Floating and Submerged Body Hydrodynamics-Book Chapter

Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

CHAPTER 6 FLOATING AND SUBMERGED BODY

HYDRODYNAMICS

6.1 TERMINOLOGY

The analysis and design of ocean structures includes ocean systems that are fixed, floating, or
submerged. Ship structure systems require a means of propelling them through the ocean waters, and for
decades naval architects have sought to incorporate better propulsion methods into the design and analysis
of ships. The terminology associated with ship design is quite different from that encountered in traditional
designs of land structures, and some of this terminology is also used in the field of ocean engineering. Some
of the common nautical terminology used aboard naval vessels and merchant ships is tabulated in Table 6-1,
and that related to naval architecture, and marine and ocean engineering for the design and analysis of float-
ing bodies is defined in Table 6-2.

6.1.1 Ship Geometry

The geometry of a ship is illustrated in what is known as lines drawings, an example of which is
shown in Fig. 6-1. A set of lines drawings includes the sheer, half-breadth, and body plans. The sheer plan
consists of vertical planes with the centerplane through the ship centerline and buttock planes at specified
distances from the centerline. The half-breadth (or waterlines) plan shows planes that are parallel to the base
plane and intersect the hull; all of these planes are called waterplanes. The body plan illustrates planes that
are perpendicular to the buttock and waterline planes, and intersect the hull.
A vertical line is drawn in the sheer plan at the intersection of the designed load waterline (DWL);
this is called the forward perpendicular (FP). A similar vertical line is drawn at the stern and is known as
the after perpendicular (AP). The distance between these two perpendiculars is the length of the ship, and
is often referred to as the length between perpendiculars (LBP). This length is used in the American Bureau
of Ships (ABS) Rules for Building and Classing Steel Vessels.
The LBP is divided into 10–40 intervals by body plan planes; these intervals are called body plan
stations, or simply stations. These stations are usually numbered from the bow, with station “0” being the
location of the FP. The molded baseline is a straight horizontal line that is used as a reference datum for
design and construction purposes. It is also the bottom of the vessel’s molded surface and coincides with the
top surface of the keel.
The draft of a vessel is the vertical distance between the waterline at which the vessel is floating
and the bottom of the vessel. The molded draft is measured vertically between the waterline and the molded
baseline, and the keel draft is measured between the waterline and the bottom of the keel. The average of
the forward and aft drafts is the mean draft. Most ships have draft marks located amidships and at both
ends as close to the forward and aft perpendiculars as possible. The difference between the forward and
aft draft readings is known as the trim. Vessel offsets are measured from the vessel centerline to the water-
line at each station, and a complete set of offsets for the various waterlines is known as a table of offsets.
Computer programs such as HULDEF, SHIPHUL, General HydroStatics (http://www.ghsport.com), and
FORAN (http://www.foransystem.com) have been developed to generate ship lines and offsets. For offshore
floating oil and gas facilities, software such as Engineering Dynamics’ StruCad and StabCad are used for
structural and stability analyses of semisubmersibles, spars, and floating production storage and offloading
(FPSO) facilities. Det Norske Veritas (DNV) produces software (SESAM) for stability, hydrodynamics,
and mooring analyses. The ABS also provides software (Safehull) for the analysis and design of cargo
ships. Mooring analysis software is available from Global Maritime (GMoor), DNV (Mimosa), and Orcina
(Orcaflex), to name a few.

173
174 Elements of Ocean Engineering

Table 6-1. Nautical Terminology


Term Definition
Aft Rear of vessel, toward the stern of vessel
Athwartship Across ship from side to side, right angle to the vessel centerline
Adrift Vessel is loose and not attached to mooring line or towline
Aground Vessel is touching sea bottom or fast to the bottom
Ahead Moving in forward direction
Alee Opposite the direction of wind, opposite of windward
Beam Extreme width of ship at widest point
Below deck Below the main deck
Bow Front of vessel
Bulkhead Wall
Capstan Electric device for winding mooring lines
Deck Floor
Draft Depth of water the vessel draws
Fathom Six feet of water depth
Focsle, Forecastle Forward part of ship above main deck
Forward Front
Freeboard Minimum vertical distance from the water surface to main deck
Galley Vessel kitchen or cooking area
Hatch Opening in vessel with a watertight cover
Head Bathroom, restroom, lavatory, marine toilet
Hull Vessel main body or structure
Keel Bottom of centerline portion of vessel
Knot Nautical mile per hour
Lee Side of ship sheltered from wind
Leeward Being in or facing the direction toward which the wind is blowing
Midship Location equally distant between bow and stern
Mooring Arrangement to secure the vessel to the seafloor, pier, or buoy
Passageway Hallway
Port Left side of the vessel when facing the bow
Rudder Vertical structure for steering the vessel
Screw Vessel propeller
Sea cock Hull valve to isolate piping between the vessel’s interior and the sea
Starboard Right side of the vessel when facing the bow
Stern Rear of the vessel
Tide Periodic rise and fall of sea level in the ocean
Topside Above the main deck
Trim Fore and aft balance of the vessel
Windlass Electric device used to raise anchor
Windward Being in or facing the direction from which the wind is blowing

6.2 HYDROSTATICS AND STABILITY

6.2.1 Displacement

The static behavior of floating and submerged bodies is governed by Archimedes’ principle,
which states that a body immersed in a fluid is buoyed up by a force equal to the weight of the fluid
it displaces. This means that the weight of the vessel is a downward force that is proportional to the
mass of the body, and the equal upward buoyant force is proportional to the mass of the displaced fluid.
Chapter 6 Floating and Submerged Body Hydrodynamics 175

Table 6-2. Symbols and Abbreviations Used in Naval Architecture and Marine and Ocean Engineering
Symbol/Abbreviation Definition
∇ Volume displacement
∆ Mass displacement
AP After perpendicular
AW Waterplane area
B Maximum molded breadth
B, CB Center of buoyancy
BM Distance between center of buoyancy and metacenter
CG, G Center of gravity
CB Block coefficient, ∇/LBT
CWP Coefficient of waterplane area, AW/LB
DWL Designed load water line
DWT Deadweight
FP Forward perpendicular
GM Distance between CG and metacenter
GZ Righting arm
I Moment of inertia
K Radius of gyration
K Keel
KB Distance between keel and center of buoyancy
KG Distance between keel and center of gravity
KM Distance between keel and metacenter
L Length of ship
LOA Ship length overall
LBP Length between perpendiculars
LCB, LCF Longitudinal center of buoyancy, flotation
LCG Longitudinal center of gravity
M Transverse metacenter
SM Simpson multiplier, section modulus
T Draft
T Period
TCG Transverse center of gravity
V Volume of displacement
VCG, VCB Vertical center of gravity, buoyancy
W Weight of ship, displacement, g∇
Waterplane Horizontal planes parallel to designed load waterplane
WL Intersection of waterplane with vessel’s form
WL Waterline

Assume that a floating body is partially immersed in a fluid (e.g., water) at rest, as shown in Fig. 6-2. At
any point or depth (D), the mass of fluid above the depth is Az, where  is the fluid density, A is the
cross-sectional area parallel to the free surface of the column of fluid, and z is the distance from the free
surface. Since a fluid does not support shear forces, static equilibrium requires that equal forces must
occur in all directions at that point. The pressure force experienced at this point (D) is gAz, which is
the weight of the column of fluid above D. If the floating body is rigid, then the integration of the vertical
components of the pressure force over the surface of the body is the buoyant force. Thus, the buoyant
force (F B) is

FB  S gz cos dS (6-1)


176
Elements of Ocean Engineering
Figure 6-1. Example of ship lines drawing. (Reprinted with permission from Lewis, E. V. (1988). Principles of naval architecture (2nd ed.).
Jersey City: Society of Naval Architects and Marine Engineers.)
Chapter 6 Floating and Submerged Body Hydrodynamics 177

Figure 6-2. Buoyant forces acting on a floating body.

where  is the angle of inclination of any portion of the body surface (S) from the horizontal. For equi-
librium, the buoyant force must equal the weight of the body. Thus, the weight of the floating body and its
contents is equal to the weight of the displaced fluid or the displacement. Similarly, the mass of a floating
body and its contents is equal to the mass of the displaced fluid. For a totally submerged body, such as a
submarine, the upward buoyant force must equal the weight of the fluid it displaces.
The volume of the underwater, or submerged, portion of a floating body can be calculated with the
use of numerical techniques such as Simpson’s or trapezoidal rules. If the submerged portion is a simple
shape, it can be calculated directly. This underwater volume is called the volume of displacement (∇).
Therefore, the weight of the displaced fluid, or the displacement weight (W) is

W g∇  ∇ (6-2)

where  is the fluid density, g is the gravitational acceleration, and  is the specific weight of the fluid.
According to Archimedes’ principle, this displacement weight is equal to the weight of the floating body
and its contents. In English units, W is usually expressed in long tons (2240 lb per ton), and in International
System of Units (SI) terms, it is usually expressed in metric tons (1000 kg per metric ton).
The centroid of the submerged portion of a floating body is called the center of buoyancy (B), and
it represents the point through which the buoyant force acts. The density of the fluid medium affects the
displaced volume of fluid and hence the draft of the vessel. Since seawater is denser than freshwater, the
draft of a vessel will increase as it travels from ocean to inland waters that are less saline or to freshwater
lakes (e.g., the Great Lakes).

6.2.2 Coefficients of Form

A number of coefficients are used to express the characteristics of a vessel’s shape, body plan sec-
tions, and waterlines. The block coefficient (CB) is defined as the ratio of the volume of displacement (∇)
of the molded form up to any waterline to the volume of a rectangular prism of length, breadth, and depth
equal to the length, breadth, and mean draft of the vessel at the waterline. In equation form, it is given as

CB  _∇ (6-3)
LBT
where L is the length, B is the breadth, and T is the mean molded draft. The values used for L, B, and
T vary, so there can be slight differences. For example, the LBP may be used for L, or the length at the
waterline may be used. Typical values of CB vary between 0.36 for high-speed vessels to 0.92 for slow-
178 Elements of Ocean Engineering

speed cargo vessels (Lewis, 1988). The midship coefficient (CM) is the ratio of the immersed area of the
midship station to that of a rectangle of equal area to the product of the molded breadth and molded draft
at the midship section. The common range of CM values is between 0.75 and 0.995 (Lewis, 1988). It is
expressed as
Immersed area at midship section
CM  ___ (6-4)
BT
The prismatic coefficient (CP) is the ratio of the volume of displacement (∇) to a prism with a length
equal to the length of the ship and a cross-section equal to the midship section area. The usual range of
values is 0.5–0.9 (Lewis, 1988). It is expressed in equation form as

CP  ____ ∇ (6-5)
L  Immersed area at midship section
or

CP  _ ∇ CB
_ (6-6)
LBTCM CM
The waterplane coefficient (CWP) is the ratio of the waterplane area (AWP) to the area of a circum-
scribing rectangle, and is expressed as
AWP
CWP  _ (6-7)
LB
This coefficient may be calculated at any draft, and typical values at the DWL range between 0.65
and 0.95 (Lewis, 1988). The vertical prismatic coefficient (CVP) is the ratio of the displacement volume (∇)
to the volume of a cylindrical solid with a depth equal to the molded mean draft and with a uniform hori-
zontal cross-section equal to the waterplane area at that draft. It is defined as
∇ CB
CVP  _ _ (6-8)
CWP LBT CWP
Finally, there is the volumetric coefficient (CV), which is defined as the displacement volume di-
vided by the cube of one-tenth of the length of the vessel, and is expressed as

CV  _∇ (6-9)
_
10( )
L 3

Common values of CV range from 1.0 for long, light ships (e.g., a destroyer) to 15 for short,
heavy vessels (e.g., a trawler). Certain ratios of the principal vessel dimensions are also frequently used
in ratio form, such as the length-to-breadth ratio (L/B, 3.5–10), length-to-draft ratio (L/T, 10–30), and
breadth-to-draft ratio (B/T, 1.8–5). Since different defi nitions of length can be used in these ratios and
coefficients, it is important to define which definition is being employed.
It is necessary to calculate areas, centroids, volumes, and other geometrical characteristics for a
vessel floating at a prescribed waterline (draft). Because the waterplane areas are not simple shapes, it is
usually necessary to use numerical techniques to calculate these characteristics. In most cases, Simpson’s
rules for integration are used. These rules are defined in Table 6-3 and further explained by Lewis (1988).
Many years ago, mechanical integrators were used to evaluate plane areas, and the planimeter is an example
of such an instrument. Numerical integration techniques are currently used.

6.2.3 Curves of Form

In the design of vessels, the curves of form are normally produced for the vessel at a series of drafts.
The curves of form consist of a number of hydrostatic properties of the vessel as a function of the different
Chapter 6 Floating and Submerged Body Hydrodynamics 179

Table 6-3. Integration Rules Using Simpson’s First and Second Rules Combined
Station Ordinate Simpson Multiplier Product
SIMPSON’S FIRST RULE PRIMARY
0 yo 1 yo
1 y1 4 4 y1
2 y2 2 2 y2
3 y3 4 4 y3
4 y4 17/8 17 y4 /8
5 y5 27/8 27 y5/8
6 y6 27/8 27 y6 /8
7 y7 9/8 9 y7/8
 product
∆y
Area  ___ ( product), where ∆y is station spacing
3
SIMPSON’S SECOND RULE PRIMARY
0 yo 1 yo
1 y1 3 3 y1
2 y2 3 3 y2
3 y3 17/9 17 y3/9
4 y4 32/9 32 y4 /9
5 y5 16/9 16 y5/9
6 y6 32/9 32 y6 /9
7 y7 8/9 8 y7/9
3∆y
____
Area  ( product), where ∆y is station spacing
8

drafts. These curves are used by the vessel operation personnel and are also known as the curves of form, or
hydrostatic curves (Fig. 6-3). The range of drafts should extend from below the lightest possible operational
draft to the deepest possible draft.

6.2.4 Hydrostatic Calculations

Hydrostatic calculations include computations of the waterplane area (AWP), tons per unit immersion
(TPI or TP cm), longitudinal center of flotation (LCF), transverse metacentric radius (BM), height of the
transverse metacenter (KM), longitudinal metacentric radius (BML), height of the longitudinal metacenter
(KML), molded displacement, total displacement (∇), displacement (D), longitudinal center of buoyancy
(LCB), and vertical center of buoyancy (KB). Table 6-4 shows example calculations for waterplane charac-
teristics. Simpson’s first rule with half multipliers was used for these computations. Similar computations
for displacement, longitudinal center of buoyancy, volume of displacement, and height of center of buoyancy
are illustrated in Table 6-5 and Table 6-6.
Two approximate expressions for determining the height of the center of buoyancy for conventional
ship forms were given by Morrish (1892) (equation 6-10) and Posdunine (1925) (equation 6-11):

1 _
KB  _
3 2 [
5T  _∇
AWP ] (6-10)

KB  T _
[
AWP

AWP  _
T
] (6-11)
180 Elements of Ocean Engineering

Figure 6-3. Example curves of form. (Reprinted with permission from Lewis, E. V. (1988). Principles of naval architecture
(2nd ed.). Jersey City: Society of Naval Architects and Marine Engineers.)

The hydrostatic calculations for the curves of form are repetitious, and consequently computers
(mainframe and personal computers) are now used extensively to perform them. The U.S. Navy (USN)’s
Ship Hull Characteristics Program (SHCP) is used widely in shipyards and design offices throughout the
United States (NAVSEA, 1976), but other types of commercial software are also available.

6.2.5 Static Stability

The stability of submerged or floating vessels is usually classified as being either static or dynamic.
The concept of static stability is discussed in this section; for further discussion about dynamic stability,
see Lewis (1988), Tupper (1996), and Zubaly (1996). Some dynamic effects, such as forces due to wind,
can be considered as static when steady conditions occur. Another stability problem occurs when vessels go
aground or when they are damaged and flooding occurs.
Floating and submerged vessels are usually rigid bodies, and for a state of equilibrium to exist, the
resultant forces and moments acting on the vessel must equal zero. Static equilibrium of a floating vessel
is concerned with the vessel in an upright position in still water. This means that the resultant gravitational
Chapter 6 Floating and Submerged Body Hydrodynamics 181

Table 6-4. Waterplane Characteristics for a Particular Waterline


Station Half- 1/2 SM Product Moment Product Moment Product (Half- Product
breadth Arm Arm breadth)
(m) Squared cubed
0 0 0.25 0 5.0 0 25.0 0 0 0
0.5 1.245 1.0 1.245 4.5 5.603 20.25 25.211 1.93 1.93
1 3.140 0.5 1.570 4.0 60280 16.0 25.120 30.96 15.48
1.5 5.359 1.0 5.359 3.5 18.757 12.25 65.648 153.90 153.90
2 7.597 0.75 5.698 3.0 17.094 9.0 51.282 438.46 328.84
3 10.956 2.0 21.912 2.0 43.824 4.0 87.648 1315.09 2630.18
4 12.007 1.0 12.007 1.0 12.007 1.0 12.007 1731.03 1731.03
5 12.039 2.0 24.078 0 0 0 0 1744.90 3489.8
6 12.039 1.0 12.039 1.0 12.039 1.0 12.039 1744.90 1744.90
7 11.899 2.0 23.798 2.0 47.596 4.0 95.192 1684.73 3369.46
8 10.271 0.75 7.703 3.0 23.109 9.0 69.327 1083.52 812.64
8.5 8.417 1.0 8.417 3.5 29.460 12.25 103.108 596.31 596.31
9 5.962 0.5 2.981 4.0 11.924 16.0 47.696 211.92 105.96
9.5 3.057 1.0 3.057 4.5 13.756 20.25 61.904 28.57 28.57
10 0 0.25 0 5.0 0 0 0 0 0
Sl  129.8 S2  34.3 S3  656.2 S4  15009
154.99  15.499 m
L _
Station spacing (s), s  _
10 10

Waterplane area, AWP  S1 _4_ s  [(129.9)(20.67)]  2,683.8 m2


3
AWP __ 2683.8
Waterplane coefficient, CWP  _   0.719
L B [(154.99)(24.078)]
(2683.8)(1.025)
Tonnes per centimeter immersion  __  27.51
100
S
( )
Longitudinal center of flotation, LCF  _2 s  _
S1
34.3 15.499  4.10 m abaft Station 5
(
129.9 )
Transverse moment of inertia, IT  S4 _ () 4 s  (15,009)(6.8884)  103,390 m4
9
Volume of displacement, ∇  17,845, from displacement curve in Fig. 6-3
I 103,390
Transverse metacentric radius, BM  _T  _ 5.79 m
∇ 17,845
Lewis, E. V. (Ed.) (1988). Principles of naval architecture (2nd rev.). Jersey City: Society of Naval Architects and Marine Engineers.

forces (weight) acting downward and buoyant forces acting upward on the vessel are of equal magnitude and
act through the same vertical line.
There are three conditions of stability for floating or submerged vessels: stable, neutral, and
unstable. In a stable condition, the vessel returns to its original equilibrium state after an external force
or moment has been applied and removed. The external moment or force causes the vessel to change its
angular attitude, and when the force or moment is removed, the vessel returns to its original equilibrium
position. When a force or moment applied to a vessel causes it to change its angular orientation, and this
new orientation is maintained after the force or moment is removed, the vessel is considered to have neu-
tral stability. An unstable condition exists when a force or moment is applied to a vessel and the change
in orientation continues even after the force or moment is removed (i.e., the vessel goes from upright to
upside-down).
Floating and submerged vessels may be inclined either athwartship (heel or list) or longitudi-
nally (trim). Athwartship or transverse stability and longitudinal stability are usually discussed separately.
182 Elements of Ocean Engineering

Table 6-5. Example Calculation of Displacement and LCB for a Particular Waterline
Station Area (m2) 1/2 Simpson Product Nondimensional Product
Multiplier (SM) Moment Arm
0.07 0 0.0175 0 5.07 0
0.035 3.0 0.07 0.21 5.035 1.1
0 4.2 0.2675 1.12 5.0 5.6
0.5 12.7 1.00 12.70 4.5 57.2
1 22.6 0.50 11.30 4.0 45.2
1.5 35.1 1.00 35.10 3.5 122.9
2 50.6 0.75 37.95 3.0 113.9
3 83.3 2.0 166.60 2.0 333.2
4 106.1 1.0 106.1 1.0 106.1
5 113.7 2.0 227.40 0 0
6 107.6 1.0 107.6 1.0 107.6
7 81.4 2.0 162.80 2.0 325.6
8 44.0 0.7813 34.38 3.0 103.1
8.5 29.1 0.8438 24.3 3.5 85.1
9 17.4 0.8438 14.68 4.0 58.7
9.5 5.3 0.3138 1.66 4.5 7.5
9.565 3.3 0.13 0.43 4.57 2.0
9.63 0 0.0325 0 4.63 0
S1  944.33 S2  95.6
Sectional area curve extended beyond stations 0 and 9.5 to extremities, as shown in Fig. 6-1, and read at midpoint between last
station and extremity. Simpson’s multipliers proportioned accordingly. At station 0.035, 1/2 SM 0.5(4)(0.035)  0.07; at
station 0, 1/2 SM  0.25  0.0175  0.2675; at station 8, 1/2 SM  0.5[1.0  (9/8)(0.5)]  0.7813 (first and second rules);
at station 8.5 and 9, 1/2 SM  [(0.5)(9/8)(3/2)]  0.8438; at station 9.5, 1/2 SM  0.5[(9/8)(1/2)  (0.65/1.0)]  0.3138; at
station 9.565, 1/2 SM  [(1/2)(4)(0.065/1.0)]  0.13.
Volume of displacement, ∇  S1(2/3) s  944.33 (2/3) 15/499  9,757 m3
Displacement, ∆  1.025(9757)  10,000 t (seawater)
S
(
Longitudinal center of buoyancy, LCB  __2 (s)  ______
S1 944.33 )
95.6 (15.499)  1.57 m forward of station 5

Lewis, E. V. (Ed.) (1988). Principles of naval architecture (2nd rev.). Jersey City: Society of Naval Architects and Marine Engineers.

Transverse stability is of the most concern because it usually determines whether a vessel will or will not
capsize, or turn upside-down. In the English system, the displacement, weight, and buoyant forces are
expressed in long tons (or lb), but in the SI system, the displacement (∆) is expressed as mass in metric tons
(or kg). Righting and heeling moments are expressed as metric ton-meters (t-m) or long ton-feet (ft-t). As
previously discussed, the weight or displacement is determined from the curves of form and draft marks.
The center of gravity (G) location is determined by experiment or calculation. Calculating the
location of G requires knowledge of the weight of all items on a vessel and their location with respect to a
selected coordinate system. The total weight of the vessel is assumed to act through the center of gravity.
An inclining experiment is conducted to experimentally determine the center of gravity location (G), which
is established relative to three reference planes. The vertical location of the center of gravity (VCG) is refer-
enced to a horizontal plane passing through the baseline (keel). A vertical transverse plane passing through
the midship location or through the forward perpendicular is the reference for the longitudinal center of
gravity (LCG). The transverse center of gravity (TCG) is referenced to a vertical plane passing through the
vessel centerline. Usually, G is located very near to the centerline plane.
The buoyant force (F b) acting on the vessel is equal to the weight of the displaced fluid, and
it acts vertically upward through the center of buoyancy (B). The submerged volume is normally
Chapter 6 Floating and Submerged Body Hydrodynamics 183

Table 6-6. Example Calculation of the Volume of Displacement and Height of the Center of Buoyancy
Height Above Waterplane Area Multiplier for Product Multiplier for Product
Baseline (m) (m2) Volume Moment
0 194 5 970 3 582
1 1714 8 13712 10 17140
2 1976 1 1976 1 1976
S1  12706 S2  15746
1 1714 5 8570 3 5142
2 1976 8 15808 10 19760
3 2137 1 2137 1 2137
S3  22241 S4  22765
Values for 1 m draft
s S  ___
Volume of displacement, ∇  ___ 1 (12,706)
[ ]
12 1 12

Moment of volume about baseline, M∇  _( )


s2  S  ___
24 2 24 ( )
12 15,746  656.1 m4

M∇ _____
___ 656.1
Height of center of buoyancy, KB    0.62 m
∇ 1059
Values for 2 m draft
s S  ___
Added volume of displacement, δ∇ (1 to 2 m)  ___ 1 (22,241)  1853 m3
[ ]
12 ( 3 ) 12
Total volume, ∇  ∇  d∇  1059  1853  2912 m3
2 2
s (S )  ___
Moment of added volume, δM∇ (1 to 2m) about 1m waterline  ___ 1 (22765) m4  948.5 m4
24 4 24
Moment of added volume about baseline  948.5  1853(1  0)  2801.5 m4
Moment of total volume about baseline, M∇  656.1  2801.5  3457.6 m4
M∇ ______
Height of center of buoyancy, KB  _____  3457.6  1.19 m
∇ 2912

converted to the weight or mass of the displaced fluid and is termed the displacement (W or ∆). The
orientation or attitude of a floating vessel is determined by the interaction of the forces of buoyancy
(F b) and weight (Fg). When no other forces are acting, the vessel will immerse until the forces of
buoyancy and weight are equal, and will rotate until the centers of gravity (G) and buoyancy (B) are
in the same vertical line. A submerged vessel will rotate until the center of gravity is directly below
the center of buoyancy. An important difference between floating and submerged objects is that for a
floating body, the center of buoyancy location changes when the object is rotated, whereas it has a fi xed
location for a submerged object.
A vessel’s position is determined by two types of hydrostatic moments that act on floating and
submerged vessels: the righting moment and the heeling moment. The righting moment occurs when the
vessel inclines to a position where the forces of weight (Fg) and buoyancy (Fb) cause the vessel to return to
an upright and vertical position. Heeling moments occur at any angle of inclination, and the forces of weight
and buoyancy cause the vessel to move away from the upright position.
The effect of the location of the center of gravity in floating and submerged vessels is illustrated
in Fig. 6-4. Lowering the center of gravity increases the stability by increasing the separation of the forces
of weight and buoyancy. Lowering the center of gravity can also change a heeling moment to a righting
moment. The longitudinal separation of B and G affects the draft and trim of the vessel. In a submerged
body, the center of buoyancy does not move, and positive stability requires that G remain below B. An
unstable condition occurs when G moves above B.
Upsetting forces may act on a vessel, causing it to heel, and the forces of weight (Fg) and buoy-
ancy (Fb) must create a righting moment to offset the heeling moment and prevent capsizing or excessive
184 Elements of Ocean Engineering

Figure 6-4. Location of center of gravity effects on stability, where MWL is the mean waterline.

heel. Examples of upsetting forces include wind, lifting of a heavy weight over the side, high-speed turns,
grounding, mooring lines, towlines, shifting of onboard weights, and entrapped water on deck.
In the design process, it is important to catalog all weights and their locations so that the overall
weight and location of the center of gravity can be estimated. The height of the center of gravity and weight
estimates are critical for determining the adequacy of the vessel stability. The weight and longitudinal cen-
ter of gravity determine the drafts at which the vessel will float. The distance G is offset from the vessel
centerplane and determines the list of the vessel. An example calculation of the weight and height of G
above the keel is shown in Table 6-7.

Table 6-7. Example Calculation of Vessel Weight and Vertical Center of Gravity
Item Weight Vertical Location Moment
(lb) (ft) (ft-lb)
1 400 2 800
2 900 1.5 1350
3 600 3.0 1800
4 750 0.5 375
5 500 2.5 1250
6 1200 0.75 900
7 300 4.0 1200
8 150 5.0 750
9 400 1.75 750
 Moments  ___  Weight  Vertical Location
VCG  _
 Weights  Weights
VCG  9125
_  1.75 ft
5200
KG  1.75 ft
Similar procedures are used to evaluate transverse and longitudinal CGs.
Chapter 6 Floating and Submerged Body Hydrodynamics 185

6.2.6 Metacentric Height

When a symmetric vessel is heeled to a very small angle , the center of buoyancy moves off the
vessel centerline as a consequence of the inclination (Fig. 6-5). The resultant forces due to weight (Fw) and
buoyancy (FB) are separated by a distance GZ called the righting arm. A vertical line is drawn through
the new center of buoyancy (B) and intersects the original vertical line drawn through the initial center of
buoyancy. The intersection of these two lines is called the transverse metacenter (M). The location of the
metacenter varies with the displacement and trim, but it is constant for a given draft. The metacenter is
essentially stationary for small vessel angles of inclination ( 7–10 degrees).
If the locations of G and M are known, the calculation of the righting arm is then accomplished by

GZ  GM sin (6-12)

The distance GM is an important index of the transverse stability at small angles of heel, and is
termed the transverse metacentric height. GM is positive when M is above G and negative when M is below
G. A vessel is stable when GM is positive, and unstable when GM is negative. The GM is used to measure
the initial stability of the vessel. Other useful relationships include the following:

IT
BM _ (6-13)

GM  KM  KG (6-14)

GM  KB  BM  KG (6-15)

For a submerged submarine, the center of buoyancy is stationary with respect to any angle of
inclination. When the submarine is submerged, B and M coincide. Therefore, the initial stability of a
submerged body is GB instead of GM. For floating bodies, the metacentric height is also used to deter-
mine the moment to heel 1 degree, moment to trim 1 cm, and moment to trim 1 in. These parameters are
expressed as

Moment to heel 1 degree  ∆ GM sin(1) (6-16)

∆ GML
Moment to trim 1 centimeter  _ (t  m) (6-17)
100 L

W GM
Moment to trim 1 inch  _L (ton  ft) (6-18)
12 L

Figure 6-5. Schematic of transverse metacenter (M) and righting arm (GZ).
186 Elements of Ocean Engineering

where L is the ship length and GML is the longitudinal metacentric height. The period of roll in still water
is estimated by

C Beam
T_ _ (6-19)
 GM

where C is a constant obtained from measured data. The value of C does not vary much and can be reason-
ably assumed as 0.8 for surface ships and 0.67 for submarines (Lewis, 1988). This expression can be used
to estimate GM when the period of roll is measured.

6.2.7 Cross Curves of Stability

The moment of weight and buoyancy that is necessary to return a vessel to its upright position is
determined by ascertaining the distance between the center of gravity and the vertical line through the cen-
ter of buoyancy for a particular inclination angle and displacement. This distance is called the righting arm
(GZ) and is illustrated in Fig. 6-6.
The cross curves of stability can be used to illustrate the righting arm for any probable value
of displacement and for several angles of heel. It is convenient to assume a location for the center
of gravity, such as point A in Fig. 6-7 or the baseline. The distance (d) is calculated for several
waterlines (displacements) and various angles of heel. If G is above A, the actual values of GZ are
obtained from

GZ  d  AG sin  (6-20)

If G is below A, then

GZ  d  AG sin  (6-21)

A is often taken at the baseline (K), and the value KG sin  is always subtracted from d. A sample
set of cross curves of stability is illustrated in Fig. 6-7, which shows the distance (d) for several displace-
ments and angles of heel.
There are different ways to obtain the cross curves of stability. The manual method uses transverse
sections of the vessel. The most common method is to use digital computers along with inputs of the vessel
offsets, such as those provided by the USN’s SHCP mentioned above. Another method is to use mechanical
integrators. All of these methods are described by Lewis (1988).

Figure 6-6. Illustration of transverse righting arms.


Chapter 6 Floating and Submerged Body Hydrodynamics 187

Figure 6-7. Example of cross curves of stability. (Reprinted with permission from Lewis, E. V. (1988).
Principles of naval architecture (2nd ed.). Jersey City: Society of Naval Architects and Marine Engineers.)

6.2.8 Curves of Static Stability

The static stability curve is a graph of the righting arm versus the angle of inclination (heel) for a
given vessel load condition. An example curve obtained from the cross curves of stability is illustrated in
Fig. 6-8. Calculations of the righting arms using values from Fig. 6-7 for a displacement of 30,000 metric
tons and KG of 8.8 m are tabulated in Table 6-8.
Vessel stability criteria are governed by several international and national organizations, including
the International Maritime Organization (IMO), U.S. Coast Guard (USCG, 1990), American Bureau of
Shipping (ABS, 2000), and USN (1987). The metacentric height is an approximate indication of the stabil-
ity, but a more accurate indication can be achieved by comparing the righting arm curve (static stability
curve) with the heeling arm curve, as illustrated in Fig. 6-9. The residual dynamic stability is represented
by the cross-hatched area between the righting and heeling arm curves. Satisfactory stability is attained if
1) the heeling arm at the intersection of the righting arm and heeling arm curves (point C) is not greater than
0.6 of the maximum righting arm, and 2) area A1 is not less than 1.4 A2 where area A2 extends 25 degrees
to windward from point C (Lewis, 1988).

Figure 6-8. Typical static stability curve with center of gravity on and off the centerline.
188 Elements of Ocean Engineering

Table 6-8. Example Calculation of Righting Arm from the Cross Curves of Stability
Angle of inclination, , degrees 0 15 30 45 60 75
Righting arms, m, from Fig. 6-8 0 1.17 2.44 3.57 3.79 3.41
Adjustment of actual KG, (6.1–8.8) sin  0 0.70 1.35 1.91 2.34 2.61
Free surface effect 0 0.05 0.06 0.05 0.05 0.04
Righting arms, m 0 0.13 0.77 1.40 1.25 0.68

Figure 6-9. Example of USN criteria for stability in wind and waves. (Reprinted with permission from Lewis, E. V. (1988).
Principles of naval architecture (2nd ed.). Jersey City: Society of Naval Architects and Marine Engineers.)

6.3 RESISTANCE AND PROPULSION

6.3.1 Resistance

Vessels that move through ocean waters, or any other fluid, must have sufficient thrust from a pro-
pulsion device to overcome the resistance to movement of the vessel through the water or fluid. The resis-
tance is different for a completely submerged vessel and a vessel that moves through the water at the surface.
Resistance forces are usually distributed over the entire wetted surface of the vehicle. Propulsion forces are
applied at the stern of the vehicle, where the propeller is located.
In real fluids, the relative motion between the marine vehicle and fluid results in the develop-
ment of a boundary layer. It is assumed that no relative motion exists at the surface of the vehicle, or, in
other words, that the velocity of the fluid and the vehicle are the same (otherwise known as the no-slip
condition). Assuming that the vehicle is moving through a fluid at rest, the velocity of the fluid near
the vehicle will vary from the velocity of the vehicle to zero far away from the vehicle. If the vehicle
is stationary and the fluid is moving by it, the velocity will vary from zero at the vehicle surface to the
fluid velocity at some distance away from the vehicle. The layer of fluid between the vehicle surface
and the location where it approaches the freestream velocity is called the boundary layer. Since the
velocity in the boundary layer approaches the freestream velocity asymptotically, the thickness is not
easily defi ned, but is usually assumed to be the distance at which the velocity in the boundary reaches
95% to 99% of the freestream velocity. In many cases, the fluid separates from the rear of the vehicle
and forms a wake, and as a result there is a pressure difference between the bow and stern of the ve-
hicle. For surface vehicles, there is an additional phenomenon resulting from the generation of waves
developed at the air-water interface.
The only horizontal force is that caused by drag or resistance when steady motion is assumed
and the effects of lift and side forces are neglected. The weight and buoyancy forces are in the vertical
direction. For the submerged vehicle, the total resistance (drag) is the summation of the forces resulting
Chapter 6 Floating and Submerged Body Hydrodynamics 189

from the summation of the pressure and shear stress distribution over the wetted hull surface. Therefore,
the total resistance (RT) is the sum of the pressure resistance (R P) and skin-friction resistance (R F), and
it is expressed as

RT  RP  RF (6-22)

A surface vehicle moving at the air-water interface results in the added complication of including
the resistance due to wave-making. In this case, the total resistance must include the wave-making contribu-
tion to the resistance (RW), and is written as

RT  RP  RF  RW (6-23)

For the vehicle, a dimensionless resistance coefficient is defined as


RT
CT  _ (6-24)
1 U2A
_
2
2
where (1/2)U A is the dynamic pressure, and A is a characteristic area of the body. The total resistance
equation is now expressed in dimensionless form as

CT  CP  CF  CW (6-25)

CP is the pressure coefficient and CF is the skin-friction coefficient, which are defined as
RP RF RW
CP  _ ; CF  _ ; CW  _ (6-26)
_1_ U2A _1_ U2A _1_ U2A
2 2 2
For bluff bodies such as a sphere, cylinder, or disc normal to flow, the boundary layer separation
produces a large wake behind the body that results in a large pressure drag, and consequently CP is much
greater than CF. When a body is streamlined, the boundary layer separation occurs near the rear or not at
all, and the skin friction resistance dominates.
It is very difficult to determine the total resistance of a vessel (RT) because there is no general re-
lationship between the coefficient of resistance (CT) and the Reynolds (Re) and Froude (Fr) numbers. The
Reynolds number is the ratio of the inertia to viscous forces, and the Froude number is the ratio of the iner-
tial to gravitational forces. The Reynolds and Froude numbers are expressed as
_
Re  UL/v, Fr  U/gL (6-27)

Equivalency of the prototype and model Re and Fr ensures a dynamic similarity between the model
and the prototype. Thus, laboratory modeling is used to determine the resistance of the model vessel, and
scaling laws are used to evaluate the prototype resistance. The CT is split into its components and each com-
ponent is evaluated using the appropriate scaling law. The Froude procedure is as follows:

1. Determine RT for the model when (Fr)M  (Fr)P, where subscripts M and P refer to model and
prototype, respectively.
2. Evaluate the skin friction resistance RF for the model using data from submerged planks of equal
surface area, length, and finish, and model Reynolds number.
3. Determine the residual resistance (RR)M from (RR)M  (RT)M  (RF)M and determine (CR)M. CR
is assumed to be a function of the Froude number only, so that (CR)M  (CR)P.
4. Compute the frictional resistance RF for the plank corresponding to the Reynolds number of the
prototype and deduce (CF)P.
5. Evaluate the prototype resistance from (CT)P  (CF)P  (CR)M.
190 Elements of Ocean Engineering

Once CT is determined, the resistance can be evaluated from


1 C AU2
RT  _ (6-28)
2 T
The assumptions are good for streamline hull forms and Froude numbers greater than 0.35. Also,
very fast digital computers, such as workstations, parallel processing computers, and super computers, can
be used to perform numerical computations of resistance and obtain good results. More-detailed descrip-
tions of resistance determination can be found in Lewis (1988) and Clayton and Bishop (1982).

6.3.2 Propulsion

Marine vessels are propelled through a fluid medium by some type of propulsion device that gener-
ates enough thrust to overcome the vessel resistance. In the early days, thrust was provided by wind (sailing
vessels) or oars (human-powered vessels). Later, steam was used as the energy source to drive propulsors.
Currently, diesel fuel, gasoline, and natural gas are commonly used to drive propulsors. Some of the differ-
ent types of propulsors used on marine vessels are tabulated in Table 6-9.
To ensure the steady flow of a fluid through a screw propeller, the propeller is replaced by a thin ac-
tuator disc. The fluid passing through the disc is assumed to experience a sudden increase in pressure while
the fluid axial velocity remains continuous. The magnitude of the thrust developed by the disc is determined
by evaluating the change in the axial momentum of the fluid. The flow relative to an actuator disc in open
water is illustrated in Fig. 6-10. The disc advances forward at a steady velocity (Va), called the velocity of
advance, that is parallel to the axis of rotation. The fluid is assumed to have a constant density and to be
stationary, inviscid, and infinite in extent. It is assumed that the propeller disc uniformly accelerates all of
the fluid passing through it, and that therefore the thrust generated is uniform over the disc. The axial veloc-
ity relative to the disc increases from Va far upstream of the disc to Va (1  a) at the disc to Va (1  b) far
downstream of the disc. The term “a” is the axial inflow factor.
As shown in Fig. 6-10, the uniform pressure on the upstream face of the disc is termed p, and that
on the downstream face is termed p  p. Using the Bernoulli theorem between the upstream location and
upstream side of the actuator disc yields
1 V2  p  _
po  _ 1 [V (1  a)]2 (6-29)
a a
2 2
On the downstream side, it is
1 [V (1  a)]2  p  _
p  p  _ 1 [V (1  b)]2 (6-30)
a o a
2 2

Table 6-9. Summary of Typical Marine Propulsors


Propulsor Advantage Disadvantage Comments
Oars, paddles human-powered small, unsteady thrust small boat use only
Paddle wheel efficient requires low-speed engine used with steam power
Vertical axis propellers efficient, high degree of weight and cost used on ferries
speed, control, and thrust
Sails free power ineffective without wind used for sport and recreation
Hydraulic jets no parts external to vessel large duct losses efficiency increases at high
speeds
Pump jet reduced duct length costly and difficult to install used in shallow draft vessels
Kort nozzle high thrust at zero speed cavitation occurs earlier than used on tugs and large
pump jet tankers
Screw propeller reliable, simple, and efficient cavitation limits performance most common propulsor
Clayton, B. R., & Bishop, R. E. D. (1982). Mechanics of marine vehicles. Houston: Gulf Publishing Co.
Chapter 6 Floating and Submerged Body Hydrodynamics 191

Figure 6-10. Flow relative to an actuator disk in open water.

Combining equations 6-29 and 6-30 and solving for p yields


1 b(2  b)V2
p  _ (6-31)
a
2
The thrust of the fluid on the disc of area (A) is
1 Ab(2  b)V2
T  Ap  _ (6-32)
a
2
The force of the disc on the fluid is written as

T  (mass flow rate) [Va (1  b)  Va] (6-33)

Using equations 6-32 and 6-33, it is shown that b  2a, and thus the thrust is expressed as

T  2 Aa (1  a) V2a (6-34)

The velocity at the disc is the arithmetic mean of the velocities well upstream and downstream of
the disc. Using energy concepts, the ideal efficiency i for the actuator disc model is

i  _ 1 (6-35)
1a
The equation for thrust can be used to define a thrust coefficient CT as

CT  _ T  4a(1  a) (6-36)
1 AV2
_
a
2
and the inflow factor is determined from

a_
2 ( [
1 1C 21
T)
_1_
] (6-37)

The ideal efficiency is now expressed as

i  __ 2 (6-38)
_1_
1  ( 1  CT )2
The above equation is illustrated in Fig. 6-11 and represents the highest possible propeller efficiency. An
ideal efficiency of 0.7 is common.
192 Elements of Ocean Engineering

Figure 6-11. Relationship between ideal efficiency (␩i) and thrust coefficient (CT). (Reprinted with permission from
Clayton, B. R., & Bishop, R. E. D. (1982). Mechanics of marine vehicles. Houston: Gulf Publishing Co.)

When a propeller is placed at the stern of the vessel, an effect is caused by the hull-propeller in-
teraction. The difference between the bare hull resistance and thrust as a result of adding the propeller is
expressed as a thrust deduction fraction defined as
TR
t  _T (6-39)
T
and

RT  T(1  t) (6-40)

where the thrust deduction factor is always positive, and thus RT is less than T. There is also an effect result-
ing from the wake behind the vessel. The overall effect is that the velocity of advance is slightly less than
the forward velocity (U) of the vessel. A wake fraction is defined to account for the wake effect by
UV
w  _a (6-41)
U
or

Va  U(1  w) (6-42)

The total propeller efficiency (T) is usually expressed as the product of the open water efficiency
(o), hull efficiency (H), and relative rotative efficiency (R). The open water efficiency is defined as
ToVa
o  _ (6-43)
2 nQo
where To is the thrust delivered at a torque Qo for a relative rotative speed of n (revolutions per unit time)
when the velocity of advance is Va. The hull efficiency is expressed as
1t
H  _ (6-44)
1w
The overall efficiency is then expressed as

T  O H R (6-45)

Some representative values of the wake fraction, thrust deduction fraction, relative rotative effi-
ciency, hull efficiency, and overall efficiency are tabulated in Table 6-10.
Chapter 6 Floating and Submerged Body Hydrodynamics 193

Table 6-10. Representative Values of Efficiency, Wake Fraction, and Thrust Deduction Factor
Vehicle Type 1-␻ 1-t ␩H ␩R ␩T
Single-propeller submarine 0.50–0.80 0.80–0.90 1.10–1.45 1.00–1.10 0.72–0.93
Single-propeller merchant ship 0.55–0.85 0.75–0.85 1.10–1.30 0.99–1.09 0.72–0.86
Dual-propeller merchant ship 0.83–0.95 0.82–0.89 0.96–0.97  0.62–0.70
Dual-propeller destroyer 0.93–1.01 0.9–0.99   0.61–0.70
Four-propeller ship 0.87–0.99 0.82–0.95   0.60–0.67
Ducted-propeller cargo ship 0.66 0.8 1.21 0.98 0.80
Clayton, B. R., & Bishop, R. E. D. (1982). Mechanics of marine vehicles. Houston: Gulf Publishing Co.

A propeller converts the power supplied by the propulsion motor into thrust as efficiently as pos-
sible, but this process always produces losses. The power delivered by the propulsion device (electric or
hydraulic motor) to the propeller shaft is the shaft horsepower (SHP). The effective horsepower (EHP) is the
product of the vessel speed and the hull resistance expressed in the units of horsepower as
RTV
EHP  _ (6-46)
550
where the constant 550 is the conversion of ft-lb/s to horsepower. The thrust horsepower (THP) is the
product of vessel speed and the thrust delivered by the propeller. The difference between EHP and THP is a
result of the thrust deduction (1  t) and the wake fraction (1  w), and is expressed as
1  t THP   (THP)
EHP  _
1w( )H (6-47)

The relationship between the SHP and EHP is illustrated in Fig. 6-12 and expressed in equation
form by
EHP
T  _ (6-48)
SHP
The total efficiency T does not include power reductions resulting from reduction gears or bear-
ing losses, which must be included if these devices are being used. In many cases, these are not present in
submersibles (Allmendinger, 1990). Typical values of the thrust deduction factor (1  t) and the wake frac-
tion (1  w) used to evaluate the hull efficiency are illustrated in Fig. 6-13. The determination of the hull
resistance requires model testing or calculation using equation 6-28 if the total resistance coefficient (CT) is
known. Values for streamlined hull shapes typical of a submarine hull are shown in Fig. 6-14.
Propeller model tests are usually conducted to assess the performance of marine propellers. The first
test is used to determine the open-water characteristics of the propeller. The model propeller is placed on a
long shaft in a water tunnel or beneath a towing carriage in a towing tank, and measurements of torque and
thrust are made for various velocities of advance and rotative speeds. For the second test, the model propeller

Figure 6-12. Propeller power relationships.


194 Elements of Ocean Engineering

Figure 6-13. Typical wake fraction and thrust deduction for the range of the propeller diameter/hull diameter ratio.
(Reprinted with permission from Allmendinger, E. E. (Ed.). (1990). Submersible vehicle systems design.
Jersey City: Society of Naval Architects and Marine Engineers.)

is installed in its proper position on a model hull and similar measurements are conducted as in the first test.
Dimensional analysis is applied using the parameters of torque, thrust, velocity of advance, diameter, rotative
speed, fluid density and viscosity, pressure, and vapor pressure of fluid. This analysis shows that

KT and KQ  function (J, ReD, N) (6-49)

where

thrust coefficient  KT  _ T (6-50)


n2D4

Figure 6-14. Resistance coefficients for streamlined hull shapes. (Reprinted with permission from Allmendinger, E. E.
(Ed.). (1990). Submersible vehicle systems design. Jersey City: Society of Naval Architects and Marine Engineers.)
Chapter 6 Floating and Submerged Body Hydrodynamics 195

Q
torque coefficient  KQ  _ (6-51)
n2D5
VA
advance coefficient  J  _ (6-52)
nD
VaD
propeller Reynolds number  ReD  _ (6-53)
_
p  pV
nominal cavitation index  N  _ (6-54)
_1_ V2
a
_ 2
and p is the upstream static pressure, D is the propeller diameter, n is the propeller rotative speed, pv is the
fluid vapor pressure, and  is the fluid dynamic viscosity. Typical open-water performance characteristics
curves for a noncavitating propeller are shown in Fig. 6-15.

Figure 6-15. Representative propeller performance curves. (Reprinted with permission from Clayton, B. R.,
& Bishop, R. E. D. (1982). Mechanics of marine vehicles. Houston: Gulf Publishing Co.)

6.4 BUOY SYSTEMS

Numerous types of buoy systems are placed in inland, coastal, and ocean waters for different pur-
poses. The objective of this section is to discuss the various types and uses of buoy systems and the static
analysis of these systems. Buoys, like other vessels, are placed in a very dynamic ocean environment and
consequently experience dynamic responses induced by ocean waves and currents. However, the dynamic
analysis of buoy systems is beyond the intended scope of this text, and the reader is referred to more
advanced publications such as that of Berteaux (1991).

6.4.1 Buoy Types and Uses

Buoys are floating or submerged objects that are typically cylindrical, spherical, discoid, or toroidal
in shape, and are moored to the seafloor with some type of anchor and mooring line (i.e., wire rope, chain,
synthetic line, or a combination thereof). Buoys are commonly used as navigation aids to mark ship chan-
nels, obstacles such as wrecks or other underwater hazards, and port entrances. It is estimated that in the
United States alone, some 24,000 buoys are deployed and maintained as aids to navigation by the USCG
(Berteaux, 1991). Buoys are often equipped with bells, whistles, and lights, and sometimes depend on wave-
induced motion to produce sounds. Heavy chain and clump anchors are frequently used to slack moor the
buoy system in depths usually not exceeding 45.7 m (150 ft). The offshore industry employs precise naviga-
tion aids for exploration and seismic surveys. Buoys may be moored with elastic tethers in water depths up to
305 m (1000 ft), and the elastic tethers are kept under high tension to limit the lateral movement of the buoy.
196 Elements of Ocean Engineering

Buoys are also used as markers and ship moorings. Marker buoys are used to locate and retrieve ob-
jects on the seafloor, such as anchors, lobster traps, sunken ships, and offshore pipelines. Large cylindrical
buoys are used to mark rig anchors for offshore drilling platform moorings. Buoys are also used to identify
the location of ship moorings used in bays and harbors when dock space is limited or when the water depth
is too shallow near the shore. Very large single-point mooring systems are used to moor oil tankers while
oil is being transferred to the tanker for eventual transport to land.
Buoys are also used for weather stations and oceanic platforms. The U.S. National Data Buoy
Center has a network of some 50 moored weather stations that measure and report wind speed and direction,
air temperature and pressure, wave conditions, and visibility. These systems benefit maritime, fishing, and

(a) (b)

(c) (d)
Figure 6-16. Examples of buoy systems. (a) USCG navigation buoy (reprinted with permission from U.S. Coast Guard
(USCG). (1990). Aids to navigation manual-technical. Department of Transportation. COMDINST MI 65003).
(b) Small watch circle marker buoy. (c) Offshore semisubmersible mooring with anchor marker buoy.
(d) USN free-swinging buoy. (Reprinted with permission from Navy Facilities Engineering Command (NAVFAC).
(1968). Design manual, harbor and coastal facilities, DM26).
Chapter 6 Floating and Submerged Body Hydrodynamics 197

Figure 6-17. Examples of buoy hull forms.

offshore industries, as well as the general public. Oceanic buoy platforms are either moored or free-drifting,
and can also be used to support sensors to measure a variety of oceanic parameters. In scientific and engi-
neering fields, buoy systems are used to measure wind, atmospheric pressure, air and surface temperature,
wave evolution and noise, ocean currents, water temperature, salinity and dissolved oxygen, distribution
of trace metals, sedimentation rates, ambient acoustic noise, light transmission and turbidity, and other
parameters. The data are transmitted by satellites and radio transmitters through the atmosphere, and by
acoustical systems for transmission through the water column. The data can be stored internally on tapes or
in computer memory. Examples of these buoy systems are shown in Fig. 6-16.
The basic hull forms for surface buoy systems are a disc (toroid or flat disc), sphere, cone, boat hull,
and spar, as shown in Fig. 6-17. Disc buoys are easier to design and build, and are more cost-efficient, but they
follow the waves in heave and slope. Sphere and cone buoys respond to the wave amplitude but not the slope.
Boat hull buoys align themselves with the direction of the waves, current, and wind, but still heave and roll
considerably. Spar buoys have the advantage of good stability. The techniques commonly used to moor surface
buoy systems in deep or shallow water (e.g., slack, taut, and three-point mooring) are illustrated in Fig. 6-18.

Figure 6-18. Example shallow water and deep water mooring techniques. (Reprinted with permission
from Berteaux, H. O. (1991). Coastal and oceanic buoy engineering. Woods Hole: Berteaux.)
198 Elements of Ocean Engineering

Typical subsurface buoy and mooring systems are shown in Fig. 6-19. These systems usually employ single-
point subsurface mooring that is recovered by disconnecting the mooring line near the anchor using a remote re-
lease mechanism (acoustic release is common). This configuration is economical and easy to deploy. When the
subsurface buoy is well below the free surface, the effect of wave motion is greatly reduced and the stability and
endurance of the mooring are improved. Multileg subsurface systems provide the advantages of reduced motion,
ample space below the buoy for instruments, and reliability, but they also cost more and are difficult to deploy.

Figure 6-19. Typical subsurface mooring configuration with an acoustic Doppler velocimeter (ADV) as the current speed sensor.

6.4.2 Static Analysis of Buoy Systems

The buoy weight (W) acts vertically downward through the buoy’s center of gravity (G). The buoy-
ancy of the buoy is equal to the weight of the displaced fluid, and this buoyant force (FB) is directed verti-
cally upward through the center of buoyancy (B). The center of buoyancy is also the centroid of the buoy’s
submerged volume. Since the buoy is either floating or submerged, it is also exposed to the flow of water
(ocean currents) past the buoy’s surface, which results in drag and lift forces acting on the buoy. The drag
force (FD) consists of friction and pressure forces as a result of tangential and normal stresses acting on the
buoy’s surface. If the flow past the buoy is laminar, the shear (tangential), or friction, stresses will predomi-
nate. If the flow is turbulent and the buoy shape is blunt, normal stresses (pressure) will predominate. The
resultant force (i.e., the resistance force) is obtained by integrating the shear and pressure stresses over the
area of the buoy. This force has two components: drag (in the direction of flow) and lift (in the direction
normal to the flow). The common “drag law” is used to determine the drag force (FD), and the similar “lift
law” is used to compute the lift force (FL). The equations for determining drag and lift are
1  C A U2 “Drag Law”
FD  _ (6-55)
2 D N

FL  _1 C A U2 “Lift Law” (6-56)


2 L L
where  is the fluid density, CD is the drag coefficient, CL is the lift coefficient, AN is the area projected on a
plane normal to the fluid flow, AL is the characteristic area used in determination of CL, and U is the average
fluid velocity. As discussed above, drag coefficients (Table 6-11) are a function of the Reynolds number
Chapter 6 Floating and Submerged Body Hydrodynamics 199

Table 6-11. Drag Coefficients for Two- and Three-Dimensional Body Shapes
Shape Reynolds Characteristic Drag
Number Length Ratio Coefficient
(R) (L/D, L/H) (CD)
Description Schematic
Rectangular plate normal to flow 103 L/H  1 1.16
H V 5 1.20
10 1.50
L  1.90
Circular plate normal to flow 103 – 1.12
D V

Circular cylinder with axis parallel 103 L/D  0 1.12


to flow D V 1 0.91
2 0.85
L 4 0.87
7 1.00
Circular cylinder with axis 105 L/D  1 0.63
perpendicular to flow 2 0.68
5 0.74
D
10 0.82
20 0.90
L V 40 0.98
 1.20
5  105 L/D  5 0.35
 0.34
Sphere 105 – 0.5
V 3  105 0.2

Hemisphere concave to flow 103 – 1.33


V

Hemisphere convex to flow 103 – 0.34


V

Ellipsoid with major axis perpendicular 5  105 L/D  0.75 0.6


D V
to flow 5  105 0.21
L

Ellipsoid with major axis parallel to 2  105 L/D  1.8 0.07


D V
flow
L
Airship hull D V 2  105 – 0.05
L

Solid cone (apex pointing in opposite – – 0.34


direction of velocity vector) D V
o
30
Solid cone (apex pointing in same – – 0.51
direction as velocity vector) D V
o
60
200 Elements of Ocean Engineering

(UD/v), where D is the characteristic buoy dimension and v is the fluid kinematic viscosity. The drag force is as-
sumed to act through the centroid of the projected area (AN). Waves acting on buoys also add a resistance force,
but this force is usually small and can be neglected.

6.4.3 Moored Subsurface Buoy

The free body diagram for a subsurface buoy moored in a current is illustrated in Fig. 6-20. For
static equilibrium to exist, the sum of the forces and moments must be zero.
By summing the forces at the attachment point, one obtains

TH  FD (6-57)

and

TV  FB  W  FL (6-58)

where TH and TV are the horizontal and vertical tension, respectively, in the mooring cable. FD, FB,
FL, and W are the drag force, buoyant force, lift force, and weight of the buoy, respectively. Equations
6-57 and 6-58 are used to estimate the tension in the mooring cable. The resultant cable tension is
determined by
_
T  T2H  T2V (6-59)

and the angle between the cable and the horizontal is found by
TV

 tan1 _
TH( ) (6-60)

Figure 6-20. Free-body diagram of a moored subsurface buoy in a current.

6.4.4 Static Cable Analysis

Buoy systems are moored with flexible cables that are subjected to ocean currents. As a result, the
cable tension increases and strong anchoring systems are needed. In this description, only two-dimensional
cases with coplanar currents are considered. The cable immersed weight (WI) is defined as

WI  FB  W (6-61)
Next Page

Chapter 6 Floating and Submerged Body Hydrodynamics 201

If WI is positive, the cable is positively buoyant (polypropylene rope), and if it is negative, the cable
is negatively buoyant (wire rope). The hydrodynamic force resulting from the current is expressed as

Rds  _ 1  C DU2 ds (6-62)


2 DN

where CDN is the normal drag coefficient and R is the drag force per unit length of cable normal to the cur-
rent. A schematic of the hydrodynamic forces is provided in Fig. 6-21.

Figure 6-21. Schematic of forces on a cable element.

When the cable is at an angle  with the horizontal direction of the current, the resistance is ex-
pressed as two components normal or tangential to the cable. The normal component FDds and tangential
component FTds are expressed as

FD ds  _ 1  C DU2 sin2 ds  Rsin2 ds (6-63)


2 DN

FT ds  _1_  CDN(D)U2 cos2 ds  Rcos2 ds (6-64)


2
where  is the ratio between the tangential and normal drag coefficients (CDT   CDN), CDN is the normal
drag coefficient, CDT is the tangential drag coefficient, and (D)ds is the surface area of the cable element
(ds). Normal (  90 degrees) and tangential (  0 degrees) drag coefficients for several cables are shown
in Fig. 6-23. The vector summation of the forces in the normal and tangential directions yields the classical
cable equilibrium equations:

Td  (FD  WI cos)ds (6-65)

dT  (WI sin  FT)ds (6-66)

These equations must be integrated numerically, and the result permits evaluation of tension and
cable geometry with known boundary conditions.
If heavy cables, such as chain or wire rope, are assumed, the resistance terms can be neglected. The
above equations reduce to
Td  WI cosds (6-67)

dT  WI sinds (6-68)

Integrating these equations from the origin where   0 to any point on the cable where the cable angle is
 yields
Tcos  To (6-69)

Tsin  WI s (6-70)

You might also like