Sie sind auf Seite 1von 109

S0300-A8-HBK-010

CHAPTER 1
NAVAL ARCHITECTURE FOR THE SALVAGE ENGINEER
1-1 INTRODUCTION
Ships are built for a wide variety of purposes, but all must meet certain fundamental requirements. They must have reserve buoyancy to enable
them to carry their designed loads and resist damage, stability to resist environmental forces or damage, and strength to withstand the stresses
imposed on their structure by their own weight, cargo, stores, and the sea. The following discussion provides the salvage engineer with the
basics of surface ship construction, stability, and strength. Submarine construction and stability are discussed in the U.S. Navy Ship Salvage
Manual, Volume 4 (S0300-MAN-A6-040).
Vessels are built to construction specifications based on stability and strength requirements, that are, in turn, based on intended service. Publicly
owned vessels (Navy, Coast Guard, etc.) are built to government specifications. Most Navy ships are built to the General Specifications for
Ships (GENSPECs), published by the Naval Sea Systems Command (NAVSEA), although some auxiliaries are built to commercial specifications.
Stability standards for Navy ships are established by Design Data Sheet (DDS) 079 issued by the Naval Ship Engineering Center. Construction
rules and stability standards for commercial vessels are established by classification societies, the International Maritime Organization (IMO),
and government regulations for the country of registry; the American Bureau of Shipping (ABS) and United States Coast Guard (USCG) establish
and enforce construction rules and stability standards for U.S. vessels. The U.S. rules are often based on IMO standards. The U.S. Maritime
Administration (MARAD) may place additional requirements on ships built with Federal financial assistance. MARAD also produces standard
designs for certain types of merchant ships. Stability and construction standards are discussed in Appendix C.
There is a basic difference in the way naval architects and salvage engineers approach the problems of ship stability and strength. Naval
architects, as designers, divide the subject into examinations of intact and damage conditions. The stability and strength of a proposed design
is examined in normal operating, or intact, conditions, which must, as matter of course, include free liquid surfaces in tanks. Damage stability
analysis examines a ship design in various hypothetical conditions of damage that include breaches in the immersed hull.
The salvage engineer on the other hand, deals with damaged stability and strength, i.e., ships in conditions of known or identifiable damage,
that may or may not include breaches in the immersed hull. There is a subtle distinction between damage and damaged stability. A salvage
engineer doesnt really deal with damage stability, or for that matter, with intact stability either. He deals with damaged stability, and conditions
that can reasonably be attained from the initial damaged condition. While the salvage engineer also examines hypothetical conditions, those
conditions usually have as a point of departure an initial damaged condition. This chapter discusses ship stability in light of those factors that
provide and enhance stability, and those that impair or degrade.
Those familiar with standard naval architecture texts may feel that this handbooks treatment of the subject glosses over the distinction between
intact and damage stability. This is true to some extent, because in the main, the distinction just doesnt matter to salvage engineers; they deal
with stabilitygood, bad, or indifferentas they find it. The fact that free surface occurs in intact ships does not obscure the fact that it always
impairs stability.
1-2 HULL FORM
A ships hull is a complex geometric form that can be defined accurately by mapping its surface in a three-dimensional orthogonal coordinate
system. If a Cartesian coordinate system is used, conventions usually set the Z-axis vertical, the X-axis longitudinal and the Y-axis athwartships.
Principal dimensions are measured along these axes. The hull form can be shown in two dimensions by a series of curves formed by the
intersection of the hull surface with planes parallel to these axes. The hull form, chosen by the designer, controls the stability and performance
characteristics of the ship in its normal environments.
1-2.1 Location of Points Within a Ship. Because a ship is a three-dimensional mobile object, references within the ship itself must be
established for locating points in, on, and about the ship. The position of any point in the ship can be described by measuring its position from
reference planes or lines. The following planes are most commonly used:

Centerplane A vertical plane passing fore and aft down the center of a ship; the plane of symmetry for most hull forms.

Design Waterplane A horizontal plane at which the hull is designed to float.

Midship Plane A transverse, vertical plane perpendicular to both the centerplane and the design waterplane, located at the
midpoint of the molded hull length between perpendiculars on the design waterplane.
Baseplane A horizontal plane passing through the intersection of the centerplane and the midships plane, or through the lowest
point of the molded hull.

1-1

S0300-A8-HBK-010

The intersections of the reference planes with specified locations on the hull create additional reference lines and points:

Forward Perpendicular (FP) A vertical line through the intersection of the stem and the design or load waterline (DWL, LWL).
After Perpendicular (AP) A vertical line at or near the stern of the ship. In naval practice, the after perpendicular passes through
the after extremity of the design waterline; in commercial practice, the after perpendicular usually passes through the rudder post,
or the centerline of the rudder stock if there is no rudder post.

Midship Section (

Centerline (C
L or CL) The projection of the centerplane in plan or end views of the hull.

or MS) An intersection of the midship plane with the molded hull.

Baseline (B
L or BL) The projection of the baseplane in the side or end views of the hull. In ships with design drag where the
baseline passes through the intersection of the midships section and the keel, parts of the hull will be below the baseline. For ships
with flat-plate keels that float on an even keel, the baseline, bottom of the molded surface, and top of the keel plate coincide; if
the keel plate is an outside strake (lapped over the adjacent strakes rather than butt-welded to them), the top of the flat-plate keel
is below the bottom of the molded surface by the thickness of the strakes on each side of it (the garboard strakes). In vessels with
hanging bar keels, the top of the keel coincides with the bottom of the molded surface.

1-2.2 Location of Points. The position of any point in the ship can be described by its:

Height above the baseplane or keel.

Athwartships position relative to the centerplane.

Longitudinal position relative to the midship section or to one of the perpendiculars.

1-2.3 Ship Dimensions. Molded dimensions, lines, etc., describe the fair surface defined by the framing and are principally of use to the
shipbuilder. Displacement dimensions and lines describe the surfaces wetted by the sea and are of principal interest to the naval architect and
salvage engineer in determining stability and performance characteristics. Extreme dimensions, such as extreme breadth, account for projections
such as overhanging decks, fender rails, etc. Molded dimensions differ from displacement dimensions by the plating, planking, or sheathing
thickness. In steel ships, this difference usually amounts to less than one percent of the total displacement. Displacement dimensions are not
usually tabulated as such; if desired, they are deduced by adding plating thickness to molded dimensions, or deducting appendage measurements
from extreme dimensions.
The principal dimensions of a ship are length, beam, and depth. Two other important dimensions are draft and freeboard. Figure 1-1 shows
the principal dimensions of a ship.

1-2

Length between perpendiculars (L, LBP or Lpp), is used for the calculation of hydrostatic properties. Length overall (LOA) is
the maximum length of the vessel, including any extensions beyond the perpendiculars, such as overhanging sterns, raked stems,
bulbous bows, etc. Length on the waterline (LWL or LWL) may or may not be the same as LBP, depending on the location of the
perpendiculars; tabulated LWL is usually taken on the design waterline.
Beam or breadth (B) is the width of the ship. Molded beam is measured amidships or at the widest section from the inside surface
of the shell plating. Maximum beam or extreme breadth is the breadth at the widest part of the ship, and is equal to the molded
breadth plus twice the plating thickness plus the width of fenders, overhanging decks, or other solid projections.
Draft (T) is the vertical distance between the waterline and the deepest part of the ship at any point along the length. Drafts are
usually measured to the keel and are given as draft forward (Tf), draft aft (Ta) and mean draft (T or Tm). A ships forward and
after draft marks are seldom at the perpendiculars and mean draft is not necessarily amidships; the slight errors introduced by using
drafts at these points can be discounted if trim is not extreme. Molded drafts are measured from the molded baseline, while keel
drafts are measured from a horizontal line though the lowest point on the bottom of the keel extended to intersect the forward and
after perpendiculars. Navigational or extreme drafts indicate the extreme depth of sonar domes, propellers, pit swords, or other
appendages which extend below the keel, and are therefore not used to calculate hydrostatic properties. Draft scales for keel drafts
are usually placed on both sides of the ship at each end as near as practical to the respective perpendiculars. The external draft
marks are generally Arabic numerals, with height and spacing arranged so that the vertical projection on the vessel of the numeral
heights and vertical spacing between numerals are both six inches. The draft figures are placed so that the bottom of the figure
indicates the keel draft. Drafts can thus be read to the nearest quarter-foot (3 inches) in relatively calm waters.
Freeboard (F) is the vertical distance between the waterline and the uppermost watertight deck.
Depth (D) is the vertical distance between the baseline and the uppermost watertight deck and is the sum of freeboard and draft.
Molded depth is measured from the top of the outer keel to the underside of the main or freeboard deck at the side. Depending
on hull form and ships attitude, both freeboard and depth can vary along the length of the ship. Unless otherwise specified,
tabulated values for depth and freeboard are usually taken at midships or at the point of minimum freeboard.

S0300-A8-HBK-010

1-2.4 Lines. The shape of a ship is


B
developed to meet specific requirements of
speed, seakeeping ability, and capacity for
the intended use of the vessel. The shape
MIDSHIPS
of the hull is defined by the plan shapes
SECTION
produced by the intersection of three
D
families of orthogonal planes and the hull
surface. Most hulls are symmetrical about
the vertical plane of the centerline. The
T
intersection of the ships molded hull
surface with this and parallel planes is
called a buttock, or buttock line. The term
CL
buttock was formerly applied only to the
AP
FP
portions of these lines aft of midships; the
DWL
forward portions were called bow lines. A
plane parallel to the baseplane and
LBP
perpendicular to the centerline plane is a
LOA
waterplane.
The intersection of
waterplanes and the molded hull are called
Figure 1-1. Principal Dimensions.
waterlines (WL).
The intersection of
transverse planes perpendicular to both
waterplanes and buttocks are termed sections. The superimposed sections (body plan), waterplanes (halfbreadth plan), and buttocks (sheer plan)
form the lines plan or lines drawing for the ship. Like other engineering drawings, the lines plan is composed of views from ahead or astern,
from above, and from the starboard side. Figure FO-1 is the lines plan for an FFG-7 Class ship.
The lines plans for steel ships usually show the molded surface. For surface ships, the molded surface is the inside of the shell plating, while
the molded surface for submarines is the outside of the hull plating. For vessels with hanging bar keels, the line of the bottom of the keel is
shown on the sheer plan to complete the lower contour of the vessel; the keel line is not usually shown for vessels with flat-plate keels because
it lies so near the line of the bottom of the molded surface. Because of the greater hull thickness, wooden ships may have separate molded and
displacement lines drawings.
1-2.4.1 The Body Plan. The body plan shows the outline of the transverse sections of a ship at equally spaced stations or ordinates along the
length of the ship. The distance between perpendiculars is commonly divided into 10 or 20 equal spaces by 11 or 21 stations, including the
forward and after perpendiculars. More or fewer stations may be used depending on the complexity of the hull shape. Half-spaced stations
may be used when the shape of the hull form changes rapidly, such as near the bow and stern. As the transverse sections are normally
symmetrical about the centerline, it is conventional to show only half sections with the forward stations on the right and after stations on the
left. Stations are numbered from forward aft, with the forward perpendicular as station zero on U.S. Navy ships. Stations forward of the forward
perpendicular (if any) may be designated by negative numbers or letters. Commercial vessels, particularly foreign-built vessels, commonly
number stations from aft forward, with the after perpendicular as zero.
1-2.4.2 Halfbreadth Plan. Due to symmetry, it is conventional to show only half of the waterplanes in a halfbreadth plan. Waterlines are
designated by their height above the baseline. The waterlines define the shape and area of the waterplane and are spaced closely enough to
accurately define the waterplane at any draft.
1-2.4.3 Sheer Plan. Superimposed buttocks form the sheer plan. They are spaced as necessary to adequately define the ships form.

1-3

S0300-A8-HBK-010

1-2.4.4 Descriptive Terms. Certain other


geometric concepts are useful in describing
a ships form. Figure 1-2 illustrates some
of the following definitions:

1-4

ONE-HALF OF
MOLDED BREADTH
TUMBLEHOME

CAMBER

Parallel midbody In many


modern ships, the form of the
hulls transverse section in the
midships region extends without change for some distance
fore and aft. This is called
parallel midbody and may be
described as extensive or
short, or expressed as a
fraction of the ships length.
Even in ships without parallel
midbody, the form of the
fullest transverse section
changes only slightly for
small distances forward or aft.

FREEBOARD

MOLDED
DEPTH

DESIGN WATERLINE
DESIGN
DRAFT
MOLDED
DRAFT

Forebody The portion of


the hull forward of the midship section.
After body The portion of
the hull abaft the midship
section.
Entrance The immersed
portion of the hull forward of
the section of greatest immersed area (not necessarily
amidships) or forward of the
parallel midbody.

DEADRISE

SHEER
AFT

CL

MOLDED
BASE LINE

SHEER
FORWARD

DEPTH

Figure 1-2. Hull Form Nomenclature.

Run The immersed portion of the hull aft of the section of greatest immersed area or aft of the parallel midbody.
Deadrise The departure of the bottom from a transverse horizontal line measured from the baseline at the molded breadth line
as shown in Figure 1-2. Deadrise is also called rise of floor or rise of bottom. Deadrise is an indicator of the ships form; fullbodied ships, such as cargo ships and tankers, have little or no deadrise, while fine-lined ships have much greater deadrise along
with a large bilge radius. Where there is rise of floor, the line of the bottom commonly intersects the baseline some distance from
the centerline, producing a small horizontal portion of the bottom on each side of the keel. The horizontal region of the bottom
is called flat of keel, or flat of bottom. While any section of the ship can have deadrise, tabulated deadrise is normally taken at
the midships section.
Knuckle An abrupt change in the direction of plating or other structure.
Chine The line or knuckle formed by the intersection of two relatively flat hull surfaces, continuous over a significant length
of the hull. In hard chines, the intersection forms a sharp angle; in soft chines, the connection is rounded.
Bilge radius The outline of the midships section of very full ships is very nearly a rectangle with its lower corners rounded.
The lower corners are called the bilges and the shape is often circular. The radius of the circular arc is called the bilge radius
or turn of the bilge. The turn of the bilge may be described as hard or easy depending on the radius of curvature. If the shape
of the bilge follows some curve other than a circle, the radius of curvature of the bilge will increase as it approaches the straight
plating of the side and bottom. Small, high-speed or planing hulls often do not have a rounded bilge. In these craft, the side and
bottom are joined in a chine.
Tumblehome The inward fall of side plating from the vertical as it extends upward towards the deck edge. Tumblehome is
measured horizontally from the molded breadth line at the deck edge as shown in Figure 1-2. Tumblehome was a usual feature
in sailing ships and many ships built before 1940. Because it is more expensive to construct a hull with tumblehome, this feature
is not usually incorporated in modern merchant ship design, unless required by operating conditions or service (tugs and
icebreaking vessels, for example). Destroyers and other high-speed combatants are often built with some tumblehome in their mid
and after sections to save topside weight.

S0300-A8-HBK-010

Flare The outward curvature of the hull surface above the waterline, i.e., the opposite of tumblehome. Flared sections cause
a commensurately larger increase in local buoyancy than unflared sections when immersed. Flaring bows are often fitted to help
keep the forward decks dry and to prevent "nose-diving" in head seas.
Camber The convex upwards curve of a deck. Also called round up, round down, or round of beam. In section, the camber
shape may be parabolic or consist of several straight line segments. Camber is usually given as the height of the deck on the
centerline amidships above a horizontal line connecting port and starboard deck edges. Standard camber is about one-fiftieth of
the beam. Camber diminishes towards the ends of the ship as the beam decreases. The principal use of camber is to ensure good
drainage in calm seas or in port, although camber does slightly increase righting arms at large angles of inclination (after the deck
edge is immersed). Not all ships have cambered decks; ships with cambered weather decks and flat internal decks are not
uncommon.
Sheer The rise of a deck above the horizontal measured as the height of the deck above a line parallel to the baseline tangent
to the deck at its lowest point. In older ships, the deck side line often followed a parabolic profile and sheer was given as its value
at the forward and after perpendiculars. Standard sheer was given by:
sheer forward = 0.2L + 20
sheer aft = 0.1L + 10

where sheer is measured in inches and L is the length between perpendiculars in feet. Actual sheer often varied considerably from
these standard values; the deck side profile was not always parabolic, the lowest point of the upper deck was usually at about 0.6L,
and the values of sheer forward and aft were varied to suit the particular design. Many modern ships are built without sheer; in
some, the decks are flat for some distance fore and aft of midships and then rise in a straight line towards the ends. Sheer
increases the height of the weather decks above water, particularly at the bow, and helps keep the vessel from shipping water as
she moves through rough seas. Some small craft and racing yachts are given a reverse or hogged sheer to give headroom
amidships without excessive depth at bow and stern.

Rake A departure from the vertical or horizontal of any conspicuous line in profile, defined by a rake angle or by the distance
between the profile line and a reference line at a convenient point. Rake of stem, for example, can be expressed as the angle
between the stem bar and a vertical line for ships with straight stems. For curved stems, a number of ordinates measured from
the forward perpendicular are required to define the stem shape. Ships designed so that the keel is not parallel to the baseline and
DWL when floating at their designed drafts are said to have raked keels, or to have drag by the keel.
Cut-up When a keel departs from a straight line at a sharp bend, or knuckle, the sloping portion is called a cut-up. High-speed
combatants usually have a long cut-up aft (extending 13 to 17 percent of LWL) to enhance propeller performance and
maneuverability. Ice-breaking vessels often have a cut-up forward to allow the ship to ride up on the ice.
Deadwood Portions of the immersed hull with significant longitudinal and vertical dimensions, but without appreciable transverse
dimensions. Deadwood is included in a hull design principally to increase lateral resistance or enhance directional stability without
significantly increasing drag when moving ahead. Sailing craft require deadwood to be able to work to windward efficiently.
Skegs or fins are fitted on barges to give directional stability. Deadwood aft is detrimental to speed and quick maneuverability
and is minimized by use of cut-up sterns in high-speed combatants and by arched keels or sluice keels (with athwartships apertures)
in tugs and workboats.
Appendages Portions of the vessel that extend beyond the main hull outline or molded surface. Positive appendages, such as
rudders, shafts, bosses, bilge keels, sonar domes, etc., increase the underwater volume, while negative appendages, such as bow
thruster tunnels and other recesses, decrease the underwater volume. Shell plating, lying outside the molded surface, is normally
the largest single appendage, and often accounts for one-half to two-thirds of the total appendage volume. Appendages generally
account for 0.2 to 2 percent of total immersed hull volume, depending on ship size, service, and configuration. Paragraph 1-4.10.2
discusses methods for estimating appendage displacement.
Hull Surfaces Hull surfaces are either warped, consisting of smoothly faired, complex three-dimensional curves, developed,
consisting of portions of cylinders or cones, or flat. Hydroconic hulls are built up of connected flat plates rather than plates rolled
to complex curves. Hydroconic construction lowers production costs and may simplify fitting patches to a casualty.

1-5

S0300-A8-HBK-010

1-2.5 Coefficients of Form. Coefficients of form are


dimensionless numbers that describe hull fineness and
overall shape characteristics. The coefficients are ratios of
areas or volumes for the actual hull form compared to
prisms or rectangles defined by the ships length, breadth,
and draft. Since length and breadth on the waterline as
well as draft vary with displacement, coefficients of form
also vary with displacement. Tabulated coefficients are
usually based on the molded breadth and draft at designed
displacement. Length between perpendiculars is most often
used, although some designers prefer length on the
waterline. Coefficients of form can be used to simplify
area and volume calculations for stability or strength
analyses. As hull form approaches that of a rectangular
barge, the coefficients approach their maximum value of
1.0. The following paragraphs describe the most commonly
used coefficients. Table 1-1 gives sample coefficients for
different type ships.
1-2.5.1 Block. The block coefficient (CB) is the ratio of
the immersed hull volume () at a particular draft to that
of a rectangular prism of the same length, breadth, and
draft as the ship:

CB =
BTL
where:

B
T
L

=
=
=
=

immersed volume, [length3]


beam, [length]
draft, [length]
length between perpendiculars, [length]

1-2.5.2 Midship Section. The midship section coefficient


(CM) is the ratio of the area of the immersed midship
section (Am) at a particular draft to that of a rectangle of the
same draft and breadth as the ship:
CM =

B
T

=
=

Block
Coefficient
CB

Type Ship

Navy Ships
Aircraft Carrier (CV-59 Class)

0.578

0.984

0.729

Battleship (BB-61 Class)

0.594

1.000

0.694

Cruiser (CGN-38 Class)

0.510

0.810

0.780

Destroyer (DD-963 Class)

0.510

0.850

0.760

Frigate (FFG-7 Class)

0.470

0.770

0.750

Replenishment Ship (AOR-1 Class)

0.652

0.981

0.777

Salvage Tug (ARS-50 Class)

0.542

0.908

0.791

Commercial Vessels
General Cargo (slow-speed)

0.800

0.992

0.880

General Cargo (medium-speed)

0.700

0.980

0.810

General Cargo (high-speed)

0.576

0.972

0.695

Tanker (35,000-ton DWT)

0.757

0.978

0.845

Large Tanker (76,000-ton DWT)

0.802

0.997

0.874

VLCC (250,000-ton DWT)

0.842

0.996

0.916

Container Ship

0.600

0.970

0.740

RO/RO

0.568

0.972

0.671

Ore Carrier

0.808

0.995

0.883

Great Lakes Bulk Carrier

0.900

0.995

0.950

Passenger Liner

0.530

0.956

0.690

Barge Carrier

0.570

0.950

0.820

0.530

0.910

0.680

Ocean Tug, Trawler

0.550

0.833

0.850

BT

Offshore Supply Vessel

0.660

0.906

0.892

Harbor Tug

0.585

0.892

0.800

Ocean Power Yacht (250 ft LWL)

0.565

0.938

0.724

area of the immersed portion of the midships


section, [length2]
beam, usually taken at the waterline, [length]
draft, [length]

Coefficients for commercial vessels are typical values; coefficients for specific ships will
vary. Coefficients of form for U.S. Navy ships can be obtained from Naval Sea Systems
Command, Code 55W. Coefficients for many merchant vessels are available from the
National Cargo Bureau, telephone (212) 571-5000. The builders hull number or name
and type of vessel must be provided to access the data files.

1-2.5.3 Waterplane. The waterplane coefficient (CWP) is


the ratio of the area of the waterplane (AWP) to that of a rectangle of the same length and breadth as the ship:
CWP =

where:

1-6

Waterplane
Coefficient
CWP

Large Car Ferry

If the vessel has bulges or blisters below the waterline, CM


may be greater than 1.

AWP =
B
LWL

Midship
Coefficient
CM

AM

where:
AM =

Table 1-1. Typical Coefficients of Form.

area of the waterplane, [length2]


= beam, [length]
= length on the waterline, [length]

AWP
LWL B

S0300-A8-HBK-010

1-2.5.4 Prismatic. The longitudinal prismatic coefficient (CP) is the ratio of the immersed volume to the volume of a prism with length equal
to the ships and cross-section area identical to the midship section:
CP =

= B
AM L
CM

where:
=
AM =
L =

immersed volume, [length3]


area of the immersed portion of the midships section, [length2]
length between perpendiculars, [length]

If length between perpendiculars and length on the waterline are equal (as they are for Navy ships), the prismatic coefficient is equal to the block
coefficient divided by the midships section coefficient. The prismatic coefficient thus indicates the longitudinal distribution of the underwater
volume of a ships hull. For a given length, breadth, draft, and displacement, a low (fine) CP indicates a hull with fine ends. A large (full) value
for CP indicates a hull with relatively full ends. For this reason, the prismatic coefficient is sometimes called the longitudinal coefficient.
The vertical prismatic coefficient (CVP) is
the ratio of the immersed hull volume to
the volume of a prism having a length
equal to the ships draft and a cross section
identical to that of the waterplane:

AWP T

100

where:

AWP =
T

immersed volume,
[length3]
area of the waterplane,
[length2]
= draft, [length]

The vertical prismatic coefficient is equal to


the block coefficient divided by the
waterplane coefficient and indicates the
vertical distribution of the underwater
volume. A full CVP indicates a concentration of volume near the keel and a fine
CVP, a concentration nearer the waterline.

BREADTH; DEPTH; DRAFT, FT

CVP =

125

75

50

H
DT
EA
R
B

PTH
DE

T
DRAF

25

1-2.6 Ship Proportions. Throughout this


handbook and many naval architecture
texts, relationships and approximations for
0
various hydrostatic and stability parameters
0
450
600
750
900
are given as applicable to ships of ordinary,
LBP, FT
or normal form. With the broad range of
FROM ELEMENTS OF SHIP DESIGN, R. MUNRO-SMITH, 1975.
ship type, size, and service requirements,
normal form is best defined by a range of
coefficients and dimension ratios. Table
Figure 1-3. Approximate Ship Proportions.
1-1 gives typical coefficients of form and
Figure 1-3 shows approximate linear
relationships between length, beam, depth,
and service draft. The relationships given below, adapted from R. Munro-Smiths Elements of Ship Design, and deadweight coefficients (defined
in Paragraph 1-3.3), are used to estimate ship dimensions during preliminary design and can help to determine whether a hull should be
considered normal.
Dimensional Ratios:
Ship type
General Cargo
Tankers
VLCC

L/B
6.3 to 6.8
7.1 to 7.25
6.4 to 6.5

B/T
2.1 to 2.8
2.4 to 2.6
2.4 to 2.6

T/D
0.66 to 0.74
0.76 to 0.78
0.75 to 0.78

1-7

S0300-A8-HBK-010

Maximum block coefficient for service conditions:


CB 1.00

0.23

Vk

(general cargo ships)

L
CB 1.00

0.19

Vk

(tankers, bulk carriers)

L
CB 1.00

Vk

0.175

(VLCC)

L
where:
Vk =
L =

service speed, knots


length between perpendiculars, ft

Beam range:
L
9
L
9
L
9

+ 20 ft B

+ 15 ft B

+ 39 ft B

+ 25 ft (cargo ships)
+ 21 ft (tankers, bulk carriers)
+ 50 ft, or
46 ft (VLCC)

where:
B = beam, ft
Beam to length relationship:
B = Ln
where B and L are given in feet and:
n

0.61 to 0.64 for general cargo ships


0.66 to 0.68 for VLCC

Length-beam product to deadweight relationship:


0.0093LB =

DWT
C
T

where:
L
B
DWT
T
C

=
=
=
=
=
=
=

length between perpendiculars, ft


beam, ft
deadweight, lton
draft, ft
0.85 to 2.0 for general cargo ships
0.525 to 0.590 for tankers
0.446 to 0.459 for VLCC

1-2.7 Offsets. The hull form can be described in tabular format by a set of measurements known as offsets. Offsets are distances measured
from the centerline to the side of the ship at each station and waterline. Molded offsets are measured to the molded surface (inside of shell
plating for steel surface ships); displacement offsets are measured to the outer hull surface. Offsets define the hull proper, without appendages.
Supplementary appendage offset tables are sometimes available. Molded or displacement offsets are usually presented in a table in the form
feet-inches-eighths. The table of offsets for an FFG-7 Class ship shown in Figure FO-1 is typical. The waterline halfbreadth entry for station
4 at the 8' 0" waterline reads 10 - 2 - 3 indicating 10 feet, 23 8 inches. Since the station spacing is given as 20.4 feet on the plan (LBP = 408
feet, 408/20 stations = 20.4), this offset precisely locates the point on the skin of the ship 81.6 feet from the forward perpendicular (4 20.4),
eight feet above the baseline and 10 feet 23 8 inches from the centerline.
Lines drawings can be constructed from tables of offsets. Of more use to the salvor is the fact that offsets can be obtained from body or
halfbreadth plans and used to determine ship volumes and areas by numerical integration (described in Paragraph 1-4). Offsets can be scaled
from arrangement drawings, or in the worst case, measured on site.

1-8

S0300-A8-HBK-010

1-2.8 Wetted Surface. The area of all or part of a ships hulls wetted surface is important to hydrodynamic resistance and pressure force
calculations. Wetted surface multiplied by average shell thickness calculates shell volume to be added to the molded volume to determine total
displacement. The area of complex hull surfaces can be calculated by numerical integration from offsets or the shell expansion plan, but this
is a tedious and time-consuming task. Wetted surface can be estimated by one of the following empirical relationships:
Denny-Mumford Formula:

= 1.7L T
T

AS = 1.7L T

Table 1-2. Taylors Coefficient.

L B CB

Taylors formula:
A S = C D L
Haslar formula for fine-lined ships:
AS

= 3.3

2/3

L
2.09 1/3

where:
AS
L

T
B
CB
D
C

=
=
=
=
=
=
=
=

B/T

3.5

16.0

4.0

16.5

5.0

17.5

8.0

20.5

9.0

21.3

10.0

22.2

11.0

23.0

12.0

23.8

13.0

24.5

14.0

25.1

16.0

26.3

18.0

27.2

wetted surface, ft2, at mean draft T, ft


length between perpendiculars, ft (immersed length)
displacement volume, ft3 = CB LBT
mean draft, ft
molded beam, ft
block coefficient
displacement, ltons
a coefficient, ranging from 15.2 to 16.0 for vessels with 0.8 Cm 0.98 and 2.5 B/T 3.5. For shallow draft vessels, C is
expressed as a function of B/T in Table 1-2.

1-3 DISPLACEMENT AND BUOYANCY


A body immersed in a fluid will experience an upward force equal to the weight of the volume of fluid displaced. This force of buoyancy is
the resultant of the normal pressures exerted by the fluid on each element of the immersed bodys surface. Buoyancy is opposed by the
downward force of gravity, or the objects weight. In order for equilibrium to exist, the two forces must be balanced. An object heavier than
an equivalent volume of water has negative buoyancy and will sink until it encounters a solid object or denser liquid, where its apparent weight
is decreased by the buoyant force acting on it. Similarly, an object less dense than water will exhibit positive buoyancy and will float with an
immersed volume such that the weight of the displaced water exactly equals the objects weight. Deeper immersion requires the application
of force. An object whose density equals that of the surrounding water is said to have neutral buoyancy and will float at whatever depth it is
placed. A ship floats by enclosing large volumes of less dense material, principally air, in a watertight skin so that its average density is less
than that of the surrounding water. To be useful, a ships effective density must be much less than that of the surrounding water to allow the
ship to support not only its own weight, but also that of crew, cargo, stores, etc.
1-3.1 Ships Weight, Displacement and Capacity. An objects displacement is the weight of the water it displaces; displacement represents
the force of buoyancy (B) acting on the object. For a ship in static equilibrium, floating free of any solid support, displacement (D) is equal
to the weight of the ship and everything in it (W), measured in long tons of 2,240 pounds. Displacement is usually given for either the
lightshipthe weight of the ship without cargo or storesor full-load conditions. A ships displacement is related to the volume of displaced
water, called the displacement volume or volume of displacement ( or V), by the weight density of water (g/gc).
D =

= W

gc

If mass density is given in slugs per cubic foot, and g in feet per second per second (ft/sec2), g/gc gives weight density in pounds-force per cubic
foot. In a standard gravitational field (g = 32.174 ft/sec2) pounds-mass and pounds-force are numerically equal. Since the worldwide variation
of gravitational acceleration is slight, weight density in pounds-force per cubic foot () can be taken as numerically equal to mass density, in
pounds-mass per cubic foot without significant error.

1-9

S0300-A8-HBK-010

With weight held constant, the product of displacement volume and water density must also be constant. For a given weight, displacement
volume varies inversely with the density of the surrounding waterdisplacement volume in water of known density can be related to
displacement volume water of any density:
1 1 = 2 2

2 =

1 1
2

The density of seawater varies with salinity and temperature, but is approximately 64 pounds per cubic foot; the density of fresh water is about
62.4 pounds per cubic foot. It is sometimes more convenient to use the inverse density, or specific volume (), of 35 cubic feet per ton of
seawater. The equivalent figure for fresh water is 35.9, commonly rounded to 36.
W =

W =

sw
35

fw
36

36
fw = sw
35
Care must be exercised not to confuse displacement, measured in long tons, with gross, net, or register tonnage. Tonnage is a measurement
of the enclosed volume of a ship used to describe her cargo capacity and does not indicate displacement. Register tonnage (gross and net) is
measured according to the rules of the country of registry or international rules, and is used as a basis for port fees, canal tolls, and similar
charges. Measurement tons were formerly equal to 100 cubic feet, but the more recent international rules determine tonnage by formulas that
do not relate volume to tonnage directly. Gross tonnage is a measure of the internal volume of the entire shipthe hull plus enclosed spaces
above the main deck. Net tonnage is derived from a formula based on the molded volume of cargo spaces, the number of passengers carried,
molded depth, and service draft; net tonnage gives an indication of the ships earning capacity. Commercial vessels engaged in international
voyages are issued a Tonnage Certificate by the country of registry. Certain special tonnages, such as Suez or Panama Canal tonnages, are
calculated by somewhat different formulae and recorded on separate certificates.
Cargo capacity may also be given in conventional volumetric units. Tank capacities are usually specified in barrels, gallons, or cubic meters.
For petroleum products and other liquids subject to thermal expansion, practical capacity is less than net capacity, to ensure that a tank "filled"
with cold oil will not overflow as the oil warms. U.S. Navy practice sets oil tank operating capacity at 95 percent of net capacity; U.S. Merchant
Marine practice at 98 percent. Dry cargo capacity is specified in cubic feet or cubic meters. Bale capacity is the volume below deck beams
and inboard of cargo battens, that is free for the stowage of bags, barrels, crates, bales, pallets, etc. Grain capacity is the net molded underdeck
volume, after deductions for the volume of frames, floors, and other structure, that is available for the stowage of granular bulk cargo. Capacity
of container ships is expressed as the number of standard 8-foot-wide by 8-foot-high containers of specified length that can be carried, often
converted to 20-foot equivalent units (TEU), or 40 foot equivalent units (FEU). Capacity for roll-on/roll-off (RO/RO) cargo and vehicle carriers
may be expressed as the number of units that can be carried or as the area of the cargo decks, in square feet or square meters.
1-3.2 Standard Loading Conditions. Displacement and stability characteristics are often referenced to certain standard conditions of loading.
1-3.2.1 U.S. Navy Ships. Characteristics are usually tabulated for the following standard conditions of loading (from NSTM Chapter 096):

1-10

Condition A - Lightship The ship complete, ready for service in every respect, including permanent ballast (solid and liquid),
onboard repair parts, aviation mobile support equipment as assigned, and liquids in machinery at operating levels, without any
items of variable load (provisions, stores, ammunition, crew and effects, cargo, aircraft and aviation stores, passengers, saltwater
ballast, fuel and other liquids in storage tanks). Formerly Condition II.
Condition A-1 - Lightship Condition A without permanent ballast. Formerly condition II-A.
Condition B - Minimum Operating Condition A condition of minimum stability likely to exist in normal operation (following
the ships liquid loading instructions). For warships, Condition B approximates the ships condition toward the end of a hostile
engagement following a long period at sea. Liquids are included in amounts and locations that will provide satisfactory stability,
trim, and limitation of list in case of underwater damage. Formerly Condition V.
Condition C - Optimum Battle Condition As formerly applied to minor combatants, the ship loaded with full ammunition
allowance and two-thirds provisions, fuel, lube oil, etc. Fuel distribution and seawater ballast are in accordance with liquid loading
instructions, except that service tanks are assumed half-full and one pair of storage tanks per machinery box are assumed empty.
Formerly Condition LS. In current practice, this condition applies only to ships with extensive underwater defense systems, such
as aircraft carriers and battleships. Liquids are carried in the amounts and locations that provide the optimum resistance to
underwater damage.

S0300-A8-HBK-010

Condition D - Full Load Two different full-load conditions are defined:


(1) Full load (contractual) The ship complete, ready for service in every respect; Condition A plus authorized complement of
personnel and passengers and their effects, full allowance of ammunition in magazines and ready service spaces, full allowance
of aircraft and vehicles with repair parts and stores, provisions and stores for the periods specified in design specifications,
sufficient fuel to meet endurance specifications, anti-roll tank liquid, liquids in tanks to required capacity in accordance with
liquid load instructions, and cargo in the amounts normally carried or a specified portion of full capacity. This condition is
used for weight estimates and reporting.
(2) Full load (departure) Same as full load (contractual) except that fuel and lube oil tanks are 95-percent full, potable and feed
water tanks 100-percent full. Formerly Condition VI. This condition is used in inclining experiment reports.

Condition E - Capacity Load The ship complete, ready for service in every respect; Condition A plus the maximum number
of crew and passengers that can be accommodated, with their effects, maximum stowage of ammunition in magazines and ready
service spaces, full allowance of aircraft and vehicles with repair parts and stores, maximum amount of provisions and stores that
can be carried in assigned spaces, tanks filled to maximum capacity (95 percent for oil tanks, 100 percent for fresh water),
maximum amounts of cargo and supplies, with the provision that the limiting drafts not be exceeded.

Data is sometimes tabulated for special or unusual loading conditions, such as special ballast conditions for amphibious warfare ships. Details
for each condition of loading are found in the ships damage control book. Standard displacement is a condition defined by the Washington
Naval Conference of 1923 as "The displacement of the ship, fully manned, engined, and equipped ready for sea, including all armament and
ammunition, equipment, outfit, provisions and fresh water for the crew, miscellaneous stores and implements of every description that are
intended to be carried in war, but without fuel or reserve feed water on board." Standard displacement was defined primarily as an aid to
ensuring compliance to restriction on warship size and total naval tonnage under international treaties, but provides a convenient means of
comparing warships and is commonly given in published summaries of naval strength, such as Janes Fighting Ships. Characteristics for standard
displacement are not normally tabulated in damage control books or similar documents.
1-3.2.2 Commercial Vessels. Two major conditions of loading are referenced in dealing with commercial vessels:

Lightship, Lightweight, or Light Displacement The ship with all items of outfit, equipment, and machinery, including boiler
water and lubricating oil in sumps, but without cargo, provisions, stores, crew, or fuel.
Fully Loaded Lightship plus cargo, fuel, stores, etc., to settle the ship to her load line. Also loaded, load, or full-load
displacement. For ships designed to carry different classes of cargo, full-load conditions may be tabulated for each type of cargo.

The trim and stability booklet will normally tabulate stability data for ballasted and partly loaded conditions, and for end of voyage and
intermediate conditions with varying amounts of fuel and stores consumed.
1-3.2.3 Loading Instructions. Specific loading instructions are provided to help operating personnel avoid loading the ship so that her stability
is dangerously low or the hull girder is overstressed. The most basic instruction is that ships shall not be loaded so heavily that their load line
(merchant) or limiting draft marks (naval) are submerged. Detailed loading instructions are given in the trim and stability booklet for merchant
ships or the damage control book for Navy ships. In certain types of ships, such as container ships, RO/RO ships, barge carriers, and ferries,
improper loading can easily reduce stability to dangerously low levels. In other ships, such as tankers and ore carriers, improper loading can
seriously overstress the hull. Transient conditions created while loading or unloading can also degrade stability or overstress the hull. Load
and stability computers supplement or replace loading instructions on many tankers, bulk carriers, and other large ships or ships with unusual
stability problems. Load computers are briefly described in Paragraph 4-2.5.3.
1-3.3 Deadweight. Deadweight (DWT) is the load carried by a ship. It is the difference between the lightship displacement and total
displacement of the ship at any time. Maximum or load deadweight is the carrying capacity of a ship measured in 2,240-pound long tons, and
is the difference between the lightweight and fully loaded displacements. Deadweight includes fuel, provisions, munitions, crew and effects,
cargo, or any other weight carried. For a merchant ship, cargo deadweight, paying deadweight, or payload is the part of the deadweight that
is cargo and therefore earning income.
It is not uncommon for the deadweight of a merchant ship to be given, but not its full-load displacement. A deadweight coefficient (CDWT) can
be defined as the ratio of full-load displacement to total deadweight:
CDWT =

FL
DWT

FL = DWT CDWT

where:
CDWT
FL
DWT

=
=
=

deadweight coefficient
full-load displacement
total deadweight

1-11

S0300-A8-HBK-010

Typical ranges for deadweight coefficient are given by R. Munro-Smith (Elements of Ship Design, 1979):
General cargo ship
Ore carrier
Bulk carrier
Oil tanker
Very large tanker, VLCC

1.39
1.30
1.19
1.16
1.28

1.61
1.39
1.28
1.25
1.32

1-3.4 Change in Draft. Draft is significant as the only principal dimension that varies routinely, while length and beam remain essentially
constant. Volume of displacement, and therefore draft, will change as a ships displacement changes due to loading or discharging cargo,
consuming or loading fuel or stores, or flooding. The new volumes and mean drafts can be computed by using the relationships shown. For
example: a box-shaped lighter 100 feet long, 30 feet wide, and 10 feet in depth, displacing 429 tons of seawater with zero trim. Because
waterplane area is constant at any draft, drafts can be found by:
= W = 35 (429) = 15,015 ft3
= L B T = 100 (30) T = 15,015 ft3
T =

LB

15,015
= 5 ft
100 (30)

where:

L
B
T

=
=
=
=
=
=

displacement volume, ft3 = LBTCB; for box-shaped lighter CB = 1.0


total weight of the barge, lton
specific volume of seawater = 35 ft3/lton
length between perpendiculars, ft
beam, ft
draft, ft

If weight (displacement) is decreased to 350 tons, the new mean draft is given by:
= 35W = 35(350) = 12,250 ft3
T =

12,250
= 4.08 ft = 4 ft 1 in.
3,000

For a complex ship shape, drafts cannot be calculated directly. The change in draft (T) can be determined if certain assumptions are made.
The increase in volume can be considered to be a prism of uniform thickness with vertical sides and horizontal section with area equal to the
waterplane area. For a wall-sided vessel (one with vertical sides, like the box-shaped lighter), this is mathematically exact; it is sufficiently
accurate for most ships for small changes in draft. The thickness of the prism is determined by dividing its volume by the area of the
waterplane:
T =

(15,015) 12,250)
=
=
AWP
L B CWP
(100)(30)(1.0)

= 0.92 ft = 11 in.
where:
T

AWP =
CWP =

1-12

= change in draft, ft
= change in displacement volume, ft3
waterplane area, ft2
waterplane coefficient

S0300-A8-HBK-010

The salvor may encounter ships in water of varying densities. The waters of harbors and estuaries might be salty, fresh or brackish; the salinity
and density of the water may depend on the state of the tide. The equalities shown can be used to relate displacement volume, draft and
displacement of any ship in water of any known density. Recalling that:
1 1 = 2 2
1
1
L B T CB

2
L B T CB

where:

L
B
T
CB

=
=
=
=
=
=
=

displacement volume, ft3


water density, lb/ft3
inverse density or specific volume, ft3/lton
length between perpendiculars, ft
beam, ft
draft, ft
block coefficient

With length and breadth constant, and CB assumed constant for a small change in draft,
T1

T2

1
T1

2
1

T2

and:
T2 =

For saltwater and fresh water:

SW
FW

T12
1
=

35
36

and:
36
TFW = TSW
35
The difference between fresh water and seawater drafts may range from 6 inches for an FFG-7 to 1.2 feet for a large aircraft carrier, or more
on a large crude carrier. Differences encountered when dealing with brackish water will be correspondingly less, and may be dealt with by using
values for fresh water and saltwater as upper and lower boundaries if the water density is unknown or variable.

1-13

S0300-A8-HBK-010

1-3.5 Tons per Inch Immersion (TPI). The foregoing analysis can be carried a step further to determine the change in displacement (D)
required to cause a change in draft of one inch. For seawater:
T =

;
AWP

= 35D

W =

T =

35 D
AWP

T AWP
35

Substituting 1 inch = 1 12-foot for T:


D =

AWP
(35)(12)

AWP
420

= TPI

where:
D

35
AWP =
TPI

= displacement, lton
= displacement volume, ft3
= water density, lb/ft3
= specific volume, ft3/lton
waterplane area, ft2
= tons per inch immersion, lton/in.

Tons per inch immersion for water of any density can be obtained by a similar calculation.
1-3.6 Reserve Buoyancy. The watertight volume between the waterline and the uppermost continuous watertight deck provides the reserve
buoyancy to the ship. Although this volume does not actually provide any buoyancy, it is available to enable the ship to take on additional
weight. Freeboard is an indication of the reserve buoyancy remaining. Freeboard and draft can be considered opposite ends of a sliding scale,
with draft representing the buoyancy in use and freeboard the buoyancy remaining.
1-3.7 Center of Gravity. A homogeneous bodys center of gravity is located at its center of volume, or centroid. The center of gravity of
a ship is not so easily definable, but can be assumed to be located on the centerline near the midship plane in a ship floating without list or trim.
The center of gravity of a ship is a function of weight distribution; its position varies with loading. With all weights stationary, the center of
gravity remains fixed regardless of the movement of the ship. Its position relative to any of the three reference planes along a perpendicular
axis (n) is given mathematically by:
n dw
=
G =
W

nw
w

where:
G
n
W

=
=
=

position of the center of gravity along any axis


distance from the origin to an incremental weight dw, or to an individual weight w
total weight = w

The location of the center of gravity greatly influences the stability characteristics of a vessel: the vertical location (VCG, or KG) influences
a vessels ability to resist heeling forces; the longitudinal location (LCG) relative to the longitudinal location of the center of buoyancy
determines trim; and a transverse location (TCG) off the centerline results in a list.

1-14

S0300-A8-HBK-010

1-3.8 Center of Buoyancy. The force of buoyancy, like gravity, can be resolved to act upwards through a single point. The center of buoyancy
(B) is located at the centroid of the submerged hull form. As the ship inclines, the shape of the underwater volume changes and the center of
buoyancy moves to the new geometric center. When a ship is at rest without list, the center of buoyancy is on the centerline directly below
the center of gravity. The location of the center of buoyancy responds directly to draft changes. As the ships displacement is increased or
decreased with a corresponding change in draft, the center of buoyancy will move to the new centroid of the redefined submerged hull form.
1-3.9 Metacenter. As shown in Figure 14, vertical lines drawn through successive
centers of buoyancy (B1, B2, and B3) as the
ship inclines slightly intersect at an
imaginary point on the centerline called the
metacenter (M). In a stable vessel, M is
located above the center of gravity. The
vertical location of M is one of the most
critical parameters affecting a ships initial
stability.

(HEEL ANGLES EXAGGERATED)


M

WL2
WL

1-3.10 Center of Flotation. The center of


flotation is the point about which the ship
trims and heels, and is at the geometric
center of the ships floating waterplane. It
is usually located aft of midships, although
it may be forward of midships in fullbodied ships.

WL

WL
WL1
B2

B1

WL2

1-3.11 Bonjeans Curves. Bonjeans


Curves or Curves of Sectional Areas are a
Figure 1-4. Relative Positions of M, B, and G During Small Inclinations.
collection of curves plotting sectional area
along the X-axis against draft on the Y-axis.
The curves are usually presented in one of the two formats shown in Figure FO-3. The section area curve may show area for either the whole
section, or for one side only, as noted on the drawing. The areas generally do not account for appendages, but may include shell plating, as
noted on the drawing. Section areas can be taken from the curves for any draft and any condition of trim or hull deflection. Section area is
converted to unit buoyancy by dividing by the specific volume of water (35 cubic feet per long ton per foot of length for seawater). Volume
of displacement and other hydrostatic properties can be determined by integration of section area or derived unit buoyancy ordinates by the
numerical methods described in Paragraph 1-4.
The rosette arrangement (Figure FO-3A), with all the curves drawn to a single set of axes, produces a more compact drawing and is favored
by some designers because lack of fairness in the hull will show itself with the curves lying side by side. Section areas are read from the
intersection of a horizontal line through the station draft on the center scale with the appropriate curve. When calculating buoyancies for varying
waterlines or wave profiles, it is sometimes more convenient to arrange the curves along the ships profile, with a vertical axis at each station
as shown in Figure FO-3B. With the section area curves arranged in this format, a trimmed waterline can be plotted as a straight line passing
through the forward draft at station zero, and the after draft at the after perpendicular, eliminating the need to determine draft at each station.
Section areas can be picked off by drawing a horizontal line from the intersection of the waterline with each vertical station marker to the
appropriate curves. If the Bonjeans Curves are not available in this format, the curves and area scale can be traced from the rosette onto a hull
profile drawn on tracing paper. The horizontal length scale for the hull profile is not critical, but should be consistent throughout its length if
buoyancy is to be calculated on waterlines that are not horizontal.

1-15

S0300-A8-HBK-010

1-4 APPROXIMATE INTEGRATION TECHNIQUES AND APPLICATIONS


The salvage engineer may be required to calculate hydrostatic data for a casualty when curves of form or other documents are not available;
for a casualty in an unusual condition, such as a ship floated upside down or on its side; or for portions of a ship that has been cut into sections.
A ships form consists of a number of intersecting surfaces, usually of nonmathematical form. Areas and volumes enclosed by these surfaces,
as well as moments of areas and volumes, and second moments of area, must be determined to calculate hull hydrostatic characteristics.
For a curve plotted on an xy coordinate system, the area under the curve and moments, second moments (moments of inertia), and location of
the centroid can be expressed as simple integrals. Since hull forms are seldom definable by mathematical equations, areas, moments, and
volumes are calculated by manual integration methods rather than by direct integration. Manual integration methods are also used to evaluate
any parameter that can be expressed as a curve of a function of some variable. For example, the total force, location of the center of effort,
and force moment of an unevenly distributed force (such as current forces) can be determined from a curve showing the force distribution.
Graphical and numerical manual integration methods are described in the following paragraphs.
1-4.1 Graphical Integration. An obvious way to calculate the area under a curve (or within a shape) is to plot the curve to scale on graph
paper and count the squares under the curve. This method can be extended to calculate the first moment of area, My = xy dx, by multiplying
the height (number of squares, y) in each column by its distance from the origin (x), and summing all such products. In the same way, the
second moment is calculated by multiplying the height of each column by x2. By adopting sign conventions and adjusting the location of the
origin, moments can be calculated about any desired axis. Graphical integration of large, complex areas is very tedious, but can be very accurate
for even the most complex or discontinuous curves.
1-4.2 Numerical Integration. Numerical integration methods, or rules, are based on the same premise as graphical integration; that the area
under a curve can be closely approximated by breaking the area up into smaller shapes whose areas can be calculated or estimated easily, and
summing the areas of these shapes. Most rules depend upon the substitution of a simple mathematical form for the actual curve to be integrated.
The accuracy of the result depends upon the accuracy of the fit between the real and assumed curves.
1-4.3 Trapezoidal Rule. The trapezoidal
rule substitutes a series of straight lines for
a complex curve to allow integration of the
curve in a simple tabular format.
Conceptually, the trapezoidal rule is the
simplest of the numerical integration rules.

A curvilinear shape can be approximated by


a series of n trapezoids bounded by n + 1
equally spaced ordinates, y0, y1, y2, y3, ...,
yn, (at stations x0, x1, x2, x3, ..., xn) as
shown in Figure 1-5. If the station spacing
is h, the area (a0,1) of the first trapezoid is:
a0,1 =

y0 + y1
2

y1

y0

y2

y3 .........................

yn-1

yn
x

x1

x0

x3 .........................

x2

x n-1

xn

Figure 1-5. Curvilinear Figure Approximated by Series of Trapezoids.

The total area of the shape (A) is approximately equal to the sum of the areas of the trapezoids:
A = a0, 1
=
=

y0 y1
2
h
y
2 0

y
= h 0
2

a1, 2

a2, 3
y1 y2

2
2y1

y1

...

1, n

y2 y3

2y2

y2

an

2 y3 . . .

y3

...

...

yn

yn

yn

yn

This expression is called the trapezoidal rule, and can be used to calculate areas of any shape bounded by a continuous curve, simply by dividing
the shape into a number of equal sections and substituting the ordinate values and the station spacing, or common interval, into the rule. The
common multiplier for the trapezoidal rule is the common interval (h). If the common interval and common multiplier (CM) are separated into
two factors, the common multiplier for the trapezoidal rule is 1.
The factors by which each ordinate is multiplied (1 2, 1, 1, 1, ..., 1 2) are the individual multipliers (m). The products of the individual multipliers
and ordinates are called functions of area, (A). The area under the curve is thus expressed as:
A = y dx = h f (A)

1-16

S0300-A8-HBK-010

Because the trapezoidal rule substitutes a series of straight lines for the curve to be integrated, it is best suited for use with smooth, long-radius
curves such as the waterlines of a ship. The rule underestimates the area under convex curves, and overestimates the area under concave curves.
Accuracy increases as station spacing is decreased. If greater accuracy is required in regions of considerable curvature, e.g. at the ends of the
ship, stations are taken at half-divisions. When half-spaced stations are used, the individual multipliers for the half-stations and adjacent stations
must be adjusted. If, for example, a half-station is inserted between ordinates 1 and 2:
y0 y1

A =

y1 y1.5 h
2
2

y1.5 y2 h
2
2

y2 y3
2

yn

h ...

yn

h
y 1.5 y1 y1.5 1.5 y2 2y3 ... yn
2 0

3
1
3
1
1
y1
y1.5
y2 y3 ...
y
= h y0
2
4
2
4
2 n

The individual multiplier for the half-station is 1 2, and 3 4 for the station on either side of it. A similar analysis will show that if several
sequential half-stations are inserted (i.e., 21 2, 31 2, 41 2, etc.) the multipliers for all stations and half-stations between the first and last half-stations
is 1 2, and the multiplier for the two outlying whole stations is 3 4. It may be more convenient to use the first form of the rule, to avoid divisors
greater than 2, in which case all the individual multipliers are doubled.
1-4.4 Simpsons Rules. The replacement
of a complex or small radius curve by a
series of straight lines limits the accuracy of
calculations, unless a large number of ordinates are used. Integration rules that replace the actual curve with a mathematical
curve of higher order are more accurate.
Simpsons rules assume that the actual curve
can be replaced by a second-order curve
(parabola). Figures 1-6 through 1-8 demonstrate the derivations of Simpsons rules.

y
2

Y = ax

+ bx + c

y0

y1

x=0

y2

x=1

x=2

1-4.4.1 Simpsons First Rule. Figure 1-6


shows a curve of the form y = ax2 + bx + c.
It is expressed by three evenly spaced
ordinates y0, y1 and y2, at x = 0, 1, and 2
(station spacing = 1). The values of the
ordinates are:

X
h
AREA = __ (y0 + 4y1 + y2 )
3

Figure 1-6. Simpsons Three-Ordinate Rule.

y0 = a (0)2

b (0)

c = c

y1 = a (1)2

b (1)

c = a

y2 = a (2)2

b (2)

c = 4a

for x = 1
b

c
2b

for x = 1
c

for x = 2

The area under the curve is:


2
A = (ax 2
0

bx

ax 3
3

c) dx =

bx 2
2

cx

2
0

8
a
3

2b

2c

Now c = y0 and y1 = y0 + a + b, and y2 = y0 + 4a + 2b. Substituting and solving for a and b:


y2

2 y1 = y0
a =

( y2

b = y1

2b
2 y1

4a

2 y0

2b

2a =

y0

2a

y0 )

2
y0

a = y1

y0

(y2

2y1
2

y0)

3
y
2 0

y2
2

2 y1

1-17

S0300-A8-HBK-010

Area (A) is expressed as:


8
a
3

2b

2c =

= 2 y0

3 y0

y2

A =

1
y
3 0

4 y1

8 y2

2 y1
2

4
y
3 2

4 y1

y0

8
y
3 1

3
y0
2
2
4
y =
3 0

1
y
3 0

2 y1

y2
2

4
y
3 1

2 y0
1
y
3 2

y2

For an ordinate spacing of h rather than unity:


h
(y
3 0

A =
This relationship is Simpsons first rule, or
3-ordinate rule, commonly called Simpsons
rule. The rule calculates correctly the area
under a second order curve and will
approximate the area under any curve that
passes through the same three points. The
accuracy depends on how closely the actual
curve approaches the parabolic form
assumed by the rule. Simpsons Rule is the
numerical integration rule used most widely
for ship calculations.

4 y1

1
1
2

y2 )

STATION

4
4

1
1
2

3-ORDINATE
MULTIPLIERS
SIMPSONS
MULTIPLIERS

Figure 1-7. Simpsons Multipliers for Long Curve.

The rule can be extended to calculate the area under a long nonparabolic curve such as a ships waterline. If the length of the curve is divided
into enough equal parts, as shown in Figure 1-7, it can be reasonably approximated by a series of parabolic segments. For a curve divided into
n equal parts, the area between the first (0) and third (2) ordinates would be given by:
A0

h
(y + 4y1 + y2)
3 0

where:
A0-2
h
L
n

=
=
=
=

area under the curve between the first and third ordinates
distance between ordinates = L/n
length of the curve
number of sections between ordinates = number of ordinates - 1

Similarly, the area between the third (2) and fifth (4) ordinates would be:
A2

h
(y + 4y3 + y4)
3 2

A4

h
(y + 4y5 + y6)
3 4

The area between the fifth (4) and seventh (6) ordinates:

and so on.
The total area is the sum of all the two section areas:
A = A0
=

A2

A4

... An

2 n

h
y 4y1 2y2 4y3 2y4 4y5 2y6 ... yn
3 0

This is the general form of Simpsons rule. Since the rule consists of a summation of areas over two sections of a curve divided into a number
of equal sections, the curve must be divided into an even number of sections (by an odd number of stations) to apply the rule. The common
multiplier (CM) is 1 3; the individual multipliers are 1, 4, 2, 4, 2, 4,..., 2, 4, 1. The derivation of the individual multipliers as a tabular summation
of the 3-ordinate rule multipliers for each two adjacent sections is shown in Figure 1-7.

1-18

S0300-A8-HBK-010

In regions where the slope of the curve changes rapidly, the accuracy of the rule can be increased by inserting intermediate (half-spaced) stations.
When half-spaced stations are used, the individual multipliers are modified. For example, a half-station could be inserted at 21 2 were there a
rapid change in form between the third and fourth stations of the curve in Figure 1-7. The area between the first and second stations is
calculated as before:
A0

h
(y + 4y1 + y2)
3 0

With the insertion of the half-station (21 2), the 3-ordinate rule can be applied to the area between the third and fourth ordinates (A2-3), with an
ordinate spacing of h/2:

A2

y
h y
2
=
y2 4y2.5 y3 = 2 2y2.5 3
32
2
3

The area between the fourth and sixth stations (A3-4) is now:
h
(y + 4y4 + y5)
3 3

A3

A3

... An

and so on. The total area is:


A = A0

A2

1 n

y2
2y3
h

2 y2.5 y
y3 4y4 y5 ... yn
y0 4y1 y2
3
2
2

h
1
1

y 4y1 1 y2 2y2.5 1 y3 4y4 2y5 ... yn


3 0
2
2

Note that unless another half-spaced station is inserted, the number of sections (n) will be even, and the rule unworkable. Intermediate stations
can be inserted at any equal division of the station spacing (third-stations, quarter-stations, etc.) and multipliers deduced in a similar manner.
Intermediate stations can be inserted anywhere along the length of the curve so long as two rules are followed:

An even number of intermediate stations must be inserted, so that the total number of segments remains even (total number of
ordinates is odd).
Intermediate stations must be inserted so there are an even number of segments in each group of consecutive whole or partial
segments (each group of whole or partial segments includes an odd number of ordinates).

Intermediate stations are commonly used


near the ends of waterlines where the hull
form changes rapidly with respect to length.
The individual multipliers can be quickly
determined by tabulating and summing the
appropriate 3-ordinate rule multipliers as
shown in Figure 1-8.
1-4.4.2 Simpsons Second Rule. Rules
can be deduced, in a similar manner, for
areas bounded by different numbers of evenly spaced ordinates, or by unevenly spaced
ordinates. For four evenly spaced ordinates:

5-1/2

2-1/2

0 STATION

1
1/2
1-1/2

1 3-ORDINATE

2
2

1
1/2
1-1/2

2
2

1
1/2
1-1/2

1/2
1/2

1 SIMPSONS

MULTIPLIER
MULTIPLIER

Figure 1-8. Simpsons Multipliers with Half-Spaced Stations.

A =

3h
(y0 + 3y1 + 3y2 + y3)
8

This is Simpsons second or three-eighths Rule. The general form is:


A =

3h
(y0 + 3y1 + 3y2 + 2y3 + 3y4 + 3y5 + 2y6 + ... + yn)
8

Simpsons second rule can be used with 4 + 3i ordinates, where i is a positive integer (i.e., 4, 7, 10, 13, etc.).
1-4.5 Applications. The derivations of Simpsons rules and the trapezoidal rule were demonstrated with area computations to aid
conceptualization, but the rules can integrate any function that can be plotted on Cartesian coordinates. If, for example, the ordinates represent
sectional areas along a ships length for a given waterline, the products of the multipliers and ordinates are functions of volume, (V), and their
summation (integral of the curve) is the volume of displacement. Calculation of areas, moments, centroids, and second moments of areas by
the are described in the following paragraphs.

1-19

S0300-A8-HBK-010

1-4.5.1 Moments and Centroids. As


shown in Figure 1-9, the moment of an
elemental strip of area about some vertical
axis YY is xydx. To determine the moment
of a larger area about the axis, the integral
M = xy dx must be evaluated. Instead of
multiplying the value of y at each station
by the appropriate multiplier, the value xy
is multiplied, where x is the distance from
the station to the reference axis, and dx is
the width of each strip, or the common
interval h. The value y dx = hyn is the area
of the strip an; the first moment of this area
about some reference axis YY is:

AREA a = ydx

yy

yn
1/2 yn

xx
dx
FOR SHADED STRIP: a = ydx
ay2 y3dx
i = ___ = ____
12
12
Myy = xa = x(ydx)
Iyy = x 2a = x2(ydx)
3
y2
y dx
y 2 ( ydx) _____
Ixx = __ a + i = __
+
2
2
12

MYY = xnhyn = xnan

Figure 1-9. Variables for Moment and Second-Moment Calculations.

The total moment is the sum of the


moments of all the strips, that is, the
integral of the incremental moments along
the length:
L
MYY = xn an dx
0

The integral can be evaluated numerically:


x a dx =
n n

xn CMf (A) = CM xn f (A)

where:
CM
(A)
mn

=
=
=

common multiplier for the appropriate integration rule


function of area = mnyn
common multiplier for the appropriate rule and station

If the reference axis is chosen to fall on an ordinate station, then the moment arms have the common interval (h) as a common factor, i.e., xn
= snh, where xn is the moment arm and sn is the number of stations from the reference axis to station n. The factor h can be brought outside
the summation:
MYY = CMh sn(A)
The products of the number of stations from the reference axis and the functions of area, sn(A), are the functions of moment (M):
MYY = CMh (M)
The distance from the centroid of the shape to the reference axis (x) is the moment divided by the area:
x

1-20

MYY
A

CM h f (M)
=
CM f (A)

f (M)
h
f (A)

S0300-A8-HBK-010

The centroid of a symmetrical shape lies on the axis of symmetry, and its location can be defined by summing moments about a single axis
perpendicular to the axis of symmetry. To precisely locate the centroid of an asymmetrical shape, moments must be summed about another,
perpendicular, axis. The calculation can be performed by taking ordinates perpendicular to the first set and integrating with respect to y rather
than x. Moments about an axis XX can also be determined using y ordinates, but with slightly less accuracy. Referring again to Figure 1-9,
the moment about axis XX of the elemental strip dx is:
y
y2
y
MX X = a = y dx = dx
2
2
2
where y is the height of the strip, and a its area. The total moment is the integral of the incremental moments along the length, and the integral
can be evaluated numerically:
MX X

Ly
n
an dx =
= 0
2

yn
2

CM f (A)n =

CM yn f (A)n
2

The product of the y ordinate and the function of area for each segment can be defined as the function of moment about x, (MXX):
f MX X = y f (A) = y 2 mn
CM
2

MX X =

f MX X

where mn is the individual multiplier for the nth ordinate. The distance from the centroid of the shape to the axis XX (y) is the moment divided
by the area:

MX X
A

CM
f MXX
f MXX
2
=
CM f (A)
2 f (A)

Moments can be summed about any axis, although it is simplest to sum them about an axis through x0 so that the number of stations from the
reference axis is simply the station number. For ship calculations, moments are often summed about the midships section to reduce the size
of the products and sums for manual calculation, and because the centers of flotation, buoyancy, and gravity normally lie near midships. When
moments are summed about a station other than an end station, a sign convention must be adopted so that distances to one side of the reference
axis (and therefore moments and functions of moments) are negative.
1-4.5.2 Second Moments of Area. The second moment of area (moment of inertia, I) of a plane shape about an axis YY parallel to the vertical
ordinates is given by:
IYY = 0L x2y dx
where:
IYY =
x
=
L =

second moment of area about some axis YY


distance from axis YY to elemental vertical strip of height y and width dx
length of the area whose second moment is desired, measured along an axis perpendicular to YY

An analysis similar to that taken for the calculation of first moments will show that the second moment of the area under a curve is calculated by:
IYY = CMh2 (IYY)
where:
CM
h
(IYY)
sn
mn
yn

=
=
=
=
=
=

common multiplier
common interval
function of second moment about axis YY = sn2mnyn
number of stations from axis YY to station n
individual multiplier for station n
height of the ordinate at station n

1-21

S0300-A8-HBK-010

The second moment of an area (moment of inertia) is always smallest about an axis through its centroid, (the neutral axis in bending stress analysis).
If moment of inertia about some axis YY, parallel to the neutral axis is known, the moment of inertia about the neutral axis (INA) is found by the
parallel axis theorem:
INA = IYY - Ad 2
where d is the distance from axis YY to the neutral axis, and A is the total area of the section.
The second moment of area about an axis XX perpendicular to axis YY can be calculated by taking ordinates perpendicular to the first set and
integrating twice with respect to y rather than x. To determine the second moment about a horizontal axis of symmetry, such as the moment
of inertia of a waterplane about its centerline, the integration can also be performed using the original set of ordinates. In Figure 1-9 (Page 120), y is the half-ordinate of an incremental strip of a waterplane measured from the centerline. The second moment of area of the incremental
strip about the centerline is:
y 2
ixx = a
2

y 2
i0 = y dx
2

1 3
1 3
y dx = y dx
3
12

where:
ixx
a
i0
dx

=
=
=
=
=

second moment of area of incremental strip about the centerline


area of the incremental strip
second moment of area of the incremental strip about a horizontal centroidal axis
(1 12)y3dx if strip is assumed to be rectangular
width of the incremental strip

The total second moment of half-waterplane area is:


L 1
1 L
IXX, half = y 3dx = y 3dx
0 3
3 0
The second moment of the total area is twice this amount, and this will be the second moment about the centerline, since the waterplane is
symmetrical about the centerline. The integration y3dx can be performed numerically:

CM h
IX X = 2

f IX X

where:
CM
h
(IXX)

=
=
=

mn

yn

common multiplier
common interval
function of second
moment about axis XX =
mnyn3
individual multiplier for
station n
height of the half-ordinate
at station n

1-4.5.3
Volumes and Centroids of
Volume. Volumes are calculated by integrating a curve of sectional areas. To calculate the volume of the tank shown in
Figure 1-10, the shape is first cut at several
stations to form section outlines. The area
of each section is calculated, and the areas
taken as ordinates along the length of the
tank. Integrating the area ordinates by the
trapezoidal rule:

y0
0

V = a dx = h (V)
where:
(V)
mn
an

1-22

=
=
=

function of volume = mnan


individual multiplier for station n
area of section at station n

4
y4 = 0

x0

ORDINATES
FOR AREA
INTEGRATION
ORDINATES
FOR VOLUME
INTEGRATION
(AREAS)

a0

x1

a1

x3

x2

a2

Figure 1-10. Determination of Volume by Numerical Integration.

a3

S0300-A8-HBK-010

The moment of volume about some axis YY is:


MYY = h2 (M)
where:
(M)
sn

=
=

function of moment of volume about axis YY = snmnan


number of stations from axis YY to station n

The distance of the centroid from axis YY:


d =

h 2 f (M)
=
h f (V)

f (M)
h
f (V)

These forms are exactly the same as those used to calculate areas and moments and centroids of areas; the only difference is that ordinate values
represent areas rather than linear distances. Integrations can be performed along additional axes to precisely locate the centroid of the shape.
1-4.5.4 General Forms for Area and Moment Calculations. Calculation of areas, moments, centroids, and second moments of area by
Simpsons first and second rules can be expressed in general forms:
A = (CM) h f (A)
MYY = (CM) h f (M)
CM
MXX =
f MXX
2
x =

(CM) h f (M)
=
(CM) f (A)

f (M)
h
f (A)

where:
A
MYY
MXX
=
x
y
IYY
IXX
CM
h
(A)
(M)
(MXX)
(IYY)
(IXX)
s
m
yn

= area under a curve between selected stations


= first moment of area about axis YY
first moment of area about axis XX
= distance from centroid of area to axis YY
= distance from centroid of area to axis XX
= second moment of area about axis YY
= second moment of area about centerline axis XX
= common multiplier for the appropriate rule (1, 1/3, 3/8, etc)
= common interval
= function of area
= mnyn
= function of moment about YY
= snmnyn = sn(A)
= function of moment about XX
= mnyn2
= yn(A)
= function of second moment about YY = sn2mnyn
= sn(M) = sn2(A)
3
= function of second moment about XX = mnyn
= number of stations from axis YY (or integration start point) to station n
= individual multiplier for station n for the appropriate rule
= height of the ordinate at station n (half-ordinate for IXX)

Examples 1-1 and 1-2 demonstrate the use of the trapezoidal rule and Simpsons rule to calculate waterplane functions for an FFG-7 Class ship.

1-23

S0300-A8-HBK-010

EXAMPLE 1-1
CALCULATION OF WATERPLANE PROPERTIES BY TRAPEZOIDAL RULE

Using 11- and 21-ordinate trapezoidal rules, calculate the waterplane area (AWP), location of the center of flotation (LCF), moment of inertia of the waterplane
about the centerline (ICL) and a transverse axis through the LCF (ICF), tons per inch immersion in saltwater (TPI), and waterplane coefficient (CWP) for the 16-foot
waterline of an FFG-7 Class ship. Compare these values with actual data.
Actual Properties:

L
Bmax
AWP
LCF

=
=
=
=

ICF
ICL
TPI
CWP =

408 ft
45.6 ft
13,860 ft2
24.1 ft aft of midships = 228.1 ft from forward perpendicular

= 135,888,480 ft4
= 1,664,145 ft4
= 33 tons/in
0.745

Since the waterplane is symmetrical about its centerline, areas and moments can be found by integrating one side of the waterplane along the centerline with
half-ordinates (halfbreadths) measured from the centerline, and doubling the results. Halfbreadths for the 16-foot waterline, in feet, inches, and eighths, are
taken from Figure FO-1. The integrations are best performed in a tabular format. To integrate on 11 ordinates, halfbreadths for stations 0, 2, 4, 6, 8, 10, 12,
14, 16, 18, and 20 are used.
Integration on 11 ordinates:
Station

Ordinate,
y
ft-in-1/8
ft
0-4-5
0.39

Integration on 21 ordinates:
Multiplier
m
1

(A)
my
ft2
0.19

Lever
(M)
(IYY)
s
s (A) s (M)
ft
ft3
ft4
0
0.0
0.0

(IXX)
m y3
ft4
0.03

6 -10 - 5

6.89

6.89

6.89

6.89

327.1

12-11 - 0

12.92

12.92

25.84

51.68

2156.7

17- 9 - 2

17.77

17.77

53.31

159.93

5611.3

20-11 - 5

20.97

20.97

83.88

335.52

9221.4

10

22- 7 - 1

22.59

22.59

112.95

564.75

11527.9

12

22- 8 - 3

22.70

22.70

136.20

817.20

11697.1

14

21- 8 - 4

21.71

21.71

151.97

1063.37

10232.4

16

19- 7 - 1

19.59

19.59

156.72

1253.76

7518.0

18

16- 8 - 6

16.73

16.73

150.57

1355.13

4682.6

20

12- 7 - 0

12.58

6.29

10

62.90

629.00

995.4

941.23

6237.65

63969.9

168.34

h
= 408/10
AWP = 2h (A)
MFP = 2h 2 (M)

=
=
=

40.8 ft
2(40.8)(168.34)
2(40.8)2(941.23)

(M)
= h
(A)

941.23
(40.8)
168.34

IFP
ICF

= 2h 3 (IYY)
= IFP - Ad 2

=
=

2(40.8)3(6237.65)
= 847,288,842 ft4
847,288,842 - 13,736.5(228.1)2 = 132,516,043 ft4

ICL = 2(h / 3) (IXX) =


TPI = AWP / 420
=
CWP = AWP / (LB)
=

= 13,736.5 ft2
= 3,133,618 ft3

= 228.1 ft from FP = LCF

2(40.8/3)(63,969.9)
13,736.5/420
13,736.5/(408 45.6)

= 1,739,981 ft4
= 32.7 tons
= 0.738

Station

0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

Ordinate,
Multiplier
y
m
ft-in-1/8
ft
0 - 4 - 5 0.39
1/2
3 - 7 - 6 3.65
1
6 -10 - 5 6.89
1
10- 0 - 2 10.02
1
12-11 - 0 12.92
1
15- 6 - 1 15.51
1
17- 9 - 2 17.77
1
19- 6 - 7 19.57
1
20-11 - 5 20.97
1
21-11 - 5 21.97
1
22- 7 - 1 22.59
1
22- 9 - 4 22.79
1
22- 8 - 3 22.70
1
22- 3 - 7 22.32
1
21- 8 - 4 21.71
1
20- 9 - 5 20.80
1
19- 7 - 1 19.59
1
18- 2 - 1 18.18
1
16- 8 - 6 16.73
1
15- 1 - 0 15.01
1
12- 7 - 0 12.58
1/2

h
= 408/20
AWP = 2h (A)
MFP = 2h2 (M)

(A) Lever
(M)
(IYY)
s
ft2
ft
ft3
ft4
0.19
0
0.0
0.0
3.65
1
3.65
3.65
6.89
2
13.78
27.56
10.02
3
30.06
90.18
12.92
4
51.68
206.72
15.51
5
77.55
387.75
17.77
6
106.62
639.72
19.57
7
136.99
958.93
20.97
8
167.76 1342.08
21.97
9
197.73 1779.57
22.59
10
225.90 2259.00
22.79
11
250.69 2757.59
22.70
12
272.40 3268.80
22.32
13
290.16 3772.08
21.71
14
303.94 4255.16
20.80
15
312.00 4680.00
19.59
16
313.44 5015.04
18.18
17
309.06 5254.02
16.73
18
301.14 5420.52
15.01
19
285.19 5418.61
6.29
20
125.80 2516.00
338.18
3775.54 50052.98

= 20.4 ft
= 2(20.4)(338.18)
= 2(20.4)2(3775.54)

AWP, ft2
LCF, ft fm FP
ICF, ft4
ICL, ft4
TPI, tons/in
CWP

1-24

13,860.0
228.1
135,888,480
1,664,145
33
0.745

ft4
0.03
48.6
327.1
1006.0
2156.7
3731.1
5611.3
7495.0
9221.4
10604.5
11527.9
11836.8
11697.1
11119.4
10232.4
8998.9
7518.0
6008.7
4682.6
3381.8
995.4
128200.7

= 13,797.5 ft2
= 3,142,457 ft3

(M)
= h
(A)

3775.54
= (20.4)
338.18

IFP
ICF

= 2h 3 (IYY)
= IFP - Ad 2

= 2(20.4)3(50,052.98)
= 849,865,964 ft4
= 849,865,964 - 13,797.6(227.8)2 = 134,155,856 ft4

= 227.8 ft from FP = LCF

ICL = 2(h / 3) (IXX) = 2(20.4 / 3)(128,200.7)


TPI = AWP / 420
= 13,797.6 / 420
CWP = AWP / (LB)
= 13,797.6 / (408 45.6)

= 1,743,529 ft4
= 32.9 tons
= 0.742

Comparison:
Actual

(IXX )

11 Ordinate
Value
Error, %
13,737.8
0.88
228.1
0.00
132,502,924
2.49
1,739,981
4.56
32.7
0.91
0.738
0.94

Value
13,797.500
227.800
134,155,856.000
1,743,529.000
32.900
0.742

21 Ordinate
Error, %
0.45
0.13
1.28
4.77
0.30
0.40

S0300-A8-HBK-010

EXAMPLE 1-2
CALCULATION OF WATERPLANE PROPERTIES BY SIMPSONS RULE
Use Simpsons first rule with 11 ordinates to calculate the waterplane properties that were calculated in Example 1-1. Compare the results with actual data
and the results by trapezoidal rule.
Ship dimensions and actual waterplane properties are the same as for Example 1-1. Halfbreadths for stations 0, 2, 4, 6, 8, 10, 12, 14, 16, 18, and 20 from
Figure FO-1 are used to integrate on 11 stations. Integration:
Station

Ordinate,

Multiplier

(A7)

Lever

(M)

(IYY)

(IXX)

my

s (A)

s (M)

m y3

ft-in-1/8

ft

ft2

ft

ft

ft4

ft4

0-4-5

0.39

0.39

0.0

0.0

0.06

6 -10 - 5

6.89

27.56

27.56

27.56

1308.3

12-11 - 0

12.92

25.84

51.68

103.36

4313.4

17- 9 - 2

17.77

71.08

213.24

639.72

22445.1

20-11 - 5

20.97

41.94

167.76

671.04

18442.7

10

22- 7 - 1

22.59

90.36

451.80

2259.00

46111.4

12

22- 8 - 3

22.70

45.40

272.40

1634.40

23394.2

14

21- 8 - 4

21.71

86.84

607.88

4255.16

40929.8

16

19- 7 - 1

19.59

39.18

313.44

2507.52

15036.0

18

16- 8 - 6

16.73

66.92

602.28

5420.52

18730.4

20

12- 7 - 0

12.58

12.58

10

125.80

1258.00

1990.9

2,833.84

18,776.28

192,702.4

508.09

h
AWP =
MFP =

IFP
ICF

=
=

ICL
TPI
CWP =

=
=

408/10
3 h (A)
2
3 h2 (M)

(M)
h
(A)

3 h 3 (IYY)
IFP - Ad2

3 (h/3) (IXX)
AWP/420
AWP/(LB)

=
=

40.8 ft
3 (40.8)(508.09)
2
3 (40.8)2(2833.84)
2

=
=

13,820.1 ft2
3,144,882 ft3

=
=

2833.84
(40.8)
508.09

2
3 (40.8)3(18,776.28)
850,156,311 - 13,820.1(227.6)2

2
3 (40.8/3)(192,702.4)
13,820.1/420
13,820.1/(408 45.6)

=
=

227.6 ft from FP

850,156,311 ft4
134,508,685 ft4

1,747,168 ft4
32.9 tons
0.743

LCF

Comparison:
Actual Value

11 Ordinate Simpsons Rule


Value

AWP, ft2

Trapezoidal Rule Error, %

Error, %

11 Ordinate

21 Ordinate

13,860

13,820.1

0.29

0.88

0.45

228.1

227.6

0.22

0.00

0.13

ICF, ft4

135,888,480

134,508,685

1.02

2.49

1.28

ICL, ft4

1,664,145

1,747,168

4.99

4.56

4.77

33

32.9

0.30

0.91

0.30

0.745

0.743

0.27

0.92

0.40

LCF, ft fm FP

TPI, tons/in
CWP

The accuracy of an 11-ordinate Simpsons rule compares favorably with that of a 21-ordinate trapezoidal rule. Simpsons rule with 21 ordinates
is only marginally more accurate than with 11 ordinates for this waterplane shape. Note that Simpsons rule calculates the moment of inertia
about the centerline with slightly less accuracy than the trapezoidal rule. The derivation of the form: ICL = (CM)(h/3) (IXX) assumes a constant
ordinate over the entire section (see Paragraph 1-4.3.3). The Simpsons multipliers do not correct for this assumption. The constant-ordinate
assumption is essentially correct for very full ships and barges with extensive parallel midbody, and will yield very accurate values for ICL.
Accuracy of ICL calculations for fine-lined ships can be increased only by using very close station spacing or integrating along an axis
perpendicular to the centerline. The 5 percent accuracy shown here should be sufficiently accurate for most salvage work.

1-25

S0300-A8-HBK-010

1-4.6 Other Simpsons Rule Forms. Simpsons rules can be derived for numbers of ordinates for which the first two rules do not apply, and
to determine areas of "left over" segments at the ends of curves.
1-4.6.1 5, 8, Minus One and 3, 10, Minus One Rules. An additional Simpsons rule, known as the 5, 8, minus one rule, is used to determine
the area between two ordinates when three consecutive ordinates are known. For ordinates y0, y1, and y2, the area between the first and second
ordinates is given by:
1
h (5y0 + 8y1 - y2)
A0-1 =
12
The area between the second and third ordinates can be found by applying the rule backwards:
1
h (-y0 + 8y1 + 5y2)
A1-2 =
12
The validity of the 5, 8, minus one rule can be verified by observing that the sum of the expressions for the two sectional areas is the 3-ordinate
rule:
1
h 5y0 8y1 y2
A = A0 1 A1 2 =
y0 8y1 5y2
12
1
h y0 4y1 y2
=
3
The 5, 8, minus one rule cannot be used for moments. The first moment of the area between the first and second ordinates (A1-2) about the first
ordinate is given by the 3, 10, minus one rule:
1 2
h (3y0 + 10y1 - y2)
M1 =
24
These two Simpsons rules are at times convenient, but are less accurate than the first and second rules.
1-4.6.2 Simpsons Rules for Any Number of Ordinates. Simpsons rules can be combined one with another to derive rules for numbers of
ordinates for which the first two rules do not apply. For example, the first rule can be used for 3, 5, 7, 9, ... ordinates, and the second rule for
4, 7, 10, .... ordinates. A rule can be deduced for six ordinates as shown below:
3
h y0 3y1 3y2 y3
A0 3 =
8
1
h y3 4y4 y5
A3 5 =
3
3
9
9
3
1
4
1
y1
y2
y3
y3
y4
y5
A
= A0 3 A3 5 = h y0
8
8
8
3
3
3
8
1
h 9y0 27y1 27y2 17y3 32y4 8y5
=
24
This is not the only rule suitable for six ordinates. By skillful use of the 5, 8, minus one rule, a rule with less awkward multipliers can be
deduced:
1
h 5y0 8y1 y2
A0 3 =
12
3
h y1 3y2 3y3 y4
A1 4 =
8
1
h y3 8y4 5y5
A4 5 =
12
A

= A0

A1

A4

5
25
25
25
25
5
y0
y1
y2
y3
y4
y5
= h
24
24
24
24
15
12
25
h 0.4y0 y1 y2 y3 y4 0.4 y5
=
24
Substituting the same values for ordinates y0 through y5 in each rule will verify that they are equivalent. Rules deduced in this manner can be
used in the general forms described in Paragraph 1-4.4.4.
1-4.7 Other Integration Rules. Simpsons rules and the trapezoidal rule are satisfactory for most manual calculations. The Newton-Cotes,
Tchebycheffs, and Gauss rules are more accurate, but require more tedious manual calculations. These rules are described in most general naval
architecture texts, such as Basic Ship Theory by K.J. Rawson and E.C. Tupper, or Muckles Naval Architecture by W. Muckle and D.A. Taylor.

1-26

S0300-A8-HBK-010

1-4.8 General Notes For Numerical Integration. The numerical integration rules presented have relative advantages and disadvantages. When
time and/or access to high-speed computers permits, the salvage engineer may select the optimum integration rule for a well-defined curve.
For curves where ordinates are tabulated for only certain stations, a rule appropriate to that number and spacing of stations must be adopted.
Some generalizations about the applicability of integration rules are listed below:

The trapezoidal rule uses constant ordinate spacing and simpler multipliers than the other rules. Any number of ordinates can be
used. The rule can accommodate half-stations at any point, and the multipliers for half-stations are easily derived. For a single
integration (area calculation) of a gentle curve, the trapezoidal rule is nearly as accurate as the Simpsons rules, but progressively
greater errors are introduced on successive integrations (for moments and moments of inertia).
Simpsons rules and the trapezoidal rule include the common interval as part of the common multiplier and can therefore calculate
areas or volumes, moments, centroids, and second moments of area (single, double, and triple integrations) directly.
Simpsons rules are the most commonly used integration rules because they are more accurate than the trapezoidal rule, but simpler
to use than the more accurate Newton-Cotes, Tchebycheffs, and Gauss rules.
Simpsons rules exactly integrate first-, second-, and third-order curves. Successive integrations produce progressively higher order
curves: the curve of area under a second-order curve is a third order curve, and the curve of the moment of areas is then a fourthorder curve. Simpsons rules will therefore exactly calculate the first moment of a second-order curve, or the second moment of
a first-order curve. Calculating the second moment of a second-order or higher curve involves integrating a fourth-order equation,
so some error is introduced even for a parabolic curve. Additional error may arise for an arbitrary curve. Experience has shown
that Simpsons rule calculates moments and second moments of relatively smooth, continuous curvessuch as those describing
ship formsaccurately if a sufficiently close station spacing is used.
An even-ordinate Simpson rule is only marginally more accurate than the next lower odd-ordinate rule; odd-ordinate Simpson rules
are therefore preferred, and almost universally used in salvage.

1-4.9 Integration of Discontinuous Curves. The integration rules discussed are applicable to continuous curves. The area under a
discontinuous curve can be obtained by applying appropriate rules to the portions of the curve between discontinuities and summing the areas.
For curves with large numbers of closely spaced discontinuities, it is simpler to divide the curve into segments at the discontinuities, approximate
each segment by a rectangle, triangle, or trapezoid, calculate the area of each segment, and sum the areas to find the total area. The centroid
of each segment can be calculated or estimated. Moments, second moments, and the centroid of the entire area can be calculated by summing
the products of each area and the lever arm from its centroid to a selected axis in a tabular format. Replacing a segment of the curve between
discontinuities (stations) with a horizontal line at a value equal to the average ordinate creates a rectangle with area equal to the area under the
curve between the two stations. If the curve between stations can be reasonably approximated by a straight line, a horizontal line intersecting
the curve midway between stations has a y value equal to the average ordinate. Repeating this process along the length of the curve creates
a stepped curve. If the discontinuities, and subsequent stations, are evenly spaced, the curve can be integrated by a modification of the
trapezoidal rule:
A = y dx = h

n
1 n

MYY = xy dz = h 2

IYY = x 2y dx = h 3

n
1
n
1

sn 1/2 yn

sn 1/2 2 yn

where:
A
MYY
IYY
h
sn
yn

=
=
=
=
=
=

area under a curve between stations 0 and n


first moment of area about axis YY
second moment of area about axis YY
common interval
number of stations from axis YY (or integration start point) to station n
height of the mid-ordinate between stations n and n-1

Weight distribution curves for ships are usually drawn assuming a constant weight distribution between stations as stepped curves. The addition
of the continuous buoyancy curve and stepped weight curve creates a discontinuous load curve. The load curve is usually stepped as described
above to facilitate integration along its length to define the shear curve. Alternatively, the buoyancy curve can be stepped before summing with
the weight curve. A stepped 10-segment (11-ordinate) buoyancy curve can be constructed from standard Navy 21-station Bonjeans Curves by
taking unit buoyancy calculated from section areas for odd station as the average unit buoyancy for segments bounded by even stationsunit
buoyancy for segment 02 is based on section area for station 1, that for segment 24 on the area for station 3, etc. Example 1-4 includes an
integration of this type.

1-27

S0300-A8-HBK-010

1-4.10 Calculation of Hull Properties. Various integrations of a ships hull form are used to determine properties such as displacement,
locations of centers, tons per inch immersion, etc., known collectively as functions of form, hydrostatic functions, or hydrostatic data. Waterlines,
buttocks, and stations of lines drawings are spaced to support numerical integration, usually by Simpsons or the trapezoidal rules. Halfbreadths
(offsets) taken along the length of a waterline provide ordinate values to define the waterplane shape; halfbreadths taken at different waterlines
at the same station provide ordinate values to define the station shape. Because ships are symmetrical about the centerline, integrations are
customarily performed for one side of the section or waterplane only, and doubled to give the total area or moment.
When working from offsets, sectional areas are usually calculated by vertical integration on horizontal ordinates from the centerline. An
integration up to a waterline gives section area corresponding to that waterline. Integrating the curve of areas along the ships length gives
volume of displacement; the centroid of the volume is the center of buoyancy.
Waterlines are integrated along the ships length to determine area of the waterplane, location of the centroid of the waterplane (center of
flotation), and moment of inertia of the waterplane about the centerline and about a transverse axis through the center of flotation. From these
properties, tons per inch immersion, location of the metacenter, etc., can be calculated. Displacement volume can be calculated by taking
waterplane areas as ordinates and integrating vertically.
Longitudinal position of the center of buoyancy (LCB) is obtained by longitudinal integration of the sectional areas. Height of the center of
buoyancy (KB) can be obtained by vertical integration of waterplane areas, or by calculating a vertical moment of area for each section. The sum
of all the vertical area moments divided by the sum of the sectional areas gives KB. Integrations of this form are included in Example 1-4 and
Appendix F.
1-4.10.1 Functions of Form. Functions of hull form are
usually calculated for each waterline so they can be plotted
as a function of draft as the ships Curves of Form, also
called Hydrostatic Curves, or Displacement and Other
Curves (D & O Curves). Figure FO-2 is a reproduction of
the curves of form for an FFG-7 Class ship. Hydrostatic data
is also recorded in the Functions of Form Diagram (Figure
B-1) for Navy ships and Hydrostatic Tables (Figure B-2) for
commercial vessels. The salvage engineer may be required
to calculate hydrostatic data when curves of form or other
documents are not available or for a casualty in an unusual
condition. Whether functions of form are calculated for a
complete range of drafts or for only a few selected drafts depends on the form of the ship and the nature of information
required by salvors. Manual calculations are best performed
on organized tabular forms called displacement sheets.

Table 1-3. Appendage Allowances.


Ship Type

Appendage allowance:
APP/FL

Single-screw, small combatant with keel sonar dome1 . . . . . .


Twin-screw, small combatant with keel sonar dome1 . . . . . . .
Single-screw, small combatant with bow sonar dome1 . . . . . .
Twin-screw, small combatant with bow sonar dome1 . . . . . . .
Twin-screw amphibious warfare ships with well decks1 . . . . . .
shell plating only . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
all other appendages . . . . . . . . . . . . . . . . . . . . . . . . . .
Twin-screw LST1
without bow thruster . . . . . . . . . . . . . . . . . . . . . . . . . .
with tunnel bow thruster (negative appendage) . . . . . . .
Single-screw merchant ships and auxiliaries of ordinary form,
less than 5,000 tons full load displacement . . . . . . . . . . . . . .
shell plating only . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
all other appendages . . . . . . . . . . . . . . . . . . . . . . . . . .
Single-screw merchant ships and auxiliaries of ordinary form,
5,000 to 15,000 tons full load displacement . . . . . . . . . . . . .
shell plating only . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
all other appendages . . . . . . . . . . . . . . . . . . . . . . . . . .
Single-screw merchant ships and auxiliaries of ordinary form,
greater than 15,000 tons full load displacement . . . . . . . . . . .
Twin-screw merchant ships and auxiliaries of ordinary form . .
shell plating only . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
all other appendages . . . . . . . . . . . . . . . . . . . . . . . . . .
VLCC, ULCC, very large bulk carriers . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

0.0167
0.0200
0.0049
0.0060
0.0106
0.0057
0.0049

...
...

0.0024
0.0014

...

0.0075

1-4.10.2 Appendage Displacement. Volumes and dis. . . 0.0060


. . . 0.0015
placements (buoyancies) based on section areas taken from
Bonjeans Curves do not include appendage volume/ dis. . . 0.0050
placement, although sectional areas from some Bonjeans
. . . 0.0040
Curves include shell plating. If known, appendage dis. . . 0.0010
placements can be added to the integrated displacement; effect on LCB can be determined by moment balance. When
. . . 0.0025
appendage buoyancy is unknown, appendage displacement
. . . 0.0081
can be estimated as a fraction of full load displacement,
. . . 0.0035
. . . 0.0046
called an appendage allowance. Appendage allowances
. . . 0.0015
vary with ship size, type, and configuration. Warships
generally have more and larger appendages than auxiliaries
or commercial vessels. Vessels with high power-to-size
Source: 1Jamestown Marine Services, 1990, unpublished; based on data from 22
ratios have larger screws and rudders than lower powered
hull types entered into ship data files for the NAVSEA POSSE Program
vessels; appendage allowance increases with the number of
screws. Large bow sonar domes on combatants are faired
into the hull, and are included in Bonjeans Curves and offsets; keel-mounted domes are appendages. For a given ship type and configuration,
appendage allowance generally increases as size decreases. Approximate appendage allowances for different ship types are given in Table 1-3.
Appendage displacement is essentially constant with draft, as most appendages (except shell plating) are low on the hull and will be emerged
only by extremely low drafts. Once determined, appendage displacement can be added to the integrated displacement for any draft that covers
the appendages to determine total displacement. Shell plating displacement can be adjusted for drafts less than full load by assuming that onehalf of the shell plating volume is concentrated in the bottom third of the draft range, and the remaining volume is evenly distributed over the
upper two-thirds of the draft range. It is usually safe to assume that LCB for the displacement with appendages is virtually the same as that
for the integrated (without appendages) displacement.
1-4.10.3 Station Spacing. In full-bodied ships (low-speed general cargo, large tankers, bulk carriers, etc.) the lengths of the waterlines between
stations in the midbody are nearly straight lines. In many modern full-bodied ships, the waterlines over the midbody are, in fact, straight lines,
forming a parallel midbody. Integration on 10 equal divisions of length (11 stations, 0-10) is sufficiently accurate for most purposes. If the
curvature of the waterlines increases sharply near the ends of the ship, half-spaced stations can be inserted to increase accuracy, for example,
at stations 1 2, 11 2, 81 2 and 91 2.

1-28

S0300-A8-HBK-010

Accuracy can be increased by reducing the station spacing throughout the length of the curve. This increases the number of calculations to be
performed, but avoids determining additional multipliers and may be simpler to program for computer calculation. For ship calculations, offsets
are usually tabulated for either 11 or 21 basic stations (10 or 20 equal divisions), with half-stations as necessary. Offsets for Navy ships are
normally tabulated for 21 basic stations, although additional tables may be prepared for very close station spacing. Offset tables for 2-foot station
spacing are available for the FFG-7, for example. Even when 21-station offset tables or Bonjeans Curves are available, integration on 11
stations is sufficiently accurate for most hull volume calculations on any smooth hull form, including fine-lined warships.
1-4.10.4 Full Sections. In full, relatively
flat-bottomed sections, special care must be
taken in calculating the area from the base
to the lowest waterline to avoid error.
Figure 1-11 shows a section near midships
where the turn of the bilge fairs into a
straight line (the rise of floor line) at point
A.
If the entire area below CD is
calculated using horizontal ordinates from
the centerline, very close ordinate spacing
must be used to avoid error because of the
rapid change of form in the shell line. The
area below CD can be calculated accurately
using vertical ordinates from CD, with halfspaced ordinates inserted near the outboard
end, or by dividing the area into two
segments, as shown. The area KABC is a
trapezoid whose area can be calculated
accurately when the position of A and rise
of floor can be determined. The area ADB
can be obtained by using Simpsons rule,
either with horizontal ordinates measured
from AB, or with vertical ordinates
measured from BD.

CL

C
K

Figure 1-11. Calculating Sectional Area Below the Lowest Waterline.

1-4.10.5 Lowest Waterlines. When displacement volume is calculated by vertical integration of waterplane areas, the volume under the lowest
one or two waterlines is calculated separately. Since the form of the ship changes so rapidly near the keel, the volume under the lowest one
or two waterlines is calculated by integrating sectional areas along the ships length. This volume is added to the volume determined by
integrating waterplane areas from the lower waterlines upward to obtain the total volume of displacement.
1-4.10.6 Ends of Full Hull Forms. On
SIMPSONS RULE
very full hulls, such as spoon-bowed
ASSUMED PARABOLIC
barges, large tankers (VLCC, ULCC), and
FORM
bulk carriers, the parallel midbody extends
WATERPLANE
nearly to the ends of the ship, where it
OUTLINE
joins to a short forebody or afterbody with
steep or sharply curving lines. The aft ends
of the lower waterlines of many fine-lined
ships also curve sharply. If the ordinate
adjacent to the end ordinate is some
2
1
FP STATIONS
distance away from the end of the parallel
midbody, the curve from this ordinate to
TRAPEZOIDAL RULE
ASSUMED STRAIGHT LINE
the end ordinate (which is 0 or very small)
assumed by Simpsons rules or the
trapezoidal rule will fall well inside the
Figure 1-12. Inherent Integration Error in Full Waterlines.
actual waterline as shown in Figure 1-12.
This will cause a serious underestimation of
area for the end sections that will lead to even greater errors in calculations of moments and second moments about axes near midships because
of the long lever arms. Intermediate stations should be inserted so that there are ordinates near the ends of the parallel midbody and at least
one or two ordinates in the forebody and afterbody. Alternatively, waterplane areas for the midbody, forebody, and afterbody can be calculated
separately and summed. The midbody area can be treated as a rectangle or integrated by a 3-ordinate Simpson or trapezoidal rule; the midbody
and forebody areas can be calculated by any convenient rule with appropriate ordinates.
1-4.10.7 Tank and Compartment Volumes. A compartments molded volume is greater than its floodable volume (the volume of liquid that
can be contained), because of the volume occupied by fittings and structure. Floodable volumes of filled holds, machinery spaces, living spaces,
etc., are estimated from molded volumes by use of permeability factors, as explained in Paragraph 1-9.1.1. Framing, sounding tubes, sea chests
and similar structures in ordinary skin tanks typically occupy about 21 4 to 21 2 percent of the molded volume in double-bottom tanks, about 1
percent in cargo tanks (i.e., permeability of empty tanks is 971 2 to 973 4 percent, and 99 percent, respectively). Heating coils, if fitted, usually
occupy an additional 1 4 percent of the molded volume. Flush tanks lie entirely within the ships framing and are externally stiffened, so floodable volume, or capacity, is essentially equal to molded volume. To calculate volumes and centroids of flush tanks, offsets are taken to the inner
surface of the tank, rather than the hull molded surface. Bale capacity of holds is calculated from offsets taken from sections showing the line
of cargo battens, line of the bottoms of deck beams, and the top of the hold ceiling (above the inner bottom) including any gratings, with deductions for stanchions and other obstructions. Grain capacity is the molded volume, less the volume of structure, hold ceiling, and shifting boards.

1-29

S0300-A8-HBK-010

1-5 TRANSVERSE STABILITY


W

Transverse stability is the measure of a


ships ability to resist rotation about its
longitudinal axis and return to an upright
position after being disturbed by an upsetting force. The following paragraphs define
the elements of transverse stability and
provide methods to calculate the transverse
stability characteristics of a vessel.

CENTER OF
GRAVITY
W
W

(b)

1-30

(c)

(a)

1-5.1 Equilibrium and Stability. A ship


floating at rest, with or without list and
trim, is in static equilibrium; that is, the
forces of gravity and buoyancy are equal
and acting in opposite directions in line
with one another. Stability is the tendency
of a ship to return to its original position
when disturbed after the disturbing force is
removed. Stability can be described as
positive, negative, or neutral.
1-5.2 Internal Forces. The internal forces
affecting floating bodies are the forces of
gravity and buoyancy. Both of these forces
act at all times on wholly or partially
submerged bodies. Figure 1-13 illustrates
the relationship between the forces of
buoyancy and gravity. Assuming the prism
floats with half its volume submerged, and
with the center of gravity located as shown,
the prism can come to rest in either
position (a), with the center of gravity
directly above the center of buoyancy, or
(c), with the center of buoyancy above the
center of gravity. In either position, the
forces of buoyancy and gravity act along
the same vertical line. If the prism is
inclined from (a) to (b), or from (c) to (d),
a couple, or righting moment, is developed
between the lines of action of buoyancy
and gravity that tends to move the body
back to its original position, i.e., the body
floats with positive stability in either
position. In position (a), with the center of
gravity above the center of buoyancy,
stability is provided by the bodys shape, or
form, and is termed form stability. If the
width of the prism is reduced while the
center of gravity remains on the centerline
at the same location, a situation arises in
which the center of buoyancy does not
move far enough to be to the right of the
center of gravity as the body is inclined
from (a) to (b). The body can then attain
positive stability only in position (c), with
the center of buoyancy above the center of
gravity. Bodies floating with the center of
buoyancy above the center of gravity
develop positive initial righting moments
regardless of shape. This mode of stability
is called weight stability. Sailing yachts
with deep weighted keels, spar buoys,
conventional ships with very low centers of
gravity, and submarines all exhibit weight
stability. Capsized ships floating upside
down very often have their centers of
gravity below the center of buoyancy, and
operate in a weight stability mode.

(d)

Figure 1-13. Stability of a Floating Object.

20

Z
B

(a) INCREASING RIGHTING ARM (GZ)

45

37
G

B
B

(b) MAXIMUM RIGHTING ARM

80

(c) DECREASING RIGHTING ARM

61

(e) UPSETTING ARM (NEGATIVE GZ)

(d) RIGHTING ARM REDUCED


TO ZERO (GZ = 0)

Figure 1-14. Development and Loss of Righting Arm.

S0300-A8-HBK-010

The center of buoyancy of a ship moves as


the ship is inclined, in a manner that
depends on the shape of the hull near the
waterline. The center of buoyancy initially
moves away from the centerline as the ship
is inclined, as shown in Figure 1-14. At
some angle of inclination, the center of
buoyancy begins to move back towards a
vertical reference line drawn through the
original position of the center of buoyancy.
The vertical line of action of the center of
gravity continues to move outward as the
ship is inclined.
At some angle of
inclination, the line of action of gravity
moves outboard of the line of action of
buoyancy, creating an upsetting moment.
Ships that have slowly heeled through
progressively greater angles of inclination
will suddenly capsize when this angle of
zero righting moment (angle of vanishing
stability) is passed.
In Figure 1-15, the prism is assumed to be
neutrally buoyant so that it is wholly
submerged but clear of the bottom. An
inclination from (a) produces an upsetting
moment that tends to rotate the prism away
from its initial position. Conversely, a
inclination from (c) produces a righting
moment. A submerged object clear of the
bottom or other restraints can therefore
have positive stability in only one position,
that is, with the center of buoyancy above
the center of gravity. Submerged objects
therefore operate in a weight stability
mode. The difference in behavior of
floating and submerged objects is due to
the fact that the center of buoyancy of a
submerged object is fixed at the center of
volume of the object, while the center of
buoyancy of a floating object will generally
shift when the object is inclined. Because
the center of buoyancy of a submerged
object is fixed, the righting moment cannot
change to an upsetting moment as the
object inclines unless the position of the
center of gravity shifts.
Stability of
submarines and other submerged objects is
discussed more completely in the U.S. Navy
Ship Salvage Manual, Volume 4 (S0300A6-MAN-040).
Figure 1-16 shows how a stable ship
subjected to normal disturbances will
develop moments tending to return the ship
to its original position. A couple is formed
as the lines of action of the opposing forces
of gravity and buoyancy are separated. The
arm of this couple, called the righting arm,
is the lever to which the ships weight is
applied to right the ship. Figure 1-17
shows the upsetting arm developed when
unstable ships are disturbed.

B
B
(b)

(a)

(c)

(d)

Figure 1-15. Stability of a Submerged Object.

M
RIGHTING
MOMENT

W1

G
L

B1

L1

CL

Figure 1-16. Righting Arm (GZ).

UPSETTING
MOMENT

G
Z

W1
M

B1

L1

CL

Figure 1-17. Upsetting Arm.

1-31

S0300-A8-HBK-010

1-5.3 External Forces. Ships are inclined by various external forces:

Wave action,

Wind,

Collision,

Grounding,

Shifting of onboard weights, and

Addition or removal of weight.

Any inclination of a ship can be termed heel, but inclinations are broadly defined as heel, list, or roll depending on the duration and nature of
the forces causing the inclination.

Heel The term heel is specifically applied to noncyclic, transient inclinations caused by forces that may be removed or reversed
quickly. Such forces include wind pressure, centrifugal force in high-speed turns, large movable weights, etc.
List A list is a permanent, or long-term inclination, caused by forces such as grounding or offcenter weight that are not likely
to be removed suddenly.
Roll When an inclining force is suddenly removed, a ship does not simply return to its upright position, but inclines to the
opposite side and oscillates, or rolls, about its equilibrium position for some time before coming to rest. The natural rolling period
(period of roll assumed by a ship free of restraints and exciting forces) is a function of weight and buoyancy distribution. Rolling
is cyclic in nature and is induced or aggravated by short duration, repetitive or cyclic forces, such as wave forces.

1-5.4 Heights of Centers. The relative heights of the centers of gravity and buoyancy and the metacenter govern the magnitude and sense
of the moment arms developed as the ship inclines. They are, therefore, the primary indicators of a ships initial stability. Nominally, the
symbols KG, KB, and KM indicate the heights of the centers of gravity and buoyancy and the metacenter above the bottom of the keel, while
the symbols VCG and VCB indicate the vertical positions of the centers of gravity and buoyancy, measured from the baseline. In practice,
KG/KB and VCG/VCB are used almost interchangeably; in steel ships with flat plate keels, the difference in height above baseline and keel for
any point is generally less than two inches and is not significant.
1-5.4.1 Height of the Center of Gravity. The height or vertical
position of the center of gravity above the keel (KG or VCG) is defined
by weight distribution. KG can be varied considerably without change
of displacement by shifting weight up or down in the ship. Conversely,
it is possible to add or remove weight without altering KG. In most
ships, the center of gravity lies between six-tenths of the depth above
the keel and the main deck:

where:
D = hull depth, keel to main deck
For barges with raked or ship-shaped bows and cut-up sterns, lightship
KG can be estimated as 0.53D. For tank barges, KG for full load varies
little from the lightship value.
Table 1-4 gives very approximate values for the height of the center of
gravity for several types of merchant ships at lightship, and for some
naval ship types at full load. Calculation of KG can be a laborious and
time-consuming process, but ignorance of the height of a ships center
of gravity invites disaster. If the height of the ships center of gravity
is known for any condition of loading (lightship, for example), and the
location of added or removed weights is known, the new height of the
center of gravity can be calculated:
Wold KGold
Wold

Ship Type

w (kg)
w

Dry Cargo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0.68D

Passenger/Cargo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0.75D

Insulated Cargo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0.72D

Cross-Channel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0.68D

Oil Tanker . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0.69D

Naval ships (KG at full load)2:


Cruiser/Destroyer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0.55D

Frigate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0.61D

Amphibious Warfare without well decks (LST/LKA/LPH) . . . .

0.63D

Amphibious Warfare with well decks (LSD/LPD/LHA/LHD) . .

0.72D

Fleet replenishment (AE/AOE/AOR/AFS/AO) . . . . . . . . . . . .

0.62D

Tender/Repair Ship . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

0.5D

Source:
1
Applied Naval Architecture, R. Munro-Smith, 1967
2
Jamestown Marine Services, 1990

where:
KG
W
w
kg

1-32

=
=
=
=

KG
(D = depth
at midships)

Merchantmen (KG at lightship)1:

0.6D < KG < D

KGnew =

Table 1-4. Approximate KG.

height of the ships center of gravity, G, above the keel


total weight of the ship and contents
individual weights added (+) or removed (-)
height above keel of centers of gravity of added or removed weights, w

S0300-A8-HBK-010

Height of the center of gravity of cargo can generally be


obtained from the ships officers, usually the chief mate.
In the absence of better information, the design
estimations proposed by R. Munro-Smith (Applied Naval
Architecture, 1967) shown in Table 1-5 may be helpful.

Table 1-5. Approximate KG of Cargo in Full Holds.

1-5.4.2 Height of the Center of Buoyancy. The height


of the center of buoyancy above the keel (KB) is solely
a function of the shape of the underwater volume. As
the centroid of the underwater hull, the center of
buoyancy is lower in flat-bottomed, full-bodied ships,
such as tankers and ore carriers, than in finer lined ships
like destroyers or frigates. Disregarding changes in the
shape of the immersed hull due to trim and heel, KB of
any ship is a function of displacement, and therefore of
draft. The height of the center of buoyancy can be
calculated by summing incremental waterplane areas
(aWP) multiplied by their heights above the keel (z) and
dividing the result by the displacement volume ():
KB =

Hold/Space

KG of Cargo (D = depth of hold)

No. 1

0.7D + depth of double bottom

No. 2

0.7D + depth of double bottom

No. 3

0.7D + depth of double bottom

No. 4

0.7D + depth of double bottom

No. 5

0.7D + depth of double bottom

tween decks

height above keel to half depth of tween


deck at mid length of the space

Based on full holds (homogeneous cargo) in general cargo ship with machinery amidships,
three holds forward and two aft. In ships with extensive parallel midbody, it may be more
appropriate to apply the expression for hold No. 3 to all holds in the parallel midbody, with
the expression for No. 1 or No. 2 (depending on fineness of forebody) applied to the
forward most hold. A similar analysis should be applied to holds aft of the machinery
space, if any.

1
a z dz
wp

This expression can be evaluated by numerical integration methods if accurate drawings or offsets are available. In practice, KB can be
approximated with sufficient accuracy for salvage work as 0.52T for full-bodied ships and 0.58T for fine-lined ships. At very light drafts, KB
is closer to the given waterline because the lower waterlines are usually much finer than the waterlines in the normal draft range. As a vessels
underwater hull form approaches a rectangular prism (CB = 1.0), KB approaches 0.5T. The following empirical relationships give estimates for
KB that are very close to calculated values for merchant vessels of ordinary form at normal drafts:
KB =

1 5T
3 2

KB = Tm

AWP

AWP
AWP

(Morrishs Formula)

(Posdunines Formula)

Tm

where:
Tm

AWP =

= mean draft, [length]


= displacement volume, [length3]
waterplane area, [length2]

1-5.4.3 Metacentric Height. The transverse metacentric height (GMT), commonly called the metacentric height, of a ship is the vertical
separation of the center of gravity and the transverse metacenter (see Figure 1-4) and is a primary indicator of initial stability. A ship with a
positive metacentric height (G below M) will tend to right itself by developing righting arms as soon as an inclining force is applied. A ship
with a negative metacentric height (G above M) will list to either port or starboard with equal facility until the centers of buoyancy and gravity
are on the same vertical line, and thereafter develop positive righting arms. This condition, known as lolling, is a serious symptom of impaired
initial stability. Metacentric height is calculated by subtracting the height of the center of gravity from the height of the metacenter above the
keel:
GMT = KMT - KG
Transverse Metacentric Radius. The transverse metacentric radius (BMT) is the vertical distance between the center of buoyancy and the
metacenter. This distance is termed a radius because for small heel angles, the locus of successive centers of buoyancy approximates a circular
arc, with the transverse metacenter as its center. Metacentric radius is equal to the moment of inertia of the waterplane about its longitudinal
centerline (transverse moment of inertia, IT) divided by the underwater volume of the hull ():
BMT =

IT

1-33

S0300-A8-HBK-010
For a rectangular waterplane, IT = LB3/12, = LBT and:

BMT =

IT

LB 3
12
LBT
B2
12T

where:
L
B
T

=
=
=

length between perpendiculars, [length]


beam, [length]
mean draft, [length]

If the waterplane shape can be accurately defined, the moment of inertia can be determined by numerical integration. If not, the transverse
moment of inertia of most ships waterplanes can be approximated by:
IT CIT LB3
where CIT is the transverse inertia coefficient and is approximated by CWP2/11.7 or 0.125CWP - 0.045. These expressions for transverse inertia
coefficient are derived from the analysis of numerous ships, and are reasonable approximations for use in salvage for ships with CWP < 0.9.
For ships with CWP > 0.9, LB3/12 is a closer approximation of the transverse moment of inertia of the waterplane.
Height of the Metacenter. The height of the metacenter above the keel is calculated by adding the metacentric radius to the height of the center
of buoyancy above the keel:
KM = KB + BM

GM = KB + BM

KG

When denoting transverse metacenter, BM, KM, and GM, the subscript "T" is often omitted as understood.
Ships with large GM develop large initial righting arms and therefore respond to moderate disturbing forces with sharp, short-period rolling.
These ships are said to be stiff. Ships with smaller metacentric heights develop smaller initial righting arms and roll more gently in a seaway.
Ships with small metacentric heights are said to be tender. Insufficient initial stability results in constant rolling in even gentle seas, making
work difficult, and may allow extreme rolling in heavier seas, perhaps causing the ship to take on water or capsize. Excessive initial stability,
or stiffness, is also undesirable because it produces an uncomfortable ride, reduces personnel effectiveness, increases requirements on weapons
stabilization systems, increases lateral acceleration loads on topside cargo and equipment, and increases hull stresses. These matters usually
do not concern the salvage engineer, but very stiff rolling of a casualty under tow may damage sensitive equipment, loosen patches, or place
excessive loads on damaged structure. The term seakindly is used to describe a ship whose metacentric height is great enough to give adequate
stability, but not large enough to cause excessive stiffness.
The natural rolling period is a function of weight and buoyancy distribution and can be expressed as a function of GM and transverse radius
of gyration (k):
TR =

2k
g GM

where:
=
=
=
GM =
g =

TR
k

1-34

natural rolling period, seconds


transverse radius of gyration of the ship mass, [length]
0.4 to 0.5 times the beam, depending on depth and transverse weight distribution
transverse metacentric height, [length]
acceleration due to gravity, [length/sec2]

S0300-A8-HBK-010

If GM and k are expressed in feet, and g is taken as 32.174 ft/sec2, the rolling period formula reduces to:
TR =

1.108 k
GM

and:
2

k
GM = 1.108

TR

If the natural rolling period is known, GM can be estimated. Taking radius of gyration k as beam (B) multiplied by a coefficient (C), a
conservative estimate of GM can be made:
GM

CB

TR

The coefficient C can be taken as 0.4 to 0.5 for naval surface ships (0.44 average), 0.4 to 0.45 for submarine hulls based on bodies of revolution,
and 0.32 to 0.37 for other submarines. Ships and Marine Engines, Volume IV, The Design of Merchant Ships (Schokker et al, 1953) gives some
experimentally derived values for commercial vessels: 0.425 for large cargo and passenger liners, 0.385 for smaller passenger liners, 0.390 for
a loaded passenger liner, and 0.405 for an ore ship in ballast. This same text references Laursens possibly more correct approach of expressing
radius of gyration as a function of both beam and depth:
k = C B2 + D2
where the constant C ranges from 0.35 to 0.39 for cargo ships of ordinary form.
The rolling period formula will not give an accurate estimate of GM for a ship rolling in a seaway because the rolling period is modified by
wave and wind forces. Significant changes in GM will be reflected by marked changes in rolling period; increased rolling period is a sign of
deteriorating stability. An empirically derived relationship holds that stability is adequate when:
TR 2 B
where:
B = beam, ft
1-5.5 Righting Arm. At equilibrium, the forces of gravity and buoyancy act equally in opposition along the vertical centerline. As the center
of buoyancy shifts with a heel, the two opposing forces act along separate and parallel lines. The forces establish the couple which tends to
return a stable ship to the upright position. The distance GZ between the lines of action of the center of gravity and the center of buoyancy,
as shown in Figure 1-16, is the righting arm. The sine of the angle of inclination () is the ratio of GZ to GM.
sin =

GZ
GM

GZ = GM sin

This relationship applies for heel angles so small that the waterplane shape is not appreciably changed, usually taken as less than 10 degrees
for wall-sided ships and 7 degrees for fine-lined ships. At greater angles of heel, the metacenter moves away from the centerline and the
relationship between GZ and GM no longer applies.
1-5.6 Righting Moment. The force applied to a righting arm (GZ) is the ships weight. The righting moment (RM) developed at any angle
of heel is given by:
RM = W GZ
At any angle of heel, the stability of the ship is measured by the righting moment developed. Since the righting moment is equal to the righting
arm times displacement and displacement normally remains constant as the ship heels, the righting arm may also be used to measure stability
for a given condition of loading. This assumption lends itself to the use of the cross curves of stability as discussed in Paragraph 1-5.9.

1-35

S0300-A8-HBK-010

1-5.7 Change of Displacement. Any change of displacement will affect the righting moments developed by the ship. An increased
displacement increases W in the expression RM = W GZ, but also affects GZ by:

Increasing draft and thereby KB.

Increasing , thereby reducing BM as I will not change significantly (BM = I/).

The height of the metacenter is normally reduced as displacement increases because the increase in KB is usually less than the reduction in BM.
The opposite effects will be noted when displacement is decreased. Additionally, the location of the added weight will affect the location of
the center of gravity and therefore GM and GZ. These effects are simultaneous but not normally compensatory. The net effect of a change
in displacement may be either an increase or a decrease in righting moments. In general, the addition of low weight or removal of high weight
will increase stability, but each change of displacement must be carefully analyzed to determine its exact effect.
1-5.8 List. List, a long-term inclination of
the ship to one side or the other, is caused
by:

DD

CV

ARS

AOR

LCC

RO/RO

Offcenter weight.

Negative GM.

6
5
4
3

A combination of offcenter
weight and negative GM.

2
1

A list caused by offcenter weight is


identified by the ships tendency to return
to its listing condition when an external
force is applied temporarily and then
removed. A list caused by negative GM is
identified by the ships tendency to loll, or
list to either side with equal facility, when
disturbed. A list caused by a combination
of offcenter weight and negative GM is
identified by the ships tendency to list with
equal facility to either side, but with a
greater degree of list to one side. Negative
GM is the most serious condition that
causes a list and should be corrected first.
Paragraph 1-9.4 discusses the effects of
negative GM in greater detail.

RIGHTING ARMS, FEET

Before attempting to correct a list on a


ship, the cause must be determined.
Inappropriate corrective measures will only
aggravate the situation.

7
6
5
4
3
2
1

8
7
6
5
4
3

1-5.9 The Stability Curve. The righting


2
arm GZ is the distance between the lines of
1
action of buoyancy and gravity at any
0
angle of heel. Since the expression GZ =
0 10 20 30 40 50 60 70 80 90
0 10 20 30 40 50 60 70 80 90
GM sin cannot be used at larger angles of
INCLINATION, DEGREES
heel, the righting arm for a given heel
angle is determined by accurately locating
Figure 1-18. Typical Stability Curves.
the centers of gravity and buoyancy, and
measuring the separation between their
lines of action. If movable weights within the ship can be neglected, the center of gravity can be assumed to be fixed. As the ship heels, the
center of buoyancy will move to the new center of the underwater volume, which can be determined by numerical integration or graphical means.
As a ship heels, it also changes its trim to some extent to maintain constant displacement. This small change in trim can usually be disregarded
when calculating righting arms. Centers of buoyancy for various inclinations, and the resulting righting arms are determined by numerical
integration. These computations can be shortened somewhat by the methods described in Paragraph 1-5.11. A plot of righting arm against heel
angle is variously called a curve of statical stability, stability curve, righting arm curve, or GZ curve. Figure 1-18 shows typical stability curves
for various ship types.

1-36

S0300-A8-HBK-010

1-5.9.1 Cross Curves of Stability. As a


ships displacement is variable, the
designers prepare stability curves for a
range of displacements. It is customary to
plot righting arm values against
displacement for each of a number of
angles of inclination to create a group of
curves known as cross curves of stability.
By entering the cross curves with the
displacement of the ship and reading the
righting arms for each angle of heel, a
stability curve for any displacement can be
developed. Since height of the center of
gravity varies with loading, an assumed
position of the center of gravity was used
by the designer to develop the cross curves
of stability. Once the stability curve has
been corrected for the true location of the
center of gravity, the following stability
data can be obtained:

SHIP WATERTIGHT TO MAIN DECK


CENTER OF GRAVITY ASSUMED 19.00 ABOVE
BOTTOM OF KEEL AMIDSHIPS
3

RIGHTING ARM - FEET

45
60

70
2
30

20
1
10

3,000

4,000

DISPLACEMENT IN SALTWATER - TONS

Range of stability.
Figure 1-19. FFG-7 Class Cross Curves of Stability.

Righting arm and moment at


any angle of inclination.
Maximum righting arm and
moment.

Angle of the maximum righting arm and moment.

Metacentric height.
Angle of deck edge immersion.

The following examples use the FFG-7


Class cross curves of stability from Figure
1-19 to develop the initial and corrected
stability curves. Figure 1-20 is the stability
curve as taken from the cross curves for a
displacement of 3,200 tons.

KG ASSUMED AT 19
DISPLACEMENT = 3200 TONS
RIGHTING ARMS IN FEET

10

20

30

40

50

60

70

80

90

DEGREES OF INCLINATION
RANGE OF STABILITY

Figure 1-20. Statical Stability Curve.

1-37

S0300-A8-HBK-010

1-5.9.2 Correction for Actual KG. If the


actual center of gravity lies above the
assumed center of gravity, the metacentric
height is decreased and the ship is less
stable; conversely, if the actual center of
gravity is below the assumed center, the
metacentric height is increased and the ship
is more stable.

G1 Z1 = GZ - G G1SIN0

G2 Z2 = GZ + G G2SIN0
0
Z1
Z

Figure 1-21 shows that the actual righting


arm, GnZn is equal to the assumed righting
arm plus or minus the vertical distance
between the actual and assumed KG,
multiplied by the sine of the angle of heel:

W1

Z2

1-5.9.3 Range of Stability. The range of


stabilitythe range of inclinations through
The new stability curve is again the difference between the two curves.which the
ship develops positive righting armsis indicated by the intersections of the stability
curve with the horizontal axis. For the
corrected stability curve in Figure 1-22, the
range of stability is from 0 to 75 degrees.

0
L
L1

CL

Figure 1-21. Assumed KG for Stability Curve.

RIGHTING ARMS - FEET

GZ for KG = 191, = 3200 TONS

GG1sin = 2sin
2

LOSS IN RIGHTING ARMS


DUE TO RISE IN G

10

20

30
40
50
60
DEGREES OF INCLINATION

70

80

90

5
KG = 21
DISPLACEMENT = 3200 TONS
RIGHTING ARMS - FEET

The assumed KG is sometimes called pole


height. It is a common practice, especially
with European designers, to develop cross
curves based on an assumed pole height of
zero. Since the assumed position of the
center of gravity coincides with the keel,
the resulting cross curves are termed KN
curves.

G
G2

The actual, or corrected, stability curve can


be constructed graphically as a sine curve
correction.

If the actual height of the center of gravity


is less than the assumed height, the
correction curve is plotted below the
horizontal axis.

GnZn = GZ GGn sin

The GGn sin curve is plotted to the same


scale as the curve of statical stability as
shown in Figure 1-22. The ordinates of the
corrected curve are the differences between
the ordinates of the two curves and can be
picked off and plotted using dividers, as
shown, or determined by tabular
calculation.

G1

G1Z1 = GZ - GG1 sin


1

10

20

30

40

50

60

70

80

90

DEGREES OF INCLINATION

RANGE OF STABILITY

Figure 1-22. Correction to Stability Curve, G Two Feet Higher Than Assumed.

1-38

S0300-A8-HBK-010

1-5.9.4 Righting Arm and Righting Moment. The righting arm at any inclination is read directly from the curve. Because each stability curve
applies only to a specific displacement and KG, the righting moment can be obtained directly for any angle by multiplying the righting arm by
the displacement. Maximum righting arm, maximum righting moment, and angle of maximum righting moment can be determined by inspection
of the stability curve. From the corrected stability curve in Figure 1-22, maximum righting arm is approximately 1.1 feet at 51 degrees of
inclination, giving a maximum righting moment of 3,520 foot-tons (1.1 ft 3,200 tons). Maximum righting arm and the angle at which it occurs
are important parameters when an upsetting moment is applied gradually or statically. Once the upsetting moment exceeds the maximum righting
moment, the ship will list past the angle of maximum righting arm. If the upsetting moment is not immediately removed, the ship will capsize,
because as the ship heels to progressively greater angles, righting moment, already less than the upsetting moment, will steadily decrease. However, ships can, and do, safely roll past their angle of maximum righting arm in response to short-term or cyclic upsetting forces.
1-5.9.5 Metacentric Height. GM is the measure of the slope of the GZ curve at the origin. The metacentric height is equal to the height of
the intersection of a tangent to the statical stability curve at the origin with a perpendicular to the horizontal axis at 57.3 degrees (one radian).
Although metacentric height can be approximated from a stability curve by this means, it is more common that GM is known and the intercept
is sketched to help draw the initial part of the stability curve. The corrected stability curve in Figure 1-22 indicates a GM of approximately
1.2 feet.
1-5.9.6 Angle of Deck Edge Immersion. For most hull forms, an inflection point in the curve corresponds roughly to the angle of deck edge
immersion. This point is not necessarily at or near the angle of maximum righting arm. The inflection results from the abrupt change in the
shapes of the waterplane and underwater volume as the deck edge is immersed. The rate of increase in righting arm has changed from positive
to negativei.e., righting arms are still increasing, but at a slower rate. The angle of deck edge immersion varies along the length of the ship,
but lies within a relatively narrow range for the large midbody sections that have the greatest influence on the stability curve. The stability curve
in Figure 1-22 shows the angle of deck edge immersion to be about 38 degrees.
1-5.9.7 Righting Energy. The area under
the stability curve, (foot-degrees, meterradians), is a measure of the ships dynamic
stabilityits ability to absorb energy
imparted by winds, waves or other external
forces.
A ship with very little area
(righting energy) under its stability curve
could be rolled past its range of stability
and capsized by even a momentary
disturbance.
1-5.10 Effects of Hull Form on the
Stability Curve. While initial stability
(righting arms at small angles of heel)
depends almost entirely on metacentric
height, the overall shape of the stability
curve is governed by hull form. Figure 123 shows how changing hull form increases
or decreases righting arm by altering the
position and movement of the center of
buoyancy.
Figure 1-24 (Page 1-40)
illustrates how altering hull form affects the
stability curve as described in the following
paragraphs.

INCREASED BEAM

INCREASED DEPTH

TUMBLEHOME AND FLARE


FINING THE BILGES
1-5.10.1 Beam. Of all the hull dimensions
that can be varied by the designer, beam
LOCAL INCREASE IN IMMERSED VOLUME
has the greatest influence on transverse
stability. Metacentric radius (BM) was
LOCAL LOSS IN IMMERSED VOLUME
shown to be proportional to the ratio B2/T
in Paragraph 1-5.4.3. BM, and therefore
Figure 1-23. Effects of Changing Hull Form.
KM, will increase if beam is increased
while draft is held constant. If freeboard is
held constant while beam is increased, the angle of deck edge immersion is decreased; righting arms at larger angles and the range of stability
are reduced. If the depth remains constant, overall stability will be reduced because KB decreases, increasing BG, although this will be offset
at small angles by the increase in BM.

1-39

S0300-A8-HBK-010

1-5.10.2 Length. If length is increased


proportionally to displacement, with beam
and draft held constant, KB and BM are
unchanged. In practice, increasing length
usually causes an increase in KG, reducing
initial stability. If length is increased at the
expense of beam, righting arms are reduced
over the full range of stability. If length is
increased at the expense of draft, righting
arms will be increased at small angles, but
decreased at large angles.
1-5.10.3 Freeboard. Increasing freeboard
increases the angle of deck edge
immersion, increasing righting arms at
larger angles and extending the range of
stability.
If draft is held constant,
increasing freeboard causes a rise in the
center of gravity, mitigating the benefits of
increased freeboard to some extent.
1-5.10.4
Draft.
Reduced draft
proportional to reduced displacement
increases initial righting arms and the angle
of deck edge immersion but decreases
righting arms at large angles.

INCREASED BEAM
INCREASED LENGTH
DECREASED DRAFT
INCREASED FREEBOARD
REFERENCE
STABILITY CURVE

INCREASED LENGTH
DECREASED BEAM

REFERENCE
STABILITY CURVE

FLARE

INCREASED
DISPLACEMENT

EXTREME TUMBLEHOME
AND/OR DEADRISE
FROM STABILITY AND TRIM OF FISHING VESSELS, J. ANTHONY HIND, 1982.

1-5.10.5 Displacement. If length, beam,


and draft are held constant, displacement
can be increased only by making the ship
fuller. The filling out of the waterline will
usually compensate for the increased
volume of displacement, and BM, as a
function of I/, will increase. Height of
the center of gravity will also be decreased
by filling out the ships form below the
waterline. These changes will enhance
stability at all angles.

Figure 1-24. Influence of Hull Form on Stability.

1
S
1-5.10.6 Side and Bottom Profile. As
M
can be seen in Figures 1-13 and 1-25, the

increase in waterplane breadth and area


L
W
caused by inclining a wall-sided ship can
Z
G
W1
be calculated by simple geometry. The
stability curve develops good early righting
B1
B
arms and range of stability. Extreme deadGZ = MS+GMsin
rise (fining the bilges) or tumblehome in
the vicinity of the inclined waterline reCL
duces the increase in waterplane area and
outward shift of the center of buoyancy, resulting in a shallow stability curve. Ships
Figure 1-25. Residuary Righting Arm.
with flaring sides develop large righting
arms because of the rapid increase in waterplane area and large shift of the center of buoyancy as the ship is inclined. A round-bottomed ship with vertical sides beginning somewhat above
the water line, such as a tug or icebreaker, will roll easily to small angles of inclination but develop strong righting moments at large angles.
In the same way, flare or watertight sponsons some distance above the water line will have no effect on initial stability, but will cause a sharp
upward turn in the stability curve at larger angles of heel.

1-40

S0300-A8-HBK-010

1-5.11 Prohaskas Method. As shown in Figure 1-25, the righting arm at large heel angles can be thought of as consisting of two parts:
GZ = MS + GMsin
The distance from the upright metacenter to the line of action of buoyancy (MS) is called the residuary stability lever. The GMsin term depends
principally on KG, while MS is essentially a function of hull form. For inclinations up to about 30 degrees in merchant hulls of ordinary beam
to draft ratio, MS can be approximated as:
MS =

BM 2
tan sin
2

where:
BM = metacentric radius of the upright ship
A more accurate approach is to define a residuary stability coefficient (CRS):
CRS =

MS
BM

where :
BM = metacentric radius of the upright ship, [length]
GZ can now be defined in terms of GM, BM, and CRS:
GZ = (BM)CRS + GMsin
Using this basic approach, a regression analysis was performed using data from 31 warship hulls to obtain expressions for CRS in terms of other
hull parameters. The following expressions give reasonable estimates for CRS at 30 degrees of heel for fine-lined ships:
CRS = 0.8566

1.2262

KB
T

CRS = 0.1859

0.0315

B
T

KB
= 0.8109
T

0.2536 CM

0.035

B
T

0.03526 CM

where:
KB
T
B
CM

=
=
=
=

height of the center of buoyancy above the keel, ft


mean draft, ft
beam, ft
midships section coefficient

1-41

S0300-A8-HBK-010

1-6 LONGITUDINAL STABILITY


Longitudinal stability is the measure of a ships ability to resist rotation about a transverse axis and to return to its original position.
Longitudinal stability is particularly important when refloating stranded ships. The effects of weight shifts, additions, and removal may not be
apparent since a grounded ship is restrained from responding as a floating ship would. The effects must be calculated to ensure that the salvor
can accurately predict trim and longitudinal stability of the vessel when afloat.
1-6.1 Trim. Because the angles of inclination about transverse axes are quite small compared to typical angles of heel about a longitudinal
axis, trim is defined as the difference between the forward and after drafts:
t = Taft

Tfwd

where:
t = trim
Regardless of the difference between forward and after drafts, if a ships waterline is parallel to the design waterline, it has zero trim. Most
ships are designed with equal forward and after drafts. Some ships are designed with a deeper draft aft, called keel drag, to keep the propellers
adequately submerged in all operating conditions, or with a slightly deeper forward draft. Drag or other designed differences in fore and aft
draft should not be confused with trim. For ships with drag, trim is defined as:
t =

Taft

Tfwd

drag

Trim greater than one percent of the ships length is usually considered excessive. Excessive trim significantly alters the shape of the underwater
volume and can adversely affect transverse stability.
1-6.2 Longitudinal Stability Parameters. The longitudinal positions of centers of buoyancy, gravity, and flotation and their movements
influence the longitudinal stability characteristics of a ship. The height of the longitudinal metacenter, similar in concept to the transverse
metacenter, is the other major parameter of longitudinal stability.
1-6.2.1 Longitudinal Position of the Center of Gravity. The longitudinal position of the center of gravity (LCG) is determined by summing
weight moments about a vertical transverse reference plane, normally through one of the perpendiculars or the midship section.
1-6.2.2 Longitudinal Position of the Center of Buoyancy. The
longitudinal position of the center of buoyancy (LCB) is the longitudinal location of the centroid of the underwater hull. For most hull
forms, LCB lies near the midships section. For low-speed, full-bodied
cargo vessels, the optimum position of the center of buoyancy (from a
hull resistance standpoint) is about 0.02LWL forward of midships. As
speed increases, the optimum position moves aft. At a speed-to-length
ratio (Vk/
L) of 1.0 the optimum position is 1 to 2 percent of LWL aft
L = 2. Table
of midships and about 4 percent aft of midships for Vk/
1-6 gives approximate ranges for the longitudinal position of the center
of buoyancy as a function of the block coefficient.

Table 1-6. Longitudinal Position of the Center of Buoyancy.


CB

In a ship at rest, the longitudinal positions of the centers of gravity and


buoyancy lie on the same vertical line. LCB and LCG are therefore the
same distance from the midship section in a ship floating on an even
keel. In a ship with trim, there is a small difference in the distances of
B and G from midships due to their vertical separation, but this
difference is so small that it can usually be ignored.

LCB Relative to the Midship Section

0.60

0.016L aft to 0.002L forward

0.65

0.011L aft to 0.009L forward

0.70

0.002L aft to 0.020L forward

0.75

0.010L forward to 0.027L forward

0.80

0.015L forward to 0.030L forward

From Ships and Marine Engines, Volume IV, Design of Merchant Ships,
Schokker et al, 1953

1-6.2.3 Longitudinal Position of the Center of Flotation (LCF). The center of flotation is the geometric center of the ships waterplane.
The ship trims about a transverse axis through the LCF. The location of the center of flotation is required to calculate final drafts after a change
in trim. This can be calculated if the shape of the waterplane is known. In ships of normal form, the center of flotation may lie either slightly
forward or slightly aft of midships. The center of flotation of fine-lined ships is usually about five percent of the ships length aft of midships.
A broad transom increases the relative proportion of waterplane area aft of midships and will tend to shift LCF aft. If unknown, the center of
flotation can be assumed to be amidships without introducing significant error to most salvage calculations.
1-6.2.4 Longitudinal Metacenter. The longitudinal positions of the centers of buoyancy and gravity are simply projections of these centers
onto the vertical centerplane. The longitudinal metacenter, in contrast, is a point distinct from its transverse counterpart. Its height is an
indication of the ships ability to resist trimming forces.
Longitudinal Metacentric Radius. The longitudinal metacentric radius (BML) is the vertical distance between the center of buoyancy and the
longitudinal metacenter. The longitudinal metacentric radius is calculated by:
BML =

1-42

IL

S0300-A8-HBK-010

If the waterplane shape is defined by ordinate stations, the moment of inertia can be determined numerically. If not, the longitudinal moment
of inertia of most ships waterplanes can be approximated by:
IL B L 3 CIL
where CIL = tegression analysis derived longitudinal inertia coefficient, approximated by 0.143CWP - 0.0659. For a rectangular barge, IL =
B(L3)/12; the value of CIL for a rectangular waterplane (the limiting value) is 1/12 or 0.0833.
Because the longitudinal moment of inertia is proportional to the cube of the ships length rather than beam, the longitudinal moment of inertia
and longitudinal metacentric radius are much greater than their transverse counterparts.
Height of the Longitudinal Metacenter. The height of the longitudinal metacenter (KML) is given by:
KML = KB + BML
Longitudinal Metacentric Height. The longitudinal metacentric height (GML) is the distance between the center of gravity and the longitudinal
metacenter.
GML = KML - KG
= KB + BML - KG
1-6.3 Trimming Arms and Moments. If
the center of gravity is displaced from its
longitudinal position in vertical line with
the center of buoyancy, as shown in Figure
1-26, a trimming moment (MT) equal to
GG1(W) tends to rotate the ship about a
transverse axis through the center of flotation. As the ship inclines, the shape of
the underwater volume changes and the
center of buoyancy moves until it is again
in line with the center of gravity. Simultaneously, the projection of the position of
the center of gravity onto a horizontal plane
moves towards the high end of the ship.
For small trim angles, the horizontal translation of the position of the center of
gravity can be neglected. The trim resulting from a known trimming moment could
be determined precisely by iterative
numerical integration, but this would be a
tedious process.
Simple methods to
estimate trim with reasonable accuracy are
described in the following paragraphs.
A ship trims about an axis through its
center of flotation because LCF lies at the
centroid of the waterplane. The moments
of volumes of the wedges immersed and
emerged as the ship trims are equal,
although the volumes are not. Because the
volumes are not equal, the ships will settle
or rise slightly as it trims to maintain
constant displacment. LCF also shifts
slifhtly as the ship trims and changes draft.

LCF

G1

W
LCF

L1

G1

W1

B1

Figure 1-26. Trim due to Shift in LCG.

ML

MT

W
W1

LCF

G
B

ZL

B1

L1
L

Figure 1-27. Trimming Moments and Longitudinal Metacenter.

1-6.4 Moment to Change Trim One Inch (MT1). A trimming moment applied to the ship in Figure 1-27 causes a longitudinal inclination
or trim angle, . The immersion and emergence of the two wedges of buoyancy causes the center of buoyancy to move forward a distance BB1.
A longitudinal righting arm GZL develops. Because the small vertical separation between B and G is much less than the longitudinal metacentric
height, GZL and BB1 are approximately equal. The moment arm GZL can be related to the longitudinal metacentric height as in transverse
inclinations:
GZL
sin =
, GZL = GML sin Mt = W GML sin
GML
where:
GZL =
GML

longitudinal righting arm, [length]


= longitudinal metacentric height, [length]

Mt = trimming moment, [length-force]


W = ships weight, [force]

1-43

S0300-A8-HBK-010

By similarity of triangles:
sin =

t
L

where:
t
L

=
=

change of trim, [length] = Tf Ta


length between perpendiculars, [length]

Setting change in trim equal to one inch or 1 12-foot:


Mt =

W (GML )
12 L

where:
GML
Mt
W
L

=
=
=
=

longitudinal metacentric height, ft


trimming moment, ft-tons
ships weight, lton
length between perpendiculars, ft

This moment is called the moment to change trim one inch (MT1); in metric units, a moment to trim one centimeter (MTCM) is similarly defined.
MT1 is useful for evaluating the effect of trimming moments so long as the change in trim is not great enough to change the waterplane area
or shape appreciably:
Mt
t =
MT1
If longitudinal metacentric height (GML) is unknown, MT1 can be closely approximated by using metacentric radius (BML), since the difference
between GML and BML is small a percentage of their values:
IL
(BML ) W
IL
35
=
(seawater)
MT1
=
12 L
420 L
12 L
This value is known as the approximate moment to trim one inch. MT1 can also be approximated less accurately by an empirical relationship:
MT l =

30 (TPI)2
B

where:
TPI =
B =

tons per inch immersion, lton/in


ships beam, ft

1-6.5 Drafts After a Change in Trim. As a ship trims about the center of flotation, the change in draft at the bow is proportional to the ratio
of the distance between the forward perpendicular and the center of flotation to the length of the ship:
Tf =

t df

L
New Tf = Tf Tf
Likewise, the change in draft aft:
Ta =

tda

L
New Ta = Ta Ta
where:
Tf
t
df
L
Ta
da

=
=
=
=
=
=

change in draft forward


change of trim
distance from forward perpendicular to LCF
length between perpendiculars
change in draft aft
distance from after perpendicular to LCF

and distance, draft, trim, and length are measured in like units.

1-44

S0300-A8-HBK-010

1-6.6 Movement of LCB and LCG with Change of Trim. As discussed in Paragraph 1-5.3, movements of LCB and LCG accompany changes
of trim. From Figures 1-26 and 1-27:
BB1
GG1
t
=
= tan =
BML
GML
l
BB1 =

BML t
L

and

GG1 =

GML t
L

where:
BML
GML
t
L

=
=
=
=
=

longitudinal metacentric radius


longitudinal metacentric height
change of trim
length between perpendiculars
trim angle

and trim and length are measured in like units.


The shift of LCG or LCB with a change in trim can be closely approximated by:
t (MT1)
BB1 or GG1 =
W
where:
t
MT1
W

=
=
=

change of trim, in.


moment to trim one inch, lton/in.
ships weight, lton

1-7 PARAMETRIC DETERMINATION OF HULL CHARACTERISTICS


The hull characteristics of a ship are determined and tabulated when the ship is designed and verified following construction. This information
is contained in a number of different documents, described in detail in Appendix B. The two most useful documents are the previously discussed
cross curves of stability and curves of form. In the absence of detailed stability information or the precise mapping of the hull form necessary
to develop hydrostatic characteristics by numerical integration, hull characteristics must be estimated. Methods of estimating some of the
required parameters have been presented in the previous sections. When information is extremely limited, an analytical method, based on a
parametric hull model, can be employed. This method has been shown to yield results within 10 percent of rigorously determined values for
most ship forms. The parametric method has its inception in a regression analysis of 31 commercial hull types published by Joseph D. Porricelli,
J. Huntly Boyd, Jr., and Keith E. Schleiffer in the Society of Naval Architects and Marine Engineers Transactions, Vol.91, pp. 307-327, August
1983. Many of the relationships were subsequently refined though further regression analysis by Herbert Engineering Corporation as part of
the NAVSEA Program of Ship Salvage Engineering (POSSE) development work in 1990 (use of POSSE is detailed in Volume 2 of the Salvage
Engineers Handbook). At the same time, relationships for stability parameters and weight distributions applicable to warships and other finelined ships were developed. The parametric factors for warships and naval auxiliaries were derived from analysis of U.S. Navy hulls and may
not apply precisely to ships of other navies. This is particularly true of amphibious warfare ships and fleet replenishment auxiliaries. U.S. Navy
amphibious warfare ships and replenishment auxiliaries are designed for a 20-knot service speed and are correspondingly finer than slower
auxiliaries and bow-door LSTs with typical speeds in the 10- to 16-knot range.
1-7.1 Parametric Model. The method creates a baseline parametric model of the hull, consisting of the following parameters for the full-load
condition:
Coefficients of form, CB, CWP, CP, CM
Displacement and weight, D, W
Height of the center of buoyancy, KB
Height of the Metacenter, KM
Height of the Center of Gravity, KG
Metacentric radius, BM
Metacentric Height, GM
Tons per Inch Immersion, TPI
Moment to Trim One Inch, MT1
Longitudinal position of the center of buoyancy, LCB
Longitudinal position of the center of flotation, LCF
Longitudinal position of the center of gravity, LCG
Parameters for other conditions are extrapolated from the baseline, or full-load model.

1-45

S0300-A8-HBK-010

1-7.1.1 Required Information. This method requires only limited information:


Length between perpendiculars, L
Breadth, B
Depth, D
Maximum summer draft amidships, T
Design sea speed at normal service draft, Vk
This information is available from sources such as the ABS Record, Janes Shipping Registry, Lloyds Register of Shipping, etc., or may be
compiled from other sources, including the ships crew or agents.
1-7.1.2 Displacement and Coefficients of Form. To determine the necessary hydrostatic characteristics of a ship, the coefficients of form
are first estimated, starting with the block coefficient:

V
CB = f1 1.10736 0.550401 k

L
where:
Vk
L
f1

=
=
=

design sea speed at service draft, knots


length between perpendiculars, ft
1.61 for destroyer type hulls (including cruisers based on destroyer hulls, such as CG-16, CG-26, CG-47, etc.)
1.41 for frigates
1.28 for cruisers
1.08 for bulk carriers
1.06 for liquid petroleum gas (LPG) carriers
1.04 for liquid natural gas (LNG) carriers
1.03 for ore-bulk-oil (OBO) carriers
1.03 for lumber ships
1.025 for product tankers/chemical carriers
1.01 for crude carriers
1.00 for breakbulk freighters and most barges with rake*
0.98 for cargo liners (16-18 kts)
0.97 for container ships
0.96 for Navy replenishment oilers (Vk 20 kts, AO/AOE/AOR)
0.95 for RO/RO ships
0.93 for Navy replenishment vessels other than oilers (Vk 20 kts, AE/AFS)
0.91 for amphibious warfare ships (LSD/LPD/LPH/LKA/LST)
0.89 for barge carriers, Navy repair ships/tenders (AR/AD/AS)

In the context of the following discussions, the phrase "barges with rake" refers to ocean going barges with raked, ship-shaped or
spoon-shaped bows, and cut-up sterns, usually with skegs. It does not apply to box-shaped lighters or to barges designed for harbor
use with identical flat rake at bow and stern.

Waterplane coefficient:
CWP = k1 0.702 CB
where:
k1

0.360
0.325
0.336
0.339
0.387
0.370
0.316
0.306

for
for
for
for
for
for
for
for

barge carriers and barges with rake


container ships
RO/RO ships
naval repair ships/tenders
destroyers, frigates, and cruisers
well deck type amphibious warfare ships (LSD/LPD)
Navy replenshishment ships and fast LKA, LST (20 kts)
other merchant ship types and slow-speed naval auxiliaries

Longitudinal prismatic coefficient:


CP = 0.917 CB + k2
where:
k2

1-46

0.073 for merchant ships and naval auxiliaries


0.075 for barges with rake
0.147 for destroyers, frigates, and cruisers

S0300-A8-HBK-010

Midships coefficient:
CB

CM =

CP

With an estimate for block coefficient, displacement volume and displacement can be estimated:
= L B T CB
L B T CB

= W

where:
=
D =

=
W =

displacement volume at full load


full-load displacement

specific volume of water = 35 ft3/lton for seawater


ships weight at full load

1-7.1.3 Heights of Centers. Height of the center of buoyancy (KB) is estimated by a form of Posdunines formula:
KB =

CWP
CB + CWP

where:
CWP =
CB

waterplane coefficient
= block coefficient

mean draft

Metacentric radius is equal to the transverse moment of inertia of the waterplane (IT) divided by the displacement volume ():
IT

BM =

IT can be expressed as:


IT = L B 3 CIT
where CIT is the transverse inertia coefficient and is a function of waterplane shape. CIT is determined from the waterplane coefficient (CWP):
CIT = 0.125 CWP

0.045

for ships

= 0.125 CWP

0.043

for barges with rake

Transverse metacentric height for the full-load departure condition (corrected for free surface) is correlated to beam, or beam to depth ratio,
depending on ship type:
B
for cargo liners and container ships
GM = 2.816 - 1.88
D
B
= 15.86 - 19.62
D

for tankers in general

B
= 0.714 + 2.2
D

for cargo ships in general

= f2 B

for other merchant ship types

B2

T
= f3 +
- 0.53 D
12 T

for barges

where:
f2

0.055 for barge carriers and RO/RO ships


0.065 for bulk carriers
0.075 for OBO carriers

f3

=
=

1.18 for barges with rake


1.00 for barges without rake

1-47

S0300-A8-HBK-010

From the estimates for KB, BM, and GM, KM and KG can be estimated:
KM = KB + BM
KG = KM - GM
Since the estimate for KG is based on the parameterized GM estimate, the value returned is the virtual, or effective KG (corrected for free
surface).
GM does not parameterize well for U.S. Navy hulls because Navy stability standards (described in Appendix C) do not include minimum GM
requirements. Uncorrected full-load KG does parameterize well, as a function of depth:
KG = f4D

(Navy hulls)

where:
f4

=
=
=
=
=
=

0.55
0.61
0.63
0.72
0.62
0.50

for
for
for
for
for
for

cruisers and destroyers


frigates
amphibious warfare ships without well decks
amphibious warfare ships with well decks (LSD/LPD)
fleet replenishment auxiliaries
repair ships/tenders

For Navy hulls, GM (uncorrected for free surface) is calculated from the estimates for KB and KG. The parametric factors were derived from
an analysis of U.S. Navy hulls and may not apply precisely to ships of other navies.
1-7.1.4 Tons Per Inch Immersion. TPI is calculated directly, using the estimated waterplane coefficient to estimate waterplane area:
L B CWP
TPI =
420
where L and B are measured in feet.
1-7.1.5 Moment to Trim One Inch. A value for MT1 is found using estimates for longitudinal metacentric height or radius:

MT 1 =

GML W
12 L

BML W
12 L

where BML is given by:


BML =

IL

IL
35 W

IL

MT1 =

420 L

The longitudinal moment of inertia, IL, of a ship-shaped waterplane can be expressed as:
IL = B L 3 CIL
where the longitudinal inertia coefficient, CIL, is given by:
CIL = 0.143 CWP - k3
where:
k3

1-48

0.0659
0.0664
0.0643
0.0634

for
for
for
for

merchant ships and slow-speed auxiliaries


replenishment auxiliaries
amphibious warfare ships
destroyers, frigates, and cruisers

S0300-A8-HBK-010

1-7.1.6 Longitudinal Positions of Centers. The distance from the forward perpendicular to LCF, LCB, and LCG can be estimated as follows.
LCF is estimated as a function of speed (Vk) and length (L):

V
LCF = 0.5L k + 0.914

160

V
= 0.485 L k
100

V
= 0.5 k
135

for tankers

0.9

for bulk carriers

0.924

for single-screw cargo ships and naval auxiliaries

0.95
= 0.5 L
Vk

1.03


V
= L 0.5 k
135

0.924

V
= 0.5 L k
135

for twin-screw cargo ships with transom sterns

0.23

for twin-screw cargo ships with cruiser sterns

0.95

for barges with rake

where Vk is given in knots and L in feet.


LCB at full load and zero trim is approximated as a function of length (L) and prismatic coefficient (CP):
LCB = L 0.5 - 0.175CP - k4
where:
k4

0.125
0.111
0.117
0.126
0.146

for
for
for
for
for

merchant ships and slow-speed auxiliaries


replenishment auxiliaries
amphibious warfare ships
destroyers, frigates, and cruisers
barges with rake

To estimate the longitudinal position of the center of gravity, trim must be known or estimated. If unknown, trim can be estimated from similar
ships as a percentage of length. Multiplying trim (t) in inches by MT1 gives the trimming moment Mt:
MT 1 (t) = Mt
Trimming moment divided by weight (W) gives the trim arm or lever (GZL):
Mt

= GZL
W
Since the trim arm is the horizontal separation between LCB and LCG prior to trimming:
LCB GZL = LCG
Upon trimming, LCB will relocate to a position in vertical line with LCG. LCG can be assumed to be directly above the estimated LCB for
a ship with zero trim at normal full-load departure condition.
1-7.2 Changes. The values calculated are for the full-load departure conditions, and must be corrected for other conditions. Floating or grounded
drafts can be observed on site. New floating displacement, drafts and location of center of gravity are determined by evaluating the effects of all
weight changes from the normal full-load departure condition. Hydrostatic properties are assumed to vary linearly with draft according to:
TPI2 = TPI1

TPI1 0.0075 T1 - T2

MT12 = MT11

MT11 0.025 T1 - T2

LCB2 = LCB1

LCB1 0.002 T1 - T2

LCF2 = LCF1

LCF1 0.004 T1 - T2

Where the subscript 1 denotes the full-load condition and the subscript 2 the new condition. The drafts T1 and T2 are taken at the LCF for each
condition.

1-49

S0300-A8-HBK-010

Longitudinal locations are referenced to the


forward perpendicular. These relationships
apply only so long as the change in draft or
trim does not cause a significant change in
the shape of the underwater hull form. KM
does not vary significantly with draft until
draft is dramatically decreased, to
approximately two-thirds full-load draft,
after which it increases.

Vk

CB

T B

CWP

KB CT
1-7.3 Calculation Hierarchy. Only CB,
GM (or KG), and LCF are calculated
directly from the basic input data (L, B, T,
D, Vk). Because other parameters are
successively calculated from previously
calculated parameters, basic data, and
empirically derived factors, there is a
hierarchy of accuracy among the calculated
parameters. This hierarchy is shown in the
two panels of Figure 1-28. Two panels are
used to reduce the complexity of the
diagram. The basic input parameters are
listed across the top of each of the two
panels.

B L

GM

TPI

IT
BM
KM
KG

B T L

1-7.4 Cautions. The parametric method


described in this paragraph was developed
through regression analysis of typical,
conventional hull forms. The less typical a
particular hull, compared to ships of the
same type, the greater the error introduced
by use of the relationships given.
As this method is based primarily on
analysis of the speed-to-length ratio, errors
will be larger for an underpowered
hullfor example, a hull designed for 20
knots but actually powered for only 16
knots.

W,

CB

CM
LCB

LCG

GZL MT

Vk

MTI

CP

TPI

BML

LCF
CWP

IL

KB
IT

BM

KM

KG

Figure 1-28. Calculation Hierarchy.

Because of the interdependence among various parameters, changing any parameter (except LCF) creates a ripple effect that necessitates
recalculation of other parameters. Mixing bits of actual data with data calculated by the analytical method in a set of salvage calculations
without recalculating lower precedence parameters tends to give poorer results than complete sets of either calculated or actual data. Specifically,
hydrostatic properties and coefficients of form must be compatible.
Within the framework of these limitations, the parametric method yields results sufficiently accurate for salvage work, and provides a means
to evaluate a casualtys condition when only limited information is available.
1-7.5 Applications to Salvage Calculations. The nature of the relationships in the analytical method dictates the methodology of their use.
From the input data, the method calculates parameters and creates a baseline ship model in the full-load condition. From the base condition,
parameters at other conditions are calculated by one of two approaches.

1-50

The new condition is defined by drafts (for example, drafts on departure from last port). Change in block coefficient is calculated
first. With the new block coefficient, mean draft and trim, a new set of parameters is calculated. The difference between old and
new displacements gives the required weight change between the full-load and new condition. If the change in draft results from
stranding, the difference between old and new displacements is the ground reaction. This approach can also be used to determine
the amount and LCG of weight that must be added or removed to reach a desired draft and trim.
The new condition is defined by change in weight (consumption of fuel and consumables, flooding, cargo discharge, etc.). The
sum of weight change and old displacement gives the new displacement. Change in draft is calculated from the total weight
change and TPI. For large weight changes, the change in draft is calculated incrementally, recalculating TPI for each intermediate
draft. Shift of LCG is calculated by moment balance. A new block coefficient is calculated from the new displacement and mean
draft. With the new block coefficient and mean draft, a new set of parameters is calculated as for the full-load condition, except
that the new LCB is calculated from the new LCG.

S0300-A8-HBK-010

1-8 WEIGHT AND STABILITY

The salvage engineer must fully appreciate the relationship between weight and ship stability. The addition and removal of weight is the most
common evolution affecting a ships stability and can be the result of onloading and offloading cargo and equipment, refueling, consuming stores
or fuel, ballasting, etc. Weight additions and removals have three effects:

Change of displacement with attendant change of draft.

Movement of the center of gravity.

Development of trimming or inclining moments.

Displacement changes cause draft


changes and changes in the hull
characteristics. The change in the
transverse metacentric radius is
particularly important because of its
potential effect on stability. Both
weight additions and removals may
change the moment of inertia of the
waterplane. Weight additions will
increase and weight removals will
decrease displacement volume. Table
1-7 illustrates the general effect of
weight changes on an intact ship. To
evaluate a weight change, it is
simplest to assume that the weight is
added or removed at the center of
gravity (G) for the purpose of
calculating the effect on mean draft,
and then moved to its final location
in a series of steps to account for the
effects of its vertical, transverse, and
longitudinal moments.

Table 1-7. Effect of Weight Movements.

STABILITY

CENTER OF
GRAVITY

CENTER OF
BUOYANCY

METACENTER*

Weight shift up
Weight shift down
Weight shift transverse

Decrease
Increase
Decrease

Up
Down
Port/Starboard

No change
No change
To low side

No change
No change
No change

Weight added at G
Weight added above G
Weight added below G

Decrease
Decrease
Increase

No change
Up
Down

Up
Up
Up

Down
Down
Down

Weight removed at G
Weight removed above G
Weight removed below G

Increase
Increase
Decrease

No change
Down
Up

ACTION

Down
Down
Down

Up
Up
Up

*Relative movement of metacenter is based on the relationship BM = I / and assumption that


waterplane shape and area do not change appreciably for moderate changes of draft and
displacement. As draft increases with added weight, the reduction in BM [I /] is greater than
the rise of B. Conversely, as draft and displacement decrease, the increase of BM is greater
than the lowering of B.

1-8.1 Weight Shifts. When weights


are moved about the ship,
displacement and mean draft remain constant; stability parameters that are functions of displacement or draft, such as height of metacenter, are
therefore unaffected. The distance the center of gravity moves when a weight is shifted is the product of the weight (w) times the distance
moved (d), divided by the total weight of the ship (W):
wd
GG1 =
W
This distance can be resolved into vertical, transverse, or longitudinal components. A single weight shift can cause any combination of
transverse, vertical, or longitudinal shifts of the center of gravity with attendant effects on longitudinal and transverse stability. Although they
occur simultaneously, each effect can be assumed to occur independently; the effects can be calculated separately as though they were occurring
sequentially. Change of KG alters GM and righting arms as discussed in Paragraph 1-5. The effects of longitudinal and transverse weight shifts
are discussed in the following paragraphs.

1-51

S0300-A8-HBK-010

1-8.1.1 Longitudinal Effects of Weight


Shifts. When a weight movement has a
longitudinal component, LCG shifts and the
ships weight acting through the new center
of gravity and buoyancy acting through the
old center of buoyancy form a couple, or
trimming moment, as shown in Figure 1-26.
The magnitude of the trimming moment is:
Mt = W GG1
where:

W1

W
Mt
GG1

=
=
=

ships weight
trimming moment
longitudinal distance from
the old LCG to the new
LCG

The trimming moment is also equal to the


product of the weight moved (w) and the
longitudinal distance moved (d).

L1

G1
B

Mt = wd

CL

1-8.1.2 Offcenter Weight. The effect of


offcenter weight is to create an inclining
moment. This effect can be evaluated by
calculating the lateral movement of the
ships center of gravity off the centerline.
The magnitude of the inclining moment is:

Figure 1-29. List Due to Transverse Shift of G.

MI = W(GG1)
where:

GG1

MI

lateral (horizontal) shift


of center of gravity,
[length]
inclining moment, [forcelength]
ships weight (including
the offcenter weight),
[force]

W1

since:

Z
Z1

wd
GG1 =
W

T
0

G1

L1

where:
d

lateral (horizontal) distance


that the weight w is moved,
[length]

CL

Figure 1-30. Reduced Righting Arm due to Transverse Shift of G.

then:
W GG1 = wd = MI

The inclining moment will cause the ship to list to an angle where the center of buoyancy is again in vertical line with the center of gravity.
The angle of list becomes the new equilibrium position; when disturbed, the ship will roll about the angle of list. The effect of a permanent
list is to reduce the righting arms and range of stability when the ship rolls towards the list, and increase them when the ship rolls away from
the list. For small angles of inclination (less than 7 to 10 degrees), list can be found by reference to the metacentric height. From Figure 1-29,
the list due to an offcenter weight can be seen to be:
GG1
tan =
GM

1-52

wd
= tan 1

W GM

S0300-A8-HBK-010

1-8.1.3 Stability Curve Correction for Offcenter Weight. Figure 1-30 shows a ship whose center of gravity has moved from G to G1. When
inclined towards G1 to some angle , the righting arm developed is not GZ, but a smaller arm, G1Z1. The reduction in righting arm (GT) is:
GT = GG1cos
As with the sine correction for actual KG, the offcenter weight correction, as a cosine curve, is plotted to the same scale as the curve of statical
stability as shown in Figure 1-31. The corrected stability curve is the difference between the two curves. The angle at which the corrected curve
crosses the horizontal axis is the angle of list caused by the offcenter weight. Extending the curve to the left of the origin shows the increased
righting arms developed on the side away from the list. In dynamic situations, the increase in righting energy on the side away from the list
does not increase stability because the ship will roll about the angle of list. If the ship is subjected to a constant upsetting force, such as a steady
beam wind, the increased righting arms provide additional stability if the ship is oriented so that the upsetting force heels the ship away from
the list, towards its strong side. The increased righting arms and energy must also be overcome if the salvage plan calls for the ship to be heeled
away from the list by external forces. It should also be remembered that if the ship is heeled towards its strong side, the area under the curve
from the point where the curve crosses the axis to the angle of heel represents stored energy. If this area is larger than the area under the
stability curve on the weak side, the ship could capsize if suddenly released.

RIGHTING ARMS IN FEET

3
COSINE
CORRECTION
CURVE FOR
OFFCENTER
WEIGHT

2
GG1 COS

INITIAL STABILITY CURVE


WITH KG OF 21

0
LOSS IN RIGHTING ARMS DUE
TO OFFCENTER WEIGHT

1
2
3
90

80

70

60

50

40

30

20

10

10

20

30

40

50

60

70

80

90

80

90

ANGLE OF INCLINATION,
RANGE OF STABILITY
ROLLS TO STBD

RIGHTING ARMS IN FEET

3
2

CORRECTED
STABILITY CURVE

ANGLE OF LIST

0
RANGE OF STABILITY
ROLLS TO PORT

1
2
POSITIVE RIGHTING ARMS, STBD

POSITIVE RIGHTING ARMS, PORT

3
90

80

70

60

50

40

30

20

10

10

20

30

40

50

60

70

ANGLE OF INCLINATION,

Figure 1-31. Correction to Statical Stability Curve for Transverse Shift of G.

1-8.2 Weight Additions and Removals. Weight addition or removal at the center of gravity changes displacement without introducing
trimming or inclining moments. The increase or decrease in mean draft in inches (T) is approximately equal to the weight added or removed
(w) in tons divided by the tons per inch immersion (TPI):
w
T =
TPI

1-53

S0300-A8-HBK-010

1-8.2.1 Weight Changes Away From the Center of Gravity. When weights are added or removed at some distance from the center of gravity,
the center of gravity moves toward the added weight, or away from a removed weight, to a new position determined by the size and location
of the weight. The weight change can be treated as an addition (or removal) at the center of gravity, followed by a shift to the location where
the weight is added:
(Gg) (w)
GG1 =
(W1)
where:
GG1
Gg
W1
w

=
=
=
=

shift of ships center of gravity, [length]


distance between ship and added weight centers of gravity, [length] = the distance d that the weight is "shifted"
new total weight of ship, [force] = W w
weight added (+) or removed ()

The new vertical, transverse, and longitudinal positions of the center of gravity can also be calculated directly, by summing moments. Height
of the center of gravity is given by:
W (KG) w (kg)
KG1 =
W w
where:
KG1
W
KG
w
kg

=
=
=
=
=

height of the ships center of gravity after weight change, [length]


original weight (displacement) of the ship, [force]
original height of the ships center of gravity, [length]
weight added (+) or removed (-), [force]
height of the center of gravity of the added or removed weight above the keel, [length]

New transverse and longitudinal positions of the center of gravity can be determined by the same method.
A longtitudinal moment caused by weight addition or removal will not necessarily trim the ship. Most ships are not symmetrical about a
transverse axis; as a ship settles or rises, the change in buoyancy is weighted towards one end, causing LCB to shift towards the fuller end.
If the buoyancy moment generated by the shift in LCB equals the trimming moment, the ship will not trim. Conversely, a weight added directly
above or below the center of gravity may cause the ship to trim to keep the centers of buoyancy and gravity in vertical line. For any weight
addition or removal, a ship will assume the trim that brings the center of buoyancy directly under the new center of gravity. The trim resulting
from a weight change can be determined very precisely by calculating LCB for trimmed waterlines at the new displacement until a trim is found
that brings LCB under LCG. Simpler approximate methods to determine trim resulting from weight changes can be derived by determining
where weights must be added or removed from a ship to change draft without changing trim. These methods are described in the following
paragraphs, and are sufficiently accurate for virtually all situations.
1-8.2.2 Weight Changes Without Change of Trim. If a weight is to be added to a ship without changing trim, it must be added at a location
that will be in vertical line with the resultant upward force of the added buoyancy. If the rise or sinkage is parallel, the added buoyancy results
from the immersion of a layer of uniform thickness between the old and new waterplanes. The center of buoyancy of this layer is very close
to the midpoint of a line connecting the centroids (centers of flotation) of the old and new waterplanes.
For small draft changes through a ships normal range of drafts, the old and new waterplanes are very nearly the same size and shape. The line
connecting the centroids is therefore essentially vertical and the center of buoyancy of the immersed layer is in line with the centroid of the old
waterplane, or center of flotation. For moderate weight changes, causing small changes in draft, at locations other than the center of flotation,
trim can be closely approximated by:
a. Taking the distance from the added or removed weight to the LCF as the trimming arm,
b. Multiplying the trimming arm by the weight to determine trimming moment, and
c. Dividing the trimming moment by MT1 to find the resulting trim.
For larger weights whose addition or removal causes draft changes large enough to appreciably change hydrostatic functions, the trimming arm
is taken as the distance from the new LCG to the LCB at the new waterline. Since TPI varies with draft, an iterative solution is required, as
shown in Example 1-3.

1-54

S0300-A8-HBK-010

EXAMPLE 1-3
WEIGHT AND TRIM
This example calculates trim resulting from moderate (causing small changes in draft) and large weight additions at various locations on an FFG-7 Class ship.
a.

Calculate the change of trim when a 100-ton weight is added to an


FFG-7 Class ship at the following locations:

b. Calculate the location for the center of gravity of 1,000 tons of weight to
be removed from an FFG-7 Class ship with initial drafts of 14 feet 6 inches
forward and aft without changing trim.

(1) Center of Flotation.


First estimate of new mean draft:
(2) Center of Gravity.
(3) 50 feet abaft the forward perpendicular.
FFG-7 Curves of Form are given in Figure FO-2. Initial
drafts are 14 feet, 6 inches, forward and aft, LBP is 408 feet.
From the curves of form:

TPI
LCF
LCB
MT1
W

=
=
=
=
=

32
23.4 feet abaft midships
LCG
=
1.4 feet abaft midships
745 foot-tons
3,495 tons

T
Tnew

=
w/TPI = 1,000/32 = 31.25 31 inches
Told - T = 14' 6" - 31" = 11 feet 11 inches

Second estimate of new mean draft:

TPI at 11' 11"


TPIavg
T
Tnew
=
LCF at 11' 9"

=
=
=
Tnew
=

28.5
(32 + 28.5)/2 = 30.25
1,000/30.5 = 33.06 33 inches
- T = 14' 6" - 33" = 11 feet 9 inches
14 feet abaft midships

Center of buoyancy of immersed layer (lcb) is approximately midway


between the old and new LCF,
(23.4 + 14)
= 18.7 feet abaft midships
lcb =
2

Calculate the increase in mean draft:

w
100
= 3.125 inches 3 inches
=
TPI
32
= Tnew T = 14 feet 6 inches 3 inches = 14 feet 9 inches

T =

Tnew

Calculate the change in trim for 100 tons added at:

Removing the 1,000 tons so that the center of gravity of the removed
weight is approximately 19 feet abaft midships will cause no noticeable
trim.
c. Calculate the change in trim for an FFG-7 Class ship with initial drafts of
14' 6" forward and aft if 1,000 tons are removed from the following
locations:
(1) LCF.

(1) Center of Flotation


(2) LCG.
The change in draft is small, so adding the weight at
LCF causes no change of trim. This is verified by
observing that the LCF at the new mean draft of 14 feet
9 inches is 23.5 feet. The center of the new waterplane
(LCF) is only 0.1 foot from the center of the old
waterplane, so the center of buoyancy of the immersed
layer is essentially directly over the old LCF.

Mt
t

=
=
=
=

distance from LCF to added weight


23.4 - 1.4 = 22 feet
w(trim arm) = 100(22) = 2,200 foot-tons
Mt / MT1 = 2,200/745 = 2.95 3
inches by the bow

(3) 50 feet abaft the forward perpendicular


50 feet abaft the forward perpendicular is 154 (204 -50)
feet forward of midships
Trim arm
Mt
t

=
=
=

Tnew
LCB at 11' 9"
MT1 at 11' 9"
MT1avg

=
=
=
=

11' 9" (from part b.)


6 feet forward of midships
565 foot-tons
(745 + 565)/2
=
655

(1) 1,000 tons removed at original LCF

(2) Center of Gravity


Trim arm

(3) 100 feet forward of midships.

23.4 + 154 = 177.4 feet


100(177.4) = 17,740 foot-tons
Mt/MT1 = 17,740/745 = 23.81 23
inches by the bow

GG1
Gg
GG1
LCG1

=
=
=
=

trim arm

=
=
=
=

Mt
t

(Gg)(w)/(W + w)
23.4 -1.4 = 22 feet
(22)(1,000) / (3,495 - 1,000) = 8.8 feet forward
-1.4 feet (aft) + 8.8 feet (forward) = 7.4 feet
forward of midships
distance from new LCG to new LCB
7.4 - 6 = 1.4 feet (LCG is forward of LCB)
1,000(1.4) = 1,400 foot-tons
Mt/MT1 = 1,400/655 2 inches by the bow

(2) 1,000 tons removed at original LCG

GG1
LCG1
trim arm
Mt
t

=
=
=
=
=

0
1.4 feet abaft midships
6 + 1.4 = 7.4 feet (LCG is aft of LCB)
1,000(7.4) = 7,400 foot-tons
Mt / MT1 = 7,400/655 11 inches by the stern

(3) 1,000 tons removed 150 feet forward of midships

Gg
GG1
LCG1

=
=
=

trim arm
Mt
t

=
=
=

150 + 1.4 = 151.4 feet


(151.4)(1,000)/(3,495 - 1,000) = 60.7 feet aft
1.4 feet (aft) + 60.7 feet (aft) = 62.1 feet abaft
midships
62.1 + 6 = 68.1 feet (LCG is aft of LCB)
1,000(68.1) = 68,100 foot-tons
Mt/MT1 = 68,100/655 104 inches by the stern

1-55

S0300-A8-HBK-010

1-8.2.3 Point of Constant Draft. When a weight is added at some point away from the LCF, the ship trims as it sinks to a new mean draft.
Drafts on the opposite side of the LCF are reduced by the effect of trim, but increased by parallel sinkage. At some point the reduction in draft
caused by trim equals the increase in draft caused by parallel sinkage:
T due to parallel sinkage = T due to change of trim
w
Tparallel sinkage =
TPI

wd1
wd1 d2
Ttrim =
t =

MT 1
MT 1 L
wd1d2
w
=
TPI
MT 1 (L)
where:
t
d1
MT1
L
TPI
d2
w

=
=
=
=
=
=
=

change of form, in.


distance from the LCF to the added or removed weight, ft
moment to change trim one inch, ft-ton/in
length between perpendiculars, ft
tons per inch immersion, lton/in
distance from the point of constant draft to the LCF, ft
weight added or removed, lton

The relationship can be solved to determine the point of constant draft for weight added or removed at a known location. It is generally more
useful to solve for d1 to find the point where weight must be added or removed to keep draft constant at some point:
(MT1) (L)
d1 =
(TPI) (d2)
Note that w cancels out of the equation. So long as the weight change is not large enough to significantly alter MT1, TPI, or the position of
LCF, the amount of weight added or removed does not affect the location of the point where weight must be added or removed to keep draft
constant at another point.
1-8.3 Inclining Experiment. The predictable and measurable effects of offcenter weight are used to determine height of center of gravity in
an inclining experiment. By shifting a known weight a specified distance, the movement of the center of gravity can be determined. The
resulting inclination (heel) observed and the tangent formula (see Paragraph 1-8.1.2):
GG1
wd
tan =
=
GMeff
W (GM)
is solved for the as inclined, or effective metacentric height, GMeff:
GMeff =

GG1
tan

wd
W tan

Inclining experiment reports are an important source of data for ship characteristics, especially a baseline vertical position for the center of
gravity.
1-8.4 Sallying Ship. Sallying ship is a procedure in which the ship is rocked, or sallied, by rapidly shifting weights back and forth, by
rhythmically heaving on the deck edge with a crane, or by personnel running back and forth. If, after inducing rolling, all exciting forces are
removed, the ship will roll with the time of roll equal to her natural rolling period. It is impossible to remove all exciting forces, but if the ship
is sallied in calm water, is clear of the bottom throughout her roll, the number of mooring lines has been reduced to the minimum acceptable
and those remaining are slack, and the ship is free of any other significant restraints, her rolling period will closely approximate the natural
rolling period, TR. GM can be estimated by means of the rolling period formula:
1.108 k 2 C B 2 0.44B 2

GM =
T
T
T
R
R
R

To determine the rolling period accurately, the ship should be timed through several rolls and the result divided by the number of rolls to find
the average rolling period. A derivation of the rolling period formula, with constants for various ship types, is given in Paragraph 1-5.4.3.
Sallying ship is often performed in conjunction with an inclining experiment as a check on the accuracy of the experiment or to provide a means
to calculate an initial estimate of GM. The accuracy of the procedure is degraded by the influences of offcenter weights, free surfaces, and
exciting or restraining forces, such as personnel moving about the ship, unslackened crane hoists or mooring lines, hydrodynamic effects of water
entrained by the moving hull surface in confined basins, etc.

1-56

S0300-A8-HBK-010

1-8.5 Ballast. A ships loading varies considerably during a voyage as fuel and stores are consumed, and for merchant ships and auxiliaries,
from one leg of a voyage to another as cargo is taken on and discharged. Ballast, liquid or solid, is carried to maintain stability or seakindliness.
As fuel is consumed from double-bottom tanks, the ships center of gravity rises and metacentric height is reduced. Saltwater ballast taken into
low tanks restores metacentric height to a safe value. All ships require certain drafts, displacement, and trim for seakindliness, propulsion
efficiency, and steering control. Discharge of cargo from forward holds and tanks trims the ship by the stern. A light draft forward causes
pounding and slamming in a seaway, reduces visibility from the bridge, and makes steering difficult in beam winds. Fuel and cargo oil tanks
were formerly used alternately as sea water ballast tanks in most ships. Environmental protection standards now prohibit discharge of oily water
in most areas, so modern ships are usually designed with dedicated or segregated ballast tanks (SBT). Normal practice is to provide ballast
capacity such that the ships displacement in ballast is 40 to 60 percent of the full-load displacement. Cargo tanks are often piped for ballast;
if the tanks have been cleaned prior to taking ballast, the ballast is clean and can be discharged overboard; otherwise the ballast is dirty and
is discharged to receiving facilities ashore. Ballast tanks are distributed over the length of the ship to provide flexibility in controlling trim and
hull bending moments. In general cargo ships, the combined center of the ballast tanks is usually near or below the combined center of the fuel
tanks. Ships designed to carry dense cargo, such as stone and ore carriers, have an excess of volume that is taken up by wing ballast tanks.
Some of these vessels are very stiff in light condition, so high ballast tanks are fitted to reduce metacentric height. Fuel tanks are still commonly
piped for saltwater ballast for emergency use. Many warships are fitted with compensating fuel tanks that admit seawater through openings in
the bottom of the tanks as fuel is drawn off the top, maintaining nearly constant weight and center of gravity in the tank.
Solid ballast, usually consisting of loose stone or sand, river mud, or other dredge spoil, is sometimes carried by cargo ships. Decomposing
organic material in mud ballast can produce flammable and toxic gases, such as methane or hydrogen sulfide. Solid ballast, carried in holds
or tween decks, can degrade stability by shifting, as explained in Paragraph 1-9.3.
Fixed solid ballast is sometimes fitted, particularly after conversions involving addition of high weight and in submarines. Ordinary concrete
or special heavy aggregate concrete is commonly used. The U.S. Navy has used cast iron ingots or lead pigs weighing about 60 pounds each.
The cast iron ingots are sometimes covered with a layer of 3 to 4 inches of cement mortar. High density drilling mud stowed in double-bottom
tanks is also used as ballast.
Ballasting instructions, where applicable, are included in the damage control book for Navy ships, and in the trim and stability booklet or loading
instructions for commercial vessels.

1-9 IMPAIRED STABILITY

A ships afloat stability can be impaired or otherwise changed by any of the following:

Addition, removal, or shift of weight, changing KG,

Change in the shape of the submerged hull from grounding or battle damage changing KM,

Free surface effect of loose liquids (FS), causing a virtual rise of G,

Free communication with the sea (FC), causing a virtual rise of G, or

Any combination of the above.

The first three conditions affect stability of the intact ship as well. Only free communication with the sea is predicated on damage to the hull.
As the primary indicator of initial stability, GM can be expressed as a function of the above effects:
GM = KM KG FS FC
The following paragraphs demonstrate the methods to calculate and apply the effect of these conditions on stability.
1-9.1 Flooding. Flooding can be caused by breaches in the hull, accumulating firefighting water, damaged saltwater systems, or any other
condition that admits uncontrolled amounts of liquid into the watertight envelope of the ship. Seawater flooding increases displacement and
reduces reserve buoyancy. Offcenter flooding causes list and reduces transverse stability. Major flooding towards the ends of the ship reduces
longitudinal stability, and in extreme cases may result in the loss of the ship by plunging. The effects of added weight on stability and trim
are addressed in Paragraph 1-8. In addition to the increased weight, loose water causes other serious consequences discussed in the following
paragraphs.

1-57

S0300-A8-HBK-010

1-9.1.1 Permeability. The effects of flooding are mitigated


by the contents of the flooded compartment. The space
occupied by solid objects or watertight volumes cannot be
occupied by floodwater, so total volume and weight of floodwater admitted is reduced. This effect is called permeability,
and a permeability factor, or ratio of the volume of
floodwater to the total volume of the space, can be defined:
available volume
=
total volume
The volume of the water entering a flooded space can be
determined by calculating the volume of the space and
multiplying by an appropriate permeability factor. The
permeability of tanks can usually be taken as the percentage
of full capacity to which they are filled to calculate the
amount of floodwater admitted. Not using a permeability
factor will result in overestimating the amount of water a
space contains. If the exact quantity of floodwater cannot be
determined, it is usually safest to err on the high side by
disregarding permeability. Permeability for cargo can be
calculated from cargo density or stowage factor, as explained
in Appendix E, U.S. Navy Ship Salvage Manual, Volume 1
(S0300-A6-MAN-010); the appendix includes an extensive
list of material densities and cargo stowage factors.
Permeabilities calculated from cargo stowage factors or cargo
densities may not be entirely accurate for breakbulk cargo in
rigid watertight packaging (cans, steel boxes, etc.) as water
will not be able to enter all void spaces in the cargo.
Permeability factors for some typical spaces and cargoes are
given in Table 1-8.
1-9.1.2 Downflooding. Downflooding occurs when a ship
heels sufficiently to immerse an opening above the normal
waterline, such as an open door or holed shell plating. This
angle of heel is defined as the downflooding angle. Righting
arms are reduced as the water accumulates on the low side,
and as an offcenter weight creates an additional upsetting
moment. A ship rolling so that it cyclically immerses a shell
opening may assume a permanent list or increase the period
and angle of roll due to the free surface effect described in
the next paragraph. As roll angle and period increase, the
time the opening is immersed increases, admitting greater
amounts of water.
1-9.1.3 Flooding into Liquid-filled Spaces. Tanks often
contain immiscible liquids, such as fuel or cargo oil, with
densities different from seawater. If an oil tank is holed,
there may be either a net inflow or outflow of liquid. There
may be an inflow even if the liquid level in the tank is above
sea level. If the density of the oil in the tank is low enough
that its head pressure at the hull penetration is less than the
seawater head pressure, water will flow into the tank. Head
pressure is a function of liquid depth and density:
Ph = h =

where:
Ph

1-58

=
=
=
=

head pressure
liquid weight density
liquid specific volume = 1/
liquid depth at point in question

Table 1-8. Selected Permeability Factors.

Miscellaneous Spaces on Naval and Commercial Ships1:


Permeability,
Space

Engine rooms (steam turbine)


fully flooded
above mid height
below mid height
lower third
Engine rooms (diesel and gas turbine)
Firerooms
Auxiliary machinery spaces
Pump rooms
Steering gear rooms
Shops
Offices, electronics spaces
Living spaces
General stores
Magazines
Powder
Small arms
Small arms ammunition
Rocket stowage
Torpedo stowage
Handling rooms
Chain locker

Full Load

Minimum Operating
Condition

0.85
0.90
0.75
0.70
0.85
0.90
0.85
0.90
0.85-0.90
0.90
0.95
0.95
0.80-0.90

0.85
0.90
0.75
0.70
0.85
0.90
0.85
0.90
0.85-0.90
0.90
0.95
0.95
0.95

0.60
0.80
0.60
0.80
0.70
0.80
0.65

0.95
0.80
0.95
0.95
0.95
0.95
0.65

Cargo Spaces:
Space
Tanks, empty, on molded volume2
Double-bottom tanks
Cargo tanks
Tanks of known capacity
Empty
With liquid contents
Bulk and breakbulk cargo (average)3
Container holds3
RO/RO holds (average)4
Liquids in cans or barrels1

Permeability,

0.97
0.99
1.00
1 - percent full
0.60-0.80
0.70
0.85
0.40

Notes:
1

From Naval Ship Engineering Center Design Data Sheet, DDS 079-1, Aug 75
See Paragraph 1-4.10.7 for discussion.
3
See Appendix E, U.S. Navy Ship Salvage Manual, Volume 1 (S0300-A6-MAN-010) for
discussion of how to calculate permeability/volume of floodwater from cargo stowage
factor/density.
4
Permeability of hold around containers; does not include space inside containers/
trailers.
2

S0300-A8-HBK-010

The equilibrium liquid level in the tank is the level that will give the same head pressure as the seawater. When there is an outflow of liquid
from the tank, the equilibrium level can be determined simply:
1 h1 = sw hsw

h1 = sw hsw
i
where the subscripts i and sw denote properties of the liquid inside the tank and of the seawater outside the hull, respectively. Since specific
gravity is directly related to density , the ratio of seawater to product specific gravities can be substituted for the density ratio. The outflow
of liquid lightens the ship, and may trim or heel it, varying hsw, so an iterative solution is required.
When there is an inflow of seawater into the tank, a water bottom forms. If the tank is holed at its bottom, hi remains essentially constant, but
lies over the water bottom of depth hsw,i. Equilibrium head pressure at the hull penetration is now expressed:
i hi + sw hsw, i = sw hsw
The inflow of seawater adds weight and may trim or heel the ship. It is possible that the liquid level will reach the tank top before equilibrium
is reached; the block of oil is held in place by sea pressure, and there can be no further weight addition, even if the ship continues to settle,
unless oil escapes through tank vents or other avenues.
Tankers carrying light oils that have suffered severe bottom damage may float in this manner, with much of the ships weight transmitted from
the tank tops to the water through the oil mass, rather than through the sides of the hull to the bottom structure. Since the lower level of the
liquid mass is above the hull penetration, and separated from it by a water bottom, there is little leakage in calm seas.
If the side of a tank is holed at a height such that the internal head pressure is less than the seawater head pressure, water will flow into the
tank. If the hole is low enough that it is covered by the water bottom, the situation is identical to that described above. If the hole is above
the top of the initial water bottom, there will be an ongoing oil-seawater exchange until the water bottom covers the opening.
The local seawater depth over a hull opening can vary with time as the ship rises, settles, trims, or lists in response to weight changes, or as
tide rises and falls around a stranded or sunken ship. Tanks may be subject to either inflow or outflow at different times. Heavily damaged
tanks will normally reach equilibrium in 20 minutes or less, although significant leakage will continue from casualties that strand at a tide that
is higher than subsequent low tides.
It is not always necessary to discharge a damaged tank completely to stop oil or other light liquids from leaking into the sea. The water bottom
formed when a tank is damaged near its bottom can prevent further discharge of liquids lighter than water. For example, in a tanker with a 50foot molded depth and a 30-foot draft, there is a 20-foot difference in head between sea level and oil level in full cargo tanks. If a full tank
is breached through its bottom plating, oil leaks out until the internal oil head balances the external seawater head. The depth of oil is
determined by converting the water head to an oil head. For the tanker described, and an oil specific gravity of 0.8:
hi =

g, sw
g, i

hsw =

1.025
30 = 38.44 ft
0.8

where:
hsw
hi
g, sw
g, i

=
=
=
=

depth to tank penetration = local draft for bottom rupture = 30 ft


oil depth with head equivalent to seawater head, ft
seawater specific gravity = 1.025
oil specific gravity = 0.8

For fresh water, specific gravity is taken as 1.0, and oil depth is found by dividing the draft or penetration depth by specific gravity; for the case
described above, the equivalent oil head is 37.5 feet. As a practical matter, the equilibrium oil depth has been reached when the cargo pumps
begin to draw water instead of oil. The thickness of the water bottom can be increased by drawing oil from the top of the tanks with portable
pumps, allowing water to flow in through the breached plating. In the initial stages of a pollution incident, salvors should attempt to create or
increase water bottoms in damaged tanks, especially if pumping or storage capacity is limited and several tanks are leaking. As operations
continue, water bottoms can be systematically increased until the tanks are completely discharged. Liquid and solid pollutants can be removed
by the methods discussed in Paragraphs 3-3 and 3-4, and the U.S. Navy Ship Salvage Manual, Volume 5 (S0300-A6-MAN-050).
The effectiveness of water bottoms is limited for water-soluble liquids or liquids with a specific gravity very near one. Water bottoms cannot
be created at all under liquids with specific gravities greater than one. Many bulk chemicals fall into this category, as well as some crude oils
and bunker fuels. Many chemicals are also highly soluble in water and cannot be contained by water bottoms.

1-59

S0300-A8-HBK-010

1-9.2 Loose Water. Liquid in a partially flooded compartment is free to move as the ship inclines. The adverse effects of loose water result
from the unrestrained movement of masses of water. The movement of significant weights causes the ships center of gravity to move off the
centerline as the ship inclines.
1-9.2.1
Free Surface Effect.
The
movement of the ships center of gravity
caused by loose water movement can be
related to the width of the free surface and
the angle of inclination. The loss of
righting arm results from the weight of a
wedge of water transferred from the high to
the low side, as shown in Figure 1-32. For
small angles, the volume of the wedge in a
rectangular tank can be calculated:
y (y tan)
dl =
Vwedge =
0
0
2
l

B
M
W1
W

Gv

L1

Z G2

y tan
dl
2

W
g

where:

l
y

=
=

length of the tank


half-width of the tank
(from its centerline)
angle of inclination

CL

Figure 1-32. Free Surface Effect.

For a rectangular tank, the centroids of the wedges are at 2 3y from the centerline of the tank; the plan area of most tanks approximates a
rectangle sufficiently to assume that the centroid of the wedge lies 2 3y from the centerline. The centroid of the transferred wedge therefore
moves a total distance of 4 3y. The moment of volume of the transferred wedge is:
l y 2 tan
l 2
4
d l y = tan
y 3 dl
moment of volume =

0
0
2
3
3

The integral 0l 2 3y3 dl is the second moment of area (moment of inertia), i, of the liquid surface (see Paragraph 1-4.5.2 for a derivation).
Substituting:
moment of volume = i tan
The weight moment of the transferred wedge is:
weight moment = f i tan
where f is the weight density of the fluid in the tank.
The weight shift and accompanying moment will cause a shift of the ships center of gravity parallel to the inclined liquid surface (and the
inclined waterline) to a new position G2:
f i tan
f i tan
=
GG2 =
W
w
Righting arms are reduced by the transverse shift of center of gravity; the transverse component of the shift GG2 is found by multiplying by
the cosine of the angle of inclination:
i tan
cos = f i sin
GG2 transverse = GG2 cos = f

w
w
The righting arm with free surface is found by subtracting the transverse shift of G from the righting arm without free surface:
f i sin
GZcorr = GZ GG2 transverse = GZ
w
where:
W
w

GZcorr

1-60

=
=
=
=

weight of the ship


weight density of the water in which the ship is floating
volume of displacement
righting arm corrected for new position of the center of gravity, G2

S0300-A8-HBK-010

The free surface correction is applied to the basic statical stability curve by graphical or tabular means in the same way the sine correction for
increased KG is applied (see Paragraph 1-5.10.1). The effect on stability of a free surface can be much greater than the effect of the weight of the
floodwater. The total correction is the sum of the corrections for each free liquid surface.
The component of the weight moment causing the transverse shift of center of gravity, f isin, is called the moment of transference. For many ships,
moments of transference are tabulated for each tank, with f expressed in long tons per cubic foot. Moments of transference are normally calculated
for a slack condition (50 percent full) and a full condition (100 percent for water tanks, 95 percent for Navy fuel tanks, 98 percent for commercial
vessel cargo tanks) for a series of heel angles. The free surface correction for each tank at each angle is obtained by dividing the moment of
transference by the ships displacement. Tabulated moments of transference are included in the damage control books of newer Navy ships.
Approximate moments of transference can be calculated by assuming a rectangular free surface:
moment of transference rectangle = f i sin = f
where:
l
b

3
lb 3
1 lb
sin =
sin
12

f 12
L1

=
=

compartment length
compartment width

For seawater flooding, where f is 35 cubic


feet per long ton, the expression reduces to:
l b 3 sin
moment of transference sw =
420

W
W1

L1

where l and b are measured in feet.

If a tank or flooded space is nearly full or


nearly empty, the liquid pockets when the
ship heels; that is, the liquid moves to
expose the deck or to cover the overhead, as
shown in Figure 1-33. Once the liquid
begins to pocket, the center of gravity, g, of
the liquid mass moves very little as heel
angle increases. The reduction in righting
arm is simply that of an offcenter weight of
known location. Model tests have shown
that pocketing normally decreases free
surface effect by approximately 25 percent.
The angle at which pocketing occurs can be
predicted by geometry. As the tank shown
in Figure 1-34 is inclined, a wedge of liquid
shifts from the high side to the low side.
The increase in water level on the low side
is equal to the decrease on the high side.
This distance (h) can be expressed as a
function of the tank breadth (b) and the
angle of inclination, :
b
h = tan
2
Pocketing occurs at angles of inclination
where h is equal to or greater than the
liquid depth in the tank (d) or the overhead
clearance (a) as shown in Figure 1-34.
Solving for :

2h
p = tan 1 p
b
where:

L
W1

Figure 1-33. Pocketing.

CL
b

l1
a
b
_ tan = h
2
y
w

d
w1

Figure 1-34. Pocketing Angle.

hp

angle of inclination where


pocketing begins
a or d, whichever is less

1-61

S0300-A8-HBK-010

Tabulated moments of transference


account for pocketing and tank shape.
When using approximate moments, a
statical stability curve can be constructed
by applying a free surface correction for
angles up to p, and an offcenter weight
(cosine) correction for larger angles.
Alternatively, the gradual diminishment
of the moment of transference can be
evaluated by defining the moment of
transference as the product of f i and
some factor C that is less than sin:
moment of transference = f i C
where:
f

fluid density (tank contents),


lton/ft3
moment of inertia of the
free surface, ft4
transference factor from
Table 1-9, 1-10, or 1-11

The moment of transference factor C


depends on the degree of fullness, ratio of
depth to breadth of the compartment, and
the angle of inclination. To simplify
evaluation of the factor C, tanks or flooded
spaces are assumed to be full or empty (no
free surface), half-full (worst case) or 95
percent full in naval practice or 98 percent
full in merchant practice (normal operating
condition). Tables 1-9 through 1-11,
reproduced from the Society of Naval
Architects and Marine Engineers
Principles of Naval Architecture, give
factors for 50, 95, and 98 percent full
tanks. These tables have been derived for
rectangular tanks but will provide
sufficient accuracy for most tanks if certain
adjustments are made to the entering
parameters of breadth and depth.
Tanks with substantial variation in
breadth, such as those that are
approximately trapezoidal in plan view,
usually have a small free surface effect;
the breadth at the narrow end should
generally be used to determine the depth
to breadth ratio. If greater accuracy is
required, breadth can be taken as:
3

b =

12 i
l

Table 1-9. Transference Factor Tanks 50 Percent Full.


Ratio of
depth to
breadth

10

20

30

40

50

60

70

80

90

0.1
0.15
0.2
0.25
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.2
1.5
2.0
3.0
4.0
5.0
6.0
7.0
8.0
9.0
10.0

0.13
0.17
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18

0.14
0.21
0.27
0.31
0.35
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36

0.14
0.21
0.27
0.34
0.40
0.50
0.57
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58
0.58

0.12
0.19
0.26
0.33
0.40
0.53
0.65
0.74
0.83
0.87
0.87
0.87
0.87
0.87
0.87
0.87
0.87
0.87
0.87
0.87
0.87
0.87
0.87

0.11
0.16
0.23
0.30
0.37
0.51
0.66
0.80
0.94
1.06
1.16
1.24
1.31
1.31
1.31
1.31
1.31
1.31
1.31
1.31
1.31
1.31
1.31

0.09
0.14
0.20
0.26
0.33
0.47
0.63
0.79
0.96
1.13
1.30
1.47
1.7
2.0
2.2
2.2
2.2
2.2
2.2
2.2
2.2
2.2
2.2

0.06
0.10
0.16
0.21
0.27
0.41
0.56
0.74
0.92
1.12
1.34
1.56
2.0
2.7
3.7
4.5
4.5
4.5
4.5
4.5
4.5
4.5
4.5

0.04
0.07
0.11
0.16
0.21
0.33
0.47
0.65
0.85
1.06
1.30
1.56
2.1
3.1
5.0
9.3
13.4
16.2
16.8
16.8
16.8
16.8
16.8

0.02
0.03
0.06
0.09
0.14
0.24
0.38
0.54
0.74
0.96
1.22
1.50
2.2
3.4
6.0
13.5
24.0
37.0
54.0
73.0
96.0
121.0
150.0

Angle of inclination, deg

Table 1-10. Transference Factor Tanks 95 Percent Full.


Ratio of
depth to
breadth

10

20

30

40

50

60

70

80

90

0.1
0.15
0.2
0.25
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.2
1.5
2.0
3.0
4.0
5.0
6.0
7.0
8.0
9.0
10.0

0.02
0.04
0.05
0.06
0.06
0.08
0.10
0.11
0.12
0.13
0.14
0.15
0.16
0.17
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18

0.02
0.04
0.05
0.06
0.07
0.09
0.11
0.13
0.14
0.16
0.18
0.19
0.22
0.25
0.30
0.36
0.36
0.36
0.36
0.36
0.36
0.36
0.36

0.02
0.04
0.05
0.06
0.07
0.09
0.11
0.13
0.15
0.17
0.19
0.20
0.24
0.28
0.35
0.46
0.53
0.57
0.58
0.58
0.58
0.58
0.58

0.02
0.03
0.04
0.06
0.07
0.09
0.11
0.13
0.15
0.17
0.18
0.20
0.24
0.29
0.38
0.52
0.64
0.74
0.80
0.85
0.87
0.87
0.87

0.02
0.03
0.04
0.05
0.06
0.08
0.10
0.12
0.14
0.16
0.18
0.20
0.24
0.29
0.38
0.56
0.71
0.85
0.97
1.09
1.16
1.22
1.27

0.01
0.02
0.03
0.04
0.05
0.07
0.09
0.11
0.13
0.14
0.16
0.18
0.23
0.29
0.38
0.58
0.78
0.96
1.14
1.30
1.46
1.6
1.7

0.01
0.02
0.03
0.03
0.04
0.06
0.08
0.10
0.12
0.13
0.15
0.17
0.22
0.28
0.39
0.62
0.87
1.12
1.36
1.6
1.9
2.1
2.3

0.01
0.01
0.02
0.03
0.04
0.05
0.07
0.09
0.11
0.14
0.16
0.18
0.23
0.31
0.45
0.77
1.12
1.5
1.9
2.3
2.7
3.2
3.6

0.00
0.01
0.01
0.02
0.03
0.05
0.07
0.10
0.14
0.18
0.23
0.28
0.41
0.64
1.14
2.6
4.6
7.1
10.3
14.0
18.2
23.0
28.5

Angle of inclination, deg

For tanks not rectangular in transverse section, the depth should normally be taken as the greatest depth. Accuracy can be increased by taking
depth as n times the distance from the free surface to the tank top, where n is 2 for tanks 50 percent full, 20 for tanks 95 percent full, or 50
for tanks 98 percent full. The tables should be entered with the next larger value for depth to breadth ratio unless interpolations are made. The
increase in accuracy gained by interpolation is usually insignificant.

1-62

S0300-A8-HBK-010

Computing moments of transference may


be time-consuming and tedious. Figure
1-32 shows that an equivalent righting
arm Gv Z can be developed by extending
the line of action of gravity back through
the ships centerline. Raising the ships
center of gravity to Gv has the same
effect on stability as shifting it to G2.
The virtual rise in the center of gravity
can be related to the actual transverse
shift:
GG2 = GGv sin
At small angles (less than 7 to 10
degrees), GZ = GMsin; the reduction in
righting arm is approximately GGvsin:
GZcorr = GM sin GGv sin
Setting the two expressions for GZcorr
equal:
f i sin
GM sin GGv sin = GZ
w
Noting that GMsin = GZ and canceling
common terms:

Table 1-11. Transference Factor 98 Percent Full.


Ratio of
depth to
breadth

10

20

30

40

50

60

70

80

90

0.1
0.15
0.2
0.25
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.2
1.5
2.0
3.0
4.0
5.0
6.0
7.0
8.0
9.0
10.0

0.01
0.02
0.02
0.03
0.03
0.04
0.05
0.05
0.06
0.07
0.08
0.08
0.09
0.11
0.13
0.16
0.17
0.18
0.18
0.18
0.18
0.18
0.18

0.01
0.02
0.02
0.03
0.03
0.04
0.05
0.06
0.07
0.07
0.08
0.09
0.11
0.13
0.16
0.22
0.27
0.30
0.33
0.35
0.36
0.36
0.36

0.01
0.02
0.02
0.02
0.03
0.04
0.05
0.06
0.07
0.07
0.08
0.09
0.11
0.13
0.17
0.24
0.30
0.35
0.40
0.44
0.48
0.51
0.54

0.01
0.01
0.02
0.02
0.03
0.04
0.04
0.05
0.06
0.07
0.08
0.09
0.10
0.13
0.17
0.24
0.31
0.38
0.44
0.49
0.55
0.60
0.64

0.01
0.01
0.02
0.02
0.02
0.03
0.04
0.05
0.06
0.06
0.07
0.08
0.10
0.12
0.16
0.24
0.31
0.38
0.46
0.52
0.59
0.65
0.71

0.01
0.01
0.01
0.02
0.02
0.03
0.03
0.04
0.05
0.06
0.06
0.07
0.09
0.11
0.15
0.22
0.30
0.38
0.46
0.54
0.62
0.70
0.78

0.01
0.01
0.01
0.01
0.02
0.02
0.03
0.03
0.04
0.05
0.05
0.06
0.08
0.10
0.13
0.22
0.30
0.38
0.48
0.58
0.67
0.77
0.87

0.00
0.01
0.01
0.01
0.01
0.02
0.02
0.03
0.04
0.04
0.05
0.05
0.07
0.09
0.14
0.23
0.34
0.45
0.58
0.70
0.84
0.98
1.12

0.00
0.00
0.01
0.01
0.01
0.02
0.03
0.04
0.06
0.08
0.10
0.12
0.17
0.27
0.47
1.06
1.9
2.9
4.2
5.8
7.5
9.5
11.8

Angle of inclination, deg

Virtual rise of G = F S = GGv =

f i

w
f i

w i
f
i
f

For flooding from the sea, the density ratio becomes one, and:
GGv =

where:
GGv
i

=
=
=

virtual rise of the center of gravity from free surface effect


transverse moment of inertia of the free surface
volume of displacement

If free surface exists in several tanks or compartments, the virtual rise of G is calculated separately for each compartment and the results summed
to determine the total virtual rise. The virtual position of the center of gravity is then used to develop a corrected stability curve, as described
in Paragraph 1-5.9.1.
Treating free surface effect as a virtual rise of the center of gravity provides a relatively quick and easy estimate of the reduction in initial
stability. The method overestimates the reduction in righting arm at larger angles because it does not account for pocketing or the reduction
in lever arms of the transferred wedge as heel angle increases, but is acceptably accurate if the sum of i for all slack tanks in ft4 is less than
twenty times the displacement in long tons. When virtually all free liquid surfaces are subject to pocketing at small angles, as in ships with
nearly full fuel load or cargo tanks, it is common practice to determine the reduction in righting arm (by transference) at an arbitrarily selected
angle of 5 or 10 degrees, and translate the reduction in righting arm into loss of metacentric height by dividing by the sine of the angle.
Equipment, cargo, or stores that pierce the floodwater surface reduce the area and effect of the free surface; this effect is called surface
permeability. The surface permeability factor is the moment of inertia of the actual free surface divided by the moment of inertia of an unpierced
plane surface with the same outer perimeter. Surface permeability is very difficult to estimate accurately. An error in estimation can cause the
salvor to believe the ship is more stable than it actually is. If, on the other hand, surface permeability is neglected, the calculations will indicate
less stability than the ship actually possesses, erring on the safe side for the salvor.

1-63

S0300-A8-HBK-010

1-9.2.2 Cross-flooding. Situations exist where, by design or damage, liquids can freely transfer, or cross-flood, between athwartships tanks:

Damaged longitudinal bulkheads.


Cross-flooding ducts fitted between shaft alleys, voids, and similar spaces in small ships to prevent the large offcenter weight
moments that would result if only one side flooded.
Faulty or inadvertently opened valves or valve manifolds, especially those connecting deep tanks where the liquid surface is above
the level of the valve.
Anti-roll tanks consisting of two tanks, normally carried about half-full, on opposite sides of the ship connected by relatively smalldiameter sluice pipes.

The shift of liquid from one space to another is treated as a moment of transference between the two tanks to determine reduction in righting
arm. The effect on initial stability, as a loss of metacentric height, is calculated for each tank separately.
1-9.2.3 Liquids of Different Densities. A tank may contain two different liquidsone of them is usually seawater. Examples include ruptured
cargo or fuel tanks and compensating tanks with water bottoms. Even if the tank is filled with liquid, there is a free surface at the interface
between the two liquids that will remain parallel to the inclined waterline. There will be a wedge of volume on the low side where the denser
liquid displaces the less dense, and a corresponding wedge on the high side where the less dense liquid displaces the denser, causing the center
of gravity of the tank to shift. This effect can be evaluated by using the difference in densities for the value f in the expressions for moment
of transference and virtual rise of G.
1-9.2.4 Bulk Cargoes. Bulk cargoes, such as grain and ore, and loose solid ballast, can produce an effect similar to that of free surface, but
the effect is modified by friction and inertia of the individual particles. In general, bulk cargo will begin to shift when the angle of inclination
is approximately equal to the angle of repose of the cargo. This is the angle between the horizontal and the slope of a granular bulk material
that is freely poured onto a horizontal surface. However, violent or cyclic ship motions or vibration can cause the cargo to shift at smaller
angles. A cargo that shifts during a heavy roll to one side will not necessarily shift back when the ship rolls to the opposite side. The tendency
to roll to greater angles on the low side can cause progressive cargo shifting that can lead to capsize. Some cargoes, especially certain ores,
may act like semi-liquid slurries in the presence of even a small amount of moisture, and shift readily when inclined.
Ships designed to carry bulk cargo, such as grain, are fitted with permanent or temporary longitudinal bulkheads in their holds that may be
supplemented with shifting boards to limit cargo movement. The cargo is normally pressed up to the tops of the holds and between the overhead
deck beams. If the cargo is not large enough to fill the hold, a portion of the grain is bagged and laid over the bulk grain to prevent shifting.
The cargo may also be tommed down by placing tomming boards, held in place by shores extending to the deck above, over the leveled cargo.
1-9.2.5 Free Communication Effect. A partially flooded, noncenterline space open to the sea introduces the effects of both offcenter weight
and free surface. In addition, floodwater is free to enter or leave the space as the ship inclines. The distribution and weight of floodwater varies
with time as the ship inclines. This creates virtual rise in the center of gravity, in addition to that caused by free surface:
Virtual rise of G = F C = GGc =

Ay2
1

where:
A =
y
=
1 =

plan area of the flooded compartment


transverse distance from the center of the flooded compartment to the ships centerline
volume of displacement to the after flooding to the waterline

Free communication exists only when the water level inside the damaged compartment remains the same as the sea level outside the hull. This
occurs only when the hull opening is relatively large compared to the volume of the space, and the compartment is vented.
1-9.3 Icing. Ice accumulation in freezing weather steadily adds high weight, increasing displacement and raising center of gravity. In severe
conditions, ice thicknesses of six inches or more can collect on weather decks in a short time. Ice builds up as spray or precipitation freeze
onto above-water structures. The rate of accumulation is therefore influenced by relative direction of winds and seas, and is seldom uniform
on both sides of the ship. The offcenter weight of accumulated ice will cause list that may cause increased ice accumulation on the low side,
especially if the primary source of ice is wind-driven spray.
High winds often accompany icing conditions; ice loading can severely degrade the ships ability to withstand heeling moments from beam
winds. As an example, a destroyer that has adequate stability for a 100-knot beam wind without ice meets the wind heel criterion (see Appendix
D) for only 80 knots with 200 tons of accumulated ice. The 200-ton ice accumulation corresponds to an average ice thickness of 5 to 6 inches
over those areas subject to icing. The effect is more severe on smaller vessels; 50 tons of topside ice on a 140-foot minesweeper reduces
maximum righting arm from 1.2 feet to 0.7 feet, and reduces maximum allowable beam wind from 85 to 40 knots.

1-64

S0300-A8-HBK-010

Once ice has started to form, it will continue to form as long as conditions favor icing. The only recourse is to remove the ice or leave the area
where ice formation is likely. Frequent heading changes can help prevent the accumulations of large weights of offcenter ice. Icing presents
particular difficulties to ships that are not free to maneuver, such as strandings and vessels under tow. The effects of accumulated weights of
ice (and snow) must be evaluated before refloating a heavily coated stranding. Removing ice from an unmanned vessel under tow may be
difficult or impossible; conditions favorable to icing are often also unfavorable for at-sea personnel transfers. At slow towing speeds, the time
needed to reach an area where conditions are significantly less favorable to icing may be considerable. Offcenter ice accumulation is likely on
towed vessels because tows follow a relatively steady course. It is important to ensure that a casualty has adequate stability under icing
conditions, or that heaters or other means to prevent icing be installed, if the casualty is to be towed through areas where icing is likely.
The U. S. Department of Commerce Publication Climatological and Oceanographic Atlas for Mariners provides guidance for expected winds and
icing conditions. In general, heavy to severe icing will occur when wind speed is greater than 30 knots and air temperature less than 28 degrees
Fahrenheit. Icing predictions can also be provided by Fleet Weather Centers and the National Oceanic and Atmospheric Administration (NOAA).

Limiting wind curves from damage control


books are based on specific loading
conditions, and the assumption that the
prescribed tank emptying/ballasting
sequence has been followed. They are not
valid for conditions that differ significantly
from these assumptions.
1-9.4
Added Weight Versus Lost
Buoyancy. The foregoing discussions have
assumed that flooding, with or without free
communication, increases the weight of the
ship by the weight of the floodwater. This
method, called the added weight method, assumes that none of the hull surface exposed
to the buoyant force of the water is lost.

9" ICE ON FOREDECK


90

5-100-3&4-F
5-250-1&2-F

85
80
WIND SPEED (KNOTS)

Damage control books for some Navy ships


include icing studies and limiting wind
velocity curves for various thicknesses of
accumulated ice.
Figure 1-35 is the
limiting wind curve for an FFG-7 Class
ship with 9 inches of ice on the foredeck;
there are also curves for 6 inches and 12
inches of ice. The fuel-ballast sequence
numbers refer to steps in the prescribed
tank emptying and ballasting sequence.
The plot is entered by reading vertically
from the appropriate fuel-ballast sequence
number to the solid wind heel curve, and
then horizontally to the maximum wind
speed for which the ship meets the Navy
wind heel criteria. The dashed lines show
the increase in allowable wind that can be
gained by ballasting the indicated tanks.
For example, at fuel sequence 6, the ship
has adequate stability to withstand 58-knot
beam winds with 9 inches of ice on the
foredeck. Continuing vertically along the
sequence 6 line shows that the limiting
wind can be increased to 62 knots by
ballasting tank 5-32-0-W, or 72 knots by
ballasting 5-32-0-W, 5-116-0-W, and 5326-1 and 2-W. If necessary, fuel tanks 5250-1 and 2-F, which are emptied by
sequence 4, can be ballasted to increase
limiting wind to 83 knots.

75

5-116-0-W
5-326-1&2-W
5-32-0-W

70
65
60
55
50
45

40
0

10

11

12

13

14

15

16

17 18

FUEL BALLAST SEQUENCE NUMBER


Figure 1-35. FFG-7 Class Ship Limiting Winds for Icing Conditions.

Table 1-12. Added Weight Versus Lost Buoyancy.

Item

Added Weight
yes

Lost Buoyancy
no

Change in volume of displacement

yes

no

Change in draft, trim, and list

yes

yes

Shift of center of gravity

yes

no

Shift of center of buoyancy

yes

yes

Shift of metacenter

yes

yes

Free surface correction required

yes

no

Free communication correction required

yes

no

Change in displacement

An alternative method, called the lost buoyancy method, can be used where floodwater in free communication with the sea is assumed to remain
part of the sea, and the flooded portion of the ship no longer contributes buoyancy. The vertical pressure forces about the flooded compartment
are assumed to act on the sea rather than on the ship.
Flooding in free communication with the sea can be assessed by either method, but the two methods cannot be mixed during calculations. Table
1-12 itemizes the important points of the two methods.
The method used is a matter of personal preference, although the added weight method is more commonly used. Unless otherwise specified,
hydrostatic and stability calculations in this book are made by the added weight method. A more complete discussion of the lost buoyancy
calculation method can be found in the Society of Naval Architects and Marine Engineers Principles of Naval Architecture.

1-65

S0300-A8-HBK-010

1-9.5 Loss of GM. Loss of GM can result


from added high weight (raised G) or increased displacement (lowered M), or both.
New initial righting arms are calculated
using the new value for GM. The stability
curve can be corrected for the new KG with
a sine curve correction as described in Paragraph 1-5.9. A ship with a very low
metacentric height will roll sluggishly. If
GM is negative, the ship is initially unstable and will loll to some angle where the
center of buoyancy has moved sufficiently
to begin to develop positive righting arms.
The ship will settle with equal facility to
the same angle of loll on either side. The
angle of loll may be estimated by:
= tan

2 GM
BM

POSITIVE RIGHTING ARMS, STBD


LOLLING RANGE
57.3

57.3
ANGLE
OF LIST

NEGATIVE
GM

POSITIVE RIGHTING ARMS, PORT

Figure 1-36. Stability Curve Showing Range of Instability (Lolling).

Where GM is the absolute value of GM.


When GM is negative, the corrected stability curve will indicate the list or angle of loll and a measure of the stability remaining beyond the
angle of loll as shown in Figure 1-36.
A warship or laden merchant vessel with negative metacentric height is in a very dangerous condition. A positive metacentric height should
be restored immediately. In general, negative metacentric height is dealt with by one of three methods:

Suppressing free surface to lower virtual height of the center of gravity,


Shifting weight downward in the ship, removing high weight or adding low weight to lower the center of gravity, or

Recovering lost waterplane to increase the transverse metacentric radius.


Free surface is suppressed by pumping from slack tanks directly overboard or by consolidating the contents of slack tanks to press up as many
tanks as possible. Partially flooded spaces should be dewatered if they can be made tight and pumped, or allowed to flood to the overhead.
When there are several slack tanks or partially flooded spaces, judicious selection of spaces to be pumped down can result in a simultaneous
suppression of free surface and a lowering of G. The effects of both the reduction of free surface and loss of low weight should be calculated
before emptying low tanks or spaces. In some cases, the net effect of pumping out is to raise the center of gravity unacceptablyflooding the
space from the sea would be more effective. The dewatering sequence should be arranged to avoid reducing GM dangerously while pumping
out. In ships with marginal stability, the transient free surface created while pumping down solid flooded spaces can cause loss of GM.
Shifting weights transversely to correct a list caused by negative GM will only aggravate an already dangerous situation. If enough weight is
shifted or added to bring the ship upright, it will list to the opposite side to an angle approximately twice that of the original list; the loll angle
is now added to the list due to offcenter weight.
1-9.6 Drydocking. A ship being drydocked is subject to an unusual loading situation; part of the ships weight is supported by keel blocks,
part by the surrounding water. This condition is complicated by changes in the size and shape of the submerged hull form as draft changes while
the dock is pumped out. This situation is analogous in many ways to that of a grounded ship, where part of the ships weight is supported by
the ground and part by water, and hull form changes with the state of the tide or passage of waves. The fundamental stability problem is to
determine whether the ship will remain stable from the time it first touches the blocks until it has completely settled, or landed, on them. On
undocking, the problem is whether the ship will be stable from the time it begins to leave the blocks until it is completely afloat. Positive GM
is taken as the indicator of adequate stability. The following discussion of docking stability is summarized from NAVSHIPS Technical Manual
(NSTM) 997, Docking Instructions and Routine Work in Drydock.
1-9.6.1 Block Reaction and Residual Buoyancy. When the keel of a ship begins to land on the blocks in a drydock, it pushes down with
an initial force w, causing a block reaction, P. A ship with trim, t, by the stern, will contact the aftermost keel block first. This block is called
the knuckle block because the ship pivots
on it. Strictly speaking, the knuckle rer1
action is not the entire block reaction, but
r
can be assumed to be in most cases. The
Gv
block reaction has two effects: a virtual
W
L
weight removal at the keel and a longG
B1
W1
L1
itudinal trimming moment. As the ship setP
tles on the blocks, P increases from zero
and is distributed over all the blocks. As
the water level falls, the distributed block
reaction increases until it equals the ships
weight, W.
The actual or residual
buoyancy, B, is equal to W - P. It is the
residual buoyancy that determines the
ships hydrostatic characteristics. Figure 1Figure 1-37. Drydocking Forces.
37 diagrams the forces on a ship during
drydocking.

1-66

S0300-A8-HBK-010

Draft at Landing. Summing longitudinal


moments about the knuckle block:

17

MEAN DRAFT, FEET

1-9.6.2 Docking Stability. Stability while


docking is analyzed either by evaluating the
effect of weight removal at the keel, or by
balancing moments about the point of first
contact. Draft at landing and draft at
instability (GM = 0) are determined and
compared. Figure 1-38 shows sample plots
for an FFG-7 Class ship.

ML = Wr B1r1

15

DRAFT AT LANDING
14.9 FEET

RESIDUAL BUOYANCY
MOMENT B1 r1

14
13

WEIGHT MOMENT, Wr
= 3,769(118.4)
446,250 FOOT-TONS

12
11
2.5

where:

3.0

4.0

3.5

4.5

5.0

MOMENT, FOOT-TONS x 10 5

=
=

B1

=
=

ships weight
distance from knuckle block to
LCG, as shown in Figure 1-38
residual buoyancy of the ship
at current draft
distance from knuckle block to
LCB, as shown in Figure 1-38

The weight moment (Wr) is constant while


the residual displacement and LCB vary
with draft. The draft at landing is the draft
where ML is zero with the keel parallel to
the tops of the keel blocks; that is, where
the weight and buoyancy moments are
equal, with B1 and r1 determined for the ship
with her keel parallel to the keel blocks.
Buoyancy moments can be calculated for a
range of drafts and plotted as shown in
Figure 1-38. The draft at landing is
indicated by the intersection of the weight
moment and buoyancy moment curves.

DRAFT AT LANDING
17

MEAN DRAFT, FEET

W
r

r1

16

16
15

W(KG)=
=3,769(18)
=67,842
FOOT-TONS

KM1 B 1

14
13
DRAFT AT INSTABILITY

13.25 FEET

12
11
5

10

9
4

MOMENT, FOOT-TONS x 10
DRAFT AT INSTABILITY

Figure 1-38. Drydocking Plots.

Draft at landing can be estimated by:


T1 = Tm -

P
12(TPI)

where:
Tl
Tm

=
=

draft at landing, ft
mean draft on entering the dock, ft

TPI

tons per inch immersion, lton/in

distance from application of P (knuckle block) to LCF, ft

The block reaction at landing, PL, is given by:


PL =

t (MT1)
h

where:
t
MT1

=
=

trim on entering the dock, in


moment to trim one inch, ft-lton/in

A rule of thumb for estimating draft at landing is:


T1 = Tmax -

2
(t)
3

where Tmax is the deepest draft on entering the dock, and Tmax and t are given in consistent units.
Draft at Instability. After touching the keel blocks, GM is given by:
GM1 = KM1 - KGv
where:
GM1
KM1

=
=

metacentric height after touching blocks


height of the metacenter after touching blocks

KGv

virtual height of the center of gravity

1-67

S0300-A8-HBK-010

The center of gravity undergoes a virtual rise due to the addition of negative weight at the keel. The height of the virtual center of gravity is:
KGv =

w(kg)
W(KG) ( P) (0)
W(KG)
=
=
W P
B1
w

It is useful to plot GM1 for various drafts to visualize the relationship between the metacentric height and draft while the ship is on the blocks.
The draft at instability is found by setting GM1 equal to zero:
W(KG)
B1

0 = KM1 KGv = KM1


KM1 =

W(KG)
B1

KM1 (B1) = W(KG)


By considering the products as moments and plotting moments against drafts as shown in Figure 1-38, the draft at instability is shown by the
intersection of the two curves. If this draft is less than the draft at landing by a comfortable margin, the ship should remain stable until firmly
supported by the keel blocks, or when it begins to leave the blocks on refloating. Example 1-4 illustrates the stability calculations for an FFG-7
Class ship entering drydock.

EXAMPLE 1-4

An FFG-7 Class ship with initial conditions as shown is to be drydocked.


Determine draft at landing and whether the ship will remain stable
throughout the docking.

=
=
=
=
=
=
=

408 ft
14 ft 3 in
16 ft 1 in
15 ft 2 in
3,769 tons
7.6 ft abaft midships
18 ft

The knuckle block will contact the keel at a point 330 feet abaft the forward
perpendicular.
From the Curves of Form (FO-2):

LCB
LCF
MT1
TPI
a.

r1
Tm
ft

Initial conditions:

L
Tf
Ta
Tm
W
LCG
KG

B1r1 as a function of draft:

=
=
=
=

3.2 ft abaft midships


23.8 ft abaft midships
773 ft-tons
32.5 tons

Initial estimates for draft at landing:


408
h = 330
23.8 = 102.2ft
2
t (MT1)
22 (773)
P =
=
= 166.4
102.2
h
P
166.4
T1 = Tm
= 14.74 ft
= 15.17
12 (32.5)
12 (TPI)

15.17
15.0
14.0
13.0
12.0
11.0

330 - [408/2 - LCB*]

B1
ltons
3,769
3,660
3,290
2,910
2,550
2,210

LCB*
ft
-3.18
-2.6
-0.08
2.6
5.32
8.1

r1
ft
122.82
123.42
125.92
128.6
131.32
134.1

B 1r 1
ft-tons
462,909
451,644
414,277
374,226
334,866
296,361

* from midships, negative values aft and positive forward

Wr and B1r1 are plotted as functions of draft in Figure 1-38, showing a


draft at landing of approximately 14.9 feet.
c. Draft at instability:

KM1(B1) =
W(KG) =

W(KG)
3769(18) = 67,842 foot-tons

KM1(B1) as a function of draft:


Tm
ft
15.17
15.0
14.0
13.0
12.0
11.0

B1
ltons
3,769
3,660
3,290
2,910
2,550
2,210

KM1
ft
122.82
123.42
125.92
128.6
131.32
134.1

KM1B1
ft-tons
84,049
81,764
73,992
65,882
58,089
50,609

or

Tl
t
Tl
b.

=
=
=

Draft at landing by plotting:

Ml =
r
=
Wr =

1-68

Tmax - 2/3(t)
22 in = 22/12 ft
16.08 - [2/3(22/12)] = 14.86 ft

Wr B 1r 1
330 - [408/2 - (-7.6)] = 118.4
3769(118.4) = 446,249.6 446,250 foot-tons

W(KG) and KM1(B1) are plotted as functions of draft in Figure 1-38,


showing a draft at instability of approximately 13.25 feet.
d. Margin between draft at landing and draft at instability:
Draft at landing - Draft at instability = 14.9 - 13.25 = 1.65 feet
Draft at landing exceeds draft at instability by 1.65 feet; the ship will be
completely settled on the docking blocks well before the residual
buoyancy ceases to provide adequate stability.

S0300-A8-HBK-010

1-10 SHIP CONSTRUCTION

Vessels are built to construction


specifications based on intended service.
Publicly owned vessels (Navy, Coast
Guard, etc.) are built to government
specifications. Most Navy ships are built to
General Specifications for Ships
(GENSPECs), although some auxiliaries are
built to commercial specifications.
Construction rules for commercial vessels
are established by classification societies
and government regulations for the country
of registry; the American Bureau of
Shipping (ABS) and United States Coast
Guard (USCG) establish construction rules
for the United States.
The hull structure consists of a watertight
grillage of stiffened plates supported by a
framework of mutually supporting
longitudinal and transverse members. The
framework and shell plating work together
to carry imposed loads. The framework
carries imposed loads and stiffens the shell
plating to allow it to function effectively as
a strength member under edge and lateral
loading. The arrangement of the structural
members is dictated by the framing system.
Structural members, with the exception of
shell plating and stanchions, are categorized
as either longitudinal, with their long axes
approximately parallel to the ships
centerline, or transverse, with their long
axes athwartships or vertical, approximately
perpendicular to the longitudinal members.
In a general context, any structural stiffener
can be called a frame, although the term is
usually reserved for the transverse frames
described in Paragraph 1-10.3.1.
1-10.1 Framing Systems. While ships
vary considerably in the details of their
construction, most conform to one of two
basic framing systems. Some reflect a
combination of the two systems. With
longitudinal and transverse structural
members crossing at right angles, only one
can be continuous. In the longitudinal
system, shown in Figure 1-39, this conflict
is resolved by the use of closely spaced
continuous longitudinal members with
intercostal transverses. The transverse
system, shown in Figure 1-40 (Page 1-70),
uses closely spaced continuous transverse
members with intercostal longitudinals.

INNER BOTTOM

CENTER
GIRDER
(KEEL)

MARGIN PLATE

BOTTOM DETAIL:
SLOT
FLOOR

FLAT
BAR

LONGITUDINALS AT NONWATERTIGHT FLOOR

FLOOR
THROUGH
BRACKET

SHORT
LENGTH

FLOOR

CONTINUOUS LONGITUDINAL
AT WATERTIGHT FLOOR

LONGITUDINAL

LONGITUDINAL CUT AT
WATERTIGHT FLOOR

Figure 1-39. Longitudinal Framing.

1-69

S0300-A8-HBK-010

In wooden ships and riveted steel


construction, continuity of the intercostal
members depends on the strength of the
joining connections; the intercostal
members contribute less direct strength to
the framing grillage and serve primarily to
stiffen the longitudinal members and shell
plating. With good alignment and modern
welding practices, full strength can be
maintained, regardless of the previous
assembly continuity of members.
In
modern, welded-construction ships, framing
systems are distinguished by the relative
size, number, and spacing of transverse and
longitudinal members.
Longitudinally
framed ships have many small, closely
spaced longitudinals, with fewer, larger,
and more widely spaced transverses;
transversely framed ships have many small,
closely spaced transverses, with fewer,
larger, and more widely spaced
longitudinals. For average merchant ships,
typical close spacing is 2 to 4 feet, typical
wide spacing is 10 to 15 feet.
Merchant ships and naval auxiliaries may
use either longitudinal or transverse
framing, depending on the service of the
ship. Generally, the same system is used
throughout the ship.
Most naval
combatants (except submarines) are
longitudinally framed, with transverse
framing near the bow and stern. Because
naval ships require a greater reserve of
strength to provide damage resistance, their
frame members are generally deeper and/or
more closely spaced than those of similarly
sized merchant vessels.
Appendix B
describes the construction and
characteristics of different types of ships.

MAIN DECK
STRINGER

TWEEN
DECK

WATERTIGHT
BULKHEAD

WATERTIGHT
BULKHEAD

GUNWALE
ANGLE

SHEER
STRAKE

DECK
BEAMS

DECK GIRDER

BEAM
KNEE

STANCHION

FRAMES

INNER BOTTOM

WATERTIGHT
FLOOR

CENTER
GIRDER

SIDE GIRDER

MARGIN PLATE

TANK
SIDE
BRACKET

LIGHTENING HOLE
PLATE
FLOOR

FLOOR

AIR HOLE

LIMBER HOLE
BRACKET

REVERSE BAR

BRACKET

BRACKET
FLOOR
FRAME BAR

1-10.1.1 Longitudinal Framing. LongiFigure 1-40. Transverse Framing.


tudinal framing systems (Figures 1-39A and
1-39B) are more efficient structurally, providing greater strength for the same weight; they are, however, less efficient in the use of internal space because of the deep web frames
supporting the longitudinals. Longitudinal framing has been widely used in tankers and bulk carriers where the disruption of internal spaces
caused by the web frames is unimportant. Modern practice tends increasingly towards longitudinal framing, or a combination system, in most
types of ships.
1-10.1.2 Transverse Framing. Transverse framing (Figure 1-40) is most often found in dry cargo vessels where deep web frames would
interfere with cargo stowage. Wooden ships are transversely framed. Given the load-carrying capacity of wood, the lack of longitudinal strength
of this system limits the maximum length of wooden vessels. Conversely, this system provides good resistance to racking stresses caused by
lateral forces that tend to distort a vessels cross section.
1-10.1.3 Combination Systems. There are framing systems that combine elements of both longitudinal and transverse framing. Figure 1-41
shows two common combination framing systems. The combination framing system was introduced to overcome the disadvantages of
longitudinal framing for dry cargo vessels. Longitudinal strength is provided by longitudinal framing in the double bottom and under the strength
deck; transverse framing is used along the side plating where longitudinal bending stresses are smaller. Plate floors and heavy transverse beams
are fitted at intervals to support the main deck and bottom longitudinals and increase transverse strength.
Cantilever framing is a modification of the combination framing system with some special features. It was developed to facilitate the building
of ships with very long and wide hatchways where the remaining deck structure provides insufficient transverse and longitudinal strength.
Transverse strength is maintained by the use of special web frames, or cantilevers, at frequent intervals abreast the hatchways. The ship is
strengthened longitudinally by heavier than normal sheerstrakes and deck stringer plates. The side plating may be extended upward at the
sheerstrake as a heavy bulwark, in place of the usual light bulwark or rails. Hatch side coamings are deep and may be continuous through the
length of the hatch deck. If the ship has two hatches abreast, a deck girder or longitudinal bulkhead is fitted on the centerline.

1-70

S0300-A8-HBK-010

1-10.1.4
Connections.
In riveted
construction, a variety of plates, angles, and
scarfs were used to create strong and rigid
joints between structural members. In
welded construction, most connections
between plates and shapes are made
directly through butt or fillet welds,
although brackets and angle bars are used
in some joints for extra stiffness.
1-10.2
Longitudinal Members.
Longitudinal structural members resist
bending about athwartships axes.

LONGITUDINALS
HATCH COAMING
STRONG BEAM BETWEEN HATCHWAYS

TRANSVERSE
BEAM

CANTILEVER

BEAM
TRANSVERSE
SIDE FRAME

LONGITUDINALS

LONGITUDINAL
FRAMING
IN BOTTOM

1-10.2.1 Keel. The keel is a major


longitudinal member that runs the length of
the ships bottom along the centerline. In
CANTILEVER FRAMING
COMBINATION SYSTEM
large ships, the keel normally consists of an
outer flat keel, the inner (plate) keel, a
vertical keel (sometimes called the center
Figure 1-41. Combination Framing Systems.
vertical keel, or CVK), and a horizontal top
flange called the keel rider plate. In small
vessels, the outer keel, vertical keel, and rider plates may consist of an I- or H-beam, while in large vessels, the keel is a built-up section. Duct
keels are flat-plate keels with two center girders, instead of one, on either side of the keel plates. Duct keels are commonly used forward of
propulsion machinery spaces to provide a pipe tunnel. The keel usually varies in cross section along the length of the ship. Some newer vessels
have no distinct keel. Instead, there is a cellular double bottom consisting of a grillage of heavy stiffeners plated over top and bottom. In this
system, the center girder is generally distinguishable from the side girders only by location. In very large, broad vessels, specially strengthened
longitudinals, called docking keels, are fitted at some distance to either side of the center keel. The docking keels help distribute docking loads
as the ship rests on three rows of keel blocks. In smaller vessels and some older merchant vessels, an outer vertical keel or bar keel is fitted.
In wooden vessels, the keel is usually a large timber, or series of timbers scarfed together. A timber keelson may be fixed atop the keel to
increase strength. In glass-reinforced plastic (GRP) vessels, the keel may be a wooden or metal member firmly bonded to the GRP skin, or may
consist of a multiple-fiber layup.
1-10.2.2 Other Longitudinal Members. Structural members that run the length of the vessel along shell plating or decks are variously termed
stringers, girders, or longitudinals. These members stiffen the entire structure against longitudinal bending loads, and reinforce shell and deck
plating against local loads. They may be built-up sections or standard structural sections. In the U.S. Navy, longitudinal members along the
side plating are called stringers; those along the bottom plating, longitudinals; and those under decks, girders. In large ships, heavy, deep, bottom
longitudinals may be fitted at some distance to either side of the keel. These members are often sized and located to carry the vertical loads
imposed by side blocks when dry docking. The heavy longitudinals are variously called sidegirders, keelsons, or docking keels.
Bilge keels may be fitted externally at the turn of the bilge to improve seakeeping by resisting rolling. Bilge keels are not usually structural
members; if they are attached by load carrying connections and extend for a significant length of the ship, they may contribute to the ships
longitudinal strength.
1-10.3 Transverse Structural Members. Transverse members are fitted primarily to stiffen the hull and enable it to resist shear and torsional
loads.
1-10.3.1 Frames. Transverse frames are analogous to ribs extending from the backbone of the keel inside the shell plating. They may continue
to the upper decks in their full cross section or be reduced in size at some height above the keel. Frame spacing and dimensions often vary
throughout the length of the ship to compensate for variations in loading. Intermediate partial frames may be added for local strengthening.
Web framesdeeper-than-normal frames with heavy flangesare often placed at intervals of several frame spaces, to stiffen and strengthen
the hull. Frames connect the longitudinal members and maintain spatial relationships in the face of shear and torsion. They also strengthen
the plating against bending under hydrostatic and dynamic loads or buckling under hull shear and bending, and act as ring stiffeners. U.S. Navy
practice is to number frames from the forward perpendicular (frame 0) aft; most foreign and many U.S. commercial vessels number frames from
aft forward. Frames forward of the forward perpendicular are designated by letters or negative numbers.
1-10.3.2 Floors. The portion of the frame from the keel to the turn of the bilge is a floor. Floors that do not continue into frames are
sometimes used for local strengthening or machinery foundations. Deep floorsdeeper than the standard floorsare used at the ends of the
ship and in high-load areas.

1-71

S0300-A8-HBK-010

1-10.3.3
Beams.
Athwartships deck
stiffeners are called beams.
They
strengthen the deck against local loads,
including hydrostatic loads for weather
decks, and contribute to overall ship
strength by increasing rigidity. Deck beams
normally join directly to frames at their
outboard ends, forming a continuous frame
ring. Triangular brackets, called beam
brackets or beam knees, are fitted to stiffen
the joint, or the beam is faired in to the
frame in a smooth arc to form a continuous
structure, as shown in Figures 1-39, 1-40,
and 1-41.

WEB

SHELL PLATING

RABBET
WRAPPER
PLATES

STEM
BAR

1-10.4 Shell Plating. Shell plating is the


side and bottom plating; i.e., those portions
of the ships skin that hold back the sea.
Bottom plating extends from the keel to the
turn of the bilge, side plating from the turn
of the bilge to or slightly beyond the upper
or main deck edge. Shell and deck plating
is arrayed in longitudinal strips called
strakes.
The strake adjacent to the keel is called the
garboard strake. The outer keel may be
incorporated into a keel strake. Strakes are
lettered from the keel outboard, starting
with the garboard strake as A. The strake
at the turn of the bilge is the bilge strake.
The uppermost strake, which joins to the
strength deck plating, is the sheer strake.
The keel, garboard, bilge, and sheer strakes
contribute significantly to longitudinal
strength, and are usually constructed of
heavier or stronger plate.

BREASTHOOK

COLLISION
BULKHEAD

PLATE STEM

SCARF
SCARF FOR
KEEL PLATE
GARBOARD
STRAKE

FLAT
PLATE
KEEL

FOREFOOT
CASTING

CASTING
BAR STEM
(OLDER, RIVETED, CONSTRUCTION)

BULBOUS BOW

1-10.5 Decks. Decks subdivide the vessel


into horizontal levels; weather decks also
close the top of the hull and maintain the
ships watertight integrity. Decks add
Figure 1-42. Stems
significant strength and rigidity to the
structure as a whole and limit the extent of
flooding after damage, provided they are or
can be made watertight. Decks may be steel or aluminum plating or wooden planking, and may be covered or sheathed with wood, tile,
linoleum, or other materials. The main deck is the highest continuous watertight deck and is usually the strength deck or upper flange of the
hull girder. Because of the main decks significance to hull strength and watertight integrity, it is used as the reference for numbering other
decks. The outboard strake of main deck plating is normally designated the main deck stringer and is either heavier or reinforced to provide
longitudinal strength. The connection of the deck to the sheer strake is critical to hull strength. Deck to sheer strake connections are often made
by means of a welded T-joint which may be backed up with an angle called the deck stringer angle or gunwale bar. Alternatively, the
connection may be made by means of a riveted gunwale bar, or the sheer strake may be rounded and butt-welded to the deck stringer. The U.S.
Navy uses the following definitions:

1-72

Platform or Platform Deck Deck extending less than the full length of the ship below the lowest complete deck; sometimes
called an orlop deck.

Flats Noncontinuous platforms between deck levels.

Half-Deck A partial deck above the lowest complete deck and below the main deck.

Forecastle Deck A partial deck above the main deck at the bow.

Poop Deck A partial deck above the main deck at the stern.

Upper Deck A partial deck above the main deck in the midships region, or one extending from the waists to either bow or stern.

S0300-A8-HBK-010

Decks above the main deck are called superstructure decks and may be referred to as levels. The term level also refers to nonwatertight
horizontal subdivision, usually by gratings of very deep compartments; for example, the upper level of a machinery space. In merchant ships
and auxiliaries, tween decks are often fitted to provide one or two levels above the hold bottom to allow cargo to be subdivided or carried high
to prevent stiff rolling.
1-10.6 Bulkheads. Bulkheads further subdivide levels or decks into compartments of varying size. Bulkheads may extend through one or several
decks and may be classed as structural, watertight, or joiner (also called partition or screen) bulkheads. Structural bulkheads are those that, by
design, contribute significantly to the ships strength. They stiffen the hull by resisting racking and torsional stresses and distribute vertical loads.
Watertight bulkheads are designed to withstand significant hydrostatic loads and are installed to increase the ships resistance to damage by
containing flooding. Transverse watertight bulkheads extend upward to a specified deck called the bulkhead deck. Bulkheads are strengthened
by angle or bar stiffeners where necessary, or are constructed of corrugated plate. Joiner or partition bulkheads separate and subdivide living,
working, storage or other spaces, but impart no watertight integrity or significant strength to the ships structure. Bulkheads often fit into more
than one class, although all bulkheads act as partitions. In practice, watertight bulkheads are almost always structural, while structural bulkheads
are often watertight.
1-10.7 Other Structural Members. The Stem Assembly (Figure 1-42) forms the bow of the ship. In its original and simplest form, still used
in wooden ships and boats, the stem or stem post consisted of a heavy, rectangular timber which is, in essence, an upward continuation of the
keel to which the side planking was attached. In ships of iron or steel construction, the stem was a rectangular forged bar attached at its base
to the keel, usually through a forefoot casting. This type of bar stem has been largely superseded by the plate stem, built up of curved wrapper
plates, although bar or heavy pipe stems are still commonly used on Great Lakes bulk carriers. The sharper portions of the stem are formed
by welding the side plates to an ordinary stem bar or length of round bar or tube, or by butt-welding the plates together. The entire assembly
is reinforced by a closely spaced network of deep floors, frames, stringers, and horizontal plate breasthooks. Vertical centerline stiffeners are
fitted in stems of large radius and bulbous bows.
Stern Assemblies, seen in Figure 1-43, close
the aft end of the hull and must accommodate propeller shafts and rudder assemblies, as well as resist the dynamic
loads imposed by the rudders. In singlerudder ships, a stern post or frame is fitted
at the aft end of the keel. It is generally
constructed of castings and forgings arranged to allow for the propeller shaft and
rudder stock bosses. The upper part of the
stern which extends past the rudder post is
supported by a special arrangement of
framing. This framing is carried by the
transom consisting of a deep, heavy transom floor in conjunction with a transverse
transom frame and beam. In counter sterns
(also called ordinary, overhanging, or elliptical sterns), which may be found in
older merchant vessels, a system of cant
framing radiates from the center of the transom like the spokes of a wheel. Cruiser
sterns have a system of transverse frames
and longitudinal girders with a number of
cant frames fitted abaft the aftermost transverse frame. Transom sterns are similar to
cruiser sterns, but end in a flat plate, called
the transom, and have no cant frames. In
twin-rudder vessels, the stern post is
omitted and the reinforced stern structure
extends forward of the rudder posts.

STERN CANT FRAMES

SIDE
GIRDER

TRANSOM
FLOOR

FLOORS
STERN
FRAME
CRUISER STERN

RUDDER
TRUNK

CENTER
GIRDER

TRANSOM
PLATE
FRAME

CRUTCH
STIFFENERS

A double bottom may be fitted to increase


FLOOR
strength and resistance to underwater
damage. The inner bottom plating is laid
over the grillage of floors and longitudinals,
forming spaces often used as tankage for
TRANSOM STERN
bunker fuel or other liquids. The outer
strake of the inner bottom is called the
margin plate, which may extend in a
Figure 1-43. Stern Assemblies.
horizontal line to the side plating, or be
inclined downward near the turn of the
bilge to form the side of the double bottom. The double bottom may or may not be continuous over the length of the ship. Large combatants
such as aircraft carriers and battleships may have more than one inner bottom.
Stanchions or pillars are used to support decks, distribute vertical loads, and stiffen the hull structure between bulkheads.

1-73

S0300-A8-HBK-010

1-10.8 Superstructures and Deckhouses. The term superstructure is applied to a portion of a ships structure above the main or upper deck
extending the width of the ship and forming an integral part of the main hull. A deckhouse is a lighter structure, usually not extending the width
of the ship, that is placed on the hull rather than forming a part of it. In practice, the two terms are often confused or used interchangeably.
In naval combatants and passenger liners, deckhouses or superstructures may extend for most of the vessels length; in most other types, they
occupy a small portion of the ships length. These structures generally house accommodation, communications, navigational, or control spaces.
They may house workshops or specialized machinery; in warships, weapons control spaces and weapons mounts are often located on or in the
superstructure or deckhouse. Deckhouses are not normally designed to contribute to overall hull girder strength, but being rigidly attached to
the hull, they carry some stresses. Superstructures, as an integral part of the hull, are normally designed to carry hull stresses.
1-10.9 Damage-resistant Features of Ships. While the entire structure of a ship is designed to resist some damage, certain features are
incorporated into ships specifically to prevent loss of the ship when damaged. Loss may result from flooding or structural failure of the hull
girder. Features enhancing a ships ability to resist damage are described in the following paragraphs.
1-10.9.1 Subdivision. Subdivision, or compartmentation, is a ships primary means of resisting damage. A system of watertight decks,
bulkheads, and an inner bottom limits the spread of flooding, fire, blast effects, weapon fragments, and fumes or gases. Extensive subdivision
is an inconvenience to everyone; production cost is increased, cargo storage is complicated, access and movement around the ship is hampered.
The degree of subdivision is therefore a compromise between safety and other requirements. Factors considered include the following:

Ability to resist battle damage.

Type Ship

Ability to resist bow collision damage.

Standard of Subdivision

Navy Ships (without side protection systems)1

Ability to resist damage from stranding.


Protection of vital spaces against flooding.

Seagoing craft under 100 ft in length

1 compartment

Ships 100-300 ft in length

2 compartments

Ships over 300 ft in length:

Ability to resist spread of fire, smoke, and


airborne contaminants.

Combatants and Personnel


Carriers, such as Hospital Ships
and Troop Transports

Withstand rapid flooding from a shell opening equal


to 15% of length between perpendiculars at any
point fore or aft

Interference of subdivision with arrangements.

All other ships

Withstand flooding from an opening equal to 12.5%


of the length between perpendiculars

Interference of subdivision with access and


systems.

Coast Guard Standards for Commercial Vessels2


Tankers over 738 ft in length

Withstand solid flooding from a shell opening with


length equal to the lesser of 0.495L2/3 or 47.6 ft,
width equal to the lesser of B/5 or 37.74 ft, from the
keel upwards without limit, at any point between
perpendiculars

Tankers between 492 and 738 ft in length

Withstand flooding from damage described above at


any point except at an aft machinery room bulkhead

Tankers less than 492 ft in length

Withstand flooding from damage described above at


any point between main transverse bulkheads,
except to an aft machinery room

Great Lakes dry bulk carriers

Withstand solid flooding from a shell opening with


length equal to the lesser of 0.495L2/3 or 47.6 ft,
width of 4.2 ft, from the keel upwards without limit,
between any two main transverse bulkheads

Barges carrying very hazardous materials

Withstand flooding from damage with length of 6 ft,


width of 30 in, from the keel upwards without limit,
at any point, including the intersection of a
transverse and longitudinal bulkhead

Provisions for carrying liquids.

1-10.9.2 Flooding. A principal concern in many casualty


situations is limiting flooding. Floodwater may be admitted
to the ship by collision, grounding, weapons strike,
firefighting, or other means. However flooding occurs, it
is necessary to limit its extent to minimize the following:

Table 1-13. Standards of Subdivision.

Ability to survive underwater damage.

Loss of transverse and longitudinal stability.


Loss of reserve buoyancy.
Damage to cargo and ship systems.

Ideally, a ship should be able to sustain increasing amounts


of flooding until it founders from loss of reserve buoyancy.
Barges carrying moderately hazardous
Withstand flooding from damage described above at
materials
any point, except on a transverse watertight bulkhead
Loss of transverse or longitudinal stability can cause a ship
1
to capsize or plunge, even when a sizable reserve buoyancy
Naval Ship Engineering Center Design Data Sheet, DDS079-1, Stability and Buoyancy of
U.S. Naval Surface Ships, 1 Aug 75
remains. Offcenter flooding and its serious effects on
2
Title 46, US Code of Federal Regulations (46 CFR), Subchapter S. Requirements have
transverse stability can be avoided by using transverse
been simplified. Additional definitions and exceptions apply. Subdivision requirements
subdivision only. Complete avoidance of longitudinal
for passenger ships are especially diverse.
watertight boundaries is not always possible or advisable,
but most modern ships follow a general pattern of transverse watertight subdivision, at the expense of admitting a larger volume of floodwater.
Some longitudinal subdivision is necessary to reduce free surface effect, especially in tanks. This subdivision normally takes the form of a
centerline bulkhead dividing the inner bottom into port and starboard tanks, or use of wing tanks smaller than the adjacent centerline tanks.
Sills, seen in Figure 1-44, or baffle plates are sometimes used to reduce the free surface effects of rolling or shallow flooding but are ineffective
against unchecked flooding. Transverse watertight bulkheads near the extremities of the ship limit flooding, and prevent the large and dangerous
trims that large amounts of floodwater at the ends of the ship would produce. Additional transverse watertight bulkheads are spaced to permit
the ship to remain afloat after a specific number of adjacent compartments, usually 1, 2, or 3, are flooded. The number of compartments that
can be flooded without causing foundering is the ships standard of subdivision or standard of flooding. For example, the FFG-7 Class frigate
shown in Figure 1-45 can remain afloat if any 3 of its 13 major watertight compartments are floodedit is said to be a 3-compartment ship.
Progressive flooding is defeated by carrying each watertight bulkhead intact from the bottom plating to a height above the expected flooding
water level. Watertight bulkheads are normally carried watertight to a specified deck, called the bulkhead deck. The bulkhead deck on most
designs is the main or weather deck and may be either a continuous or stepped deck. For the FFG-7 Class ship shown in Figure 1-45, the main
deck is the bulkhead deck. Standards of subdivision for Navy and commercial ships are given in Table 1-13.

1-74

S0300-A8-HBK-010

Ships are assigned a minimum freeboard based on the reserve buoyancy required to sustain flooding to their standard of subdivision without
foundering. This freeboard is measured from a margin line that represents the highest allowable waterline in a damaged condition. The margin
line is usually established near the bulkhead deck or a designated freeboard deck. Load lines for cargo ships and tankers or limiting draft marks
for warships are marked at a distance below the margin line corresponding to the required freeboard. If the load line or limiting draft mark is
not immersed before damage, and flooding is equal to or less than the standard of subdivision, the ship will remain afloat at a waterline at or
below the margin line after damage. Salvors may not be able to restore a ships required minimum freeboard; reduced freeboard must be
recognized as a loss of reserve buoyancy and damage resistance. This is particularly important if the casualty is to be towed some distance to
safe haven. In such a case, a salvage engineer may wish to calculate the standard of subdivision for the ship in its actual condition.
1-10.9.3 Likely Damage. Certain features
are incorporated into ships to isolate
common or likely forms of damage.
Because the ends of the ship are more
vulnerable to damage from collision or
grounding, a collision bulkhead is required
at about five percent of the ships length
from the bow, along with an afterpeak
bulkhead near the stern, enclosing the
propeller shaft penetration into the hull. A
second collision bulkhead may be required
in large ships. Watertight double bottoms
are required in some classes of vessels to
provide protection against grounding and
limited protection against underwater
weapons. Machinery spaces are segregated
from the rest of the ship by watertight bulkheads that (1) protect the ship from intense
machinery space fires, and (2) protect vital
equipment located in the machinery spaces
from flooding in other parts of the ship.
Sheer can prevent or delay progressive
flooding through deck openings when trim
is extreme, as shown in Figure 1-46.
Wing tanks, common in tankers, ore
carriers, and large combatants, limit flooding from damage to the sides. The effect
of offcenter flooding can be mitigated by
constructing the wing tanks with volumes
that are small compared to the center tanks
or holds, or by keeping wing tanks filled at
least to the waterline. A system of wing
tanks combined with a double bottom
produces, in effect, a double hull.
1-10.9.4 Structural Damage. Structural
failure is resisted by the use of materials of
consistent and known strength, and by building in reserve strength. Ships scantlings are
selected to result in bending stresses on the
order of 15,000 to 22,000 pounds per square
inch, considerably less than the yield stress
of shipbuilding steels (32,000 psi or greater).
This stress level is often contingent on
specified loading sequences and conditions,
particularly in very large tankers or bulk
carriers. Hull strength is addressed in greater
detail in Paragraph 1-11.

Figure 1-44. Effects of a Sill.

SECOND DECK (DC dk)

MAIN DECK
(BULKHEAD DECK)

AP

368

328

292

250 212 180

140

100 84 64

32 20

FP

Figure 1-45. FFG-7 Transverse Subdivision.

SHIP WITHOUT SHEER

SHIP WITH SHEER TRIMMED TO SAME DEGREE

Figure 1-46. Sheer Defeating Progressive Flooding.

1-10.9.5 Additional Features of Naval Ships. Both naval and merchant ships use the damage-resistant features previously described. Naval ships,
intended to go "in harms way," incorporate additional damage-resistant features in their construction. Naval ships will usually have more extensive
subdivision than merchant vessels, although some naval auxiliaries are built to classification society standards. Combatants are built with a much
greater degree of subdivision and greater reserve of strength than auxiliaries or merchant ships of the same size. Naval vessels often have multiple
machinery spaces segregated by watertight bulkheads, as well as auxiliary machinery spaces located remotely from the main machinery rooms.
Additional vital spaces, such as ship control stations or weapons spaces, are designated and protected by watertight subdivision. In all commissioned
vessels of the U.S. Navy, a damage control (DC) deck is designated. The DC deck, on which damage control equipment and stations are located,
is considered a vital space and is made watertight where feasible. Remote operators for certain vital piping and electrical systems are located on
the DC deck. The damage control deck is located high in the ship and is usually covered; fore and aft access is provided through watertight openings
in the main transverse bulkheads. Doors and nonwatertight fittings in main transverse bulkheads are not permitted below the DC deck. Doors
through transverse bulkheads into shaft alleys are not allowed; no penetrations are allowed through the collision bulkhead. In addition to armored
decks and side armor, large combatants, such as aircraft carriers and battleships, are fitted with underwater defense systems (also called side protective
or torpedo protection systems) consisting of layered wing and bottom tanks. These are alternately empty or liquid-filled to absorb the shock of
underwater explosions. The tank boundaries form a series of barriers that must be breached before major spaces are flooded.

1-75

S0300-A8-HBK-010

1-11

SHIP STRENGTH

1-11.1 Stresses in Ships. Ships, like all


structures, are subject to load-induced stress
and the resulting strains. Simple beam
theory is employed to predict ship
responses to various conditions of loading
by treating assuming the ships structure as
a built-up box girder bearing an distributed
load (weight of the ship and contents) and
supported by a distributed reaction
(buoyancy). Of principal concern are the
compound bending and shear stresses
resulting from the ships loading and wave
action.
Torsional stresses are also
important, and can be severely aggravated
by grounding in large ships. Stresses may
be divided into three groups:

Primary or Structural
Affecting the hull girder.

PRIMARY

SECONDARY

Secondary or Local
Affecting major substructures
or definable areas of the hull,
such as a hold or bulkhead.

TERTIARY

Figure 1-47. Deflections from Primary, Secondary, and Tertiary Stresses.

Tertiary Very localized,


affecting small areas of
plating or single stiffeners.

BUOYANCY

The distinctions among primary, secondary,


and tertiary stresses are illustrated by the
character of the accompanying structural
deflections, as shown in Figure 1-47. The
total stress on any portion of structure is
the sum of primary, secondary and tertiary
stresses that may tend to either reinforce or
cancel one another.
1-11.1.1
Structural Stresses.
The
principal structural stresses are caused by
the following conditions:

1-76

WEIGHT

HOGGING

WEIGHT
Weight and Buoyancy Distribution. Differences in buoyBUOYANCY
ancy and weight distribution
cause longitudinal bending
stresses and accompanying
shear stresses. An excess of
buoyancy in the midships
region with an excess of
weight near the ends of the
ship places the deck in tension and the keel in compression. The resulting convex
deflection is called hog or
SAGGING
hogging.
An excess of
weight in the midships region
and excess buoyancy near the
Figure 1-48. Hull Girder Bending.
ends places the deck in compression and the keel in tension. The concave deflection is called sag, or sagging. Long waves can impose hogging or sagging conditions as shown in Figure
1-48. Bending stresses are resisted by the longitudinal strength members, particularly those of the strength deck, sheer strake and
bottom. Bending stresses are normally greatest in the midships region of an intact ship, while maximum shear stresses occur in
the quarter length regions.

S0300-A8-HBK-010

Water Pressure. The distributed force of buoyancy, as


water pressure, is resisted by
the side and bottom plating
stiffened by a network of
frames, floors, longitudinals,
etc. All weight loads are
ultimately transmitted through
the ship structure to be borne
by water pressure. The differences in weight and water
pressure distribution produce
varying loads as shown in
Figure 1-49.

WATER PRESSURE

Racking. Transverse waves


alter the water pressure distribution around the ship, as
shown in Figure 1-50. The
unequal pressure distribution
tends to bend side plating and
transverse frames about a
horizontal longitudinal axis.
The transverse distortion is
called racking and is resisted
by shear stresses in the ships
structure. Racking stresses
are highest on the corners of
a ships cross section.
Racking is resisted by transverse bulkheads and frame
ring, particularly the corner
brackets.
Drydocking. Ships supported
by a single line of drydock
keel blocks will hog transversely. A cellular double
bottom stiffens the hull
against such hog, but additional lines of side blocks
are more effective.
Stranding.
Stranding
changes the bending stress
distribution in the hull girder
by altering the buoyancy distribution and introducing concentrated loading along the
bottom. Point loads similar
to those caused by docking
blocks, but naturally much
less predictable, result if the
ship strands on uneven or
rocky ground. Large ships
may sag transversely if
stranded over a narrow width
near the centerline.

WEIGHT LOADS

DEFLECTION
(EXAGGERATED)

Figure 1-49. Water Pressure.

DEFLECTION
(EXAGGERATED)
WATER
PRESSURE
WAVE
PROFILE

Figure 1-50. Racking.

1-77

S0300-A8-HBK-010

1-11.1.2 Local Stresses. Secondary and tertiary stresses result from localized loads such as the following:

Panting. Panting is an oscillatory motion of the shell plating, principally near the bow and stern of a ship, caused by uneven water
pressure as the ship passes through waves. The fore-end (and sometimes the after) structure is reinforced with a system of panting
beams, panting stringers, panting frames, breasthooks, and deep floors to withstand panting loads.
Pounding or slamming. Pounding occurs when the bows of a pitching ship clear the water and come down heavily. Pounding
is most severe in full-bowed ships in the bottom structure in the forward quarter length of the ship. In this pounding region,
plating and bottom stiffeners are often heavier and/or more closely spaced than in the rest of the ship.
Local Loads. Local strengthening enables the ship structure to carry loads caused by large local weights, such as machinery or
cargo. Similar measures are used to strengthen structure in way of fittings that transmit high loads, such as padeyes, winch
mounts, and kingpost foundations. The geometry of portions of the hull or fittings may cause stress raisers, requiring local
reinforcement to increase load-carrying capacity. Figure 1-51 shows some forms of local reinforcement.
Vibration. Vibration from engines, propellers, etc., causes stresses in various parts of the ship. Vibration-induced stresses are
resisted by local stiffening of areas in way of vibration sources.

1-11.1.3 Weapons Effects. Impact and


shock effects of airborne, underwater, and
contact explosions can cause severe and not
wholly predictable loads on ship structure.
Warships are constructed with this kind of
loading in mind, and are therefore strengthened to withstand blast and impact loads
over much of their structure. The exact
nature of this strengthening varies from
ship to ship but generally consists of closer
stiffener and bulkhead spacing than would
be found in an equivalent-sized merchant
ship or auxiliary. Weapons effects are
discussed in greater detail in the U.S. Navy
Ship Salvage Manual, Volume 3 (S0300A6-MAN-030).
1-11.2 Longitudinal Bending Stress. The
magnitude of the longitudinal bending
stresses in the hull girder is a function of the
total bending moment, cross-sectional area
distribution. The bending moment is a
function of the shear force distribution along
the ships length, which is in turn a function
of the ships load distribution. The hull is
assumed to be a statically loaded beam that
behaves in accordance with the theory of
flexure (see Paragraph 2-3). The downward
loads on the beam are the weights of the
component parts of the ship and any weights
carried on the ship. Upward loads are the
forces of buoyancy (and ground reaction or
block reaction for stranded, beached, or dry
docked ships).
Bending moment is
calculated by a double integration of the
static load curve.
The steps in the
longitudinal stress calculation are:

1-78

FREE-EDGE STIFFENING

FACE STRAP

PLATE STIFFENING

TRIPPING
BRACKETS

GUSSET

DOUBLER PLATE
MACHINERY
FOUNDATION

BILGE
KEEL

Determine longitudinal
weight and buoyancy
distributions.
Statically balance the ship on
still water or a wave.
Develop the longitudinal load
distribution or curve.

Figure 1-51. Local Strengthening.

Integrate the load curve to give shear forces.


Integrate the shear curve to give bending moments.
Determine which structure in sections of interest is effective.
Determine moment of inertia, section modulus and location of the neutral axis for sections of interest.
Calculate bending and shear stresses in sections of interest.

DEEP
FRAME

S0300-A8-HBK-010

These steps are examined separately in the following paragraphs. Amplifying information can be found in the Naval Ship Engineering Center
Design Data Sheet DDS 100-6, or any good naval architecture text. Examples 1-5, F-3, and F-5 demonstrate longitudinal strength calculations.
1-11.2.1 Load Curve. The load on the hull girder at any point is the difference between the buoyant force and weight at that point. This is
graphically represented by superimposing buoyancy and weight curves. The areas under the curves represent total buoyancy and total weight.
For a floating ship, the two areas must be equal, with their geometric centers in vertical line. Figure 1-61 shows the load curve developed for
Example 1-5. For the shear and bending moment integrations to close properly, the ship must be statically balanced; that is, weight and
buoyancy, as calculated by integration of the respective curves, should be within 0.5 percent, and LCB and LCG should be within one foot of
each other.
It is important to adopt sign conventions for the directions of forces and distances, and carry them through subsequent calculations. The
calculations in this handbook follow the intuitive convention that downward forces (weight) are negative and upward forces (buoyancy) are
positive, resulting in load curves that are predominantly positive over the middle portion for hogging hulls, and predominantly positive at the
ends for sagging hulls.
1-11.2.2 Buoyancy Curve. The magnitude of the buoyant force at any point is a function of the cross-sectional area below the water line and
the water density. The buoyancy curve will therefore follow the curve of areas. Areas of sections are most easily obtained from Bonjeans
Curves, shown in Figure FO-3 and described in Paragraph 1-3.11. Lines drawings, offsets, or general plans can also be used to determine
sectional areas by numerical integration. The still water buoyancy curve is developed by dividing sectional areas by 35 (cubic feet per long
ton of seawater) to convert to unit buoyancy (tons per foot) and plotting these values as ordinates.
A buoyancy curve based on ordinates taken from Bonjeans Curves will not include appendage buoyancy. If known, appendage buoyancies
can be added to the basic curve as rectangles or trapezoids. When appendage buoyancy is unknown, a simpler and generally adequate solution
is to assume that an appropriate appendage allowance (a fraction of full-load displacement) is distributed over the length of the ship. Final buoyancy ordinates are determined by an appendage allowance adjusted for the ships condition, i.e., the appendage allowance divided by actual displacement. Buoyancy ordinates multiplied by the adjusted appendage allowance plus one give adjusted buoyancy ordinates. Integrating the
adjusted buoyancy ordinates should give a correct total buoyancy equal to total weight. Appendage allowances are discussed in Paragraph
1-4.10.2.
As part of the regression analysis described
in Paragraph 1-7, Porricelli, Boyd, and
Schlieffer developed a method of
approximating the buoyancy curve for
merchant and auxiliary hulls with a series
of trapezoids. The method is reasonably
accurate for full-bodied ships (CB > 0.6).
The ship is first divided into three
segments: the parallel midbody (pmb), the
forebody (fb), and the afterbody (ab). The
forebody and afterbody are then divided
into two sections each.
A uniform
buoyancy distribution is assumed for the
parallel midbody and represented by a
rectangle. Ordinates are plotted at the
forward and after perpendiculars and the
boundaries of the sections of the hull and
connected by straight lines to form the
buoyancy curve. Buoyancy of the parallel
midbody (Bpmb), lengths of sections (Ln, bn)
and heights of ordinates (yn) are calculated
as shown in Figure 1-52.

y3

y3

y4
y2

y5

y1
b5
AP

b4

b3

Lab

L pmb

b2

b1
Lfb

L pmb = (1.74CB - 1.002)L

b1 = (0.61 - 0.615 CB)L

y1 = 0.04y3

Lfb

b2 = Lfb - b1
b3 = Lpmb

y2 = CB y3

Lab

= (1.186 - 1.17CB)L

= L - Lpmb - Lfb
Lpmb B Tm Cm
B pmb =
35

b4 = Lab - b5
b = 0.2L
5

FP

y3 = Bpmb /Lpmb
y4 = CB y3
y5 = 0.08y3

To facilitate summing weight and buoyancy


Figure 1-52. Approximate Buoyancy Curve for Full-Bodied Ship.
curves to develop the load curve, the
buoyancy curve is often stepped, that is,
approximated by a series of horizontal segments at a height corresponding to the mean buoyancy ordinate for that segment. The procedure for
stepping a curve is described in Paragraph 1-4.9. It is not necessary for the buoyancy curve to have the same number of segments as the weight
curve, although it is convenient for all of the bounding stations for the curve with fewer segments to coincide with stations on the other curve.
The load curve resulting when the two curves are summed will have the same number of segments as the curve with the most segments.
1-11.2.3 Weight Curve. Weight distribution tables or curves are often difficult to obtain, even though they are developed during the design
of the hull girder. For U.S. Navy ships, a Longitudinal Strength and Inertia Sections drawing is prepared, showing weight distribution, usually
for full load. A portion of the longitudinal strength drawing for FFG-7 Class ships is reproduced in Figure FO-4. The complete drawing includes section scantlings, similar to Figure 1-58, for a number of stations along the ships length. Format and content of longitudinal strength
drawings for Navy ships are more completely described in Appendix B.

1-79

S0300-A8-HBK-010

Weight distributions for Navy ships are tabulated or drawn for 20 standard ship segments between perpendiculars, plus one segment forward
of the forward perpendicular and one aft of the after perpendicular (22 segments). The segments forward and aft of the perpendiculars extend
from the perpendiculars to the ends of the ship and are not necessarily the same length as the segments between perpendiculars. Segments are
identified by the stations that bound them, numbered from 0 at the forward perpendicular to 20 at the after perpendicular. Weight distribution
is assumed to be uniform within each segment, producing a stepped curve. For cargo ships, tankers, etc., where loading may vary by
compartment, it may be more convenient to segment the ship by compartments. Weight distributions for a number of Navy ships are given in
Appendix B.
The weight curve from a longitudinal strength drawing or other source must be corrected for the ships actual weight distribution, including any
major alterations (SHIPALTS). Often this information is not available and weight change estimates must be made until the weight distribution
sums to the known ship displacement. If detailed weight curves are not available, weight distribution can be estimated by one of the methods
described in Paragraph 1-11.13.
1-11.2.4 Shear and Bending Moment Curves. A fundamental principle of beam theory is that at any point in an elastic beam:
P =

dS
dx

d 2M
dx 2

S = Pdx and M = Sdx = Pdx

where:
P
S
M

=
=
=

load
shear
bending moment

Vertical shear at any section is the sum of


the vertical forces to one side of the
section; the shear curve is therefore
developed by integrating the load curve (the
sum of the weight and buoyancy curves)
along its length, starting from either end of
the ship. The total positive area under the
shear curve should equal the total negative
area for static equilibrium. Shear is zero at
the ends of the ship; for most ships, shear
will be maximum near the quarter-lengths
and change signs near midships.

W >B
SAGGING SHIP
B >W

1-80

When P is 0, S is a maximum or
minimum and M is at an
inflection point.

LOAD, P

AP

FP

SHEAR, S

Bending moment at any section is the sum


of force moments about the section. The
bending moment curve is developed by
integrating the shear curve along its length.
Bending moment is zero at the ends of the
ship, and is maximum where shear changes
sign. The load and shear curves cannot be
defined mathematically, so graphical or
numerical methods are used to perform the
integrations, as shown in Paragraph 1-4 and
Appendix F. Several important relationships between the load, shear and bending
moment curves, illustrated in Figure 1-53,
act as checks on the completed curves:

B >W
S = 0, M AT
LOCAL MAX

P AT MAX,
S AT INFLECTION

MOMENT, M

P = 0, S = LOCAL
MAX/MIN, M AT
INFLECTION

AP

FP

B >W

B >W

HOGGING SHIP

When P is a maximum, S is at an
inflection point.
When S is 0, M is a maximum or minimum.

W >B

Figure 1-53. Load, Shear, Bending Moment Curve Relations and Conventions.

S0300-A8-HBK-010

The load and shear curves can be integrated from either end. Each integration should close to zero at the end opposite the beginning. Small
errors in closing are unavoidable if the areas under the weight and buoyancy curve are not precisely equal, and LCG and LCB are not coincident.
It is sometimes useful to integrate each curve twice, once in each direction, and compare the results. If the integrations close to zero, integrating
in the opposite direction will reverse the sign of the ordinate at each station, but will not change the magnitude. If the integrations do not close
precisely, integrating in the opposite direction will change the magnitude of the shear and moment ordinates at each station, and is a means of
estimating the error range of the calculated values. If the shear curve does not close, the sections of maximum shear and bending moment will
also shift somewhat when integrating in the opposite direction. For small errors in closing, the magnitude of the shear and bending moment
ordinates in the middle portion of the curve will be fairly reliable, but the ordinates near the ends of the ship should not be trusted.
A useful convention is to integrate the load curve from left to right (from aft forward) to develop the shear curve, and the shear curve from right
to left (from forward aft) to develop the moment curve. Following this convention, along with taking downward forces as negative, will result
in shear and moment curves with the features shown in Figure 1-53:

For sagging hulls:


(1) Positive shear on the left side of the plot (aft).
(2) Negative shear on the right side of the plot (forward).
(3) Negative (convex downwards) bending moment.

For hogging hulls:


(1) Negative shear on the left side of the plot (aft).
(2) Positive shear on the right side of the plot (forward).
(3) Positive (convex upwards) bending moment.

This convention is useful because the bending moment curves superficially resemble a sagging or hogging hull, as appropriate. Other
conventions may be encountered in ship design data. Shear curves that are the mirror image of the convention described above are common
and result when both shear and moment integrations are run in the same direction. U.S. Navy longitudinal strength drawings disregard the sign
of bending moments and shear forces and show all curves above the axis to save space. Example 1-5 calculates still water bending moment
and shear curves for an FFG-7 hull; the curves are illustrated in Figures 1-62.
1-11.3 Variations in Loading. Any change in weight or buoyancy distribution will alter the load curve.
1-11.3.1 Changes in Weight Distribution. Changes in weight distribution generally result from deliberate actions, such as taking on or
discharging cargo, ballasting, launching or recovering aircraft and boats, use of fuels or other consumables, or shifting weights. Weight
distribution can also be changed in a casualty by:

Flooding.

Major fires which consume flammable materials.

Spilled cargo.

Loss of structure or fittings.

Weight additions or removals change total weight, and therefore affect total buoyancy and buoyancy distribution. Weight shifts that significantly
alter trim also affect the buoyancy distribution.
Buoyancy distribution can change without an accompanying change in weight distribution. Such changes result from:

Waves.

Grounding.

Drydocking.

1-81

S0300-A8-HBK-010

1-11.3.2
Wave-induced Buoyancy
Distribution. In all but the stillest water,
buoyancy distribution changes constantly in
proportion to the variations in draft along
the ships length as successive wave trains
pass. A wave-induced buoyancy curve is
developed by superimposing a wave profile,
or series of profiles, on the ship profile,
instead of using a horizontal waterline, to
determine drafts at stations. An infinite
number of waves are possible; in practice,
it is usually sufficient to examine only
worst-case situations. Maximum midships
bending moments result from the two
situations shown in Figure 1-48. Ships are
designed to carry the stresses imposed by
these conditions, based on a trochoidal or
sinusoidal wave form with length equal to
the ships length (L). Standard wave
heights were formerly taken as L/20, and
then 1.1
L as ship size increased, but with
steady increases in ship length, these
formulae yield unrealistically large waves.
More recent ABS construction rules specify
different formulae for different ranges of
length, although Navy design practice still
uses the 1.1
L wave. Although artificial,
these assumed conditions have proven
adequate for design work; they are used
here to illustrate the procedures for
analyzing wave-induced stresses in ships.
The salvage engineer who finds it necessary
to evaluate the strength of a casualty
exposed to wave action should base his
worst cases on observed or expected waves
and the actual loading and structural
condition of the casualty.
Total bending moment is sometimes spoken
of as the sum of a still water bending
moment and a wave-induced bending
moment. The total bending moment is
simply the bending moment resulting from
the load distribution at that instant. The
bending moment can be evaluated by
adding to or subtracting from the still water
buoyancy curve or by starting from scratch
by superimposing a wave profile over the
Bonjeans curves to develop the buoyancy
curve, as shown in Figure 1-54. As before,
the area under the buoyancy curve must
equal the area under the weight curve.

STILL WATERLINE

DECREASED BUOYANCY

WAVE PROFILE

INCREASED BUOYANCY

SECTIONAL AREAS

WAVE PROFILE

BONJEANS CURVES
Figure 1-54. Wave-Induced Buoyancy.

R
h

L
r2
2R

EQUAL AREAS

LINE OF
CENTERS
r2
2R

STILL
WATERLINE
INITIAL PLACEMENT OF WAVE ON HULL PROFILE

Figure 1-55. Trochoidial Wave Form.

To ensure that shear and bending moment


integrations close, the ship must be
statically balanced on the wave; that is, the waterline must be adjusted until weight equals buoyancy and the center of buoyancy is in vertical
line with the center of gravity. When using Bonjeans Curves in the profile format, this is most easily accomplished by plotting the wave profile
to the same vertical and horizontal scales as the Bonjeans Curves on a piece of tracing paper. The wave profile is laid over the Bonjeans
Curves, with either the crest or trough at the midship station, as appropriate. Section areas are picked off as ordinates to a trial buoyancy curve,
which is integrated to determine buoyancy and LCB. If the first guess does not match buoyancy and weight within limits, successive calculations
are made, moving the wave up and down and trimming it until a position is found where buoyancy is within one percent of weight, and LCB
is within one foot of LCG. When the final position of the ship on the wave is determined, the section areas are converted to unit buoyancies
to plot a precise buoyancy curve that is used to determine the mean unit buoyancy over each segment of the ships length. Buoyancy and weight
curves are then summed to calculate the load curve; shear and bending moment integrations are conducted as for the still water condition. When
the rosette format Bonjeans Curves are used, drafts at each station must be determined by interpolation so the section areas can be read from
the curves. Alternatively, rosette format curves can be traced onto a profile of the ship. The horizontal scale of the ship profile (not the same
as the Bonjeans Curve area scale) is not critical, but should not be more than twice the vertical scale; if the horizontal scale is too great, portions
of the wave profile will be steep enough that small errors in plotting will cause significant errors in reading sectional areas.

1-82

S0300-A8-HBK-010

A trochoid is the curve traced by a point inside a circle as the circle rolls along a horizontal line, as shown in Figure 1-55. Coordinates for the
trochoidal wave form are developed from the relationships:
x = L

y = h

sin
+ h
360
2
1

cos
2

The relationships are not linear, so there is no fixed interval that will match the x interval to station spacing; x and y coordinates are
determined for values of from 0 to 360 at convenient increments, such as 30 degrees. Because the ordinates to the trochoidal wave do not
fall on Bonjeans stations, it is important to plot the curve carefully to minimize error. The area under a sagging trochoid is less than the length
multiplied by half the height, so the line of centers (see Figure 1-55) must be placed above the still water line for buoyancy to equal ships
weight (for a hogging wave, the line is placed below the still waterline). The area under a trochoid is equal to that of a rectangle with the same
length and an upper boundary formed by a line r2/2R below the line of centers. Since the circle describing the trochoid makes one revolution
in the ships length, L = 2R, and 2R = L/. For an L/20 wave, r = L/40, and:

r
2R

L 2

40
L

L2
1,600 L

L
1,600

= 0.00196 L

L
wave

20

As an initial estimate, the line of centers of the trochoidal wave should be placed 0.00196L above the still waterline.
If r is expressed as 0.55
L, L will cancel out of the ratio, giving no solution. For a 1.1
L wave, r is expressed as h/2, and:
2

r
2R

h 2

2
L

h 2
4L

0.785 h 2
L

1.1 L wave

For manual calculations, it is often simpler to use sinusoidal waves (y = Lsin), as they are not horizontal-scale dependent. The full wave form
is developed in 180 degrees, and ordinates calculated at even increments of are plotted at evenly spaced stations. If increments of are set
equal to 180 divided by the number of segments, the wave ordinate stations correspond to the Bonjeans curve stations, simplifying determination
of section areas. Sinusoidal waves are somewhat steeper than trochoidal waves. For fine-lined ships, maximum hogging moments will be lower
and maximum sagging moments higher than moments based on trochoidal waves of the same length and height. For full-bodied ships, both
hogging and sagging moments will be higher when based on sinusoidal waves. For a ship with block coefficient of 0.46, the standard 1.1
L
sine wave bending moment is 6 percent less than trochoidal for hogging and 2 percent higher for sagging. For a block coefficient of 1.0, the
standard sine wave bending moment is 11 percent higher for hogging and 9 percent higher for sagging.
1-11.4 Curve Scales. It is sometimes convenient to draw the load, shear, and bending moment curves on the same plan. To standardize
drawing size and simplify manual integration, the U. S. Navy has adopted the following scaling criteria for longitudinal strength drawings like
that shown in Figure FO-4.

Base length for all curves is 20 units. Base length corresponds to the length between perpendiculars, so the horizontal scale is
one unit = L/20 feet.
The mean heights of the weight and buoyancy curves are three units for the full load condition. Vertical scale for weight,
buoyancy, and load curves is one unit = W/3L tons per foot of length.
One square unit of area under the weight, buoyancy, or load curves represents L/20 W/3L = W/60 tons.
The shear curve is drawn so that one unit of ordinate represents two square units of area under the load curve; the vertical shear
scale is one unit = W/30 tons.
One square unit of area under the shear curve represents L/20 W/30 = WL/600 foot-tons.
The bending moment curve is drawn so that one unit of ordinate represents three square units under the shear curve; the vertical
moment scale is one unit = WL/200 foot-tons.

1-83

S0300-A8-HBK-010

Navy drawings use one inch as the base unit, but any convenient unit or multiple can be used. When there is no requirement to plot curves
on the same plan, it is more convenient to make all the integration calculations in the base units without scale conversions.
1-11.5 Section Modulus. From beam
theory, the bending stress () at any point
is given by:
My
=
I

LONGITUDINAL BULKHEAD
BELOW
DECK
OPENING

where:
1:4 SLOPE

bending moment at the


section in question
vertical distance from the
neutral axis to the fiber
(element) in question
moment of inertia of the
section in question about the
neutral axis

TRANSVERSE BULKHEAD BELOW


PLAN VIEW
SHADOW
IN DECK

SHADOW
IN DECK
DECK OPENING

This relationship shows that the maximum


tensile and compressive stresses will occur
in the beam elements furthest from the
neutral axis. The distance from the neutral
axis to the outer fibers is designated c. The
term I/c is sometimes calculated separately
and called the section modulus (Z or SM).
Substituting:
max =

Mc
M
=
I
Z

EFFECTIVE
BULKHEAD

1:4 SLOPE

SHADOW IN
BULKHEAD

STRENGTH DECKS

BRACKET
NONSTRENGTH DECK

TRANSVERSE BULKHEAD

If, as is common, bending moment is exSECTION A-A


pressed in foot-tons, moment of inertia in
in2-ft2, and distances from the neutral axis
Figure 1-56. Ineffective Shadow Zones at Discontinuities.
in feet, the calculation yields bending stress
in long tons per square inch. It is best to
convert tons per square inch to pounds per square inch for comparison with material strengths (normally tabulated in psi) and to avoid confusion
between long, short, and metric tons.
1-11.5.1 Effective Structure. Calculating the moment of inertia for a simple girder is straightforward; the relatively complex cross section
of a ship is another matter. Judgement must be used to determine which elements of the ships structure effectively contribute to longitudinal
strength. Elements that are subject to buckling, tripping and other forms of load shirking, or that are inadequately joined to the overall structure,
cannot be assumed to contribute to longitudinal strength. As load shirking by panels with a width-to-thickness ratio greater than 70 is likely,
contribution of unsupported plating panels should be limited to 70 times the thickness. Material not structurally continuous for at least 40 percent
of the length of the ship about the section being examined is assumed to be ineffective.
Only the net cross-sectional area of longitudinally continuous components of longitudinal strength members, excluding openings and ineffective
shadow areas forward and aft of openings or other discontinuities, are included when calculating the moment of inertia. The shadow area of
an opening is the area forward and aft of the opening between converging lines drawn tangent to the radiused corners at a slope of one transverse
unit to four longitudinal units, as shown in Figure 1-56. All structures, including longitudinal framing and other connected structures within
this area, are considered ineffective. For openings caused by damage or with sharp corners, lines bounding shadow areas should be drawn
tangent to points outside the area of wrinkled or upset plating, or at a distance equal to 30 times the plating thickness from the edge of the
opening, whichever is greater. Shadow areas adjacent to discontinuities such as the ends of longitudinal bulkheads, strength decks, and inner
bottoms, are bounded by lines with a 1:4 slope, as shown in Figure 1-56.

1-84

S0300-A8-HBK-010

1-11.5.2 Calculating Section Modulus. After the elements to be included have been selected, moment of inertia, I, is calculated by summing
second moments of area (ay2) of individual elements about an arbitrary axis. It is most convenient to sum moments about the keel (some
authorities prefer to use an assumed neutral axis). Moments of inertia (i) of elements with significant vertical dimensions are added to the
summed second moments of elemental areas. Moment of inertia about the keel (IK) is then:
(ay 2)

IK =

(i)

where:
IK
a
y
ay2
i

=
=
=
=
=

moment of inertia of section about the keel, in2-ft2


area of individual section element, in2
height of centroid of section element above the keel, ft
second moment of area of individual section element, in2-ft
moment of inertia of individual section elements, in2-ft2

Measuring areas in square inches and vertical distances from the axis in feet gives second moments of area (moments of inertia) in in2-ft2, rather
than the in4, ft4, cm4, etc., customarily used in other branches of engineering. Moment of inertia of a rectangle is equal to bh3/12, where h is
the height and b the breadth of the rectangle:
i =

bh 3
(bh) h 2
=
12
12

If area is given in square inches, and height


in feet, the units of moments of inertia of
individual elements are consistent with the
units of ay2.
Individual moments of inertia for inclined
or curved plates with significant vertical
dimensions are determined by calculating
the square of the radius of gyration (k) as
shown in Figure 1-57. Moment of inertia
can then be calculated from the definition
of radius of gyration.
i

h2
12

133.33
144

402
12

REFERENCE AXIS

y
y
h
g

ak2

To obtain i in in2-ft2, a must be given in


square inches, and k in feet. If the
inclined flat-plate section shown in
Figure 1-58 is 5 8-inch thick, 54 inches
wide, and inclined so that h is 40 inches,
then:
k2 =

ah 2
12

= 133.33 in2

h
h2

k2 = 12

i = ak2

Figure 1-57. Moment of Inertia for Inclined Plates.

= 0.926 ft2

i = a k 2 = 54 (0.926) = 31.25 in2 ft2


8

Since the neutral axis of the ships section passes through the centroid of the section, height of the neutral axis above the keel is found by
dividing the first moment of areas by the sum of areas of the section. The moment of inertia about the neutral axis is found by the parallel axis
theorem:
INA = IK Ad 2
where:
INA
IK
A
d

=
=
=
=

moment of inertia about the neutral axis, in2-ft2


moment of inertia about the keel, in2-ft2 = (ay2) + (i)
total area of individual section elements, in2 = (a)
height of the neutral axis above the keel, ft = (ay2)/(a)

1-85

S0300-A8-HBK-010

Once INA and height of the neutral axis are known, section modulus (INA/c) is easily calculated. The neutral axis is not usually equidistant from
the top and bottom flanges of the hull girder (strength deck and keel), so each flange has its own value for c and therefore Z. The summations
required to find height of the neutral axis and moment of inertia can be methodically performed in a tabular format. Table 1-14 is a sample
section modulus calculation for the ship section shown in Figure 1-58.
In an intact ship of uniform cross section, maximum bending stress occurs at the location of maximum bending moment. A vessels cross section
is not normally uniform throughout its length, but the scantlings at each section are selected by the designer to keep bending stresses within
acceptable limits based on the anticipated bending moment.

5 x 4 x 6.00#T
5 x 5 3/4 x 13.0#T

CL
20.5

2 7-1/2" x
.500 PL HY-80
SHELL
DOUBLER

5.75
15.3#

2 6"x0.75"
PL HY-80
SHELL
DOUBLER

30
SHADOW
25.5#

PL

HY-80

L 20

30.0 ABV BL
L 19

6 x 6 1/2 x 13.0#T
25

L 18

5 x 4 x 6.00#T

7.65#

"E"-20.4 PL
HY-80#

L 17
4 x 4 x 5.00#T

20

SHADOW

SHADOW

L 16

10.2# PL

21.0 ABV BL

L 15

FEET

6 x 4 x 7#T

L 14
15

6 x 4 x 8.00#T

L 13

"D"-12.75# PL

L 12

6 x 6 1/2 x 13.0#T

L 11
7 x 6 3/4 x 15#T

10

L 10
18 x 7 1/2 x 50#I-T

L9

8 x 7 x 22.5#T

L8

5
9 x 7 1/2 x 25#T
L6

25 x 13 x 162#
I-T CVK
0

L5
L3
L1

35.7#

"C"-15.3# PL HY-80

L7

L4

L2

PL F.K.

"B"-20.4# PL
2 9" x 0.75 PL M.S.
SHELL DOUBLER

"A"-38.25# PL HY80

CUTS

NOTE:
I - T SHAPES ARE FORMED FROM W SHAPES BY
CUTTING LOWER FLANGE FROM WEB, USUALLY
WITH TWO VERTICAL CUTS

Figure 1-58. Frigate Hull Section at Station 10.

1-86

BL

S0300-A8-HBK-010

Table 1-14. Section Modulus for FFG-7, Station 10.

Component

Mn Dk Girders, Inbd (7) - T


Mn Dk Grdrs, Outbd (4) - T
2nd Dk Girders, (10) - T
Mn Dk Plating, Inbd, less shadow zones
Mn Dk Plating, Outbd
2nd Dk Plating, Inbd, less shadow zones
2nd Dk Pltg, Outbd
"E" Strake
"D" Strake
"C" Strake
"B" Strake
"A" Strake
"E" Doubler, upper
"E" Doubler, lower
"A" Doubler
Side Stringers
L20 - T
L19 - T
L18 - T
L17 - T
L16 - T
L15 - T
L14 - T
L13 - T
L12 - T
L11 - T
L10 - T
L9 - T
L8 - T
Bottom Longitudinals
L7 - T
L6 - T
L5 - I - T
L4 - T
L3 - T
L2 - T
L1 - T
CVK (1/2) I - T
Flat Keel (1/2)
Totals

Dimensions

5 4 6#
5 5.75 13#
4 4 5#
(246 - 75) 0.375
84 .625
(225 - 90) .25
51 .25
93 .3125
162 .3125
84 .375
93.25 .5
96 .75
31.5 .5
30 .75
33 .75
6
6
5
5
6
6
6
6
6
6
6
6
6

6 13#
6 13#
4 6#
4 6#
4 7#
4 7#
4 8#
4 8#
6.5 13#
6.5 13#
6.5 13#
6.5 13#
6.5 13#

7 6.75 15#
7 6.75 15#
18 7.5 50#
8 7 22.5#
8 7 22.5#
9 7.5 25#
9 7.5 25#
25 13 162#
14 .875

y
(ft)

ay
(in2 ft)

ay2
(in2 ft2)

12.39
15.24
14.80
64.13
52.50
25.31
12.75
29.06
50.63
31.50
46.63
72.00
15.75
22.50
24.75

29.613
26.613
20.746
30.000
30.000
21.000
21.000
17.875
16.500
7.000
3.000
0.875
28.000
26.500
0.500

366.91
405.58
307.04
1923.75
1575.00
531.56
267.75
519.49
835.31
220.50
139.88
63.00
441.00
596.25
12.38

10865.16
10793.76
6369.87
57712.50
47250.00
11162.81
5622.00
9285.92
13782.66
1543.50
419.63
55.13
12348.00
15800.63
6.19

3.82
3.82
1.77
1.77
2.08
2.08
2.36
2.36
3.82
3.82
3.82
3.82
3.82

28.000
26.500
24.500
22.750
19.250
17.500
16.000
14.750
12.500
11.750
9.000
7.500
6.250

106.96
101.25
43.37
40.27
40.04
36.40
37.76
34.81
47.75
44.89
34.38
28.65
23.88

2994.88
2682.60
1062.44
916.09
790.77
637.00
604.16
513.45
596.88
527.40
309.42
214.88
149.22

4.42
4.42
10.60
6.63
6.63
7.33
7.33
16.38
6.13
598.95

5.500
4.500
4.25
2.750
2.000
1.500
1.000
1.500
0.073

24.31
19.89
45.05
18.23
13.26
11.00
7.33
24.57
0.45
8989.85

133.70
89.51
191.46
50.14
26.52
16.49
7.33
36.85
0.03
215549.69

a
(in2)

(ay)/a

8,985.85/598.95

IK for half-section =
INA for half-section =

(ay 2) + i
IK - Ad 2

=
=

215,549.69 + 971.75
=
216,521.44 - (598.95 15.012) =

216,521.44 in2 ft2


81,577.95 in2 ft2

INA for full section =

2INA for half-section

2(81,577.95)

163,155.90 in2 ft2

Depth - d
INA/c t
d
INA/c b

=
=
=
=

30 - 15.01
163,155.90/14.99
15.01 ft
163,155.90/15.01

=
=

14.99 ft
10,884.32 in2 ft

10,869.81 in2 ft

cDK
ZDK
cK
ZK

=
=
=
=

h or k*
(ft)

i = ah2/12
or ak2*
(in2 ft2)

7.75
12.50
1.88*
0.42*
0.26*
2.63
2.50

145.45
659.18
111.33*
8.23*
4.87*
9.04
11.72

1.33*

18.75*

2.08

3.18
971.75

15.01 ft

Notes: Areas and centroids for T-shapes taken from AISC Manual for Steel Construction, 8th Edition.
i of vertical web only

1-87

S0300-A8-HBK-010

Q = FIRST MOMENT OF AREA OF STRUCTURE OUTSIDE


AXIS WHERE STRESS IS DESIRED

B
MAX
NEUTRAL
AXIS

VERTICAL SHEAR

HORIZONTAL SHEAR

CROSS-SECTION
SHEAR STRESS
DISTRIBUTION

b = TOTAL THICKNESS OF HULL PLATING AND


EFFECTIVE LONGITUDINAL BULKHEADS
I NA = MOMENT OF INERTIA ABOUT NEUTRAL AXIS
S = SHEAR ON SECTION

= ISQb
NA

SQ

MAX = I MAX
NAb

Figure 1-59. Shear Stress in the Hull Girder.

1-11.6 Shear Stress. Shear stresses result from vertical shear, caused by the uneven force distribution along the ships length, and horizontal
shear, caused by longitudinal bending and racking, as shown in Figure 1-59. The shear force is distributed over the section, each element
contributing to the total. Shear stress distribution can be modeled by the theory of thin-walled sections, as explained in the Society of Naval
Architects and Marine Engineers Principles of Naval Architecture, but this method requires the evaluation of indefinite line integrals, and may
be too tedious for field calculations. For salvage calculations, shear stress, , along any horizontal axis BB can be adequately approximated by
the expression:
SQ
=
INA b
where:

S
Q
a
y
INA
b

=
=
=
=
=
=
=
=

shear stress
shear at the section in question
first moment of area about the neutral axis of the area of effective structure above axis BB
ay
area of individual structural element
vertical distance of individual structural elements from neutral axis
moment of inertia of the section about the neutral axis
total width of material resisting shear along axis BB, in

Moment of inertia is obtained as part of the section modulus calculation. The first moment of area, Q, is determined by summing the products
of areas and their distances from the neutral axis in the same manner that ay about the keel is determined in the section modulus calculation.
The material width, b, is normally twice the shell-plating thickness (to account for both sides), plus the thickness of effective longitudinal
bulkheads, i.e., those that extend from the strength deck to the bottom of the ship and are firmly anchored at both top and bottom. Consistent
units must be used, along with appropriate conversion factors. If moment of inertia and first moment of area are in the customary units of in2-ft2
and in2-ft, a conversion factor of 12 must be applied to obtain stress in units of force per square inch:
=

1-88

SQ
12 INA b

S0300-A8-HBK-010

SHEAR
STRESSES
SHEAR
ELEMENT

TRANSVERSE
FRAMES

SHEAR
FORCE, S

LONGITUDINALS

SHEAR STRESSES

SHEAR WRINKLES IN PLATE PANELS

Figure 1-60. Shear Stress.

Shear is normally determined in long tons, giving shear stress in long tons per square inch; shear stress, like bending stress, is converted to
pounds per square inch by multiplying by 2,240 pounds per long ton.
Shear stresses act in pairs, are equal on all four faces of a plane element, and are maximum on planes parallel and perpendicular to the shear
force, as shown in Figure 1-60. Because the paired stresses tend to change the angle between faces of an element and lengthen the diagonal,
shear yield in plating panels is evidenced by diagonal wrinkles.
The form of the expression implies that shear stress in any section is zero at the deck and keel and maximum at the neutral axis, where Q is
maximum:
SQmax
max =
12INAb
where:
Qmax = first moment of the area above neutral axis about the neutral axis
Although shear stress in the deck is very low, and may approach zero near the centerline, shear stress is not usually zero at the deck edge; the
expression does estimate shear stress in the middle portion of the side shell (where it is normally of greatest concern) accurately.

1-89

S0300-A8-HBK-010

EXAMPLE 1-5
STILL WATER BENDING MOMENT CALCULATION

This example illustrates the detailed still water strength calculations for an
FFG-7 Class ship, including steps to reconcile inconsistent data, and to
balance weight and buoyancy. Examples 4-5 through 4-12 in the U.S. Navy
Ship Salvage Manual, Volume 1 (S0300-A6-MAN-010) illustrate simplified
calculations for a simple barge.
For an FFG-7 Class ship in the 1/3 Consumed Stores loading condition,
calculate:
Deck and keel bending stresses for stations 3 through 17
Maximum shear stress
From the Damage Control Book (DC Book) loading summary (Appendix F):
1/3 Consumed Stores, Sequence 6 Fuel/Ballast:

Tf
Ta
TLCF
W
LCG
LCB
MT1

=
=
=
=
=
=
=

14' 8"
15' 8"
15.23' (LCF 23.79 ft abaft midships)
3748.15 tons
5.53 ft abaft midships
3.06 ft abaft midships
769.01 ft-tons

Tank

Clean Ballast:
5-34-0-W
5-116-0-W
5-328-1-W
5-328-2-W

Oily Ballast:
5-100-3-F
5-100-4-F
5-250-1-F
5-250-2-F

tons

lcg fm Comments
midships
ft

32.04
53.56
19.62
19.62

161.8
80.0
-141.1
-141.1

9.47
9.47
9.9
9.9

92.3
92.3
-59.8
-59.8

Weight

5-132-0-F
5-164-0-F
5-170-0-F
4-170-0-W

Total

19.21
44.00
16.31
11.84

68.4
36.9
29.0
29.7

Contaminated Oil Settling Tank


Waste Oil Retention Tank
Oily Waste Water Holding Tank
Sewage Collection, Holding and Transfer
(CHT) Tank

254.94 tons

From Curves of Form (FO-2) for TLCF = 15.23':


=
=
=
=

Fuel/ballast tanks, filled with fuel


for departure full load. Listed weights
are differences between weights of equal
volumes of fuel and seawater

Miscellaneous Holding Tanks:

Full-load Displacement = 3,951.79 tons

W
LCB
LCF
MT1"

Saltwater ballast tanks listed


as empty for full load

W = 4,224.83 - 254.94 = 3,969.89

3,750 tons
3.1 ft abaft midships
23.8 ft abaft midships
770 ft-tons

The difference between the corrected longitudinal strength drawing


displacement and the full-load departure displacement from the DC Book is:
3,969.89 - 3,951.79 = 18.1 tons

From Longitudinal Strength and Inertia Sections Drawing (FO-4):

W = 4,224.83 tons

or 4.6 percent. The discrepancy cannot be resolved further without


additional data. It is not necessary to constuct a corrected full-load curve
that would then be corrected for the actual loading condition. The two
corrections can be made simultaneously.

Scale Factors:
b. Initial Weight Curve for 1/3 Consumed Stores condition (3,748.15 tons)
Length
1 in.
Weight Ordinates 1 in.
Weight Area
1 in2
Shear Ordinates 1 in.
Shear Area
1 in2
Moment Ordinates 1 in.
a.

=
=
=
=
=
=

408/20
4,224.83/3L
4,224.83/60
4,224.83/30
4,224.83(408)/600
4,224.83(408)/200

=
=
=
=
=
=

20.4 ft
3.45 tons/ft
70.41 tons
140.83 tons
2,872.88 ft-tons
8,618.65 ft-tons

Resolution of discrepancies in raw data

The data from the DC Book and Curves of Form are in good agreement.
However, at equilibrium, LCB and LCG must be aligned vertically. The
Curves of Form give LCB for the ship with 0 trim. Assuming the same to be
true for the DC Book, the initial trim arm (BGL) is 2.47 feet (5.53 - 3.06).
The resulting trim would be:

t = W(BGL)/MT1 = 3,748.15(2.47)/769.01 = 12.04 in by the stern


This is consistent with the tabulated drafts. In constructing the weight and
buoyancy curves, it will be assumed that the actual centers of gravity and
buoyancy are on a vertical line 5.53 feet aft of midships.
There is a discrepancy of 273 tons between the full-load weights as given
by the DC Book (3,951.79 tons) and the longitudinal strength drawing
(4,224.83 tons). This discrepancy must be resolved as completely as
possible before proceeding. The longitudinal strength drawing is prepared
for the most extreme loading conditions. It is therefore likely that items of
weight were included that are not included in the operating full-load
departure condition described in the DC Book. The most probable items
that would be included for the longitudinal strength drawing but deleted from
the operational full load are saltwater ballast and waste-holding tanks that
would be presumed empty for the departure condition. An examination of
the full-load condition and tank capacity tables from the DC Book reveals the
following potential weights.

1-90

The weight curve is created by deducting the weight differences between the
full-load condition and the actual condition from the full-load curve at their
locations. The corrections to the full-load curve described in Paragraph a.
above are deducted at the same time. Examination of the DC Book loading
summaries for the full load and 1/3 consumed stores conditions reveals the
following weight differences:
Item

Full Load
Weight
tons

1/3 Consumed
Weight
tons

Difference
tons

lcg from
Midships
ft

Provisions and Stores


Dry provisions
Frozen
Chill
Clothing, Small Stores
Ship Stores
General Stores

13.95
4.84
4.79
0.31
3.49

9.29
3.23
3.19
0.21
2.33

4.66
1.61
1.60
0.10
1.16

9.0 fwd
20.0 fwd
20.0 fwd
145.5 fwd
4.0 fwd

Deck Gear
Flammable Liq & Paints
Bosun Storeroom
Medical Stores
Misc Storerooms

2.37
3.77
4.13
1.00
7.46

1.58
2.51
2.75
0.67
4.98

0.79
1.26
1.38
0.33
2.48

81.3 fwd
115.5 fwd
137.1 fwd
-176.0 aft
-68.5 aft

8.73
7.88
7.88

8.71
2.37
2.37

0.02
5.51
5.51

-94.4 aft
115.8 fwd
115.8 fwd

Potable Water
5-292-3-W
5-308-1-W
5-308-2-W

S0300-A8-HBK-010

Segment
Item

Full Load
Weight
tons

1/3 Consumed Difference


Weight
tons
tons

lcg from
Midships
ft

Lubricating Oil
3-272-2-F
3-278-2-F
3-286-2-F
3-208-4-F
3-236-1-F
3-236-2-F
3-292-8-F

3.50
4.00
2.75
0.95
1.05
1.05
0.92

2.35
2.68
1.84
0.63
0.70
0.70
0.61

1.15
1.32
0.91
0.32
0.35
0.35
0.31

-70.7 aft
-77.9 aft
-85.0 aft
-6.0 aft
-33.9 aft
-33.9 aft
-89.3 aft

32.12
32.12
65.69
28.43
33.60
33.60

0.00
0.00
63.60
22.21
0.00
0.00

32.12
32.12
2.09
6.22
33.60
33.60

92.3 fwd
92.3 fwd
75.5 fwd
51.8 fwd
-59.8 aft
-59.8 aft

46.47
2.54
1.21
1.33

23.18
0.53
0.25
0.28

23.29
2.01
0.96
1.05

-4.0 aft
-40.9 aft
-89.3 aft
1.7 fwd

29.81

8.54

21.27

Fuel Oil, Storage


5-100-3-F
5-100-4-F
5-116-1-F
5-140-1-F
5-250-1-F
5-250-1-F
Fuel Oil, Service
5-204-2-F
3-240-2-F
3-292-6-F
5-201-3-F

-1.4-0
0-1
1-2
2-3
3-4
4-5
5-6
6-7
7-8
8-9
9-10
10-11
11-12
12-13
13-14
14-15
15-16
16-17
17-18
18-19
19-20
20-20.6

5-132-0-F
5-170-0-F
5-164-0-F
Total:

0.00
0.00
0.00

9.61
4.08
2.12

-9.61
-4.08
-2.12

68.4 fwd
29.0 fwd
37.0 fwd

The ordinates for the weight curve are calculated by consolidating the
differences by weight segments, distributing the weight difference over the
length of the segment, and dividing the distributed weight difference by the
scale factor (3.45). The new weight curve ordinates are calculated in the
following table:

-1.4-0
0-1
1-2
2-3
3-4
4-5
5-6
6-7
7-8
8-9
9-10
10-11
11-12
12-13
13-14
14-15
15-16
16-17
17-18
18-19
19-20
20-20.6

0.21
0.62
1.37
1.59
2.93
3.12
3.11
3.56
2.86
2.01
3.76
3.76
3.49
1.99
4.21
3.55
2.57
2.40
1.88
2.35
1.95
0.29

lcg from FP
Moment
at midordinate lcg area
in.
in2
-0.70
0.50
1.50
2.50
3.50
4.50
5.50
6.50
7.50
8.50
9.50
10.50
11.50
12.50
13.50
14.50
15.50
16.50
17.50
18.50
19.50
20.30

-0.15
0.31
2.06
3.98
10.26
14.04
17.11
23.14
21.45
17.08
35.72
39.48
40.14
24.88
56.84
51.48
39.84
39.60
32.90
43.48
38.03
5.97

53.58

557.59

LCG of the curve is more than one foot from the known LCG (5.53 ft aft of
midships), so the curve must be adjusted to move the LCG forward. The
initial buoyancy curve is developed for comparison before correcting the
weight curve.
c. Initial Buoyancy Curve for 1/3 Consumed Stores condition (3,748.15
tons)

203.64

W1/3 = 3,951.79 - 203.64 = 3,748.15

Segment

1.40
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
0.60

Area
yl
in2

W
= area scale factor = 53.58(70.41) = 3772.85 tons
centroid = moment/area = 557.59/53.58 = 10.41 in fm FP
LCG
= centroid scale factor = 10.41(20.4) = 212.28 ft fm FP
= 212.28 - 204 = 8.28 ft aft of midships

-150.9 aft

Miscellaneous Tanks

0.15
0.62
1.37
1.59
2.93
3.12
3.11
3.56
2.86
2.01
3.76
3.76
3.49
1.99
4.21
3.55
2.57
2.40
1.88
2.35
1.95
0.49

Length
l
in.

Totals

JP-5
5-344-0-J

Ordinate
y
in.

Old
Weight
Dist Load Ordinate New Ordinate
Ordinate Difference wt diff/20.4 Difference Old ord - diff
dl/3.45
in.
tons
tons/ft
in.
in.
0.15
0.62
1.37
2.05
2.95
3.29
4.29
4.50
2.95
2.95
3.90
4.10
3.50
3.25
4.28
3.58
2.57
2.96
2.18
2.35
1.95
0.49

0.00
0.00
0.00
-32.14
-1.38
-12.28
-83.18
-66.04
-6.22
-65.95
-10.08
-23.61
-0.70
-89.01
-4.95
-2.20
0.00
-39.24
-21.27
-0.33
0.00
0.00

0.00
0.00
0.00
-1.58
-0.07
-0.60
-4.08
-3.24
-0.30
-3.23
-0.49
-1.16
-0.03
-4.36
-0.24
-0.11
0.00
-1.92
-1.04
-0.02
0.00
0.00

0.00
0.00
0.00
-0.46
-0.02
-0.17
-1.18
-0.94
-0.09
-0.94
-0.14
-0.34
-0.01
-1.26
-0.07
-0.03
0.00
-0.56
-0.30
0.00
0.00
0.00

0.15
0.62
1.37
1.59
2.93
3.12
3.11
3.56
2.86
2.01
3.76
3.76
3.49
1.99
4.21
3.55
2.57
2.40
1.88
2.35
1.95
0.49

The weight curve is integrated on these ordinates to determine total area


(weight) and longitudinal position of the centroid (center of gravity). The
integration is carried out in a tabular format:

The buoyancy curve ordinates are calculated by determining section areas


for each station from the Bonjeans Curves (FO-3), dividing the area by 35
to convert to unit buoyancy (tons per foot), and dividing the unit buoyancy
by the scale factor (3.45). Drafts at each station are calculated assuming
no hog or sag. Before calculating ordinates from the section areas, the area
curve is integrated to compare total buoyancy and LCB with total weight and
LCG from the weight curve. The integration is performed by Simpsons rule
on 21 stations:
Station Draft
T
ft
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

14.67
14.72
14.77
14.82
14.87
14.92
14.97
15.02
15.07
15.12
15.17
15.22
15.27
15.32
15.37
15.42
15.47
15.52
15.57
15.62
15.67
Sums

Ordinate
Multiplier
(Section Area)
y
m
ft2
2
55
131
205
270
326
379
428
471
499
515
519
500
470
418
357
285
215
153
95
41

1
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
1
18,963

h
V
W
LCB

=
=
=
=
=

f(V)

Lever

f(M)

ym
ft3

s
ft

s (V)
ft4

2
220
262
820
540
1,304
758
1,712
942
1,996
1,030
2,076
1,000
1,880
836
1,428
570
860
306
380
41

0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

0
220
524
2,460
2,160
6,520
4,548
1,1984
7,536
1,7964
10,300
22,836
12,000
24,440
11,704
21,420
9,120
14,620
5,508
7,220
820
193,904

20.4
(h/3) (V) = (20.4/3)(18,963) = 128,948.4 ft3
V/35 = 128,948.4/35 = 3684.24 tons
h(M)/(V) = 20.4(193,904)/(18,963) = 208.6 ft fm FP
208.6 - 204 = 4.6 ft aft of midships

1-91

S0300-A8-HBK-010

d. Adjusting Weight and Buoyancy Curves


The weight and buoyancy curves disagree by 88.61 tons on total area. This
error is undesirable, but probably tolerable. The 3.68-foot separation
between the centers of gravity and buoyancy is excessive and must be
corrected. The ordinates of both curves must be adjusted to bring the
centers of gravity and buoyancy to within one foot of each other and within
one foot of the point 5.53 feet abaft midships.
Total buoyancy is corrected first by gradually increasing the area curve
ordinates until the buoyancy (area under the curve divided by 35) equals
total weight. There is a greater probability of error in reading the section
areas for the middle stations because the Bonjeans Curves for the middle
stations slope more gently than those near the ends. The corrections are
therefore weighted towards the center of the curve. LCB is then moved aft
by transferring a strip of uniform thickness from the forward half of the curve
to the aft half. The thickness of the strip is determined by trial and error.
After several iterations, the following section areas were determined:
Station

Ordinate Multiplier
(Section Area)
y
m
ft2

0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

0
54
134
204
274
329
379
430
475
510
524
524
509
479
429
369
299
229
167
109
59

Sums

h
V
W
LCB

1
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
2
4
1

f(V)

Lever

f(M)

ym
ft3

s
ft

s (V)
ft4

0
216
268
816
548
1,316
758
1,720
950
2,040
1,048
2,096
1,018
1,916
858
1,476
598
916
334
436
59

0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

19,387
=
=
=
=
=

0
216
536
2,448
2,192
6,580
4,548
12,040
7,600
18,360
10,480
23,056
12,216
24,908
12,012
22,140
9,568
15,572
6,012
8,284
1,180
199,948

20.4
(h/3) (V) = (20.4/3)(19,387) = 131,831.6 ft3
V/35 = 131,831.6/35 = 3,766.62 tons
h(M)/(V) = 20.4(199,948)/(19,387) = 210.4 ft fm FP
210.4 - 204 = 6.4 ft abaft midships

Now that the total buoyancy and location of LCB are both acceptably near
the known values, the buoyancy curve ordinates are calculated:
Station
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

1-92

Section
Area
ft2
0
54
134
204
274
329
379
430
475
510
524
524
509
479
429
369
299
229
167
109
59

Unit Buoyancy
B = A/35
tons/ft
0.00
1.54
3.83
5.83
7.83
9.40
10.83
12.29
13.57
14.57
14.97
14.97
14.54
13.69
12.26
10.54
8.54
6.54
4.77
3.11
1.69

Ordinate
B/3.45
in.
0.00
0.45
1.11
1.69
2.27
2.72
3.14
3.56
3.93
4.22
4.34
4.34
4.22
3.97
3.55
3.06
2.48
1.90
1.38
0.90
0.49

LCG of the initial weight curve is moved forward by transferring strips of


uniform thickness from segments in the after half of the curve to the
corresponding segments in the forward half, and by reducing some
ordinates in the after half to lower total weight slightly. The thickness of the
strips are determined by trial and error. After several iterations, ordinates
were determined and integrated as follows:
Segment

-1.4-0
0-1
1-2
2-3
3-4
4-5
5-6
6-7
7-8
8-9
9-10
10-11
11-12
12-13
13-14
14-15
15-16
16-17
17-18
18-19
19-20
20-20.6

Ordinate
y

Length
l

Area
yl

in.

in.

in2

0.15
0.66
1.41
1.63
2.97
3.16
3.15
3.60
2.90
2.05
3.80
3.72
3.45
1.95
4.17
3.51
2.53
2.36
1.82
2.29
1.86
0.49

1.40
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
0.60

0.21
0.66
1.41
1.63
2.97
3.16
3.15
3.60
2.90
2.05
3.80
3.72
3.45
1.95
4.17
3.51
2.53
2.36
1.82
2.29
1.86
0.29

Totals

53.49

lcg from FP Moment


at
lcg Area
Midordinate
in.
in2
-0.70
0.50
1.50
2.50
3.50
4.50
5.50
6.50
7.50
8.50
9.50
10.50
11.50
12.50
13.50
14.50
15.50
16.50
17.50
18.50
19.50
20.30

-0.15
0.33
2.12
4.08
10.40
14.22
17.33
23.40
21.75
17.42
36.10
39.06
39.68
24.38
56.30
50.90
39.22
38.94
31.85
42.37
36.27
5.97
551.90

W
= area scale factor = 53.49(70.41) = 3766.51 tons
centroid = moment/area = 551.90/53.49 = 10.32 in fm FP
LCG
= centroid scale factor = 10.32(20.4) = 210.53 ft fm FP
= 210.53 - 204 = 6.53 ft aft of midships
The adjusted weight and buoyancy curves are shown in Figure 1-61.
e. Shear and Bending Moment Curves
Ordinates to the load shear and bending moment curves are determined by
a continuous tabular calculation. Curve segments are identified by the
bounding stations in the first column. The weight ordinates are written in the
second column. The mean buoyancy ordinates for each segment are
written in the third column. The load ordinate in the fourth column is found
by subtracting the weight ordinate (column 2) from the mean buoyancy
ordinate (column 3). The load curve is integrated along its length by
keeping a running total of the area under the load curve in the fifth column.
In keeping with the convention of integrating the load curve from left to right,
the area total is run from bottom to top in this table. The area for each
segment is the ordinate multiplied by the segment length (1 inch for all but
the two end segments). The area total is the area up to the forward station
of the segment. The shear ordinates in the sixth column are determined by
dividing the areas in column 5 by two. The shear curve defined by these
ordinates is shown in Figure 1-62.
The shear ordinates are carried into the following table and written in the
second column, next to the appropriate station (column 1). It is necessary
to interpolate the x intercept (station 10.41) to properly integrate the curve
and to determine the section of maximum bending moment. The mean
shear ordinate for each segment is written in the third column. The shear
curve is integrated along its length from forward aft (top to bottom); the
running total is written in the fourth column. The shear areas are divided by
3 and written in the fifth column as the moment ordinates. The resulting
bending moment curve is shown in Figure 1-62. Bending moments for use
in the bending stress calculations are determined by multiplying the moment
ordinate by the scale factor, 8,618.65 ft-tons/in.

S0300-A8-HBK-010

Segment Weight Mean


Load Cum. Area Shear
Ordinate Buoyancy Ordinate under Ordinate
w
Ordinate b - w
Load Area/2
in.
b
in.
Curve
in.
in.
in2
-1.4-0
0.15
0.00
-0.15
-0.02 -0.012
0-1
0.66
0.23
-0.43
0.19
0.093
1-2
1.41
0.78
-0.63
0.62
0.308
2-3
1.63
1.40
-0.23
1.25
0.623
3-4
2.97
1.98
-0.99
1.48
0.738
4-5
3.16
2.49
-0.67
2.47
1.233
5-6
3.15
2.93
-0.22
3.14
1.568
6-7
3.60
3.35
-0.25
3.36
1.678
7-8
2.90
3.75
0.85
3.61
1.803
8-9
2.05
4.07
2.02
2.76
1.378
9-10
3.80
4.28
0.48
0.74
0.368
10-11
3.72
4.34
0.62
0.26
0.128
11-12
3.45
4.28
0.83
-0.36 -0.182
12-13
1.95
4.09
2.14
-1.19 -0.597
13-14
4.17
3.76
-0.41
-3.33 -1.667
14-15
3.51
3.31
-0.20
-2.92 -1.462
15-16
2.53
2.77
0.24
-2.72 -1.362
16-17
2.36
2.19
-0.17
-2.96 -1.482
17-18
1.82
1.64
-0.18
-2.79 -1.397
18-19
2.29
1.14
-1.15
-2.61 -1.307
19-20
1.86
0.69
-1.17
-1.46 -0.732
20-20.6 0.49
0.00
-0.49
-0.29 -0.147
1

2
3
4
Shear Mean
Area
Station Ordinate Shear under
Ordinate Shear
Curve
in.
in.
in2
-1.4 -0.120
0
0.41
0
0.093
0.57
0.62
1
0.308
1.19
0.47
2
0.623
1.65
0.68
3
0.738
2.33
0.99
4
1.233
3.32
1.40
5
1.568
4.72
1.62
6
1.678
6.34
1.74
7
1.803
8.08
1.59
8
1.378
9.67
0.87
9
0.368
10.55
0.18
10
0.128
10.73
0.06
10.4 0.000
10.76
-0.09
11
-0.182
10.70
-0.39
12
-0.597
10.31
-1.13
13
-1.667
9.18
-1.56
14
-1.462
7.62
-1.41
15
-1.362
6.20
-1.42
16
-1.482
4.78
-1.44
17
-1.397
3.34
-1.35
18
-1.307
1.99
-1.02
19
-0.732
0.97
-0.44
20
-0.147
0.53
-0.07
20.6 0.000
0.46

5
6
Moment Moment
Ordinate Mom. Ord
Shear x 8618.65
Area/3
in.
ft-tons
0.00
0
0.19
0.40
0.55
0.78
1.11
1.57
2.11
2.69
3.22
3.52
3.58
3.59
3.57
3.44
3.06
2.54
2.07
1.59
1.11
0.66
0.32
0.18
0.15

1,629
3,407
4,745
6,700
9,531
13,554
18,217
23,217
27,787
30,295
30,823
30,899
30,744
29,625
26,373
21,879
17,822
13,737
9,601
5,717
2,788
1,526
1,314

5
4

5
4

BUOYANCY

3
SCALE IN INCHES

2
WEIGHT

LOAD

-1

-1
AP 19 18 17 16 15 14 13 12 11 10 9

1 FP

STATIONS
Figure 1-61. Buoyancy, Weight, and Load Curves for FFG-7.

10.4
5
4

5
4

MOMENT

3
SCALE IN INCHES

SHEAR

3
2

-1

LOAD

-2

-1
-2

-3

-3

-4

-4

-5

-5
AP 19 18 17 16 15 14 13 12 11 10 9

1 FP

STATIONS
Figure 1-62. Still Water Load, Shear, and Bending Moment Curves for FFG-7.

f. Bending Stresses
Bending stresses are calculated using the tabulated moments of inertia from
the Longitudinal Strength and Inertia Sections Drawing (FO-4):
Station Moment
M
ft-tons

INA
2

in -ft

ckeel

ft

deck
Mc/I
tons/in2

15.09
15.68
14.18
15.32
15.37
15.53
15.36
14.62
14.90
14.16
14.27
15.00
12.87
11.70
10.12
9.46

0.91
1.32
1.88
2.04
2.74
3.12
2.92
2.64
2.85
2.60
2.70
2.92
2.56
2.33
2.01
1.59

20.58
18.84
19.27
17.23
16.38
15.59
15.21
15.45
15.02
15.51
15.10
14.27
15.29
14.83
14.32
13.07

cdeck
2

ft

keel
Mc/I
tons/in2

Since the ship is hogging, the deck is in tension and the keel in
compression. All weight and buoyancy forces were given in long tons, so
the stresses are in long tons per square inch. Stresses are converted to psi
by multiplying by 2,240. Deck and keel bending stresses are plotted in
Figure 1-63 (Page 1-94). Note that the maximum bending stresses do not
occur at the section of maximum bending moment.
g. Maximum Shear Stress

3
4
5
6
7
8
9
10
10.4
11
12
13
14
15
16
17

6,700
9,531
13,554
18,217
23,217
27,787
30,295
30,823
30,899
30,744
29,625
26,373
21,879
17,822
13,737
9,601

110,681
112,994
102,384
136,770
130,123
138,267
159,477
170,416
161,280
167,165
156,553
135,444
110,066
89,467
69,084
57,188

1.25
1.59
2.55
2.29
2.92
3.13
2.89
2.79
2.88
2.85
2.86
2.78
3.04
2.95
2.85
2.19

S(Q)
12 INA b

Shear stress is a function of shear force (S), moment of inertia (I), and plating thickness (b), and is maximum at the neutral axis for any section. Maximum shear occurs at station 7. Moments of inertia for adjacent stations and
other stations of high shear are equal to or greater than that for station 7.
Side-plating thickness at the neutral axis is constant between stations 3 and
17 (information taken from the section drawings of the Longitudinal Strength
and Inertia Sections drawing - not reproduced in this handbook). Maximum
shear stress can therefore be assumed to occur at or near station 7 at the
neutral axis. The first moment of area about the neutral axis and shear
stress for station 7 are calculated in a tabular format as shown on the
following page.

1-93

S0300-A8-HBK-010

Mn Dk Girders
5 x 4 x 6#
(12)
2nd Dk
4 x 4 x 5#
Girders (11)
Mn Dk Plating,
192 x .25
Inbd
Mn Dk Plating,
84 x .375
Outbd
2nd Dk
202.5 x .1875
Plating, Inbd
2nd Dk Pltg,
52.5 x .25
Outbd
"D" Strake
105 x .375
"C" Strake
above N.A.
84 x .3125
(16.33')
"D" Doubler
30 x .75
Side Stringers
L20
6 x 4 x 7#
L19
6 x 4 x 7#
L18
5 x 4 x 6#
L17
5 x 4 x 6#
L16
6 x 4 x 7#
L15
6 x 4 x 7#
L14
6 x 4 x 8#
Totals

y
ft

ay
in2ft

15.93 15.100 240.54


10.66 6.062 64.61
48.00 15.370 737.76
31.50 15.370 484.16

STRESS, TONS/IN2

a
in2

Dimensions
in.

37.97 6.312 239.66

17

b
S

16

14

13

12

26.25 3.500 91.88


22.50 11.750 264.38
1.67
1.67
1.33
1.33
1.67
1.67
1.96
270.5

13.625
11.750
9.875
8.125
4.625
2.875
1.250

22.77
19.64
13.11
10.79
7.73
4.80
2.45
2856.2

16

1-11.7 Bending Stress in Inclined Ships.


If a ship is inclined, as shown in Figure 164, the depth of sections is increased and
bending stresses at the "corners" may be
increased. For a ship heeled to an angle ,
the new axis of bending is parallel to the
water line. The bending moment, M, can
be resolved into Mcos about the old
(horizontal) neutral axis and Msin about
the centerline of the ship. Each component
produces stress as if it acted independently,
and the total stress at some point P, with
coordinates (x,y), is:
My cos
INA

9
8
7
6
5
4
3
2
1

11
10
9
STATIONS
MAIN DECK

=
=

INA =
ICL =

1-94

Mx sin
ICL

15

14

13

12

9
8
7
6
5
4
3
2
1

11
10
9
STATIONS
KEEL

total bending stress at (x,y)


distance from the old
neutral axis to the point in
question
distance from the centerline to the point in
question
moment of inertia about the old neutral axis
moment of inertia about the centerline

Figure 1-63. Still Water Bending Stresses for FFG-7.

Y
X

N
NEU
T
AXI RAL
S
HORIZONTAL

D
OL RAL
T
U
NE AXIS

where:
t
y

10.4

9
8
7
6
5
4
3
2
1

ay = 2,856.2 in2-ft
2(2,856.2) = 5,712.4 in2-ft
=
130,123 (from Longitudinal Strength
and Inertia Sections drawing)
=
2 plate thickness @ NA = 0.625 in.
=
1.803 x 140.83 = 255.61 tons
=
S(Q)/12INAb = 1.5 tons/in2
=
1.5 2,240 = 3,360 psi

t =

15

39.38 10.995 432.93

17
Qhalf-section =
Qwhole section =
INA

10.4

9
8
7
6
5
4
3
2
1

13.13 6.312 82.85

STRESS, TONS/IN2

Component

CL

Figure 1-64. Stresses in Inclined Ships.

S0300-A8-HBK-010

Since the section is not symmetrical about its new bending axis, the neutral axis is not parallel to the waterline (horizontal) but is inclined to
it by some angle , as shown in Figure 1-64. The angle, , between the new neutral axis and the horizontal can be found from:
INA
ICL

tan = tan

If the point farthest from the neutral axis has coordinates (x1, y1) referenced to the centerline and the old neutral (horizontal) axis, maximum
bending stress in the section is:
max =

My1 cos

Mx1 sin

INA

ICL

Tabulated section moments of inertia about the centerline are not normally available to the salvage engineer, and must be calculated. Calculating
ICL is somewhat simpler and shorter than calculating INA because the incremental second moments are taken about a known axis (the centerline).
There is therefore no need to sum first moments about an arbitrary axis to locate the neutral axis. For intact sections, only the incremental
second moments of area for one side need be summed; the moment of inertia is twice the sum for one side. Distances from the centerline are
scaled from section drawings.
Maximum bending stresses in an inclined ship may be 20 percent greater than when the ship is upright.
1-11.8 Combined Stresses. The bending (tensile or compressive) and shear stresses in a ship or other beam combine to form the principal
stress at any point. It can be shown that:
s (s ) = 2
where:
s

=
=
=

principal stress at any point


simple tensile or compressive stress at the point in question
shear stress at the point in question

This relationship does not solve for s so iterative or trial and error methods are used to determine principal stress.
The presence of shear in the hull girder distorts the sections so that the conditions on which simple beam theory are based are not strictly
fulfilled (see Chapter 2 for an explanation of basic beam theory). This alters bending stress distribution across the section from that predicted
by beam theory. Analysis of this problem is beyond the scope of this book, but the general effect is to increase bending stress at the corners
of the section, i.e., the deck edges and the bilge, and reduce bending stresses at the center of the deck and bottom. This effect is appreciable
only when the ratio of length to depth is small.
1-11.9 Acceptable Stress Levels. The stress that any material can withstand without failure is a function of the properties of that material and
the definition of failure. Fracture is an obvious and final form of failure. Permanent or plastic deformation, or unacceptable extents of deflection
or elastic deformation can also be considered failure.
1-11.9.1 Failure Definition. In many engineered systems, deflection or deformation of a component in excess of certain limits interferes with
the operation of the mechanism and is considered failure. Plastic deformation is often considered failure because of the discontinuous behavior
of the material as it yields. Plastic behavior may be acceptable in components subjected to in-line, tensile loading where elongation will not
cause interference with any other components. The deformation may render the component unsuitable for continued use, but many salvage
evolutions are one-time events. Plastic behavior or excessive deflection/deformation should be carefully examined, as such deformation in
components can alter stress levels in other components in unforeseen or unpredictable ways. Plastic failure in ship hulls is unacceptable because
it unpredictably alters load responses.
Failure of a given component must be defined accurately, so that limiting stress values for that component can be set. The limiting stress values
define limiting loads for components; the degree of load sharing among components will define system load limits.
1-11.9.2 Factors of Safety. Use of an appropriate factor of safety keeps stresses well below the failure point and allows for manufacturing
defects and inconsistencies in loading. Safety factors are specified by various regulatory agencies, depending on intended use of systems and
components. In salvage it is not always possible to use a standard safety factor, so reduced factors of safety must often be accepted. This does
not mean that salvors can disregard safety factors. Each situation must be examined to determine acceptable stresses and loads. A reduced
safety factor represents an increased chance of failure. The consequences of failure must be considered and precautions taken.

1-95

S0300-A8-HBK-010

1-11.9.3 Common Materials. The most commonly used shipbuilding materials are:

Steel and Iron.

Aluminum.

Wood.

Glass Reinforced Plastic (GRP).

Copper and Copper Alloys.

In addition to encountering these as components of a ship, the salvor may use any of them in on-site repairs or fabrication of salvage systems,
along with concrete or other materials. The ultimate or yield stresses of many materials vary depending on whether tensile, compressive or shear
stress is experienced. This is an important factor in salvage operations, where components may be loaded in ways other than those anticipated
by the designer. The mechanical properties of commonly used materials are given in Appendix E.
Steel, in the form of rolled plate, rolled or forged structural shapes, or complex castings, is the most commonly used shipbuilding material.
Shipbuilding steel meeting ABS and Navy specifications has a yield stress of not less than 32,000 psi and an ultimate stress of 58,000 -70,000
psi. In the United States, structural shapes and plates for general use are usually manufactured to American Society for Testing of Materials
(ASTM) Standard A36, requiring a tensile yield strength of not less than 36,000 psi. Unless otherwise specified, mild steel can be assumed to
have a yield strength of about 30,000 psi, although some alloys have yield strengths as low as 20,000 psi.
Plating thickness is often specified by weight per square foot. Steel weighs approximately 490 pounds per cubic foot, so a 40.8-pound plate
is approximately 1-inch thick. Iron weighs 480 pounds per cubic foot, so 1-inch iron plate weighs exactly 40 pounds per square foot. In
common usage, the decimal fraction is often dropped when naming steel plate; 1-inch steel plate is called 40-pound plate, quarter-inch steel plate
is called 10-pound plate, etc. This practice can sometimes lead to confusionsteel plate and shapes are sometimes fabricated to dimensions
specified by weight per area or linear dimension. The thickness of plate so manufactured will be slightly less than assumed by dividing the
weight by 40. Table E-15 correlates steel-plate thickness to weight per square foot.
Major load-bearing members, such as sheer and garboard strakes, main deck stringers and bottom girders, etc., and submarine pressure hulls
are frequently fabricated of high-stength steels. High-strength steels are designated by an "HY" (high yield), "HSLA" (high-strength, low-alloy)
or number, i.e., HY80, HSLA80, HY100, HY140, etc.; the number specifying the nominal yield stress in thousands of pounds per square inch.
High-strength steels are difficult to weld and cut. Intermediate-strength steels, with yield stresses in the 35,000 - 45,000 psi range, are often
used for the major strength members of larger merchant hulls to provide the required strength with lighter scantlings. These steels have been
called high-tensile (HTS) or higher strength steels by classification societies to avoid confusion with truly high-strength steels.
Corrosion-resistant steels (CRES), sometimes called stainless steels, are used extensively where corrosion or appearance are important factors.
Strength and other properties vary widely, depending on composition. Because of their resistance to oxidation, corrosion-resistant steels are
considered nonferrous metals, and are difficult to cut with oxygen-fuel or oxygen-arc cutting equipment. Low magnetic signature alloys are
sometimes used on mine countermeasures ships.
Cast iron is used occasionally for complex shapes not subject to tensile loads. Wrought iron is more malleable and corrosion-resistant than mild
steel, and nearly as strong. Wrought iron is no longer produced in the United States, but was formerly used in place of steel in ship
construction, and may be encountered in older ships. Wrought iron stud-link chain is found occasionally.
Aluminum is used extensively in small ships, boats, and landing craft. The yield stress of pure aluminum is about 5,000 psi, but some alloys
have yield stresses as high as 78,000 psi. Aluminum alloys used in shipbuilding have yield stresses in the range of 12,000 - 20,000 psi. Because
of aluminums low density, aluminum alloy members are lighter, but bulkier, than steel members of the same strength; aluminum is often used
in superstructures to reduce topside weight.
Wood is used in the construction of mine countermeasures ships and small craft. The hardness and density of wood vary with species and water
content. Green wood contains varying amounts of water as sap; wood absorbs water in humid climates or when immersed. The strength
characteristics of wood vary with species and type of stress; all species are much stronger against normal stresses than against shear; most are
stronger in tension than in compression.
Glass Reinforced Plastic is used in the hulls of small craft and some mine countermeasures ships, in piping systems, as sheathing over wooden
hulls and in joiner bulkheads. It is also frequently used as a patching material for other materials. Strength varies depending on the orientation
of the glass fibers and plastic resins used.
Copper and its alloys, such as brass, bronze, monel, and copper-nickels, are used in piping systems, propellers, and fittings where corrosion
resistance or low magnetic signature are required. Although certain copper alloys are very strong, they are seldom used as structural members
or fittings, except on mine countermeasures ships, because of their high cost.

1-96

S0300-A8-HBK-010

1-11.10 Hull Girder Deflection. Hull


girder deflection is a function of the fourth
integral of the load curve with respect to
ship length, and girder stiffness, indicated
by the product of moment of inertia (I) and
modulus of elasticity (E). Deflection is
determined by double integration of the
curve of bending moment divided by EI.
Since I, and sometimes E vary along a
ships length, M/EI is calculated at several
stations to construct an M/EI curve. The
curve is integrated from left to right to
determine the ordinates to the first integral
curve, which is again integrated from left to
right to determine ordinates to the second
integral curve. A straight line is drawn
between the ends of the second integral
curve, as shown in Figure 1-65. The
vertical separation between the straight line
and the second integral curve at any station
is the deflection at that station. As shown
in Figure 1-65, the straight line in the
deflection plot corresponds to a straight line
connecting forward and after drafts in a
floating ship, i.e., deflection is assumed
zero at the fore and after perpendiculars.

SECOND
INTEGRAL

DEFLECTION AT B
FIRST
INTEGRAL

AP

xB

FP

__
M
EI

DEFLECTION AT B

Figure 1-65. Hull Girder Deflection Determination.

Table 1-15. Hull Deflection.

The hull deflection of a stranded or


damaged casualty is readily observable; a
salvage engineer does not usually calculate
hull deflection unless unusually extreme
loadings are contemplated and the degree of
hull deflection may affect salvage work or
conditions. Observed deflection is a rough
indicator of hull stress; a first estimate of
stress can be obtained by comparing a
casualtys deflection with the stress
corresponding to similar deflections in ships
of similar form and size. Table 1-15 gives
stresses and deflections calculated for four
different ships in various conditions.
1-11.11
Approximate Strength Calculations. Lack of detailed ship data or
time for rigorous calculations may
necessitate the approximation of all or part
of the strength calculations. The following
paragraphs describe methods to estimate
weight distribution, section properties, and
still water or wave bending moment.
1-11.12 Weight Curve Approximations.
There are a number of empirically derived
approximations for weight distribution,
none of which is equally applicable to all
ship types. The station coefficient method,
presented below, is probably the most
accurate, but is applicable to only three
ship types at present. Less accurate, but
more generally applicable methods are
presented in the following paragraphs.

Ship Type:

VLCC

CONTAINER SHIP
(1400 TEU)

0.56

0.87

0.49

650

1050

673

47

97.5

175.9

105.7

30

50

90.5

66.5

FFG-7

T-AO 187

CB

046

LBP, ft

408

Beam, ft
Depth, ft

Characteristics:

Deflection conditions:

Stresses and Deflections:

Full Load
Maximum stress, ksi

5.6

-3.4

15.6

-4.3

Maximum deflection, in.

2.4

2.5

6.2

-1.1

Maximum stress, ksi

5.5

-11.1

13.0

10.3

Maximum deflection, in.

2.3

5.9

5.3

8.1

Maximum stress, ksi

17.8

-18.7

29.3

23.4

Maximum deflection, in.

7.2

10.6

17.7

11.2

Ballast

Full load w/hogging wave

Stranded on one pinnacle (hogging) deflection at 34 ksi

10.6

14.0

17.7

11.0

Stranded on two pinnacles (sagging) deflection at 34 ksi

-20.3

-27.2

-42.3

-20.4

Note: Positive stresses indicate tension, negative stresses compression


From Hull Deflection Versus Bending Moment Study for Supervisor of Salvage, U.S. Navy, Herbert Engineering
Corporation, 5 March 1991

1-97

S0300-A8-HBK-010

Breakbulk (general) cargo


ship with engine room
and accommodations
three-quarters aft of the
forward perpendicular.
Container ship with
forward and aft accommodations.
Tanker with engine room
and accommodations aft.

The length between perpendiculars (LBP) is


divided into 20 basic segments. The breakbulk cargo ship has a segment forward of the
forward perpendicular and the tanker and
container ship each have segments aft of the
aft perpendicular. Station coefficients (CSN)
from Table 1-16 for the appropriate ship type
are used to determine the weight ordinate
(OSN) for each half segment:
OSN = CSN W1s

Table 1-16. Station Coefficients, CSN.


CSN

Station
A

CSN

Station
C

AFT

0.022542
0.022542
0.022542
0.022542
0.022542
0.022542
0.022542
0.022542
0.022542
0.022542
0.022542
0.022542
0.022542
0.021834
0.021352
0.022349
0.021834
0.021352
0.020387
0.017493
0.016496

0.024942
0.024942
0.024942
0.024942
0.024942
0.024942
0.024942
0.024942
0.024942
0.024942
0.024942
0.023199
0.022668
0.021606
0.021114
0.020052
0.019522
0.018991
0.01793
0.017399
0.016338

AFT
20.5-21
20-20.615
20-20.5
19.5-20
19-19.5
18.5-19
18-18.5
17.5-18
17-17.5
16.5-17
16-16.5
15.5-16
15-15.5
14.5-15
14-14.5
13.5-14
13-13.5
12.5-13
12-12.5
11.5-12
11-11.5
10.5-11

0.006303

10-10.5
0.023068
9.5-10
0.023068
0.012377
0.010676
9-9.5
0.023068
0.014333
0.015049
0.01793
8.5-9
0.023068
0.022157
0.017975
0.021114
8-8.5
0.023068
0.020875
0.020387
0.034267
7.5-8
0.023068
0.020875
0.022831
0.038513
7-7.5
0.023068
0.020875
0.02476
0.039536
6.5-7
0.023068
0.021516
0.028169
0.034267
6-6.5
0.022157
0.022157
0.029616
0.025321
5.5-6
0.021516
0.023472
0.025243
0.025321
5-5.5
0.020875
0.032612
0.025243
0.025321
4.5-5
0.020201
0.033252
0.038845
0.024942
4-4.5
0.019560
0.041076
0.038845
0.024942
3.5-4
0.018919
0.053453
0.040774
0.024942
3-3.5
0.018245
0.055409
0.042736
0.024942
2.5-3
0.017604
0.052172
0.043701
0.024942
2-2.5
0.016963
0.028025
0.022542
0.024942
1.5-2
0.016289
0.023068
0.022542
0.024942
1-1.5
0.015142
0.023068
0.022542
0.024942
0.5-1
0.014333
0.023068
0.022542
0.024942
0-0.5
0.013692
0.023068
0.022542
0.024942
-0.55-0
0.013051
FWD
FWD
Ship A Breakbulk cargo ship - engine room and accommodations three-quarters aft from
Ship B Container ship with forward and aft accommodations
Ship C Tanker with engineroom aft
0.015807

FP

Osn

1-11.12.1 Station Coefficient Method.


This method was developed as part of the
Pouricelli-Boyd-Schleiffer regression
analysis discussed in Paragraph 1-7 and
provides a means to approximate lightship
weight distribution of three types of
merchant hulls:

where:
Wls = lightship weight

The hull steel weight is


calculated or estimated by
deducting weights of
machinery, propellers, and
superstructure from the
lightship weight, or by the
methods described below.
The bare hull weight distribution is estimated by
one of the methods described in the following
paragraphs.
The deducted items are
added at their locations to
complete the lightship
weight curve.

FP

Osn

1-11.12.2 Bare Hull Estimates. For ship


types other than the three mentioned above,
the lightship weight curve is approximated
in three steps:

AP

BREAK-BULK CARGO SHIP-ENGINE ROOM AND ACCOMMODATIONS


THREE-QUARTERS AFT FROM FP

AP

FP

CONTAINERSHIP WITH FORWARD AND AFTER ACCOMMODATIONS

Osn

The weight ordinates are plotted as shown


in Figure 1-66 to develop the lightship
weight curve. Variable weights (cargo,
flooding, etc.) are added as rectangles or
trapezoids at the appropriate station for the
ships actual load condition.

AP

FP

TANKER WITH AFT ENGINEROOM


Figure 1-66. Station Coefficient Weight Curves.

1-98

S0300-A8-HBK-010

After the distribution of the hull weight of


the ship has been estimated, the variable
weights of fuel, stores, cargo, boats,
aircraft, ballast, ammunition, crew and
effects, etc., are added by superimposing
rectangles or trapezoids on the curve at
their locations.

280
260
240
220

Hull steel weight for commercial vessels


can be estimated by the two relationships
shown below:

WH L (B

or,

2D) k2

where:
WH =
L =
B
D
k1

=
=
=
=
=

hull weight, ltons


length between perpendiculars, feet
molded beam, feet
molded depth, feet
weight coefficient
0.0027 for welded construction
0.0030 for riveted construction

POWER DENSITY LB/SHP

W H L B D k 1,

200
DIRECT DRIVE DIESEL
NUCLEAR
STEAM TURBINE
REHEAT STEAM TURBINE
GEARED DIESEL
COMBINED DIESEL AND
GAS TURBINE
GAS TURBINE
REGENERATIVE

180
160
140
120
100
80
60
40

k2

=
=
=

weight coefficient
0.0433 for welded construction
0.0558 for riveted construction

20
4

10

12

14

18

22

26

30

34

38

42

SHP RATING OF PROPULSION PLANT (THOUSANDS)


COMPILED FROM VARIOUS SOURCES, INCLUDING SHIPS AND SHIPBUILDING OF TOMORROW,
SCHNKNECKT, LSCH, SCHELZEL & OBENHAUS, 1983; SHIP DESIGN AND CONSTUCTION,
TAGGART, 1980; MARINE ENGINEERING, HARRINGTON, 1955 AND MANUFACTURERS DATA

Weights of machinery and outfits can


sometimes be obtained from the ships
information book (SIB), operating and
Figure 1-67. Machinery Weight.
technical manuals, or manufacturers data.
Machinery weight for commercial vessels
can be estimated very approximately by use of the "power density" factors taken from Figure 1-67.

There is no standard definition of what is included in the term machinery weight, so figures given in ships data must be investigated to
determine what items are included. Values taken from the curves in Figure 1-67 include the weight of main propulsion units, shafting, bearings,
propellers, boilers, stacks, condensers, generators, switchboards, and pumps; all piping, floors, ladders and gratings in the machinery spaces;
water in boilers, engines, and piping; and refrigerating and steam heating systems for a normal vessel. Weights of steering gear, deck machinery,
and piping outside the machinery spaces are not included. Machinery weights are subject to variation, depending on the ship type and service.
In ship types that require particularly rugged or reliable machinery, machinery weight will be about 10 percent higher than the values from Figure
1-67. Different makes of diesel engine of the same horsepower will
vary in weight by as much as 50 percent. Total machinery weight in
Table 1-17. Machinery Weights for Combatants.
specialized vessels will include items not fitted on ordinary ships, or
larger numbers of common items. Examples are the refrigeration plant
on a refrigerated cargo ship, additional pumps and generators on
salvage and service vessels, dredge machinery, etc.
BB, CV
50-60 pounds/SHP
Because of their high speed and correspondingly powerful machinery,
the weight of machinery of naval combatants is a large portion of the
total weight of the ship. Emphasis on machinery weight savings during
design results in lower weight per horsepower than in the average
commercial vessel. Machinery dry weight for different types of
combatants can be taken from Table 1-17.

CG, CL, CA
DD, FF
DD, FF, CG (gas turbine)

35-40 pounds/SHP
27-30 pounds/SHP
20-25 pounds/SHP

1-99

S0300-A8-HBK-010

Table 1-18 gives weights of bronze propellers as a function of shaft horsepower and rpm. Table 1-19 gives summarized weight lists for different
types of ships to illustrate general trends in weight distribution. Additional weight summaries are included in Appendix B.
Table 1-18. Weights of Bronze Propellers (lbs).

SHP

100

120

140

500
1,000
2,000
3,000
5,000
10,000
15,000
20,000

3,415
6,545
12,080
17,410
24,905
55,100
62,155

2,775
5,270
9,830
14,105
20,495
35,705
50,030

2,315
4,475
8,265
11,680
17,190
29,315
40,335
50,910

Shaft RPM
160
2,030
3,880
7,140
10,360
14,545
24,245
32,400
40,115

180

200

250

300

350

1,785
3,460
6,350

1,585
3,150
5,730

1,255
2,445
4,630

970
1,915
3,670

750
1,520
2,975

From Ships and Marine Engineers, Volume IV, The Design of Merchant Ships, Schokker, Newerburg, Bossnack, and Burghgracf, The Technical Publishing
Company H. Stam, 1953

Table 1-19. Lightship Weight Summaries.


Ship Type

Steel
Outfit
Machinery10
Fixed Ballast

5,115
2,586
1,039
---

5,011
2,230
867
---

Combination
Passenger/
Reefer
Container
Ship2
5,482
3,959
982
---

Lightship

8,746

8,108

10,4235

Item

Mariner
With Added General
Features, Cargo Ship1
1962

Container
Ship3

BargeBargecarrying
carrying
ship
Ship
(LASH)4 (SEABEE)5

Tanker6

Ore Carrier7

Small
Freighter8

Passenger Container
Ship
Vessel9

10,282
2,525
1,911
---

9,588
2,937
1,105
---

12,983
2,979
1,421
---

11,519
1,844
831
---

12,137
1,600
980
---

2,248
574
398
---

11,850
6,875
2,525
---

4,557
1,739
837
3,329

14,718

13,630

17,383

14,194

14,717

3,220

21,250

10,452

Weights in Long Tons

From Ship Design and Construction, Amelio M. DArcangelo; Society of Naval Architects and Marine Engineers, 1969 and Princples of Naval Architecture,
Society of Naval Architects and Marine Engineers, Second Edition, 1967 and Third Edition, 1988
Notes:
1
573 LOA, machinery and house 3 4 aft, 6 holds, 2 tween decks, 24,000 SHP, 23 Kts.
2
574 LOA, machinery and house midships, 19,800 SHP, 20 Kts.
3
752 LOA, machinery 3 4 aft, house forward, 1,920 TEU, 60,000 SHP, twin screw, 27 Kts.
4
820 LOA, machinery 3 4 aft, house forward, 79 LASH barges, 32,000 SHP, 27.5 Kts.
5
824 LOA, machinery 3 4 aft, house forward, 38 SEABEE barges, 36,000 SHP, 20 Kts.
6
810 LOA, machinery and house aft, single bottom, 5 center and 8 wing tanks, 19,000 SHP, 17 Kts.
7
765 LOA, machinery and house aft, 7 holds, 19,000 SHP, 16.5 Kts.
8
390 LPB, two deck, three-island design, 3,150 SHP, 13 Kts.
9
661 LPB, ten deck, 1,200 passenger, 650 crew, 30,000 SHP, 20 Kts.
10
Steam turbine plants in all cases, single screw unless otherwise noted.

Table 1-20. Prohashas Ordinates for the Coffin Diagram.

Type of Ship
Tanker
Full-bodied cargo ships w/o erections
Fine-lined cargo ships w/o erections
Full-bodied cargo ships with erections

Prohaskas ordinates
a&c
b
0.75WH / L
1.125WH / L
0.65WH / L
1.175WH / L
0.60WH / L
1.20WH / L
0.55WH / L
1.225WH / L

where: WH = Hull weight, ltons (less propelling machinery)


L = Length overall, ft

Reproduced from Applied Naval Architecture, R. Munro, 1967

1-100

Type of Ship
Fine-lined cargo ships with erections
Small passenger ships
Large passenger ships

Prohaskas ordinates
a&c
b
0.45WH / L
1.275WH / L
0.40WH / L
1.30WH / L
0.30WH / L
1.35WH / L

S0300-A8-HBK-010

1-11.12.3 Coffin Diagram. Bare hull


weight distribution for ships with parallel
midbody can be approximated by a line
diagram, commonly called a coffin
diagram, consisting of a rectangle over the
length of the midbody and trapezoids at the
bow and stern.
Three hull weight
distribution methods are based on the coffin
diagram. The Biles and Prohaska methods
each divide the length overall into three
equal segments as shown in Figure 1-68. A
third method, that may be termed the
general parallel midbody method, divides
the length into three segments based on the
observed length of the parallel midbody.

L/3

L/3

L/3

AP

FP
BILES METHOD ORDINATES
W
a = 0.566 __H
L
W
b = 1.195 __H
L

Biles method ordinates for ordinary cargo


and passenger vessels are shown in the
Figure 1-68, Prohaska method ordinates for
different ship types are given in Table 1-20.

W
c = 0.653 __H
L
WHERE WH = HULL WEIGHT (LESS MACHINERY)

The centroid of the Biles diagram is


0.0056L abaft midships. Small adjustments
can be made to the end ordinates so that
the centroid of the diagram corresponds to
the longitudinal position of the center of
gravity of the hull. LCG of the bare hull is
not at the same location as the light ship
LCG. The position of the centroid of the
coffin diagram must be chosen so that LCG
will shift to a known or estimated position
as weights are added, corresponding to the
condition where LCG is known.

FROM APPLIED NAVAL ARCHITECTURE, R. MUNRO-SMITH, 1967.

Figure 1-68. Coffin Diagram.

__
7 L
9

By shortening one end ordinate and


lengthening the other by an equal amount,
a triangle is transferred from one trapezoid
to the other, as shown by the dotted lines in
Figure 1-69. The centroid of each triangle
lies one-third of its length from its base:
1 L
=
3 3

x
G

L/3

L
9

G1

L/3

L/3

AP

FP
54(WH )GG1
x = ____________
7L2

where L is the length of the diagram, corresponding to length overall (LOA) of the ship.
The shift of the centroid of the total area is
therefore (7/9)L. If the base of the triangle
is taken as x, and its height as L/3, then,

Figure 1-69. Adjusting LCG of the Coffin Diagram.

1 L
Area of triangle = x =
2 3
xL 7L
Moment of the shift = =
6 9

xL
6
7xL 2
54

The shift of the centroid of the diagram, representing the LCG of the hull is thus:
L2
7
Shift of LCG = ( x )
WH
54
where WH is the bare hull weight. The triangle base, x, required to give the desired shift of LCG is:
x =

54 (WH) (desired shift of LCG)


7L 2

1-101

S0300-A8-HBK-010

In the general parallel midbody method, the


beginning and end points and length of the
parallel midbody are determined by
inspection. The middle ordinate (b) is
defined as shown in Figure 1-70. The end
ordinates are chosen so that the centroid of
the entire diagram corresponds to the bare
hull LCG. Figure 1-71 shows how to select
end ordinates for a trapezoid to place the
center of the trapezoid in a desired location.

A line parallel to the base is drawn through


the centroid of the area under the parabolic
curve. A second line is drawn from the
base of the parabola at its midlength to
intersect the first line at a distance from the
midships ordinate equal to twice the desired
shift in LCG. This line is extended beyond
the contour of the parabola.
The
intersection of this line with a horizontal
line drawn from the center of the parabolic
curve defines one point on the new curve.
Parallel lines drawn at other ordinates
define other points on the new curve, as
shown in Figure 1-72.
For ships without parallel midbody, a bare
hull weight curve can also be generated by
assuming that two-thirds of the hull weight
follows the still water buoyancy curve and
distributing the remaining one-third in the
form of a trapezoid so arranged that the
center of gravity lies above the center of
buoyancy, as shown in Figure 1-73. This
method has been found to yield close
approximations to the hull weight
distribution for large warships.

b
l
1.3

W
__H
L
W = Hull Weight
b = b1 x

l = LENGTH OF TANK SECTIONS


OR PARALLEL MIDDLE BODY
b1

1-11.12.4
Ships Without Parallel
Midbody. An approximate weight curve
for ships without parallel midbody can be
constructed as a parabola over a rectangle,
with the area under each representing half
the bare hull weight (Cole, reproduced in
Applied Naval Architecture, R. MunroSmith, 1967).
The ordinate for the
rectangle is WH/2L; the maximum
(midships) ordinate for the parabola is
3WH/4L, as shown in Figure 1-72. LCG of
this figure is amidships. Correction for
LCG lying forward or aft of midships is
made by swinging the parabola.

1.4

1.2

L = LENGTH OF SHIP OVERALL

1.1

1.0
0.3

0.2

0.4

0.5

0.6

0.7

l
__
L
FROM PRINCIPLES OF NAVAL ARCHITECTURE, SNAME, 2ND EDITION, 1967.

Figure 1-70. General Parallel Midbody Weight Curve.

l = LENGTH OVER WHICH THE WEIGHT IS DISTRIBUTED


x = LONGITUDINAL DISTANCE FROM THE SMALLER
END OF THE TRAPEZOID TO ITS CENTER OF GRAVITY
a = AREA OF THE TRAPEZOID
= TONS FOR WEIGHT AND LOAD CURVES
b1, b2 = END ORDINATES

b1

b2

1-11.13 Wave Bending Moment with


Nonstandard Waves.
The salvage
engineer must often assess the ability of a
damaged casualty to withstand wave
l
bending loads, either during the salvage
operation or during transit to a repair
__ ( 3x
__ -1)
__ ); b = 2a
__ (2 - 3x
b1 = 2a
2
l
l
l
l
facility. Because of the tedious nature of
the calculations, the usual first task is to
determine the stresses imposed by a
Figure 1-71. Centroid of a Trapezoid.
standard L/20 or 1.1
L wave with length
equal to ships length. If the ship can carry
loads imposed by a standard wave, no further calculations need be performed in most cases. If, however, the stresses imposed by the standard
wave are excessive, calculations must be performed for trial wave heights and lengths until the maximum acceptable wave is determined, unless
bending moment caused by waves with differing length and height can be correlated to those caused by the standard wave.

1-102

S0300-A8-HBK-010

A 1991 analysis by Herbert Engineering


Corporation of five hull forms with block
coefficients ranging from 0.46 to 1.0
developed factors that relate nonstandard
wave bending moments to normalized
standard bending moment. The factors are
functions of block coefficient, wavelength,
and wave height.

ORIGINAL
CURVE
CORRECTED
CURVE

WBM =
35

B
h

=
=

X = DESIRED SHIFT
OF LCG
FROM APPLIED NAVAL ARCHITECTURE, R. MUNRO-SMITH, 1967.

Figure 1-72. Parabolic Weight Curve.

PARALLEL TO
STILL WATER
BUOYANCY CURVE
2/ W
3 H

1/ W
3 H

Figure 1-76 (Page 1-104) can be entered


with block coefficient to get an estimate of
the standard bending moment (waveheight
= 1.1
L, wavelength = L).

AP

FP

FROM BASIC SHIP THEORY, RAWSON AND TUPPER 3RD EDITION, 1983.

Figure 1-73. Alternate Weight Distribution for Ships Without Parallel Midbody.

0.015
0.01
HOG

normalized wave
bending moment,
dimensionless
wave bending moment,
ft-lton
standard seawater
specific gravity, ft3/lton
length between
perpendiculars, ft
beam, ft
wave height, ft = 1.1
L

WH
2L

0.005
0
-0.005

SAG

NBM =

3WH
4L

a=

BASE LINE

NORMALIZED MOMENT VS. WAVELENGTH (Cb = 0.46)

where:

b=
2
b
5

Figure 1-76 (Page 1-104) shows normalized


maximum and standard hogging and
sagging moments as a function of block
coefficient. All curves are based on 1.1
L
trochoidal waves. The normalized bending
moment is given by:
WBM 35
L 2Bh

WH
L

2X

The analysis revealed that for fine-lined


ships, maximum wave bending moment
occurs at wavelengths slightly less than the
ships length (approximately 0.75L), and
may be as much as 15 percent higher than
bending moment for the standard wave.
Figure 1-74 shows the relationship between
wavelength and bending moment for an
FFG-7 Class ship (CB = 0.46) for a 1.1
L
wave height. Figure 1-75 (Page 1-104)
shows the relationship between standard
wave bending moment and maximum wave
bending moment as a function of block
coefficient.

NBM =

a + b = 1.25

-0.01
-0.015
-0.02
0.9

0.8

0.7

0.6
0.5
0.4
LOCATION FROM FP (X/LBP)

0.3

0.2

0.1

L = 1.0 LBP

L = .50 LBP

L = 1.0 LBP

L = .50 LBP

L = .75 LBP

L = .25 LBP

L = .75 LBP

L = .25 LBP

L = LENGTH OF TROCHOIDAL WAVE


FROM WAVEHEIGHT AND WAVELENGTH VERSUS BENDING MOMENT STUDY FOR SUPERVISOR
OF SALVAGE U.S. NAVY, HERBERT ENGINEERING CORP., 20 FEBRUARY 1991

Figure 1-74. FFG-7 Bending Moment with Varying Wavelength.

1-103

S0300-A8-HBK-010

1-11.14 Murrays Method for Approximating Maximum Bending Moment. An


approximation of determining maximum
bending moment has been developed by J.
M. Murray, former Chief Ship Surveyor to
Lloyds Register of Shipping. Murrays
method computes still water bending
moment by taking moments of weight and
buoyancy about midships. Wave bending
moment is calculated by use of empirical
coefficients. The sum of the two gives
total bending moment at midships, which
can be taken as the maximum bending
moment in most cases. The method is
reasonably accurate for ships floating at a
trim of less than one percent of their
length.
1-11.14.1 Still Water Bending Moment.
Still water bending moment (SWBM) is
given by:
SWBM = MW
where:
MW

1.14

MOMENT RATIO

1.12
1.1
1.08
1.06
1.04
1.02
1
0.4

0.5
HOG

0.6

0.7

0.8

BLOCK COEFFICIENT

0.9

1
SAG

FROM WAVEHEIGHT AND WAVELENGTH VERSUS BENDING MOMENT STUDY FOR SUPERVISOR
OF SALVAGE U.S. NAVY, HERBERT ENGINEERING CORP., 20 FEBRUARY 1991

Figure 1-75. Ratio of Maximum to Standard Wave Bending


Moment as a Function of Block Coefficient.

0.025
0.02
0.015
0.01
0.005
0
-0.005
-0.01
-0.015

mean moment of
weight
-0.02
Mwf + Mwa
-0.025
= __________
0.8
0.9
1
0.4
0.5
0.6
0.7
2
BLOCK COEFFICIENT
MAX. HOG
MAX. SAG
Mwf
= moment of weight
STD. HOG
STD. SAG
forward of midships,
ft-lton or m-tonne
FROM WAVEHEIGHT AND WAVELENGTH VERSUS BENDING MOMENT STUDY FOR SUPERVISOR
OF SALVAGE U.S. NAVY, HERBERT ENGINEERING CORP., 20 FEBRUARY 1991
= Wf(LCGf)
Mwa = moment of weight aft of
midships, ft-lton or mFigure 1-76. Normalized Wave Bending Moment as a Function of Block Coefficient.
tonne
= Wf, a(LCGfa)
Wf, a = weight of the forebody or afterbody, lton or m-tonne
LCGf, a = LCG of the forebody or afterbody, measured from midships, ft or m
Mbf + Mba
MB
= mean moment of buoyancy = _________
2
Mbf
= moment of buoyancy forward of midships, ft-lton or m-tonne
= B f (LCBf )
Mba
= moment of buoyancy aft of midships, ft-lton or m-tonne
= Ba(LCBa)
Bf,a
= buoyancy of the forebody or afterbody, lton or m-tonne
LCBf, a = LCB of the forebody or afterbody, measured from midships, ft or m

1-104

MB

1.16

NORMALIZED BENDING MOMENT

The plots in Figures 1-77 and 1-78 are


entered with wavelength expressed as a
function of ship length to determine the
ratio between wave bending moment for the
wavelength and the standard wave bending
moment. The ratio is then applied to wave
bending moment determined from Figure 176 or by rigorous calculation to estimate
wave bending moment for the nonstandard
wavelength. Figure 1-79 (Page 1-106)
gives normalized bending moments for
wavelengths equal to L with nonstandard
waveheight.

WAVE MOMENT / STANDARD WAVE MOMENT

S0300-A8-HBK-010

1.25
1
0.75
0.5

HOG

0.25
0
0.25
SAG
0.5
0.75
1
1.25
1.5
2

1.5

0.5

0.5

1.5

WAVELENGTH / LBP
FROM WAVEHEIGHT AND WAVELENGTH VERSUS BENDING MOMENT STUDY FOR SUPERVISOR
OF SALVAGE U.S. NAVY, HERBERT ENGINEERING CORP., 20 FEBRUARY 1991

WAVE MOMENT / STANDARD WAVE MOMENT

Figure 1-77. Ratio of Wave Bending Moment to Standard Bending Moment, CB = 0.46.

1
0.8
0.6
0.4
HOG
0.2
0
0.2
SAG
0.4
0.6
0.8
1
2

1.5

0.5

0.5

1.5

WAVELENGTH / LBP
FROM WAVEHEIGHT AND WAVELENGTH VERSUS BENDING MOMENT STUDY FOR SUPERVISOR
OF SALVAGE U.S. NAVY, HERBERT ENGINEERING CORP., 20 FEBRUARY 1991

Figure 1-78. Ratio of Wave Bending Moment to Standard Bending Moment, CB = 1.0.

1-105

S0300-A8-HBK-010

Since total weight and buoyancy moments are mean moments, they are numerically equal to the product of the mean weight or buoyancy and
the mean lever arm:
MW =

MB =

Mwf + Mwa
2
Mbf + Mba
2

W + Wa
= f
LCGm
2

B + Ba
= f
LCBm
2

where:

LCBm

mean distance from


midships of the centers
of gravity of the fore and
after bodies
= mean distance from
midships of the centers
of buoyancy of the
fore and after bodies

Since the sum of the weights of the fore


and after bodies is equal to the total weight,
which is equal to displacement, which is
similarly equal to the sum of the
buoyancies of the fore and after bodies, still
water bending moment can be expressed:
SWBM =

(LCGm
2

LCBm)

0.025
0.02
NORMALIZED BENDING MOMENT

LCGm =

0.015
0.01
0.005
HOG
0
SAG
-0.005
-0.01
-0.015
-0.02
-0.025
-1

If the mean of the centers of gravity is


greater than the mean of the centers of
buoyancy, the weight levers are longer than
the buoyancy levers, and the net moment is
hogging, as shown in Figure 1-80. If the
mean of buoyancy centers is greater, the
net moment is negative, and sagging.
Forward and after weight moments are determined by summing the moments of
individual weights. Weights and centers of
variable weights can be obtained from ships
officers or estimated with reasonable accuracy. Machinery weight can be approximated from the factors given in Paragraph
1-11.12.2; machinery lcg is determined by
inspection. Hull weight can be estimated as
described in Paragraph 1-11.12.2. The mean
distance from midships of the centers of
gravity of the forward and after bodies of the
hull can be expressed as a portion of length
between perpendiculars:

-0.8

-0.6

-0.4

-0.2

0.2

0.4

0.6

0.8

WAVE HEIGHT / STANDARD WAVE HEIGHT


Cb = 0.46

Cb = 0.58

Cb = 0.78

Cb = 0.84

Cb = 1.0

FROM WAVEHEIGHT AND WAVELENGTH VERSUS BENDING MOMENT STUDY FOR SUPERVISOR
OF SALVAGE U.S. NAVY, HERBERT ENGINEERING CORP., 20 FEBRUARY 1991

Figure 1-79. Normalized Wave Bending Moment as a Function of Wave Height.

LCGm

GF

BA

BF

GF

LCBm
MEAN DISTANCE TO FORE AND AFT LCGs GREATER
THAN MEAN DISTANCE TO LCBs - HOGGING

mean lcg = aL
where:

BA

=
=

=
=

an empirical coefficient
0.223 for a cargo ship with
forecastle and poop; deckhouse and machinery
amidships
0.24 for a tanker with forecastle, bridge, and poop
0.233 for a cargo ship with
machinery aft

GA

GF

BF

MEAN DISTANCE TO FORE AND AFT LCBs GREATER


THAN DISTANCE TO LCGs - SAGGING
Figure 1-80. Determination of Still Water Bending Moment by Murrays Method.

Values of a for different configurations can be estimated from those given above. For example, 0.225 might be used for a cargo ship with
machinery slightly aft of midships.

1-106

S0300-A8-HBK-010

Mean buoyancy moment can be estimated as:


MB =

cL
2

Table 1-21. Coefficient c for Mean LCB in Murrays Method.

Draft
0.06L
0.05L
0.04L
0.03L

where:

cL
L
c

=
=
=
=

total buoyancy (displacement), lton or tonne


mean position of LCB, ft or m
length between perpendiculars, ft or m
empirical coefficient based on block coefficient and
draft from Table 1-21

c
0.179CB + 0.063
0.189CB + 0.052
0.199CB + 0.041
0.209CB + 0.030

L = length between perpendiculars, block coefficient, CB is taken at draft equal to 0.06L

EXAMPLE 1-6
CALCULATION OF STILL WATER BENDING MOMENT BY MURRAYS METHOD

Calculate the still water bending moment for a cargo ship with machinery
and accommodations three-quarters aft with the following characteristics:
length between perpendiculars
570 feet
beam
80 feet
molded depth
55 feet
full load draft
35 feet
block coefficient
0.71
displacement
32,400 lton
deadweight
23,800 lton
hull weight
6,250 lton
weight of propulsion machinery 1,200 lton
center of machinery room
145 ft aft of midships
Variable Weight Distribution:
Weight
lton

item
Cargo: Hold 1
Hold 2
Hold 3
Hold 4
Hold 5
Oil fuel in deep tank
Oil fuel in double bottom tanks
Feed water
Potable water
Crew & effects, stores

3000
4200
6100
6800
3700
370
435
20
250
75

lcg from
midships
ft
231 F
142 F
60 F
95 A
250 A
200 F
85 A
170 A
122 A
165 A

Calculation:
Mean distance from midships of centers of buoyancy
The load draft of 35 ft is approximately 0.06L, CB = 0.71,

Weight moments, after body:


item
Weight
lton
Hold 4
Hold 5
O.F. (double bottom)
Feed water
Potable water
Machinery
Crew & effects, stores
Total:

lcg from
midships
ft
95 A
250 A
85 A
170 A
122 A
147 A
165 A

6,800
3,700
435
20
250
1,200
75
12,480

Weight moments, fore body:


item
weight
lton
Hold 1
3,000
Hold 2
4,200
Hold 3
6,100
O.F. (deep tank)
370
Total:
13,670

lcg from midships


ft
231 F
142 F
60 F
200 F

Total weight moments:


item
hull
after body
fore body
Total:

weight
lton
6,250
12,480
13,670
32,400

Moment
ft-lton
646,000
925,000
36,975
3,400
30,500
176,400
12,370
1,830,645

moment
ft-lton
693,000
596,400
366,000
74,000
1,729,400

moment
ft-lton
819,375
1,830,645
1,729,400
4,379,420

Mean distance from midships of centers of gravity:

LCGm

= Total moment/total weight


= 4,379,420/32,400 = 135.2 ft

= 0.179CB + 0.063 = 0.190


Still water bending moment:

cL = 0.19(570) = 108.3 ft = LCBm


Hull weight moment = WHaL (take a to be 0.23)
= 6,250(0.23)(570) = 819,375 ft-lton

SWBM = /2 (LCGm - LCBm)


= (32,400/2)(135.2 - 108.3)
= 435,780 ft-lton

LCGm is greater than LCBm; the net moment is positive, or hogging

1-107

S0300-A8-HBK-010

1-11.14.2 Wave Bending Moment. Wave bending moment,


for a standard wave with length equal to the ships length,
can be estimated as:
WBM =

bL 3B
1,000,000

for wave height =

2.2 b L 2.5 B
100,000

for wave height = 1.1 L

Table 1-22. Wave Bending Coefficient for Murrays Method.

L
20

Wave bending coefficient


b
Block Coefficient
CB
Hogging
(wave crest at midships)

Sagging
(wave trough at midships)

25.00
24.25
23.55
22.85
22.10
21.35
20.65
19.90
19.20
18.45
17.75

28.00
27.25
26.50
25.70
24.90
24.10
23.35
22.60
21.80
21.05
20.30

where:
WBM
L
B
b

=
=
=
=

0.80
0.78
0.76
0.74
0.72
0.70
0.68
0.66
0.64
0.62
0.60

wave bending moment, ft-lton


length between perpendiculars, ft
beam, ft
empirical coefficient based on block
coefficient and wave position, from Table
1-22

1-11.15 Section Property Design Rules. In the absence of


better information, empirical relationships and construction
standards can be used to estimate section modulus or moment
of inertia. The following design rules are taken from Applied
Naval Architecture, R. Munro-Smith, 1967.
A first approximation of the midships section moment of
inertia can be made from:
I =

CB taken at draft = 0.06L

cBD3

where:
I
B
D
c

=
=
=
=

moment of inertia, ft4 or m4


molded beam, ft or m
depth to strength deck, ft or m
empirical coefficient, ranging from 0.14 to 0.16
0.18 for cargo ships
0.22 for large tankers
0.175 to 0.21 for small tankers

An estimate for section modulus and/or moment of inertia can be made by reference to preliminary design expressions for maximum shear force
and bending moment, and assuming the ship was built to withstand that force and moment.

12

Mmax

L
C

Smax

LBTCD
35

L
C

L 2BTCB
35C

where:
Smax = maximum shear, lton

= displacement, lton
Mmax = maximum bending moment, ft-lton
L
= length between perpendiculars, ft
= block coefficient
CB
C
= a constant, generally ranging from 20 to 40
35 for most auxiliaries, merchant ships, and vessels with large longitudinal prismatic coefficient Mmax = LBT/1600 (CB taken
as 0.75)
20 for destroyers, and vessels with small longitudinal prismatic coefficient Mmax = LBT/1490 (CB taken as 0.47)
These relationships give a good approximation for the full-load condition on a standard hogging wave. For most merchant ships, hogging
moments are greater than sagging moments.

1-108

S0300-A8-HBK-010

1-11.16 By Rule Section Modulus. Classification society rules set minimum standards for midships section modulus. Midships section
modulus of an in class ship will not be lower than the minimum standard, and is unlikely to be much higher. Bending stresses in the midships
region can be roughly estimated without determining section modulus rigorously, provided the following are true:

The ship was built to classification society standards or other specifications requiring minimum section modulus, and is currently
in class.

The minimum section modulus standards are known.

The ship has not suffered damage that will reduce section modulus in the sections where stresses are to be determined.

A summary of section modulus requirements established by the American Bureau of Shipping (ABS) is given in Appendic C.
1-11.17 Strength Considerations in Salvage Operations. A ship is designed and constructed to withstand expected shear forces and bending
moments. In an intact floating ship, maximum bending moment occurs in the midships region and maximum shear near the quarter-length points.
These sections are designed to ensure that stresses remain below acceptable limits. Three conditions common to salvage operations may require
that the stress levels be examined at other points:

The ship may be loaded in ways not foreseen by the designer. Because of flooding, grounding or other unusual conditions of
loading, maximum bending moment can occur at some section other than midships. Similarly, maximum shear may be at some
point other than at the quarters.
Damage can alter the stress distribution at a section so that maximum stress can occur in some section other than where maximum
bending moment or shear occurs. Damage, even over a short distance, disrupts the continuity of longitudinal members and reduces
the section modulus for some distance on either side of the damaged section.
Local damage or distortion can render plating and stiffeners more susceptible to tripping, buckling, or other forms of load shirking,
thereby reducing effective moment of inertia.

The load, shear, and bending moment


curves of a casualty must be carefully
examined:

Stresses should be determined


wherever shear or bending
moment are maximum or the
effective moment of inertia is
reduced.
The effects of salvage actions
on load, shear and bending
moment should be examined
before taking the action.
Accesses should not be cut in
locations that will reduce the
section modulus or strength
member continuity.

A useful salvage technique is to calculate


and plot the maximum acceptable shear and
bending moments along the length of the
ship. The bending moments and shear
resulting from planned actions can be
compared with the allowable limits to
determine if the planned action is safe.
Figure 1-81 shows maximum acceptable
bending moments for an FFG-7 Class ship.

BENDING MOMENT, FOOT-POUNDS x 10

MAXIMUM MOMENT, DECK

2
MAXIMUM MOMENT, KEEL

17

16

15

14

13

13

12

11

10

STATIONS
MAXIMUM BENDING MOMENTS BASED ON ASSUMED
MATERIAL YIELD OF 32,000 PSI WITHOUT SAFETY FACTOR

Figure 1-81. Maximum Bending Moment for FFG-7.

1-109 (1-110 blank)

Das könnte Ihnen auch gefallen