The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact...

110
The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of Naval Architects and Marine Engineers 601 Pavonia Avenue Jersey City, New Jersey 07306

Transcript of The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact...

Page 1: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

The Principles of

Naval Architecture Series

Intact Stability

Colin S. Moore

J. Randolph Paulling, Editor

Published by The Society of Naval Architects

and Marine Engineers 601 Pavonia Avenue

Jersey City, New Jersey 07306

Page 2: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Copyright O 2010 by The Society of Naval Architects and Marine Engineers.

The opinions or assertions of the authors herein are not to be construed as of cia1 or re ecting the views of SNAME or any government agency.

It is understood and agreed that nothing expressed herein is intended or shall be construed to give any person, rm, or corporation any right, remedy, or claim against SNAME or any of its

of cers or member.

Library of Congress Cataloging-in-Publication Data

Moore, Colin S. Intact stability 1 Colin S. Moore. -- 1st ed. p. cm. -- (Principles of naval architecture)

Includes bibliographical references and index. ISBN 978-0-939773-74-9

I. Stability of ships. I. Title. VM159.M59 2010 623.8'171--dc22

2009043464

ISBN 978-0-939773-74-9

Printed in the United States of America

First Printing, 2010

Page 3: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Nomenclature

A &I

A P B B B I BL -

BM -

BML

b C CL CB CG G c p

D D DWT E e F F

F P FW G G, GM -

GML -

GZ 9 9 H h I IL IT i ,

i,.

K - K B KG KM - KMI, k L L

stands for area, generally area of waterplane after perpendicular maximum molded breadth center of buoyancy etc., changed positions of center of buoyancy molded baseline transverse metacentric radius, or height of M above B longitudinal metacentric radius, or height of ML above B width of a compartment or tank constant or coefficient centerline; a vertical plane through centerline block coefficient, VILBT center of gravity waterplane area coefficient, &/LB molded depth diameter, generally deadweight energy, generally base of Naperian logarithms, 2.7183 force, generally center of flotation (center of area of waterplane) forward perpendicular fresh water center of gravity of ship's mass etc., changed positions of the center of gravity transverse metacentric height, height of M above G longitudinal metacentric height, height of MI, above G righting arm; horizontal distance from G to Z acceleration due to gravity center of gravity of a component head depth of water or submergence moment of inertia, generally longitudinal moment of inertia of waterplane transverse moment of inertia of waterplane longitudinal moment of inertia of free surface in a compartment or tank transverse moment of inertia of free surface in a compartment or tank any point in a horizontal plane through the baseline height of B above the baseline height of G above the baseline height of M above the baseline height of ML above the baseline radius of gyration length, generally length of ship

LBP LPP LOA LwL Lw LCB LCF LCG LWL I M M ML MT MTcm MTI m m

rnL

0 ox OY OZ P P P Q R S SW T T Tw TCB TCG TPcrn TPI t t v Vk v c

VCB VCG vcg W

WL WL1 v v W

length between perpendiculars length between perpendiculars length overall length on designed load waterline length of a wave, from crest to crest longitudinal position of center of buoyancy longitudinal position of center of flotation longitudinal position of center of gravity load, or design, waterline length of a compartment of tank moment, generally transverse metacenter longitudinal metacenter trimming moment moment to trim 1 em moment to trim 1 inch mass, generally (W/g or w/g) transverse metacenter of liquid in a tank or compartment longitudinal metacenter of liquid in a tank or compartment origin of coordinates longitudinal axis of coordinates transverse axis of coordinates vertical axis of coordinates (upward) force of keel blocks pressure (force per unit area) in a fluid probability, generally fore and aft distance on a waterline radius, generally wetted surface of hull salt water draft period, generally period of a wave transverse position of center of buoyancy transverse position of center of gravity tons per em immersion tons per inch immersion thickness, generally time, generally linear velocity in general, speed of the ship speed of ship, knots speed of a surface wave (celerity) vertical position of center of buoyancy vertical position of center of gravity vertical position of g weight of ship equal to the displacement (pgV) of a ship floating in equilibrium any waterline parallel to baseline etc., changed position of WL volume of an individual item linear velocity weight of an individual item

Page 4: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

xvi NOMENCLATURE

x distance from origin along X-axis Y distance from origin along Y-axis x distance from origin along Z-axis Z a point vertically over B, opposite G

A,, displacement mass = pV A displacement force (buoyancy) = pgV 6 specific volume, or indicating a small change

0 angle of pitch or of trim (about OY-axis) P permeability P density; mass per unit volume 4) angle of heel or roll (about OX-axis) $ angle of yaw (about OZ-axis) V displacement volume cc) circular frequency, 2r/T, radians

Page 5: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Preface Intact Stability

During the twenty years that have elapsed since publication of the previous edition of this book, there have been remarkable advances in the art, science and practice of the design and construction of ships and other floating structures. In that edition, the increasing use of high speed computers was recognized and computational methods were incorporated or acknowledged in the individual chapters rather than being presented in a separate chapter. Today, the electronic computer is one of the most important tools in any engineering environment and the laptop computer has taken the place of the ubiquitous slide rule of an earlier generation of engineers.

Advanced concepts and methods that were only being developed or introduced then are a part of common engineering practice today. These include finite element analysis, computational fluid dynamics, random process methods, numerical modeling of the hull form and components, with some or all of these merged into integrated design and manufacturing systems. Collectively, these give the naval architect unprecedented power and flexibility to explore innovation in concept and design of marine systems. In order to fully utilize these tools, the modern naval architect must possess a sound knowledge of mathematics and the other fundamental sciences that form a basic part of a modern engineering education.

In 1997, planning for the new edition of Principles of Naval Architecture was initiated by the SNAME publica- tions manager who convened a meeting of a number of interested individuals including the editors of PNA and the new edition of Ship Design and Construction on which work had already begun. At this meeting it was agreed that PNA would present the basis for the modern practice of naval architecture and the focus would be principles in preference to applications. The book should contain appropriate reference material but it was not a handbook with extensive numerical tables and graphs. Neither was it to be an elementary or advanced textbook although it was expected to be used as regular reading material in advanced undergraduate and elementary graduate courses. It would contain the background and principles necessary to understand and to use intelligently the modern analytical, numerical, experimental and computational tools available to the naval architect and also the fundamentals needed for the development of new tools. In essence, it would contain the material necessary to develop the understanding, insight, intuition, experience and judgment needed for the successful practice of the profession. Following this initial meeting, a PNA Control Committee, consisting of individuals having the expertise deemed necessary to oversee and guide the writing of the new edition of PNA, was appointed. This committee, after participating in the selection of authors for the various chapters, has continued to contribute by critically reviewing the various component parts as they are written.

In an effort of this magnitude, involving contributions from numerous widely separated authors, progress has not been uniform and it became obvious before the halfway mark that some chapters would be completed before others. In order to make the material available to the profession in a timely manner it was decided to publish each major sub- division as a separate volume in the "Principles of Naval Architecture Series" rather than treating each as a separate chapter of a single book.

Although the United States committed in 1975 to adopt SI units as the primary system of measurement the transi- tion is not yet complete. In shipbuilding as well as other fields, we still find usage of three systems of units: English or foot-pound-seconds, SI or meter-newton-seconds, and the meter-kilogram(force)-second system common in engineer- ing work on the European continent and most of the non-English speaking world prior to the adoption of the SI system. In the present work, we have tried to adhere to SI units as the primary system but other units may be found particu- larly in illustrations taken from other, older publications. The symbols and notation follow, in general, the standards developed by the International Towing Tank Conference.

Several changes from previous editions of PNA may be attributed directly to the widespread use of electronic com- putation for most of the standard and nonstandard naval architectural computations. Utilizing this capability, many computations previously accomplished by approximate mathematical, graphical or mechanical methods are now car- ried out faster and more accurately by digital computer. Many of these computations are carried out within more com- prehensive software systems that gather input from a common database and supply results, often in real time, to the end user or to other elements of the system. Thus the hydrostatic and stability computations may be contained in a hull form design and development program system, intact stability is often contained in a cargo loading analysis system, damaged stability and other flooding effects are among the capabilities of salvage and damage control systems.

Page 6: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

x PREFACE

In this new edition of PNA, the principles of intact stability in calm water are developed starting from initial stability at small angles of heel then proceeding to large angles. Various effects on the stability are discussed such as changes in hull geometry, changes in weight distribution, suspended weights, partial support due to grounding or drydocking, and free liquid surfaces in tanks or other internal spaces. The concept of dynamic stability is introduced starting from the ship's response to an impulsive heeling moment. The effects of waves on resistance to capsize are discussed not- ing that, in some cases, the wave effect may result in diminished stability and dangerous dynamic effects.

Stability rules and criteria such as those of the International Maritime Organization, the US Coast Guard, and other regulatory bodies as well as the US Navy are presented with discussion of their physical bases and underlying assump- tions. The section includes a brief discussion of evolving dynamic and probabilistic stability criteria. Especial atten- tion is given to the background and bases of the rules in order that the naval architect may more clearly understand their scope, limitations and reliability in insuring vessel safety.

There are sections on the special stability problems of craft that differ in geometry or function from traditional seagoing ships including multihulls, submarines and oil drilling and production platforms. The final section treats the stability of high performance craft such as SWATH, planing boats, hydrofoils and others where dynamic as well as static effects associated with the vessel's speed and manner of operation must be considered in order to insure adequate stability.

J. RANDOLPH PAULLING Editor

Page 7: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Table of Contents

An Introduction to the Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

Foreword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . AuthorlsBiography xiii

Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv

ElementaryPrinciples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

Determining Vessel Weights and Center of Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

MetacentricHeight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

CurvesofStability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

EffectofFreeLiquids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

Effect of Changes in Weight on Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

Evaluation of Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

Draft, Trim. Heel. and Displacement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

The Inclining Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

SubmergedEquilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

TheTrimDive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

Methods of Improving Stability. Drafts. and List . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

StabilityWhenGrounded . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

AdvancedMarineVehicles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

Page 8: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

1 Elementary Principles

1.1 Gravitational Stability. A vessel must provide adequate buoyancy to support itself and its contents or working loads. It is equally important that the buoyancy be provided in a way that will allow the vessel to float in the proper attitude, or trim, and remain upright. This involves the problems of gravitational stability and trim. These issues will be discussed in detail in this chapter, primarily with reference to static conditions in calm water. Consideration will also be given to criteria for judging the adequacy of a ship's stability subject to both internal loading and external hazards.

It is important to recognize, however, that a ship or offshore structure in its natural sea environment is sub- ject to dynamic forces caused primarily by waves, wind, and, to a lesser extent, the vessel's own propulsion sys- tem and control surfaces. The specific response of the vessel to waves is typically treated separately as a ship motions analysis. Nevertheless, it is possible and advis- able to consider some dynamic effects while dealing with stability in idealized calm water, static conditions. This enables the designer to evaluate the survivability of the vessel at sea without performing direct motions analyses and facilitates the development of stability criteria. Evaluation of stability in this way will be ad- dressed in Section 7.

Another external hazard affecting a ship's stability is that of damage to the hull by collision, grounding, or other accident that results in flooding of the hull. The stability and trim of the damaged ship will be considered in Subdivision and Damage Stability (Tagg, 2010).

Finally, it is important to note that a floating struc- ture may be inclined in any direction. Any inclination may be considered as made up of an inclination in the athwartship plane and an inclination in the longitudi- nal plane. In ship calculations, the athwartship inclina- tion, called heel or list, and the longitudinal inclination, called trim, are usually dealt with separately. For float- ing platforms and other structures that have length to beam ratios of nearly 1.0, an off axis inclination is also often critical, since the vessel is not clearly dominated by either a heel or trim direction. This volume deals pri- marily with athwartship or transverse stability and lon- gitudinal stability of conventional ship-like bodies hav- ing length dimensions considerably greater than their width and depth dimensions. The stability problems of bodies of unusual proportions, including off-axis stabil- ity, are covered in Sections 4 and 7.

1.2 Concepts of Equilibrium. In general, a rigid body is considered to be in a state of static equilibrium when the resultants of all forces and moments acting on the body are zero. In dealing with static floating body sta- bility, we are interested in that state of equilibrium as- sociated with the floating body upright and at rest in a

still liquid. In this ease, the resultant of all gravity forces (weights) acting downward and the resultant of the buoyancy forces acting upward on the body are of equal magnitude and are applied in the same vertical line.

1.2.1 Stable Equilibrium. If a floating body, ini- tially at equilibrium, is disturbed by an external mo- ment, there will be a change in its angular attitude. If upon removal of the external moment, the body tends to return to its original position, it is said to have been in stable equilibrium and to have positive stability.

1.2.2 Neutral Equilibrium. If, on the other hand, a floating body that assumes a displaced inclination be- cause of an external moment remains in that displaced position when the external moment is removed, the body is said to have been in neutral equilibrium and has neutral stability. A floating cylindrical homogeneous log would be in neutral equilibrium in heel.

1.2.3 Unstable Equilibrium. If, for a floating body displaced from its original angular attitude, the dis- placement continues to increase in the same direction after the moment is removed, it is said to have been in unstable equilibrium and was initially unstable. Note that there may be a situation in which the body is stable with respect to "small" displacements and unstable with respect to larger displacements from the equilibrium position. This is a very common situation for a ship, and we will consider cases of stability at small angles of heel (initial stability) and at large angles separately.

1.3 Weight and Center of Gravity. This chapter deals with the forces and moments acting on a ship afloat in calm water. The forces consist primarily of grav- ity forces (weights) and buoyancy forces. Therefore, equations are usually developed using displacement, A, weight, W, and component weights, w. In the "Eng- lish" system, displacement, weights, and buoyant forces are thus expressed in the familiar units of long tons (or lb.). When using the International System of Units (SI), the displacement or buoyancy force is still expressed as A=pgV, but this is units of newtons which, for most ships, will be an inconveniently large number. In order to deal with numbers of more reasonable size, we may express displacement in kilonewtons or meganewtons. A non-SI force unit, the "metric ton force," or "tonnef," is defined as the force exerted by gravity on a mass of 1000 KG. If the weight or displacement is expressed in tonnef, its numerical value is approximately the same as the value in long tons, the unit traditionally used for ex- pressing weights and displacement in ship work. Since the shipping and shipbuilding industries have a long history of using long tons and are familiar with the nu- merical values of weights and forces in these units, the tonnef (often written as just tonne) has been and is still commonly used for expressing weight and buoyancy.

Page 9: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

2 INTACT STAB1 llTY

With this convention, righting and heeling moments are then expressed in units of metric ton-meters, t-m.

The total weight, or displacement, of a ship can be determined from the draft marks and curves of form, as discussed in Geometry of Ships (Letcher, 2009). The position of the center of gravity (CG) may be either cal- culated or determined experimentally. Both methods are used when dealing with ships. The weight and CG of a ship that has not yet been launched can be established only by a weight estimate, which is a summation of the estimated weights and moments of all the various items that make up the ship. In principle, all of the compo- nent parts that make up the ship could be weighed and recorded during the construction process to arrive at a finished weight and CG, but this is seldom done ex- cept for a few special craft in which the weight and CG are extremely critical. Weight estimating is discussed in Section 2.

After the ship is afloat, the weight and CG can be ac- curately established by an inclining experiment, as de- scribed in detail in Section 9.

To calculate the position of the CG of any object, it is assumed to be divided into a number of individual components or particles, the weight and CG of each be- ing known. The moment of each particle is calculated by multiplying its weight by its distance from a refer- ence plane, the weights and moments of all the particles added, and the total moment divided by the total weight of all particles, W. The result is the distance of the CG from the reference plane. The location of the CG is com- pletely determined when its distance from each of three planes has been established. In ship calculations, the three reference planes generally used are a horizontal plane through the baseline for the vertical location of the center of gravity (VCG), a vertical transverse plane either through amidships or through the forward per- pendicular for the longitudinal location (LCG), and a vertical plane through the centerline for the transverse position (TCG). (The TCG is usually very nearly in the centerline plane and is often assumed to be in that plane.)

1.4 Displacement and Center of Buoyancy. In Sec- tion 1, it has been shown that the force of buoyancy is equal to the weight of the displaced liquid and that the resultant of this force acts vertically upward through a point called the center of buoyancy, which is the CG of the displaced liquid (centroid of the immersed volume). Application of these principles to a ship, submarine, or other floating structure makes it possible to evaluate the effect of the hydrostatic pressure acting on the hull and appendages by determining the volume of the ship below the waterline and the centroid of this volume. The submerged volume, when multiplied by the specific weight of the water in which the ship floats is the weight of displaced liquid and is called the displacement, de- noted by the Greek symbol A.

1.5 Interaction of Weight and Buoyancy. The attitude of a floating object is determined by the interaction of the forces of weight and buoyancy. If no other forces are acting, it will settle to such a waterline that the force of buoyancy equals the weight, and it will rotate until two conditions are satisfied:

1. The centers of buoyancy B and gravity G are in the same vertical line, as in Fig. l(a), and

2. Any slight clockwise rotation from this position, as from WL to WILl in Fig. l(b), will cause the center of buoyancy to move to the right, and the equal forces of weight and buoyancy to generate a couple tending to move the object back to float on WL (this is the condi- tion of stable equilibrium).

For every object, with one exception as noted later, at least one position must exist for which these conditions are satisfied, since otherwise the object would continue to rotate indefinitely. There may be several such posi- tions of equilibrium. The CG may be either above or be- low the center of buoyancy, but for stable equilibrium, the shift of the center of buoyancy that results from a small rotation must be such that a positive couple (in a direction opposing the rotation) results.

An exception to the second condition exists when the object is a body of revolution with its CG exactly on the

Fig. 1 Stable equilibrium of floating body

Page 10: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

I

Fig. 2 Neutral equilibrium of

axis of revolution, as illustrated in Fig. 2. When such an object is rotated to any angle, no moment is produced, since the center of buoyancy is always directly below the CG. It will remain at any angle at which it is placed (this is a condition of neutral equilibrium).

A submerged object whose weight equals its buoy- ancy that is not in contact with the seafloor or other ob- jects can come to rest in only one position. It will rotate until the CG is directly below the center of buoyancy. If its CG coincides with its center of buoyancy, as in the case of a homogeneous object, it would remain in any position in which it is placed since in this case it is in neutral equilibrium.

The difference in the action of floating and sub- merged objects is explained by the fact that the center of buoyancy of the submerged object is fixed relative to the body, while the center of buoyancy of a floating ob- ject will generally shift when the object is rotated as a result of the change in shape of the immersed part of the body.

As an example, consider a watertight body having a rectangular section with dimensions and CG as illus- trated in Fig. 3. Assume that it will float with half its volume submerged, as in Fig. 4. It can come to rest in either of two positions, (a) or (c), 180 degrees apart. In either of these positions, the centers of buoyancy and gravity are in the same vertical line. Also, as the body is inclined from (a) to (b) or from (c) to (d), a moment is developed which tends to rotate the body back to its original position, and the same situation would exist if it were inclined in the opposite direction.

1- 20 cm -4 Fig. 3 Example of stability of watertight rectangular body.

floating body.

If the 20-em dimension were reduced with the CG still on the centerline and 2.5 em below the top, a situation would be reached where the center of buoyancy would no longer move far enough to be to the right of the CG as the body is inclined from (a) to (b). Then the body could come to rest only in position (c).

As an illustration of a body in the submerged condi- tion, assume that the weight of the body shown in Fig. 3 is increased so that the body is submerged, as in Fig. 5. In positions (a) and (c), the centers of buoyancy and gravity are in the same vertical line. An inclination from (a) in either direction would produce a moment tending to rotate the body away from position (a), as illustrated in Fig. 5(b). An inclination from (c) would produce a mo- ment tending to restore the body to position (c). There- fore, the body can come to rest only in position (c).

A ship or submarine is designed to float in the upright position. This fact permits the definition of two classes of hydrostatic moments, illustrated in Fig. 6, as follows:

Righting moments: A righting moment exists at any angle of inclination where the forces of weight and buoy- ancy act to move the ship toward the upright position.

Overturning moments: An overturning moment exists at any angle of inclination where the forces of weight and buoyancy act to move the ship away from the upright position.

The center of buoyancy of a ship or a surfaced sub- marine moves with respect to the ship, as the ship is inclined, in a manner that depends upon the shape of the ship in the vicinity of the waterline. The center of buoyancy of a submerged submarine, on the contrary, does not move with respect to the ship, regardless of the inclination or the shape of the hull, since it is station- ary at the CG of the entire submerged volume. This con- stitutes an important difference between floating and submerged ships. The moment acting on a surface ship can change from a righting moment to an overturning moment, or vice versa, as the ship is inclined, but this cannot occur on a submerged submarine unless there is a shift of the ship's CG.

It can be seen from Fig. 6 that lowering of the CG along the ship's centerline increases stability. When a righting moment exists, lowering the CG along the cen-

Page 11: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

A

(c) (d) Fig. 4 Alternate conditions of stable equilibrium for floating body.

terline increases the separation of the forces of weight and buoyancy and increases the righting moment. When an overturning moment exists, sufficient lowering of the CG along the centerline would change the moment to

a righting moment, changing the stability of the initial upright equilibrium from unstable to stable.

In problems involving longitudinal stability of undam- aged surface ships, we are concerned primarily with de-

Fig. 5 Singe condition of stable equilibrium for submerged floating body

Page 12: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STABl llTY

SURFACE -\ SHIP -\

POSITIVE STABl LlTY

NEGATIVE STAB l LlTY

(a) RIGHTING MOMENT WHEN HEELED (b) OVERTURNING MOMENT WHEN HEELED

POSITIVE STABl LlTY

SUBMERGED SUBMARINE

NEGATIVE STAB l LlTY

(c) RIGHTING MOMENT WHEN HEELED (d) OVERTURNING MOMENT WHEN HEELED Fig. 6 Effect of height of CG on stability.

termining the ship's draft and trim under the influence of various upsetting moments, rather than evaluating the possibility of the ship capsizing in the longitudinal direc- tion. If the longitudinal centers of gravity and buoyancy are not in the same vertical line, the ship will change trim as discussed in Section 8 and will come to rest as illus- trated in Fig. 7, with the centers of gravity and buoyancy in the same vertical line. A small longitudinal inclination will cause the center of buoyancy to move so far in a fore and aft direction that the moment of weight and buoy- ancy would be many times greater than that produced by the same inclination in the transverse direction. The lon- gitudinal shift in buoyancy creates such a large longitudi- nal righting moment that longitudinal stability is usually very great compared to transverse stability.

Thus, if the ship's CG were to rise along the center- line, the ship would capsize transversely long before there would be any danger of capsizing longitudinally. However, a surface ship could, theoretically, be made to founder by a downward external force applied toward

one end, at a point near the centerline, and at a height near or below the center of buoyancy without capsizing. It is unlikely, however, that an intact ship would encoun- ter a force of the required magnitude.

Surface ships can, and do, founder after extensive flooding as a result of damage at one end. The loss of buoyancy at the damaged end causes the center of buoy- ancy to move so far toward the opposite end of the ship that subsequent submergence of the damaged end is not

Fig. 7 Longitudinal equilibrium.

Page 13: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

6 INTACT STAB1 l lTY

adequate to move the center of buoyancy back to a posi- tion in line with the CG, and the ship founders, or cap- sizes longitudinally. The behavior of a partially flooded ship is discussed in Tagg (2010).

In the case of a submerged submarine, the center of buoyancy does not move as the submarine is inclined in a fore-and-aft direction. Therefore, capsizing of an in- tact submerged submarine in the longitudinal direction is possible and would require very nearly the same mo- ment as would be required to capsize it transversely. If the CG of a submerged submarine were to rise to a posi- tion above the center of buoyancy, the direction, longi- tudinal or transverse, in which it would capsize would depend upon the movement of liquids or loose objects within the ship. The foregoing discussion of submerged submarines does not take into account the stabilizing effect of the bow and stern planes which have an impor- tant effect on longitudinal stability while the ship is un- derway with the planes producing hydrodynamic lift.

1.6 Upsetting Force. The magnitude of the upsetting forces, or heeling moments, that may act on a ship deter- mines the magnitude of moment that must be generated by the forces of weight and buoyancy in order to prevent capsizing or excessive heel.

External upsetting forces affecting transverse stabil- ity may be caused by:

Beam winds, with or without rolling. Lifting of heavy weights over the side. High-speed turns. Grounding. Strain on mooring lines. Towline pull of tugs.

Internal upsetting forces include:

Shifting of on-board weights athwartship. Entrapped water on deck.

Section 7 discusses evaluation of stability with re- gard to the upsetting forces listed above. The discussion below is general in nature and illustrates the stability principles involved when a ship is subjected to upsetting forces.

When a ship is exposed to a beam wind, the wind pressure acts on the portion of the ship above the water- line, and the resistance of the water to the ship's lateral motion exerts a force on the opposite side below the wa- terline. The situation is illustrated in Fig. 8. Equilibrium with respect to angle of heel will be reached when:

The ship is moving to leeward with a speed such that the water resistance equals the wind pressure, and

The ship has heeled to an angle such that the moment produced by the forces of weight and buoyancy equals the moment developed by the wind pressure and the wa- ter pressure.

As the ship heels from the vertical, the wind pres- sure, water pressure, and their vertical separation re- main substantially constant. The ship's weight is con-

PRESSURE /

CL

Fig. 8 Effect of a beam wind

stant and acts at a fixed point. The force of buoyancy also is constant, but the point at which it acts varies with the angle of heel. Equilibrium will be reached when sufficient horizontal separation of the centers of grav- ity and buoyancy has been produced to cause a balance between heeling and righting moments.

When a weight is lifted over the side, as illustrated in Fig. 9, the force exerted by the weight acts through the outboard end of the boom, regardless of the angle of heel or the height to which the load has been lifted. Therefore, the weight of the sidelift may be considered to be added to the ship at the end of the boom. If the ship's CG is initially on the ship's centerline, as at G in Fig. 9, the CG of the combined weight of the ship and the sidelift will be located along the line GA and will move to a final position, GI, when the load has been lifted clear of the pier. Point GI will be off the ship's centerline and somewhat higher than G. The ship will heel until the

Fig. 9 Lifting a weight over the side

Page 14: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 7

CL Fig. 10 Effect of offside weight.

center of buoyancy has moved off the ship's centerline to a position directly below point GI.

Movement of weights already aboard the ship, such as passengers, liquids, or cargo, will cause the ship's CG to move. If a weight is moved from A to B in Fig. 10, the ship's CG will move from G to GI in a direction parallel to the direction of movement of the shifted weight. The ship will heel until the center of buoyancy is directly be- low point GI.

When a ship is executing a turn, the dynamic loads from the control surfaces and external pressure accel- erate the ship towards the center of the turn. In a static evaluation, the resulting inertial force can be treated as a centrifugal force acting horizontally through the ship's

CENTRIFUGAL FORCE ? G I ' o B,

CG. This force is balanced by a horizontal water pres- sure on the side of the ship, as illustrated in Fig. ll(a). Except for the point of application of the heeling force, the situation is similar to that in which the ship is acted upon by a beam wind, and the ship will heel until the moment of the ship's weight and buoyancy equals that of the centrifugal force and water pressure.

If a ship runs aground in such a manner that contact with the seafloor occurs over a small area (point con- tact), the sea bottom offers little restraint to heeling, as il- lustrated in Fig. ll(b), and the reaction between ship and seafloor of the bottom may produce a heeling moment. As the ship grounds, part of the energy due to its forward motion may be absorbed in lifting the ship, in which case a reaction, R, between the bottom and the ship would de- velop. This reaction may be increased later as the tide ebbs. Under these conditions, the force of buoyancy would be less than the weight of the ship because the ship would be supported by the combination of buoyancy and the reaction at the point of contact. The ship would heel until the moment of buoyancy about the point of contact became equal to the moment of the ship's weight about the same point, when (W - R) x a equals W x 6.

There are numerous other situations in which ex- ternal forces can produce heel. A moored ship may be heeled by the combination of strain on the mooring lines and pressure produced by wind or current. Tow- line strain may produce heeling moments in either the towed or towing ship. In each ease, equilibrium would be reached when the center of buoyancy has moved to a point where heeling and righting moments are balanced.

In any of the foregoing examples, it is quite possible that equilibrium would not be reached before the ship

(a) EFFECT OF A TURN (b) EFFECT OF GROUNDING Fig. 1 1 Effect of a turn and grounding.

Page 15: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

8 INTACT STAB1 llTY

capsized. It is also possible that equilibrium would not be reached until the angle of heel became so large that water would be shipped through topside openings, and that the weight of this water, running to the low side of the ship, would contribute to capsizing which otherwise would not have occurred.

Upsetting forces act to incline a ship in the longitudi- nal as well as the transverse direction. Since a surface ship is much stiffer, however, in the longitudinal direc- tion, many forces, such as wind pressure or towline strain, would not have any significant effect in inclining the ship longitudinally. Shifting of weights aboard in a longitudinal direction can cause large changes in the attitude of the ship because the weights can be moved much farther than in the transverse direction. When very heavy lifts are to be attempted, as in salvage work, they are usually made over the bow or stern rather than over the side, and large longitudinal inclinations may be involved in these operations. Stranding at the bow or stern can produce substantial changes in trim. In each ease, the principles are the same as previously discussed for transverse inclinations. When a weight is shifted longitudinally or lifted over the bow or stern, the CG of the ship will move, and the ship will trim until the center of buoyancy is directly below the new position of the CG. If a ship is grounded at the bow or stern, it will assume an attitude such that the moments of weight and buoyancy about the point of contact are equal.

In the case of a submerged submarine, the center of buoyancy is fixed, and a given upsetting moment pro- duces very nearly the same inclination in the longitudi- nal direction as it does in the transverse direction (Fig. 12). The only difference, which is trivial, is because of the effect of liquids aboard which may move to a differ- ent extent in the two directions. A submerged subma- rine, however, is comparatively free from large upset- ting forces. Shifting of the CG as the result of weight changes is carefully avoided. For example, when a tor- pedo is fired, its weight is immediately replaced by an equal weight of water at the same location.

1.7 Submerged Equilibrium. Before a submarine is submerged, considerable effort has been expended, both in design and operation, to ensure that:

The weight of the submarine, with its loads and bal- last, will be very nearly equal to the weight of the water it will displace when submerged.

The CG of these weights will be very nearly in the same longitudinal position as the center of buoyancy of the submerged submarine.

The CG of these weights will be lower than the center of buoyancy of the submerged submarine.

These precautions produce favorable conditions that are described, respectively, as neutral buoyancy, zero trim, and positive stability. A submarine on the surface, with weights adjusted so that the first two conditions will be satisfied upon filling the main ballast tanks, is said to be in diving trim.

I i ~ d = ~ ~ ~ ~ I I

INITIAL POSITION ZERO HEEL, UPSETTING NEW

NEUTRAL BUOYANCY MOMENT EQUILIBRIUM

a) TRANSVERSE STABILITY

INITIAL ZERO TRIM POSITION - NEUTRAL BUOYANCY

I-d---I TRIMMING MOMENT - wd= WXG,

NEW EQUILIBRIUM POSITION

b) LONGITUDINAL STABILITY Fig. 12 Effect of weight shift on the transverse and longitudinal stability of

a submerged submarine.

The effect of this situation is that the submarine, in- sofar as transverse and longitudinal stability are con- cerned, acts in the same manner as a pendulum. This imaginary pendulum is supported at the center of buoy- ancy, has a length equal to the separation of the ten- ters of buoyancy and gravity, and a weight equal to the weight of the submarine.

Page 16: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 9

It is not practical to achieve an exact balance of It is also not necessary, since minor deviations can be weight and buoyancy or to bring the CG precisely to the counteracted by the effect of the bow and stern planes same longitudinal position as the center of buoyancy. when underway submerged.

Determining Vessel Weights and Center of Gravity

2.1 Weight and Location of Center of Gravity. It is im- portant that the weight and the location of the CG be estimated at an early stage in the design of a ship. The weight and height of the CG are major factors in deter- mining the adequacy of the ship's stability. The weight and longitudinal position of the CG determine the drafts at which the ship will float. The distance of the CG from the ship's centerline plane determines whether the ship will have an unacceptable list. It will be clear that this calculation of weight and CG, although laborious and tedious, is one of the most important steps in the suc- cessful design of ships.

During the early stages of design, the weight and the height of CG for the ship in light condition are estimated by comparison with ships of similar type or from coef- ficients derived from existing ships. At later stages of design, detailed estimates of weights and CGs are re- quired. It is often necessary to modify ship dimensions or the distribution of weights to achieve the desired op- timum combination of a ship's drafts, trim, and stability, as well as to meet other design requirements such as motions in waves and powering. Sample lightship, full load, and ballast load conditions are shown in Table 1.

2.2 Detailed Estimates of Weights and Position of Cen- ter of Gravity. The reader is referred to Chapter 12, by W. Boze, of Ship Design and Construction (Lamb, 2003) for a detailed discussion of the methodology of weight estimating for each design stage, starting with concept design and ending with detail design.

Ordinarily in design, the horizontal plane of refer- ence is taken through the molded baseline of the ship, described in Letcher (2009). The height of the CG above

this base is referred to as KG and its position as VCG. Sometimes, after a ship's completion, the reference plane is taken through the bottom of the keel, which, depending on the definition of the molded surface, may be a few centimeters below the molded surface.

The plane of reference for the longitudinal position of the CG may be the transverse plane at the midship section, which is midway between the forward and af- ter perpendiculars. In this case, the LCG is measured forward or abaft the midship section. This practice in- volves the possibility of inadvertently applying the mea- surements aft instead of forward, or vice versa, and a more desirable plane of reference is one through the af- ter or forward perpendicular.

The plane of reference for the transverse position of the CG is the vertical centerline plane of the ship, the transverse position of the CG being measured to port or starboard of this plane.

In weight estimates, it is essential that an orderly and systematic classification of weights be followed. Two such classifications are in general use in this country: Classification of Merchant Sh ip Weights by the U.S. Maritime Administration (MARAD, 1995), and Expanded Sh ip Work Breakdown Structure (ESWBS) by the U.S. Navy (NAVSEA, 1985). The MARAD system uses three broad classifications of hull (steel, outfit, and machinery) each further subdivided into 10 subgroups. The ESWBS uses nine major classifications reflecting the mission requirements of military vessels. Further recommendations on weight control techniques can be found in the Recommended Practice No. 12 produced by the International Society of Allied Weight Engineers

Table 1 Sample summaries of loading condition weights and centers.

Post-Panamax Containership Aframax Tanker 132,000 m3 LNG (Membrane Type) Handymax Bulk Carrier Carrier

Mass* Displace- VCG' LCG** Mass Displace- VCG LCG Mass Displace- VCG LCG Mass Displace- VCG LCG ment' ment ment ment

Lightship 24,510 240,223 61% -7% 19,004 186,258 53% -5% 28,017 274,595 75% -5% 7289 71,439 73% -7%

Full Load 76,318 747,993 71% -3% 129,032 1,264,643 57% 3% 99,899 979,110 73% 0% 35,453 347,475 60% 2%

Ballast 49,275 482,944 45% -4% 62,070 608,348 39% 1% 75,561 740,573 64% 1% 25,944 254,277 56% 2%

LBP (m) 262 239 275 160

Depth 24.3 21.0 20.1 13.6

(m)

*In tonnes. 'In Kilo-Newtons. :Percent of depth. **Percent of LBP fonvard (+) or aft (-) of midships.

Page 17: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 9

It is not practical to achieve an exact balance of It is also not necessary, since minor deviations can be weight and buoyancy or to bring the CG precisely to the counteracted by the effect of the bow and stern planes same longitudinal position as the center of buoyancy. when underway submerged.

Determining Vessel Weights and Center of Gravity

2.1 Weight and Location of Center of Gravity. It is im- portant that the weight and the location of the CG be estimated at an early stage in the design of a ship. The weight and height of the CG are major factors in deter- mining the adequacy of the ship's stability. The weight and longitudinal position of the CG determine the drafts at which the ship will float. The distance of the CG from the ship's centerline plane determines whether the ship will have an unacceptable list. It will be clear that this calculation of weight and CG, although laborious and tedious, is one of the most important steps in the suc- cessful design of ships.

During the early stages of design, the weight and the height of CG for the ship in light condition are estimated by comparison with ships of similar type or from coef- ficients derived from existing ships. At later stages of design, detailed estimates of weights and CGs are re- quired. It is often necessary to modify ship dimensions or the distribution of weights to achieve the desired op- timum combination of a ship's drafts, trim, and stability, as well as to meet other design requirements such as motions in waves and powering. Sample lightship, full load, and ballast load conditions are shown in Table 1.

2.2 Detailed Estimates of Weights and Position of Cen- ter of Gravity. The reader is referred to Chapter 12, by W. Boze, of Ship Design and Construction (Lamb, 2003) for a detailed discussion of the methodology of weight estimating for each design stage, starting with concept design and ending with detail design.

Ordinarily in design, the horizontal plane of refer- ence is taken through the molded baseline of the ship, described in Letcher (2009). The height of the CG above

this base is referred to as KG and its position as VCG. Sometimes, after a ship's completion, the reference plane is taken through the bottom of the keel, which, depending on the definition of the molded surface, may be a few centimeters below the molded surface.

The plane of reference for the longitudinal position of the CG may be the transverse plane at the midship section, which is midway between the forward and af- ter perpendiculars. In this case, the LCG is measured forward or abaft the midship section. This practice in- volves the possibility of inadvertently applying the mea- surements aft instead of forward, or vice versa, and a more desirable plane of reference is one through the af- ter or forward perpendicular.

The plane of reference for the transverse position of the CG is the vertical centerline plane of the ship, the transverse position of the CG being measured to port or starboard of this plane.

In weight estimates, it is essential that an orderly and systematic classification of weights be followed. Two such classifications are in general use in this country: Classification of Merchant Sh ip Weights by the U.S. Maritime Administration (MARAD, 1995), and Expanded Sh ip Work Breakdown Structure (ESWBS) by the U.S. Navy (NAVSEA, 1985). The MARAD system uses three broad classifications of hull (steel, outfit, and machinery) each further subdivided into 10 subgroups. The ESWBS uses nine major classifications reflecting the mission requirements of military vessels. Further recommendations on weight control techniques can be found in the Recommended Practice No. 12 produced by the International Society of Allied Weight Engineers

Table 1 Sample summaries of loading condition weights and centers.

Post-Panamax Containership Aframax Tanker 132,000 m3 LNG (Membrane Type) Handymax Bulk Carrier Carrier

Mass* Displace- VCG' LCG** Mass Displace- VCG LCG Mass Displace- VCG LCG Mass Displace- VCG LCG ment' ment ment ment

Lightship 24,510 240,223 61% -7% 19,004 186,258 53% -5% 28,017 274,595 75% -5% 7289 71,439 73% -7%

Full Load 76,318 747,993 71% -3% 129,032 1,264,643 57% 3% 99,899 979,110 73% 0% 35,453 347,475 60% 2%

Ballast 49,275 482,944 45% -4% 62,070 608,348 39% 1% 75,561 740,573 64% 1% 25,944 254,277 56% 2%

LBP (m) 262 239 275 160

Depth 24.3 21.0 20.1 13.6

(m)

*In tonnes. 'In Kilo-Newtons. :Percent of depth. **Percent of LBP fonvard (+) or aft (-) of midships.

Page 18: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

(ISAWE, 1997). Some design offices may use systems differing in detail from either of these, but the general classification will be similar.

2.3 Weight and Center of Gravity Margins. The weight estimate will of necessity contain many approxi- mations and, it may be presumed, some errors. The er- rors will generally be errors of omission. The steel as received from the mills is usually heavier, within the mill tolerance, than the ordered nominal weight. It is impossible, in the design stages, to calculate in accurate detail the weight of many groups such as piping, wiring, auxiliary machinery, and many others.

For these and similar reasons, it is essential that mar- gins for error be included in the weight estimate. The amount of these margins is derived from the experience of the estimator and varies with the accuracy and ex- tent of the available information.

Table 2 is a composite of the usual practice of sev- eral design offices. In each instance, the smaller values apply to conventional ships that do not involve unusual features and for which there is a reliable basis for the estimate. If the estimate is reviewed by several inde- pendent interested agencies, there is less chance of substantial error and smaller margins are in order. The

Table 2 Weight margins.

Margin of Weight (in percent of lightship weight)

Cargo ships 1.5-2.5

Tankers 1.5-2.5

Cargo-passenger ships 2.0-3.0

Large passenger ships 2.5-3.5

Small naval vessels 6.0-7.0

Large naval vessels 3.5-7.0

Margin in VCG Meters

Cargo ships 0.15-0.23

Tankers 0.15

Cargo-passenger ships 0.15-0.23

Large passenger ships 0.23-0.30

Small naval vessels 0.15-0.23

Large naval vessels 0.15-0.23

larger values apply to vessels with unusual features or in which there is considerable uncertainty as to the ulti- mate development of the design.

The amount of margin will also depend on the seri- ousness of misestimating weight or CG. For example, until the advent of the double bottom for tankers, there was no real need for any margin at all in the VCG of a conventional tanker because such ships generally have considerably more stability than is needed. On the other hand, if there were a substantial penalty in the contract for overweight or for a high VCG, a correspondingly sub- stantial margin in the estimate would be indicated.

The above margins apply to estimates made in the contract-design stage, where the calculations are based primarily on a midship section, arrangement drawings, and the specifications. In a final, detailed finished- weight calculation, made mostly from working draw- ings, a much smaller margin, of 1% or 2%, or even, if extremely detailed information is available, no margin at all may be appropriate.

Margins assigned to U.S. military ships (NAVSEA, 2001) are called acquisition margins and include Pre- liminary and Contact Design Margins, Detail Design and Build Margins, Contract Modifications Margin, and Government Furnished Material Margin. The U.S. Navy also includes Service Life Allowances that range from 5% to 10% for weight and 0.5 to 2.5 ft (0.15 to 0.75 m) for the VCG to allow for future modifications and additions to the ship.

For more detailed information on margins and allow- ances, the reader is referred to Chapter 12 in Ship De- sign and Construction (Lamb, 2003).

2.4 Variation in Displacement and Position of Center of Gravity With Loading of Ship. The total weight (displace- ment) and position of the CG of any ship in service will depend greatly on the amount and location of the dead- weight items discussed in Letcher (2009): cargo, fuel, fresh water, stores, etc. Hence, the position of the CG is determined for various operating conditions of the ship, the conditions depending upon the class of ship (see Sec- tion 3.8). These are usually calculated using an onboard loading computer that has capabilities for tracking cargo weight, ship stability, and strength (Fig. 13).

Page 19: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

Fig. 13 Sample loading computer display.

Metacentric Height

3.1 The Transverse Metacenter and Transverse Meta- centric Height. Consider a symmetric ship heeled to a very small angle, 64, shown, with the angle exagger- ated, in Fig. 14. The center of buoyancy has moved off the ship's centerline as the result of the inclination, and the lines along which the resultants of weight and buoyancy act are separated by a distance, m, the right- ing arm. In the limit 64 + 0, a vertical line through the

Fig. 14 Metacenter and righting arm

center of buoyancy will intersect the original vertical through the center of buoyancy, which is normally in the ships centerline plane at a point M, called the transverse metacenter. The location of this point will vary with the ship's displacement and trim, but, for any given drafts, it will always be in the same place.

Unless there is an abrupt change in the shape of the ship in the vicinity of the waterline, point M will remain practically stationary with respect to the ship as the ship is inclined to small angles, up to about 7 degrees.

As can be seen from Fig. 14, if the locations of G and M are known, the righting arm for small angles of heel can be calculated readily, with sufficient accuracy for all practical purposes, by the formula

- - GZ = GM sin 64 (1)

The distance m i s therefore important as an index of transverse stability at small angles of heel, and is called the transverse metacentric height. Since m is consid- ered positive when the moment of weight and buoyancy tends to rotate the ship toward the upright position, ?% is positive when M is above G, and negative when M is below G.

Metacentric Height (rm is often used as an index of stability when preparation of stability curves for large an-

mgray
3
mgray
3
Page 20: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

Fig. 13 Sample loading computer display.

Metacentric Height

3.1 The Transverse Metacenter and Transverse Meta- centric Height. Consider a symmetric ship heeled to a very small angle, 64, shown, with the angle exagger- ated, in Fig. 14. The center of buoyancy has moved off the ship's centerline as the result of the inclination, and the lines along which the resultants of weight and buoyancy act are separated by a distance, m, the right- ing arm. In the limit 64 + 0, a vertical line through the

Fig. 14 Metacenter and righting arm

center of buoyancy will intersect the original vertical through the center of buoyancy, which is normally in the ships centerline plane at a point M, called the transverse metacenter. The location of this point will vary with the ship's displacement and trim, but, for any given drafts, it will always be in the same place.

Unless there is an abrupt change in the shape of the ship in the vicinity of the waterline, point M will remain practically stationary with respect to the ship as the ship is inclined to small angles, up to about 7 degrees.

As can be seen from Fig. 14, if the locations of G and M are known, the righting arm for small angles of heel can be calculated readily, with sufficient accuracy for all practical purposes, by the formula

- - GZ = GM sin 64 (1)

The distance m i s therefore important as an index of transverse stability at small angles of heel, and is called the transverse metacentric height. Since m is consid- ered positive when the moment of weight and buoyancy tends to rotate the ship toward the upright position, ?% is positive when M is above G, and negative when M is below G.

Metacentric Height (rm is often used as an index of stability when preparation of stability curves for large an-

mgray
3
mgray
3
Page 21: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 l lTY

Fig. 15 Locating the transverse metacenter.

gles (Section 4) has not been made. Its use is based on the assumption that adequate in conjunction with ade- quate freeboard, will assure that adequate righting mo- ments will exist at both small and large angles of heel.

3.2 Location of the Transverse Metacenter. When a symmetric ship is inclined to a small angle, as in Fig. 15, the new waterline will intersect the original waterline at the ship's centerline plane if the ship is wall-sided in the vicinity of the waterline because the volumes of the two wedges between the two waterlines will then be equal, and there will be no change in displacement. If v is the volume of each wedge, V the volume of displacement, and the CGs of the wedges are at gl and g,, the ship's center of buoyancy will move:

In a direction parallel to a line connecting g, and g,. A distance, Dl, equal to (v . glg2)lV.

As the angle of heel approaches zero, the line 9291, and therefore m, becomes perpendicular to the ship's centerline. Also, any variation from wall-sidedness be- comes negligible, and we may say

If y is the half-breadth of the waterline at any point of the ship's length at a distance x from one end, and if the ship's length is designated as L, the area of a sec- tion through the wedges is i(y)(y tan 64) and its cen- troid is at a distance of 2 x y from the centroid of the corresponding section on the other side v x 'g1g2 =

I,

S+(y)(y tan 64)(2 x y)dx or - o tan64 3 ,

The right hand side of this expression, 4 y" dx, is recognized as the moment of inertia of an area bounded by a curve and a straight line with the straight line as the axis. If we consider the straight line to be the ship's cen- terline, then the moment of inertia of the entire water- plane about the ship's centerline (both sides) designated

z L as IT, is 1, = - S y%d?: = 9' and, therefore, when 3 0 tan 64

This theorem was derived by the French hydrographer Pierre Bouger while on an expedition to Peru to mea- sure a degree of the meridian near the equator. It ap- peared in his Trait6 du Navire, published in Paris in 1746. It can be shown that BM is equal to the radius of curvature of the locus of B as 64 - 0.

The height of the transverse metacenter above the keel, usually called is just the sum of m, or IT& and m, the height of the center of buoyancy above the keel. The height of the center of gravity above the keel,

is found from the weight estimate or inclining ex- periment. Then,

- - - GM = K M - KG

3.3 The Longitudinal Metacenter and Longitudinal Metacentric Height. The longitudinal metacenter is similar to the transverse metacenter except that it in- volves longitudinal inclinations. Since ships are usually not symmetrical forward and aft, the center of buoyancy at various even keel waterlines does not always lie in a fixed transverse plane but may move forward and aft with changes in draft. For a given even keel waterline, the longitudinal metacenter is defined as the intersec- tion of a vertical line through the center of buoyancy in the even keel attitude with a vertical line through the new position of the center of buoyancy after the ship has been inclined longitudinally through a small angle.

The longitudinal metacenter, like the transverse metacenter, is substantially fixed with respect to the ship for moderate angles of inclination if there is no abrupt change in the shape of the ship in the vicinity of the waterline, and its distance above the ship's CG, or the longitudinal metacentric height, is an index of the ship's resistance to changes in trim. For a normal surface ship, the longitudinal metacenter is always far above the CG, and the longitudinal metacentric height is always positive.

3.4 Location of the Longitudinal Metacenter. Locating the longitudinal metacenter is similar to, but somewhat more complicated than, locating the transverse meta- center. Since the hull form is usually not symmetrical in the fore-and-aft direction, the immersed wedge and the emerged wedge usually do not have the same shape. To maintain the same displacement, however, they must have the same volume. Fig. 16 shows a ship inclined lon- gitudinally from an even keel waterline WL, through a small angle, 84, to waterline WILl. Using the intersec- tion of these two waterlines, point F, as the reference for fore and aft distances, and letting:

L = length of waterplane & = distance from F to the forward end of waterplane y = breadth of waterline WL at any distance x from F

Page 22: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

Fig. 16 Longitudinal metacenter

the volume of the forward wedge is

and the volume of the after wedge is

Equating the volumes

Q L-Q

t a n ~ ~ J , x y d x = t a n ~ O J , x z j d x 0 0

These expressions are, respectively, the moment of the area of the waterplane forward of F and the mo- ment of the area aft of F, both moments being about a transverse line through point F. Since these moments are equal and opposite, the moment of the entire wa- terplane about a transverse axis through F is zero, and therefore F lies on the transverse axis through the centroid of the waterplane, called the center of f lotation.

In Fig. 16, AB is a transverse vertical plane through the initial position of the center of buoyancy, B, when the ship was floating on the even keel waterline, WL. With longitudinal inclination, B will move parallel to gig,, or as the inclination approaches zero, perpendicu- lar to plane AB, to a point B,. The height of the metacen- ter above B will be

The distance of g,, the centroid of the after wedge, from F is equal to the moment of the after wedge about F divided by the volume of the wedge, and a similar for- mula applies to the forward wedge. If the moments of

the after and forward wedges are designated as ml and m,, respectively, then the distance

The moments of the volumes are obtained by inte- grating, forward and aft, the product of the section area at a distance x from F and the distance x, or

rn, =SOQ(y)(stan SO)(s)ds = tan 68 f? yds

L-Q

rn, = tan SO J,, s"ds

The integrals in the expressions for ml and m2 are recognized as giving the moment of inertia of an area about the axis corresponding to x = 0, a transverse axis through F, the centroid of the waterplane. Therefore, the sum of the two integrals is the longitudinal moment of inertia, I,, of the entire waterplane, about a transverse axis through its centroid. Then

m, + m, = v -g,g, = 11, tan SO

or v . QQ, IL=- tan 68

In the limit when S+ + 0

where 11, is the moment of inertia of the entire water- plane about a transverse axis through its centroid, or center of flotation.

Page 23: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

14 INTACT STAB1 llTY

The height of the longitudinal metacenter above the keel is given by an expression similar to equation (3) by replacing the transverse metacentric radius by the longitudinal metacentric radius.

- - - KML = KBL + K B

and

GML = KML - KG

3.5 Metacenter for Submerged Submarines. When a submarine is submerged, as noted in Section 1, the center of buoyancy is stationary with respect to the ship at any in- clination. It follows that the vertical through the center of buoyancy in the upright position will intersect the vertical through the center of buoyancy in any inclined position at the center of buoyancy, and the center of buoyancy is, there- fore, both the transverse and longitudinal metacenter.

To look at the situation from a different viewpoint, the =of a surfaced submarine is equal to m p l u s or plus IIV. As the ship submerges, the waterplane disappears, and the value of I, and hence is reduced to zero. The value of becomes plus zero, and B and M coincide.

The metacentric height of a submerged submarine is usually denoted rather than

3.6 Effects of Trim on the Metacenter. The discussion and formulas for and all assumed that the waterline at each station was the same; namely, no trim existed. In cases where substantial trim exists, values for -- BM, KM, and m w i l l be substantially different from those calculated for the zero trim situation. It is important to calculate metacentric values for trim for many ship types, and tables for various trims are often included in trim and stability books. The use of computers makes these tables less useful as the effects of trim are included directly in the computation of the righting arm by maintaining longitu- dinal moment equilibrium; thus, mis computed directly when needed. Section 4.4 includes the effects of trim in computing cross curves. Letcher (2009)) in describing the calculation of also discusses the effects of trim.

3.7 Applications of Metacentric Height 3.7.1 Moment to Heel 1 Degree. A convenient and

frequently used concept is the m o m e n t to heel 1 degree. This is the moment of the weight buoyancy couple, or WW when the ship is heeled to 1 degree, and is equiva- lent to the moment of external forces required to pro- duce a 1-degree heel. For a small angle, the righting arm is given by m sin 4 and, after this is substituted for we have:

Moment to heel 1 degree = A?%? sin(1deg) (5)

Within the range of inclinations where the metacenter is stationary, the change in the angle of heel produced by a given external moment can be found by dividing the moment by the moment to heel 1 degree.

3.7.2 Moment to T r i m 1 Degree. The same theory and formula apply to inclinations in the longitudinal di- rection, and we may say:

Moment to trim 1 degree = A m L sin(1 deg) (6)

where mL is the longitudinal metacentric height. We are more interested, however, in the changes in draft produced by a longitudinal moment than in the angle of trim. The expression is converted to moment to trim 1 cm by substituting 1 cm divided by the length of the ship in centimeters for sin 1 deg. The formula becomes, with metric ton units,

AGM I, MTcm = - t - m l0OL

where L is ship length in meters. As a practical matter, mL is usually so large compared to m t h a t only a negli- gible error would be introduced if mI, were substituted for GMI,. Then II,/V may be substituted f o r m , where IL is the moment of inertia of the waterplane about a transverse axis through its centroid, and A = pV, where p is density. Then, moment to trim 1 cm:

For fresh water, p = 1.0; for salt water, p = 1.025 (t/m3). Since the value of this function is independent of the position of G but depends only on the size and shape of the waterplane, it is usually calculated together with the displacement and other curves before the location of G is known. Although approximate, this expression may be used for calculations involving moderate trim with satisfactory accuracy for ships of normal proportions.

3.7.3 Period of Roll. The period of roll in still wa- ter, if not influenced by damping effects, is given by:

Period = constant>< k - CxB

E -JZ where k is the radius of gyration of the ship's mass about a fore and aft axis through its CG.

The factor "constant x k" is often replaced by C x B, where C is a constant obtained from observed data for different types of ships.

This formula may be used to estimate the period of roll when data for ships of the same type are available, if it is assumed that the radius of gyration is the same per- centage of the ship's beam in each case. For example, if a ship with a beam of 15.24 m and a m of 1.22 m has a period of roll of 10.5 seconds, then

If another ship of the same type has a beam of 13.72 m and a of 1.52 m, the estimated period of roll would be:

Page 24: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

The variation of the value of C for ships of different types is not large; a reasonably close estimate can be made if 0.80 is used for surface types and 0.67 is used for subma- rines. In almost all cases, values of C for conventional, homogeneously loaded surface ships are between 0.72 and 0.91. This formula is useful also for estimating rn when the period of roll has been observed.

A snappy, short period roll may be interpreted as in- dicating that a ship has moderate to high stability, while a sluggish, slow roll (long period) may be interpreted as an indication of lesser stability, or that other factors such as free surface or liquids in systems may be in- fluencing the roll period. However, the external rolling forces due to waves and wind and the effects of forward speed through the water tend to distort the relationship

of T =- CB . Hence, caution must be exercised in cal- E

culating m v a l u e s from periods of roll observed at sea, particularly for small and/or high-speed craft.

The case of the ore carrier is an interesting illustra- tion of the effect of weight distribution on the radius of gyration, and therefore on the value of C. The weight of the ore, which is several times that of the lightship, is concentrated fairly close to the CG, both vertically and transversely. When the ship is in ballast, the ballast wa- ter is carried in wing tanks at a considerable distance outboard of the CG, and the radius of gyration is greater than that for the loaded condition. This can result in a variation in the value of C from 0.69 for a particular ship in the loaded condition to 0.94 when the ship is in ballast. For most ships, however, there is only a minor change in the radius of gyration with the usual changes in loading.

If no other information is available, the metacentric height, in conjunction with freeboard, is a reasonably good measure of a ship's initial stability, although it must be used with judgment and caution. On ships with ample freeboard, the moment required to heel the ship to 20 degrees may be larger than 20 times the moment to heel 1 degree, but on ships with but little freeboard it may be considerably less. Little effort may be required to capsize a ship with large m but with small freeboard. When the metacentric height is zero or negative, certain types of ship would capsize, while other types might de- velop fairly large righting moments at the larger angles of heel. The metacentric height may be used, however, as an approximate index of stability for an undamaged ship with reasonable confidence if the ship can be com- pared to another with similar lines and freeboard for which the stability characteristics are known.

3.8 Conditions of Loading. A ship's stability, and hence may vary considerably during the course of a voyage or from one voyage to the next, and it is nec- essary during its design to determine which probable condition of loading is the least favorable and will there- fore govern the required stability. (The general effect of variations in cargo and liquid load during a ship's op-

eration is further discussed in Section 6). It is custom- ary to study, for each design, a number of loaded condi- tions with various quantities, locations, and densities of cargo and with various liquid loadings. When a ship is completed, the builder usually provides such informa- tion for the guidance of the operator in the form of a trim and stability booklet. Typical booklets contain a general arrangement of the ship, curves of form, capaci- ties and centers, and calculations of and trim for a number of representative conditions and blank forms for calculating new conditions. The information contained in such a booklet is required for all general cargo ships, tankers, and passenger ships by international conven- tions, including both the International Convention on Load Lines and International Convention for the Safety of Life at Sea (SOLAS) (IMO, 2006). Similar information is furnished for naval ships and mobile offshore drill- ing units where it is often referred to as the operating manual. An onboard loading computer is allowed as a supplement to the trim and stability booklet, but cannot replace it. Type approval requirements for loading com- puter software or systems vary internationally from none to explicit version approval.

The range of loading conditions that a ship might ex- perience varies with its type and the service in which it is engaged. Typical conditions usually included in the ship's trim and stability booklet are:

Full load departure condition, with full allowance of cargo and variable loads. All the ship's spaces are filled to normal capacity with load items intended to be car- ried in these spaces, which usually implies minimum density homogeneous cargoes, whether general, dry bulk, liquid, or containerized. A typical example is given in Table 3.

Naval combatant ships do not carry cargo in the usual sense. Instead, cargo equivalent variable load on such ships would be ammunition or fuel for onboard air- craft.

Additional conditions may be included for other heavier cargo densities, involving partially filled or empty holds or tanks. For ships that carry deck cargoes such as container ships and timber carriers, conditions with cargo on deck should be included, since they may be critical for stability. Some ships may have minimum draft requirements, which may include immersion of propulsors or minimum draft forward to limit slamming in heavy seas.

Partial load departure conditions, such as half car- go or no cargo. When no cargo is carried, solid or liquid ballast may be required, located so as to provide suffi- cient draft and satisfactory trim and stability.

Arrival or minimum operating conditions. These describe the ship after an extended period at sea and are usually the lowest stability conditions consistent with the liquid loading instructions (see Section 6.8). Certain cargo ships might be engaged in point-to-point service, while others might make many stops before returning

Page 25: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

16 INTACT STAB1 llTY

Table 3 Typical full load departure condition-post-Panamax con- tainership.

Item Weight VCG (m) LCG TCG FSMom (KN) (m-MS) (m-CL) (m-KN)

Lightship

Constant

Low Sulphur

Fuel Oil

Diesel Oil

Lube Oil

Fresh Water

SW Ballast

Misc.

In Hold

On Deck

Misc. Weights

Displacement 716,703 16.409 6.625A 0.157P 179,406

Stability Calculation Trim Calculation

KMt 16.409 m LCF draft 11.967 m

VCG 16.409 m LCB (even 6.025A m-MS keel)

GMt (Solid) 3.05 m LCF 11.414A m-MS

FSc 0.25 m MT 1 cm 11,470 m-KNIcm

GMt 2.799 m Trim 0.375 m-A (Corrected)

GM t 0.50 m List 3P degrees (Required)

GMt (Margin) 2.299 m Propeller 145 % immersion

to home port. The amount of cargo and consumables would vary, depending on the service. Conditions for naval ships would reflect the most adverse distribution of ammunition, along with reduced amounts of other consumables.

In all of the above conditions of loading, it is neces- sary to make appropriate allowances for the effects on stability of the free surface of liquids in tanks, as ex- plained in Section 5.

U.S. Coast Guard (USCG) stability requirements are given in the Code of Federal Regulations (2006).

3.9 Suitable Metacentric Height. The stability of a ship design, as evidenced approximately by its metacen- tric height ( m ) , should meet at least the following re- quirements in all conditions of loading anticipated:

It should be large enough in passenger ships to pre- vent capsizing or an excessive list in case of flooding a portion of the ship as a result of an accident. The effect of flooding is described in Tagg (2010).

It should be large enough to prevent listing to unpleas- ant or dangerous angles in case all passengers crowd to

one side. This may require considerable ?%? in light dis- placement ships, such as excursion steamers carrying large numbers of passengers.

It should be large enough to minimize the possibil- ity of a serious list under pressure from strong beam winds.

For passenger ships, the first bullet point is often the controlling consideration. The International Convention requirements for stability after damage, or other crite- ria for sufficient stability, may result in a metacentric height that is larger than that desirable from the stand- point of rolling at sea. Since the period of roll in still water varies inversely as the square root of the metacen- tric height, larger metacentric heights produce shorter periods of roll, resulting in greater acceleration forces which can become objectionable. The period of roll may also be a factor in determining the amplitude of roll, since the amplitude tends to increase as the period of roll approaches the period of encounter of the waves.

Of these two conflicting considerations, that of safety outweighs the possibility of uncomfortable rolling, and adequate stability for safety after damage must be pro- vided for passenger ships and is desirable for cargo ships. However, the metacentric height should not be permitted to exceed that required for adequate stability by more than a reasonable margin.

Numerous international and national maritime orga- nizations have established stability criteria which cover to some degree almost all types of ships, be they com- mercial or military. These are discussed further in Sec- tion 7.

Since the required stability will vary with displace- ment, it is convenient to express the required stability as a curve of required Wplo t ted against displacement or draft. Actual values for various loading conditions including corrections for free surface of liquids in tanks (Section 5) are compared to the required m. Condi- tions of loading that are unsatisfactory must avoided by issuing loading instructions that will prevent a ship from loading to an unsatisfactory stability condition. Required ?%? curves must be used with caution since analysis of the righting arm curve, which defines the stability at large angles, is the only rigorous method of evaluating adequacy of stability. The righting arm curve takes into account freeboard, range of stability, and the other features discussed in Section 4. Hence, stability criteria are usually based on righting arm curves, rather than on m alone. Further recent positions taken by na- tional authorities are increasingly requiring the direct evaluation of stability for the specific loading condition rather than a single criterion for a specific draft (see Section 7).

Navy ships must meet all the stability requirements of commercial ships, including the ability to operate safely in severe weather. In addition, they must have the ca- pability of withstanding considerable underwater hull damage as a result of weapons effects. For these rea-

Page 26: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 17

sons, navy ships may have larger initial ?%?than similar sized commercial ships.

An alternative approach is to make use of the "allow- able KG" curve, derived from the righting arm curves, which has the advantage that no stability calculations are necessary to judge the suitability of a loading condi- tion. Thus, it is more amenable to implementation as a criterion in load-planning software that does not have access to the hull geometry information.

While loading computer software can rapidly evalu- ate a potential loading condition against stability crite- ria, a useful tabulation (NAVSEA,1975), can be prepared for ships to permit a quick judgment as to whether a proposed weight change will generally be acceptable or unacceptable with regard to the limits on draft and stability. The most useful part of this is the gauge on sensitivity of the ship stability to weight changes. This tabulation is titled Ship Status for Proposed Weight Changes and takes on the following format:

1 Ship 1 Status 1 Allowable KG for Governing Loading Conditions

Status 1 means that the ship has adequate weight and stability margins with respect to these limits. Thus, a

reasonable weight change at any height is generally ac- ceptable.

Status 2 means that a ship is very close to both the limiting drafts and the stability (E) limits. Thus, any weight increase or rise in the CG is unacceptable.

Status 3 means that a ship is very close to the stabil- ity limit but has adequate weight margin. If a weight ad- dition is above the allowable m v a l u e and would thus cause a rise in the ship's CG, the addition of solid ballast low in the ship may be a reasonable form of compensa- tion.

Status 4 means that adequate stability margin exists but that the ship is operating at departure very close to its limiting drafts. Tankers and beach landing ships usu- ally fall into this category. A weight addition is at the expense of cargo deadweight, or else may adversely af- fect the ability of a landing ship to land at a designated beach site.

To reduce any necessary compromise between the requirements of a large amount of initial stability to withstand underwater hull damage and the desire to reduce ??%f to obtain more comfortable rolling char- acteristics, many large ships have antirolling tanks or fin stabilizers which operate to reduce roll ampli- tude. Antiroll tanks operate on the principle of active or passive shifting of liquids from side to side out of phase with the ship's rolling. The liquids may cause a free surface effect problem (discussed in Section 5) which must be taken into account when evaluating a ship's stability.

Curves of Stability

4.1 Righting Arm. To determine the moment of weight and buoyancy tending to restore the ship to the A upright position at large angles of heel, it is necessary to know the transverse distance between the weight vec- tor and the buoyancy vector. This distance is called the righting arm and is usually referred to as

An illustration of the ship drawn with waterlines in- clined at angles +, and +2 is shown in Fig. 18. The figure shows the initial upright center of buoyancy B, and new 1

centers of buoyancy B, and B,, corresponding to +, and +2, respectively. The corresponding righting arms are then GZ, and GZ,, computed for reference point 0. The reference point 0 would normally be the CG. Sometimes,

+ l a \I] cc' B1

righting arms are calculated for an assumed location of the reference point 0 (usually taken at the keel), and in A , this case they are often referred to as righting arms for D a poleheight of zero. Fig. 17 Transverse righting arms.

Page 27: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
Page 28: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
Page 29: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
Page 30: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
Page 31: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
Page 32: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
Page 33: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
Page 34: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
mgray
4-24
Page 35: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
mgray
4-25-1
mgray
4-25-2
Page 36: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
mgray
26-1
Page 37: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
mgray
27-1
Page 38: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
mgray
fig 36
Page 39: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of
Page 40: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

5 Effect of Free Liquids

5.1 Free-Surface Effect. The motion of the liquid in a tank that is partially full reduces a ship's stability be- cause, as the ship is inclined, the CG of the liquid shifts toward the low side. This causes the ship's CG to move toward the low side, reducing the righting arm. Ships with bulwarks that trap water on deck experience simi- lar reductions in the righting arm.

Prior to the widespread use of computers, direct evaluation of the free-surface effect on a normal ship required an excessive amount of calculation. In many cases, the tanks are not rectangular, and when formed by the shell and a longitudinal bulkhead, they are not symmetrical. With computer models of ships easily de- veloped, direct calculation of the free-surface effect is practical through computation of the fluid CG in the heeled (or trimmed) condition. However, approximate methods are useful when assessing the impact of free surfaces prior to performing computer-based calcula- tions or when they are not available as is the case early in the design process. Furthermore, while loading com- puter software can easily handle direct calculation of the shift in CG, some regulations explicitly require du- plication of approximate methods.

Consider a tank containing liquid in ship that is heeled to a small angle 4, as shown in Fig. 40.

As a result of the heel, some of the liquid will flow from the high side to the low side of the tank and a heel- ing moment will exist equal to the weight of the shifted liquid, w, multiplied by the lateral distance between the original and final position of the CG of w. If the tank is approximately wall-sided, the weight of liquid shifted

in an incremental length of tank is

the total weight, w, is given by the following expres- sion.

where b(x) is the width of the tank free surface at longi- tudinal position x.

The moment arm of this couple, b,, equals the lateral separation of the centroids of the two triangular wedge- shaped volumes representing the original and final loca-

Fig. 40 Fluid shift for a wall-sided tank

tions of w. In order for the triangular wedges to be of equal volume, the initial and final free-surface planes must intersect in a longitudinal line that passes through the centroid of the free-surface area. The heeling couple is then given by

Here, b(x) is the breadth of the tank free surface at loca- tion x along the length; i , is the moment of inertia of the free-surface area of the tank about a longitudinal axis through the centroid of that area.

This formulation points out that the mathematical processes applying to the motion of the CG of the liquid are similar to those applying to the motion of the center of buoyancy of a ship. At small angles of inclination, the liquid in each tank has a metacenter located at a dis- tance equal to i,lv above its CG in the upright position, where i, is the moment of inertia of the surface of the liquid about an axis through its centroid and parallel to the centerline, and v is the volume of the liquid. It is common practice to develop a tank table that shows the moment of inertia for varying quantities of liquid.

In evaluating the effect of free surface in a ship's tanks, the usual practice is to assume the most unfavor- able disposition of liquids likely to occur. If a tank is empty or completely full, there is no effect. The maxi- mum effect occurs when a tank is about half full. There- fore, it is customary to assume that the largest tank in each of the systems, or the largest pair of tanks if they are in pairs, is half full. This assumption is made even when a full-load condition is being studied, since a free surface will develop shortly after the ship leaves port. In the fuel-oil system, the settling tanks also are assumed to be half full. Fuel-oil tanks which are nominally full of fuel are considered to be either about 95% full in naval practice or about 98% full in the ease of a merchant ship, to allow for expansion and off-gassing of the oil, and will therefore have some free-surface effect. A ballasted fuel-oil tank or a nominally full water tank should be completely full and have no effect. If "split plant" op- eration is practiced, which involves dividing the system into two or more independent sections to enhance reli- ability in the event of damage, the largest tank or pair of tanks in each section is assumed to be half full.

5.2 Evaluation of Effect of Free Surface on Metacentric Height. For ships on which the effect of free liquid is

Page 41: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 3 1

relatively small, an effective metacentric height may be computed by considering the effect of free liquid as a virtual rise of the CG of the ship. By equation (17), it is seen that the free surface results in a heeling moment for small angles of heel. The net righting moment is then given by subtracting this heeling moment from the right- ing moment computed as though the liquid in the tank is "frozen,"

After substituting A = pgV in the denominator of the sec- ond term in parentheses, we obtain equation (18) for the net righting moment after correction for free surface,

This is seen to be an expression which is independent of the quantity of liquid in the tank, depending only on the moment of inertia of the free surface, the total ship buoyant volume, and the ratio of density of the tank liq- uid to the density of the water in which the ship is float- ing. Therefore, for any condition of loading, free surface may be evaluated for small angles of heel by adding the

values of bL for all tanks in which a free surface P v

exists. This summation represents a virtual rise in the ship's CG caused by the free-surface effect. This rise, called the free-surface correction, is added to KC;, the height of the ship's CG above the keel, resulting in an equivalent reduction in the metacentric height. The ef- fective GM, corrected for free surface, is then given by equation (19) as

It has been pointed out previously that the metacen- tric height, when multiplied by the displacement of the ship and the sine of the angle of inclination, provides a fairly accurate evaluation of the righting moment up to 7, or perhaps 10, degrees. The same is true of the metacentric height when reduced in the manner just described to take account of the effect of free liquid, provided that the surface of the liquid does not reach the top or bottom of the tank during this inclination. In cases where the tank is about half full, this treatment of free surface is accurate except in the case of a wide, shallow tank. When a tank is nearly full, this treatment would be accurate only in the ease of a narrow, deep tank where the surface of the liquid would not reach the top of the tank.

In making the loading calculations discussed in Sec- tion 3.8, free-surface corrections to m must be made on the basis of reasonable assumptions regarding the condition of all tanks. In the full load departure condi- tion, for example, it is customary to modify the condi-

tion to allow for free surfaces shortly after departure by assuming that service fuel tanks are half-full. The larg- est pair of storage fuel tanks are customarily assumed to be empty, but with maximum free surface effect act- ing (for a noncompensating system), and potable and reserve feed water are reduced to two-thirds full. For naval ships with a compensating system, the same as- sumptions are made except that 100% seawater ballast is carried in one pair of storage fuel tanks instead of assuming a pair of empty tanks. Ships with tank gaug- ing systems linked to loading computers can account for actual fill levels on a real time basis, but these are reasonable assumptions for design evaluation.

5.3 Evaluation of Effect of Free Surface on Righting Arm. The effect of free liquid in a wall-sided tank is to cause the CG of the liquid to shift through a certain dis- tance, d, parallel to the inclined waterline. If the weight of the liquid is w and the displacement of the ship A, the CG of the ship will move parallel to the inclined wa- terline through a distance (dw)lA, reducing the righting arm by that amount. The quantity (d - w) is known as the m o m e n t of transference. When free surface is present in a number of tanks, the summation of (d - w) for the various tanks, divided by the displacement of the ship, gives the total reduction in righting arm.

If the surface of the liquid has not reached the top or bottom of the tank and the tank has vertical walls, the distance d is equal to the distance from the CG of the liquid with the ship in the upright position to the meta- center of the liquid, which is equal to i,lv, multiplied by the sine of the angle of inclination for small angles. Therefore,

or since w = pgv

The moment of transference is thus seen to be inde- pendent of the quantity of liquid in the tank.

If the surface of the liquid has reached the top or bot- tom of the tank, the moment of transference will be re- duced and may be expressed by the product of p - i, and some factor less than sin +, or

The value of C depends on the degree of fullness, the ratio of depth to breadth of the tank, and the angle of inclination, each of which has some influence on the de- gree to which the motion of the liquid is suppressed.

Fig. 41 illustrates the method used in determining the factor C for a prismatic tank. Since the depth of the tank is 1.22 m, the CG of the liquid in the upright position (as used in the weight estimate) is on the centerline and 0.3 m above the bottom. At the 30 degree inclination, the CG of the liquid is 0.54 m above the bottom and 1.46 m from the centerline.

Page 42: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

32 INTACT STAB1 llTY

Fig. 41 Free surface effect at large angles

Considering the shift of the CG both parallel and nor- mal to the bottom, the horizontal component when in- clined is

1.46 . cos 30 deg + (0.54 - 0.3) . sin 30 deg = 1.38 m

The weight of the liquid is

1 -x6.1x1.22x 1 x p g = 3.72pg1 2

where 1 is the length of the tank and p is the density of the liquid.

The moment of transference is

1.388 x 3.72 - I - pg = 5.13 pgl

1 x pgx 1 x (6.1)" The value pg . i, = = 18.9 pgl , where i,

12 is the transverse moment of inertia. Thus, the ratio of the moment of transference to p . i , which is the factor

5.13pgl C, is - =0.27. This factor is independent of the

18.9pg1 length of the tank and the density of the liquid.

5.4 Top and Bottom Effects on Free Surface. Many ships have several wide double-bottom tanks contain- ing fuel oil, most of which will be 98% full. In such cases, if the effect of free liquid on metacentric height were to be evaluated by assuming that the CG of the liquid in each of the tanks were at its metacenter, a gross ex- aggeration of the loss of righting moment would be ob- tained at angles beyond 1 or 2 degrees due to the effects of topping off or bottoming out. The practice which has been adopted to produce a more reasonable value for the free-surface effect is to determine the effect of free liquid on the righting arm at an arbitrarily selected an- gle of 5 degrees and to translate the reduction in right- ing arm at 5 degrees into a reduction in metacentric height by dividing it by the sine of 5 degrees. The effect of this assumption is to produce a value of metacentric height, adjusted for free liquid, which, when multiplied by the weight of the ship and the sine of the angle of in- clination, will give, very nearly, the correct value for the righting moment at 5 degrees inclination.

To illustrate this, Fig. 42 shows the effect of topping off or bottoming out is shown for two rectangular tanks. One has a depth/breadth ratio of 0.2, while the other has a ratio of 0.8. The ratio of the moment of transference at each fill level for a 5 degree inclination to the maximum moment is plotted. For the shallow tank, the reduc- tion is significant for the top and bottom 15% fill levels, whereas for the deep tank the effect is concentrated in the top and bottom 2% or 3%.

Wide double-bottom tanks will show significant reductions in free-surface effect when nearly full or empty. Modern on-board loading computers that have direct linkage between the tank gauging system and the loading computer will take this into account.

For crude oil carriers, it is usually assumed that nom- inally full cargo tanks have a free surface due to boil-off of the volatile elements in the crude oil. Typically, these tanks are relatively deep, and credit for the reduction in free-surface effects is restricted to fill levels of 98% and above. The 2008 Intact Stability Code (IMO, 2008) requires that free-surface effects be considered when-

-

Relative Free-Surface Effect

Relative Filling Level

Fig. 42 Effect of tank depth to breadth ratio on relative free surface inertia.

mgray
fig 42
Page 43: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

Fig. 43 Wirefl-ame model for wing ballast tank in a containership.

ever the fill level is less than 98%. For nominally full cargo tanks, the GM correction for free surface is made at 98%, based on an inertia moment of liquid surface at 5 degrees and the righting lever corrected on the real shifting moment of cargo liquids.

5.5 Determination of Moment of Inertia of Free Sur- face. The process of determining the moment of inertia of the surface in a rectangular tank is relatively simple, since the moment of inertia about an axis through the centroid of the surface is given by

. 1b3 2 =- " 12

Many tanks, however, are trapezoidal in planview, while others may have one or more curved sides; in such cases, the process is more complex and generally performed us- ing a computer. In most cases for non-computer-based calculations, it will be sufficiently accurate to convert each tank shape to the best equivalent rectangular shape and then approximate the moment of inertia for the rect-

angular shape. This procedure is particularly practical when other calculations are approximate and the free- surface correction is used to provide an indication of the reduction in ?%!due to free surface when making an early determination of adequacy of stability.

In loading computer applications, the moment of trans- ference is often computed for several fluid levels at a small angle of heel, typically 5 degrees, and stored in the ship database. Then, a table lookup function can be used to include the free-surface correction for the actual fill level that is accurate for small heel angles associated with nor- mal loading conditions. The fluid density is normally set to 1 in this computation and the moment of transference corrected for the actual density at the time of application. This tabular approach does not require maintaining the tank geometry models in the ship database.

A sample of a complicated ballast tank in a container- ship is shown in Fig. 43, with the corresponding table of moment of transference stored in equivalent moment of inertia shown in Fig. 44.

Tank Table - SW WING 3F S

Fig. 44 Variation of free-surface effective moment of inertia with volume

mgray
fig 43
Page 44: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

34 INTACT STAB1 llTY

5.6 Numerical Example of Free Surface Calcula- tion. The following sample calculation is for a small ship, with tank layout and dimensions shown in Fig. 45.

5.7 Approximate Versus Exact Calculation of Free-Sur- face Effects. As noted earlier, the approximation to use a free-surface correction to the metacentric height works for small angles of heel. The following two figures show the effect of using a free-surface correction to w o n the righting arm curve. In Fig. 46, the righting arm curve for a 100 m x 20 m x 10 m barge with a four by two tank arrange- ment is shown for two methods of calculation. The tanks are 50% full and have a full slack free surface. Applying the free-surface correction computed using the 5 degree moment of transference calculation as a correction to the righting arm curve leads to the curve labeled "GZ" in Fig. 46, whereas the direct calculation of the shift in CG of the fluids leads to the curve labeled "GZ direct." Both curves are similar until about 30 degrees. After that, the direct calculation leads to a larger effect of the shift of fluid.

Diesel Oil Tanks

2 tanks

1 tank:

1 tank:

Fresh Water Tanks

2 tanks

However, in Fig. 47, the tanks are now 90% full. Shortly after heeling to 15 degrees, the tanks start to top-off, and the free surface is limited.

For normal operations, where heel angles are ex- pected to be small using a constant free-surface correc- tion to is accurate, but this will not be true if large heels are anticipated. Further, effects due to trim will not be accounted for.

5.8 Effect of Free Surface on Trim. Free liquid on a ship acts in the fore-and-aft direction in the same man- ner as in the transverse direction. For an intact ship with normal tankage, the effect of free liquid on trim is so small that it is often ignored. Its magnitude is small in comparison to the assumption in the formula for mo- ment to trim 1 em, described in Section 3, so that the CG is at the same height as the center of buoyancy. How- ever, for some vessels, such as some river barges with- out extensive transverse subdivision, free liquids may have an important effect on trim. Free-surface effects

4 3 i p = 30.8 m . $55 t/m = 26.3 m-t

x 1/12 [3.6 m x (1.9 mf]

i = 4.15 m4

i p = 4.15 m4 . 1.0 t/m3 = 4.15 m-t Free Surface Correction

Total ip = 26.3 + 4.15 = 30.5 m-t Displacement = 741t

i p = 30.5 m-t = 0.04m A 741 t

Fig. 45 Sample free-surface effect calculation

Page 45: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 35

I - GZd irect

Heel (degrees)

Fig. 46 Influence of direct calculations for 50% full (slack) tanks

also play a role in the evaluation of damage stability for double hull tankers subject to raking bottom dam- age (Tagg, 2010). There, the large trims associated with the damage can lead to significant free-surface effects in the cargo tanks. It is not unusual for this to make a difference in whether a double hull tanker meets the ex- isting damage stability criteria.

For a submerged submarine, since GM, and GMT are equal, the longitudinal free-surface effect may be more significant than the transverse effect since many tanks having free surface may be longer than they are wide. Under normal circumstances, however, the trim is carefully adjusted to zero in the submerged condition.

5.9 Effect of Longitudinal Subdivision on Free Sur- face. The subdivision of large tanks into two or more smaller tanks may be an effective method of improving stability by suppressing the motion of free liquids.

For a rectangular tank, the free surface varies as the cube of the width of the tank. Fig. 48 shows typical tanker designs with different levels of internal subdi- vision. When a centerline bulkhead is introduced into a double hull tanker design, the free-surface effect is reduced by a factor of four. That is, the combined free surface of the port and starboard tanks is one-fourth of the free surface of the single tank. The "three tank across" arrangement is typical of many small and mid- size single hull tankers. For a vessel with the propor-

Heel (degrees)

Fig. 47 Influence of direct calculation of free surface on 90% full tanks.

tions shown, the free-surface effect is about one-sev- enth of the "single tank across" arrangement.

Subdividing tanks in order to improve stability in- volves a compromise as it requires considerable in- crease in structure, piping, and fittings.

5.10 Effect of Double Hulls on Free Surface. Many barges and all new crude oil carrying tankers, as well as many other vessels, are now constructed with double hulls. This leads to the potential for a double free-sur- face effect. If there is significant liquid in the double hull tanks, they can exhibit a significant free surface at the same time that there is a free surface in the main tanks. Fig. 49 shows this effect for a combination of cargo tanks at a 65% fill level and "U" ballast tanks at 15% filling. The illustrated vessel would be unstable if loaded this way.

Early in the use of double hull tankers, there were instances of tankers lolling during cargo loading or off- loading operations. This is illustrated by the curve in Fig. 50 where the ?%?is initially negative but at about 12 degrees of heel, the GZ curve becomes positive. In Fig. 50, the GZ curve is plotted for both port and star- board heel directions. It shows a flat, near zero righting arm region between 12 degrees to port and 12 degrees to starboard. This indicates that it requires little heel- ing moment to make the vessel roll from one side to the other. This is characteristic of ships prone to lolling.

This behavior led to some concern about the overall intact stability of double hull tankers. A SNAME ad hoc

(Single Tank Across) (With Centerline Bulkhead) (Three Tanks Across) Free-Surface Effect = 1 Free-Surface Effect = 1 /4 Free-Surface Effect = 1;7

Fig. 48 Effect of longitudinal subdivision.

mgray
fig 46 and 47
mgray
fig 48
Page 46: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

Fig. 49 Section views of double hull vessel showing two free surfaces

panel investigated this issue (Moore, Neuman, & Pip- penger, 1996). While lolling was possible for some single tank-across tankers, inclusion of a centerline bulkhead in the cargo tanks prevented this problem.

On the other hand, barges have capsized due to water in the double hull space creating additional free surface, especially in conjunction with failure of the watertight centerline division in the double hull space.

Statical Stability Curve With Two Free Surfaces

Angle (deg)

Fig. 50 Negative (GM) stability leading to an angle of loll

5.11 Effect of Tank Fill Level on Free Surface. For nonrectangular tanks, the free-surface effect will vary with the level of the liquid in the tank. For instance, Fig. 51 shows a ballast "U" tank which is 35% full with the water level at one-half the double bottom height, and a tank in which the water level is increased so that the

ballast extends into the wings. The free-surface effect changes by a factor of three for the arrangement shown. Note that the moment of inertia for a U-tank is the mo- ment of inertia of the two separated free surfaces about their common centroid (i.e., the ship centerline), not the sum of the two inertias computed about their individual centers. Relatively small changes in ballast can have a dramatic effect on the overall stability.

5.12 Effect of Two Liquids. In some cases, a tank will contain two different liquids. For example, some mixed liquid cargos are not true solutions and may separate. Compensating tanks are known to have dif- ferent liquids in the same tank. Examples are gaso- line tanks in which seawater is introduced at the bottom as gasoline is drawn from the top to avoid the accumulation of explosive mixtures, submarine diesel oil tanks in which the oil is replaced by sea- water to preserve submerged equilibrium, and tanks on some diesel-driven surface type ships in which a compensating system is used to improve stability. In the latter case, the stability improvement results from maintaining low weight in the ship, reduction of free surface, and reduction in the effect of possible off- center flooding after damage.

Although these tanks are always completely full of liquid, a free-surface effect exists at the interface which will remain parallel to the waterline as the ship inclines. There will be a wedge on the low side in which the lighter liquid will be replaced by the heavier, and a wedge on the high side in which the heavier liquid will be replaced by the lighter. This effect may be evaluated by using the free-surface correction for the heavier liquid minus the correction for the lighter.

"U" Tank: 3 5% Full "U" Tank: 60% Full Free-Surface Effect = 1 Free-Surface Effect = 113

Fig. 51 Effect of fill level in a "U" tank.

Page 47: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 37

PLAN VIEW Cross-connection

y Anti-Roll Tmk ! /

I

CL

SECTION Fig. 52 ~ n t i d tanks.

5.13 Effect of Antiroll Tanks. One antiroll device is the antiroll tank. Such a tank may have a configuration in plan view, as shown in Fig. 52. The tank is usually filled to about 50% capacity and, thus, has a significant free-surface effect when the ship is not rolling. Prior to use of computers, the best calculation method was to treat the shift of liquid from one side to the other as a moment transference at each angle of heel and to make the appropriate correction to the righting arm curve. Now, direct evaluation of the shift in CG is possible. Free-surface effect on initial at zero heel should be calculated as though there are two separate tanks, each filled to 50% capacity if the cross connection is closed and a single large tank if it is open. This free-surface correction is added to the other free-surface correc- tions for tanks to obtain a ??@ (or KC;) corrected for free surface.

5.14 Bulk Dry Cargo. Bulk dry cargo, such as ore, coal, or grain, may redistribute itself if the ship rolls or heels to an inclination greater than the angle of repose of the substance carried (angle of repose is the angle between a horizontal plane and the cone slope obtained when bulk cargo is freely poured onto this plane). Thus, a ship may start a voyage with the upper surface of such a cargo horizontal and with the cargo evenly distributed throughout the space. But if the ship rolls sufficiently to cause a cargo shift, a list will result. A ship that has listed due to even a slight shift of cargo is susceptible to further rolling to increasing angles on the low side with further shifting of the cargo. Ships have been known to capsize from such progressive shifting of cargo.

Virtually all cargoes are directly influenced by the seaway-induced motions of the ship, which produce sig-

nificant angular and lateral accelerations. In a rapidly rolling ship, bulk cargoes may shift even when the maxi- mum angle of roll is less than the angle of repose be- cause of the dynamic effects of rolling. Calculations of motion dynamics show that the accelerations involved in rolling result in a greater likelihood of cargo shifting when the cargo is located above the ship's CG (as in the 'tween decks) rather than below (in the hold).

Some guidance in the design of ships to carry solid car- goes in bulk is provided by the IMO Code of Safe Practices for Solid Bulk Cargoes, 2004 (IMO, 2005), also known as the BC Code. Solid bulk cargoes, such as concentrates and coals, may contain high moisture and, when subjected to cyclic forces, pore water pressure can rise resulting in abrupt loss of shear strength (Green, 1980; Tanaka, 1990). To minimize the risk of this phenomenon (i.e., liquefac- tion), the BC Code introduces an upper bound of the mois- ture content of cargo called the transportable moisture limit (TML). The TML is defined as 90% of theflow mois- ture point, which depends on the characteristics of cargo and should be measured experimentally.

Grain has long been recognized as a dangerous cargo because of its tendency to flow or shift in the hold of a roll- ing ship. Grain freely poured into a compartment arranges itself into a pile of conical shape. The angle of the cone with horizontal varies with the specific variety of grain but may be as much as 30 degrees. This is the angle of repose. If the pile is static, the surface will remain undisturbed; if subjected to motions at sea, the grain surface could move in response to this motion, shifting the cargo CG. This mo- ment leading to heel of the ship is called the grain heel- ing moment. However, if the surface of the grain is lev- eled to a zero degree angle with horizontal, then the ship would have to roll in excess of the angle of repose before the grain would shift. To minimize the possibility that bulk grain will shift at sea, the International Grain Code (IMO, 1991) requires that the grain be trimmed (i.e., leveled after it has been loaded). The magnitude of the potential grain shift is dependent upon the amount of open space above the grain in which it can move. When a compartment is filled to the maximum extent possible, the adverse ef- fect of a grain shift will be less than if the compartment is partially filled. The Grain Code recognizes this in the determination of the grain heeling moment by assuming a 15 degree shift of grain when the compartment is full as opposed to a 25 degree shift when the compartment is par- tially filled. Once the grain heeling moment is calculated, the angle of heel due to grain shift and residual stability can be compared to regulatory requirements.

The grain heeling moment depends upon three fac- tors: the angle of shift, the internal geometry of the grain hold, and the weight of the grain. In practice, a vol- umetric heeling moment, measured in m" is computed for the various cargo compartments and the different ullages of grain in the compartments. For any loading condition, the volumetric heeling moment is then con- verted to a grain heeling moment by dividing by the stowage factor, measured in m3/t, to obtain the moment

mgray
fig 52
Page 48: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

38 INTACT STAB1 l lTY

due to the weight shift. The stowage factor utilized here has a different definition than is usually applied in com- mercial marine practice. Here, it is the volume per unit weight accounting for the interstices between the grain particles rather than a factor including the effects of broken stowage (i.e., the space left vacant when a com- partment is nominally filled). As the weight of grain that shifts is not constrained by spaces normally left vacant, this volumetric stowage factor is appropriate.

The IMO regulation, along with other useful informa- tion, is given in General Information on Grain Loading (National Cargo Bureau, 1994, Rev. 2002). In the United States, the National Cargo Bureau represents the USCG for review of grain-loading plans and safe bulk cargo stowage on U.S. flag vessels. This reference also pro- vides alternative means for achieving compliance with stability requirements including ballasting, overstowing of cargo, use of saucers, bundling of bulk grain, strap- ping or lashing, securement with wire meshes, tempo- rary longitudinal division, and self-trimming holds with sloping longitudinal bulkheads.

5.15 Suspended Cargo or Weight. In some cases, an item of cargo is suspended from a point above its CG. This method of stowage calls for special correction in the cal- culation of m A weight suspended from a boom is a simi- lar case and serves as a convenient explanatory example.

The CG of a weight suspended freely from a boom will remain vertically below the end of the boom, re- gardless of the list of the ship. The point of suspension, therefore, is the metacenter through which the weight acts. It makes no difference in the stability of the vessel whether the weight itself hangs high above the deck or not, provided the point of support remains the same. A suspended weight may be treated as though its CG was at the point of support. Therefore, if individual items of a full cargo, such as sides of beef, were suspended from several feet above each item's CG, the metacentric height of the vessel would be appreciably less than it would have been with an equal weight of unsuspended cargo at the same actual CG.

Effect of Changes in

6.1 Effect on Displacement and Center of Gravity. The effect of changes in weight on the ship's displacement and CG may be evaluated by the method used in the weight estimate (Section 2) and loading calculations (Section 3). This is a tabulation of items of weight with corresponding arms from the three reference planes, the resulting moments, and a summation of weight and moments. An item that is removed is entered as a nega- tive weight, and, if the arm is positive, the moment will be negative. If the arm is negative, as it may be in the case of port and starboard moments about the center- line plane, the sign of the moment will be opposite to that of the weight. If the present weight and moments of the ship are entered in the tabulation, followed by those of the items to be added and/or removed, the totals will represent the ship after the changes are accomplished.

The use of the tabular form (such as in typical spread- sheet software) is convenient when a number of items are involved. When only a single item already aboard ship is moved, the ship's CG will move in a direction parallel to the motion of the CG of the item moved. The shift GIG, parallel to any axis can be determined from the principle that

where w is the weight (mass = w/g) of the item moved and d is the distance moved.

A convenient method of finding the equilibrium angle following a weight shift is to treat the weight movement by adding a negative (upward) weight at the original position of the weight item and a positive (downward)

Weight on Stability

weight at the final location. The moment of the couple so formed is the heeling moment. If weight w is shifted transversely through a distance a and upward through a distance b, the resulting moment, at an angle of heel of 4 is given by w(a cos 4 + b sin 4). This expression, when divided by the ship displacement A, is the heeling arm varying with angle of heel. It may be plotted on the righting arm curve of GZversus 4 and the point of inter- section of the two curves gives the equilibrium angle of heel after the weight shift.

A problem that frequently arises is to find how much weight must be added in a given position to move the ship's CG a given distance. Consider the following nu- merical example. Assume that just enough ballast is to be added at a level 0.5 m above the keel to lower the CG of a 10,000 KN ship from 7 m above the keel to 6.5 m above the keel. Let the added ballast = w, KN.

I Displacement (KN) I VCG (m) I Vertical Moment (m-KN) I Original

Added Ballast

Since the required VCG = 6.5 m, the required vertical moment equation is:

(10,000 + W) . 6.5 = 70,000 + 0.5 - w

Therefore, w = 833 KN. 6.2 Effects on Stability. The effect of weight changes

on a ship's stability can be evaluated by recalculating the righting arms and the metacentric height for the re-

New condition 1 10,000 + w I

10,000

w

70,000 + 0 . 5 ~

7.0

0.5

70,000

0 . 5 ~

Page 49: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

38 INTACT STAB1 l lTY

due to the weight shift. The stowage factor utilized here has a different definition than is usually applied in com- mercial marine practice. Here, it is the volume per unit weight accounting for the interstices between the grain particles rather than a factor including the effects of broken stowage (i.e., the space left vacant when a com- partment is nominally filled). As the weight of grain that shifts is not constrained by spaces normally left vacant, this volumetric stowage factor is appropriate.

The IMO regulation, along with other useful informa- tion, is given in General Information on Grain Loading (National Cargo Bureau, 1994, Rev. 2002). In the United States, the National Cargo Bureau represents the USCG for review of grain-loading plans and safe bulk cargo stowage on U.S. flag vessels. This reference also pro- vides alternative means for achieving compliance with stability requirements including ballasting, overstowing of cargo, use of saucers, bundling of bulk grain, strap- ping or lashing, securement with wire meshes, tempo- rary longitudinal division, and self-trimming holds with sloping longitudinal bulkheads.

5.15 Suspended Cargo or Weight. In some cases, an item of cargo is suspended from a point above its CG. This method of stowage calls for special correction in the cal- culation of m A weight suspended from a boom is a simi- lar case and serves as a convenient explanatory example.

The CG of a weight suspended freely from a boom will remain vertically below the end of the boom, re- gardless of the list of the ship. The point of suspension, therefore, is the metacenter through which the weight acts. It makes no difference in the stability of the vessel whether the weight itself hangs high above the deck or not, provided the point of support remains the same. A suspended weight may be treated as though its CG was at the point of support. Therefore, if individual items of a full cargo, such as sides of beef, were suspended from several feet above each item's CG, the metacentric height of the vessel would be appreciably less than it would have been with an equal weight of unsuspended cargo at the same actual CG.

6 Effect of Changes in Weight on Stability

6.1 Effect on Displacement and Center of Gravity. The effect of changes in weight on the ship's displacement and CG may be evaluated by the method used in the weight estimate (Section 2) and loading calculations (Section 3). This is a tabulation of items of weight with corresponding arms from the three reference planes, the resulting moments, and a summation of weight and moments. An item that is removed is entered as a nega- tive weight, and, if the arm is positive, the moment will be negative. If the arm is negative, as it may be in the case of port and starboard moments about the center- line plane, the sign of the moment will be opposite to that of the weight. If the present weight and moments of the ship are entered in the tabulation, followed by those of the items to be added and/or removed, the totals will represent the ship after the changes are accomplished.

The use of the tabular form (such as in typical spread- sheet software) is convenient when a number of items are involved. When only a single item already aboard ship is moved, the ship's CG will move in a direction parallel to the motion of the CG of the item moved. The shift GIG, parallel to any axis can be determined from the principle that

A x GIG, = wd or GIG, = wdlA (20)

where w is the weight (mass = w/g) of the item moved and d is the distance moved.

A convenient method of finding the equilibrium angle following a weight shift is to treat the weight movement by adding a negative (upward) weight at the original position of the weight item and a positive (downward)

weight at the final location. The moment of the couple so formed is the heeling moment. If weight w is shifted transversely through a distance a and upward through a distance b, the resulting moment, at an angle of heel of 4 is given by w(a cos 4 + b sin 4). This expression, when divided by the ship displacement A, is the heeling arm varying with angle of heel. It may be plotted on the righting arm curve of GZversus 4 and the point of inter- section of the two curves gives the equilibrium angle of heel after the weight shift.

A problem that frequently arises is to find how much weight must be added in a given position to move the ship's CG a given distance. Consider the following nu- merical example. Assume that just enough ballast is to be added at a level 0.5 m above the keel to lower the CG of a 10,000 KN ship from 7 m above the keel to 6.5 m above the keel. Let the added ballast = w, KN.

1 New Condition 1 10,000 + w I 1 70,000 + 0 . 5 ~ I

Original

Added Ballast

Since the required VCG = 6.5 m, the required vertical moment equation is:

(10,000 + W) . 6.5 = 70,000 + 0.5 - w

Therefore, w = 833 KN. 6.2 Effects on Stability. The effect of weight changes

on a ship's stability can be evaluated by recalculating the righting arms and the metacentric height for the re-

Displacement (KN)

10,000

w

VCG (m)

7.0

0.5

Vertical Moment (m-KN)

70,000

0 . 5 ~

Page 50: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

vised values of displacement and vertical and transverse positions of the CG. The effect of any changes in free liquid should be included. A large change in the ship's displacement will have some effect on the change in righting arms and metacentric height. The free-surface effect will also change, even though there is no change in tankage, because this effect is the moment of the free liquid surface divided by the displacement.

In general, weights added above the ship's VCG will reduce stability, while weights added below the VCG will improve stability. Weights removed will have the op- posite effects. An exception to the above generalization would be those cases that involve large weight changes. The changes in righting arms due to the change in draft accompanying the weight addition (as read from the cross curves) may have effects opposite to the effects of the vertical location of the changes. Off-center weight changes that result in the CG being shifted from the cen- terline plane to a position off-center, port, or starboard will result in heel either to port or starboard. Section 3.7 illustrates the determination of small heel angles caused by moderate athwartship weight movements (moment to heel 1 degree).

Movement of a weight that is already aboard has no effect on displacement, and, at a given angle of heel, no effect on the center of buoyancy. The effect of such a weight movement on the righting moment at this angle is to move the ship's CG a direction parallel to the move- ment of the CG of the weight. The distance through which ship's CG moves depends only on the magnitude of the weight, the distance through which its CG moves, and the weight of the ship.

Thus, the effect on transverse stability when shifting an on-board weight depends on whether the shift results in a raising or lowering of the ship's VCG and whether or not the shift causes the upright ship's CG to move from the centerline plane to a position to either side of the centerline plane. A large weight shifted a moderate fore and aft distance, or a smaller weight shifted a large fore and aft distance, would be necessary to affect lon- gitudinal stability in the form of a change in trim. Longi- tudinal moment changes affect trim (produce a change in longitudinal inclination) in the same manner that athwartship moment changes affect heel.

6.3 Compensation for Initial Heel. Since the angle of heel of an undamaged ship may be caused either by an external moment, such as is produced by a beam wind, or by an off-center location of the CG, the measures to increase stability by means of weight changes would be different in these two cases:

6.3.1 Case 1. The CG is on the ship's centerline and the buoyant volume is symmetrical with respect to the ship's centerline; that is, the ship is undamaged. Since an external moment may act in either direction, any measure used should increase the righting moment in both directions and should be effective at small angles. The measures available, therefore, consist of adding weight with its CG on or near the ship's centerline be-

low the current CG, removing weight with its CG on or near the ship's centerline above the CG, or moving weight downward in a direction parallel to the ship's centerline. Weight removals or additions that increase the ship's waterplane area can also improve stability.

6.3.2 Case 2. The ship's CG has moved off the cen- terline as a result of (a) transverse shift of weight aboard ship, (b) the addition or removal of an off-center weight, or (c) the rise in the ship's VCG due to weight additions or weight removals that result in negative metacentric height. Corrective action should be taken after a deter- mination is made as to the cause of the initial heel. For example, in the ease of negative GM, the recommended action is to lower the ship's VCG rather than adding a mo- ment in a direction opposite to the heel, since in such a case the ship would heel to a larger angle to the other side and might even capsize. If the initial heel is caused by an off-center weight, then an appropriate action would be the addition of an opposing transverse moment by means of off-center weight additions or removals or athwartship shifts of on-board weights. If empty tankage were avail- able on the opposite side from the weight, adding liquids to such tanks on the opposite side might provide the off- setting athwartship moment. However, if adding liquids in empty low tanks is selected as the means to lower the ship's CG (and, thus, improve stability), the adverse effect of free surface in the process of filling the tank must be taken into consideration before a final decision is made. The large free-surface effect, reducing stability, often off- sets the stability gain in filling low, wide double-bottom tanks especially in the early stages of filling.

6.4 Large Trim Changes. If a change in weight results in a very large change in trim, the shape of the underwa- ter body may be quite different from that in the even keel attitude. The trim may be so large as to make the con- ventional displacement and other curves and the cross curves inapplicable, since they are based on an even keel condition. Older trim and stability books for ships that were expected to trim significantly often included tables of KMT for several trims. These tables are often included in loading computer software. An alternative approach is to reevaluate the hydrostatic properties di- rectly from the hull geometry for the trimmed waterline. This requires an iterative approach (discussed further in Section 8) to match the hydrostatic properties to the final resulting trim. Modern stability software will have the capability of computing the righting arms for a ship with initial trim and will take into account changes of trim that accompany heel.

6.5 Weight Changes in Submarines. Changes in weight on submarines are limited by the requirement that the weight and buoyancy must be very nearly equal in the submerged condition. Compensation for changes in the ship's fixed weight is usually obtained by an equal change in the lead ballast. This process may increase or decrease the righting moment in both the surfaced and submerged conditions, depending on the relative heights of the items added and removed. If it is not feasible to re-

Page 51: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 l lTY

move high weight and add ballast low in the ship, stabil- ity can be increased by adding submerged buoyancy and adding an equivalent weight of solid ballast at a lower level than the added buoyancy.

6.6 Effect of Cargo on Stability. In the typical cargo ship, the weight of the cargo may be twice the weight of the lightship; in a tanker, it may be more than four times the lightship weight. If the cargo is not homogeneous or does not fill the available space, a large variation in the CG of the loaded ship can result from varying the distribution of cargo. If a ship with a full load displace- ment of 10,000 tons has a metacentric height of 0.6 m when loaded with 5000 tons of cargo at the centroid of the holds, the metacentric height would be reduced to zero if the same weight of cargo were loaded with its CG 1.2 m above the centroid of the holds. On the other hand, the ship might be quite uncomfortable if the cargo were stowed with the heavy items in the lower holds.

The optimum height of the CG of the cargo may vary with the displacement and, therefore, with the total weight of the cargo to be loaded. A light cargo might be stowed with safety higher than a heavier load, since the lighter cargo would result in a greater freeboard. In many ships, there is little option in the disposition of cargo, in which case the stability characteristics can be predicted quite accurately.

A factor which should be considered, particularly in stowing ore cargoes, is the effect on radius of gyra- tion and, thus, on rolling characteristics. If most of the heavy cargo were stowed topsides or very low in the ship, rather than homogeneously, the effect on the ra- dius of gyration could be sufficient to affect rolling of the ship. Another factor affecting the stability of a large containership with several layers of topside stowage is the additional sail area exposed to beam winds.

6.7 Consumption of Liquids and Stores. During the course of a voyage, there may be a considerable varia- tion in displacement and in the position of the ship's CG resulting from the reduction in the consumable load. The major factor in this variation is usually the con- sumption of fuel and water, which are usually carried low in the ship and constitute a significant percentage of the ship's full load displacement. The consumable stores are usually a smaller item and are usually stowed fairly close to the vertical position of the ship's CG. In any case, the effect of consumables requires consider- ation in the evaluation of stability (see Section 3.8).

The low weight represented by the fuel can be re- placed by taking aboard seawater ballast. In older ships, this could have been either in the empty fuel tanks or in tanks provided for this purpose, usually designated as clean or segregated ballast tanks. The discharge of oily ballast water, even at sea, has become such a pollution hazard that current international convention standards require ships to be designed with segregated ballast tanks, thus obviating the need to add seawater ballast to empty fuel oil tanks. In any case, fuel contamination issues make this practice unattractive.

Oil tanker operators usually reserve a pair of deep tanks (identified as slop tanks) on their ships for stor- ing and chemically treating washings from the cargo oil tanks so that the effluent can be safely discharged over- board without polluting the waters, and the recovered oil can be reused as desired.

For a specific ship design, the designer should check with the governing regulatory organizations in order to make sure that all required criteria have been included in the calculations. USCG requirements are published regularly in the USCG Code of Federal Regulations (2006; 46 CFR 170-174; for small passenger vessels in part 178 and for fishing vessels in 46 CFR 28). Inter- national requirements are contained the International Code on Intact Stability, 2008 (IMO, 2008a).

The U.S. Navy has, in the past, taken the approach that in diesel-driven ships, ballasting of empty fuel tanks is less objectionable because of the greater differ- ence in density between the fuel and the seawater than in the case of Bunker C oil. Hence, a compensating sys- tem was developed for early naval diesel ships in which each bunker tank is always kept completely full of oil or water. This is accomplished by piping a group of tanks in series with a seawater ballast connection in the tank farthest from the machinery space. As oil is drawn from the top of the tank closest to the machinery space, sea- water enters the tank farthest away at the bottom. When this tank is full of seawater, the ballast passes through a pipe from the top of the ballasted tank to the bottom of the tank next in the series, and this process is continued through the successive tanks. With this system, there is an increase in the weight of liquid in the fuel system as the fuel is consumed. The rate of increase may exceed the rate of consumption of stores and water, resulting in a slight increase in the ship's displacement during a voyage, along with a lowering of the ship's CG. Environ- mental concerns have reduced the use of this practice.

Consumption of liquids and stores on diesel submarines has very little effect on stability. Nearly all fuel tanks are of the compensating type. Continuous compensation is ob- tained for variation in weight of the variable load by adjust- ing the quantity of seawater in tanks provided for this pur- pose in order to maintain a balance of weight and buoyancy when the ship is submerged. This process involves only mi- nor changes in the vertical position of the CG.

However, because of the previously mentioned pol- lution problems and resulting antipollution regulations, new and future naval surface ship designs usually adopt the use of clean ballast tanks to achieve the necessary stability, immersion, and trim. Meanwhile, for existing surface ships and submarines with compensating sys- tems, better warning devices, controls, and other equip- ment have been installed to prevent spillage of oily bal- last or fuel overflows during refueling. There is not much of a problem on nuclear submarines, which carry little fuel oil. While the fuel oil tanks are of compensating type, refueling is usually done in port where facilities are avail- able to receive the dirty ballast water.

Page 52: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

7 Evaluation

The main aim of stability criteria is to enhance safety against capsizing. However, stability evaluation plays a role in assessing operational effectiveness as well as ul- timate survival. While the discussion of stability factors in this chapter is largely limited to static or quasistatic factors, the reality is that intact stability evaluation rep- resents an evaluation of a capability to resist dynamic events. Human factors are also an important part of stability criteria. Both training standards and the con- sequence of failure play a role in the establishment of stability standards. Further, to assure absolute safety by built-in features is unfeasible: no ship can be built that cannot be capsized by negligent decisions or faulty operation.

Numerous criteria exist, and some are reviewed be- low. These include international commercial criteria codified via the IMO (2008), U.S. commercial and U.S. Navy criteria, and an introduction to newly developed probabilistic standards. In general, these current stan- dards are quite old and have been largely unchanged for decades. The criteria reflect partly the ships of their times and the calculation methods available. Caution should be exercised when using computer-based evalu- ation of righting arm curves to minimally meet existing standards. Further, the IMO is currently pursuing devel- opment of new generation intact stability criteria using state-of-the-art methods. In addition, in 2008, SNAME established a Dynamic Stability Task Group, which is intended to be a long-range, cross-disciplinary study of dynamic stability for vessels that operate in significant sea conditions.

Meanwhile, it is incumbent on the designer to ensure that the ship as designed meets all the applicable na- tional and international stability requirements of the country in which the ship is to be certified. The designer should make sure that all anticipated operations of the ship are investigated for stability, even though not re- quired for certification. A general skepticism of stability standards is encouraged. This is in part because these standards sometimes take a one-size-fits-all approach that may not be fully appropriate to the particular ship design, especially one that incorporates unusual geom- etry, proportions, arrangements, or mode of operation. Finally, it is important that guidance regarding the sta- bility characteristics of the ship be communicated to the master effectively through clear presentation in sta- bility booklets. The IMO has developed a "Model Load- ing and Stability Manual" (IMO, 1999).

7.1 Evaluation of Stability Based Upon GM and GZ Curves. As explained in Section 3.9, stability is often evaluated approximately based on metacentric height alone, without the benefit of a complete righting arm curve. This is equivalent to assuming that the righting

of Stability

arm curve has the form of a sine curve because it is as- sumed that the righting arm is equal to m s i n &. For ships with large freeboard and the type of form that pro- duces a righting arm curve concave upward near the or- igin, this practice is usually safe but may result in under- estimation of the ship's stability. For ships having little freeboard and the type of form that results in a righting arm curve concave downward near the origin, this prac- tice may not be acceptable because it does not assure an adequate range of positive stability or adequate residual dynamic stability (defined below).

This section will deal mainly with criteria based on consideration of actual shape and other characteristics of the curves of righting and heeling moment (or arm) for an undamaged ship through large angles of heel. One of the most important criteria in the evaluation of stabil- ity, a ship's ability to survive flooding due to damage, is discussed in Tagg (2010). So far as the intact ship is con- cerned, the righting arm curve for the least favorable condition of loading and the heeling arm curves for the various upsetting forces provide useful data for judg- ing the adequacy of the ship's stability. Features of the curves that warrant consideration from a purely static viewpoint are:

The angles of steady heel under the influence of a static heeling moment, as indicated by point A in Fig. 33.

The range of positive stability (point B in Fig. 33). The relative magnitudes of the heeling arm and the

maximum righting arm.

The angle of steady heel is important from two stand- points: first, its absolute value determines its adverse effect on personnel and the operation of the ship. Also, its value with respect to the angle of heel at which the deck edge will be submerged is a measure of the ship's resistance to capsizing, since the righting arm increases at a lesser rate after the deck edge is awash.

The range of residual positive stability is important since it limits of the angle to which the ship can heel without capsizing.

The excess of maximum righting arm over the heel- ing arm, in addition to providing a margin for the up- setting forces of wind and wave, is essential as an al- lowance for inaccuracies in calculating the heeling and righting arms.

Some stability criteria have been based only on such static considerations (Wendel, 1960), but most also in- clude the work and energy considerations (dynamic stability) discussed in Sections 4.9 and 4.10. In par- ticular, comparisons are made of the areas under the two curves up to certain angles and sometimes of the residual dynamic stability, which represents the work required, in addition to the effect of the heeling moment

Page 53: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

42 INTACT STAB1 l lTY

being considered, to capsize the ship. Such additional energy might be supplied by wave action, ship rolling, or by wind pressure in eases where upsetting forces other than wind are being considered. The residual dynamic stability is represented by the area below the righting arm curve and above the heeling arm curve beyond the angle of static equilibrium (i.e., between points A and B in Fig. 33).

The optimum criterion would allow full flexibility of op- eration while maintaining enough stability to avoid capsiz- ing in the roughest seas. To date, there is no such criterion applicable to all sizes of ships with differently shaped hulls in all of the varying types of marine operations.

In addition to the different methods of statical evalu- ation outlined in previous sections, it must always be borne in mind that a floating object leads a dynamic life, reacting continuously to the forces and moments im- posed by the wind and sea. In addition to meeting pub- lished regulations, the ship designer should review the available criteria and use those that fit each particular design best. It may be necessary to combine the best facets of several criteria for a specific design.

Various international and national organizations have established criteria for determining adequacy of intact stability to withstand specific upsetting forces. These organizations include:

IMO National Authorities (e.g., USCG, Transport Canada

Marine Safety) Navies International Association of Classification Societies

(IACS) and member classification societies

7.2 Merchant Ship Stability Criteria. The development of intact stability criteria for merchant ships has been slow due to the great variety of geometric forms, range of loadings, regional and national differences in type of ship used for various commercial enterprises, rapid development of new marine transport systems, and the naturally slow process of international agreement.

Since 1966, stability evaluation satisfactory to each nation has been required by the International Conven- tion Load Lines. Thus, while almost every nation has adopted stability criteria, each is formed upon the type of ships normally in use.

The IMO has adopted mandatory and recommended criteria on intact stability for ships of various sizes, dynamically supported craft, mobile offshore drilling units, offshore supply ships, and fishing vessels, all of which refer to intact stability or include intact stability standards (IMO, 2008). In each code and recommenda- tion, both GM and stability curve evaluations are uti- lized. Many national codes, such as the 2008 IS Code, IACS, etc., make these recommendations mandatory.

Current U.S. stability standards for U.S. flag merchant ships are contained in the USCG Code of Federal Regu- lations (2006). These standards include intact stability criteria for:

46 CFR 170-General Weather (steady wind heeling): GM (see Section

4.11) General criteria for vessels less than 100 meters

other than tugboats: RA curve 46 CFR 171-Passenger ships

Passenger heel: w Wind heel: w Sailing ships (monohull): RA curve Sailing catamaran: moment limit

46 CFR 172-Bulk cargoes Barges with hazardous cargo: RA curve, trans-

verse and longitudinal GM 46 CFR 173-Special use ships

Lifting: Counter ballast, RA curve Towing: w a n d RA curve

46 CFR 174-Specific vessel types Barges with deck cargo: RA curve Offshore drilling units: RA curve Towboats: RA curve

46 CFR 28-Commercial fishing vessels Lifting: RA curve Water on deck: RA curve Intact righting energy: RA curve Severe wind and roll: RA curve

Finally, the IMO is currently developing new gen- eration intact stability criteria that may focus on per- formance-based criteria. Some aspects of this are dis- cussed in Section 7.9.

The ship designer should be aware that these codes and rules are continuously evolving and should be fa- miliar with the latest versions that are applicable to the design under consideration.

7.3 International Maritime Organization Intact Stabil- ity Criteria Applicable to All Ships. The term "stability" in ship stability criteria does not have the same mean- ing as in theoretical mechanics. It is more appropriately represented by "boundedness" of a relevant motion, usually rolling (Francescutto, 1992, 1993). The IMO in- tact stability provisions have developed over time, start- ing in 1968 with IMO Resolution A.167 and are now col- lected in two key publications: the International Code on Intact Stability, 2008 (2008 IS Code, IMO (2008)) that revised and restructured the previous "Code on Intact Stability for All Types of Ships Covered by IMO Instru- ments" as amended in 1999 (IMO, 1999) and MSC Cir- cular 707 regarding guidance to the ship master (IMO, 1995). Among various criteria developed in the past, mainly for small craft, the one by Rahola (1939) is of particular interest because it formed the basis for most of the subsequent work by IMO.

The general outcome of these criteria is typically in the form of a limiting curve for GM or KG as a function of ship draft. Comparisons have been made for families of ships of same typology between statistical and weather criteria requirements (Boccadamo, Cassella, Russo Krauss, & Scarnardella, 1994; Francescutto, 1990), gen-

Page 54: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 43

erally finding that the second one is more severe. Com- parisons have also been made between A.562, A.167, and SOLAS790 for particular types of ships, like the modern large passenger cruise ships. In this case, it was found that the weather criterion is exceedingly more severe than SOLAS'9O (Francescutto & Serra, 2001; Frances- cutto, Serra, & Scarpa, 2001; IMO 2003).

In contrast to the previous IMO Code on Intact Stabil- ity, the 2008 IS Code contains criteria that are manda- tory for new ships that must meet either SOLAS 1974 or the 1988 Load Lines Protocol. The mandatory nature of the criteria is intended to reduce stability-related ship casualties. The 2008 IS Code only addressed criteria that were already included in the previous code. De- velopment of new criteria as well as revision of criteria contained in the 2008 IS Code is underway at IMO and is discussed further in Section 7.9.

Z3.1 IMO Resolution A1 67 1968, ("General Crite- rion'y, Res. No. 749 (18) 3.1 (1995, amended 1999), and Section 2.2 i n the 2008 IS Code. This criterion, urged by a recommendation of the 1960 SOLAS Conference, originates from the studies of Rahola, who developed an empirical criterion that specified required righting arms at several angles, a minimum value of the maximum angle, and a minimum area under the curve of righting arms (dynamical arm) up to a specified maximum angle on the basis of data on capsize casualties of 30 small ships. This empirical approach was based upon stability characteristics alone and did not attempt to specify the heeling moments that might be expected from the vari- ous hazards encountered in service.

The current implementation of this regulation in- cludes the following criteria:

1. The area under the righting arm curve should be not less than 0.055 m-radians up to 4 = 30 degrees, where 4 is the heeling or inclining angle (degrees).

2. The area under the righting arm curve should be not less than 0.09 m-radians up to 6 = 40 degrees or up to an angle where the nonweathertight openings come under water (whichever is less).

3. The area under the righting arm curve should be not less than 0.03 m-radians between the angles of heel 4 = 30 to 4 = 40 degrees or such lesser angle mentioned under Standard 2.

4. The righting arm should be at least 0.2 m at an an- gle of heel equal to or greater than 30 degrees.

5. The maximum righting arm should occur at an an- gle of heel exceeding 30 degrees.

6. The initial metacentric height should be not less than 0.15 m.

This general criterion practically gives rules so that the righting arm in calm water stays "sufficiently high" in a "sufficiently large range of angles" to provide a safe ship. It cannot be reasonably improved since it contains no relationship between demand and capacity as the regulation includes no physical modeling and no men- tion of sea state (Francescutto, 2002). Further, the level

I Heel Angle(deg)

Fig. 53 IMO Statistical Criterion GZ curve areas.

of safety is unknown for current ships and is one of the considerations included in the development work dis- cussed further in Section 7.9.

The Explanatory Notes to the International Code on Intact Stability (IMO, 2008a) provides background into the origin of the criteria regarding righting lever curve properties and shows that two methods were used in the reassessment of the criteria in 1985: the first was identical to that used by Rahola and the second used a discrimination analysis. In this work, it becomes clear that some of the ships that "passed" the criteria actually were involved in a capsizing. The Explanatory Notes also contain guidance for the application of the 2008 IS Code with respect to ships typically of wide beam and small depth that may have a difficulty meeting the crite- rion for the angle of heel at which the maximum righting lever occurs as contained in the code.

In spite of the fact that the original IMO resolution (Res. A.l67[ES IV]) was addressed to ships less than 100 m in length, the general practice of many shipbuilders and national authorities to accept it as criteria for ships greater than 100 m in length meant that the 100 m limita- tion was removed when the criteria were included in the 1995 Code on Intact Stability.

Z3.2 IMO Resolution A.562 ("Weather Criterion'y, 1985, now Res. No. 749 (18) 3.2 (1995, amended 1999) and Section 2.3 i n the 2008 IS Code. This recommen- dation originated as an answer to a recommendation given in the conclusions of SOLAS774: "[IMO] recom- mends that steps be taken to formulate improved inter- national standards on intact stability of ships taking into account, amongst other things, external forces affecting ships in a seaway which may lead to capsizing or to unac- ceptable angles of heel." As such, the weather criterion is based on physical considerations and assumes that a ship with complete loss of power is turned by the wind into a beam seas position where it experiences heavy resonant roll caused by wave and wind gust action. Even though it relies on a significant amount of empirical information, the weather criterion can be regarded as semiempirical, based on a simplified physical model.

mgray
53
Page 55: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 l lTY

Fig. 54 IMO Weather Criterion GZ curve areas

The IMO version of the weather criterion follows the work of Watanabe (1938) and Yamagata (1959) in Ja- pan. The ship is assumed to obtain a stationary angle of heel, QO, due to side wind loading represented by a lever, I,,,,, which is not dependent on the heel angle and is a result of a 26 mls wind. The ship is assumed to perform resonant rolling around this angle due to side wind loading. As a result of this rolling, it momentarily reaches a maximum windward roll of 4,. As at this posi- tion the ship is most vulnerable in terms of excitations from the weather side, it is further assumed that it is acted upon by a gust wind represented by arm, lW2 = 1.5 lW1. This is translated into an increase of the wind ve-

locity, assumed to affect the ship for a short period of time but at least equal to one half the natural roll period. The choice of wind corresponds to an extreme storm condition. The stability requirement is that should the ship roll freely from its windward roll angle, +,, start- ing with zero angular velocity, the limiting angle, &, to the lee-side should not be exceeded during the ensuing half-cycle. This limiting angle is the lesser of three, the downflooding angle where significant openings are sub- merged, the vanishing angle, @,, or the angle of 50 de- grees (an explicit safety limit). When expressed as an energy balance, the work done by the wind excitation as the ship rolls from the windward side to the leeward side should not exceed the potential energy at the limit- ing angle, &. Area "b" must be greater than area "a."

For details of the criterion, the reader is referred to the 2008 IS Code and the associated explanatory notes (IMO, 2008a). The essential components include a wind heeling arm developed from the projected lateral area of the ship and deck above the waterline and the wind- ward angle of roll that includes factors: k for roll damp- ing (including the effect of bilge keels), XI and X, for hull form (Beam/Depth and C,), r for the effective wave slope coefficient, and s for wave steepness based upon the ship natural roll period. The formula for the wind- ward roll angle, in degrees, is:

0, = i o 9 l c x , x 2 ~ s (21)

This criterion is simple to use, is based on a physi- cal (although rough) modeling (Kuo and Welaya, 1981),

and can be improved and "updated" to some extent for new ship typologies. On the other hand, the simplified modeling takes into account only capsizing in beam waves and wind, while no internal degree of freedom is introduced (shifting of cargo, water on deck, etc.). The level of safety is to a large extent unknown, although some studies (Dudziak and Buczkowski, 1978; Umeda & Ikeda, 1994; Umeda, Ikeda, & Suzuki, 1992) try to quan- tify it, resulting in quite different values, also depend- ing on ship typologies. In principle, the level of safety (probability of noncapsizing) is "computable" once a mission profile is assumed (Iskandar, Umeda, & Hama- moto, 2001) or in general terms connected with the ex- pected operational life of the ship in the environment assigned by the criterion rules. The application of the Weather Criterion to ships with large values of both KG/ Draft and Beam/Draft, like the modern large passenger cruise ships, leads to more stringent requirements than the application of current subdivision and stability rules for damaged ships (Bertaglia, Serra, & Francescutto, 2003). It was for this reason that the IMO issued "In- terim Guidelines for the Alternative Assessment of the Weather Criterion" (2006a) that provide procedures for the determination of the wind heeling lever by means of direct measurements and for the experimental deter- mination of the angle of roll +,. IMO also developed ex- planatory notes (IMO, 2007) for these interim guidelines to provide an example of how they can be applied.

There are several criticisms of this stability crite- rion, which, on the other hand, has proven to be quite effective in improving the safety of navigation. They are connected with the mathematical modeling of the roll motion equation and the effects of bilge keels, the evaluation of roll period which is based on a regression formula, the overestimation of the wave steepness s at large roll periods due to an extension of the range of applicability beyond the original limits, the overestima- tion of the effect of waves through the effective wave slope coefficient r, and possibly the overestimation of the wind and wind gusts effects.

The IMO development of new generation intact sta- bility criteria may recommend changes to update the Weather Criterion and to introduce a new philosophy based on performance standards.

Z3.3 Revised Guidance to the Master to Avoid Dan- gerous Situations i n Adverse Weather and Sea Con- ditions (MSC Circular 1228) (IMO, 2OO7a) -Super- cedes MSC Circular 707. While the stability curve is a good indication of a ship's stability and its resistance to capsizing, certain ship types have been known to cap- size under certain wave and following sea conditions even though their stability curves indicated good sta- bility. Considerable experimental work has been done in this field, as reported by Oakley, Paulling, and Wood (1974), Amy, Johnson, and Miller (1976), and De Kat, Brouwer, McTaggart, and Thomas (1994). Such work provides plausible explanations of the mechanisms that cause unexpected capsizing. This includes the special

Page 56: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

cases of capsizing in following seas (Oakley, Paulling, & Wood, 1974) and severe rolling in head seas (France et al., 2003). New developments can be expected in the future, with IMO playing an important role in the evolu- tion of stability criteria toward greater safety for both large and small craft.

The MSC Circular 1228 addresses the operational indications to avoid dangerous phenomena occurring in longitudinallquartering and head waves. Stability is not directly addressed but safety in the mentioned environ- mental conditions is the primary concern.

The guidance aims at giving seafarers caution on dangerous phenomena that they may encounter during navigation in following, quartering, and head seas and providing the basis for a decision on ship handling in order to avoid such dangerous situations. It provides ad- vice on safe and unsafe combinations of ship speed and course relative to waves in a simplified form of a polar diagram. However, the guidance was designed to ac- commodate all types of merchant ships and is therefore of a general nature; it may be too restrictive for ships with favorable stability characteristics and too gener- ous for certain other ships.

When sailing in severe following, quartering, or head seas, a ship is likely to encounter various kinds of dan- gerous phenomena which may lead to capsizing or other hazardous conditions. The sensitivity of a ship to dan- gerous phenomena will depend on the actual stability parameters, hull geometry, ship size, and ship speed. This implies that the vulnerability to capsizing and its probability of occurrence in a particular sea state may differ for each ship.

The period with which a ship traveling in following and quartering waves encounters the waves becomes longer than in head or bow waves, and principal dangers caused in such situations are as follows:

Surf-r iding and broaching-to. When a ship is situ- ated on a steep forefront of high wave in following and quartering sea condition, the ship can be accelerated to ride on the wave; this is known as surf-riding. When a ship is surf-ridden, the so-called broaching-to phenom- enon may occur, which puts the ship in danger of capsiz- ing as the result of sudden change of ship's heading and unexpected large heeling.

Reduction of intact stability caused b y r id ing o n the wave crest at midship. When a ship is riding on the wave crest, the intact stability will be decreased sub- stantially depending on the ship form. The amount of stability reduction is nearly proportional to the wave height, and the ship may lose its stability when the wave length is one to two times of ship length and wave height is large. This situation is especially dangerous in follow- ing and quartering seas because the duration of riding on wave crest (i.e., the time of diminished stability) be- comes longer.

Synchronous rolling motion. Large rolling motions may be excited when the natural rolling period of a ship

coincides with the encounter wave period. In case of navigation in following and quartering seas, this may happen when the transverse stability of the ship is mar- ginal, and therefore, the natural roll period becomes longer.

Parametric rolling motion. Unstable and large am- plitude roll motion will take place if the encounter wave period is approximately one half the natural roll period of the ship. In order for this type of rolling to occur in head and bow seas, the ship normally must operate at a relatively low speed for the critical wave encounter pe- riod to occur. But steering the ship head into the seas at low speed is often a standard heavy weather operat- ing procedure. The infrequent occurrence of parametric roll in such circumstances can probably be attributed to other factors such as a ship form and appendages con- ducive to large roll damping and small wave-induced stability variations. In following and quartering seas, this can occur particularly when the initial metacentric height is small and the natural roll period is very long.

Combination of various dangerous phenomena. The dynamic behavior of a ship in following and quartering seas is very complex. Ship motion is three-dimensional and various detrimental factors or dangerous phenom- ena, such as additional heeling moment due to deck sub- mergence, shipping, and entrapment of water on deck or cargo shift due to large roll motions, may occur in com- bination with the above-mentioned phenomena simulta- neously or in sequence. This could create an extremely dangerous combination which may cause ship capsize.

7.4 US. Navy Criteria. U.S. Navy criteria for all sizes of ships are based on efforts to make a physical assess- ment of the heeling moments expected due to various specific hazards, in conjunction with the curve of right- ing moments.

Z4.l General. U.S. Navy criteria are intended to ensure the adequacy of stability of all types and sizes of naval ships, as evidenced by sufficient righting energy to withstand various types of upsetting or heeling mo- ments. The fundamental energy relationships have been discussed in Section 4.10, and this subsection deals with specific criteria and factors used by the U.S. Navy (NSWCCD, 2008; Sarchin & Goldberg, 1962; Goldberg & Tucker, 1975). The U.S. Navy criteria are of value for designers of commercial as well as naval ships, for al- though the limits were established through the particu- lar experience of the U.S. Navy, the basic principles ap- ply to all ships. These criteria have formed the basis for those of many other navies (Hayes, 2002; MOD, 1999)

It is first necessary to establish the loading condi- tions for which the ship will be expected to withstand the upsetting forces (Section 3.8 discusses conditions of loading). It is important to note at this point that stabil- ity righting arm curves should be prepared for all con- ditions and corrected for free surface, as discussed in Section 5.3. The curve having the least area is then se- lected for use. Each heeling arm curve is superimposed

Page 57: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

46 INTACT STAB1 llTY

on this graph so that the resulting plot has the appear- ance of Fig. 33.

The various types of upsetting moments considered by the U.S. Navy are discussed in the following, begin- ning with the special case of beam winds.

Z4.2 Beam Winds Combined w i th Rolling 7.4.2.1 BEAM WINDS AND ROLLING. Beam winds and

rolling are considered simultaneously, since a fairly rough sea is to be expected when winds of high velocity exist. If the water were still, the ship would require only sufficient righting moment to overcome the heeling mo- ment produced by the action of the wind on the ship's "sail area.'' When the probability of wave action is taken into account, an additional allowance of dynamic stabil- ity is required to absorb the energy imparted to the ship by the rolling motion.

7.4.2.2 WIND VELOCITIES. The wind velocity which an intact ship is expected to withstand depends upon its service. Specific wind velocities to be assumed should be obtained from the appropriate regulatory body that governs the ship design, if available. If not, U.S. Navy values may be used, as given in Table 4.

7.4.2.3 WIND HEELING MOMENT. A general formula that is used to describe the unit pressure on a ship due to beam winds is as follows in SI units:

Table 4 Wind velocities assumed by the U.S. Navy.

Service Minimum Wind Velocity for Design Purposes

(in knots)

1. Ocean

(a>

(b)

2. Coastwise

(a>

(b)

(c>

3. Harbor

Ships which must be expected to 100 weather the full force of tropical cyclones. This includes all ships which will move with the amphibious and striking forces.

Ships which will be expected 80 to avoid centers of tropical disturbances.

Ships which must be expected to 100 weather the full force of tropical cyclones.

Ships which will be expected 60 to avoid centers of tropical disturbances, but to stay at sea under all other circumstances of weather.

Ships which will be recalled to 60 protect anchorage if winds of over Force 8 are expected.

60

where C = dimensionless coefficient for ship type pa = air density (mass per unit volume) V,, = wind velocity

There is considerable uncertainty regarding the value of C. Similarly, the variation of wind velocity at different heights above the waterline is not universally agreed upon.

The most widely used expression for P, in English units (lb force per sq ft), is P = 0.004 - V,2 (where V is in knots) or P = 0.0195 - V,2/1000 in tonnes force per square meter (Vin knots). Heeling arm H.A. due to wind is,

0.0195 .V;AI cos'" H.A. =

lOOOA

where (see Fig. 55): A = projected sail area, m2 1 = arm (vertical distance) from center of lateral re-

sistance of underwater hull (usually assumed at half draft) to centroid of hull and superstructure lateral area, m

V,,, = wind velocity, nominal (knots) 4 = angle of inclination A = displacement in metric tons, (t)

It is recognized that as the ship heels to large angles, the use of A . I cos2 4 is not rigorous since the exposed area varies with heel and is not a cosine function. How- ever, other effects are also ignored and the above for- mula may be used to obtain gross effects. Wind tunnel tests at the David W. Taylor Naval Ship Research and Development Center on models representing different ship types and superstructure forms have indicated that an average coefficient of 0.0035 rather than 0.004 (0.017 vice 0.0195 for metric units) should be used in the fore- going formula, which assumes a constant wind gradient. In order to account for actual full-scale velocity gradi- ent effects, an average coefficient value of 0.004 in con- junction with the wind velocity gradient curve is used by the U.S. Navy. This curve is a composite of various values described in the literature. The nominal velocity

I

WIND, V w I

Lateral Area !

I e - -

Center of ! DRAFT Lateral 1 I

Resistance ! . -. -. -. -. -. -. .

Fig. 55 Heeling effect of wind

Page 58: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

Wind Pressure vs. Height (100 Knot Wind at 10 m)

Height (m)

Fig. 56 Wind pressure for nominal 100 knot wind (at 10 m above waterline).

is assumed to occur at about hl-@ = 10 m (33 ft) above the waterline. The U.S. Navy assumes the wind gradient is proportional to (h/h,f)1'7.5, where h is the height above the waterline. Tabulated values for the heeling moment have been prepared for a nominal 100 knot wind as an aid in determining wind heeling moments in DDS 079-1 (NWSCCD, 2008). Data from this table is well repre- sented (R2 = 0.9978) by a logarithmic expression:

where h is the height above the waterline, m, and P is the pressure in tonnes/m2.

7.4.2.4 ADEQUATE STABILITY. The criteria for adequate stability when encountering adverse wind and wave con- ditions are based on a comparison of the righting arm and heeling arm curves, as shown in Fig. 57 (NWSCCD, 2008).

Stability is considered satisfactory if:

The heeling arm at the intersection of the righting arm and heeling-arm curves (Point C) is not greater than six tenths of the maximum righting arm; and

Area A, is not less than 1.4 A,, where area A, extends 25 degrees or 4, (if roll angle is determined from model tests) to windward from Point C.

ANGLE OF NCLINATION DEG

+ LL 4

Fig. 57 U.S. Navy criterion of stability in wind and waves.

ui r

The foregoing criteria for adequate stability with re- spect to adverse wind and sea conditions are based on the following considerations:

CURVE A INTACT RlGETlNG ARM CURVE CURVE B HEELING ARM CURVE I

A wind heeling arm in excess of the ship's righting arm would cause the ship to capsize in calm water. The requirement that the heeling arm be not greater than six tenths of the maximum righting arm is intended to pro- vide a margin for gusts and for inaccuracies resulting from the approximate nature of the heeling arm calcula- tions.

In the second criterion, the ship is assumed to be heeled over by the wind to Point C and rolling 25 de- grees or 4,. from this point to windward; the 25 degrees being an arbitrary but reasonable roll amplitude for heavy wind and sea conditions. Area A, is a measure of the energy imparted to the ship by the wind and the ship's righting arm in returning to point C. The margin of 40% in A, is intended to take account of gusts and for calculation inaccuracies. Energy losses mentioned in Section 4.10 are ignored, and it is assumed that no downflooding occurs.

Upsetting moments caused by lifting weights over- side, personnel crowding, and high-speed turning are presented, followed by U.S. Navy combined criteria for all three.

Z4.3 Lifting of Heavy Weights Over the Side 7.4.3.1 EFFECT OF LIFTING WEIGHTS. Lifting of weights

will be a governing factor in required stability only on small ships which are used to lift heavy items over the side. Lifting of weights has a double effect upon trans- verse stability. First, the added weight, which acts at the upper end of the boom, will raise the ship's CG and thereby reduce the righting arm. The second effect will be the heel caused by the transverse moment when lift- ing over the side.

7.4.3.2 HEELING ARMS. For the purpose of applying the criteria, the ship's righting arm curve is modified by

Page 59: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

48 INTACT STAB1 llTY

correcting VCG and displacement to show the effect of the added weight, assumed to be at the end of the boom. The heeling arm curve is calculated by the formula:

Heeling arm = wa cos 4

W where

w = weight of lift a = transverse distance from center1

boom W = displacement, including weight of 4 = angle of inclination, degrees

ine to end of

lift

Z4.4 Crowding of Personnel to One Side 7.4.4.1 EFFECT OF CROWDING OF PERSONNEL. The move-

ment of personnel will have an important effect only on smaller ships that carry a large number of personnel. The concentration of personnel on one side of a small ship can produce a heeling moment which results in a significant reduction in residual dynamic stability.

7.4.4.2 HEELING ARMS. The heeling arm produced by the transverse movement of personnel is calculated by Eq. (26):

wacos4 Heeling arm = -

W where

w = weight of personnel a = distance from centerline of ship to center of

gravity of personnel W = displacement 4 = angle of inclination, degrees

In determining the heeling moment produced by the personnel, it is assumed that all personnel have moved to one side as far as possible. Each person occupies -0.2 m (2 sq ft) of deck space.

Z4.5 High-speed Turning 7.4.5.1 HEELING ARMS PRODUCED BY TURNING. The cen-

trifugal force acting on a ship during a turn may be ex- pressed by the formula (English or metric):

WV2 Centrifugal force = -

gR (27)

The arm in conjunction with this force to obtain the heel- ing moment is the vertical distance between the ship's CG and the center of lateral resistance of the underwa- ter body. This arm is assumed to vary as the cosine of the angle of inclination. The center of lateral resistance is taken vertically at the half draft.

If the centrifugal force is multiplied by the arm and divided by the ship's displacement, an expression for heeling arm is obtained.

v2a Heeling arm = -cos 4

gR (28)

a = distance between ship's CG and center of lateral resistance (half draft) with ship upright

4 = angle of inclination, degrees

For all practical purposes, R may be assumed to be one half of the tactical diameter. If the tactical diameter is not available from model or full-scale data, an estimate is made.

7.4.5.2 CRITERIA FOR ADEQUATE STABILITY. The U.S. Navy criteria for adequate stability for lifting weights, personnel crowding, and high-speed turning are based on a comparison of the righting arm and heeling arm curves (see Fig. 58).

Stability is considered satisfactory if:

The angle of heel, as indicated by point C, does not exceed 15 degrees;

The heeling arm at the intersection of the righting arm and heeling arm curves (point C) is not more than six tenths of the maximum righting arm; and

The reserve of dynamic stability (shaded area) is not less than four tenths of the total area under the right- ing-arm curve.

The criteria for adequate stability are based on the following considerations:

Angles of heel in excess of 15 degrees will interfere with operations aboard the ship and adversely affect safety and comfort of personnel.

The requirements that the heeling arm be not more than six tenths of the maximum righting arm and that the reserve of dynamic stability be not less than four tenths of the total area under the righting arm curve are intended to provide a margin against capsizing. This margin allows for possible overloading and for possible inaccuracies resulting from the empirical nature of the heeling arm calculations.

7.5 Topside Icing. Icing can occur due to four me- teorological effects: freezing sea spray, freezing of par- tially melted snow, freezing rain, and freezing of fog. It is difficult to estimate an upper limit for accumulation of ice. Once ice has started to form, it will continue to accumulate under unfavorable conditions and the only

0, ANGLE OF INCLINATION, DEG

Fig. 58 U.S. Navy criteria of stability for weights overside, personnel crowding. where

Page 60: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 49

recourse is to institute ice removal measures or leave the area. High winds are likely to occur during periods of icing, and it is appropriate to consider combined icing and wind effects.

A ship of destroyer size, which is capable of withstand- ing a 100 knot beam wind without ice, can withstand a beam wind of only 80 knots with an ice accumulation of 200 tons. A cruiser type in service, which can with- stand a 90 knot beam wind without ice, can withstand a beam wind of only 78 knots with an accumulation of 600 tons of ice. The foregoing ice weights correspond roughly to a 15 em coating on horizontal and vertical surfaces where ice would build up. An actual build-up of ice would be nonuniform, but the ice weights deter- mined on the basis a uniform 15 cm coating may be used in estimating maximum beam wind velocity for which the stability criterion will be met. For destroyer sizes and above, the criteria will be met for a 70 knot wind in combination with topside icing. For smaller ships, topside icing results in a more significant reduction in righting arms and the allowable beam wind velocity is accordingly less. For example, a 59 m patrol ship, which can meet the wind criterion for a 75 knot beam wind without ice, will have to avoid beam winds in excess of 50 knots if there has been substantial ice accumulation. In the ease of a smaller mine sweeper of 46 m, 50 tons of topside ice reduces the maximum righting arm from 0.4 m to about 0.2 m with a reduction in range from 90 to 56 degrees. The maximum allowable wind is reduced from 85 to about 40 knots.

The design approach to topside icing is to determine the maximum allowable beam winds combined with ic- ing for a ship whose stability has been established from other governing criteria. The design would be consid- ered satisfactory if the allowable wind at time of icing was in excess of winds that are likely to be encountered in the intended service.

U.S. (NAVSEA, 1975) and U.K. naval standards (MOD, 1999) assume 150 mm of ice with a density of 950 kg/& on all exposed decks, horizontal platforms, and roofs. It is derived from traditional approaches where the mass and VCG of ice actually experienced, on all exposed vertical and horizontal surfaces, has been equated to a representative thickness solely on horizontal surfaces. Under this approach, ice accretion on vertical surfaces should not be included as this is inherent in the assump- tion of 150 mm of ice.

The Torremolinos Convention for the Safety of Fish- ing Vessels (IMO, 1977) contains recommendations of minimum requirements for icing of fishing vessels with specific guidelines as to amounts of ice accumulation to be assumed. This is reproduced in the 2008 IS Code as well as the Voluntary Guidelines for the Design and Construction of Fishing Vessels between 12 and 24 m in length. In each case, consideration of the increased windage area due to icing is recommended.

The U.K. Maritime and Coastguard Agency full icing allowance assumes all exposed horizontal surfaces are

subject to an ice weight of 30 kg/m2, and that the vertical surfaces corresponding to the lateral area of one side of the vessel are subject to a weight of 15 kg/m%ith this weight acting on centerline. This approach is also taken in the High Speed Craft Code (IMO, 2000) and U.S. fish- ing vessel regulations (USCG, 2009) where the same ice load is effectively applied.

Lower icing standards are applicable for areas where icing is not expected to be so severe or the possibility of seeking shelter is higher, such as coastal waters. Ar- eas of applicability are included in the Code on Intact Stability. Icing is to be included in the loading condi- tions for which the stability of the vessel is evaluated. In general, the icing assumed in stability calculations will almost always be specially considered by the regulatory administrations involved, and thus, the ice loads should be developed in agreement with them.

The University of Alaska has published comparative icing charts developed from experience in the Gulf of Alaska and the Aleutian Islands which show ice ac- cumulation as a time dependent phenomenon (Wise & Comiskey, 1980).

7.6 Stability Criteria for Certain Ship Types Z6.1 Fishing Vessels. There are several particular

features of the design and operation of fishing vessels that can result in stability problems. Among these are col- lection of water on the deck from large waves and severe motions, free surface in the fish holds or tanks, and the ef- fects of deck loads and nets suspended from a boom. The training, experience, and attention of the personnel who operate such vessels are likely to be slanted towards fish- ing rather than safe vessel operation. There are records of fishing vessel casualties that are directly attributable to such unwise practices as stowing a very heavy catch of fish on deck that reduces freeboard to dangerously low levels and, at the same time, reduces GM as well.

For vessels 24 m (79 feet) or longer, the primary means for determining the adequacy of a fishing ves- sel's intact stability is evaluating the characteristics of its static righting arm curve. The principal stability cri- teria are contained in the IMO 1993 Torremolinos Proto- col (IMO, 1995a), which is based on the work by Rahola. Various countries have adopted versions of this protocol for their own use. In the United States, the stability re- quirements are included in 46 CFR subpart 28.500, and include effects of water on deck, stuck gear, lifting, and a righting energy criterion. In general, the modifications to the IMO version are the addition of a minimum range of positive stability, typically 60 degrees or more (Wom- ack, 2002). IMO's Voluntary Guidelines for the Design, Construction, and Equipment of Small Fishing Vessels (IMO, 2005) between 12 and 24 m in length provide rec- ommended criteria for vessels in this size range. IMO is developing recommendations for smaller decked fishing vessels (length less than 12 m) and for undecked fishing vessels of any length.

Womack (2002) presented a critique of the application of this static, one-size-fits-all approach to the diverse

Page 61: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

50 INTACT STAB1 l lTY

fleet of small commercial fishing vessels in a very com- plex dynamic environment. For example, the smaller the boat, the more significant the sea conditions are. This is shown by contemplating the effect of 6 m seas on a 300 m tanker, a 45 m trawler, or a 15 m offshore lobster boat. The 6 m seas are no concern for the tanker, mini- mal concern for the trawler, and significant concern for the lobster boat. The existing stability criteria do not reflect this conflict due to scalability problems with the Torremolinos area criteria and the lack of true dynamic analysis methods. Womack (2002) goes on to outline the steps required for a satisfactory stability analysis and evaluation and the equally important presentation of the stability guidance and stability concepts to the crews.

Z6.2 Towboats. Towboats may also be prone to sea-motion-related capsizing. In addition, these ship types are characteristically designed with low free- board, which enhances the danger of taking on sea wa- ter through topside openings. Other hazards frequently experienced by tugs are the towline forces generated by the tug's own propeller thrust, called self-tripping, and by the movement of the ship being towed, called tow-tripping. Towboats must meet the general stabil- ity criteria, such as the USCG stability criteria (CFR 174.145); however, the heeling arm developed below will usually dominate.

7.6.2.1 HEELING ARM. The formula for calculating the transverse heeling arm curve for tow-line pull, used by the USCG (CFR 173.095), is as follows (metric units):

Heeling Arm = 2 N ( P ~ D ) ~ / ~ x s x h x c o s ~

13.93A (29)

where: N = number of propellers P = shaft horsepower per shaft, kilowatts D = propeller diameter, m s = effective fraction of propeller slip stream de-

flected by the rudder, assumed to be that frac- tion of the propeller circle cylinder which would be intercepted by the rudder if turned to 45 de- grees from the vessel's centerline

h = vertical distance from propeller shaft centerline at rudder to towing bitts, m

A = displacement, t qb = angle of inclination

7.6.2.2 CRITERIA FOR ADEQUATE STABILITY. The U. S. Navy criteria for adequate stability are based on the an- gle of heel and a comparison of the ship's righting arm and the heeling arm curve (see Fig. 58). Stability is con- sidered satisfactory if:

The angle of heel, as indicated by point C, does not ex- ceed the angle at which unrestricted downflooding may occur, or 40 degrees, whichever is less. The limit on range is to provide a margin of safety in the event a watertight door or vent duct is open and could be a pathway for seri- ous downflooding due to wave and heel action.

The heeling arm at the interception of the righting arm and heeling arm curves (point C) is not more than six tenths of the maximum righting arm.

The reserve of dynamic stability (shaded area) is not less than four tenths of the total area under the righting arm curve. The USCG criterion is similar to the navy criterion except that it requires the reserve of dynamic stability to be 0.61 m-degrees (0.01065 m-radians).

7.6.3 BULK CARRIERS CARRYING GRAIN. Once the grain heeling moment has been computed (see Section 5.14), the IMO (1991) and national regulations require that the angle of heel due to the shift of grain be less than the lesser of 12 degrees or the angle at which the deck edge immerses. The grain heeling moment is then applied as a heeling arm curve for comparison to the righting arm curve. The residual area between the two curves (as limited by 40 degrees, the angle of maximum difference between the righting and heeling righting arms, or angle of flooding) must be not less than 0.075 m-radians. Fi- nally, the mincluding free-surface effects shall not be less than 0.30 m. Recognizing the importance of angle of repose and trimming the cargo, the vessel is required to be upright before proceeding to sea.

7.7 Evaluation of Mobile Offshore Drilling Units. Cri- teria for stability evaluation of mobile offshore drill- ing units (MODUs) are included in the 2008 IS Code (IMO, 2008) and updated in classification society rules for construction (American Bureau of Shipping, 2008). Semisubmersible drilling platforms obtain static stabil- ity from surface-piercing columns that connect their submerged flotation bodies to the above-water platform. Analysis of their ability to withstand the upsetting forces of winds and waves (under varying loading conditions) is similar to the type of analysis made for conventional ship forms. However, the wind heeling moments applied include the form drag of the various components of the structure, such as the drilling derrick, and the increased exposure of superstructure deck as the MODU heels. In addition, the overturning moment must be considered for all axes of heel (see Section 4.12).

Stability must be evaluated for all modes of operation of the vessel, including transit, operating, and severe storm conditions. Alternative criteria that take into ac- count the environmental conditions, dynamic response (through wind-tunnel and wave-tank tests or nonlinear simulation, as appropriate), the potential for flooding, the susceptibility for capsizing, and adequate margin for un- certainties may be allowed. The wind speeds associated with the MODU severe storm criteria are 100 knot winds. Hurricanes in the Gulf of Mexico in 2005, where winds far in excess of 100 knots were experienced (NOAA, 2005), led to the loss of several MODUs indicating that sufficient margins may not be included in these criteria.

7.8 Evaluation of Stability of Submarines. The forego- ing principles apply to a surfaced submarine as well as to surface ships. There are some peculiarities of subma- rines, however, which should be mentioned.

Page 62: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 5 1

The form of the hull of a submarine is such that the righting arms in the surfaced condition are positive at angles well beyond 90 degrees, a condition that is sel- dom found in surface ships other than sailing yachts and self-righting rescue boats. The only significant heeling moment to which a surfaced submarine is sub- jected results from wind and wave action. Unlike sur- face ships, all topside openings can be closed to prevent shipping of water during heavy rolling except for the old diesel-powered submarines that must operate with open diesel engine manifolds. Consequently, capsizing of an intact submarine is extremely unlikely. The ma- jor stability problem is rolling to very large angles with adverse effects on personnel and the operation of the ship.

The righting arm curve for a submerged submarine is equal to the metacentric height, m, multiplied by the sine of the angle of inclination. Its maximum value, therefore, occurs at 90 degrees. Except for the minor ef- fect of shifting of liquids and loose items in the ship, the range of positive stability would be 180 degrees. A submarine is subjected to only minor heeling moments when submerged. Therefore, there is no danger of cap- sizing an intact submerged submarine, provided the metacentric height has at least a small positive value.

During the period while a submarine is submerging or surfacing, its transverse stability is less than when either surfaced or submerged because of the free liquid in the main ballast tanks. On the surface, there is only a small free-surface effect in the main ballast tanks, caused by the small quantity of residual water that can- not be blown whose surface remains above the tops of the flood openings. When the submarine is submerged, there is no free surface in the main ballast tanks be- cause they are completely full.

An approximate evaluation of stability during sub- merging and surfacing can be made by a series of calcu- lations of displacement, height of the CG of the ship, and the free-surface effect, assuming that the main ballast tanks are filled to successively greater depths. The only variables in these calculations are the weight, VCG, and vertical moment of free surface of the water in the main ballast tanks. The effect of the water in the main bal- last tanks at each assumed level is added to the weight, vertical moment, and vertical moment of free surface of the ship in the surfaced condition, after the vertical moment of free surface of the residual water has been deducted.

The results of these calculations, consisting of the dis- placement and height of the CG of the ship, adjusted for free-surface effect, are plotted in Fig. 59, together with the height of the metacenter; the minimum metacentric height is determined as the smallest vertical distance between the two curves. Stability is satisfactory if the metacentric height has a small positive value because the nature of the righting arm curve during submergence is such that positive values will be developed at small an- gles of heel when the metacentric height is zero.

I

0 A W W 0 W & X

CENTER OF

\, Q CO

CENTER OF GRAVITY #

DISPLACEMENT

Fig. 59 Submarine stability while submerging

The height of the metacenter drops as displacement is increasing from its value in the surfaced condition, shown to the left of Fig. 59, until it meets the curve of the height of the center of buoyancy. The vertical sepa- ration of these two curves, B- is equal to IJV, which has been reduced to zero as the ship submerges, owing to the disappearance of the waterplane when the hull is submerged.

The assumption in these calculations that all main ballast tanks are filled to the same waterline is some- what unrealistic because the actual levels in the vari- ous tanks depend on the area of the flood openings, the shape of the individual tanks, and the depths to which the openings are submerged. The flood openings are sized to flood the forward tanks faster than the after tanks to produce a down angle on the submarine and expedite submerging. In addition, any rolling of the sub- marine will increase the depth to which the tanks on the low side are submerged, causing them to fill faster than those on the high side.

When the main ballast tanks are arranged in pairs, the moment of inertia of the individual port and star- board tanks is used, rather than the moment of inertia of the pair considered as a single tank because there is no flow from one side to the other.

7.9 Review of the International Maritime Organiza- tion Intact Stability Code. Several problems have been identified in the existing procedures for stability assess- ment, especially for vessels that deviate from the ship forms upon which the standards have been based. Op- erational practice and experience are ways to judge the effectiveness of current stability criteria. Fortunately, accidents that are clearly related to a failure of a ship's intact stability are very rare. The ability to investigate the origin of such accidents is often severely hampered by the depth of water in which the lost ship is located. Those accidents that do avail themselves to full investi- gation are often associated with several failure events,

Page 63: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 l lTY

including internal flooding that impairs a ship's intact stability.

Further, several investigations have shown that the level of safety associated with the 2008 IS Code criteria is not consistent (Dudziak & Buczkowski, 1978; Rakhm- anin, 1986; Umeda & Ikeda, 1994; Umeda, Ikeda, & Su- zuki, 1992; Kobylinski, 1993; Kruger, Hinrichs, & Cramer, 2004; IMO, 2004a; Van Daalen, Boonstra, & Blok, 2005; De Kat, Van Walree, & Ratcliffe, 2006). For example, in Van Daalen, Boonstra, and Blok (2005), the probability of capsizing for a containership is presented for which a complete route analysis was computed using simulated ship motions. This work highlights the importance of op- erator input since the results indicate a very high prob- ability of capsize without operator input and a moderate to low probability with appropriate operator input. The lowest Wpresented in the analysis is the actual stability limit according to the code. While some vessels would not operate at this limit due to damage stability require- ments, this analysis indicates that the code does not rep- resent a unique safety level because low GM can lead to low probability of capsize and vice versa.

In response to concerns such as the above, the IMO Subcommittee on Stability, Load Lines, and Fishing Ves- sels Stability (SLF) initiated a review of the Intact Sta- bility Code in 2002, using a two-phased approach: short- term and long-term. The short-term phase involved the restructuring of the code into mandatory and recom- mended parts, resulting in the 2008 IS Code and also the revision of MSC Circular 707 (see Section 7.3.3), result- ing in MSC Circular 1228. The long-term effort, which is titled "Development of new generation intact stability criteria," is focused on establishing minimum require- ments for ship design, applicable to unconventional types of ships and major dynamic modes of stability. The SLF group refined the scope of the long-term effort into a framework, which includes:

Distinction of intact stability failures into two types: total stability failure, or capsizing, and partial stability failure, which would generally be the occurrence of a large amplitude roll angle or excessive accelerations that would impair the ship's normal operation.

Definition of the types of criteria into deterministic and probabilistic and into parametric and performance- based.

Identification of the major dynamic modes to be ad- dressed in the new generation intact stability criteria development:

Righting arm variation in waves associated with problems such as parametric excitation and pure loss of stability,

Stability under a "deadship" condition, and Maneuvering related problems in waves such as

broaching-to.

Description of vulnerability criteria that would iden- tifv the susce~tibilitv of a s h i ~ to different modes of

stability failures and that would be based on simplified models, analytical solutions, or statistical data. These criteria would likely require separate development for different failure modes.

Agreement on the meaning of "safety level" and re- quiring its evaluation for any proposed new generation intact stability criteria.

The framework also discusses the form and use of performance-based criteria. The chief benefit of perfor- mance-based criterion is that it provides a physically robust solution for a particular stability failure mode and can be formulated to use procedures such as model tests, numerical simulations, and analytical solutions or a combination of these methods. If formulated as deter- ministic criteria, the performance-based criteria would result in a "pass-fail" output. If formulated as probabi- listic criteria, the result would be expressed in terms of probability of failure during a specified time or as an average rate of failures. In each case, to quantify the safety level, a performance-based assessment would be required. In principle, two formulations of this assess- ment are possible: one is a short-term formulation in which the assessment would be performed for each of a list of assumed situations (environmental conditions and loading conditions) and the other is a long-term for- mulation in which the safety level is evaluated for a se- ries of assumed operational scenarios, and the analysis would take into consideration the probabilities of these assumed situations.

The group agreed that the development of probabilis- tic performance-based criteria requires consideration of two key issues: time dependence and the problem of rarity. The influence of time on such criteria is clear from the fact that an increase in the exposure time of a ship in adverse wind and wave conditions increases the probability of a stability failure. The problem of rarity arises when the average time before a stability failure may occur is very long in comparison with the natural roll period. Because stability failures are rare, there is a need to obtain estimates of the rate of stability failures, which, in turn, means performing many time-domain numerical simulations or conducting model tests for long durations. To address this problem, the framework suggests several possible solutions.

The plan of action for the development of new genera- tion intact stability criteria includes the development of direct assessment procedures, standard requirements for onboard, ship-specific guidance, criteria for cer- tain types of ships or operations that may not already be included in the 2008 IS Code, and a plan by which the criteria may be implemented into the 2008 IS Code. Several research efforts aimed at developing alternative methods of demonstrating compliance through experi- mental testing (Bertaglia, Scarpa, Serra, Francescutto, & Bulian, 2004; Bulian, Francescutto, Serra, & Umeda, 2004; Hua, 2004; Clauss, Hennig, Brink, & Cramer, 2004), ~erformance-based criteria (Cramer. Kruger. & Mains.

Page 64: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

I NTACT

2004; Kruger et al., 2004), and probabilistic stability criteria (Bertaglia et al., 2004) have been stimulated by the IMO opening of consideration of alternatives to the prescriptive criteria currently in place. Belenky, de Kat, and Umeda (2008) addressed the principle issues related to development of performance-based criteria for intact stability including the motivation for the effort and the physics of the three modes of stability failures under consideration, and paid special attention to the problems involved in probabilistic performance-based criteria development. Belenky et al. (2008) also pro- vided a review of the current methods and techniques for simulating all three modes of stability failures.

7.10 Dynamic Stability Assessments. While dynamic stability has been discussed for decades within the context of equilibrium concepts of stability, the study of ship motions in winds and waves has been tradition- ally within the purview of seakeeping assessments. In recent years, the use of more advanced computational tools has permitted researchers to evaluate the use of these tools for complete assessment of ship stability performance in extreme sea conditions. Casualties as- sociated with parametric roll resonance also focused attention on dynamic stability.

The preamble of the 2008 IS Code acknowledges this situation and recommended that the code should not re- main static but should be reevaluated and revised, as necessary, to take account of rapidly evolving modern ship design technology. It further recognized:

that in view of a wide variety of types, sizes of ships and their operating and environmental conditions, problems of safety against accidents related to

stability have generally have not yet been solved. In particular, the safety of a ship in a seaway in- volves complex hydrodynamic phenomena which up to now have not been fully investigated and un- derstood. Motions of ships in a seaway should be treated as a dynamical system and relationships between ship and environmental conditions like wave and wind excitations are recognized as ex- tremely important elements. Based on hydrody- namic aspects and stability analysis of a ship in a seaway, stability criteria development poses com- plex problems that require further research.

The IMO framework for new generation intact stabil- ity criteria development includes several key elements. Among these are the development of vulnerability cri- teria to identify susceptibility to partial or total stabil- ity failures and development of procedures for direct stability assessment. Initially, this development would address major dynamic stability failure modes includ- ing stability under the deadship condition, righting arm variation problems such as parametric excitation and pure loss of stability, and maneuvering related prob- lems in waves such as broaching-to. The focus would be on ships that are susceptible to a stability failure that is neither explicitly nor properly covered by the exist- ing intact stability regulations. Long-term work will in- clude development of on-board, ship-specific guidance requirements.

A comprehensive list of references at the time of this writing can be found in Belenky et al. (2008) and in Belenky and Sevastianov (2007).

Draft, Trim, Heel, and Displacement

Computation of the draft, trim, heel, and displace- ment of a ship or other vessel is a calculation rarely done by hand since the advent of personal computers. However, the algorithms in the computer software often apply the same methods used in hand calculations. In fact, loading computer software is often required to du- plicate results obtained by hand. Rather than utilizing curves of form to obtain hydrostatic properties, com- puter software utilizes table lookup and interpolation routines to interrogate the hydrostatic tables or direct computations from the hull geometry model to perform the same functions. The calculations outlined here are easily adapted to spreadsheet use.

8.1 Trim. Trim, as used in this section, defines the longitudinal inclination of the ship. Trim may be ex- pressed as the angle between the baseline of the ship and the waterplane, but it is usually expressed as the difference in drafts at the bow and at the stern.

8.2 Center of Flotation. The center of flotation is the point in the ship's waterplane through which the axis of rotation passes when the ship is inclined by a pure moment without change of displacement, either trans- versely, longitudinally, or both. It is shown in Section 4 that, for longitudinal inclinations, this point is the cen- troid of the waterplane, and similar reasoning would ap- ply to inclinations in any direction.

The center of flotation is useful in the determination of drafts for two reasons. When the ship is trimmed with no change in displacement, as when a weight is moved forward or aft, there is no change in draft at the center of flotation. If the change in trim is moderate and the origi- nal waterline and the change in trim are known, the new waterline can be established. Also, if a small weight is added to the ship at the center of flotation, there is no change in trim because the increment of weight is added directly above the location of the increment of buoyancy.

Page 65: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

I NTACT

2004; Kruger et al., 2004), and probabilistic stability criteria (Bertaglia et al., 2004) have been stimulated by the IMO opening of consideration of alternatives to the prescriptive criteria currently in place. Belenky, de Kat, and Umeda (2008) addressed the principle issues related to development of performance-based criteria for intact stability including the motivation for the effort and the physics of the three modes of stability failures under consideration, and paid special attention to the problems involved in probabilistic performance-based criteria development. Belenky et al. (2008) also pro- vided a review of the current methods and techniques for simulating all three modes of stability failures.

7.10 Dynamic Stability Assessments. While dynamic stability has been discussed for decades within the context of equilibrium concepts of stability, the study of ship motions in winds and waves has been tradition- ally within the purview of seakeeping assessments. In recent years, the use of more advanced computational tools has permitted researchers to evaluate the use of these tools for complete assessment of ship stability performance in extreme sea conditions. Casualties as- sociated with parametric roll resonance also focused attention on dynamic stability.

The preamble of the 2008 IS Code acknowledges this situation and recommended that the code should not re- main static but should be reevaluated and revised, as necessary, to take account of rapidly evolving modern ship design technology. It further recognized:

that in view of a wide variety of types, sizes of ships and their operating and environmental conditions, problems of safety against accidents related to

stability have generally have not yet been solved. In particular, the safety of a ship in a seaway in- volves complex hydrodynamic phenomena which up to now have not been fully investigated and un- derstood. Motions of ships in a seaway should be treated as a dynamical system and relationships between ship and environmental conditions like wave and wind excitations are recognized as ex- tremely important elements. Based on hydrody- namic aspects and stability analysis of a ship in a seaway, stability criteria development poses com- plex problems that require further research.

The IMO framework for new generation intact stabil- ity criteria development includes several key elements. Among these are the development of vulnerability cri- teria to identify susceptibility to partial or total stabil- ity failures and development of procedures for direct stability assessment. Initially, this development would address major dynamic stability failure modes includ- ing stability under the deadship condition, righting arm variation problems such as parametric excitation and pure loss of stability, and maneuvering related prob- lems in waves such as broaching-to. The focus would be on ships that are susceptible to a stability failure that is neither explicitly nor properly covered by the exist- ing intact stability regulations. Long-term work will in- clude development of on-board, ship-specific guidance requirements.

A comprehensive list of references at the time of this writing can be found in Belenky et al. (2008) and in Belenky and Sevastianov (2007).

Draft, Trim, Heel, and Displacement

Computation of the draft, trim, heel, and displace- ment of a ship or other vessel is a calculation rarely done by hand since the advent of personal computers. However, the algorithms in the computer software often apply the same methods used in hand calculations. In fact, loading computer software is often required to du- plicate results obtained by hand. Rather than utilizing curves of form to obtain hydrostatic properties, com- puter software utilizes table lookup and interpolation routines to interrogate the hydrostatic tables or direct computations from the hull geometry model to perform the same functions. The calculations outlined here are easily adapted to spreadsheet use.

8.1 Trim. Trim, as used in this section, defines the longitudinal inclination of the ship. Trim may be ex- pressed as the angle between the baseline of the ship and the waterplane, but it is usually expressed as the difference in drafts at the bow and at the stern.

8.2 Center of Flotation. The center of flotation is the point in the ship's waterplane through which the axis of rotation passes when the ship is inclined by a pure moment without change of displacement, either trans- versely, longitudinally, or both. It is shown in Section 4 that, for longitudinal inclinations, this point is the cen- troid of the waterplane, and similar reasoning would ap- ply to inclinations in any direction.

The center of flotation is useful in the determination of drafts for two reasons. When the ship is trimmed with no change in displacement, as when a weight is moved forward or aft, there is no change in draft at the center of flotation. If the change in trim is moderate and the origi- nal waterline and the change in trim are known, the new waterline can be established. Also, if a small weight is added to the ship at the center of flotation, there is no change in trim because the increment of weight is added directly above the location of the increment of buoyancy.

Page 66: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

54 INTACT STAB1 l lTY

The longitudinal position of the center of flotation is plotted on the curves of form, as described in Letcher (2009). The curve may be labeled center offiotation or CG of waterplane.

8.3 Moment to Trim 1 Cm. The formulas for moment to trim 1 em, AGMJlOOL, and approximate moment to trim 1 em, pIL/lOOL, are derived in Section 3.7.

The trim produced by a moderate longitudinal mo- ment can be calculated by dividing the moment by the approximate moment to trim 1 ern with sufficient accu- racy for normal ships. There may be some unusual craft for which the difference between GM, and BML is large enough to make the use of the approximate moment to trim unacceptable, particularly if there is a large longi- tudinal free-surface effect which affects the position of the CG substantially.

The moment to trim 1 em is usually plotted on the curves of form.

8.4 Tons per Centimeter Immersion. The tons per cen- timeter immersion, which is the displacement of a layer of water 1 ern thick at the waterplane, is used to calcu- late the change in draft at the center of flotation caused by a moderate change in displacement. The increase or decrease in draft is equal to the change in displacement divided by the tons per centimeter immersion.

This function, TPcm, is plotted on the curves of form. In practice, TPcm is often replaced in computer software algorithms by taking the derivative of the dis- placement versus draft relationship.

8.5 Determination of Drafts from Weight and Location of Center of Gravity. There are two methods by which the drafts forward and aft may be obtained when the displacement and longitudinal position of the ship's CG are known. The first involves the displacement and other curves and is used when the trim is moderate. As noted earlier, this approach is often implemented in loading computer software to duplicate hand calcula- tions. The second is based on direct computation of hull properties from the geometry based upon the current waterplane and is used when the trim is so large that the approximations used in the first method are not accept- able. Alternatively, use of the Bonjean curves is possible, but rarely implemented when a computer is used. When using the curves in these methods, the ship is assumed to be at the angle of heel for which the hydrostatic prop- erties have been evaluated. The second method is the basis for software algorithms with the extension that the hull hydrostatic properties are evaluated at an arbi- trary, specified angle of heel.

The steps in determining the forward and after drafts from the curves of form or hydrostatic tables are as fol- lows:

1. The even keel draft, or draft at LCF, is determined at the value indicated by the weight summary.

2. At this draft, the longitudinal location of the center of buoyancy, the longitudinal location of the center of flota- tion, and the approximate moment to trim 1 ern are found.

3. Assume, for the moment, that the ship's CG is at the longitudinal position of the center of buoyancy as read from the displacement and other curves. If this were the case, the ship would be floating at even keel, and the drafts forward and aft would be equal to the draft read from the displacement curve. If, as is usually the case, the CG obtained from the weight estimate is not at the longitudinal position of the center of buoyancy, there is a trimming moment, MT, equal to the weight of the ship multiplied by the distance, parallel to the keel, from the center of buoyancy to the CG.

4. The trim, in centimeters, produced by the moment, MT, is calculated by dividing MT by the approximate mo- ment to trim 1 em. This is the difference between the forward and after drafts. The trim will be by the bow if the CG is forward of the center of buoyancy in the even keel attitude; otherwise, the trim will be by the stern.

5. The slope of the trimmed waterline with respect to the even keel waterline is determined by dividing the trim by the length between perpendiculars.

6. The draft at the center of flotation will be equal to the even keel draft found in step 1, since the trimmed waterline will intersect the even keel waterline at this point. The draft at either perpendicular will be equal to the draft at the center of flotation plus or minus the product of the slope of the trimmed waterline and the distance from the center of flotation to that perpen- dicular.

As an illustration of this process, the following ex- ample is presented. Given:

Length between perpendiculars = 161 m Displacement, SW = 19,000 KN Even keel LCF draft = 8.32 m LCG = -3.66 m

At 8.32 m draft: LCB from amidships = -2.42 m LCF from amidships = -5.73 m MT = 214 KN-m

Trimming moment: 19,000(-3.66 + 2.42) = 23,500 KN-m

Center of flotation to after perpendicular: (80.5 - 5.73) = 74.77 m

Slope of trimmed WL:

Draft aft = 8.32 + 74.77 x 0.0068 = 8.32 + 0.51 = 8.83 m

The curves of form would indicate whether these drafts are drafts to the bottom of the keel or to some other baseline.

In rare cases, the moment to trim 1 ern found on the curves of form or in the hydrostatic tables may be based on the length between draft marks rather than the length between perpendiculars, in which case the slope

Page 67: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

I NTACT

of the trimmed waterline is equal to the trim divided by the length between the draft marks.

The foregoing method for determining drafts involves two approximations that are sufficiently accurate for a small trim but become less accurate as the trim increases. One is the assumption that the trimmed and even keel waterlines intersect at the center of flotation of the even keel waterline. The other is that there is no change in the moment to trim 1 em as the ship is trimmed. If the drafts must be determined very accurately, if the trim is more than about 11150 of the ship's length, or if the trim pro- duces a marked change in the shape of the waterplane, the following method is appropriate.

This iterative method involves finding, by trial and er- ror, the forward and after drafts that correspond to the given displacement and produce a center of buoyancy in the same longitudinal position as the given CG. The steps are:

1. Assume a starting draft and trim. This initial trial may be based upon the first method.

2. Determine the difference between displacement and buoyancy, and between the longitudinal moments of the weight and buoyancy.

3. Estimate the change in draft at the center of flo- tation and the change in trim. These estimates can be based upon calculated values of tons per centime- ter and moment to trim 1 em. These derivatives of the evaluation function can also be calculated using a finite difference approach; however, that requires additional computations.

4. Convergence is achieved when the difference be- tween displacement and buoyancy, and in their longi- tudinal moments are within tolerance, or alternatively when the changes in draft and trim between iterations become acceptably small. Satisfactory results are usu- ally obtained on the third iteration.

The final solution will be within tolerances, but will differ for different initial estimates. For this reason,

even when doing direct calculations, computer soft- ware algorithms must base the calculations on a consis- tent starting point to insure repeatability. This method is as accurate for a trimmed waterline as for an even keel waterline.

8.6 Determining Displacement and Center of Gravity from Drafts

8.6.1 Methods. The displacement and the longi- tudinal location of the ship's CG can be determined if the drafts forward and aft are known. Direct evaluation of the buoyancy and coordinates of the center of buoy- ancy from the hull geometry model for the waterplane defined by the drafts is recommended. For moderate trim, with no abrupt changes in the waterplane between the trimmed and corresponding even keel waterlines, the displacement and other curves may be used with reasonable accuracy. Use of the Bonjean curves to de- termine the longitudinal center of buoyancy is possible, but rarely implemented when computers are used.

When the curves of form are used, the displacement may be obtained by either of two methods, one of which depends upon the availability of the function increase (or decrease) in displacement per centimeter of trim aft. The derivation of this function is illustrated in Fig. 60, which shows a ship with a trim by the stern, the midship perpendicular, and the center of flotation for an even keel waterline.

For moderate trim, the displacement is equal to that under an even keel waterline passing through point F and can be read from the displacement curve at draft T,. If the draft readings are taken at the perpendiculars or at two sets of draft marks nearly equidistant from amidships, it is convenient to enter the displacement curve with the mean of the draft readings, T,, read the displacement, and add (or subtract) the correction rep- resenting the difference in displacement at the even keel drafts Tl and 71,. The value of this correction is simply the difference in drafts Tl and T2 (in m) multiplied by the TPm. Note: in SI units, this is actually Newtons or

LBP Fig. 60 Trimmed waterline and center of flotation.

Page 68: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

56 INTACT STAB1 l lTY

Kilo-Newtons per meter, but the original nomenclature is retained. If Tl and T2 are in meters, this is

The value of (T, - Ir,) or h in Fig. 60, may be obtained conveniently from similar triangles,

hld = tlL, or h = d . tlL

Hence, the change in displacement becomes

d - (tlL) - TPm

For t = 1 m, the change in displacement per m of trim is (dlL) TPm.

In cases where the center of flotation is at a consider- able distance from amidships (resulting in an apprecia- ble difference between T, and T,) and where the values of the other functions, such as transverse metacenter above baseline, longitudinal center of buoyancy, or mo- ment to trim 1 cm are changing rapidly with change in draft, more accurate values of these functions will be obtained if they are read at T, rather than at T2 because they will correspond to the draft at which the displace- ment is equal to the actual displacement.

After the displacement has been found, the CG may be located as follows:

1. The location of the center of buoyancy and the moment to trim 1 ern at the even keel draft Tl, or as an approximation, Ir,, are read from the displacement and other curves.

2. The trim between the perpendiculars is found by taking the difference between the drafts at the forward and after perpendiculars, or, if the draft marks are not located at the perpendiculars, by multiplying the differ- ence in the readings by the ratio of the length between perpendiculars to the distance between the marks.

3. The ship is initially assumed to be floating in equi- librium on an even keel at the displacement just found, in which case the CG would be in the same longitudinal position as the center of buoyancy. If we now assume that the given trim was produced by movement of the CG, the moment generated by this movement would be equal to the trim multiplied by the moment to trim 1 em. The movement of the CG would be equal to this moment divided by the ship's displacement.

4. If the trim is by the bow, the CG would have moved forward; otherwise, it would have moved aft.

8.6.2 Numerical Examples. Illustrative examples of the above procedures are shown below.

8.6.2.1 To DETERMINE DISPLACEMENT Given: LBP = 161 m

Draft at fwd perp = 7.45 m Draft at aft perp = 8.38 m

Mean draft = (7.45 +8.38)

=7.92 m 9

Trim = 8.38 - 7.45 = -0.93 m

At the 7.92 m waterline:

Displacement = 17,960 KN Increase in displacement per meter trim aft = 88.6 KN

Increase in displacement:

0.93 x 86.6 = 82.4 KN (say 82)

Corrected displacement = 18,042 KN

OR

Given: Mean draft = 7.92 m

LCF at 7.92 m draft from amidships = -5.3 m Trim by stern = -0.93 m

Then

-0.93 Slope of trimmed waterline = - - - - 0.0058 161

Draft at center of flotation = 7.92 + 5.3 x .0058 = 7.95 m Displacement at 7.95 m waterline = 18,042 KN

8.6.2.2 To FIND THE LONGITUDINAL CENTER OF GRAVITY. At the 7.95 m waterline:

LCB from amidships = -2.26 m Moment to trim 1 ern = 207 KN-m Trimming moment = -0.93 x 207 = -19,251 KN-m

Shift of CG =

Location of CG from amidships =

8.7 Determining Draft After Change in Loading. When a change in loading is contemplated, there are two ap- proaches for determining the drafts after the change is made. The approach prevalent in computer software is to consider the new loading independently and utilize the methods described in Section 8.5.

A second approach, which is appropriate when com- puter software is not available, the changes in draft and trim are small, and no large initial trim is present, in- volves determining the difference~ in drafts before and after the change.

When this method is used, the change in loading is reduced to a single equivalent item by assuming a single equivalent weight is added at the center of flotation. The weight is then shifted longitudinally such that the mo- ment due to the shift equals the sum of moments of the individual changed items. The draft would change by an amount equivalent to the changed weight divided by the TPm, termed "parallel sinkage." It is next assumed that the equivalent item is moved from the center of flota- tion through the longitudinal shift distance computed above. This would generate a trimming moment equal to the weight of the equivalent item times the distance, parallel to the keel, from the center of flotation to its

Page 69: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 57

actual location. The change in trim is equal to this trim- ming moment divided by the moment to trim 1 em. For example, if the weight is added forward of the center of flotation so that the trim will be by the bow, the increase in draft forward will be equal to the parallel sinkage plus the change in trim multiplied by the ratio of the distance of the center of flotation aft of the forward perpendicu- lar to the length between perpendiculars. The change in draft aft, which may be negative, will be equal to the parallel sinkage minus the change in trim multiplied by the ratio of the distance of the center of flotation for- ward of the after perpendicular to the length between perpendiculars.

The following example illustrates the procedure. Start with the ship given in Section 8.6. Assume 300 KN are to be added 61 m forward of amidships. From the example in Section 8.6, at the 7.95 m waterline:

Tons per meter = 2650 KN LCF from amidship = -5.3 m Moment to trim 1 m = 20,700 KN-m Parallel sinkage = 300 KN I 2650 KNIm = 0.1132 m CG of added weight, from the center of flotation

= (61 + 5.3) = +66.3 m Trimming moment = 300 x 66.3 = 19,890 KN-m Change in trim = 19,890 KN-m/20,700 KN-m = 0.96 m Center of flotation from fwd perpendicular

= (80.5 + 5.3) = 85.8 m Center of flotation from aft perpendicular

= (80.5 - 5.3) = 75.2 m Change in draft forward

= 0.96 m x 85.8 ml161 m = 0.51 m Change in draft aft = 0.96 m x 75.2 ml161 m = 0.45 m Draft forward = (7.45 + 0.11 + 0.51) = 8.07 m Draft aft = (8.38 + 0.11 - 0.45) = 8.04 m

8.8 Navigational Drafts. The navigational draft of a ship in any condition is the draft to the bottom of the greatest projection below the waterline of the hull

or any appendage. This draft determines the depth of water required for safe navigation. The navigational draft may be increased somewhat over its static value due to settling of the ship caused by forward motion. It is sometimes necessary for a ship to pass through a channel or enter a dry dock in which the depth of water is less than the ship's normal navigational draft. This situation requires a study of the possible adjustments of the load to minimize the navigational draft. In extreme cases, it may be desirable to remove or, in the case of an uncompleted ship, to omit some items of lightship weight. The problem may be complicated by the projec- tion of appendages below the keel.

Fig. 61 shows a ship on which the rudder projects be- low the line of the keel a distance a, and an appendage forward projects a greater distance b. The navigational draft will occur either at A or B, depending upon the direction and amount of trim. For any given displace- ment, the navigational draft will be minimized when the drafts at A and B are equal. The slope of the trimmed waterline which will produce this condition is equal to (b - a)lc with the ship trimmed by the stern, which is derived as follows:

Let the fwd navigational draft = yl + b Let the aft navigational draft = y2 + a For equal drafts at A and B,

Then

tan 19 = (y2 - yJc = (b - a)lc

The trim for minimum navigational draft will be the slope of the waterline multiplied by the length between perpendiculars (L) or (b - a) Llc.

Assume that the drafts of the ship in some condition of loading are known, and the problem is to reduce the navigational draft to a certain figure, based on the chan- nel depth. Assume that the draft at the rudder in Fig.

Fig. 61 Projections determine navigational drafts.

Page 70: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

58 INTACT STAB1 l lTY

61 exceeds that at the forward appendage. In this ease, there is a plane, P, located in the forward portion of the ship, at which an addition or removal of weight will have no effect on the draft at the rudder. If a weight is added in plane P, the effect on the after draft may be consid- ered, as discussed in Section 8.7, to consist of parallel sinkage as if the weight were added at the center of flo- tation, and the effect of the trim produced by moving the weight from the center of flotation to plane P. If these two effects are equal and in opposite directions, there will be no change in draft at the rudder. This may be expressed as follows.

w w.e d - - -.- where w = added weight

TPL M T ~ L

Then the plane P can be located by solving for e,

This equation applies not only to the rudder, but to any point on the ship, if d is the distance from the point at which no change of draft is desired to the center of flotation.

If the weight were removed rather than added, these equations would still apply because the signs of both sides would be reversed. If the moment to trim 1 cm is based on the length between draft marks, rather than the length between perpendiculars, the length between draft marks would be substituted for L in the equa- tions.

Having determined the optimum trim and the loca- tion at which weight changes have no effect on naviga- tional draft, we can reduce the navigational draft by removing weight aft of plane P, adding weight forward of plane P, or both until the optimum trim is attained. If the draft is still excessive, further reduction can be obtained only by changes which are equivalent to the removal of weight from the portion of the ship between plane P and the corresponding plane Q in the after part of the ship in which changes have no effect on the draft at the forward appendage. This does not preclude the use of weight additions as a means for further reduction of navigational draft, but any addition must be made in conjunction with a removal of greater magnitude, such that the CG of the equivalent removal will fall between planes P and Q.

In selecting items to be removed or the location at which an item is to be added, it should be realized that the effect of changes in weight fairly close to plane P or Q, although favorable, may be too small to justify the effort involved.

8.9 Hog and Sag. Deviation of the keel from a straight line, which may be of a permanent or temporary nature or a combination of both, is known as hog when the keel is concave downward or sag when the keel is concave upward. Permanent deflection may be caused by shrinkage associated with welding, while temporary

deflections are caused by flexure due to the ship's load- ing or by thermal expansion. If hog or sag is appreciable, an adjustment should be made when determining dis- placement from draft readings.

Hog or sag is detected by calculating the draft amid- ships as the mean of the readings of the forward and after draft marks and comparing the calculated value with the actual reading of the midship draft marks. Hog or sag is measured by the difference in the calculated and actual drafts amidships. If the actual reading is greater than the calculated value, the ship is sagging; if the actual reading is less, the ship is hogging.

To find the displacement in a known hogged or sagged condition, using the hull geometry model or Bonjean curves, it is customary to enter the curves at the actual drafts forward and aft and to determine the drafts at the intermediate stations by the assumption that the keel deflection is parabolic. If the deflection is explicitly defined, from shipboard measurements or a deflection analysis of a structural model of the ship, then the dis- placement can be calculated based upon the actual im- mersed volume. When the approximate method based on displacement and other curves is used, the practice is to increase the draft at the center of flotation or at amidships, calculated as described in Section 8.5, by 75% of the sag or to decrease it by 75% of the hog.

For a ship with a rectangular waterplane, the percent- age would be 67, since the area under a parabola is two thirds of the area of the circumscribing rectangle, and this percentage would tend to increase as the waterline becomes finer at the ends. The 75% figure is an approxi- mation which has been found to give good agreement with direct calculations for a normal ship form.

8.10 Drag. Some ships, particularly tugs and small high-speed craft that have considerable power for their size, are designed to float deeper at the stern than at the bow in order to submerge their relatively large pro- pellers. Craft designed to come ashore on beaches are also designed with a permanent stern drag so that when grounded along the ship bottom, the propeller will re- main submerged. The designed drag is the difference between drafts to the bottom of the keel at the forward and after perpendiculars.

When a ship has a designed drag, the waterlines used in preparing the displacement and other curves are not parallel to the keel but are parallel to the design water- line. The ship is considered to have zero trim when float- ing with the keel sloped to suit the designed drag.

Great care should be used in relating drafts to dis- placement and center of buoyancy when the ship has a designed drag. There is no fixed convention for locating the baseline from which drafts are measured in drawing the displacement and other curves. It may be a horizon- tal line at the intersection of the bottom of the keel with the after perpendicular or the midship station, or the intersection of the molded baseline with one of the per- pendiculars. It may be completely below the ship. The draft marks installed on the ship may indicate drafts

Page 71: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 59

above any of these baselines or above the bottom of the keel at the location of the draft marks.

8.1 1 Computation of Heel. When computing the equi- librium heel angle of a ship using hand calculations, the cross curves (see Section 4.4) are useful. However, when computed manually, cross curves usually neglect effects due to change in trim with heel. Examples of the errors that may occur in computed righting arms if trim is neglected are found in Paulling (1959). Computer soft- ware currently available makes the formerly laborious task of including the effects of trim on the righting arm no longer an issue.

One approach is to:

1. Specify a range of heel angles to evaluate. 2. For each heel angle, fix the ship at that heel angle

and compute the longitudinal equilibrium position (i.e., mean draft and trim).

3. Evaluate the righting moment based upon the po- sitions of the transverse and vertical centers of gravity and buoyancy. Compute m b y dividing righting moment by displacement. See Fig. 17 for the derivation of m .

4. Fit a spline curve through the mversus heel angle data and determine locations where m i s zero. The low- est angle where = is zero with a positive slope to the - GZ curve is the angle of equilibrium under static load- ing. A = of zero in the upright position with negative slope to the = curve can be an example of unstable equilibrium. There is a trade-off in accuracy of the es- timate of the equilibrium angle from the spline curve

fit to the m c u r v e between number of angles and com- putation time. Some algorithms use the spline curve to estimate the location and use a search routine to fine tune the answer.

It is possible to directly search for the equilibrium angle without computing the =curve, but computation of this curve is usually done anyway as it is necessary for most stability criteria evaluations.

8.12 Reference Planes. Calculations such as draft, displacement, heel, and longitudinal position of the CG involve the use of longitudinal, transverse, and horizon- tal reference planes which are a potential source of seri- ous error if not specified clearly and accurately. It is of critical importance for the user to clearly understand the exact definitions of these planes of reference when using the information generated from such calculations. If, for example, the weight estimate and the curves of form do not use the same reference plane, it is neces- sary to take into account the separation of the two refer- ence planes in finding the distance from the CG to the center of buoyancy. A frequent example is when the wa- terlines on the displacement and other curves are mea- sured from the molded baseline while the draft marks on the ship use the bottom of the keel as a reference. On wooden ships, these reference planes may be as much as 0.3 m apart. Once again, the user must be careful to ensure that the reference planes are consistent or that the proper adjustment has been made.

The Inclining Experiment

9.1 Basic Principles. Toward the end of the construe- of inclining, draft readings also are taken in order to es- tion period, an inclining experiment is conducted to es- tablish its displacement. tablish, experimentally, the weight of the ship and the The displacement and the longitudinal position of the vertical and longitudinal coordinates of its CG. The re- ship's CG, in the condition in which it is inclined, are sults of the inclining experiment customarily supersede found from the observed drafts as discussed in Section the corresponding figures in the weight estimate. 8. The metacentric height, is then determined in the

The International Convention on Safety of Life at Sea following manner: - (S0LAS'747 as amended) requires that every passenger 1. As discussed in Section 3, the righting arm, m, at ship regardless of size and every cargo ship having a a small angle of inclination, 4, is length of 24 m and u~wards should be inclined on com- " - -

pletion and the elements of its stability determined. An GM = GM sin 4 (31) inclining experiment is also to be conducted where any From which it follows that the righting moment is alterations are made to a s h i ~ so as to materiallv affect the stability.

The inclining experiment entails heeling the ship to where W is the displacement of the ship. In the inclining a small angle by the movement of a known weight in experiment, this is determined from the drafts. a direction perpendicular to the ship's centerline plane 2. The heeling moment, M, produced by moving a through a measured distance. The ship is allowed to at- weight, w, aboard ship perpendicular to the ship,s ten- tain a state of static equilibrium in heel, at which point terline plane through a distance, d, is the angle of inclination is measured. The process is then repeated at several angles in both directions. At the time M = w . d c o s + (33)

Page 72: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 59

above any of these baselines or above the bottom of the keel at the location of the draft marks.

8.1 1 Computation of Heel. When computing the equi- librium heel angle of a ship using hand calculations, the cross curves (see Section 4.4) are useful. However, when computed manually, cross curves usually neglect effects due to change in trim with heel. Examples of the errors that may occur in computed righting arms if trim is neglected are found in Paulling (1959). Computer soft- ware currently available makes the formerly laborious task of including the effects of trim on the righting arm no longer an issue.

One approach is to:

1. Specify a range of heel angles to evaluate. 2. For each heel angle, fix the ship at that heel angle

and compute the longitudinal equilibrium position (i.e., mean draft and trim).

3. Evaluate the righting moment based upon the po- sitions of the transverse and vertical centers of gravity and buoyancy. Compute m b y dividing righting moment by displacement. See Fig. 17 for the derivation of m .

4. Fit a spline curve through the mversus heel angle data and determine locations where m i s zero. The low- est angle where = is zero with a positive slope to the - GZ curve is the angle of equilibrium under static load- ing. A = of zero in the upright position with negative slope to the = curve can be an example of unstable equilibrium. There is a trade-off in accuracy of the es- timate of the equilibrium angle from the spline curve

fit to the m c u r v e between number of angles and com- putation time. Some algorithms use the spline curve to estimate the location and use a search routine to fine tune the answer.

It is possible to directly search for the equilibrium angle without computing the =curve, but computation of this curve is usually done anyway as it is necessary for most stability criteria evaluations.

8.12 Reference Planes. Calculations such as draft, displacement, heel, and longitudinal position of the CG involve the use of longitudinal, transverse, and horizon- tal reference planes which are a potential source of seri- ous error if not specified clearly and accurately. It is of critical importance for the user to clearly understand the exact definitions of these planes of reference when using the information generated from such calculations. If, for example, the weight estimate and the curves of form do not use the same reference plane, it is neces- sary to take into account the separation of the two refer- ence planes in finding the distance from the CG to the center of buoyancy. A frequent example is when the wa- terlines on the displacement and other curves are mea- sured from the molded baseline while the draft marks on the ship use the bottom of the keel as a reference. On wooden ships, these reference planes may be as much as 0.3 m apart. Once again, the user must be careful to ensure that the reference planes are consistent or that the proper adjustment has been made.

The Inclining Experiment

9.1 Basic Principles. Toward the end of the construe- of inclining, draft readings also are taken in order to es- tion period, an inclining experiment is conducted to es- tablish its displacement. tablish, experimentally, the weight of the ship and the The displacement and the longitudinal position of the vertical and longitudinal coordinates of its CG. The re- ship's CG, in the condition in which it is inclined, are sults of the inclining experiment customarily supersede found from the observed drafts as discussed in Section the corresponding figures in the weight estimate. 8. The metacentric height, is then determined in the

The International Convention on Safety of Life at Sea following manner: - (S0LAS'747 as amended) requires that every passenger 1. As discussed in Section 3, the righting arm, m, at ship regardless of size and every cargo ship having a a small angle of inclination, 4, is length of 24 m and u~wards should be inclined on com- " - -

pletion and the elements of its stability determined. An GM = GM sin 4 (31) inclining experiment is also to be conducted where any From which it follows that the righting moment is alterations are made to a s h i ~ so as to materiallv affect the stability.

The inclining experiment entails heeling the ship to where W is the displacement of the ship. In the inclining a small angle by the movement of a known weight in experiment, this is determined from the drafts. a direction perpendicular to the ship's centerline plane 2. The heeling moment, M, produced by moving a through a measured distance. The ship is allowed to at- weight, w, aboard ship perpendicular to the ship,s ten- tain a state of static equilibrium in heel, at which point terline plane through a distance, d, is the angle of inclination is measured. The process is then repeated at several angles in both directions. At the time M = w . d c o s + (33)

Page 73: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

60 INTACT STAB1 LlTY

3. Since the righting moment and heeling moment are equal at the time the inclination is measured,

- w.dcos4 GM =

Wsin qb

The height of the ship's CG is found by subtracting the metacentric height, from m, the height of the

metacenter above the keel.

When the ship as inclined has a considerable trim, it is usually necessary to calculate displacement and directly from the hull geometric model for the actual trimmed condition. However, if the ship is inclined in a nearly zero trim condition, the height of the metacenter may be read from the displacement and other curves at the draft at the center of flotation, after the draft at the center of flotation has been obtained from the observed drafts.

It is desirable to perform the inclining experiment when the ship is as nearly complete as is practicable, and at this late stage of construction there are usually some tanks containing fuel oil or fresh water. It is de- sirable to adjust these liquids before the experiment in such a way that the tanks are either (a) completely full or completely dry to avoid any free surface, or (b) to avoid having any tank nearly full or nearly empty, so that there will be no appreciable change in the moment of inertia of the surface during the expected inclination as a result of the liquid surface reaching either the tank top or bottom. When this is done, the value of "as inclined" is the virtual rn and the height of the CG is the virtual height, which includes the free-surface ef- fect. The value of the free-surface effect, in meters, is equal to the summation of i,/S for the various tanks di- vided by the ship's displacement, where i, is the moment of inertia of the free surface and S the specific volume of the liquid in the tank in cubic meters per ton, as dis- cussed in Section 5. To find the real CG of the ship, the free-surface effect, as well as the virtual rn must be subtracted from the height of the metacenter, or

Occasionally, there may be an unavoidable free sur- face that is not constant throughout the range of inclina- tion and therefore cannot be treated as a virtual rise of the CG. Its effect may be taken into account by consid- ering the shifted liquid as part of the inclining weight by adding the weight of the liquid, multiplied by the dis- tance its CG moves in the direction perpendicular to the ship's centerline plane, to the moment of the inclining weight.

In almost all cases, a new ship is inclined in a condi- tion that is not representative of a real operating condi- tion in that the construction or conversion work has not yet been completed, there is a considerable amount of foreign material aboard, such as staging, and the ship may be partially loaded with fuel oil and fresh water. However, it is desirable to delay the inclining until most of the foreign material is no longer needed and then re- move most of it from the ship. Then, as an immediate preparation for the test, it is necessary to make an esti- mate (inventory) of the weight and the vertical and longi- tudinal moments of all items which are part of the light- ship and have not yet been put aboard, all foreign items which will be removed, and all items of load which are aboard, and to apply these estimates to the weight and the vertical and longitudinal moments of the ship as de- termined from the inclining experiment to produce the lightship condition. If there are items of lightship weight aboard but not in their final positions, the moments that will result from shifting these items must be included. The importance of a thorough and accurate inventory cannot be overemphasized because an accurate deter- mination of the position of the CG from the inclining experiment is critical to the determination of the safety of the ship during the remainder of its lifetime.

9.2 Preparation for Inclining. The following points require attention prior to the inclining experiment. The latest IMO instructions on preparing a ship for inclining and conducting the experiment may be found in the IMO Code on Intact Stability, 2008.

9.2.1 Schedule. If the inclining test is to be of- ficially witnessed by the government administration (e.g., the USCG) for purposes of stability approval, it is necessary to submit a schedule of the major tasks and procedure to an administration inspection or technical office in advance.

9.2.2 Drafts and Trim. While it is possible, and sometimes necessary, to incline a ship with a large trim, there are advantages in reducing trim nearly to zero, as previously noted. Excessive trim will make it necessary to correct readings of tank capacities and centroids given in the tank capacity tables and make it more dif- ficult to adjust the levels in partially filled tanks so that the liquid level will not reach the tops or bottoms of the tanks. Drafts at which abrupt changes in the waterplane will occur as the ship is inclined should be avoided.

9.2.3 List. While a small initial heel is not objec- tionable, it should be small enough so that the list, plus the expected inclination, will not exceed the angle at which the relationship = sin 4 no longer ap- plies.

9.2.4 Metacentric Height. The ship must have pos- itive metacentric height at the time of the experiment, after allowances are made for free surface and the ef- fect of the inclining weights and gear.

9.2.5 Free Surface in Tanks. Free liquids can be dealt with provided the surface does not reach the top or bottom of the tank as the result of the combination

Page 74: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

of list and trim and provided that the moment of inertia of the free surface can be determined accurately. How- ever, consideration of free surface can be eliminated entirely if the tanks are either completely full or com- pletely empty, and if possible these conditions should be obtained. A tank cannot be assumed to be empty un- less it is known that the liquid below the suction has been substantially removed, nor assumed full unless the sounding is well above the top of the tank, and it is known that no large air pockets exist. To accomplish this, an air escape must be available at the highest point of the tank, but even this will not eliminate the numer- ous small air pockets between the structural members. It may be possible to heel the ship while the tank is fill- ing to assist the escape of air. The best procedure is to have all tanks that must contain liquid about half full, provided the resulting free-surface effect can be accu- rately calculated and will not produce negative meta- centric height.

9.2.6 Personnel Aboard. Arrangements should be made to reduce the personnel aboard to a minimum nec- essary for the test. Those permitted to remain should not be allowed to conduct any work involving movement of the people, their equipment, or items in or attached to the ship.

9.2.7 Transfer of Liquids. Arrangements should be made to prevent changes in the liquid load during the experiment. Valves next to the tanks in all systems should be closed. Precautions should be taken to pre- vent both deliberate and accidental transfer. The latter could occur as the ship is inclined if port and starboard tanks are inadvertently cross-connected through a pip- ing system.

9.2.8 Swinging Weights. Items such as boats and booms, which are normally fixed in a stowed position, should be secured to prevent swinging during the ex- periment.

9.2.9 Forces Affecting Heel. During the experi- ment, the inclination of the ship should not be influ- enced appreciably by any forces other than the effect of the inclining weights. Gangways should be lifted clear of the ship during the experiment. The effect of floats, fenders, and submerged objects should be eliminated. The effect of wind, current, pier, mooring lines, cable, and hose should be reduced to a minimum. If possible, the experiment should be performed at slack tide, or, if feasible, in a dry dock. Consideration should be given to the possibility of heading the ship into the wind or cur- rent. Lines, cable, and hose from ship to shore should be well slacked while inclination readings are taken.

9.2.10 Selection of Inclining Weights. Inclining weights should be selected that will produce an angle of heel sufficient to ensure accurate results, but inclina- tions should not be carried to an angle at which m n o longer equals sin 4 (not more than 4 degrees, per IMO). In practice, it is customary to estimate in advance the probable m a t the time the inclining experiment is to be performed, and a weight is then selected which

will give an angle of heel of about 1 degree on each side of the upright for large vessels, 1 112 degree for 30 m vessels of normal form, and 2 to 3 degrees for very small craft of normal form. This practice assures that no ap- preciable change in will occur during the experi- ment. In the case of ships whose sides flare apprecia- bly at the waterline amidships, the angle of inclination should not exceed 1 degree from the upright. The fol- lowing expression derived from equation (34) gives the weight required:

It is advisable to incline the ship, by means of the weights selected, some time prior to the experiment to ensure that a suitable angle can be attained. A car car- rying the weights and rolling on transverse rails gives excellent results because little rolling of the ship is in- duced and the movement of the weights can be measured accurately. Handling weights by a crane is a practical method, although not as satisfactory. The weight of the inclining weights and car, if used, must be accurately determined and recorded.

9.2.11 Measurement of Inclination. Provision should be made for measuring the angle of inclination indepen- dently at three stations. This will permit rejecting a read- ing that is obviously inconsistent with the others. Numer- ous devices have been used for this purpose, some of which give a direct reading of the tangent of the angle. A pendulum of string or fine wire, 4 to 6 m long, with a heavy bob damped in a bucket of oil, will give excellent results and is required for inclinings witnessed by the USCG, unless prior approval is obtained for another de- vice. To obtain a pendulum of this length in a location protected from the wind, it is usually necessary to run the line through one or more hatches. A check should be made to ensure that the pendulum is free to swing to the expected angle without interference. A horizontal trans- verse batten, fixed to the ship's structure, is provided at the lower end, above the bob, for recording pendulum de- flections. The length of each pendulum, from the point of suspension to the batten, is recorded so that tangents of angles of inclination may be calculated.

9.2.12 Reading Draft Marks. Provision should be made for reading the forward, midship, and aft draft marks. A glass tube with a small hole in the bottom and a scale inside, or a similar device, is recommended to damp out minor wave action.

9.2.13 Measurement of Water Density. Provision should be made for obtaining samples of the water in which the ship is floating at the time of the experiment. A weighted bottle which can be opened while submerged is useful in obtaining samples from various depths. Den- sity is then obtained by use of a hydrometer.

9.3 Conducting the Inclining Experiment. The opera- tions involved in conducting the inclining experiment are as follows:

Page 75: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 l lTY

9.3.1 Inventory. A comprehensive definition of the lightship condition is necessary as the basis for the inventory. The inventory consists of three summa- tions: first, the weight and the vertical, longitudinal, and transverse moments of those items which would be removed to bring the ship to the light condition, then similar figures for items to be added, and last, the mo- ments produced by moving to their final positions those items of lightship weight which are aboard but not in their proper locations. The weights to be removed in- clude all items of load and all material aboard which are foreign to the ship, determined by a thorough survey of all spaces. Each tank and void should be sounded, preferably both before and after the experiment, and the specific gravity of the contents recorded. Spaces which are presumably empty should be investigated. The in- clining weights and gear should not be overlooked.

9.3.2 Draft Readings. Draft readings should be taken simultaneously on the port and starboard sides, at the forward, amidships, and after draft marks at the time of inclining. If no midship draft marks are avail- able, the midship drafts should be obtained by measure- ment from the deck amidships.

9.3.3 Determination of Water Density. Hydrom- eter readings should be taken for several samples from various locations along the ship's length and at various depths. The hydrometer readings should be converted to indicate the density of the water in air.

9.3.4 Movement of Personnel. Movement of per- sonnel during the experiment should be restricted.

9.3.5 Weight Movements. Inclining weights are customarily moved to produce at least two inclinations to port and two to starboard, the intermediate inclina- tions being about half the maximum. Weights should be moved slowly or set down easily to avoid inducing a roll. The transverse displacement of each weight from its initial position is measured and recorded after each movement.

9.3.6 Measurement of Inclination. An initial mark is made simultaneously on each batten, if pendulums are used, or any other device is set to zero, while the inclin- ing weights are in their initial position. Thereafter, read- ings of the inclination are taken, simultaneously at each station, after each weight movement. The signal to read the inclination should be given after allowing sufficient time for the ship to come to a position of equilibrium after the weight movement. The ship should be clear of the pier and all lines should be well slacked. If the ship is not absolutely steady, the reading of inclination should be taken at the midpoint of the residual motion.

It is essential to incline to both sides of upright, and it is highly desirable to use either multiple weights or mul- tiple arms so as to heel the ship to at least two different angles on either side of upright. Thus, if anything occurs to cause an erroneous reading, it can be identified as an error more easily.

9.3.7 Plot of Tangents. During the inclinations, the tangents of the angles of inclination should be plotted

against the moments of the inclining weights. Varia- tion of the resulting plot from a straight line indicates that conditions are not favorable or that an error has been made, in which case a check should be made to determine the cause. Trials should be repeated until a satisfactory set of readings has been obtained. The plot of tangents will indicate only certain types of er- ror. For example, an error in measuring the length of one of the pendulums would be apparent since the tan- gent would be consistently larger or smaller than those from the other pendulums. Or, if a weight movement was measured incorrectly, the corresponding point on the plot would be out of line with the others, but this lack of alignment might be caused by an external force that was acting only at this particular inclination. On the other hand, if a single inclining weight is used and its weight is recorded incorrectly, this would not be ap- parent from the plot.

9.4 Inclining Experiment Report. This report con- sists of a recording of the observed data, the calcula- tions necessary to determine the displacement and CG at the time the ship was inclined, and the calculations made to arrive at the lightship condition by modifica- tions to the condition of the ship at the time of inclin- ing. It is advisable to record the basic data, such as the weight of each inclining weight and the distance it was moved and the lengths and deflections of each pendu- lum, rather than only the moments and tangents, in or- der to permit further checking in case any data appear later to be questionable. Actual tank soundings should be recorded, and determination of liquid weights from sounding tables shown-including trim corrections, if any. Table 5 shows a condensed summary of a typical inclining report.

The displacement and the longitudinal position of the ship's CG are calculated by one of the methods described in Section 8. If the midship draft readings indicate that the ship has hog or sag and the hydrostatic properties are being obtained from undeflected tables or curves, correction for this should be made as discussed in Sec- tion 8.9. The displacement thus obtained is multiplied by the ratio of the specific gravity of the water in which the ship was floating to 1.025 (the specific gravity of salt water) to obtain the displacement as inclined.

The plot of tangents is prepared showing the tangents measured at each of the three stations, plotted at the ap- propriate inclining moment for each inclination. Under ideal conditions, a straight line can be drawn through the plotted points. In most cases, however, some judg- ment must be applied in drawing the straight line that best represents the information plotted on the plot of tangents. If at any particular trial, one measurement of inclination does not agree well with the other two, it may be appropriate to disregard it. Or, it may be possible that one of the recorded moments does not represent the ac- tual moment acting at that time because of the effect of wind or current. It is usually necessary to draw the line that best represents the slope corresponding to the plot-

Page 76: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 63

ted points. This line does not necessarily pass through the origin, since this point carries no more weight than the others, and it is possible that some upsetting force was acting at the time that the initial reading was made. After this line has been established, its slope w - dltan 4 is divided by the displacement to find the metacentric height, which, from Section 9.1, is

w.d GM=-

Wtan d

After the metacentric height has been obtained, the vertical position of the CG in the inclined condition can be found and the characteristics of the lightship condi- tion developed as described in Section 9.1 (see Table 5).

Many ships have significant deflections in their in- clining condition. If the displacement and associated m v a l u e s have been computed from the deformed ship geometry, it may be appropriate to correct the results of the vertical CG to account for the deflection. In the example in Table 5, the metacentric height is measured from the undeflected baseline (i.e., the metacentric height includes the deflection). In some computational methods, the metacentric height will be reported from the local keel location. In either case, the location of the VCG for the undeflected ship will be different from the value computed here as the vertical location of the ship's lightship weight components will change as the deflection is removed. If the deflection is assumed to be parabolic, the adjustment will be from 67% of the de- flection for a uniform longitudinal weight distribution to 87% of the deflection for a more tapered longitudinal weight distribution. A conservative approach would be to use the lower adjustment for hog deflections and the larger one for sag deflections. These corrections can have a significant impact on the ability of vessels with low m s u c h as some containerships to meet stability requirements and thus limit the cargo carried.

It is desirable to record any major features of the lightship at the time of inclining, such as the weight and CG of any permanent ballast, for future reference.

9.5 Inclining in Air. When a small boat is to be in- clined, it is preferable, and may be more convenient, to perform the experiment in air rather than in water.

The boat is suspended by slings forward and aft which pass over a knife edge, and the slings are adjusted so that the base line used for calculations is parallel to the knife edge. The knife edge is supported by two scales: one forward and one aft. Inclining weights and pendu- lums are provided and used in the same manner as for inclining in water. See Fig. 62(b), where 4 is the inclina- tion due to the movement of the weight w through dis- tance A, G is the CG of the boat with inclining weight, and H is the height of knife edges above the keel.

The weight of the boat, W,, is obtained by adding the two scale readings, W, + W2. The distance x of the CG of the boat from the forward perpendicular is obtained by taking moments. Fig. 62(a):

Trim and Displacement Summary

Property Value Units Notes

Draft: Aft Perp. 1 5.428 1 m I from calculated waterline

Draft: Midship 1 5.708 1 m I from calculated waterline

Draft: Fwd Perp 1 6.151 1 m I from calculated waterline

Trim b/w 1 0.723F 1 m 1 from calculated waterline Perpendiculars

Table 5. Condensed inclining experiment report.

-

-

-

-

-

-

-

-

-

-

0.082H from calculated waterline

Draft a t LCF o n st. line between perps.

Corrected Draft 1 5.725 1 m 1 correction = 213 .'HoglSag at LCF

Displacement a t 1 19,527 1 MT 1 corrected for HoglSag LCF Draft

Sp Gr used for Displ Calc

Sp Gr of Flotation 1 1.0133 1 1 at 11.8 degrees C Water

Total Displacement 1 19,305 1 MT I in flotation water

Stability and Center of Gravity Summary

Property Value Units Notes

Virtual Metacentric 4.277 m from trials results Height (GMv)

Free-Surface 1 0.098 1 m 1 for liquids as inclined Correction

Transverse m GMv+ FS Metacentric Height 1 4'375 1 1 (GMt)

KMt Above 1 13.461 1 m I ;;;defined KMt a t trial Baseline I 381:5 m KMl- VCG Long'l Metacentric Height (GM1)

Moment t o Trim m-MT/ GMl x Displacement1 1 c m c m (100 x LBP)

Trimming Lever I 0.000F I m I not required

LCB 1 102.284F 1 m-AP I user-defined LCB at trial time

LCG 1 102.284F 1 m-AP 1 same as LCB

TCG 0.004P m GMv x ListIBeam at ref. waterline

Radius of Gyration and Rolling Constant

Period of Complete Roll I O 1 seconds I Apparent Radius of I 0 1 m I T x Sqr(GMv)/l.lOX Gyration

Rolling Constant l o l I T x Sqr(GMv)lBeam at ref. waterline

(continued)

Page 77: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

Table 5. Condensed inclining experiment report. (continued)

Weight Data I ID Weight VCG LCG TCG Description

Number I (MT) 1 (m-BL) I (m-AP) I (m-CL) I Location

74.463F 12.287P Row 10 port 1 1 19'849 1 1 1 ride

2 1 25 1 19.833 1 74.463F 1 12,2873 1 Row 10 stbd side I 3 1 30 1 20.044 1 61.082F 1 12.287P 1 Row 12 port side I 4 1 30 1 19.922 1 61.082F 1 12.2873 1 Row 12 stbd side I

Trial Data Entry/Summary I

5

6

SHIP AT THE TIME OF STABILITY TEST PLOT OF MOMENT vs TANGENT

Stbd Tan

30

30

Trial Number

1

Port Heel~ng Moment (ft-LT)

-7000 -6000 -5000 -4000 -3000 -2000 -1OOO/

' 1000 2000 3000 4000 5000 6000 7000

Stbd Heel~ng Moment (ft-LT)

-0.0075

19.995

19.922

Trial Time

1700

Port Tan

Condition 1-Lightship I

47.701F

47.701F

Total Moment (m-MT)

0

Item

12.287P

12.2878

Average Tangent

0

Ship in Condition 0 1 19,305 1 9 1 175,404 1 102.284F 1 1,974,628F 1 0.004P I 72P 1 1897 1

Row 13 port side

Row 13 stbd side

Trial GMt (m) -

I Weight (MT) I VCG (m-BL) I VMoment I LCG (m-AP) 1 L Moment I TCG (m-CL) I T Moment (m-MT) (m-MT) (m-MT )

FS Moment (m-MT )

Liquids as Inclined to be Deducted

Items to be reduced for Lightship 1

-4330

Dry Items to be Deducted for Lightship

Hence,

-272 -4991 64.044F

Dry Items to be Added for Lightship

Lightship (Condition 1)

x = A + B e WII(Wl + W.,) 9.6 Accuracy. Consideration of the subject of accu- racy will not only tend to improve the reliability of ex-

The accuracy of the scales may be checked by re- periments, whether in water or air, but may also result weighing the boat with the scales interchanged. in avoiding laborious refinements which do not have an

The location of VCG of the boat is obtained by the appreciable effect on the results. same process used when inclining in water, except that An error in measurement of pendulum length, pendu- the metacenter is located at the knife edge. Hence, the lum deflection, inclining weights, or weight movement will CG above the keel is, result in a proportional error in the metacentric height.

-17,433F

15

14,718

0.247P 1 6 7 p 1 - 1

12.5

10

188

150,186

54.175F

91.618F

813F

1,348,443F

0

0.0583

0

8483

-

0

Page 78: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 l lTY

! (Scale) w2 (Scale)

b- A 4 !

! ! ! I

! I CG (boat only) - - - - +$ - - - - ! !

Fig. 62 Inclining experiment in air.

The effect of inaccuracy in draft readings on displace- ment can be evaluated by considering the ship's tons per centimeter immersion. If the height of the metacenter is changing rapidly with a change in draft, as it often does at light displacements, an error in the height of the CG may result from inaccuracy in draft readings.

Errors in inventory appear as equal errors in dis- placement. The degree of accuracy used in locating CGs for items of the inventory depends on the weight of the item. Exact location of centers of light items is not nec- essary.

The contribution of different tanks to the total free- surface effect varies widely. If a tank has a width equal to about half the ship's beam, the shape of the surface should be determined accurately and precise methods used in finding the moment of inertia. For smaller tanks, less accurate methods may be used. There may be small tanks or tanks containing small quantities of liquid for which the movement of transference at the expected angle of heel is negligible compared to the moment of the inclining weights.

The proper degree of importance to be attached to any item can be evaluated by an approximate calcu- lation of its effect on the ship's displacement and the height of the CG.

9.7 Induced Rolling (Sallying). Rolling may be in- duced, for small ships, by sallying, in which a group of people moves across the deck in synchronism with the ship's natural period, or a vertical force may be applied on one side and suddenly released. For larger ships, a weight may be landed on one side and then lifted and

lowered several times, again in synchronism with the natural period of roll.

The period of roll may be found quite accurately by measuring the total elapsed time of a number of rolls.

From Section 3.7, the period of roll,

where C is the rolling constant and B is the beam of the ship.

Since the sally experiment gives the period of roll for a full cycle, using the as inclined m y i e l d s the roll- ing constant C. At subsequent times, sallying the ship in calm waters again yields a period of roll, and an ap- proximate m e a n be determined by using T,, C, and B in the above formula.

If at the time of inclining, the rolling constant, C, which is determined from the data, is out of line with constants found for previous similar ships, this would indicate a significant difference in construction of the new ship or a possible error in the experiment. The con- struction drawings and calculations should be closely checked.

IMO notes that it may be desirable to use the ship's pe- riod of roll as a means of approximately judging a ship's stability at sea, and in fact ship masters often do so as a check of loading condition calculations. Due care must be taken to discard readings departing significantly from the majority of other observations or being caused by forced oscillations due to the seaway.

Page 79: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STABlllTY

10 Submerged Equilibrium

10.1 Definition. A submerged submarine is in equi- librium when its weight is equal to the buoyancy of the total hull and when its CG is in the same longitudinal po- sition as its center of buoyancy so that it has zero trim. Actually, there may be very small differences between the weight and buoyancy and between the longitudinal positions of the centers of gravity and buoyancy. These differences are overcome by use of submarine's planes when underway in a submerged condition. When a sub- marine must be dead underwater to maintain silence, it is important that the weight equals the submerged buoy- ancy exactly.

On the surface, the weight of the submarine is lighter than its submerged displacement, and it is necessary to take on seawater in the ballast tanks to enable the sub- marine to submerge and be in a condition of equilibrium with zero trim while submerged. There are certain load- ing conditions on the surface that tend to make the sub- marine "heavy" or "light." In such cases, ballast water must be carried when on the surface so that upon com- plete flooding of the main ballast tanks, a submerged condition of equilibrium and zero trim will result. A sub- marine that is properly designed with respect to weight, buoyancy, and variable ballast tank capacity will always be in diving trim on the surface regardless of the actual load variations and will be able to successfully dive and be in equilibrium with zero trim in a submerged condi- tion with the main ballast tanks full of seawater.

Diving trim, diving ballast, and variable ballast are discussed in Sections 10.2, 10.3, and 10.4.

10.2 Items of Weight. The items of weight that are considered in studies of submerged equilibrium are il- lustrated in Figs. 63 and 64, and are defined as follows:

10.2.1 Submerged Displacement. The submerged displacement is the displacement of the entire envelope of the ship minus any free flooding spaces. The sub- merged displacement is fixed by geometry rather than by weight. For a given configuration of the ship, the sub- merged displacement will vary only with the density of the sea. Weight must be adjusted to conform to the sub- merged displacement.

10.2.2 Lightship. Lightship is calculated by taking the sum of the weights of the components making up the ship. This weight is fixed unless some alteration to the ship is made.

10.2.3 Lead. Solid ballast is the margin in the weight estimate. In submarine design, the volume of the submerged displacement is made larger than the antici- pated weight in the submerged condition by a generous margin. Some of this margin is usually needed to com- pensate for inaccuracies in the weight and displace- ment calculations or for unexpected installations and future modifications. When the ship is completed, any

unexpended margin must be installed as solid ballast to achieve submerged equilibrium. Solid ballast is part of lightship. Lead is most often used as solid ballast mate- rial because of its high density and because it causes few corrosion problems.

10.2.4 Load to Submerge. The load to submerge is the weight that must be added to the lightship-with lead to bring the ship to a condition of submerged equi- librium. Assuming no changes are made that affect the geometry or weight of the ship, the load to submerge will vary only when there is a change in the density of the seawater.

10.2.5 Normal Fuel-Oil Tanks. Fuel-oil storage tanks are fitted with a seawater-compensating system so that they are always full of oil or seawater.

10.2.6 Main Ballast Tanks. Main ballast tanks are tanks that are flooded to submerge and blown to sur- face. They are fitted with vents at the top, which are opened to flood the tanks, flood openings at the bottom, and air connections for blowing.

10.2.7 Fuel Ballast Tanks. Fuel ballast tanks are tanks that may be rigged either as normal fuel-oil tanks or as main ballast tanks. When used for fuel, they are compensating tanks and are handled in the same man- ner as the normal fuel tanks. After the oil in the fuel ballast tanks has been burned, they may be converted to serve as main ballast tanks, and thereafter flooded

1 SUBMERGED DISPLACEMENT

I LIGHT SHIP-WITH LEAD I LOAD TO SUBMERGE I

NORMAL CONDITION- SURFACED

I RESERVE

BUOYANCY VARIABLE BALLAST ,

VARIABLE LOAD- C

LIGHT SHIP

LEAD ' RESIDU~L I ! WATER I

I LIGHT snw

ABBREVIATIONS: NFO- NORMAL FUEL OIL WATER SEAL IN FBT FBT-FUEL BALLASTTANKS MBT- MAIN BALLAST TANKS

I I

Fig. 63 Weights, diesel-powered submarine.

MAXIMUM CONDITION-SURFACED RESERVE

BUOYANCY

! V A R I A ~ L E I DIVING L E A D 1 i BALLAST 1 I BALLAST

Page 80: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STABlllTY

SUBMERGED DISPLACEMENT

I RESERVE j ~uoyAycy

LIGHT SHIP-WITH LEAD

SURFACED CONDITION

VARIABLE L O A D A I

LOAD TO SUBMERGE

VARIABLE BALLAST RESIDUAL WATER

DIVING BALLAST

Fig. 64 Weights, nuclear-powered submarine.

upon submerging and blown upon surfacing. Conver- sion to main ballast tanks reduces the surface displace- ment and increases the reserve buoyancy until the ship is refueled. Fuel ballast tanks are not fitted on nucle- ar-powered submarines because of the small amount of fuel oil.

10.2.8 Residual Water. Residual water is the water in main ballast tanks and fuel ballast tanks, located be- low the top of the flood opening, which cannot be blown upon surfacing.

10.2.9 Water Seal in Fuel Ballast Tanks. The wa- ter seal in fuel ballast tanks is the layer of water above the top of the flood opening and below the bottom of the compensating water pipe which is maintained when the fuel ballast tanks are serving as normal fuel-oil tanks to prevent spilling of oil through the flood openings as the ship rolls. When the tank is nominally full of fuel, its contents, starting from the top, consist of fuel, water seal, and residual water.

10.2.10 Diving Ballast. Diving ballast is the term ap- plied to the water that is admitted to the ship upon diving and blown to bring the ship to the surface. When the fuel ballast tanks are used for fuel, the diving ballast is equal to the capacity of the main ballast tanks above the residual water. When fuel ballast tanks are rigged for main ballast, the diving ballast is equal to the capacity of the main bal- last tanks and the fuel ballast tanks above the residual water. Some submarines have a tank near amidships des- ignated as safety tank, which is blown upon surfacing and is considered to be a part of the diving ballast. The safety tank may be only partly filled in the submerged condition when the ship is heavily loaded. The safety tank, when in- stalled, is intended to be equal in volume and with about the same location as the topside conning area and may be blown while the submarine is submerged to regain buoy- ancy in case of topside damage.

10.2.11 Normal Condition-Surfaced. This is the condition of the displacement of the ship on the sur- face when fuel ballast tanks are rigged as main ballast tanks.

10.2.12 Maximum Condition-Surfaced. This is the condition of the displacement of the ship on the sur- face when fuel ballast tanks are rigged as normal fuel oil tanks.

10.2.13 Reserve Buoyancy. This is the condition of the displacement of the volume of the envelope of the ship above the waterline in the surfaced condition, mi- nus any free-flooding spaces.

10.2.14 Variable Load. Variable load includes such items as personnel and their effects, missiles, torpedoes, provisions, stores, cargo, passengers, potable water, re- serve feedwater, battery water, reserve reactor coolant, lubricating oil, oxygen, reserve hydraulic oil, contents of sanitary tanks and hovering tanks, and fuel oil. The variable load in the normal and maximum conditions is identical except that the oil in the fuel ballast tanks is included in the maximum condition.

10.2.15 Variable Ballast. Variable ballast is sea- water ballast that is adjusted continuously at sea to compensate for changes in variable load or in seawa- ter density. Variable water ballast is carried in forward and after tanks called trim tanks and in midship tanks called auxiliary tanks to permit adjustment of the lon- gitudinal moment as well as the weight. Some diesel- powered boats have variable fuel-oil tanks that are noncompensating tanks sized so that the weight of oil that they carry is approximately equal to the increase in weight that occurs when the oil in the compensating tanks is replaced by seawater. Burning oil from the vari- able fuel oil tanks so that the percentage remaining in these tanks is the same as the percentage remaining in the compensating tanks will tend to keep the weight of the contents of all oil tanks nearly constant. Variable fuel-oil tanks are piped so that oil may be transferred between them and the compensating tanks. This trans- fer involves an increase or decrease in the ship's weight as the compensating water is drawn or expelled from the normal fuel oil tanks. Because of this capability, they are considered as part of the variable ballast rather than part of the variable load.

10.3 Relationships Between Items of Weight. AS con- ditions for submerged equilibrium, each of the bars in Figs. 63 and 64 must represent the same weight and each must have its CG in the same longitudinal posi- tion. Two other equations are indicated by the vertical lines in Figs. 63 and 64. First, the reserve buoyancy is equal to the diving ballast, and the longitudinal posi- tions of their centroids must coincide. Also, the load to submerge (the difference between the submerged dis- placement and the lightship with lead) is equal to and has the same longitudinal CG as the sum of the variable load, variable ballast, residual water and diving ballast, and, in the maximum condition, the water seal in the fuel ballast tanks.

These necessary equalities are achieved by varia- tions in weight and longitudinal position of the CG of the lead ballast and the variable ballast. To conserve space, the variable ballast tanks are sized to accommo-

Page 81: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STABlllTY

date only the probable variation in the variable load plus the variation in the submerged displacement caused by changes in seawater density. Lead ballast, since it occu- pies less space per ton, is used for the required adjust- ment beyond the capacity of the variable ballast tanks. In general, changes in lead are made in the shipyard to compensate for changes in the lightship weight or in the volume of the submerged displacement, while the vari- able ballast is used to compensate for changes that oc- cur at sea.

10.4 Diving Trim. A submarine on the surface at sea is normally kept in diving trim, which means that the weights aboard are adjusted so that completion of flood- ing of the main ballast tanks, and any fuel ballast tanks rigged as main ballast tanks, will submerge the ship in a condition of equilibrium, with the ship's weight equal to the submerged displacement and the CG in the same longitudinal position as the center of buoyancy.

It can be seen from Fig. 63 that the weight and lon- gitudinal moment of the surfaced ship in diving trim must be equal to the difference between the figures for the submerged displacement and those for the reserve buoyancy, and that the reserve buoyancy corresponds, in weight and center, to the diving ballast. Since the vol- ume and moment of both the submerged displacement and the diving ballast depend only on the configuration of the ship, the surface drafts in diving trim, in either the normal or the maximum condition, are determined by the geometry of the hull.

Diving trim is maintained at sea by adjustment of water in the variable ballast tanks. Where variable fuel tanks are fitted, the oil in these tanks may also be ad- justed. While the submarine is submerged at very low speeds, variable ballast may be admitted, discharged, or transferred longitudinally until any fore-and-aft in- clination is eliminated and any appreciable tendency to rise or settle disappears. Between such experimen-

tal adjustments, the proper quantity and disposition of variable ballast is maintained by recording all changes in weight such as the replacement of fuel by seawater, ejection of trash or blowing of sanitary tanks, and mak- ing compensating changes in the variable ballast.

In addition to such gradual or minor changes in weight, there are large and abrupt changes that may oc- cur in the submerged condition, due to firing of weap- ons, which require immediate compensation. This is ac- complished by admitting a quantity of water, as part of the firing operation, equal to the weight of the weapon ejected.

The moment diagram, illustrated in Fig. 65 is a conve- nience in finding the change in weight that must be made in the variable ballast to compensate for a change in the variable load. If a weight, w, is added at some point, P, along the ship's length, reading the scales directly below point P will indicate the percentage of the weight, w, to be removed from the after trim and the auxiliaries, or from each of the two trim tanks, to compensate for the added weight. In this diagram, which is usually plotted below an inboard profile of the ship to which it applies, points A, B, and Care located, respectively, at the loca- tions of the centroids of the after trim, auxiliaries, and the forward trim tanks. The scales are constructed by dividing the distances between A and B, B and C, and A and C into 100 equal divisions.

As an example of the use of the moment diagram, if 1000 kg is added at point P in Fig. 65, there will be no change in the weight or in the longitudinal position of the CG of the ship if either:

1. Two hundred and ninety kg of water are blown from the forward trim tank and 710 kg are blown from the after trim tank, or

2. Five hundred and thirty kg are blown from the auxil- iary tanks and 470 kg are blown from the after trim tank.

PERCENT IN FORWARO TRIM 0 I 0 20 30 4 0 5 0 6 0 70 8 0

I 9 0

I 100

I I I I I I 1 -

I 100 9 0 80 70 60 50 4 0 30 20 10

PERCENT IN AFTER TRlM 0

C.G. OF AFTER TRlM

A

0 10 2 0 30 4 0 XI

AUXILIARIES 0

PERCENT IN AUXILIARIES 60 70 80 9 0 1 0 0 90 00 70 60 XI 4 0 30 20 10 0

Fig. 65 Moment diagram for submarine.

I I I 1 1 1 1 I" I 1 I I 1 I

I00 9 0 8 0 70 6 0 XI 4 0 30 20 10 0 I0 2 0 30 40 50 6 0 70 8 0 90 100

t PERCENT IN AFTER TRlM t PERCENT IN FORWARD TRIM f

Page 82: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

10.5 Equilibrium Conditions. Since the variable bal- last must be adjusted to compensate for changes in seawater density and for changes in the variable load, it is necessary to evaluate the magnitude of probable changes in both the weight and the longitudinal moment of these two items in order to select the proper size for the variable ballast tanks.

To ensure that diving trim can be achieved with any possible variation in seawater density and variable load, it would be necessary to develop the lightest possible condition in the heaviest seawater, the heaviest pos- sible condition in the lightest water, and conditions with maximum longitudinal moments in each direction in both heavy and light seawater and to make the vari- able ballast tanks large enough to compensate for these changes. This would result in very large variable ballast tanks on a ship having a limited amount of space. By using judgment to eliminate improbable extreme condi- tions, the variable ballast tanks can be held to a reason- able size.

In displacement calculations for surface ships, the seawater-specific volume is assumed to be 0.975 m3 per ton, and the normal variations from this figure are neg- ligible, since they would produce only a small change in draft. A small change in the displacement of a sub- merged submarine, such as 10 tons, would result in an unacceptable imbalance between weight and buoyancy. Since 10 tons is a small percentage of the submerged displacement, only a small change in the density of the seawater is required to produce such an unbalance.

The specific volume of seawater has been found to vary from 0.981 to 0.971 m"er ton. The extreme vari- ation in variable ballast to compensate for this effect would occur if the ship were to dive in light water, fill- ing the main ballast and fuel ballast tanks, and then pass, submerged, into heavy water. It is customary to assume that this extreme, or the opposite, will not oc- cur, and that the diving ballast is of the same density as that in which the submarine is operating. Under this assumption, the quantity of variable ballast needed for variation in seawater density is equal to the submerged displacement less the diving ballast, in cubic meters, multiplied by the change in density. As an example, if the submerged displacement is 4000 t and the diving ballast 500 t, the quantity of variable ballast needed to counteract this effect would be:

For the purpose of studying the additional variation in variable ballast necessary to compensate for changes in the variable load, calculations are made for a series of equilibrium conditions representing heavy loads in light water, light loads in heavy water, and heavy for- ward and heavy aft loadings in both light and heavy wa-

ter. In calculating the loads aboard for any particular condition, judgment and familiarity with operating pro- cedures are necessary in deciding on the quantities of the various items of variable load if very large variable ballast tanks are to be avoided.

The heavy forward and heavy aft conditions are not necessarily heavy. The term heavy forward, for exam- ple, means that loads in the forward end are heavy while those aft are light. In a ship that carries most of the vari- able load forward, the heavy aft condition might be quite light. In the heavy aft condition, the quantities of torpe- does and dry cargo, for example, would be assumed to be zero in the forward portion of the ship while a full load of such items would be assumed aft. As in the ease of the heavy and light conditions, it is advisable to inves- tigate two heavy forward and two heavy aft conditions: one with a large percentage of fuel aboard in which only the oil in the fuel ballast tanks at the heavy end has been burned and a condition occurring later when all oil in the fuel ballast tanks and in the normal fuel-oil tanks at the heavy end has been burned. Since the heavy for- ward and heavy aft conditions are not necessarily ei- ther heavy or light, calculations should be made for both heavy and light seawater.

The final result of the equilibrium-condition calcula- tions is the weight and longitudinal moment of the vari- able ballast to balance, which is the variable ballast re- quired under the assumed loading and seawater density to bring the ship to diving trim on the surface and to submerged equilibrium after diving.

As shown in Figs. 63 and 64, the variable ballast to balance can be established by subtracting the summa- tion of weight and longitudinal moment of the variable load from the figures for the load to submerge. Figures 63 and 64 also show the load to submerge may be found by deducting the weight and longitudinal moment of the lightship, with lead, from the figures for the submerged displacement. Two sets of values for the load to sub- merge are found by using figures for the submerged dis- placement at both 0.981 and 0.971 m3/t. The load to sub- merge in light water is used for the heavy conditions, the figures for heavy water used for the light conditions, and both are used for the heavy forward and heavy aft con- ditions. Two summations of variable load are required for the heavy forward and heavy aft conditions, so that the density of the diving ballast will correspond, in each ease, to that used for the submerged displacement.

10.6 The Equilibrium Polygon. The equilibrium poly- gon of a typical diesel-powered submarine, illustrated in Fig. 66, is a device for presenting graphically the en- velope of variation in weight and longitudinal moment which can be obtained by adjusting the variable ballast. In Fig. 66, the weight of variable ballast is plotted verti- cally and the longitudinal moment, about the transverse reference plane used for the equilibrium conditions, is plotted horizontally. Each side of the polygon represents the effect of filling one of the variable ballast tanks. The polygon is constructed by adding, algebraically and sue-

Page 83: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STABlllTY

cessively, the weights and moments of each of the vari- able ballast tanks, starting with the forwardmost tank and proceeding aft, then repeating the process starting with the aftermost and proceeding forward. Each sum- mation is plotted as in Fig. 66, where line OA represents the weight and moment developed as the forward trim tank is filled, line AB the effect of filling the forward variable fuel oil tank after the forward trim tank has been filled, and so forth until point E, representing the weight and moment of all the variable ballast tanks, is reached. The same point E is reached by a different route by plotting the various stages of the summation starting with the aftermost tank and proceeding forward.

The weight in each of the variable ballast tanks is taken to be equal to the net capacity of the tank at spe- cific volume of 0.975 m3 per ton. This volume is applied to the variable fuel tanks, even though they contain oil, because the transfer of 0.975 m3 of oil from a variable fuel tank to a normal fuel tank would force 0.975 m" sea- water, or l ton, overboard. The change in weight is as- sumed to occur at the location of the variable fuel tank, although there will also be a small change at the inde- terminate location of the normal fuel-oil tank to which the oil is transferred.

The exterior broken line in Fig. 66 shows the effect of considering the negative tank as part of the variable ballast. This tank, located forward of the center of buoy- ancy and normally empty in the surfaced and submerged conditions, is customarily filled just prior to diving in order to expedite the operation. This causes the weight of the ship to exceed the submerged displacement by a few tons and produces a down angle on the ship, both of which are favorable to rapid submerging. The nega- tive tank is blown when the ship reaches the ordered depth, restoring equilibrium. If necessary to meet very

light equilibrium conditions, the negative tank may be treated as part of the variable ballast, if the loss of the advantage of its normal function in the light condition is accepted.

The variable ballast can be adjusted so that its weight and moment correspond to the coordinates of any point within the polygon. Point P in Fig. 66, for example, can be reached by filling the after trim tank, moving from O to I; part of the after variable fuel tank, moving from I to S; then partially filling the auxiliaries, forward variable fuel tank, and the forward trim tank. Line SR is paral- lel to and not longer than HG, RQ is parallel to and not longer than GF, and QP is parallel to and not longer than FE. This is only one of many ways in which point P can be reached.

Fig. 67 is the polygon of Fig. 66, with the weight and moment of the variable ballast to balance for the vari- ous equilibrium conditions plotted. It is immediately apparent from Fig. 67 that the ship cannot be brought to submerged equilibrium in Condition Heavy No. 2 in light seawater, and that all other conditions can be met without the use of the negative tank, although there is but little margin in the ease of Light No. 2.

10.7 Adjustment of Lead and Variable Ballast Tankage. Under the conditions shown in Fig. 67, it is apparent from Figs. 63 and 64 that, if some lead ballast were re- moved, the amount of variable ballast required would be increased, and Condition Heavy No. 2 in Fig. 67 would move upward into the polygon. Also, if some lead were shifted aft with no change in the total amount, this point would move horizontally to the right into the polygon. All other points in Fig. 67 representing equilibrium con- ditions would, in either case, move the same distance and in the same direction. This shows that all points may be moved, as if they were plotted on a separate

AFTER YOYCYTS 0 FORWARD MOMENTS

Fig. 66 Equilibrum polygon for a subrnarne.

Page 84: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STABlllTY

AFTEH MOMENTS FORWARD MOMENTS

Fig. 67 Equilibrium polygon for Fig. 66, with required weight and moment plotted for several loading conditions.

piece of paper, either fore and aft or up and down, or a combination of both, by adjustment of the lead ballast. There may be, of course, some physical limitation on the adjustment of lead, or its removal may be precluded by considerations of transverse stability.

In the case shown in Fig. 67, it is apparent that the spread of the points representing the equilibrium condi- tions is beyond the capacity of the variable water and variable fuel tanks, but that the constellation of points can be embraced by the polygon that includes the effect of the negative tank. If the ship were already built, the loss of complete effectiveness of the negative tank in div- ing in the light condition in heavy water would probably be accepted and Light No. 2 allowed to move into the area representing the effect of the negative tank. Other- wise, it would be prudent to consider an increase in the size of the polygon. In this situation, the polygon is use- ful in deciding which tank or tanks should be enlarged. It is apparent from Fig. 67 that no improvement would result from increasing the capacity of the after variable fuel tank or the after trim tank, since this would only extend the polygon to the left. The greatest gain, per ton increase in capacity, would be obtained by increasing the size of the auxiliary tanks, but if this were not fea- sible, increasing either the forward trim or the forward variable fuel tank would be effective.

10.8 Stability in Depth. While the forces of weight and buoyancy can be brought very nearly to equilibrium when a submarine is submerged, most submarines have no inherent stability with respect to depth since, as the

ship settles or rises, no force is generated to return the ship to the original level. A situation may exist in which the water at greater depths may be appreciably denser than that near the surface because of differences in temperature and salinity, which will enable the subma- rine to rest on the interface if its weight is greater than its displacement in the less dense water but less than its displacement in the denser water. Otherwise, unless some force is applied, as by the planes or a hovering sys- tem, most submarines would eventually either rise to the surface or settle to the bottom.

On the normal submarine, the pressure of the sea- water on the hull tends to produce an unstable condi- tion. The loss of buoyancy due to compression of the hull as the ship settles exceeds the gain in buoyancy due to compression of the seawater and the resulting slight increase in its density. The net result is that buoyancy is decreased as the ship settles and increased as the ship rises. The effect of sea pressure would be aggravated if there were a partially filled tank open to the sea, since the air therein would expand and compress readily with changes in depth, expelling water when the ship was rising or admitting water settling.

On some very rigid hulls, the effect is reversed, since the effect of compression of the hull is less than the ef- fect of the compression of the seawater in increasing the water density. This results in a small gain in buoyancy as the ship settles, a small loss when it rises, and hence a minor stabilizing effect.

Page 85: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 l lTY

11 The Trim Dive

11.1 Basic Principles. The trim dive is an experimen- tal determination of the weight and longitudinal mo- ment of the load to submerge, as defined in Section 10.

Theoretically, the load to submerge could be ob- tained, as illustrated in Figs. 63 and 64 by deducting the lightship with lead, as determined from the inclining ex- periment, from the calculated figures for the submerged displacement. The submerged displacement, however, cannot be calculated accurately because of numerous topside appendages. Also, the load to submerge, deter- mined in this manner, would represent a small differ- ence between two large quantities, and therefore be subjected to a larger error than if it were determined directly. It is therefore customary to find the load to submerge experimentally by an inventory of all weights aboard that comprise the load to submerge, taken while the ship is in submerged equilibrium.

The load to submerge is used as the basis for cal- culating the variable ballast to balance in the various equilibrium conditions, which, in turn, determines the optimum weight and disposition of the lead ballast.

11.2 Conducting the Trim Dive. The ship is completely submerged in an area that is free from strong currents and sharp density gradients. The variable ballast is carefully adjusted to bring the ship to submerged equi- librium. The ship is held at rest long enough to ensure that there is no fore and aft inclination and no appre- ciable tendency to rise or settle.

While the ship is in submerged equilibrium, a sample of seawater is taken, preferably from a circulating sys- tem in operation, and the density determined.

An inventory is taken of the weight and longitudinal moment of all items aboard (other than lead ballast) that are not part of the lightship weight. As in the case of the inclining experiment, this inventory must be based on a comprehensive definition of the lightship condition. The total weight and moment resulting from this inventory are the load to submerge and its longitudinal moment at the seawater density observed concurrently with the inventory.

11.3 Report of the Trim Dive. The calculations made in the report of the trim dive involve converting the load to submerge at the density of the seawater in which the ship was submerged to its values at specific volumes of 0.981, 0.975, and 0.971 m3/t. As mentioned in Section 10, these values represent the variation in seawater-spe- cific volume and are used in the equilibrium conditions, where small variations in specific volume are important. The value of 0.975 m3/t is used in stability calculations, as in the case of surface ships.

When the inclining experiment and the trim dive have been completed, the weights and longitudinal moments of the lightship with lead and of the load to submerge are known. Theoretically, the sum of these two items should correspond to the volumetric calculations for the submerged displacement, but minor discrepancies are to be expected due to the inaccuracies involved in each of the three items. It is customary to regard the submerged displacement from the inclining experiment and trim dive as being more accurate than that obtained from the volumetric calculations.

The values of the load to submerge at the various sea- water-specific volumes are obtained as follows:

1. The submerged displacement and its longitudinal moment at the time of the trim dive are obtained by add- ing the weights and moments of the lightship with lead from the inclining experiment and the load to submerge from the trim dive.

2. The submerged displacement and its longitudinal moment at 0.981, 0.975, or 0.971 m3/t are found by mul- tiplying the weight and moment obtained in step l by the ratios of 0.981, 0.975, and 0.971 m"/t to the specific volume of the outside seawater.

3. The load to submerge and its longitudinal moment at specific volume of 0.981, 0.975, and 0.971 m3/t are ob- tained by subtracting the figures for the lightship with lead from the submerged displacement at those specific volumes.

Page 86: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY

12 Methods of Improving Stability, Drafts, and List

12.1 Changes in Form. In the early stages of design, variations in the ship's form, usually changes in beam or depth, may be used effectively to obtain optimum sta- bility and drafts. The effects of changes in form on sta- bility are discussed in detail in Section 4. Draft can be varied by changes in beam or fullness, and trim can be varied by increasing the fullness at one end of the ship and decreasing it at the other, thus moving the center of buoyancy in the desired direction. This must be done with caution, however, as it may have a serious adverse effect on resistance. Moreover, in cargo ships, and espe- cially in tankers and ore carriers, the benefit is partially offset by the resulting movement, in the same direction, of the CG of the total cargo space.

When extensive topside weight is to be added to an existing ship, drafts and stability may be improved by the installation of blisters. These have the effect of in- creasing the ship's beam and may be used to adjust trim by moving the center of buoyancy longitudinally. There is also at least one case in which unsymmetrical blisters have been used successfully to remove an inherent list, with improvement in drafts and stability.

The full effect of intact stability improvement due to widening the ship's beam by blisters or other means is not a complete measure of the net improvements in the case of damage stability. Where damage stability is governing in the ship's required initial stability, the net stability improvement must be considered rather than the improvement in intact stability alone.

12.2 Adjustment of Load. The effectiveness of adjust- ing the load, as a means of improving stability, correct- ing trim and minimizing list, depends upon the ratio of the load to the total weight of the ship and the freedom that the operator has in the distribution of loads within the ship. One extreme is the cargo ship, loading mixed cargo of various densities, which can be stowed to ob- tain a wide variation in the ship's CG. The other extreme is a ship such as a tug, which carries very little load and each item is loaded in a specific location.

In the former case, it is advisable to calculate the ef- fect of each proposed loading on the stability and drafts and to ensure that the cargo is stowed symmetrically. The forward draft should be adequate to ensure against pounding in a seaway, and the after draft adequate to provide sufficient propeller immersion. The expected variation in the consumable load during the voyage should be taken into account. Any prescribed draft limi- tation should be observed. Modern ships' loading soft- ware programs include the capability to monitor these considerations and to propose ballasting to maximize stability within these constraints.

12.3 Permanent Ballast. Permanent ballast, judi- ciously located, can often be used to improve stability

or trim, or to remove list. When installed primarily to improve one of these characteristics, ballast can often be used to improve the others as well.

Permanent topside or 'tween-deck ballast has also been used to decrease the metacentric height of ships found to be too "stiff" for satisfactory operation.

When the object is to increase stability, the use of a dense material, such as lead, iron, or concrete made with a dense aggregate, is usually appropriate as this will permit the installation of ballast at a lower CG than a less dense material, and therefore with a greater ef- fect, per ton of ballast, in lowering the ship's CG.

If solid ballast is to be located in fuel tanks, there is a marked advantage in using metallic ballast, preferably lead. The effective weight of such ballast is its weight in oil. A dense material will displace less oil per ton of ef- fective weight than a less dense type.

When ballast is used to improve drafts or remove list, and lowering of the ship's CG is not required, ordi- nary concrete may be effective. In many cases, the bal- last may be installed on several levels, in relatively thin layers, and the additional volume occupied by the less dense ballast is not important.

The use of liquid ballast, preferably fresh water with a rust inhibitor, is often effective, and has the advan- tages of a lower material cost and easy removal. In some eases, however, considerable expense may be involved in reinforcing the structure to contain the ballast when, for example, it is located in a cargo hold and the strength of the deck above is not adequate to withstand the head imposed when the ship rolls or heels to a large angle. When permanent liquid ballast is used to improve sta- bility, its location and effectiveness in case of underwa- ter hull damage must be considered.

The use of permanent ballast in surface ships occurs most often in ships that are being converted for some purpose other than that for which they were originally designed. A common example is the naval auxiliary which has been converted from a merchant type, car- ries little or no cargo, and has had considerable topside weight added.

There is a common feeling that the installation of permanent ballast in new designs is undesirable, and an indication that the design is less than optimal. This is not necessarily the case. The outstanding example of essential ballast installation is the case of the subma- rine, as previously discussed.

12.4 Weight Removal. Removal of topside weight is the most effective method of improving stability be- cause it will not only lower the ship's CG, but will also, in most cases, cause the center of buoyancy to move farther to the low side as the ship is inclined as a result of lighter draft. Examination of the slope of the cross

Page 87: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

74 INTACT STAB1 l lTY

curves will indicate the increase in righting arm that will be obtained because of the decrease in displace- ment, to which will be added the increase in righting arm produced by lowering of the ship's CG.

Removals of topside weight are particularly advan- tageous in improving the stability of a ship that has little freeboard, when the addition of low weight may be counterproductive in increasing the righting arms at significant angles of heel. In addition, weight removal is generally beneficial in improving the ship's resistance to foundering after underwater damage.

12.5 Loading Instructions. The studies of stability and trim made during the design stage may indicate that certain limitations must be placed on the ship's loading in order to obtain satisfactory characteristics. These must be clearly transmitted to the operating personnel.

This is particularly important in the case of stability problems because unsatisfactory stability, unlike unfa- vorable drafts or list, is an invisible source of potential danger, and the ship may be operated unknowingly in an unsafe condition.

These limitations, together with information regard- ing the drafts and stability resulting from a variety of probable or possible loading conditions are usually presented to the owner by the designer in the form of a stability booklet, which, in the United States, must be approved by the USCG for all merchant ships before a certificate of approval is issued. This booklet also usu- ally contains the information regarding required GM, re- ferred to in Section 3 and in Tagg (2010). The U.S. Navy provides similar loading instructions to its operators.

Stability When Grounded

The problem of the stability of grounded ships is lim- ited, in general, to the dry docking of ships of relatively small w a n d to salvage operations.

13.1 Stability During Dry Docking. When a vessel en- ters dry dock, it generally has a trim; hence, the keel makes an angle with the keel blocks. Frequently, the keel blocks of dry docks are given a slight slope to fa- cilitate dock drainage and also to make it easier to dock ships which trim by the stern.

As the water level falls, due to the pumping out of the dry dock, the keel of the ship comes into contact with the keel blocks. Vessels are usually trimmed by the stern, in which case the after part of the keel will ordinarily touch first. The weight that is supported by the keel blocks at any subsequent time is the difference between the displacement when fully water borne and the displacement to the waterline in the aground condi- tion. As the water continues to recede, the slope of the keel gradually approaches the slope of the keel blocks. The force exerted by the keel blocks has the same effect on the of the ship as would the removal of a cor- responding weight from a position at the vessel's keel. If the bilge blocks are considered as contributing to the ship's stability, the condition of minimum stability oc- curs when the keel blocks first make contact throughout the entire length of the keel just before the bilge blocks are hauled into place.

If the weight on the keel blocks materially reduces the effective GM, a ship with a small w fully waterborne may become unstable and list to an appreciable angle before the bilge blocks can be hauled. To avoid this situ- ation, a ship with little m s h o u l d be trimmed as nearly as practicable to the slope of the keel blocks before it enters the dry dock. The vessel can then be lowered in such a manner that all of the blocks will touch the keel at approximately the same time. The bilge blocks and the

shores to the dock sides can be placed in their proper positions before much weight has been placed on the blocks, and hence before appreciable loss in effective -

GM has taken place. Often, it is not practicable to place the vessel in the condition indicated. That event must be handled with special care during the critical interval between the grounding at one end of the ship and the hauling of bilge blocks. If the trim is not excessive, the ship can usually be held upright by judicious use of side shores until the keel bears fore and aft, when the bilge blocks are hauled.

When a ship has just landed on the keel blocks, part of its weight is borne by the blocks and part of it is water borne. Consider the ship in Fig. 68 resting on the keel blocks with the waterline at WILl, which is below the waterborne waterline WL.

P = upward force exerted by keel blocks M1 = metacenter at waterline W,Ll G = CG of the ship A, = displacement of the grounded ship to waterline

WlLl

Fig. 68 Stability in dry dock.

Page 88: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

74 INTACT STAB1 l lTY

curves will indicate the increase in righting arm that will be obtained because of the decrease in displace- ment, to which will be added the increase in righting arm produced by lowering of the ship's CG.

Removals of topside weight are particularly advan- tageous in improving the stability of a ship that has little freeboard, when the addition of low weight may be counterproductive in increasing the righting arms at significant angles of heel. In addition, weight removal is generally beneficial in improving the ship's resistance to foundering after underwater damage.

12.5 Loading Instructions. The studies of stability and trim made during the design stage may indicate that certain limitations must be placed on the ship's loading in order to obtain satisfactory characteristics. These must be clearly transmitted to the operating personnel.

This is particularly important in the case of stability problems because unsatisfactory stability, unlike unfa- vorable drafts or list, is an invisible source of potential danger, and the ship may be operated unknowingly in an unsafe condition.

These limitations, together with information regard- ing the drafts and stability resulting from a variety of probable or possible loading conditions are usually presented to the owner by the designer in the form of a stability booklet, which, in the United States, must be approved by the USCG for all merchant ships before a certificate of approval is issued. This booklet also usu- ally contains the information regarding required GM, re- ferred to in Section 3 and in Tagg (2010). The U.S. Navy provides similar loading instructions to its operators.

Stability When Grounded

The problem of the stability of grounded ships is lim- ited, in general, to the dry docking of ships of relatively small w a n d to salvage operations.

13.1 Stability During Dry Docking. When a vessel en- ters dry dock, it generally has a trim; hence, the keel makes an angle with the keel blocks. Frequently, the keel blocks of dry docks are given a slight slope to fa- cilitate dock drainage and also to make it easier to dock ships which trim by the stern.

As the water level falls, due to the pumping out of the dry dock, the keel of the ship comes into contact with the keel blocks. Vessels are usually trimmed by the stern, in which case the after part of the keel will ordinarily touch first. The weight that is supported by the keel blocks at any subsequent time is the difference between the displacement when fully water borne and the displacement to the waterline in the aground condi- tion. As the water continues to recede, the slope of the keel gradually approaches the slope of the keel blocks. The force exerted by the keel blocks has the same effect on the of the ship as would the removal of a cor- responding weight from a position at the vessel's keel. If the bilge blocks are considered as contributing to the ship's stability, the condition of minimum stability oc- curs when the keel blocks first make contact throughout the entire length of the keel just before the bilge blocks are hauled into place.

If the weight on the keel blocks materially reduces the effective GM, a ship with a small w fully waterborne may become unstable and list to an appreciable angle before the bilge blocks can be hauled. To avoid this situ- ation, a ship with little m s h o u l d be trimmed as nearly as practicable to the slope of the keel blocks before it enters the dry dock. The vessel can then be lowered in such a manner that all of the blocks will touch the keel at approximately the same time. The bilge blocks and the

shores to the dock sides can be placed in their proper positions before much weight has been placed on the blocks, and hence before appreciable loss in effective -

GM has taken place. Often, it is not practicable to place the vessel in the condition indicated. That event must be handled with special care during the critical interval between the grounding at one end of the ship and the hauling of bilge blocks. If the trim is not excessive, the ship can usually be held upright by judicious use of side shores until the keel bears fore and aft, when the bilge blocks are hauled.

When a ship has just landed on the keel blocks, part of its weight is borne by the blocks and part of it is water borne. Consider the ship in Fig. 68 resting on the keel blocks with the waterline at WILl, which is below the waterborne waterline WL.

P = upward force exerted by keel blocks M1 = metacenter at waterline W,Ll G = CG of the ship A, = displacement of the grounded ship to waterline

WlLl

Fig. 68 Stability in dry dock.

Page 89: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 75

G,fll = virtual metacentric height of ship at waterline WlL,

If the force P is considered as a weight removed from the ship at point K, the virtual CG of the ship rises to point G,, termed the virtual C.G. of the grounded ship, and the virtual weight is W - P.

Taking moments of the real and virtual removed weights about the keel, the virtual height of the CG is given by

KG, = - E . W - P . 0 - KG.W

-

W - P W - P

The virtual metacentric height of the grounded ship, based on displacement, A, at W,Ll is

- - E2.w G,Ml=KM, --

W-P

The righting moment at a small angle of heel, 64, is given by

RM = AIGfll sin 64 = (W - P)GJIl sin 64

The righting moment may also be defined in terms of the full, ungrounded displacement, A. The moment given by such an expression must still be equal to that given by the above, therefore, the corresponding expres- sion for effective metacentric height, a, is obtained by equating the expression for righting moment in terms of A and ml,, equating RM = ~m,, sin 64 to the RHS of the above expression for RM and rearranging,

- -

p p PKM - W

PKM, GMle = GUM, - 2 = GUMl -- A .

If the value of G a l is low or negative, there is danger that the ship will assume a list before the bilge blocks are hauled unless it is otherwise supported.

When a vessel enters dry dock with a trim, generally the most critical stage of docking is at the time when the vessel's keel comes in contact with the keel blocks throughout its length. To investigate this special condi- tion and to determine the force P at this stage, the fol- lowing calculations may be made:

1. Calculate the displacement and LCB of the fully wa- terborne ship for the trim at which it enters dry dock.

2. If the vessel is trimming by the stern and hence the after end of the keel touches the keel blocks first, calculate the moment of weight about the after end of the keel. This will be the product of the waterborne dis- placement and the horizontal distance from the after keel block to a point vertically below the LCB.

3. Draw several waterlines at an inclination to the keel equal to the slope of the dry dock blocks represent- ing different displacements both greater and less than the waterborne displacement. Calculate the displace- ment and moment of buoyancy about the after end of the keel for each of these sloping waterlines. If the keel

blocks should be level, then these waterlines will be par- allel to the keel.

4. On graph paper, with displacements as abscissas and the moments as ordinates, plot a curve of moment of buoyancy about the after end of the keel. Also plot the moment of weight on the same graph (this will be a straight horizontal line). The point at which the mo- ment-of-buoyancy curve crosses the moment-of-weight line gives the displacement at the time the keel first bears fore and aft on the keel blocks. The difference be- tween this displacement and the fully waterborne dis- placement is the weight supported by the keel blocking (i.e., force P in Fig. 68). Having determined P, one can calculate the virtual metacentric height from equation (39) above. Current salvage engineering software per- mits automation of this procedure.

The problem of stability during docking is often elimi- nated entirely, especially for large heavy ships, by provid- ing a cradle of bilge blocks and cribbing, shaped to con- form to the contour of the bottom of the ship. This requires accurate control of the fore-and-aft position of the ship, which is not necessary if sliding bilge blocks are used.

13.2 Stability Stranded. When a ship is stranded on a fairly flat bottom, there is usually no issue of transverse stability. The possibility of a stranded ship capsizing is greatest as the result of ebbing tide if the ship is grounded on a bottom that offers no restraint to heeling. This would be the case if grounding occurs on a peak which was con- siderably higher than the surrounding bottom, as illus- trated in Fig. ll(b). When a ship is aground in this man- ner, the heel increases as the tide ebbs. The attitude of the ship would always be such that the moment of buoyancy equals the moment of weight, or W . b in Fig. ll(b) would be equal to (W - R) a. If the ship were to capsize and, as a result, lose contact with the point of support, the reac- tionR would be zero. Since W . b = (W - R) a, when R = 0, a and b would be equal, or G would be directly above B,. The situation would be the same if the ship were heeled to its range of positive stability, and the angle of heel could be determined from the statical stability curve cor- responding to the ship's weight and the position of its CG. From this, it is seen that a stranded ship will not capsize, in the absence of other upsetting forces, until it reaches an angle of heel equivalent to its range of positive stabil- ity if it were afloat.

The following conclusions may be drawn:

1. It is unlikely that a stranded ship will capsize un- less its range of positive stability is much less than usual. Unless impaled, the ship would slide from the point of support when the tangent of the angle which the bottom of the ship makes with the horizontal exceeds the coef- ficient of static friction between the bottom of the ship and the support. This angle is generally much smaller than the range of positive stability.

2. If the angle of inclination approached the range of positive stability, only a relatively small strain would be required to free the ship as the reaction of the support

Page 90: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

76 INTACT STAB1 llTY

approached zero. The point of application of the tow- of support, and estimating the displacement below this line should be low, since only a small heeling moment waterline. If this displacement exceeds the weight of the would be required to capsize the ship. ship, the range of positive stability will not be reached.

3. The likelihood of capsizing with the expected Current salvage engineering software makes it possible variation in tide can be evaluated by assuming the ship to directly evaluate this situation by computing the sta- heeled to its range of positive stability, drawing the wa- bility for a range of waterlines. terline at the lowest expected level relative to the point

Advanced Marine Vehicles

Various types of advanced marine vehicles are able to attain unusually high speeds by means of special de- sign features or devices, such as dynamic lift, fan-gener- ated lift, narrow, twin hulls, or wing-in-ground effects. A number of these are shown diagrammatically in Fig. 69 and more information may be found in Lamb (2004). For such craft, separate treatment is required when dealing with intact stability for low-speed-displacement and high-speed modes of operation. Multihull vehicles, such as catamarans including SWATHS, trimarans, and pentamarans, may generate additional stability at high speeds but in general operate in a displacement mode.

When operating at low speeds in the hullborne or displacement mode, their stability problems are similar to those of conventional displacement ships. In general, high-performance craft are exposed to the same haz- ards as conventional ships, and thus the requirements for adequate stability and buoyancy to resist the effects of these hazards are similar. Hence, the same general approach to evaluating stability is used for the advanced vehicles. Additionally, special problems peculiar to some of these vehicles may arise, all of which should

be considered in the stability and buoyancy analysis. Examples of such problems are low freeboard, with the hazard of shipping and trapping seawater, large CG shifts because of extension or retraction of foils, weight constraints that might limit the number of watertight bulkheads, thin shell structure which is susceptible to damage from impact with waves or debris at high speeds or from collisions, and for the low waterplane catamaran types, the potential for large unsymmetrical flooding. Although the submerged foils of hydrofoil craft generally have a small effect on static stability, they do have a favorable effect on damping of rolling motion.

14.1 Stability Criteria and Hazards. At the high speeds for which advanced marine vehicles are designed, how- ever, their stability characteristics and the means of achieving stability are very different from those of con- ventional ships and vary considerably from one type to another. In November 1977, IMO published a proposed code of safety for submerged foilborne craft (IMO, 1977), but only very general guidance on the subject of stability for high-speed operation was included. In 1994, the IMO released the first High Speed Craft (HSC) code

Fig. 69 Types of advanced marine vehicles

Page 91: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

76 INTACT STAB1 llTY

approached zero. The point of application of the tow- of support, and estimating the displacement below this line should be low, since only a small heeling moment waterline. If this displacement exceeds the weight of the would be required to capsize the ship. ship, the range of positive stability will not be reached.

3. The likelihood of capsizing with the expected Current salvage engineering software makes it possible variation in tide can be evaluated by assuming the ship to directly evaluate this situation by computing the sta- heeled to its range of positive stability, drawing the wa- bility for a range of waterlines. terline at the lowest expected level relative to the point

Advanced Marine Vehicles

Various types of advanced marine vehicles are able to attain unusually high speeds by means of special de- sign features or devices, such as dynamic lift, fan-gener- ated lift, narrow, twin hulls, or wing-in-ground effects. A number of these are shown diagrammatically in Fig. 69 and more information may be found in Lamb (2004). For such craft, separate treatment is required when dealing with intact stability for low-speed-displacement and high-speed modes of operation. Multihull vehicles, such as catamarans including SWATHS, trimarans, and pentamarans, may generate additional stability at high speeds but in general operate in a displacement mode.

When operating at low speeds in the hullborne or displacement mode, their stability problems are similar to those of conventional displacement ships. In general, high-performance craft are exposed to the same haz- ards as conventional ships, and thus the requirements for adequate stability and buoyancy to resist the effects of these hazards are similar. Hence, the same general approach to evaluating stability is used for the advanced vehicles. Additionally, special problems peculiar to some of these vehicles may arise, all of which should

be considered in the stability and buoyancy analysis. Examples of such problems are low freeboard, with the hazard of shipping and trapping seawater, large CG shifts because of extension or retraction of foils, weight constraints that might limit the number of watertight bulkheads, thin shell structure which is susceptible to damage from impact with waves or debris at high speeds or from collisions, and for the low waterplane catamaran types, the potential for large unsymmetrical flooding. Although the submerged foils of hydrofoil craft generally have a small effect on static stability, they do have a favorable effect on damping of rolling motion.

14.1 Stability Criteria and Hazards. At the high speeds for which advanced marine vehicles are designed, how- ever, their stability characteristics and the means of achieving stability are very different from those of con- ventional ships and vary considerably from one type to another. In November 1977, IMO published a proposed code of safety for submerged foilborne craft (IMO, 1977), but only very general guidance on the subject of stability for high-speed operation was included. In 1994, the IMO released the first High Speed Craft (HSC) code

Fig. 69 Types of advanced marine vehicles

Page 92: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 77

and subsequently updated it to reflect experience in this rapidly expanding fleet in 2000 (IMO, 2000). This code also covers the requirements in displacement and tran- sition modes.

Special concern in the HSC code is paid to the effects of passenger crowding and high-speed turns. Explicit criteria are provided for these situations. These crite- ria can be met in terms of a static stability curve, or a "dynamic" stability curve. In this context, the dynamic stability curve is the integral of the static stability curve from zero to the angle of interest as defined in Section 4.10 and shown in Fig. 32. As noted by Courser (2003), it is rather ambiguously defined in the HSC code.

Stability hazards to which high speed craft identified in the HSC code include:

Directional instability, which is often coupled with roll and pitch instabilities;

Broaching and bow diving in following seas at speeds near to wave speed, applicable to most types;

Bow diving of planing monohulls and catamarans due to dynamic loss of longitudinal stability in relatively calm seas;

Reduction in transverse stability with increasing speed of monohulls;

Porpoising of planing monohulls, consisting of cou- pled pitch and heave oscillations, which can become violent;

Chine tripping, being a phenomenon of planing monohulls occurring when the immersion of a chine generates a strong capsizing moment;

Plough-in of air-cushion vehicles, either longitudinal or transverse, as a result of bow or side skirt tuck-under or sudden collapse of skirt geometry, which, in extreme cases, can result in capsize;

Pitch instability of SWATH craft due to the hydrody- namic moment developed as a result of the water flow over the submerged lower hulls;

Reduction in effective metacentric height (roll stiff- ness) of surface effect ships (SES) in high-speed turns compared to that on a straight course, which can result in sudden increases in heel angle and/or coupled roll and pitch oscillations; and

Resonant rolling of SES in beam seas, which, in ex- treme cases, can result in capsize.

The HSC code requires numerical simulation and model and/or full-scale testing of the stability of the craft. In general, the above effects are issues addressed in dynamic analysis of the motions of the vessel (Faltin- sen, 2006).

Some notes on stability for the various types follow. The stability problems of high performance craft are of- ten a combined result of the geometric and high-speed characteristics of such craft.

14.1.1 Planing Hull. Compared with other types of advanced marine vehicles, the planing boat is well known and has been in wide use for very many years. For chine boats, the question of transverse stability has,

over the years, been given very limited treatment. How- ever, the fact that these craft were always beamy, with very large transverse GM when planing, seems to have ensured satisfactory stability in service. For the deep-V, high deadrise planing hulls, it has been found that lon- gitudinal strakes or spray strips must be used to obtain satisfactory transverse stability. Planing hulls tend to be considerably stiffer in roll when operating at high speeds than when at rest. This is particularly true of low deadrise hull forms. As deadrise is increased, the stabil- ity is decreased. Longitudinal instability can result in porpoising as noted above.

14.1.2 Catamarans. Catamarans have higher wa- terplane inertias than monohulls of similar size. As a con- sequence, m i s higher and results in excellent stability at low angles of heel. This stiffness can lead to high roll accelerations. As is often the case with high ~ v e s s e l s , the peak of the righting arm curve usually occurs at rela- tively low angles. The HSC code accounts for this in its requirements that limit the area contributing to the sta- bility to the angle at which the maximum occurs (or downflooding or 30 degrees if they govern).

14.1.3 Trimarans and Pentarnarans. The main role of the side hulls in these vessels is to provide sta- bility while still allowing the main hull to be slender enough to result in low wave-making resistance. The stability performance is much like that of a catamaran, although there is greater flexibility in distributing the contributions to the waterplane inertia. A further dis- cussion of the trimaran stability can be found in An- drews and Zhang (1996).

14.1.4 Small Waterplane Area Hull. The SWATH is essentially a displacement catamaran having much of its buoyant volume deeply submerged and only slender struts extending through the water surface to support the main working deck. The small waterplane area fea- tured in its design serves to decrease wave drag and to provide good seakeeping qualities. Intact stability must be provided in the same way as for a conventional ship. Ballast tanks are usually provided to allow trim and draft to be carefully controlled and to provide damage control in the event of loss of buoyancy in any compart- ments. Some SWATH designs include horizontal control surfaces to provide additional trim control and stabi- lization at high speed. Since longitudinal metacentric height, m, will generally be of a similar magnitude to transverse metacentric height, m, stabilization of heave, pitch, and roll are possible. This is unlike a con- ventional monohull for which only roll stabilization is usually feasible.

14.1.5 Hydrofoil Craft. In the case of hydrofoil craft, intact stability while foilborne is attained entirely by hydrodynamic means, although there may be some incidental aerodynamic effects. Surface-piercing foils make it possible to attain stability passively (i.e., with- out automatic controls) since heel to one side will result in increased immersed foil area and hence lift on that side-and less on the opposite side-which produces a

Page 93: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

righting moment. However, some adjustment of one or more of the fixed foils is usually possible, particularly to control running trim. Sometimes, electronic stability augmentation systems are provided, actuating trailing edge flaps, to improve the ride quality.

For hydrofoil craft with fully submerged foils, an au- tomatic control system or autopilot is essential, not only for transverse stability but for attaining proper eleva- tion and trim relative to the water surface. The control system operates either by changing the angles of inci- dence of the foils individually or by adjusting trailing edge flaps. Such automatic systems can also control to some extent the wave-induced motions of pitch, heave, and roll in moderate seas.

Practice indicates that the nondimensional foilborne longitudinal metacentric height ( ~ I L , . ) for a surface- piercing foil craft should lie between 3.5 and 5.5, where L,. is the longitudinal distance between foils. Also, the yawing moment contribution of the aft foils should be at least 20% greater than the moment contribution of the forward foils for adequate directional stability.

One approximate method for determining the foil- borne transverse metacentric height of a surface-pierc- ing V-foil configuration as shown in Fig. 70 is given in the High Speed Code IMO (2000):

IH = angle at which aft foil is inclined to horizontal S = height of CG above water

Boats with fully submerged foils depend almost en- tirely on active controls for transverse stability. The available righting moment is usually so large at normal foilborne speeds that wind loads and off-center passen- ger loads have negligible effect.

One important source of roll disturbance is the wa- ter velocity due to the orbital motion in waves. In beam seas, the horizontal component of the orbital velocity produces sideslip and resultant side forces on the struts, while the vertical component alters the angle of attack, and hence the lift, on the foils; the phase relation of this effect, port and starboard, may induce appreciable roll- ing if not counteracted by the controls. At any foilborne speed, the righting moment obtainable from deflection of the ailerons (and the rudder, if heel-to-steer control is provided) must be larger than the heeling moment pro- duced by any possible beam-sea condition within the craft's operational envelope.

14.1.6 Air Cushion Vehicles. When an air cushion vehicle (ACV) operates in the cushionborne mode, the craft is supported aerostatically by the air pressure in the cushion system. The transverse (or longitudinal) stability of a simple craft with a single air cushion and no flexible skirts or fixed sidewalls is inherently nega- tive because the uniform air pressure acting across the undersurface provides an upward buoyancy force that always acts through the center of pressure (CP) of the upper surface of the air chamber (which is on the centerline of a symmetrical hull), while the downward weight force acting through the CG (usually above the CP) will be displaced in the direction of heel (or trim), thus producing an upsetting moment.

One method of improving stability at low speeds is to subdivide the air cushion by means of air jets or flexible skirt keels-at least one fore-and-aft and one athwart-

where:

n, = percentage of hydrofoil load borne by front foil nH = percentage of hydrofoil load borne by aft foil L, = clearance of front foil LH = clearance of aft foil a = clearance between bottom of keel and water g = height of CG above bottom of keel I, = angle at which front foil is inclined to horizontal

Fig. 70 Surface-piercing hydrofoil craft parameters ( HSC code)

Page 94: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 79

ships. In the hypothetical case of little or no air leakage, when the craft heels (or trims), the air in the heeled- down compartment would be compressed, thus increas- ing the pressure, while the air in the raised compart- ment would expand, thus decreasing the pressure. The result would be a righting moment of buoyancy which would tend to counteract the effect of the off-center weight vector.

In actual ACVs, air is supplied by a lift system of fans and leaks continuously through the open gap between the water and the craft's seal system. In early designs, the seal system consisted of peripheral jets or skirts that were flexible and designed so that heel or trim will give rise to a skirt deflection that would shift the geo- metrical center of pressure to cause a righting moment (Fig. 71). In some early designs, these air leakage forces were partially controlled with variable geometry skirts. As skirt designs have progressed, the complex periph- eral jets and controllable geometry designs have been replaced by much simpler and more effective skirts of the bag-and-finger type. If cushion compartmentation as well as flexible skirts are provided, additional stabil- ity will result. When heeled, the gap and hence air leak- age in the down side are reduced and the air pressure increases while the leakage increases on the high side, with a corresponding pressure drop. Again, a righting moment results.

Underway, particularly at higher speeds, the ACV develops aerodynamic and hydrodynamic forces and moments affecting the craft's stability, and these must be considered in design. Incidents of trim and heel in- stability have occasionally caused some craft to plow-in and/or capsize, usually during turning maneuvers under severe weather conditions. Hence, efforts must be made by experiments and full-scale trials to ensure that dy- namic effects on stability are favorable.

ACV designers can make use of a nondimensional transverse stiffness measure for evaluating stability, ~ I B , where is the transverse metacentric height on cushion and B is the overall beam. In calculating or estimating m, the principal effect to be taken into ac- count is the shift of the center of cushion pressure with

Fig. 71 ACV Stability

heel relative to the vertical through the CG. For conve- nience, approximate empirical relations have been de- vised for the design evaluation of stiffness.

Criteria for stability are found in the High Speed Craft code and further background is available in the bibliography of Lamb (2004).

1-4.1.7' Surface Effect Ships. SES utilize another method of obtaining transverse stability: immersed sidewall structures containing sufficient volume to pro- vide significant buoyancy and dynamic lift when under- way in the cushionborne mode. The static shift of center of buoyancy when the craft heels can then be calculated and stability determined in a similar manner as for a conventional hull, except that the waterline inside the hulls is lower than on the outside. Longitudinal stability is attained by the use of flexible seals at bow and stern, supplementing the buoyant moments of the sidewalls.

Transverse stability must be carefully evaluated during the design stages. This is particularly true for SES with high ratios of length to beam. The U.S. Navy's XR-5, for example, which has a length-to-beam ratio of 6.54, was subjected to a series of rudder-reversal tests, in simulated collision-avoidance avoidance maneuvers with a range of CG heights to determine acceptable ra- tios of CG height-to-beam.

Criteria for stability are found in the High Speed Craft code and further background is available in the bibliography of Lamb (2004).

1-4.1.8 Wing in Ground. A wing-in-ground (WIG) craft is defined as a vessel capable of operating com- pletely above the surface of the water on a dynamic air cushion created by aerodynamic lift due to the ground effect between the vessel and the water's surface. WIG craft are capable of operating at speeds in excess of 100 knots. Stability of these craft is governed by aero- dynamic effects during flight. Longitudinal stability is the primary concern as the center of lift changes rap- idly if the WIG transitions from ground-effect operation to fully airborne, leading to the possibility of flipping in pitch. This is controlled generally by large horizontal tail surfaces. The IMO has produced Interim Guidelines for WIG craft (IMO, 2002).

References

American Bureau of Shipping. (2008). Rules for build- i n g and classing offshore mobile drilling un i t s . Houston, TX.

Andrews, D. J., & Zhang, J-W. (1996). A novel solution to stability-the trimaran ship. International RINA S y m p o s i u m o n Watertight Integrity and Sh ip Sur- vivabili ty. London, England.

Amy, J. R., Johnson, R. E., and Miller, E. R. (1976). De- velopment of intact stability criteria for towing and fishing vessels. Trans. SNAME, 8-4, 75-114.

Belenky V. L., de Kat, J. O., & Umeda, N. (2008). Toward performance-based criteria for intact stability. Ma- r i n e Technology, -45, 101-120.

mgray
71
Page 95: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 79

ships. In the hypothetical case of little or no air leakage, when the craft heels (or trims), the air in the heeled- down compartment would be compressed, thus increas- ing the pressure, while the air in the raised compart- ment would expand, thus decreasing the pressure. The result would be a righting moment of buoyancy which would tend to counteract the effect of the off-center weight vector.

In actual ACVs, air is supplied by a lift system of fans and leaks continuously through the open gap between the water and the craft's seal system. In early designs, the seal system consisted of peripheral jets or skirts that were flexible and designed so that heel or trim will give rise to a skirt deflection that would shift the geo- metrical center of pressure to cause a righting moment (Fig. 71). In some early designs, these air leakage forces were partially controlled with variable geometry skirts. As skirt designs have progressed, the complex periph- eral jets and controllable geometry designs have been replaced by much simpler and more effective skirts of the bag-and-finger type. If cushion compartmentation as well as flexible skirts are provided, additional stabil- ity will result. When heeled, the gap and hence air leak- age in the down side are reduced and the air pressure increases while the leakage increases on the high side, with a corresponding pressure drop. Again, a righting moment results.

Underway, particularly at higher speeds, the ACV develops aerodynamic and hydrodynamic forces and moments affecting the craft's stability, and these must be considered in design. Incidents of trim and heel in- stability have occasionally caused some craft to plow-in and/or capsize, usually during turning maneuvers under severe weather conditions. Hence, efforts must be made by experiments and full-scale trials to ensure that dy- namic effects on stability are favorable.

ACV designers can make use of a nondimensional transverse stiffness measure for evaluating stability, ~ I B , where is the transverse metacentric height on cushion and B is the overall beam. In calculating or estimating m, the principal effect to be taken into ac- count is the shift of the center of cushion pressure with

heel relative to the vertical through the CG. For conve- nience, approximate empirical relations have been de- vised for the design evaluation of stiffness.

Criteria for stability are found in the High Speed Craft code and further background is available in the bibliography of Lamb (2004).

1-4.1.7' Surface Effect Ships. SES utilize another method of obtaining transverse stability: immersed sidewall structures containing sufficient volume to pro- vide significant buoyancy and dynamic lift when under- way in the cushionborne mode. The static shift of center of buoyancy when the craft heels can then be calculated and stability determined in a similar manner as for a conventional hull, except that the waterline inside the hulls is lower than on the outside. Longitudinal stability is attained by the use of flexible seals at bow and stern, supplementing the buoyant moments of the sidewalls.

Transverse stability must be carefully evaluated during the design stages. This is particularly true for SES with high ratios of length to beam. The U.S. Navy's XR-5, for example, which has a length-to-beam ratio of 6.54, was subjected to a series of rudder-reversal tests, in simulated collision-avoidance avoidance maneuvers with a range of CG heights to determine acceptable ra- tios of CG height-to-beam.

Criteria for stability are found in the High Speed Craft code and further background is available in the bibliography of Lamb (2004).

1-4.1.8 Wing in Ground. A wing-in-ground (WIG) craft is defined as a vessel capable of operating com- pletely above the surface of the water on a dynamic air cushion created by aerodynamic lift due to the ground effect between the vessel and the water's surface. WIG craft are capable of operating at speeds in excess of 100 knots. Stability of these craft is governed by aero- dynamic effects during flight. Longitudinal stability is the primary concern as the center of lift changes rap- idly if the WIG transitions from ground-effect operation to fully airborne, leading to the possibility of flipping in pitch. This is controlled generally by large horizontal tail surfaces. The IMO has produced Interim Guidelines for WIG craft (IMO, 2002).

References

American Bureau of Shipping. (2008). Rules for build- i n g and classing offshore mobile drilling un i t s . Houston, TX.

Andrews, D. J., & Zhang, J-W. (1996). A novel solution to stability-the trimaran ship. International RINA S y m p o s i u m o n Watertight Integrity and Sh ip Sur- vivabili ty. London, England.

Amy, J. R., Johnson, R. E., and Miller, E. R. (1976). De- velopment of intact stability criteria for towing and fishing vessels. Trans. SNAME, 8-4, 75-114.

Belenky V. L., de Kat, J. O., & Umeda, N. (2008). Toward performance-based criteria for intact stability. Ma- r i n e Technology, -45, 101-120.

mgray
71
Page 96: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

80 INTACT STAB1 l lTY

Belenky, V. L., & Sevastianov, N. B. (2007). Stability and safety of ships, r isk of capsizing. (2nd Ed.). Jersey City, NJ: SNAME.

Bertaglia, G., Scarpa, G., Serra, A., Francescutto, A., & Bulian, G. (2004). Systematic experimental tests for the IMO weather criterion requirements and further development towards a probabilistic intact stability approach. Proceedings of the 7th International Ship Stability Workshop. Shanghai, China.

Bertaglia, G., Serra, A., & Francescutto, A. (2003). The intact stability rules are changing: impact on the de- sign of large cruise ships. International Conference o n Passenger Sh ip Safety, 45-54. London, England.

Boccadamo, G., Cassella, P., Russo Krauss, G., & Sea- mardella, A. (1994). Analysis of IMO stability crite- ria by systematic hull series and by ship disasters. Proceedings of the 5th International Conference o n Stability of Ships and Ocean Vehicles. Melbourne, Florida.

Boze, W. (2003). Mass properties. In T.Lamb (Ed.), Ship design and construction. Jersey City, NJ: SNAME.

Bouger, P. (1746). Trait6 d u Navire, et de ses mouve- mens. Paris: Jombert.

Breuer, J. A., & Sjolund, K-G. (2006). Orthogonal tip- ping in conventional offshore stability evaluations. Proceedings of the 9 th International Conference o n Stability of Ships and Ocean Vehicles. Rio De Janei- ro, Brazil.

Bulian, G., Francescutto, A., Serra, A., & Umeda, N. (2004). The development of a standardized experi- mental approach to assessment of ship stability in the frame of weather criterion. Proceedings of the 7 th International Sh ip Stability Workshop. Shang- hai, China.

Clauss, G. F., Hennig, J., Brink, K.-E., & Cramer, H. (2004). A new technique for the experimental investigation of intact stability and the validation of numerical simu- lations. Proceedings of the 7th International Ship Stability Workshop. Shanghai, China.

Courser, P. (2003). A software developer's perspective of stability criteria. Proceedings of the. 8 th Inter- national Conference o n the Stabili ty of Sh ip and Ocean Vehicles (STAB 2003). Madrid, Spain.

Cramer , H., Kruger, S., & Mains, C. (2004). Assessment of intact stability-revision and development of sta- bility standards, criteria and approaches. Proceed- ings of the 7th International Sh ip Stability Work- shop. Shanghai, China.

De Kat, J. O., Brouwer, R., McTaggart, K. A., & Thomas, W. L. (1994). Intact survivability in extreme waves: new criteria from a research and Navy perspective. Fifth International Conference o n Stability of Ships and Ocean Vehicles, STAB '94. Melbourne, Florida.

De Kat, J. O., Van Walree, F., & Ratcliffe, A.T. (2006). Fo- rensic research into the loss of ships by means of a time domain simulation tool. Proceedings of the 6th International Conference o n Stability of Ships and Ocean Vehicles. Rio de Janeiro, Brazil.

Dudziak, J., & Buczkowski, A. (1978). Probability of ship capsizing under the action of the beam wind and sea as a background of stability criteria. Proceedings of the I IMAEM Congress, Vol. I, 678-701. Istanbul, Tur- key.

Faltinsen, 0 . (2006). Hydrodynamics of high-speed mar ine vehicles. Cambridge, MA: Cambridge Press.

France, W., Levadou, M., Treakle, T. W., Paulling, J. R., Michel, R. K., & Moore, C. (2003). An investigation of head sea parametric rolling and its influence on con- tainer lashing systems. Marine Technology, 40, 1-19.

Francescutto, A. (1992). Towards a reliability based ap- proach to the hydrodynamic aspects of seagoing ves- sels safety. Proceedings of the 11th Int. Conference of OMAE'92,2, 169-173. Calgary, Canada.

Francescutto, A. (1993). Is it really impossible to design safe ships? Trans. RINA, 135, 163-173.

Francescutto, A., & Nabergoj, R. (1990). Parametric analysis of the stability of a family of fishing vessels. Bulletin de 1lAssociation Technique Maritime et Akronautique, 90,63-81.

Francescutto, A., & Serra, A. (2001). Weather criterion for intact stability of large passenger vessels. Pro- ceedings of the 5th International Workshop o n Ship Stability and Operational Safety, 4.6.1-4.6.4. Tri- este, Italy.

Francescutto, A., Serra, A., & Scarpa, S. (2001). A criti- cal analysis of weather criterion for intact stability of large passenger vessels. Proceedings of the 20 th International Conference OMAE'OI. Rio de Janeiro, Brazil.

Francescutto, A. (2002). Intact stability-the way ahead. Proceedings of the 6th International Sh ip Stabili ty Workshop. Glenn Cove, New York.

Goldberg, L. L., & Tucker, R. G. (1975). Current status of stability and buoyancy criteria used by the U.S. Navy for advanced marine vehicles. Naval Engineers Journal, 8 7 33-46.

Green, P. V. (1980, January). The hazard of flow in bulk mineral cargoes. (Warren Springs Lab., U.K.) Safety at Sea.

Hayes, P. (2002). The Royal Australian Navy stability standard. Proceedings of the 7th International Con- ference o n Stabili ty of Ships and Ocean Vehicles, Launceston, Tasmania.

Hua, J. (2004). Model test for quantitative analysis of ship dynamics in waves. Proceedings of the 7th Interna- tional Ship Stability Workshop, Shanghai, China.

IMO. (1974). International convention on load lines and international convention for the safety of life at sea (SOLAS). London, England.

IMO. (1977). International convention for safety of fishing vessels (Res. A.168). London, England.

IMO. (1991). International code for the safe carriage of grain in bulk (IMO-240E). London, England.

IMO. (1995). Guidance to the master for avoiding dan- gerous situations in following and quartering seas (MSCICirc.707). London, England.

Page 97: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

INTACT STAB1 llTY 8 1

IMO. (1995a). 1993 Torremolinos protocol and Torre- molinos international convention for the safety of fishing vessels. London, England.

IMO. (1999) Amendments to the code on intact stabili- ty for all types of ships covered by IMO instruments (IMO Res. MSC.75 (69)). London, England.

IMO. (2000). International code of safety for high-speed craft. London, England.

IMO. (2002). Interim guidelines for wing-in-ground (WIG) craft (MSCICirc. 1054). London, England.

IMO. (2003). Development of a generalized S factor (IMO Document SLF 45lINF.6). London, England.

IMO. (2004). International convention for the safety of life at sea, 1975, and its protocol of 1988: articles, annexes and certificates. London, England.

IMO. (2004a). Review of the intact stability code: to- wards the development of dynamic stability crite- r i a (IMO Document SLF 471416). London, England.

IMO. (2005). Code of safe practice for solid bulk car- goes. London, England.

IMO. (2005a). Voluntary guidelines for the design, con- struction and equipment of small fishing vessels. London, England.

IMO. (2006) International convention on load lines and international convention for the safety of life at sea (SOLAS). London, England.

IMO. (2006a). Interim guidelines for the alterna- tive assessment of the weather criterion (MSC.11 Circ.1200). London, England.

IMO. (2007). Explanatory notes to the interim guide- lines for alternative assessment of the weather cri- terion (MSC.llCirc.1227). London, England.

IMO. (2007a). Revised guidance to the masterfor avoid- ing dangerous situations i n adverse weather and sea conditions (MSC.llCirc.1228). London, England.

IMO. (2008). Code on intact stability (Approved at IMO SLF50, May 2007). London, England.

IMO. (2008a). Explanatory notes to the international code on intact stability (MSC.llCirc.1281). London, England.

ISAWE. (1997, May 21). Weight control technical re- quirements for surface ships, Recommended Prac- tice No. 12, Rev. B. Los Angeles, California: Interna- tional Society of Allied Weight Engineers.

Iskandar, B. H., Umeda, N., &Hamarnoto, M. (2001). Cap- sizing probability of an Indonesian RoRo passenger ship in irregular beam seas. Journal of the Society of Naval Architects, 189,31-37.

Kobylinski, L. (1993). Discussion in Fransecutto (1993), Trans. RINA. 135, 163-173.

Kruger, S., Hinrichs, R., & Cramer, H. (2004). Perfor- mance based approaches for the evaluation of intact stability problems. 9th Symposium on Practical De- sign of Ships and Other Floating Structures. Lue- beck-Travemuende, Germany.

Kuo, C., & Welaya, Y. (1981). A review of intact ship sta- bility research and criteria. Ocean Engineering, 8, 65-84.

Lamb. T. (2004). Ship design and construction. Jersey City, NJ: SNAME.

Letcher, J. (2009). The geometry of ships. In J. R. Paulling (Ed.), Principles of naval architecture: the series. Jersey City, NJ: SNAME.

MARAD. (1995). Maritime administration classifica- tion of weights of merchant ships. Washington, DC: U.S. Maritime Administration.

MOD. (1999). Stability of surface ships, part 1, conven- tional ships (Report SSP 24). London, England: U.K. Ministry of Defence, Sea Technology Group.

Moore, C., Neuman, J., & Pippenger, D. (1996). Intact sta- bility of double hull tankers. Marine Technology, 33, 435-450.

National Cargo Bureau. (1994, Rev. 2002). General in- formation for grain loading. Baltimore, MD: Nation- al Cargo Bureau, Inc.

NAVSEA. (1975). Stability and buoyancy of U.S. Naval surface ships (Design Data Sheet DDS 079-1). Wash- ington, DC: U.S. Department of the Navy.

NAVSEA. (1985). Expanded ship work breaksown struc- ture for all ships and ship/combat systems (29040- AA-1DX-0101SWBS 5d and S9040-AA-1DX-020lSWBS 5d (I and 11)). Washington, DC: U.S. Department of the Navy.

NAVSEA. (2001). Policy for weight and vertical center of gravity above botton of keel (KG) margins for surface ships (NAVSEAINST 9096.6B). Washington, DC: U.S. Department of the Navy.

Niedermair, J. C. (1932). Stability of ships after damage. Trans. SNAME, 40.

NOAA. (2005). 2005 Atlantic Ocean tropical cyclones. Retrieved July 7, 2009, from http:llwww.ncdc.noaa. gov/oa/c1imate/research/200512005-at1antic-trop- cyclones. html

NSWCCD. (2008). Stability and buoyancy of US. Naval surface ships (Design Data Sheet DDS 079-1, Version 2.01). Bethesda, MD: Naval Surface Warfare Center, Carderock Division, NSWCCD-20-TR-2007105.

Numata, E., Michel, W. H., & McClure, A. C. (1976). As- sessment of stability requirements for semisubmers- ible drilling units. Trans. SNAME, 84, 56-74.

Oakley, 0 . H., Paulling, J. R., & Wood, P. D. (1974). Ship motions and capsizing in astern seas. Proceedings of the Tenth Symposium on Naval Hydrodynamics. Cambridge, MA: MIT.

Paulling, J. R. (1959). Transverse stability of tuna clip- pers. Proceedings of the Second International Fish- ing Boat Congress, Rome, Italy. In J.O. Traung (Ed.), Fishing boats of the world II. London, England: Fish- ing News (Books) Ltd.

Paulling, J.R. (2007). On parametric rolling of ships. Proceedings of the International Conerence. On Practical Design of Ships and Other Floating Struc- tures, 2, 707-715. Houston, TX: ABS.

Rahola, J. (1939). Thejudging of the stability of ships and the determination of the minimum amount of stabil- ity. Unpublished doctoral dissertation, Helsinki.

Page 98: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

82 INTACT STAB1 l lTY

Rawson, K. J., & Tupper, E. C. (1965). Basic ship theory, Vol. I. Oxford, England: Butterworth-Heinemann.

Sarchin, T. H., & Goldberg, L. L. (1962). Stability and buoyancy criteria for U.S. Naval surface ships. Trans. SNA ME, 70.

Tagg, R. D. (2010). Subdivision and damage stability. In J. R. Paulling (Ed.), Principles of Naval architecture: the series. Jersey City, NJ: SNAME.

Tanaka, M. (1990). Development of new criteria against shifting of bulk cargoes (IMO, BC311312). Tamaki Ura, Japan.

Umeda, N., & Ikeda, Y. (1994). Rational examination of stability criteria in the light of capsizing probability. Proceedings of the 5th International Conference on Stability of Ships and Ocean Vehicles. Melbourne, Florida.

Umeda, N., Ikeda, Y., & Suzuki, S. (1992). Risk analysis applied to the capsizing of high-speed craft in beam seas. In J. B. Caldwell, & G. Ward (Eds.), Proceed- ings of the International Conference on Practical Design of Ships and Mobile Units-PRADS'92, Vol. 2. Amsterdam: Elsevier, 2.1131-2.1145.

USCG Code of Federal Regulations. (2009). Title 46- Shipping, chapters I and 11, containing U S . Coast Guard stability regulations applicable to commer- cial ships. Washington, DC.

Van Daalen, E. F. G., Boonstra, H., & Blok, J. J. (2005). Capsize probability analysis for a small container vessel. Stability Workshop, Istanbul, Turkey.

Van Santen, J. (1986). Stability calculations for jack-ups and semi-submersibles. Proceedings of the Confer- ence on Computer Aided Design, Manufacture and Operation in the Marine and Offshore Industries. Washington, DC.

Vassalos, D. (1986). A critical look into the development of ship stability criteria based on worklenergy bal- ance. Trans. RINA, 128,217-234.

Watanabe, Y. (1938). Some contributions to the theory of rolling. I.N.A. Transactions, 408-432.

Wendel, K. (1960). Die Wahrscheinlichkeit des Ueberste- hens von Verletzungen. Schiffstechnik, z47-61.

Wise, J. L., & Comiskey, A. L. (1980). Superstructure ic- ing in Alaskan waters (NOAA Special Report). Boul- der, CO: NOAA.

Womack, J. (2002). Small commercial fishing vessel sta- bility analysis. Where are we now? Where are we go- ing? Proceedings of the 6th International Ship Sta- bility Workshop. Webb Institute, Glenn Cove, New York.

Yamagata, M. (1959) Standards of stability adopted in Japan. Transactions of Institution of Naval Archi- tects, 101, 417-443.

Page 99: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

This page has been reformatted by Knovel to provide easier navigation.

INDEX

Index Terms Links

A

Adequate stability, U.S. Navy criteria 47 47

Advanced marine vehicles 76

ACVs 78 79

catamarans 77

hydrofoil craft 77 78

planning hull 77

SES 79

small waterplane area hull 77

stability criteria, hazards 76

trimarans, pentamarans 77

WIG crafts 79

Air cushion vehicle (ACV) 78 79

Antiroll tanks, free-surface effect, free liquids 37 37

B

Beam effect 20 20

Beam winds 6 6 46

Bilge fining 21 21

Bulk carriers carrying grain 50

Bulk dry cargo, free-surface effect, free liquids 37

C

Cargo effect on stability, changes in weight 40

Catamarans 77

Center of buoyancy 2

Center of floatation 13

Center of gravity (CG)

changes in weight effect 38

height, stability 4 5

weight and 1

see also Displacement, CG location determinations

Page 100: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Changes in weight

cargo effect on stability 40

CG impact 38

displacement impact 38

initial heel compensation 39

large trim changes 39

liquid and stores consumption 40

stability impact 38

submarines 39

Critical roll axis 27

Cross curves of stability, statical stability curves 19 19

Crowding of personnel, U.S. Navy stability criteria 48 48

D

Dangerous phenomena combinations 45

Depth effect, statical stability curves 20 20

Displacement, buoyancy interaction

floating body equilibrium 2 3

longitudinal equilibrium 5 5

overturning moments 3 5

righting moments 3 5

stable equilibrium, floating body 3 4

submerged floating body 3 4 6

watertight rectangular body 3

see also Draft, trim, heel, and displacement calculations

Displacement, CG location

CG margins 10 10

changes in weight effect 38

classification systems 9

detailed estimates 9

sample loading computer display 11

sample summaries 9

variation with ship loading 10 11

weight margins 10 10

see also Draft, trim, heel, and displacement calculations

Diving ballast 67

Diving trim 68

Double hulls, free-surface effect, free liquids 35 36

Page 101: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Draft readings, inclining experiment 61

Draft, trim, heel, and displacement calculations

center of flotation 53

displacement and CG from drafts 55 55

draft after change in loading 56

drag 58

heel computation 59

hog and sag 58

moment to trim 1 cm 54

navigational drafts 57 57

reference planes 59

tons per centimeter immersion 54

trim 53

weight, CG and 54

Drafts, trim, inclining experiment 60

Drag 58

Dynamic stability assessments 53

Dynamic stability, rolling effect 24 25

E

Equilibrium concepts 1

Equilibrium polygon, submerged equilibrium 69 70 71

F

Fishing vessels, stability criteria 49

Fluid shift for wall-sided tank, free-surface effect, free liquids 30 30

Form changes, statical stability curves 21 21

Free surface in tanks, inclining experiment 60

Free-surface effect, free liquids

antiroll tanks 37 37

approximate vs. exact calculations 34 34 35

bulk dry cargo 37

double hulls 35 36

fluid shift for wall-sided tank 30 30

large angles 31 32

longitudinal subdivision 35 35

metacentric height 30

moment of inertia 33 33

Page 102: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Free-surface effect, free liquids (Cont.)

moment of transference 31 32

numerical example 34 34

relative filling level 31 32

relative free-surface effect 31 32

righting arm effect 31 32

tank fill 36 36

top and bottom effects 32 32

trim 34

two liquids 36

wing ballast tank 33 33

Fuel ballast tanks, submerged equilibrium 66

Full load departure condition, metacentric height 15 16

G

GM, GZ curves, stability criteria 41 43 44

Gravitational stability 1

Grounded ships

dry dock stability 74 74

stranded stability 75

Grounding effect 7 7

H

Heel

computation 59

forces, inclining experiment 61

heeling moment, statical stability curves 25 26

initial heel compensation, changes in weight effect 39

see also Draft, trim, heel, and displacement calculations;

Upsetting forces, heeling moments

High-speed turning, U.S. Navy stability criteria 48

Hog and sag 58

Hydrofoil craft 77 78

I

IMO Resolution A.167 43

IMO Resolution A.749 43

Page 103: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Impulsive moment response 25 25

Inclining experiment

accuracy 64

basic principles 59

draft readings 61

drafts, trim 60

forces affecting heel 61

free surface in tanks 60

inclination measurement 61 62

inclining in air 63 65

inclining weights selection 61

induced rolling, sallying 65

inventory 62

list 60

metacentric height 60

personnel aboard 61

personnel movement 62

plot of tangents 62

preparation for inclining 60

report 62 63

schedule 60

swinging weights 61

transfer of liquids 61

water density 61

weight movements 62

Inclining in air 63 65

Inclining weights selection 61

Induced rolling, sallying 65

Initial heel compensation, changes in weight effect 39

International Convention for the Safety of Life at Sea (SOLAS) 15

International Convention on Load Lines (ICLL) 15

International Marine Organization (IMO)

MSC Circular 1228 44

Resolution A.167 43

Resolution A.749 43

stability criteria 42 51

Page 104: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

J

Jack-up platform, statical stability curves 28 28 29

L

Lead, solid ballast, submerged equilibrium 66

Lead, variable tankage adjustment, submerged equilibrium 70 71

Lifting weights effect, U.S. Navy stability criteria 47

Lightship, submerged equilibrium 66

Liquid and stores consumption, changes in weight effect 40

List, inclining experiment 60

Load to submerge 66

Loading conditions, metacentric height 15 16

Longitudinal equilibrium 5 5

Longitudinal metacentric height 12 13

Longitudinal stability, upsetting forces, heeling moments 8 8

Longitudinal subdivision, free-surface effect, free liquids 35 35

M

Main ballast tanks, submerged equilibrium 66

Maximum righting moment, statical stability curves 22 23

Merchant ship stability criteria 42

Metacentric height (GM)

applications 14

arrival conditions 15

center of floatation 13

free-surface effect, free liquids 30

full load departure condition 15 16

inclining experiment 60

loading computer software 17

loading conditions 15 16

longitudinal metacenter 12 13

metacenter, submerged submarines 14

minimum operating conditions 15

moment to heel 1 degree 14

moment to trim 1 degree 14

Navy ships 16

partial load departure conditions 15

Page 105: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Metacentric height (GM) (Cont.)

passenger ships 16

period of roll 14

Ship Status for Proposed Weight Changes 17 17

stability curves 16

statical stability curves 16

submerged submarines 14

suitable conditions 16

transverse metacenter 11 11 12

trim effects 14

Midcolumn draft, statical stability curves 27 27 28

Minimum operating conditions, metacentric height 15

Mobile offshore drilling units (MODUs) 50

Moment diagram, submerged equilibrium 65

Moment of inertia, free-surface effect, free liquids 33 33

Moment of transference, free-surface effect, free liquids 31 32

Moment to heel 1 degree, transverse metacentric height 14

Moment to trim 1 degree 1cm, longitudinal metacentric height 14

MSC Guidance to Masters, Circular 1228 44

N

Neutral equilibrium 1 3

Normal fuel-oil tanks, submerged equilibrium 66

O

Offshore structures, non-ship-shape vessels

statical stability curves 27 27 28 29

Offside weight, upsetting forces, heeling moments 6 6

Overturning moments, weight and buoyancy interaction 3 5

P

Parametric rolling motion 45

Passenger ships, metacentric height 16

Period of roll, metacentric height 14

Personnel aboard, inclining experiment 61 62

Planning hull 77

Pontoon-based offshore structure, statical stability curves 27 27

Page 106: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Potential energy surface, jack-up platform, statical stability

curves 28 29

Preparation for inclining, inclining experiment 60

R

Reference planes, draft, trim, heel, and displacement

calculations 59

Relative filling level, free-surface effect, free liquids 31 32

Relative free-surface effect, free-surface effect, free liquids 31 32

Reserve buoyancy, submerged equilibrium 67

Residual water, submerged equilibrium 67

Righting arm (GZ), statical stability curves 17 18

Righting arm curve computation, statical stability curves 18

Righting arm effect, free-surface effect, free liquids 31 32

Righting moments

statical stability curves 26 26

weight and buoyancy interaction 3 5

Roll axis, critical, statical stability curves 27

Rolling effect, dynamic stability, statical stability curves 24 25

S

Sallying, see Induced rolling, sallying

Ship main body appendages, buoyancy contribution 19

Ship Status for Proposed Weight Changes 17 17

Slope of GZ curve at origin, statical stability curves 22 22 23

Small waterplane area hull 77

Stability criteria

bulk carriers carrying grain 50

dangerous phenomena combinations 45

dynamic stability assessments 53

fishing vessels 49

GM, GZ curves 41 43 44

hazards 76

IMO 42 51

merchant ships 42

MODUs 50

parametric rolling motion 45

surf-riding, broaching-to 45

Page 107: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Stability criteria (Cont.)

synchronous rolling motion 45

topside icing 48

towboats 50

wave crest at midship 45

Stability criteria, U.S. Navy

adequate stability 47 47

beam winds, rolling 46

crowding of personnel 48 48

general 45

heeling arms 47

high-speed turning 48

lifting weights effect 47

wind heeling moment 46 46

wind pressure vs. height 47

wind velocities 46 46

Stability evaluation standards 41

Stability in depth, submerged equilibrium 71

Stability, drafts, and list improvement

changes in form 73

load adjustment 73

loading instructions 73

permanent ballast 73

weight removal 73

Stable equilibrium 1

floating body 2 4

submerged floating body 3 4 6

Statical stability curves

beam effect 20 20

bilge fining effect 21 21

cross curves of stability 19 19

depth effect 20 20

dynamic stability, rolling effect 24 25

form changes 21 21

heeling moment 25 26

impulsive moment response 25 25

jack-up platform 28 28 29

maximum righting moment 22 23

Page 108: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Statical stability curves (Cont.)

metacentric height 16

midcolumn draft 27 27 28

offshore structures, non-ship-shape vessels 27 27 28 29

pontoon-based offshore structure 27 27

potential energy surface, jack-up platform 28 29

righting arm curve 17 18

righting arm curve computation 18

righting moment 26 26

roll access, critical 27

rolling effect, dynamic stability 24 25

ship main body appendages, buoyancy contribution 19

significance 22 22 23 24

slope of curve at origin 22 22 23

static stability curves 18 18

transverse righting arms 17 17 18

tumble-home flare effect 21 21

typical stability curves, different ships 24

waves effect 21 22

work and energy determination 24 25

Submarines

changes in weight effect 39

stability criteria 50 51

submerged, metacentric height 14

Submerged equilibrium 8

definition 66

diving ballast 67

diving trim 68

equilibrium conditions 69

equilibrium polygon 69 70 71

fuel ballast tanks 66

lead, solid ballast 66

lead, variable tankage adjustment 70 71

lightship 66

load to submerge 66

main ballast tanks 66

maximum condition, surfaced 67

moment diagram 65

Page 109: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Submerged equilibrium (Cont.)

normal condition, surfaced 67

normal fuel-oil tanks 66

reserve buoyancy 67

residual water 67

stability in depth 71

submerged displacement 66

variable ballast 67

variable load 67

water seal, fuel ballast tanks 67

weight items 66 66 67

weight items relationship 66 67 67

Submerged floating body, weight and buoyancy interaction 3 4 6

Surf-riding, broaching-to 45

Surface effect ships (SES) 79

Suspended cargo or weight, effect on stability 38

Swinging weights, inclining experiment 61

Synchronous rolling motion 45

T

Tank fill level, free-surface effect, free liquids 36 36

Top and bottom effects, free-surface effect, free liquids 32 32

Topside icing, stability criteria 48

Towboats, stability criteria 50

Transfer of liquids, inclining experiment 61

Transverse metacenter, metacentric height 11 11 12

Transverse righting arms, statical stability curves 17 17 18

Transverse stability, upsetting forces, heeling moments 8 8

Trim

changes in weight effect 39

free-surface effect, free liquids 34

metacentric height 14

see also Draft, trim, heel, and displacement calculations

Trim dive 67

basic principles 72

calculations, report 72

conducting 72

Trimarans, pentamarans 77

Page 110: The Principles of - eclass.snd.edu.gr€¦ · The Principles of Naval Architecture Series Intact Stability Colin S. Moore J. Randolph Paulling, Editor Published by The Society of

Index Terms Links

This page has been reformatted by Knovel to provide easier navigation.

Tumble-home and flare effects, statical stability curves 21 21

Turn effect, upsetting forces, heeling moments 7 7

Two liquids, free-surface effect, free liquids 36

U

Unstable equilibrium 1

Upsetting forces, heeling moments

beam wind 6 6

grounding effect 7 7

longitudinal stability 8 8

offside weight 6 6

transverse stability 8 8

turn effect 7 7

weight lifting over the side 6 6

V

Variable ballast, submerged equilibrium 67

Variable load, submerged equilibrium 67

W

Water density, inclining experiment 61

Water seal, fuel ballast tanks, submerged equilibrium 67

Watertight rectangular body, stability 3

Waves effect, statical stability curves 21 22

Waves effects, dynamic 45

Weight estimate 9

Weight items, submerged equilibrium 66 67 67

Weight lifting over the side, upsetting forces, heeling moments 6 6

Weight movements, inclining experiment 62

Wind heeling moment, U.S. Navy stability criteria 46 46

Wind pressure vs. height, U.S. Navy stability criteria 47

Wind velocities, U.S. Navy stability criteria 46 46

Wing ballast tank, free-surface effect, free liquids 33 33

Wing-in-ground (WIG) crafts 79

Work and energy determination, statical stability curves 24 25