
acted on by a constant heeling moment m (due perhaps to a beam wind or
asymmetric loading) she will heel to an angle given by in = W.GZ =
W.GM.8 if is small. Here W and m are both constant, so that GM. has
to be constant too. It follows that will vary if GM varies. But we have
already seen that the effective GM of a flared ship will vary as a wave
travels along the length. Thus, a ship in these circumstances will
experience a varying - that is, a rolling motion - even when the seas
encountered are coming from dead ahead or astern. The amplitude could
be troublesome near resonance, which occurs when the wave encounter
period is half the natural roll period of the ship. As above, the short-term
treatment is to avoid resonance by adjusting course or speed, and also to
avdid if possible carrying any list. (Fig. 21)

c
w

Figure 21 Loss of stability with wave crest amidships;


note loss of waterplane area and inertia. (C indicates
waterplane in calm water, W water-plane in wave.)

Free Surface
We have noted that BM, and hence GM, depend upon waterplane inertia.
Effective inertia can be lost from internal free surfaces of liquid, such as in
slack tanks. The effective loss of GM from a free surface is given by i/V,
where l is the moment of inertia of the free surface about its own centre of
area. One can think of the ship's waterplane inertia being reduced from I
to I - i. If i is large, the loss of GM could be serious.

One can see from this that it is quite easy to design a ship in which
the initial stability can be destroyed at a stroke. One does this by
arranging an internal compartment or cargo space which is so long and

85
wide that its moment of inertia i is comparable with that of the ship's
waterplane I. If we put this compartment low down, we can expect it to
flood following collision or other mishap. In that event BM shrinks
drastically from I/V to (I - i)/V, GM and GZ become negative and the ship
capsizes. Don't the regulations prevent this? At the time of writing, they do
not.

Stabilisers
As already mentioned, every ship has Its own natural period of roll; if
heeled and released in calm water, it will roll back and forth like a
pendulum. The motion is relatively lightly damped, so that the rolling
persists over a number of cycles. It is a characteristic of lightly damped
systems that the amplitude of oscillation will build up markedly if the
system is subjected to a cyclic disturbance with frequency close to the
natural frequency of the system. In ship terms, the roll will build up to
large angles If beam seas are encountered at the right frequency; this is
the condition of roll resonance.

Sizes, shapes and materials being what they are, most conventional
ships finish up with natural periods somewhere in the range 8 to 16
seconds. It so happens, as a result of one of nature's pleasantries, that
most of the bigger waves in the world's oceans also have periods in the
range 8 to 16 seconds. The result is that resonance in roll is a common
phenomenon at sea.

As with other engineering systems, the amplitude of resonant


response is very sensitive to damping; it can be sharply reduced by a
moderate increase of the damping, much as a pendulum can be made less
responsive by immersing the bob in a bucket of water (whence the term
damping). Resonant rolling in a ship can be usefully decreased by applying
roll moments with the correct sense and phasing to achieve daiSping.
Quite moderate moments properly applied can produce useful reductions
in rolling. A device designed and installed to provide such damping
moments is properly called a roll damper. However, in popular usage it is
often referred to as a "stabiliser"; this is a thoroughly bad term, as the
device does not increase the stability of a ship, in fact it may actually
dccrcHsc it*

There is a difference here between the rolling and pitching motions of


a ship. While rolling is lightly damped, pitching is already heavily damped
by the shape of the ship. Resonant pitching can occur, but the extra
damping forces needed to achieve useful reductions in pitching motion are
enormous. Attempts have been made to devise pitching dampers, eg in the
form of flat plates attached to the forefoot, but such devices have to be
extremely large and heavy to be of any use, and they have not been found
cost effective. Effective roll dampers of many different types have, however,
been developed.

86
The first roll damper exploited the free surface effect. Every tank in a
ship has a natural period, the time taken for the water or other liquid
content to slosh from port to starboard and back again. One can design a
tank to have a natural period of slosh equal to the ship's natural period of
roll. Then, when the ship rolls resonantly in a beam sea, the fluid in the
tank will rush back and forth with great vigour, dissipating its energies by
thumping against the end bulkheads and by swirling round internal
obstructions, which may be inserted for the purpose. The energy so spent
in the tank is abstracted from the ship's rolling motion, which is thereby
reduced in amplitude. Such a tank constitutes a simple roll damper, well
known and exploited in ships since Victorian days. This device still works
when the ship is not under way, a useful virtue for ships which require to
lie stopped or move slowly at sea, eg cable ships, light ships

* If a tank roll damper is added to an existing ship, it should be noted


that\ the free surface of the water in the tank actually results in a
reduction in the effective GM of the ship; the device could even be called a
"de-stabiliser", although it remains an effective roll-damper. A damping
tank should be provided with means to dump its contents rapidly to sea,
for use when circumstances arise, eg collision damage, where GM needs to
be maximised.

The tuned tank was the first of the so-called stabilisers. Since its
Invention great energy and ingenuity have gone into devising other
systems, involving pumps, air blowers, rolling weights, gyroscopes,
activated fins and so on. With the right control gear one can even use the
rudder to reduce roll, by exploiting the cross-coupling between roll and
yaw.

Inventors proliferate and patent lawyers profit, but ships still roll. Any
new inventor entering the field should bear in mind that a roll-damping
device should generate a moment in quadrature with the roll angle, not in
phase with it; that is to say, the moment should oppose the roll velocity,
not the roll angle. Some very elaborate control systems have been
produced, with integral and differential as well as angle terms in the
control signal, but the basic requirement remains simple - to suck energy
out of the rolling moment, and that means opposing the roll velocity.
Complications introduced into the control signal usually have the aim of
countering the effects of time-delays in the operation of, eg, fin
mechanisms.

The simplest, and probably the most cost-effective, roll damper of the
lot is probably the h ^ bilge keel; no maintenance, power
consumption, noise, inboard volume or patent fees.

87
CHAPTER EIGHT

VIBRATION

Sources of Vibration

The mariner is often uncomfortably aware of the fact that ships vibrate.
Sometimes the vibration is clearly attributable to an inboard mechanical
source, such as a diesel engine; we shall not discuss vibrations of inboard
origin here. Other kinds of vibration, however, arise from various kinds of
interaction between the ship and the water, and so do concern us. Most of
these are associated in one way or another with the propeller.

Consider first the situation where a propeller suffers damage, such as the
loss of part or all of one blade, or distortion of a blade. When the shaft
rotates the centrifugal forces on the propeller boss no longer balance out,
and we have in effect a radial force on the shaft bearing rotating at the
speed of the shaft. This induces the whole hull of the ship to flex and
vibrate as a beam, in both the horizontal and vertical planes. If the shaft
frequency picks up one of the hull natural frequencies the vibration
amplitude may be severe. Hie effect is similar if the propeller is intact but
the shaft is bent.

The centrifugal fluctuation is not the only vibratoiy stimulus. The


damaged propeller still generates thrust, but this thrust is now offset from
the shaft axis/This is equivalent to a thrust up the axis {as with an intact
propeller) combined with a couple which rotates with the shaft. The
moment of the rotating couple exerts another vibratory stimulus upon the
hull. Note that both vibratory mechanisms arising from damage excise the
ship at shaft frequency. \
i
Wake Effects
Consider now the intact propeller. A properly shaped and balanced screw
propeller, rotating in a uniform flow of water moving parallel to the shaft
axis and remote from any solid body, will generate a pure and constant
thrust force acting exactly up the axis of the propeller shaft, without any
vibratory component.
i
If, however, the velocity of the incoming water stream is not uniformly
distributed over the disc swept out by the propeller, each propeller blade
wiH varying angles of incidence as it
rotates about the shaft. The lift and drag forces exerted by the water on
the blade will fluctuate in response to the changes in incidence.

88
Consider, for example, the effect of a local "shadow" extending over a
small sector of the disc, where the inflow is somewhat slower than
elsewhere. Since the blade sweeps round the disc at constant velocity, it
will find the incidence angle in the shadow increased, with the result that
lift and drag forces rise temporarily to a local peak. These forces have
components up the shaft axis and at right angles to the axis. Since they
fluctuate they are vibratory stimulus forces, and the component square to
the axis can excite vibration of the ship's hull as a beam. (The component
up the shaft represents only a marginal fluctuation superimposed on the
normal propeller thrust, and has only minor effects on the ship).

f In a single-screw vessel the propeller operates in the frictional wake of


the hull, which can produce an inflow velocity which is quite non-uniform
over the disc swept out by the propeller. The actual pattern of velocity
distribution varies from ship to ship, but in most ships there Is usually a
district shadow of strong wake (ie, of low inflow velocity) at 12 o'clock,
whei|e the blade passes behind the sternpost. This feature is a common
sourde of vibration in merchant ships.
*

It will be noted that the frequency of stimulus from the wake effect is
the blade-passage frequency; ie the shaft rotational frequency times the
number of blades. A four-bladed screw turning at 120 RPM has a blade-
passage frequency of 480 cycles per minute, or 8 hertz.
If wake-effect vibration is encountered, various palliative measures
are possible. Fins can be added to< the hull to modify the velocity
distribution through the disc; the number of propeller blades can be
increased to raise the frequency and reduce the amplitude of the stimulus;
and the blades can be "skewed back" so that each blade transits the
shadow progressively and not all at once. However, a complete cure can
rarely be achieved by these means,

Pressure Effects -
We can regard the wake effect as the hull acting upon the propeller
through the influence of its frictional wake. With pressure effects, we have
the propeller acting upon the hull through the influence of its pressure
field.

Each blade of a working propeller is generating hydrodynamic lift. As


we have seen in discussing foils, this means that the flow over the blade
back is speeded up in relation to the flow over the face; the pressure on
the back is correspondingly reduced. This is the origin of the lift force.

But the suction field is not confined to the back skin of the propeller
blade; it extends some way out into the water, and acts upon any solid
body which happens to be placed there. Thus, if the blade suction field
impinges on any part of the hull plating as the blade sweeps round, then

89
that area of plating will suffer a transient suction force.

This may not be a significant mechanism of vibration stimulus in


many single-screw merchant ships, where there is little plating in the
swept suction field; but in other hull forms, such as twin screw-vessels
and particularly in shallow-draught craft with propellers in tunnels under
the bottom, we may well find propeller blade tips sweeping quite close to
bottom plating. Designers try to keep these clearances as large as possible
- a clearance of at least one-half the propeller diameter is usually aimed at
- but nevertheless pressure-field forces can become large enough to set up
hull vibration. In extreme cases they can also cause local fatigue failures
in the plating immediately exposed to the suction fields. The frequency of
pressure-stimulated vibration Is again that of blade-passage.

Cavitation
The nature of cavitation and the manner of its occurrence on and near
propeller blades has been discussed in earlier chapters. It will be recalled
that cavitation is undesirable for a variety of reasons, and it is for the most
part avoided by limiting the load per unit area of the propeller blade, and
by choosing the blade section to give a uniform rather than peaky suction
distribution over the back. Tip-vortex cavitation can be minimised by
adjusting the distribution of lift along the radius to avoid a peak near the
tip-

Nevertheless, cavitation is still met on occasion. In the vibration


context the most troublesome sort is that which occurs as the blade
suffers a lift peak on passing through a wake shadow. The cavity produces
a transient change in lift and drag as it forms, and then radiates a sharp
pressure pulse through the water as it collapses. This pulse acts oh local
bottom plating rather like a hammer blow. V

The forces from cavitation are sharp and peaky, and are most
noticeable inboard rather as noise than as vibration. Doubtless they
contribute to the general level of vibration, as an addition to that already
present from wake and pressure effects; the noise, however, is quite
separate and is distinctive from its sizzling and banging nature.

Singing
Occasionally one meets a propeller which emits a continuous Audio-
frequency tone or whistle through the water. This is known as singing; it is
a source of annoyance to crew and passengers, and a gift to hostile
submarines, though hardly a serious detriment to performance. The
singing is attributable to the mechanical elastic vibration of individual
blades, much in the manner of a bell or gong.

90
For such vibration to be sustained it needs to be fed with mechanical
energy. The origin of this energy can be sought in the nature of the water
flow round the trailing edge of the blade. The hydrodynamically ideal
shape for the blade trailing edge would be a sharp thin knife-edge, but in
the real world such an edge would be highly vulnerable to damage from
floating objects and impossible to maintain. It is much more practical to
round off the blade edge into a blunt circular or elliptical section, and this
is usually done. But we have already seen that bodies with rounded sterns
will suffer from separation of flow, and that the position of the point at
which separation takes place may be unstable. Consider what happens if
the separation point moves forward on one surface; it will then move aft on
the,other. Recovery pressures are greater on the unseparated side than on
the other, so that there is a net force deflecting the blade tail towards the
separated side.

The blade now twists eiastieally under this force, so changing the
incidence of flow so that the separation point on the separated side slips
aft and that on the other side moves forward in response. The force
reverses, pushing the blade in the opposite direction. The cycle repeats,
and the vibration is sustained with energy abstracted from the water flow.

It will be seen that there are some similarities with the flow
phenomena encountered in the elliptical-sterned craft discussed in an
earlier chapter. The curative treatment is also similar. The elliptical
trailing edge on the propeller blade is chiselled away, leaving a bevelled
edge with sharply-angled corners. The corners anchor down the separation
points and the cyclic action disappears.

It is notable that singing only occurs in a propeller cast from bronze


or similar material with a low internal damping within the metal structure.
Such a screw will ring like a bell when struck with a hammer. Cast-iron
propellers, at one time common in merchant ships, cannot be persuaded
to sing. A particularly hysteretic or "dead" copper-manganese alloy has
been considered for specialised applications where the least risk of singing
cannot be accepted.

Wave-induced Vibration
We now come to the only serious kind of water-related vibration that is not
associated with the propeller. This is the beam-flexure type of vibration of
the whole hull which is stimulated by repeated encounter of the bow with
the waves of the sea. It is sometimes referred to as springing, and is
usually associated only with the largest ships.

The possibility of springing vibration in smaller ships has always


existed in theory, but it has not been troublesome because - in pre-VLCC
days - resonance was not approached. In discussing stability, we have
already noted that the wave energy in the world's wave systems lies mainly

91
in waves with periods in the range 8 to 16 seconds. Note that these figures
are for wave crests passing a point fixed in the sea; they must be
distinguished from ship wave encounter periods, which are affected by the
speed of the ship.

For ships smaller than VLCCs the period of the fundamental vibration
of a ship's hull as a beam is of the order of a second or less, so that such a
ship is well clear of resonance with energetic waves. She may perhaps
resonate with shorter waves, but these have so little energy in them that
the response is not enough to worry about.

With the very large vessels now in service the situation is different.
The increase in dimensions and mass, together with the acceptance of
higher stresses in high-tensile materials, has lengthened the natural
period of beam-style vibration so that natural periods as long as 3 or even
4 seconds are possible.

Consider such a vessel lying stopped, head to weather. A train of


waves with wavelength between crests of 100 metres approaches. These
waves travel with a crest celerity of around 24 knots and strike the bow at
8 second intervals. This seaway is at the lower end of our "energetic"
spectrum. With a ship period of, say, 4 seconds we are still well clear of
resonance.

The ship now gets under way and heads into wind at 20 knots. The
speed relative to wave crests now goes up to 44 knots, and as a result the
encounter period drops to 4^2 seconds. We are now very close to
resonance. If there are waves in the sea spectrum marginally below 100
metres long, or if we increase ship speed to 24 knots, we shall have full
resonance, and the ship will spring continuously as long as these s^as are
beingmet. ..
\
It will be seen that the very large ships, particularly in the loaded
condition when their periods are longest, can quite easily get intd the
resonant condition and experience springing. This form of oscillation, can
persist for hours on end and be a source of crew discomfort and fatigue
failure in stressed parts of the hull.
As a preventive measure against springing, the building of structural
damping or damping mechanisms into the ship is a theoretical possibility.
A system using a resonant tank of water supported on an air-cushion,
analogous in a way to the tank roll-damper, was demonstrated son the
model scale and in a naval destroyer in the early 1970s, but was not
developed further. A simpler and more economical treatment for springing
is the operational one: the ship's officers should simply alter course or
speed or both so as to change the period of wave encounter away from the
resonant value.

92
Diagnosing vibration causes
Naval architects seeking the source of vibrations experienced in a ship will
usually begin by measuring the frequencies present in various parts of the
structure. Predominance of shaft frequencies suggests propeller or shaft
damage, or defects such as manufacturers' errors in blade dimensioning.
Blade passage frequencies suggest wake or pressure effects. Main-engine
crankshaft or cylinder-firing frequencies point to inboard mechanical
origin^. Sometimes all frequencies are present together, making life more
difficult. However, the predominant frequencies usually give pointers as to
where to investigate in more detail.
*

Ship's officers reporting vibrations should do their best to pick out the
predominant frequencies; this Is difficult without specialised instruments,
but sometimes a good indication can be obtained by means as simple as
drawing a pencil line freehand across a piece of paper stuck to deck or
bulkhead over a period of, say, 10 seconds; the squiggles may be clear
enough to count. The record of any such observations should indicate the
propeller shaft RPM at which the records were taken, together with ship's
loading conditions, sea state and other relevant circumstances.

From an the foregoing it will be seen that a ship interacts with the water
that supports her in a most complex and intricate manner. Many experts
have studied the process, and no doubt will continue long to do so,
because none among them would claim that every last subtlety is yet fully
understood, or that ship performance and safety have reached anything
like perfection.

This short book does not claim to have done more than introduce the
general reader, somewhat briefly and superficially, to selected aspects of
the enormous subject of ship hydrodynamics. The book will, however, have
served its purpose if it encourages mariners and those who have business
in ships to look over the side with an enlivened curiosity at the water
rushing past, and to wonder at the remarkable way in which it behaves,
sometimes to our benefit, sometimes to our injury. And perhaps some
readers will be persuaded to seek further enlightenment by plunging into
the more solemn and profound depths of the technical literature.

93
INDEX
Added mass 46 Group velocity 17
Agulhas Current 36 GRP 78
Air lubrication 7 Helicoid 54
Angle of entry- 28 HMS CAPTAIN 82
Angle of vanishing stability- 80 Hovercraft 7, 30
Archimedes of Syracuse 2 Hydrofoil craft 30
Bank effects 39 Hydrofoils 51
Bathythermographs 4 Hydrophobic paint 12
Bernoulli 37 Ice 1
Berthing loads 46 Ideal fluids 70
Bilge keel 76, 87 Inclining experiment 83
Blade-passage frequency 89 Knuckles 78
Boiling 10 Laminar flow 69
Bore 33 Lateral-thru st unit 65
Bottom suction 38 Lift 47
Boundary layer 69 Lines-generating systems 68
Bow Long-chain molecules 7
-and-buttock lines 68 Main hump 18
bulbous 76 Measured mile speed trials 32
cushion 40 Metacentre 82
rudder 65 Metacentric height 82
Broaching 24 Naval replenishment-at-sea 43
Bubble collapse 50 Ogival 53
erosion 58 Overtaking 42
Canal barges 33 Paddle-wheel effect 59
Cat's paws 14 Paradox of D'AIembert 70
Cavitation 11, 49, 90 Pitch 54
Celerity 15 Pitching 86
Centripetal acceleration 63 Pivoting point 66
Clapotis 20 Pneumatic breakwater 35
Clearance 38 Pooping 24
Competitive rowing and sailing events 32 Propellers 53
Corresponding speed 29 controllable pitch or CP 58
Currents 33 damaged 88
Damping 86 raked , 57
tank 87 variable pitch 1 58
Density 1 Properties of water

Design pitch setting 59 Proximity A 37


Diagonals 68 PTFE 5 12
Dispersion effect 34 Races 1 36

Displacement sections 68 Recovery pressure 71


Draught survey 2 Refraction effect "20
Dredged channel 41 Residuary resistance 27, 29, 71
Drogue 24 Resistance
Eddies 72 force 25
Effect of roughness 26 prediction 29
Effective power 25 Resonance 84,86
Erosion 50 Righting lever GZ 80
Fairness 68, 76 Roll s
Fences 51 damper 86
Fins 51 parametric 85
Following sea 23, 84 period of 83
Free surface 85 Rudder 52, 62
Frictional coefficients 26 high-angle 52
FRP 78 Separation of flow 28, 72
GM 82 Shear

94
strain 4 Tugs 42
stress 4 Tunnels 78
Sheer 41 Turbulent
Ship-ship interaction 42 flow 69
Shoals and steering 44 wake 72
Singing 90 Under-keel clearance 38, 44, 66
Skew 56 Unloaded tips 58
Skin Ventilation 51
friction 5 Vibration 88
friction force 27 Virtual mass 45
Sound waves 3 Viscosity 4
Spray Viscous drag 5
generation 8 W. Froude 29
rails 9 Wake
Springing 91 effects 88
Squat 37 frictional 71
Stabilisation 80 shadow 56, 60, 89
Stabilisers 86 Waterlines 68
Stability 80 Wave
directional 65 -energy recoveiy devices 15
Stall \ 49, 52 -induced vibration 91
Steadiness 80 -making resistance 17, 26
Steering 62 capillary 12, 14
Stream lines 70 encounter periods 92
Submarine 3 gravity 13
Suction field 89 in shallow water 20
Surface tension 7 ocean 13
Tactical diameter of translation 21
Tip vortex 57 propagation 15
Trailing vortices 75 steep-fronted 22
Transom sterns 75 tidal 23
Transverse stability 39 Wind against tide 36
Tsunamis 23 Yaw 64

95

You might also like