You are on page 1of 28

Weight and buoyancy

Hydrostatic forces

Archimedes' principle
A ship floating at rest in calm water is acted upon by two forces, weight and
buoyancy. Weight is the downward force on the ship. The total weight force (W)
acts on the ship as if it were concentrated at the balancing point or the centre of
gravity (G). Buoyancy is the upward force of all the hydrostatic pressures on the
hull. The horizontal components of the water pressures on unit areas of the ship’s
sides and bottom, increasing with depth, act in opposite directions and cancel
each other. The vertical components of the water pressures on unit areas combine
to form an upward force (B) equal to the weight of the water displaced by the
underwater hull volume. This weight varies slightly with the specific gravity of the
water. The centre of buoyancy (B) lies at the geometric centre of the immersed
volume. The ship sinks in the water until the force B exactly equals the force W,
in accordance with Archimedes’ principle.
Calculation of ship weight and buoyancy volume

terms used in ship design


In an early stage of the design, the ship weight is estimated as the sum of the
weights of the cargo, hull, fittings, equipment, propelling
and auxiliary machinery, piping systems, electrical and electronic gear, fuel,
water, consumable stores, passengers, and crew, plus a margin of a few percent
for weights that are underestimated. At a later stage the weights are calculated
more precisely or are taken from actual weights of similar items. In many cases,
the weight estimates are revised constantly as the design proceeds to avoid an
ultimate overweight that might detract seriously from the ship’s performance.

The underwater volume of the ship under design must be adequate not only to
displace a weight of water that will support the entire ship, but it must be so
disposed in length, breadth, and height and so shaped in every part that all the
other operating and naval architectural requirements are fulfilled. When the ship
is built and fully laden, it must float level and upright at the designed waterline
(typically indicated by a Plimsoll line).

As the underwater and above-water portions of the hull are fashioned, the naval
architect maintains a running check of the estimated weights and
calculated buoyancy volumes, as well as the products of these weights and
volumes times the horizontal fore-and-aft distances or “moment arms” of each
from the transverse vertical reference plane at mid-length. These products are
known as the longitudinal weight and buoyancy moments.

naval architecture
To carry out these operations systematically, the underwater hull is divided into
segments by imaginary transverse planes called stations. There may be 10 such
segments for a boat, 40 or more for a large ship. The volume of each segment is
computed together with the position of the centre of volume for each. The forward
and after moments of volume are then computed in the same way as for the fore-
and-aft moments of weight. A summation of the individual segment volumes
gives the total underwater hull volume. The fore-and-aft positions of the centres
of gravity of the individual weight groups are then estimated. Separate sums are
kept of the moments of these groups forward of and behind the mid-length.
Dividing the total underwater hull volume by the volume per unit weight of the
fresh, brackish, or salt water in which the ship is to run gives the weight of water
displaced. This must equal the total weight if the ship is to float at the designed
waterline. The net weight moment, forward of or abaft mid-length, is divided by
the total weight to give the distance at which the centre of gravity (G) lies forward
of or abaft the mid-length. The same operation for the volume moments gives the
fore-and-aft position of the centre of buoyancy (B).
Achieving level attitude or trim
Buoyancy, gravity, density, and water displacement explained
See all videos for this article
For the ship to float at the level attitude or zero trim desired, G and B must lie in
the same vertical transverse plane. If their calculated positions are different, and
the size, proportions, and shape of the underwater hull are satisfactory, it is
customary to shift the weights within the hull until the desired trim is attained.

In practice, the record of estimated weights and fore-and-aft moments is


accompanied by a record of vertical moments above the keel (K) or the base
plane. From this it is possible to estimate the position of G above K. At the same
time, a record is made of vertical moments of buoyancy. When summed up and
divided by the volume, these give the position of B above the keel. Both the
distances KG and KB are required for estimating the metacentric stability.

If, when the ship is built, the actual weights and volumes, or their centres, do not
agree exactly with the estimated values (some equipment may have been added
during the construction period), the ship floats at a waterline slightly different
from that contemplated by the operator and designer. For a surface ship this
difference is usually of no great importance. However, for
a submarine, W and B must equal each other exactly. It is also important to
ensure that, when submerged, the centres G and B are in the same transverse
plane, so that the craft floats level when stopped underwater.

The weights and weight moments for a submarine are estimated and calculated
exactly as for a surface ship, but two separate volumes must be calculated, one for
the surface condition, with main-ballast tanks empty, and one for the submerged
condition, involving principally the volume of the pressure proof hull. To the
volume of the latter there must be added the water-excluding volumes of all parts
external to it. Among these are the outer hull structure, shafting, propellers,
rudders and diving planes, anchors and chains, masts and periscopes, and the
great multitude of external items. For every seven tons of solid steel in this
category, about one ton of buoyancy force is gained.

Metacentric stability
Concept of the metacentre

buoyancy in ships
One would think, at first sight, that the average surface ship, with its weight
concentrated above its point of support (considered as the centre of buoyancy),
would fall over like a top that has stopped spinning. If properly designed it does
not do so, because as the ship is inclined transversely, say by a strong wind, the
centre of buoyancy of the immersed portion of the hull shifts sideways, in the
same direction as the inclination. This is because the volume of the wedge of
emersion shifts to the low side. At the new inclined waterline, the centre of
buoyancy moves to the low side as well. The total buoyancy force continues to act
upward along the true vertical. Provided no weights shift as the ship inclines,
the centre of gravity remains at G, in the keel-to-deck centreline. The buoyancy
force acting upward and the weight force acting downward through G produce a
righting moment which acts to return the ship to its upright position.
static stability of a ship
For small angles of inclination, say less than 10°, the verticals through the shifted
centres of buoyancy intersect the ship centreline at or close to a point M, called
the metacentre. The stability provided by the action described is therefore called
metacentric stability. For the situation most often encountered in practice, it is
transverse metacentric stability. For a similar situation in which the ship is
inclined or trimmed in its fore-and-aft centre plane, it is longitudinal metacentric
stability.

If the centre of gravity of the ship is too high, the righting moment for any
inclination is negative; that is, it acts to incline the ship still further. The ship then
has transverse metacentric instability. Whether it will capsize or not depends
upon whether the revised position G is overtaken by the vertical through the
revised B as the inclination increases. If so, the ship remains in
that inclined position, with a righting moment that is practically zero.
Indexes of metacentric stability
The distance GM for positive stability, known as the metacentric height, is taken
as one index of the degree of metacentric stability. The other is the range of
stability, or the angle of inclination at which the metacentric height diminishes to
zero. For example, when a ship is heeled transversely until the depressed deck
edge goes underwater, the centre of buoyancy cannot move to the inclined side
far enough to keep the metacentre well above the centre of gravity on the ship
centre plane. At a critical inclination the metacentre lies at the centre of gravity
and the righting moment disappears. For inclinations beyond this the
metacentric height becomes negative, the righting moment becomes a capsizing
moment and the ship rolls over.

A greater range of metacentric stability can be built into a ship by raising the
uppermost watertight deck to a higher position above the calm-water plane of
flotation. The ship can then heel to a greater angle before water comes over the
lower deck edge. For a craft like a sailing yacht, with a deep, heavy, stabilizing
keel, the deck edge can actually go underwater and the range of positive stability
can extend to large angles, provided water is kept out of the hull as the water level
climbs higher and higher over the inclined deck.

The desirable value of transverse GM varies with the type, size, and service of the
ship, but within limits it is still subject to the experience and judgment of the
naval architect. It averages from 0.04 to 0.06 of the beam but may be as high as
0.10 of the beam for combatant vessels subject to heavy damage. For fishing
vessels it may have two values, one for the outgoing and one for the return or
loaded part of the voyage. The range of positive transverse GM for a normal ship
to run in the open sea is usually in excess of 40° and may run as high as 70° or
more, provided the hull remains intact and the weights do not shift.
Vertical position of metacentre
The height of the metacentre above the keel, or other selected point, depends
upon the shape and size of the underwater body and of the at-rest waterline. The
total height KM is the sum of the height KB of the centre of buoyancy above the
keel and the height BM of the metacentre above the centre of buoyancy. The latter
is known as the metacentric radius. The distance KB can be estimated by an
approximate formula, it can be calculated by procedures applying to irregular
volumes, or it can be determined by a mechanical integrator. The metacentric
radius BM depends entirely on the geometry of the underwater hull and can be
calculated from the formula, BM = I/V where I is the transverse moment of
inertia of the waterline plane about the centreline axis, and V is the immersed
volume of the hull. This relationship was first derived by Pierre Bouguer in the
18th century. For a given ship length and underwater volume, both I and BM are
proportional to the cube of the beam, so that the latter is an important factor in
transverse metacentric stability.
Vertical position of centre of gravity
The vertical position of the centre of gravity is as important as that of the
metacentre in determining metacentric stability and the behaviour of the ship at
sea. For a cargo ship, or even for a warship, this position may change by many
feet, depending upon the nature, amount, and vertical position of the loads, fuel,
and stores. Ballast may be used to increase the total mass moving in a seaway or
to place G in a more advantageous position. Ice accumulating on the upper works
of fishing boats, trawlers, and other small craft may be so thick, so heavy, and so
high above the normal centre of gravity, as well as so difficult to remove at sea,
that G rises above M and the craft capsizes.
Inclining experiment
Fortunately, the naval architect is able to make a full-scale check of the predicted
or calculated metacentric stability before the completed ship goes to sea. By
shifting liquids or solid masses whose weights and offset positions are known
accurately, the centre of gravity of the whole ship is shifted and the ship is heeled.
This shift is sideways for a determination of transverse and lengthwise for a
measurement of longitudinal. The angle of inclination of the ship for each such
shift is measured accurately with special devices. Then the actual metacentric
height is determined for that loading condition. The height of the centre of gravity
above the keel is then calculated from the relation.
Stability of submarines

submarine
When a submarine is in surface condition, its metacentric stability situation is the
same as that of any surface ship. When diving or submerging, its water-plane area
diminishes progressively to a negligible amount. The transverse moment of
inertia of this plane likewise diminishes and with it the value of BM. Moreover,
as the main-ballast tanks are flooded and the pressure hull is taken under,
buoyancy is transferred from a low to a high position, so B rises in relation to the
hull. Since G remains in its original position, M drops to the vicinity of G. Indeed,
for some moments during diving and surfacing, it may be below G, with a
negative GM. The flooding, venting, and blowing of the main-ballast tanks, with
separate port and starboard controls, enable the submarine crew to counteract
any list that may develop during this brief transition period.

When the submarine is under the surface, the water-plane area is zero, as is BM.
This means that the centre of buoyancy of the pressure-proof hull (B1) and M
coincide. In this condition B1 is higher than G, so that the action of the buoyancy
force and the weight force produces a stable situation, and the craft is said to have
pendulum stability.

General arrangement
Despite the many ships of each type that have been designed over the years and
the general similarity of various spaces and their locations within the
types, ship operators still find advantages in particular arrangements. This
situation reveals the variety of combinations possible when the designer
endeavours to make large-scale compromises with both major and minor
features. Propelling machinery at the stern with crew accommodations and
navigating spaces in one group aft over the machinery represents efforts to devote
the most useful spaces to the cargo and to concentrate services and living spaces
in a region clear of cargo-stowing and cargo-handling areas. Naval architectural
requirements impose limitations concerning weight distribution, metacentric
stability, hull strength and stiffness, and subdivision and damage control which
can rarely be disregarded.
General arrangement features by ship type
A brief tabulation of principal ship types serves to highlight the arrangement
features characteristic of each.

Nassau, Bahamas

Titanic: first-class lounge


Passenger liners for ocean crossings, carrying only passengers, baggage, and
incidental cargo, devote large volumes in the most comfortable part of the ship to
passenger accommodations, with large additional volumes for public spaces in
deckhouses and superstructures. The propelling machinery, uptakes, and hatches
are placed clear of the accommodations. Passenger ships for service on rivers and
in protected waters utilize deck and superstructure volume as passenger spaces
for practically the entire length. Excursion ships for day service extend the
accommodations to overhangs beyond the main hull.

Combined passenger and cargo ships devote the most comfortable positions to
the passengers without encroaching unduly on storage and handling facilities for
cargo.
General dry-cargo ships with machinery amidships have not always allocated the
best available spaces and facilities for the cargo hatches and holds. The propelling
machinery is preferably aft, to keep the best cargo spaces clear, an arrangement
becoming increasingly popular. Means are provided to trim the ship with liquids
in ballast tanks.

ton: Colombo Express container ship


Container ships, roll-on-roll-off ships, sea trains, barge carriers, and car ferries
embody special arrangements of structure, machinery, and crew spaces to keep
them clear of the spaces for large containers, wheeled vehicles, or barges.

Bulk-cargo carriers, for solids or liquids or both, are the ultimate in large single-
purpose ships, with everything possible sacrificed to cargo capacity.

USS Carl Vinson


Aircraft carriers have flight decks of the greatest practicable area, even to the
extent of using overhangs beyond the main hull. High hangars under the flight
deck provide storage and repair space for aircraft. Internal and deck-edge
elevators move these craft to and from the flight deck.

submarine
Submarines are of the double-hull type, with a ship-shaped outer hull of relatively
light construction, if their mission calls for high speed and good seakeeping
qualities on the surface. If, however, submerged performance is the primary
function of the craft, as in the case of modern nuclear-powered submarines, they
have single hulls of suitable shape. The volume between the heavy inner and light
outer hulls of a double-hull craft is devoted to carrying fuel and ballast liquids
which need not be protected from hydrostatic pressure.
Cargo handling

tanker
The ship arrangement must lend itself to getting the cargo in and out as well as to
carrying it from one port to another. Indeed, speed in loading and unloading
cargo is just as important as speed through the water. Access to the holds and to
the internal deck spaces is provided by hatches through the decks and in some
cases by doors in the ship’s side leading to the deck storage areas.

Ships carrying dry general cargo are sometimes equipped with their own handling
gear. This enables them to transfer cargo in any port and to load to and from
lighters in places where they must anchor offshore. Containers may be loaded
and discharged from special-purpose container ships through oversize hatches
by either shipboard or shore-based cranes. Some bulk-cargo ships carry a huge
swinging boom with a belt conveyer running on it, by which material may be
dumped in high piles at a distance from the ship’s side. Freight cars are loaded
and unloaded from sea trains by special dock cranes that pick up an entire loaded
car. Liquid cargo is pumped aboard through flexible pipes from storage tanks on
shore; the unloading is invariably done by the ship’s own high-capacity pumps.
Watertight closures
Whatever the mission of a craft, or the arrangement of major and minor features
adopted, water must definitely be excluded from the hull under severe operating
conditions. This calls for strong, tight closures for openings, including doors, port
covers, and protectors for glass windows. It also requires the watertight and
wave-resistant sealing of large openings such as cargo hatches. On many ships
these openings are closed by heavy metal covers handled by mechanical power
and capable of secure sealing and locking. The structure surrounding these
openings must be so rigid that its deformation under wave or sea loads or other
service conditions does not jeopardize the water-tightness of the cover.
Ballast tanks
When a ship is running rather light with its hull relatively high out of water, it is
at a disadvantage in winds and waves. It needs added inertia to help it drive
through waves, added weight to put the hull farther down in the water, and more
mass high in the ship to reduce the righting moment and to ease the rolling. These
needs are met by building in tanks that can be filled with fresh water or reserve
fuel. The tanks are easily emptied when the weight is no longer desired. Awkward
and inaccessible places in the hull, where neither cargo, machinery, nor useful
load can be placed to advantage, can often be used for these tanks.

All submarines, whether they have two separate hulls or not, carry main-ballast
tanks. These are empty when the craft is on the surface; they help to lift the bridge,
the deck, and the hatches above the water and to provide reserve-buoyancy
volume when rolling and pitching in waves. By opening flood valves at the bottom
and air-vent valves at the top, these tanks may be completely flooded with
seawater to make the craft submerge. To raise the submarine, the vent valves are
closed, and the water is blown out by compressed air. Another set of tanks, called
the variable-ballast tanks, have water taken into or pumped out of them from time
to time to keep the weight of the submarine always equal to the weight of the
water displaced by the buoyant volume. When a submarine runs from salt
water into brackish water having less weight per unit volume, some water must
be pumped out of the variable-ballast tanks because the supporting forces are less
in the lighter water.

Resistance and propulsion

nuclear submarine
The resistance to forward motion of a ship is of three principal kinds: friction;
wave making; and separation or eddy making. Friction or viscous resistance is
caused by the acceleration of liquid particles in a forward direction as the bow
continually runs into a region of liquid at rest. The layer of accelerated particles,
augmented by vortex motion and turbulence, becomes progressively thicker as it
moves aft, forming what is known as the boundary layer. The vortexes and
disturbances in this layer are visible in the belt of “confused” water around a
moving ship at the waterline. The energy in this layer represents the work done
by the ship in overcoming viscous resistance. It is eventually dissipated as heat
and is not recovered.
turbulent flow
Wave-making resistance is caused by transferring kinetic energy in the ship to
energy in the surface or gravity wave system which accompanies it. While
the configuration of this system near the ship remains fixed for a given speed,
waves are continually left astern and the energy in them is lost. Consequently, the
large wave pressure buildup over the forward part of the ship is only partially
balanced by the buildup aft.

Separation is caused by the lack of sufficient pressure in the water in a given


region to force this water laterally inward and to make it flow closely along all
parts of the ship, especially in the tapering or blunt after portion. In the region
known as the separation zone, water is dragged in from astern to fill the gap that
would be left because the flow does not close in from the sides. Resistance is
generated by the forward acceleration of water that would otherwise flow aft and
be left behind. The confused and eddying mass of water being dragged along in
the separation zone behind the square transom stern of a motorboat is clearly
visible at low and moderate speeds. The added drag due to separation behind the
square stern of a skiff, immersed deeply by passengers sitting in the stern, is very
real to the rower in that skiff.
Computing friction resistance
The friction resistance of a ship can be computed from a knowledge of its wetted
area and a friction value per unit area derived from the towing of flat planks or
friction planes of various lengths at various speeds. By using very thin sections,
sharply pointed at the ends, wave making and eddy making are eliminated. From
the known towing forces and wetted area of the plank or plane there are derived
a set of friction values per unit surface area of the plane, in terms of the towing
speed. For calculating the friction resistance of a ship at any given speed, it is
usually assumed that the friction value for each unit of wetted-surface area is
equal to that for a friction plane having the same length as the ship and towed at
ship speed. The wetted area of the ship is calculated by averaging the girth at a
series of stations equally spaced along the length and multiplying by the wetted
length. The flat-plate friction data cannot be applied indiscriminately to the
curved surfaces of ships.

Rough areas on wetted ship surfaces are caused by uneven plating and planking;
laps, butts, rivet points and weld beads; anticorrosive and antifouling coatings of
plastic paint and other materials; and fouling due to marine organisms. All of
them increase friction resistance and the thickness of the boundary layer. For
resistance calculations their effects are lumped in a general roughness allowance,
which is added to the value of the friction for a given area of smooth surface.
Wave-making resistance
Information available to the naval architect on the surface waves generated by a
moving ship is derived originally from the observations of John Scott Russell in
the 1840s, the experimental work of William Froude and Robert Edmund Froude
in the 1870s and 1880s, and the analytic studies of Lord Kelvin in the latter
decade. These showed that: (1) A gravity wave system is formed by a moving
pressure disturbance. For example, drawing one’s finger across a water surface
makes waves. (2) Pressure disturbances exist where there are changes in
curvature around a ship, such as those at the extreme bow and stern and at the
“shoulders.” (3) The progressive or traveling wave system caused by each
pressure disturbance consists of two parts: (a) a diverging group of waves,
with crests and troughs lying at a small angle to the direction of motion of the
disturbance, and (b) a transverse group of waves, with crest lines slightly convex
forward, where they cross the path of the moving disturbance. The diverging
waves at the bow are easily seen on any moving boat or ship, as are the transverse
waves abaft the stern on any craft which is traveling rapidly. The transverse waves
of the bow system, modified by the forward shoulder system, are also indicated
by the crests and troughs in the wave profile alongside the ship.

In addition to the progressive waves, whose shape remains the same for a given
speed but which spread outward and aft, there is a water-level disturbance that
moves along with the ship and whose elevations at the bow and stern and
depression amidships are not radiated as gravity waves. There may thus be six or
eight or more sets of water-level changes generated by the movement of one ship.
The changes of elevation due to each are superposed so that two crests coinciding
produce a sort of double crest, while a crest and a trough coinciding act to cancel
each other.

From a resistance standpoint, the most important progressive wave systems are
generated at the bow and stern. The length of a gravity wave depends upon its
velocity, and the velocity of a wave whose crest travels along with the bow must
correspond to the ship speed. It follows, therefore, that the second, third, and
succeeding crests of the transverse bow series move aft along the ship as the speed
increases. This means that, at certain ship speeds, a transverse crest of the bow
system is superposed on the stern system in such manner as to build up a
traveling mound of water at the stern. The internal hydrostatic pressure in this
mound acts to push the ship forward and hence to diminish its wave-making
resistance.

At other ship speeds, the superposition of the bow and stern wave systems drops
the water level at the stern, with no compensation for the hydrostatic pressure
which the bow of the ship must push against at this speed. As a result, the total
resistance of a ship fluctuates above and below what is known as its "natural"
resistance as the speed is increased and as the various progressive wave systems
combine to produce beneficial or harmful effects.
David Watson Taylor
The velocity of gravity waves varies as the square root of the product of the
acceleration of gravity and the wavelength. The forward speeds of the transverse
waves generated by a ship correspond to the ship speed V. The interference effects
described depend upon a relation between the wavelengths LW and the ship
length L. Hence, the wave systems are geometrically similar if the ratio of V to the
product Square root of√gL remains constant, where g is the acceleration of
gravity. This ratio is the Froude number = V/Square root of√gL.

David Watson Taylor simplified this relation in the 1900s to the ratio of the ship
speed V in knots to the square root of the ship length L in feet. Thus, the speed-
length quotient = Taylor quotient Tq = V/Square root of√L.

USS Cole; destroyer


When the estimated wave-making resistance is plotted on a basis of Froude
number or Taylor quotient, humps and hollows show up in the curves. The naval
architect selects a ship length whose wave-making resistance will be less than its
“natural” resistance when the ship is traveling at its most efficient speed. The
extreme case in this category occurs with the destroyer-like craft which, at a
speed-length quotient of about 2.0 or a Froude number of about 0.6, rides largely
on the back of its own first bow-wave crest with its stern in the first trough
following. It is, in fact, constantly running uphill; part of its resistance, called the
slope drag, is due to this action. A planing boat such as a speedboat is in a
corresponding position, with bow high in the air and stern squatting deeply, when
about to pass through what is known as its hump speed. As this speed is reached
and exceeded, if the engine has ample power and the boat is not too heavy, the
boat approaches and reaches full planing speed. Here it is literally riding on top
of the first crest of its own bow-wave system. With its flat stern sliding gracefully
over the water there is, in effect, no stern-wave system.

Separation resistance
The drag due to separation of the boundary layer from a ship surface, and to
eddying and backwash in the separation zone, is a form of pressure resistance.
This means that, like wave-making resistance and some types of roughness
resistance, it is due to forces exerted at right angles to the hull surface. Like these
resistances, it varies as a power of the ship speed.

Hydrodynamic knowledge of separation phenomena and the physical laws which


govern them has not progressed to the point where the onset of separation can be
predicted in advance with certainty and where the magnitude of separation
resistance can be calculated. It is known, however, that the pressure in such a
zone is less than atmospheric, so that the water literally sucks backward on the
ship. If air can be led to the zone to displace the eddying water, the suction is
removed. When a motorboat with a square or transom stern extending below the
water is speeded up until the stern “clears,” the backwash and eddying disappear.
With the square stern exposed to the atmosphere, the separation resistance also
disappears.
Resistance of submarines
When a submarine submerges to a depth below the surface equal to four or more
times its maximum diameter or its hull depth, the surface disturbance resulting
from its forward motion becomes negligible and its wave-making resistance
practically disappears. This is a great advantage, especially at high speed, despite
the increase in wetted surface and friction resistance caused by taking the whole
craft under the water. However, old-style submersible craft have considerable
amounts of gear above the water line, such as flat decks, rails, anchors, capstans,
chocks, and similar fittings, put there for operation on the surface. It is difficult
to streamline them for low resistance under water. Modern true submarines,
intended to spend almost all of their operational time fully submerged
under nuclear power, have dispensed with most of these irregularities.
Furthermore, their length-to-diameter ratios have been reduced so that frictional
resistance at high speeds is minimized.
Resistance in shallow and restricted waters
The forces on a ship traversing shallow waters are governed by the presence of
solitary waves caused by ship motion and other disturbances. If the ship speed is
slightly less than the solitary wave speed, the ship runs uphill on the back of this
wave so that its hydrodynamic resistance is increased by the slope drag. If it can
be speeded up so as to run slightly faster than the wave, it slides downhill on the
face of the wave and its resistance is reduced below that of its deepwater
resistance. The speed of progressive waves of a given length is less in shallow than
in deep water. If a tug, for example, is running at a speed in shallow water at which
it has a crest at the bow and another at the stern, its speed must be decreased if
the two crests are to be kept at the advantageous positions indicated. At the same
time, the crests may be higher and the trough may be lower because waves
become steeper as they enter shallow water. A fast craft also squats more deeply
at the stern when running in shallow water. In fact, this increase in squat may be
sufficient to cause the craft to scrape bottom even though it has plenty of water
under it when at rest.

When the clearance between the bottom of the ship and the bed of the water body
is initially small, the water that flows under the ship is speeded up, with an
increase in friction resistance on the ship. When the sides or walls of the channel
are close to the ship, the lateral constriction speeds up this flow still further.
Methods of approximating the increased resistance and the depth of water
necessary to give the equivalent of deepwater resistance are available.

Self-propelled craft designed for efficient operation in shallow and restricted


waters must have: (1) provision for adequate flow of water to the propellers; (2)
adequate shielding to prevent drawing air from the surface; and (3) rudders of
extra-large area, usually one rudder behind each propeller, to overcome the
horizontal forces resulting from the closeness of adjacent banks or of other craft
being met in a channel.
Ship form for minimum resistance

boat
Certain general rules for ship form based upon hydrodynamics are available: (1)
The use of easy and fair surfaces along the general paths followed by the water
flow. Small changes of curvature in the flow lines are particularly important. (2)
At and near the surface the flow lines must follow the surface or the wave profile.
Since most of the wave-making resistance is generated by pressure disturbances
near the surface, easy curvature is important there. Proof of good design in this
respect is low wave crests and shallow troughs around the ship when running. (3)
Most of the flow in almost any type of ship goes under the bottom rather than
around the sides, hence the ship form must not interfere with it. (4) Submerged
bulbs intended to produce surface-wave systems that will partly neutralize the
crests and troughs produced by pressure disturbances elsewhere require careful
design and positioning. (5) Probably the most important feature in shaping the
hull of a self-propelled craft is to provide a good flow of water to the propulsion
devices. So far as known, this calls for the highest practicable degree of uniformity
of relative velocity over the whole thrust-producing area, the greatest possible
degree of flow opposite to the direction of advance of the blades of the propulsion
device, and the greatest mass density of the water in which the device is to work.
Concerning the last item, it is known that the water entering the propeller disks
of destroyers and other high-speed craft contains many air and gas bubbles. In
the aggregate, the reduction of mass density due to them can be appreciable.

It is important to note, however, that the optimum ship design for a given mission
may not be the one that has the form of minimum resistance. In the overall
economic picture, gains from better stowage of cargo, for example, may outweigh
a modest increase in fuel consumption. Developable hull forms have been
designed for small craft that are cheaper to build and offer little, if any, resistance
penalty.

Action of propulsion devices


Thrust by a ship propulsion device acting on the water (or on the air) is produced
by imparting sternward acceleration to a mass of that water or air. The forward
thrust is proportional to the product of the mass of fluid acted upon and the
accelerating rate. For the most efficient propulsion, the mass should be large and
the acceleration small. In a screw propeller, this calls for a large diameter and a
small increase in relative backward velocity when water is passing through the
propeller.

The thrust per blade of a propulsion device is measured by the reduction in


pressure on the back or advancing side of the blade and the increase in pressure
on the face or after side. As a rule, the former is much larger than the latter so
that the blade draws or pulls rather than pushes itself through the fluid in which
it works.

Follow the Delta Queen down the Mississippi River and learn how steam power advanced
naval architecture
See all videos for this article
Modern propulsion methods for boats and ships include oars, sails, paddle tracks,
paddle wheels, hydraulic and pump jets, airscrews, rotating-blade propellers, and
screw propellers. Screws are usually run in the open, but for producing high
thrusts at low ship speeds, as when towing, they may be surrounded by a
fixed shrouding such as the Kort nozzle. Vertical axis propellers with adjustable
blades offer the great advantage that the magnitude and direction of thrust can
be varied at will, making them vastly more versatile than any known combination
of screw propeller and rudder, and giving the craft exceptional maneuverability.
A tug fitted with one or more such propellers can exert a pull equally well in any
direction. The number of propulsion devices depends upon the available power
in each engine, the need for reliability or maneuverability, the limiting draft and
many other factors.
Interactions between propeller and ship
The operation of a screw propeller involves a number of interactions that are by
no means fully understood. Part of the water through which the propeller moves
is the boundary layer moving aft past the hull, with a relative velocity less than
the ship velocity. Another part of it lies within the wave crest (or trough) that runs
along above the propeller. Because of these and other effects, the water moves at
different velocities and in different directions in different parts of the propeller
disk. In general, the ship drags the water along with it to a certain extent, so that
its average speed Va past the propeller is less than the ship speed V. The
difference V−Va is the wake velocity, and the ratio of this velocity to the ship speed
is the wake fraction w = V−Va/V.

There are reduced pressures in the region forward of the propeller, resulting from
corresponding pressures on the forward sides of the blades. These act to increase
the hull resistance R and require a greater thrust T to overcome it. This resistance
augment or loss of thrust is T−R and the thrust deduction fraction t = T−R/T.
Efficiency of propulsion
The efficiency with which any mechanical propulsion device drives a ship is a
product of three separate ratios. The first is propeller efficiency, the ratio of input
to output when the device is running in open water by itself, as when a model is
tested in a model basin. The second, known as the hull efficiency, is the
ratio 1−t/1−w, indicating the average effect of hull-propeller interaction. The
third, known as the relative rotative efficiency, is the ratio of the propeller
efficiency when the device is running in open water to the propeller efficiency
operating in the irregular ship’s wake.

For ships having screw propellers, the efficiency of propulsion decreases as more
propellers are added. It varies from 0.76 to 0.80 or more for a well-placed and
well-designed single screw, from 0.65 to 0.72 for twin screws, and from 0.60 to
0.64 for quadruple screws such as are carried by large liners and warships.

In practice, the open-water efficiency for a given size of propulsion device is found
to vary in almost predictable fashion with the thrust loading coefficient, T/Va2.
Starting with this factor, it is possible to estimate the shaft power required to
drive a ship having a known resistance at any given speed.

Cavitation
cavitation
Any moving submerged body, like a screw-propeller blade, has to push the water
aside as it moves. If it moves so fast that the surrounding pressure is not sufficient
to cause the water which has been pushed aside to close in around the body and
follow its contours, or if the pressure is so low that the same thing occurs when
the blade moves slowly, the water either “opens up” or it leaves the blade. In the
first case, bubbles are formed in it, each filled with water vapour. When they move
along into a region of increasing pressure, they collapse suddenly. The resulting
severe pressure fluctuations may cause pieces of the metallic blade surface to
break off in an action known as erosion. In the second case, a relatively large
vapour-filled cavity is formed next to the blade. This may collapse on the blade or
at a distance behind it.

For screw propellers of normal form, any cavity next to the blade interferes with
proper flow around it and usually has a harmful effect on thrust and propulsion.
Cavitation can be minimized by proper attention to the design of the propeller.
The shape selected for the section should be one known to be relatively free from
cavitation and one on which the reduced pressure is as uniform as possible along
the chord (length) of the section, from leading to trailing edge.

At each radius the blade is made wide enough to carry the local thrust load at the
velocity of and at the average water pressure for that radius. The use of large blade
areas to delay cavitation must be balanced against the loss of efficiency caused by
greater friction drag on the wider blades.

On supercavitating propellers of special design, the blades travel so fast that the
water pressure is never sufficient to permit the flow to follow the blade. The
vapour cavity is then allowed to expand until it covers the whole back of the blade.
The pressure on the back approaches absolute zero while the friction on that side
disappears, since the water no longer touches it. Propeller blades of this type, with
sharp leading edges and blunt or square trailing edges, have been used
successfully on racing motorboats since the 1920s.
General design and positioning of propellers
Ship propeller.
The propulsion device should be treated as an essential part of the ship, not as a
sort of appendage to the hull, and should be designed with it. The flow to and
from the propulsion device, whatever its form, is a most important feature from
the standpoint of efficient propulsion as well as avoidance of objectionable
vibration. Fortunately, it is possible to "see" this flow on medium and large
models in circulating-water channels, to study it at length and to correct
unsatisfactory features of it while the ship is still in the design stage. Model
techniques are also available to give the designer a reasonably good preliminary
warning of excessively large periodic forces which may be generated on the ship
if corrective measures are not taken.

Because of the great thrust sometimes exerted by the single blades of powerful
propulsion devices and the rapid changes of pressure and velocity which take
place near them, adequate clearance spaces must be allowed between these
blades and the adjacent parts of the ship. Propulsion devices mounted in
transverse ducts or tunnels, extending through the thin ends of the ship from one
side to the other, apply transverse forces or swinging moments when the ship is
moving or stationary. These thrusters are usually installed at the bow where they
greatly improve the ship’s handling qualities around docks and piers. On shallow-
draft vessels, screw propellers are fitted inside fore-and-aft arch-shaped recesses
called tunnels. A large proportion of the propeller area is often above the at-rest
waterline, but if air is excluded from entering, the tunnel fills with water when
the propeller starts rotating, permitting the latter to develop thrust over its entire
area.
In many cases it is possible to select the principal features and proportions of a
screw propeller by the use of one or more of the many sets of series charts based
upon test results of systematic series of propeller models. The disadvantage of
this method is that the designer is restricted to the number of blades, blade
profiles, and blade-section shapes of the models that have been tested. However,
there are usually two or three sets of models which approximate what the
designer has in mind so that with the data from these some combination of
tentative characteristics can be rather well bracketed. If the designer feels that
cavitation may be encountered, a propeller may be designed from first principles
on the basis of circulation theory and published airfoil data.
Model experiments
The towing of ship models to determine their resistance and similar
characteristics was initiated in 1872 by William Froude to take the place of
limited knowledge of physical laws governing ship behaviour, complexity of the
interactions encountered, and lack of understanding of the effects of changes in
shape and proportions. To make the procedure workable at all, Froude had to
separate the friction resistance from the total observed resistance. After
subtracting the friction resistance, estimated on the basis of tests which he made
by towing flat planks, Froude called what was left the residuary resistance. For
corresponding ship and model speeds, where the Froude numbers V/(gL)0.5 (V is
the speed, g the acceleration of gravity, and L the waterline length) were the
same, he extrapolated the residuary resistance on the basis that this resistance
per ton of displacement was the same for both ship and model. The calculated
ship friction and expanded residuary resistances were then added to give the total
ship resistance.

In later years, techniques were developed for the testing of propellers, for self-
propelling ship models, for determining lines of flow and wave profiles, and for
measuring the effects of minute changes upon the total resistance. Nevertheless,
many of the old problems remain, despite all the time, thought, and effort devoted
to their solution. Indeed, it appears that advances in knowledge in the field of
hydrodynamics raise new problems faster than the old ones are solved. In spite of
this, the model-test procedure has been of great assistance to naval architects
and, in general, of great engineering value. All the maritime countries of the
world have ship-model testing establishments, and their staffs compare
techniques at the International Towing Tank Conference held regularly every
three years. Very few large and important ships are built without first testing one
or more models of them.

Maneuverability
St. Lawrence Seaway
All self-propelled craft, of whatever size, shape, form, or type, are required to
steer a reasonably straight course in both smooth and rough water, to turn so as
to change course or heading or to take emergency evasive action, to start, stop
and back, and to perform any other desired maneuvers. Submarines are required
to maneuver similarly in a vertical plane, including the operations of diving,
depth keeping, hovering in one spot, and surfacing.
Dynamic stability of route
The ease and reliability of steering of a ship depend, among other things, upon
whether or not it has dynamic stability of route. A self-propelled ship that is
stable in this sense will, if left to itself with no rudder angle applied, continue
generally on its original course. If disturbed by some external force, it may swing
slightly or moderately to a new course, whereupon it will continue along that
course or route until again disturbed. Most slender ships like destroyers are
dynamically stable in route. Others of fat or chubby shape, if left to themselves
and then disturbed, will swing farther and farther from the original route. A sign
of route instability is the persistence of the ship in swinging one way after
moderate corrective or opposite rudder is applied to stop the swing. A ship of this
type may become positively unmanageable in shallow water, where the
sluggishness of any ship is intensified.
Steering and turning
Steering involves corrections to bring a ship back to a given course or heading
after it has deviated as a result of some disturbance. Steering by hand control is
easier and more efficient if instruments in front of the steersman show almost
instantly when these deviations begin. Gyrocompasses are far more satisfactory
than magnetic compasses for this purpose. Further, a ship that is dynamically
stable in route, but not too much so, and one that is not oversteered, requires only
a small rudder angle and relatively infrequent use of the rudder. Automatic
steering by gyro pilot is available for all sizes and types of ships.

Turning is involved when changing course; when maneuvering in formation with


other ships, and when following a curved channel. However, the most important
turning maneuver for any ship is to sheer off suddenly and to get clear of its
original course when danger is unexpectedly sighted ahead along that course. To
clear the extension of its original path in the shortest distance and the least time,
assuming that the ship is going too fast to be stopped completely, requires rapid
laying of the rudder to the emergency angle, rapid response of the ship in starting
to turn, and rapid motion of the ship to the right or left of the course to clear the
danger ahead.

The rudder action serves not only to swing the ship in the desired direction but
also to keep its bow pointed inside the path of its centre of gravity so that a
turning moment is generated. The inward-acting hydrodynamic force on the hull
equals the outward-acting centrifugal force resulting from motion in a curved
path. The amount by which the ship heads inside the instantaneous direction of
motion is the drift angle.

In the course of turning, especially with a large drift angle, the increased hull
resistance causes the ship to slow down, sometimes involving a reduction of 40
percent or more of the speed with which it approached the turn. After the average
ship has turned at least 90°, conditions become steady and its centre of gravity
moves at uniform speed in a circular path.

In the steady-state portion of the turn the inward force caused by the drift angle
exactly balances, in both magnitude and moment about the centre of gravity, the
outward rudder force and the centrifugal force at the centre of gravity caused by
the turning action. If the wind and sea were entirely quiet, the ship would
continue to turn in a steady circle as long as the rudder was held at a constant
angle and the speed remained constant.

Ability to steer a straight course or to turn readily is achieved in any given ship
design by the use of a large rudder area. When the rudder is at zero angle, it serves
as a vertical stabilizing fin. When angled, the large area provides the large
swinging moment necessary for good turning.
Stopping and reversing
Stopping in an emergency, as contrasted with normal coasting and gradual
retardation, is achieved by slowing the propulsion device to less than driving
speed and then by reversing its direction of thrust. If reversed too rapidly, it
is liable to overload the engine, to draw air down from the surface to the propeller
in large quantities and to churn the air-water mixture into excessive turbulence
without developing the maximum astern thrust. Capacity to start and stop quickly
is built into a craft by providing an engine that will reverse readily and by using a
propulsion device with a large thrust-producing area. Both these features are
stressed in the design of tugs.

Rudders and planes


rudder
Rudders and other control surfaces are usually placed at the stern of a ship for
several reasons. When placed behind screw propellers, they benefit from the
increased velocity in the propeller outflow jet or race. If the rudder is attached to
the bow, it is ineffective hydrodynamically in producing a swinging moment. Such
positioning causes the ship to turn with a smaller drift angle and hence a larger
turning radius. In fact, a normal ship when moving backward steers only
indifferently or not at all. The rudder also receives better mechanical protection
at the stern than it would at the bow.

Hovercraft
For craft that are required to back out of long slips, or even to back into harbour
entrances, like the English Channel ferries at Dover, a rudder is fitted at the bow.
This becomes the trailing end when backing, and the ship steers satisfactorily
with a rudder at that end. A centreline rudder mounted between two widely
spaced wing propellers benefits only little or perhaps not at all from the
augmented water velocity in the propeller outflow jets. Adequate swinging effect
is then achieved by mounting two rudders abreast, one abaft each propeller.

The diving planes for controlling the rise and dive angle of a submarine are placed
at the stern, directly abaft the propellers, to benefit from the higher water velocity
in that region. Bow planes, if fitted, are used principally to control the depth at
which the craft runs. They are effective as control surfaces because only vertical
forces, not swinging or diving moments, are desired and because they project
from the hull and create up or down forces independent of the hull forces.

Control surfaces called flanking rudders are placed forward of the screw
propellers on shallow-draft push boats to assist the normal rudder in producing
side forces. They enable these craft, when pushing groups of barges 1,000 or more
feet in overall length, to maneuver around river bends and through channel turns.
Heel when turning
In a turn, the inward hydrodynamic force produced by the drift angle is applied
at a point well below the waterline. The outward centrifugal force is applied at
the centre of gravity, usually located at or above the waterline. This couple acts to
heel the ship outward to an angle at which it is balanced by the righting moment
resulting from the transverse metacentric stability. The contribution of the
rudder to this pattern is a force acting to reduce the angle of heel. Thus, in a steady
turn, if the rudder angle is suddenly removed, the outboard heel is momentarily
increased. Ships with small metacentric stability and comparatively large rudders
have capsized through this cause.

Submarines with large, highly streamlined fairwaters around the periscopes and
masts heel inward on submerged turns, especially if running at more than low or
moderate speeds. This is because a large part of the inward hydrodynamic force
is generated by the drift angle on the fairwater. This force acts inward at a level
well above the centre of gravity, where the outward centrifugal force is applied.
The outward lateral force on a rudder mounted below the main hull acts at the
same time to increase the inward heel.
Effect of propulsion-device action on maneuverability
The individual thrusts of independent wing propellers, with axes offset from the
centre of gravity, exert a swinging moment about that centre. Ships with the
rudder damaged or lost have been steered by suitable operation of the wing
propellers. On some ships, pushing ahead on one screw and pulling astern on the
other acts to turn the ship around almost on its own centre. Tugs with port and
starboard paddle wheels driven independently, or with rotating-blade propellers,
can maneuver even more readily in this fashion.

Blades of stern propellers that encounter cross flow under the ship when swinging
or yawing produce lateral forces that counteract the swinging motion and
increase the diameter of the turn. If air is drawn into the upper blades of the
propeller on a single-screw ship, excess lateral forces on the lower blades swing
the stern in the direction that the upper blades are moving, say from port to
starboard. To a certain extent, these forces can be counteracted by the rudder,
but, for the most part, the operator of a single-screw ship must foresee their
existence and make adequate allowance for them.
Maneuverability of submarines in the vertical plane
Many of the factors involved in the steering and turning of ships in the horizontal
plane apply also to the depth keeping, rising, and diving of submarines in the
vertical plane. The problem is much more severe here, however, because of the
extreme relative thinness of the layer of water between the surface and the
permissible working depth. As submarines have been built for greater depths,
their speeds have increased, and therefore the problem has been accentuated.
The undersea craft, required to run at almost constant depth for extended
periods, requires reasonable dynamic stability of route in the vertical plane. It
also requires controllability at extremely low submerged speeds so that it may
hover at one spot or creep along slowly, without making any noise. Should the
submarine crew lose vertical control with the craft headed for dangerous depths,
a high-pressure air-blowing system serves to expel some of the water in the main-
ballast tanks. The additional buoyancy thus gained checks and stops its descent.
Maneuvering predictions and model experiments
The ultimate aim of the naval architect is to formulate and collect rules and
formulas by which a ship may be designed directly or by which its behaviour and
performance may be predicted directly. The first are available in small part; some
data for the second have been derived by tests under model towing carriages and
rotating arms. These serve to determine the forces and moments resulting from
elementary motions such as ahead motion with yawing deviations and motion at
various drift angles when the centre of gravity is moving in a circular path,
simulating a steady turn. The forces and moments are then fed into established
equations of motion and the integrated performance is predicted therefrom. This
approach has been used primarily for the determination of dynamic stability of
route, which involves only fairly small angles of attack and angular velocities.
When these motions become large, as they do for large course departures, this
approach can be used only with large empirical corrections.

Free-running self-propelled ship models, sometimes radio controlled, can


simulate turns and other maneuvers, permitting derivation of the path of the
centre of gravity, changes in forward speed, rudder angles, angles of heel, and
related data. Self-propelled models, supplied with power and steered by distant
control from a towing carriage following, provide experimental checks on
steering, dynamic stability of route, effectiveness of rudders and certain
maneuvers which can be performed within the limited width of a model
testing basin.

Ships in waves
Considered as the environment for boats and ships of all kinds and sizes, the
term sea is used to denote all waters large enough for the operation of these craft,
from creeks and ponds to lakes and oceans. The wind and the ships moving across
the sea create a pattern of undulations ranging from minute ripples to waves of
gigantic size. The currents moving through it must also be taken into account in
all ship operations and in some ship-design problems. The variations in density,
resulting from the amount of salts in solution, determine the variable-ballast tank
capacity of submarines and the ability of a submarine to “sit” on a layer of dense
water while largely supported by a less dense layer above.

Considering the overall surface configuration, termed the seaway, the classical
concept of a train of regular waves is highly unrealistic, but it has some practical
uses. The normal seaway is highly irregular, with waves of different heights and
lengths traveling in many directions. For analytic purposes, it may be considered
as made up of a multitude of very low waves, having a wide range of lengths and
periods and traveling in various directions, superposed to produce the actual
seaway. When this is done, a useful approach is to use statistical methods to
define the seaway by its spectrum, which indicates the amplitudes of its many
(theoretically infinite) wave components.

The sea is also home to teeming masses of marine life, many of which
are detrimental to ships. Marine borers attack wood exposed on underwater
portions of the hull. Barnacles cling to the underwater hull, roughening its surface
and increasing the ship’s resistance to travel through the water. Sea water is
highly corrosive to most materials, and severe electrochemical effects cause rapid
disintegration of submerged metals that are unprotected.
Ship motions in waves
Treated as a rigid body, a ship partakes of six oscillatory motions in a seaway.
Three are translatory motions of the whole ship in one direction: (1) surge is the
oscillation of the ship fore and aft; (2) sway is the motion from side to side; and
(3) heave is the up-and-down motion. The other three oscillations are rotary: (4)
roll is the angular rotation from side to side about a fore-and-aft axis; (5) pitch is
the bow-up, bow-down motion about an athwartships axis; and (6) yaw is the
swing of the ship about a vertical axis. Yawing is not necessarily oscillatory for
every service condition. All six of these motions can and do take place
simultaneously in a confused sea, so the situation is most complex.

The forces and moments caused by waves are balanced by three types of forces
and moments opposing them: (1) those inertia reactions developed by the
acceleration of the ship and cargo and the adjacent water; (2) those that result in
damping the oscillatory ship motion or reducing its extent by the generation of
surface gravity waves, eddies, vortexes, and turbulence; the energy required for
setting up these disturbances is carried away and lost; (3) those of hydrostatic
nature that act to restore the ship to a position of equilibrium as, for example,
when the ship rolls to an angle greater than that called for by the exciting moment.

The behaviour of a ship in waves is too complex for the motions in all six degrees
of freedom to be completely described mathematically. However, the longitudinal
motions of pitching and heaving can be treated as a coupled system (neglecting
surging), under the assumption that lateral motions do not exist at the same time
or are reduced by stabilization to minimal values. Similarly, rolling can be treated
along with heaving and swaying on the assumption that pitch and heave do not
occur or have negligible effect. Equations of motion can then be set up that equate
the wave exciting forces and moments to the three types of forces associated with
the motions that were described above.

The theory of rolling was developed in the 19th century by Froude. The theory of
coupled pitching and heaving is more recent, stimulated by the work of Boris
Korvin-Kroukovsky in the 1950s, who applied a so-called “strip” method in which
the ship was divided longitudinally into strips or segments. The total force and
moment acting on the ship and the resulting motions were assumed to be the
result of the integration of all the forces in the individual strips without
appreciable interference. Model tests in many laboratories have confirmed the
basic soundness of this approach, although refinements are continually being
made. Computer programs for solving the equations and calculating the pitch-
heave motions of any ship are commonly used in the design stage.

The pioneering work of Manley St. Denis and Willard J. Pierson, Transactions of
the Society of Naval Architects and Marine Engineers (1953), showed how the
motions of a ship in an irregular seaway can be statistically described by assuming
that the irregular motions are the sum of the ship’s response to all the regular
component waves of the seaway described by its spectrum. This powerful tool has
permitted the extension of calculated motions (or those measured in a model
tank) to the prediction of realistic irregular sea responses and hence to the
comparative evaluation of alternative ship designs under realistic conditions.

Work by various investigators along the above lines has shown that longitudinal
weight distribution and overall ship proportions have a much greater effect than
details of hull form on pitching and heaving, and on the associated shipping of
water, slamming, and high accelerations. In general, a short pitching period in
relation to ship length is found to be advantageous in raising the limit of speed in
rough head seas. This suggests concentration of heavy weights amidships, if
possible, and favours long, slender hulls over short, squat ones

You might also like