Você está na página 1de 41

P1: KKK

November 11, 1997 10:39 Annual Reviews AR049-20


Annu. Rev. Fluid Mech. 1998. 30:61353
Copyright c 1998 by Annual Reviews Inc. All rights reserved
FLUID MECHANICS FOR SAILING
VESSEL DESIGN
Jerome H. Milgram
Massachusetts Institute of Technology, Cambridge, Massachusetts 02139;
e-mail: jmilgram@MIT.edu
KEY WORDS: sailing vessels, ships, numerical hydrodynamics, model testing, numerical
simulation, velocity prediction, hull design
ABSTRACT
The design of sailing vessels is an ancient art that places an ever-increasing re-
liance on modern engineering sciences. Fluid mechanics shares the forefront of
this reliance along with structural mechanics. This review focuses on the appli-
cation of uid mechanics in modern sailing vessel design. It is now common
practice to predict sailing performance with what are called velocity prediction
computer programs. The validity of the predictions is crucially dependent on ac-
curate modeling of the air and water forces on the vessel. This article reviews
existing methods of force modeling that include theory, experimentation, and
numerical uid mechanics and aerodynamics. The accuracy and reliability of
the numerical methods are considered on the basis of experimental results and
full-scale performance in areas for which the information is available.
1. INTRODUCTION
The last sailing vessel article to appear in the Annual Reviewof Fluid Mechanics
was the excellent and thorough article by Larsson (1990). Because considerable
literature both before and since has appeared on the subject, this article includes
only that information the author believes to be the most interesting, important,
and timely. Portions of this review are similar, if not identical, to earlier work
(Milgram 1996); however, this article is more concise and also includes new
information from more recently published literature as well as from important
older articles.
During the past seven years, the author participated in the design of several
International Americas Cup Class (IACC) racing yachts. Because of the high
613
0066-4189/98/0115-0613$08.00
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
614 MILGRAM
level of racing competition in the use of these vessels, all available capabilities
must be brought to bear on their design. This includes considerable experimen-
tal and numerical uid mechanics. Much of what is contained in this article
was developed for that activity.
2. EVALUATION OF DESIGNS AND DESIGN IDEAS
A sailing vessel is a complex interconnected system, and most design changes
inuence more than one kind of uid force. For example, to reduce the frictional
resistance of the hull by reducing its wetted surface while still maintaining its
length requires a reduction in beam that causes a reduction in heeling stability,
which, for prescribed sail shapes, leads to an increase in heel angle. The change
in heel angle not only changes the hull shape, it also changes the sail forces.
How does one determine whether the sum of all these effects is advantageous
or not? More importantly, how can one evaluate the effects if the sail shapes
are simultaneously changed to optimize them for the altered hull? Short of
complete, full-scale sailing experiments, answering these questions requires a
numerical method of predicting performance. A computer program that does
this is a velocity prediction program (VPP).
A brief description of VPPs follows, and a consideration of the individual
forces a VPP needs in order to work properly forms the framework for much
of the remainder of this article. However, there is a vast amount of litera-
ture about VPPs. The interested reader is referred to Kerwin (1975), where
the basis of the rst fundamentally sound VPP is described, and to Larsson
(1990), Milgram (1993), and Van Oossanen (1993) for additional information
and various perspectives on these programs.
2.1 Fundamental Principles for a Velocity
Prediction Program
The primary purpose of a VPP is to predict the boat speed for any prescribed
wind conditions and sailing angle (
T
) between the wind direction and the course
of the boat. In a computational model this is achieved by balancing counteract-
ing aerodynamic and hydromechanic forces and moments. The course of the
vessel differs from the heading of its centerline by the yaw (leeway) angle, .
Figure 1 shows the aerodynamic and hydromechanic force and moment com-
ponents in the deck plane, which is perpendicular to the center plane of the
vessel. Those involved in the VPP force and moment balance are: F
af
, the
aerodynamic forward force in the course direction; F
ah
, the aerodynamic heel
force, which is perpendicular to the forward force and parallel to the deck plane;
M
ah
, the aerodynamic heeling moment, whose vector is along the centerline
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 615
Figure 1 Forces and moments in the deck plane.
of the yacht; F
wr
, the resistance of the yacht in the direction opposite to the
course direction; F
wh
, the hydromechanical heel force component, which is
perpendicular to the course and parallel to the deck plane (F
wh
is exclusive of
components of that part of the buoyancy force that balances the weight of the
yacht); and M
wh
, the righting moment of the water on the yacht, whose vector
is in the direction of the yacht centerline and which includes both hydrostatic
and hydrodynamic components.
For any equilibrium sailing condition there are three balance equations in-
volving these forces and moments:
F
wr
(V
b
, , ) = F
af
(V
b
, , ),
M
wh
(V
b
, , ) = M
ah
(V
b
, , ),
F
wh
(V
b
, , ) = F
ah
(V
b
, , ),
(1)
where V
b
is the boat speed in the direction of the course, is the heel angle,
and is the leeway (yaw) angle.
For prescribed values of the wind speed and the sailing angle
T
, all six terms
in Equation 1 depend on the boat speed, the heel angle, and the leeway angle.
Figure 2 shows a block diagram of the VPP model, which solves the three
equations for these three unknowns.
Numerical solutionof the balance Equations 1is the most straightforwardpart
of a VPP. Conversely, modeling all the forces involved is an approximate and
imperfect science. Hydrodynamics, as we do it in terms of theory, experiment,
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
616 MILGRAM
Figure 2 A block diagram of the velocity prediction program (VPP). This emphasizes the fact
that solving the force balance equations is the minor and more ordinary part of the process. The
modeling of the forces is necessarily imperfect and requires most of the effort in developing a
faithful VPP.
and numerical computation, makes its greatest contribution to this eld by
predicting forces and teaching us how to model them.
In addition to the basic force models, two additional VPP features, which
involve feedback, are shown in Figure 2: the sail shape optimizer, which ad-
justs the sail shapes and their associated aerodynamic characteristics to maxi-
mize the boat speed for the prescribed wind conditions; and the direction opti-
mizer, which adjusts the angle
T
between the course and the wind to maximize
speed made good, VMG = |V
b
cos
T
|, when the desired course is upwind or
downwind.
2.2 Use of a Velocity Prediction Program
Figure 3 shows the effect on time that a 1% change in total resistance has on
an International Americas Cup Class (IACC) yacht sailing a course 17.2 km
upwind and 17.2 km downwind. A 1% change in resistance corresponds to a
change in race course sailing time of 2468 s, depending on wind speed. These
times relate to substantial margins of victory or defeat. When the tactical ad-
vantages of the faster boat are considered as well, the inuences of the speed
differences are even greater.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 617
Figure 3 Time differentials, in sailing a 34.3-kmcourse, that result froma 1%change in resistance.
The time differentials showninFigure 3correspondtoabout 0.3%differences
in the average speed. Even smaller speed differences can be meaningful for
racing vessels, so differences of very small magnitude need to be considered in
methods of evaluating candidate designs.
It is not possible to predict absolute boat speeds for a prescribed design to
within 0.3% or less of the actual sailing speed. However, this extreme accuracy
is not required on an absolute basis, only on a relative basis, and can be achieved,
to a greater or lesser degree depending on the design area under consideration,
if the technology is pushed to its limit for some experimental and numerical
methods.
3. DECOMPOSITION OF THE FORCE
COMPONENTS
Avessel under sail with non-zero heel and yawangles involves water ows that
are asymmetric port and starboard and that are inuenced by sea waves and
unsteady ship motions. It also involves a complicated air ow over the sails,
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
618 MILGRAM
hull, and waves. In the face of these complications, to effectively apply the
present state of the art in uid mechanics to analyze the ows and to support
the design process, it is necessary to make simplifying approximations.
3.1 Hydrodynamic Resistance
The essential goal in modeling hydrodynamic resistance is determination of
the function F
wr
(V
b
, , ), or alternatively F
wr
(V
b
, , F
wh
), for any prescribed
hull form in prescribed sea conditions. A useful approach, described in detail
by Milgram & Frimm (1993), is to use an additive resistance model of the
following form:
F
wr
= D
h f
+ D
r
+ D
af
+ D
hi
+ D
w
T
d
, (2)
where F
wr
is the total hydrodynamic resistance (drag); D
h f
is the frictional
drag of the hull; D
r
is upright residuary resistance of the entire vessel; D
af
is
the friction and interference drag of the appendages; D
hi
is the drag resulting
from heel and yaw (leeway) or, equivalently, from heel and heel force produc-
tion; D
w
is the resistance resulting from sea waves (added resistance); and T
d
is the mean dynamic thrust resulting from interactions of appendages with the
unsteady ow, which is due to vessel seakeeping motions and sea wave orbital
velocities. Nondimensional force coefcients, C, are obtained by dividing cor-
responding forces by
1
2

w
V
2
b
S
h
, where
w
is the density of the water and S
h
is
the wetted surface.
Figure 4 shows the fraction of resistance contributed by each component,
exclusive of T
d
, versus wind speed from VPP computations for an IACC yacht
sailing upwind using tank test data and measured sea spectra in San Diego,
California, as input. Figure 5shows the fractions for sailingdownwind. Tacking
angles for optimum speed-made-good are used both upwind and downwind.
Although the general features exhibited in Figures 4 and 5 are common to a
broad range of vessel types, precise values of resistance components depend
on vessel type and sailing conditions.
The hull friction is always the largest component for upwind sailing, and
in light winds it is largest for downwind sailing. For higher wind speeds in
downwind sailing, the residuary resistance becomes the largest because of the
high boat speeds in these conditions. The effects of resistance resulting from
heel, side force, and added resistance are negligible in downwind sailing and
are neglected in Figure 5. For upwind sailing, in the stronger winds, all other
resistance components are similar in magnitude, with the exception that the
appendage friction becomes less consequential as the wind speed increases.
Strictly speaking, the terms in Equation 2 are not independent of each other.
For example, the viscosity, which is mainly responsible for friction drag,
causes boundary layers whose displacement thickness inuences the residuary
wavemaking resistance. If each term in Equation 2 is estimated independently
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 619
Figure 4 Fractions of total resistance for each component for upwind sailing.
from the others, the interactions between terms are not accounted for. How-
ever, if scale model tests are conducted, the interactions are captured if the
model scale is not too small; they get mixed into the various terms in the de-
composition. For example, if the upright residuary resistance is dened as the
measured resistance minus the presumed frictional resistance, the sum of these
two resistances automatically includes the interactions. The work of Kirkman
& Pedrick (1974) suggests that scale model waterline lengths need to be 5 m or
more for a reasonable assurance of reliable results in the experimental process
and expansion of its data to full scale.
Once a set of resistance components that includes these interactions is deter-
mined by model testing, it is often useful to estimate numerically the difference
in a resistance component between two designs. Although this may not capture
differences in the interactions, it often provides a good approximation to the
differences that are the most signicant.
3.2 Aerodynamic Forces
The forward and heeling aerodynamic forces, F
af
and F
ah
, are given in terms of
the aerodynamic lift and drag forces, L
a
and D
a
, and the apparent wind angle,
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
620 MILGRAM
Figure 5 Fractions of total resistance for each component for downwind sailing.

a
, as
F
af
= L
a
sin
a
D
a
cos
a
and (3)
F
ah
= L
a
cos
a
+ D
a
sin
a
. (4)
Similarly, the aerodynamic heeling moment, M
ah
, is determined from these
forces and the heights of their centers.
On a moving vessel, the apparent wind speed and angle depend on height,
since the components due to true wind speed depend on height and the compo-
nent due to vessel speed is independent of height. The reference wind speed,
V
a
, and angle,
a
, are commonly taken as the speed and angle at a height of
10 m above the water.
The lift force, L
a
, can be accurately determined from lifting surface theory,
as described in Section 6 for upwind and close-reach sailing, where the local sail
cambers and incidence angles are modest. For offwind sailing, ow separation
is almost always large enough to materially inuence the lift, so experimental
data are required to construct a mathematical model for it.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 621
The aerodynamic drag force includes the induced drag of the sails as well
as the frictional and parasitic drag on the sails, mast, rigging, and hull. For
windward and close-reach sailing, the induced drag data can come from the
same computational implementation of lifting surface theory that provides the
lift. However, all the drag for offwind sailing and the friction and parasitic drag
for upwind sailing must come from experiments or empirical estimates.
Aerodynamic lift and drag coefcients, C
L
a
and C
D
a
, are
C
L
a

L
a
1
2

a
V
2
a
S
a
and C
D
a

D
a
1
2

a
V
2
a
S
a
, (5)
where
a
is the air density and S
a
is the actual sail area.
Each sailing condition has a different lift and drag coefcient for optimal
performance. The usual modeling approach is to determine a maximal-allowed
lift coefcient as a function of apparent wind angle, C
L
max
(
a
). For each appar-
ent wind angle and operating lift coefcient, which can be any positive value
less than or equal to C
L
max
(
a
), there is an associated drag coefcient. The VPP
chooses the amount of sail area to set, up to a maximal-allowed amount, for
optimal performance. To complete the specication, the drag coefcient needs
to be modeled as a function of C
L
and
a
.
The author has had success in modeling the drag coefcient as
C
D
(C
L
,
a
) = C
D
o
(
a
) +C
2
L
C
i
(
a
) +C
2
L
C
D
p
(
a
), (6)
where C
D
o
(
a
) includes the friction drag of the sails and the prole drag coef-
cient of the hull, mast, and rigging; C
i
(
a
) is a coefcient of induced drag;
and C
D
p
is a coefcient of lift-dependent prole drag.
The wind tunnel data of Campbell (1997) indicates that using an exponent
greater than 2 on the lift coefcient in the last term in Equation 6 improves
the description of the sail forces in his experiments. Euerle & Greeley (1993)
developed procedures for modeling sail forces for differing vertical distributions
of lift by altering C
i
in ways that can be well approximated theoretically.
Two approaches can be taken for estimating C
D
o
(
a
) and C
D
p
(
a
). One is
to add estimates of the drags from existing data. As examples, rigging drag
can be based on published drag coefcients for cylinders, and sail parasitic
drag can be based on section data (Milgram 1971). The other approach is to
measure the drag and subtract the induced drag computed from lifting surface
theory for the sail shapes in use to obtain the other drag components. This
has been done by Milgram (1993) and by Masuyama & Fukasawa (1997) with
instrumented sailing vessels. Each of these vessels has all parts of the sails and
rig afxed to a rigid frame that is connected to the hull through a six-component
computer-interfaced force and moment dynamometer as described by Herman
(1988). Further information on combining data from a sailing dynamometer
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
622 MILGRAM
with numerical results to develop an upwind VPP sail force model, and for
using the sailing dynamometer data alone for developing an offwind VPP sail
force model, is provided by Peters (1992) and by Milgram et al (1993).
4. TOWING TANK TESTING
A thorough review of towing tank testing of model scale sailing vessels is
provided collectively in the works of Larsson (1990), Van Oossanen (1993),
and Milgram (1993). Here, the process is outlined and some special problems
are described.
When data are obtained by model testing, the frictional terms, D
h f
and D
af
,
are subject to Reynolds scaling, whereas the other terms, D
r
, D
hi
, D
w
, and T
d
,
are subject to Froude scaling. The upright quantities, hull friction, appendage
friction, and residuary resistance are determined in the same way in ordinary
resistance tests of vessels that are not powered by sails. Appendage friction is
estimated on the basis of appendage geometry, and the hull friction coefcient
is taken as
C
h f
(Re) = (1 +k)C
f
(Re), (7)
where Re is the Reynolds number based on length, k is the formfactor evaluated
from the tank data by the method of Prohaska (1966), and C
f
(Re) is the at
plate frictional resistance. The difference between the measured resistance
and the estimated frictional resistance is taken as the residuary resistance, D
r
.
In addition to straight-ahead tests with the vessel upright, a sailing vessel
model needs to be tested with non-zero heel and yaw (leeway) angles with
both resistance and side force measured. This greatly increases the number
of tank runs required for a sailing vessel as compared to an engine-propelled
vessel. About 135 test combinations of speed, heel, and leeway are typically
required to fully quantify the hydrodynamic forces on a sailing vessel. This
number of test runs is based on determination and use of a single form factor,
k, obtained from a large number of low-speed upright (zero heel and zero yaw)
runs and a single value of wetted surface and length for expanding the model
tank data to full scale.
In principle, a different form factor applies to each different heel angle, and
different values of wetted surface and length apply to each combination of heel
and speed. The additional measurements and the additional test runs for deter-
mination of variable form factors, lengths, and wetted surfaces would greatly
increase the complexity and cost of a model test program. Mantzaris (1992) and
Milgram (1993) showed that for 31% scale tank models with waterline lengths
of about 6 m, the effect of a variable form factor in the analysis for the time for
an IACC yacht to sail a 47.3-km race course is about 4 s; the effect of including
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 623
the speed-dependent wetted surface on the sailing time is between 7 and 10 s,
depending on the wind speed. These effects increase as the model size is made
smaller because in the tank data analysis, errors in the frictional resistance ap-
pear in the residual resistance, which is expanded to full scale differently than
the frictional resistance.
A conclusion is that for model waterline lengths of at least 5 m and for scale
factors on the order of 1:3, it is not necessary to use heel-dependent form fac-
tors. The errors in predicted speed and sailing times caused by estimating the
frictional resistance from a constant length and wetted surface for the cases ex-
amined by Mantzaris and Milgramare similar to those fromforce measurement
errors on the order of 0.5%, which is the best currently achievable (Parsons &
Pallard 1997). Thus, we would need to be concerned if two very different
vessel types were being compared. However, for similar vessel types, these
errors would be similar for candidate designs and can be neglected in com-
parisons. The ability to obtain satisfactory full-scale performance-prediction
comparisons fromtank data with estimated frictional resistance based on single
values of length, wetted surface, and formfactor for each design when the model
is large enough is of major importance in making model-scale tank testing a
practical endeavor.
In conducting tank tests of vessels to be used for racing, accuracy and re-
peatability are of paramount importance. Section 2.2 describes the sensitivity
of racing performance to small changes in resistance. Since total accuracy
is impossible, a reasonable approach is to strive to limit measurement errors
or lack of repeatability to 1% or less, and to take special measures when one
has to choose between designs whose predicted performances differ by lesser
amounts. For example, the scale model of each design can be tank tested at
four separate times and the results then averaged together. This reduces the
erroneous data variability by a factor of 2.
One way to minimize some of the errors in tank data is to use speed, heel
angle, and heel force as independent variables instead of speed, heel angle, and
leeway angle. The primary inuence of leeway angle on hull drag is the drag
induced by the heel force associated with the leeway. Therefore, if there is a very
small error in leeway measurement or an unmeasured cross-tank current during
part of a data acquisition run in a tank test, erroneous effects are minimized
because the correct relationship between the induced drag and the heel force
still occurs. Details of this approach are provided by Milgram (1993).
5. INDIVIDUAL COMPONENTS OF RESISTANCE
This section focuses on the components of resistance individually, with em-
phasis on numerical methods for estimating some of them. The most reliable
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
624 MILGRAM
ways of obtaining several of the estimates are thorough and carefully done ex-
periments. Existing numerical methods cannot provide speed predictions that
are precise on an absolute basis, but they can provide differences in some of
the resistance components for differing designs and thereby aid in the design
process.
5.1 Hull Friction
Although hull friction is often the largest of the resistance components, it is the
one that is least amenable to numerical hydrodynamics. Recently, a number of
investigators applied computational Reynolds-averaged Navier Stokes (RANS)
methods to the viscous resistance of ships and boats (cf Larsson et al 1989,
Farmer et al 1995, Miyata 1996). None of these references provided a compar-
ison between computation and experiment for hull friction drag. Larsson et al
(1989) gave a comparison for pressure coefcient and wall friction velocity at
several locations on a large commercial ship hull form with an error between
theory and experiment for the wall friction of as much as 30% in the aft part
of the ship where the boundary layer is thick. In addition to uncertainty about
accuracy, one problem with using RANS codes to compare forces between dif-
fering designs is that they require an extreme amount of computer time. This
issue is discussed at some length by Farmer et al (1995), who describe ongoing
research to deal with it.
Currently, none of the computational methods for hull friction and its inter-
action with free surface effects are adequate to provide support to the design
process. We can expect improvements in RANS codes for free surface ows
on ships and boats as time goes on. A future alternative to RANS codes for
faster computation of viscous free surface ows may be outer inviscid solutions
strongly coupled with inner integral boundary layer equations, as is described
for appendages in Section 5.2. It would be helpful to have data from broad
systematic experimental studies of the relationship between sailing vessel hull
parameters and the friction drag coefcient, but these have not been done, at
least not to my knowledge.
Without either a robust numerical tool or an experimental database for fric-
tion drag and its interaction inuence on the other drag components, tow-
ing tank experiments with the friction drag estimated by Prohaskas method
is state of the art. This cannot estimate the friction drag precisely, since it
yields a coefcient that depends only on the Reynolds number, whereas there
are Froude number effects on the friction drag. Thus, the error in the esti-
mate gets mixed into the other resistance components, which are all subject to
Froude scaling, except for the appendage friction. Since the scale corrections
diminish with model size, large models give more reliable results than small
ones.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 625
5.2 Computation of Viscous Drag on Appendages
Appendages on a n-keeled sailing vessel include the keel n and the rudder,
and possibly a ballast bulb and winglets on the keel or rudder, or both (for
an example showing such appendages, see Figure 14 later in the article). In
principle, the viscous drag on appendages could be computed either with a
RANS method or with a method that strongly couples an outer inviscid ow
with inner boundary layer equations.
With the existing state of the art, the drag forces provided by RANS codes
have not had the accuracy needed for the design of racing vessels. Develop-
ment of outer inviscid solutions strongly coupled with inner integral boundary
layer equations is in the formative stage for three dimensional ows and shows
considerable promise for the future. On the other hand, the strongly coupled
method for two-dimensional (2D) ows is very advanced and shows excellent
agreement with experiments. It can be used in support of design, because the
rudder and keel n and optional winglets of a high-performance sailing vessel
are high-aspectratio lifting surfaces, so their friction drag can be estimated
from 2D section analysis.
A review of the coupled method in two and three dimensions is given here.
The ow away from the immediate vicinity of the lifting surface is largely in-
viscid, but viscous effects are important in the boundary layers. For many years
researchers tried to iterate between the inviscid and boundary layer solutions.
The idea was to compute an inviscid ow, use its pressure gradients in solving
the integral boundary layer equations, solve again for the inviscid ow with the
airfoil thickened by the displacement thickness of the boundary layer, including
the viscous wake, solve the boundary layer equations again, etc. This approach
failed when the boundary layer thickened rapidly, such as often occurs close to
the trailing edge.
The strongly coupled method, which solves for the outer and boundary layer
ows simultaneously, is much more robust. Advanced 2D developments have
been achieved by Drela & Giles (1987) and Drela (1989). Recently, Nishida
(1996) and Milewski (1997) developed the methodology and associated com-
puter codes for three-dimensional (3D) ows and applied them to some simple
cases. For purposes of explanation, with a focus on incompressible water ow,
the outer inviscid ow is described by its velocity potential, , continued to the
body surface in the boundary integral equation form used by Milewski (1997).
The perturbation velocity potential is , to which the potential of the free stream,

inf
, must be added to obtain the total velocity potential,
2
(2)
=
_
s
b
_
G

n

G
n
_
ds
_
s
w

w
G
n
ds +
_
(s
b
+s
w
)
Gds, (8)
where is 2 for 2Dowor 3 for 3Dow, Gis the Rankine source Green Function
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
626 MILGRAM
(logr for 2Dor 1/r for 3D, with r the distance between eld and source points),
n is the normal into the lifting surface, ds is the differential element of length or
area (2- or 3D), s
b
is the path or surface (2- or 3D) around the airfoil, s
w
is the
path or surface (2- or 3D) on the center of the wake,
w
is the jump in potential
across the wake fromtoptobottom(constant alongeachwake streamline), and
is a ctitious transpiration source strength distribution on the lifting surface and
wake that has to be determined so as to make the outer ow the same as the real
boundary layer would cause. The solution to Equation 8 with the last term re-
moved, andsubject tothe usual Neumannboundaryconditionandthe Kutta con-
dition, is called the inviscid potential,
inv
. The total velocity potential is then
=
inf
+
inv
+
_
(s
b
+s
w
)
G
v
ds, (9)
where G
v
is the sum of G and a body-shape-specic dipole distribution on the
surface, chosen such that the normal derivative of G
v
is zero except where the
source and eld points coincide.
TWO-DIMENSIONAL FLOW The surface velocity, which corresponds to the tan-
gential velocity at the outer edge of the boundary layer, is called U
e
and is
obtained as the derivative of the total potential with respect to the tangential
coordinate, s.
U
e
(s) =
(
inf
+
inv
)
s
+
_
(s

b
+s

w
)
dM
ds

G
v
(s, s

)
s

ds

, (10)
where the mass defect, M, the transpiration source strength, , and the boundary
layer displacement thickness,

=
_
(1u/U
e
)d ( is the coordinate normal
to the surface), are related by
=
dM
ds
and M = U
e

. (11)
The objective is to nd a solution for M(s) and the other boundary layer
parameters that satisfy the integral boundary layer equations
d
ds
+(2 + H)

U
e
dU
e
ds

C
f
2
= 0 and (12)

d H

ds
+ H

(1 H)

U
e
dU
e
ds
2C
D
+ H

C
f
2
= 0, (13)
where
_
(u/U
e
)(1 u/U
e
)d is the momentum thickness, H

/
is the shape parameter, C
f
2
wall
/[(U
e
)
2
] is the skin friction coefcient,


_
(u/U
e
){[1 (u/U
e
)
2
]d} is the kinetic energy thickness, H

/
is the kinetic energy shape parameter, C
D

_
(du/d)d/[(U
e
)
3
] is the
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 627
dissipation coefcient, and is the shear stress and u is the local velocity in the
boundary layer.
Drela & Giles (1987) give the semi-empirical equations for all the above
boundary layer parameters in terms of , M, and U
e
for laminar ow, which
is then entirely specied by these parameters and the simultaneous solution
of Equations 10, 12, and 13. The user must either specify the location of the
transition point from a laminar to a turbulent boundary layer or use a semi-
empirical relation to estimate where natural transition occurs. One common
method is based on an estimate of the ratio of the amplitude of the most unstable
Tollmien-Schlichting wave at the transition point to its value at the rst location
of growth with the ratio expressed as e
n
. The value of n at transition has been
correlated with T
f
, the ratio of root mean squarefree stream turbulence speed
to mean speed, by Mack (1977) as
n = (8.43 +2.4 log T
f
). (14)
Drela & Giles (1987) give a semi-empirical function, f
1
, for the rate of change
of n along the chord
d n
ds
f
1
(H, ) = 0. (15)
For turbulent boundary layers, they provide semi-empirical relations for all
of the boundary layer parameters in terms of , M, U
e
, and the coefcient
of maximum shear stress in the boundary layer, C

, for which they provide a


semi-empirical differential equation in the form of
d log C

ds
f
2
(c

, H, H

, ) = 0. (16)
Therefore, where the boundary layer is turbulent, the ow is entirely specied
by four parameters and the simultaneous solution of Equations 10, 12, 13, and
16. The solution is obtained by discretizing the airfoil surface and Equations
10, 12, 13, and 16 and solving them by Newtons method using the inviscid
solution for a starting point.
Figure 6 demonstrates the capabilities of the method. M Drela (private
communication) has provided experimental results and those from XFOIL, his
program that implements the method, for an LA203A section exhibiting some
complicated ow features whose experimental characteristics were measured
by Liebeck & Camacho (1985). Drela set the angle of attack in the compu-
tations to the one that achieved the same lift coefcient as was achieved in
the measurements. A laminar separation bubble occurs at x/c 0.62, with
turbulent reattachment thereafter, as can be seen in both the suction side pres-
sure distribution and in the upper boundary layer streamline. Near the trailing
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
628 MILGRAM
Figure 6 Calculated and measured pressure coefcients on an airfoil section. The upper part of
the gure shows the pressure coefcients with the inviscid calculation (dashed line) and the coupled
boundary layer calculation (solid line). The lower part of the gure shows the airfoil section and
the streamlines at the outer edge of the boundary layer.
edge, there is very substantial boundary layer thickening on the suction side
and reduced thickness on the pressure side. This is shown by deviations from
the inviscid pressures in both the experimental and computed results.
The remarkably good comparison between theory and experiment is due to
several factors. Equations 10, 12, and 13 are fundamentally correct. However,
they are not all that matters. The solution is crucially dependent on the various
relations between boundary layer parameters (cf. Drela & Giles 1987) and
these are semi-empirical. The reason for their accuracy is that vast amounts of
experimental data, gathered over many years, have been used in generating the
semi-empirical relations.
We do have XFOIL computations and experimental data for the forces on
foil sections used for sailing vessel appendages. Figure 7 shows drag coef-
cient versus lift coefcient for an airfoil section as calculated by this method
using the program XFOIL (Drela 1989) and as measured in a water tunnel (see
Kerwin 1994). For the calculation, n was set to 3.5, which corresponds to the
tunnel turbulence level of 0.7%. Experimental forces were based on velocities
around a large rectangle surrounding the section, measured with a laser doppler
velocimeter and applying the Bernoulli equation outside the wake and momen-
tum conservation principles to the ow (Kinnas 1991). Agreement between
numerical prediction and experiment is excellent.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 629
Figure 7 Calculated (line) and measured (circles) section drag versus lift coefcients.
Figure 8 (left) shows the airfoil section characteristics calculated by this
method for two section shapes that could be used for the keel n of a sailing
vessel. One section has a thickness fraction of 0.13, and the other has a
thickness fraction of 0.17. The calculation is done for n = 9, which cor-
responds to an incident stream with negligible turbulence. The gure indicates
that for lift coefcients in excess of 0.33, the thicker section has less fric-
tion drag as a result of more laminar ow. Design experience runs counter
to this result. A boat goes faster with a keel thickness fraction of 0.13 than
with a thickness fraction of 0.17, even though typical keel lift coefcients
exceed 0.33 for windward sailing. The reason for the disparity between the
computed results and sailing experience is most probably that the combina-
tion of free stream turbulence, particulate matter, and small gas bubbles in
ocean water near the surface cause transition near the leading edge on the
suction sideno matter what the section shape (cf Lauchle & Gurney 1984).
Figure 8 (right) shows the airfoil characteristics with suction side transition
set at a maximum downstream location of 5% of the chord. For this forced
transition condition, the thinner section has less drag up to a lift coefcient
of 0.88. This is consistent with design experience, as operating lift coef-
cients are designed somewhat lower than 0.88. The combined effects of free
stream turbulence and particulate matter in the ow is an area in need of further
study.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
630 MILGRAM
Figure 8 Characteristics for two airfoil sections: (left) a clean, low-turbulence inow; (right)
forced transition at 5% of the chord.
THREE DIMENSIONAL FLOW Whereas Nishida (1996) coupled the boundary
layer equations with outer-ow full-potential compressible ow equations for
3D, Milewski (1997) coupled the boundary layer equations to inviscid incom-
pressible ow panel methods, as described above, to 3D problems.
The 3D form of Equation 10 is

U
e
=

i u
e
+

kw
e
=

U
inv
+
_
(a

b
+a

w
)

a
G
v
( a,

a

)da

, (17)
where x and z are local coordinates in planes parallel to the surface in the chord-
wise and cross-stream directions, respectively, a =

i x +

kz, and a signies
surface area on the 3D airfoil or its wake. a
b
refers to the object surface and a
w
refers to the wake surface. The coordinate y is normal to the surface and takes
the role of used in the 2D development.
The literature for the 3D method gives the integral boundary layer equations
in a slightly different form than the literature for the 2D method: The 3D
literature uses shear stress and dissipation rather than their coefcients. The
three integral boundary layer equations for the 3Dcase that followfor chordwise
momentum (18), cross-stream momentum (19), and energy (20) in the form
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 631
given by Mughal (1992) are used by both Milewski and Nishida.

x
_
q
2
e

xx
_
+

z
_
q
2
e

xz
_
+q
e

x
u
e
x
+q
e

z
u
e
z
=

x
w

, (18)

x
_
q
2
e

zx
_
+

z
_
q
2
e

zz
_
+q
e

x
w
e
x
+q
e

z
w
e
z
=

z
w

, (19)
and

x
_
q
3
e

x
_
+

z
_
q
3
e

z
_
=
2D

, (20)
where q
e
= |

U
e
|,
x
w
and
z
w
are the two components of shear stress at the sur-
face, and Dis the energy dissipation per unit area. For the 3Dcase, there are four
momentum thicknesses,
xx
,
xz
,
zx
, and
zz
; two displacement thicknesses,

x
and

z
; and two kinetic energy thicknesses,

x
and

z
. Integral expres-
sions for these thicknesses in terms of the velocity components through the
boundary layer, and for the dissipation, D, in terms of shear stresses and strain
rates in the layer, are given by Mughal (1992), Nishida (1996), and Milewski
(1997).
In developing the numerical method, Nishida and Milewski both rotate the
(x, z, y) local coordinate system to an (x
1
, x
2
, y) system with x
1
parallel to
the ow at the outer edge of the boundary layer and x
2
perpendicular to this
streamline. The integral boundary layer equations18, 19, and 20are easier
to express and intuitively understand in the (x, z, y) system, but the discretized
solution is carried out more naturally in the (x
1
, x
2
, y) system. Both Nishida
(1996) and Milewski (1997) provide expressions for the displacement, mo-
mentum, and energy thickness in this system. The mass defects, displacement
thicknesses, and transpiration source strength, which is required in the outer
ow Equation 17, are related by
M
1
= q
e

1
, M
2
= q
e

2
, and =
M
1
x
1
+
M
2
x
2
. (21)
This is part of the strong coupling between the integral boundary layer equations
and the outer ow equation.
The method provides the tangential friction drag and the normal pressure
drag individually. Figure 9 shows these drag coefcients (based on surface
area) versus Reynolds number, as calculated by Milewski (1997), for a form
that might be used as the ballast bulb on a sailing vessel, with forced tran-
sition at 5% of the bulb length. The form is an NACA 0020 section rotated
to form a body of revolution. Figure 10 shows near-surface streamline di-
rections on a rectangular wing of aspect ratio 5 at an angle of attack of 9.5
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
632 MILGRAM
Figure 9 Drag coefcient, based on bulb area, for an axisymmetric ballast bulb.
degrees and a Reynolds number of 2 10
6
. The spanwise ow at the tip,
outward on the pressure side and inward on the suction side, is captured in the
computation.
Although the 3D method has been used only on bulb and wing forms thus
far, the thoughts for the future involve hull-keel-bulb and fuselage-wing-tail
combinations. If this can be done, the next step would be to include the effects
of the free surface.
5.3 Residuary Resistance
The residuary resistance determined in a towing tank experiment is the sum of
the wavemaking resistance, interactions between resistance components that are
not explicitly modeled, and errors in the presumed frictional resistance. Because
of the Reynolds number scalingof the frictional resistance, the inuence of these
errors on predicted full-scale resistance is minimized by using large models for
acquiring towing tank data. For example, if the full-size vessel is 20 m long and
the frictional resistance is in error by 4% because of an error in estimated form
factor, the full-scale predicted resistance is in error by about 0.5%if a one-third
scale model is used, but it is about 1.5% if a one-tenthscale model is used.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 633
Figure 10 Flow directions of wall streamlines on a rectangular wing from a three-dimensional
coupled integral boundary layer equation and outer potential ow computation. Angle of attack is
9.45 degrees, and Reynolds number 2.1 10
5
.
NUMERICAL METHODS A possible use of numerical hydrodynamics for min-
imizing residuary resistance is computation of the wavemaking resistance for
differing hull geometries. The state of the art of these computations has im-
proved greatly in recent years, but comparisons between experiments and the
computations demonstrate uncertainties that lead one to turn to experiments
whenever possible. The existing computational difculties appear to lie both in
the wave resistance computation itself and in the inability to accurately predict
resistance component interactions.
Computer-based solutions to the RANS equations with free surface bound-
ary conditions can, in principle, capture wavemaking and viscous ow inter-
actions. The application of this method to hulls with sailing vessel forms,
whose overhangs inuence the ow, is just beginning. Farmer et al (1995)
provide RANS code results for ship wave elevations and forces, and Miyata
(1996) shows RANS-computed wave elevations and hull pressures. How-
ever, comparisons between computed and measured forces, which are the
crucial quantities for design comparisons, are not generally available. Force
comparisons between experiments and the more mature area of numerical
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
634 MILGRAM
determination of wave resistance for inviscid ow do exist and are reviewed
here.
The rst numerical procedure giving results that were generally accurate
enough to be considered for use in detailed design of real ships and boats is that
of Dawson (1977). Whereas all the prior methods linearized the mathematical
problem about the at free surface with a uniform stream, Dawson linearized
the ship wave problem about the double-body ow, which corresponds to the
submerged portion of the ship beneath a rigid free surface. This basis ow
contains many of the inuences of the ow around the displacement form of
the ship. Only the surface waves are left out, and these are approximated as a
linear perturbation on the double-body ow.
Extensions and applications of Dawsons method to sailing vessel hulls
include those of Larsson (1987), Rosen et al (1993), Nakos et al (1993),
Sclavounos (1995), Raven (1994), and Rosen & Laiosa (1997).
Comparisons of the wave drag from Dawson-type codes with experimental
residuary resistance showdifferences on the order of 30%over the sailing speed
range. If this error were the same for differing designs, the results could still
be used on a comparative basis for design. However, the analysis by Milgram
(1996) of the computed and experimental results for two IACC yacht designs
given by Rosen et al (1993) shows the error differs by up to 20% for the two
designs at some speeds.
Larsson (1987) recognized that much of the difculty associated with pre-
diction of the wave resistance of sailing vessel hulls by Dawson-type codes
involved lack of consideration of the portion of the hull above the still wa-
terline. He demonstrated this by showing that the error was much larger for
a 12-mclass yacht hull form with long and at overhangs than it was for a
5.5-mclass yacht with steeper and shorter overhangs.
Two approaches have been used to improve wave resistance calculations on
vessels with overhangs: the articial lengthening of the vessel, developed by
Sclavounos (1995); and a nonlinear method that treats the actual wetted portion
of the vessel while it is moving forward, developed independently by Raven
(1994) and by Rosen & Laiosa (1997). The approach of Sclavounos solves the
initial value problem in which the vessel is brought to speed from rest, and
the time domain computation is continued until the wave resistance becomes
nearly steady. In this process, the surface elevation is computed at each time
step. Because of the shallow slope overhang proles of the vessels under con-
sideration, the wetted length changes throughout the computation. An artice
that is used is to stretch the vessel longitudinally during the computation so
the still-water length matches the computed wetted length, followed by recom-
puting the double-body basis ow. Another artice used by Sclavounos is to
alter the boundary condition on some of the near-centerline free surface pan-
els that border on the separation line of the stern underbody prole. Instead
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 635
of using the usual kinematic free surface boundary condition on these pan-
els, the condition of tangent separation was used inasmuch as this is what is
observed on real vessels. In spite of these special features in the numerical
methods, they show considerable overprediction of the wave resistance at high
speed.
Raven (1994) has taken the procedure one step further by calculating the
wave resistance for the nonlinear problem through a set of iterations where a
linear boundary integral equation (panel) method is used at each iteration, but
it is linearized about the solution for the previous iteration. This is continued
to convergence, and the actual wetted shape ends up being used. Figure 11
shows the percentage difference between computation and experiment for an
offshore racer-cruiser yacht with a still waterline length of 21.5 m based on
the data given by Raven. Both linear (Dawson-based) and nonlinear results
are given. If one discounts the low Froude number result, for which the much
smaller wave resistance is difcult to separate from the frictional resistance
in the experimental data, the nonlinear theory is better. This is consistent
with a comparison made by Larsson et al (1989) between linear and nonlinear
computations for large powered ships.
Figure 11 Percentage error inthe wavemakingresistance of anIACCracer-cruiser yacht computed
by Raven (1994).
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
636 MILGRAM
Rosen & Laiosa (1997) developed a nonlinear method equivalent to the
method of Raven (1994) in principle, although details of how the boundary
value problems are solved are different. They show experimental residuary
resistances and both linear (Dawson-type) and nonlinear (actual wetted hull)
computations of the wave resistance for two IACC yacht designs, which differ
considerably in their bow overhangs. One is designated as a destroyer bow and
the other as a spoon bow. The experiments show less residuary resistance for
the spoon bow, and the linear code shows less wavemaking resistance for the
destroyer bow. However, the nonlinear code gives the opposite result and ranks
the boats for design purposes in the same order as do the experiments. Figure 12
Figure 12 Percentage error in the wavemaking resistance computed by Rosen & Laiosa (1997):
(a) the linear computation results; (b) the nonlinear computation.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 637
Figure 13 Total resistance drag areas for two IACC yachts predicted from towing tank model test
data. (left) Upright; (right) 20 degrees heel.
shows the percentage error in the numerical predictions of wave resistance in
the upright condition for these two hulls, based on the data given in the gures
of Rosen & Laiosa (1997). The nonlinear code brings the errors for each of
the designs signicantly closer together. The percentage error, E
%
, is dened
as
E
%
100
computed wave resistance measured wave resistance
measured wave resistance
. (22)
THE NEED FOR CONSIDERING HEEL ANGLE IN RESISTANCE PREDICTION Figure
13 (left) shows the upright resistance drag areas [resistance/(0.5
w
V
2
b
)] deter-
mined by expanding towing tank test data to full scale for two IACC yachts,
USA 9 and USA 23. For boat speeds slower than 11.5 knots, USA 9 is shown
to have slightly less resistance. Yet, in actual use, USA 23 sailed a typical day
race course about 2 min faster than USA 9 for these wind speeds. The speed
advantage of USA 23 was even more at higher wind speeds, in which the boats
sailed faster.
The differences in resistance when the vessels are heeled is a principal reason
for the relative performances of the two vessels. Figure 13 (right) shows the
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
638 MILGRAM
drag areas vs speed, as determined fromthe tank tests, for the same two designs
at a heel angle of 20 degrees and zero side force. Differences in resistance up
to 4% are seen and the design with superior performance is the one with lower
resistance. Of course, in actual sailing there would be non-zero side forces
when the heel angle is 20 degrees, but the zero side force cases are shown here
to show the effects of heel most clearly. It is not at all uncommon for sailing
vessels of differing performance potential to have similar upright resistances,
but signicantly different resistances when heeled.
5.4 Induced Drag of the Hull and Its Appendages
The form of resistance decomposition shown in Equation 2 is chosen, in part,
for convenience. The rst three terms provide the calm water resistance in the
upright condition, the fourth is the change due to the vessel operating at non-
zero heel and leeway (yaw) angles, and the fth adds the resistance increase due
to the presence of sea waves. The focus of this section is the use of theoretical
and numerical methods for estimating the differences in induced drag between
candidate designs. It is a portion of the fourth term. Induced drag occurs
because the circulation around all side-forceproducing portions of the hull
and appendages induces changes in the ow direction everywhere and the local
lift is perpendicular to the induced ow direction. The induced ow is in exact
analogy to the downwash of an airplane, which is responsible for its induced
drag.
In theory, the induced drag of a lift (heel force) producing object is very
nearly proportional to the square of the lift for ows in the absence of a free
surface. However, towing tank tests show that the actual induced drag versus
lift-squared function of a sailing vessel differs from the theoretical one based
on no free surface, both in its mean slope and in its linearity. Examples can
be found in Greeley & Cross-Whiter (1989). The disparity is certainly related
to free surface effects, although it is likely that some nonlinearity is due to the
leeway-dependent location of the aft stagnation zone on the heeled hull.
Completely numerical determination of induced drag for use in speed pre-
diction at the required level of accuracy for racing vessels is beyond the state
of the art. It awaits development of robust numerical methods that accurately
solve for the entire heeled and yawed vessel, including both the boundary layers
and the largely inviscid outer ow. However, there is good reason to believe
that presently available numerical hydrodynamics can evaluate the difference
in induced drag between identical canoe-body hulls with different appendages.
Although Larsson (1987) has developed a numerical method for calculating the
combined wave resistance and induced drag for a vessel moving with leeway
but no heel, the more common approach is to solve for the inviscid ow about
the heeled and yawed wetted portions beneath a rigid free surface. In practice,
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 639
this is done by solving for the double-body owin an innite uid. Two reasons
justify this approach.
1. The free surface effects onthe side force andinduceddragare strongest onthe
canoe body. With the same canoe body used in comparisons, the difference
in these effects derived from appendage variations should be small.
2. Most of the hull-induced owvariations on the appendages are due to the hull
displacement effects, which are captured in the rigid free surface problem.
The ship wave ow effects on the appendages, which are not included in
the rigid free surface problem, are very nearly the same when only the ap-
pendages are varied so that they have little effect on the differences between
appendage forces.
Two effective methods are in general use for calculating the heel force (lift)
and induced drag for the double body problem: the boundary integral equation
method (BIEM), often called a panel method, and the vortex lattice method.
The BIEM (cf. Katz & Plotkin 1991, Morino & Gennaretti 1992) is based on
the usual application of Greens theorem, leading to the integral equation
2(x) =
_ _
S
b
_
(x

)
G(x, x

)
n

+ G(x, x

V

n

_
dS

+
_ _
S
w
(x

)
G(x, x

)
n

dS

, (23)
where is the disturbance velocity potential caused by the body; S is the surface
of the body and its vortex (equivalently dipole) wake;

V is an onset owvelocity
vector; x and x

represent 3D locations of eld and source points, respectively,


and all points lie on the body or on the wake; G is the Rankine source Green
function, 1/|x x

|; subscript S
b
refers to an integral over the body with normal
into the body; and subscript S
w
refers to an integral over the wake on which the
potential jump from side 1 to side 2 is , and the normal to the wake is in the
direction from side 2 to side 1.
Discretization of the integral Equation 23 leads to the set of linear algebraic
equations that is the panel method. In formulating it, the jump in velocity
potential across any vortex/dipole wake streamline has to be set equal to the
jump in velocity potential at the position on the trailing edge of the body from
which the streamline originates (cf Lee 1987). Initially, the position of the
wake sheet is unknown and needs to be determined interactively with streamline
tracing for maximum accuracy.
The vortex lattice method, as originally developed by Greeley & Cross-
Whiter (1989), is a simplication of the panel method. It requires simpler
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
640 MILGRAM
program input data and provides much faster computation. Its origins stem
from the lifting body panel method rst developed by Hess (1972). That ap-
proach was to use both surface source and surface dipole panels, as indicated in
Equation 23, on lifting surfaces such as aircraft wings and tails, but to use only
source panels on fuselages, which are boat canoe bodies and keel bulbs here.
To properly model the circulation at the lifting surface roots, it was carried into
the fuselage to the centerline.
The Greeley & Cross-Whiter method represents the lifting surfaces as their
centerplane distributions of sources and vortex lattices (or, equivalently, dipole
panels) carried into connecting bodies and hull canoe bodies and keel bulbs
as surface distributions of sources. Both Greeley & Cross-Whiter (1989) and
Ramsey (1996) have done extensive comparisons between the results of panel
methods and vortex lattice methods, with the nding that the two methods yield
very nearly the same induced drag for a prescribed lift.
An example of the use of the vortex lattice code of Greeley & Cross-Whiter
by J Kleene (private communication) is given. Figure 14 is a drawing of the
submerged portion of an IACC yacht at a heel angle of 20 degrees. It is
customary to adjust the fore and aft location of the rig (sails) such that the rudder
carries about 20% of the heel force exerted by the sails. This corresponds to
rudder angles of about 1degree inlight winds and3degrees instrongwinds. The
particular vessel geometry under consideration is more efcient at generating
heel (side) force on the keel n than on the rudder because the keel is not
only deeper, it also has winglets. Use of a vortex lattice code has shown that
minimum induced drag occurs when the rudder side force is very nearly zero.
For typical sailing conditions, the numerical computation predicts the induced
drag to be reduced by 8% when the rudder side force is reduced from 20% of
the total to zero and the keel side force increased accordingly. Figure 15 shows
the time gained versus wind speed of an IACC yacht, when sailing a 34.4-km
windward-to-leeward course, resulting from an 8% reduction in induced drag
as determined by a VPP. All of these gains, except for a fraction of a second,
occur on the 17.2-km windward portion of the course.
Figure 14 Submerged portion of an IACC racing yacht. It is shown at a heel angle of 20 degrees.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 641
Figure 15 Time gained when sailing a 34.4-km windward-to-leeward course when 20% of the
side force is moved from the rudder to the keel and hull.
5.5 Added Resistance Due to Sea Waves
Prediction of the added resistance of a sailing vessel due to sea waves at
a level of accuracy comparable to important differences between vessels in
competition is still beyond the state of the art. Apart from the condition of
surng downwind, which can be viewed as temporarily negative-added resis-
tance, the added resistance is largest and of primary importance for sailing
to windward, which is the case considered here. Fully modeled experiments
would be difcult and costly because in addition to the requirement of model-
ing the mass and its rst and second moments, the tests should be conducted
with the waves coming from the oblique wind direction. It would require
a large wave basin capable of generating waves that are oblique to the tow-
ing direction, with a model support and force measurement system that al-
lowed unsteady motions and that obtained mean forces with errors limited to
about 1%. Such a facility probably does not exist, which explains why trust-
worthy data, such as that provided by Gerritsma & Moeyes (1973), are for
head seas only. Even if the experiments could be done, an accurate simu-
lation would require a mechanism to provide the pitch damping associated
with unsteady sail aerodynamics, which Skinner (1982) showed is signicant.
Because of these experimental difculties, added resistance is generally esti-
mated from numerical methods, even though they ignore some of the important
effects.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
642 MILGRAM
The numerical methods in use are of second order in wave amplitude, which
is the lowest contributing order. Because of this, the numerical results must
be treated with caution when determining the inuence of added resistance
resulting from large waves, as has been concluded by Caponnetto (1993). The
results do improve speed predictions, however, and are therefore commonly
used in VPP.
The added resistance operator, A(, ), is the ratio of the added resistance
to the square of the amplitude of a wave of circular frequency and angle ,
which is 0 for stern seas and 180 degrees for head seas. In a directional sea
wave spectrum S(, ), the added resistance to lowest (second) order is
D
w
=
_

_

0
2A(, )S(, ) dd. (24)
The role of numerical hydrodynamics is determination of A(, ). Until
recently, the procedure of Gerritsma & Beukelman (1972) gave the best agree-
ment between computation and experiment. This procedure uses strip theory
to estimate the added resistance from the rate of energy imparted to radiated
waves caused by relative vertical motions between the ship and the water. Part
of the contribution to added resistance from sea wave diffraction by the ship is
neglected in the procedure.
Recently, Sclavounos & Nakos (1993) and Sclavounos (1995) calculated
A(, ) by panel methods. As in their wave resistance computations men-
tioned above, the linearized free surface problem about the double body ba-
sis ow is determined. This is a more accurate approach than earlier proce-
dures and provides a more representative treatment of the effects of diffracted
waves. Sclavounos reported (private communication) that theoretical results
were brought into improved agreement with model scale measurements in head
seas when the vessel length was stretched to the wetted length associated with
the steady ow.
Rosen & Laiosa (1997) formulated the unsteady ship motion and added-
resistance problem in the form of a frequency domain panel method linearized
about their nonlinear calm water forward speed solution. They show excellent
agreement between computed and experimental values of the added resistance
operator in head seas, A(, ), for one sailing vessel hull, whereas for a dif-
ferent, but similar, hull the computations overpredict experimental values by as
much as 100%. Likewise, both experimental and computed heave-frequency
responses are provided that are obviously wrong for one vessel in long waves
and obviously correct for two other vessels. In spite of these shortcomings, the
approach of linearizing the unsteady problem about the nonlinear calm water
result has merit and is likely to be the next step in development of numerical
methods for ship motions.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 643
Figure 16 Added resistances of two IACC designs in fully developed seas.
Rosen & Laiosa (1997) applied their method to both upright and heeled
vessels. They show that the computed added resistance of an IACC yacht is not
changed signicantly from the upright condition when the vessel is heeled 20
degrees. However they report (private communication) that this is not the case
with all vessels. This is consistent with the results of Gerritsma & Keuning
(1988), who found that the variation of added resistance with heel angle depends
on the vessel type.
Figure 16 shows the added resistance based on added resistance operators
computed with the program of Sclavounos & Nakos (1993) and Equation 24
for two IACC yachts. One is a base boat; the other has the waterline beams
increased by 10%and the canoe-body depth decreased by 10%so as to maintain
displacement. This pair of shapes is chosen for demonstration here because for
xed displacement and length, beam is the dominant parameter for variations
in numerically determined added resistance. In the gure, the added resistance
is shown as a fraction of the total resistance of the base boat sailing to windward
as determined through a VPP. For this example, the long crested oblique seas
are propagating in the wind direction and specied by the Pierson-Moskowitz
spectrum, which is
S
PM
() = 0.081
g
2

5
exp
_
0.74
_
g
U
w

_
4
_
, (25)
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
644 MILGRAM
where g is the acceleration due to gravity and U
w
is the wind speed at a height of
19.5 m. As indicated in Figure 16, the added resistance in typical sea conditions
is of a magnitude such that it must be included in speed prediction procedures if
theyare tobe accurate. Inclusionof the addedresistance computedbythe second
order theory and applied to the wave spectrum of interest brings predicted and
measured sailing speeds into better agreement, but it does not distinguish be-
tweenthe performance of differingdesigns inwaves tothe extent that is observed
in practice. It is believed that one reason for this lies in the effects of particularly
large waves, which are not entirely captured by the second order theory.
5.6 Mean Forward Thrust Resulting
from Unsteady Motions
When an appendage is operating in the unsteady owassociated with sea waves
and unsteady vessel motions, it has the possibility of generating a mean forward
thrust resulting fromthe unsteady relative motion of the uid. This thrust is very
small for vertical appendages such as n keels and rudders. However, it can be
signicant for horizontal appendages such as keel or rudder winglets, which is
the subject considered here. There is very little published information about the
mean thrust caused by unsteady motions of sailing vessels. The mathematical
analysis that follows here was rst derived for the 1992 Americas Cup defense,
and it appears in similar form in Milgram (1996). A similar result was derived
independently by Sclavounos & Huang (1997), who did a thrust analysis and
coupled it with a panel method for the unsteady motion to obtain predictions
for the effect of rudder winglets on unsteady motions and added resistance in
addition to the unsteady thrust of the winglets.
As is sketched in Figure 17, the lift on a hydrofoil is perpendicular to the
incident ow and the drag is parallel to it. Foils with symmetrical thickness
Figure 17 Forces on a wing section at an angle of attack.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 645
forms are used to simplify the presentation here, although the effect occurs
on nonsymmetrical foils as well. Lift occurs when the inow has an angle of
attack,
w
+
v
, with respect to the nose-tail line of the foil, and a component of
this lift force is forward when
w
= 0.
v
is the pitch angle. The inow angle
of attack on horizontal appendages occurs when there is a relative vertical
velocity resulting from vessel motions or vertical components of the orbital
velocities of the sea waves. As shown in the gure, there is a net forward force
if Lift sin(|
w
|) > Drag cos(
w
). This net forward force occurs if
w
is either positive or negative. As the vessel undergoes unsteady motions in
waves, the inow angle to a horizontal appendage oscillates between upward
and downward. The lift vector direction can be mainly up or down, but it almost
always has a forward component. If the average of this forward component
exceeds the average of the aft component of the drag, positive average thrust
occurs. Otherwise, there is net drag.
The dynamic thrust is estimated here with lifting line theory, largely because
horizontal appendages that are efcient at its generation necessarily have large
aspect ratios. For relative motions of frequency f
e
with boat speed V
b
and
appendage dimensions on the order of , the thrust-producing effects have
reduced frequencies, f
e
/V
b
, which are low enough for quasi-steady lifting
line or lifting surface theory to be accurate.
Yawmotions of the vessel are neglected, and the dynamic thrust on generally
vertical appendages, such as a keel n or a rudder, are not considered. Just the
dynamic thrust on horizontal appendages such as a rudder wing or keel winglets,
which have the greatest potential for thrust production, is analyzed. The effects
of spanwise ow are neglected.
The relative angle of incidence on the wing is
w
+
v
cos , where is the
(presumed steady) heel angle. The lift coefcient, C
L
, is
C
L
= 2(
w
+
v
cos ). (26)
Ignoring a small increase due to thickness of the wing, the lift slope factor,
(cf. Glauert 1959, Thwaites 1960), is
=
R
A
cos
2 + R
A
, (27)
where R
A
is the aspect ratio and is the sweep angle of the wing.
Using the approximation of an elliptical spanwise circulation distribution,
the induced drag coefcient, C
D
i
, is
C
D
i
=
1
R
A
C
2
L
. (28)
Thus, the forward thrust, T, to lowest order in
w
and temporarily neglecting
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
646 MILGRAM
viscous effects, is
T =
1
2
V
2
b
A
p
[2
w
(
w
+
v
cos )
2
(
w
+
v
cos )
2
], (29)
where A
p
is the planform area of the wing and 4/R
A
.
The angle of incidence, to lowest order, is

w
=
w
V
b
cos , (30)
where wis the relative vertical velocity at the fore and aft location of the wing. It
is the difference between the vertical component of the sea wave orbital velocity
and the heave velocity of the boat at the wing location. Since the boat is both
heaving and pitching, w depends not only on time but also on the fore and aft
location of the wing. Denoting statistical (or time) averages by overbars, taking
the average of Equation 29 and using Equation 30 provides the fundamental
equation for the dynamic thrust, T
d
:
T
d
=

2
V
2
b
A
p
cos
2

_
2
V
2
b
w
2
+
2 2
V
b
w
v

2
v
_
. (31)
The ship motions are approximated as being linearly dependent on the sea
state, with complex transfer functions (frequency responses) from sea wave
elevation to heave of H
w
(
e
) and to pitch of H

(
e
). In practice, these
transfer functions are determined numerically from strip theory (cf Salvesen
et al 1970) or from a panel method (cf Nakos & Sclavounos 1990). Then, for
a one-sided sea wave encounter frequency spectrum S(
e
),
w
2
=
_

0
|H
w
(
e
)|
2
S(
e
)d
e
, (32)

2
v
=
_

0
|H

(
e
)|
2
S(
e
)d
e
, (33)
and
w
v
=
_

0
Re
_
H

w
(
e
)H

(
e
)

S(
e
)d
e
. (34)
Equation 31 is the fundamental equation for dynamic thrust and can be used
when the viscous and parasitic drag of the thrust-producing wing are accounted
for in the appendage friction drag. However, for force comparisons of various
designs of wings whose principal purpose is to generate dynamic thrust, or to
decide whether to use such a wing at all, it is useful to consider the difference
of the dynamic thrust and the friction and parasitic drags of the wing.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 647
Here, an example is provided for a rudder wing located at a depth of 3 m and
8 m aft of midship on an IACC yacht. The wing area is 1 m
2
. It has an aspect
ratio of 9, a taper ratio (tip chord/root chord) of 1/2, and a sweep angle of 25
degrees. The wing has an NACA 0012 section, and its friction drag coefcient
is approximated from XFOIL (Drela 1989) results at a Reynolds number of
1.5 million as 0.008 + 0.164
2
, where the relative angle of incidence, , has
units of radians. The angle of incidence used in the friction drag calculation is
the root mean square relative incidence angle, which is determined in terms of
the quantities used in this section as

2
=
_
1
V
2
b
w
2
+
2
V
b
w
v
+
2
v
_
cos
2
. (35)
The interference drag area, A
I
, between the wing and the rudder is taken as that
given by Hoerner (1965) for strut junctions,
A
I
= t
2
_
17
_
t
c
_
2
0.05
_
, (36)
where t is the thickness and c is the chord length at the root intersection.
Figure 18 shows both the dynamic thrust and the difference between the
dynamic thrust and the sum of the friction and interference drags for upwind
sailing in fully developed seas with spectra given by Equation 25. The angle
Figure 18 Dynamic thrust from a rudder wing for windward sailing.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
648 MILGRAM
between the seas and the course of the boat is taken as 142 degrees (38 degrees
fromahead). The pitch and heave frequency responses used to generate the data
in the gure were calculated by the SWAN program of Nakos & Sclavounos
(1990).
Numerical experiments show that dynamic thrust is greatly diminished by
reducing the wing aspect ratio. So, it is important that the aspect ratio be as
large as possible for a thrust-producing wing.
Actual sailing experience shows less windward advantage than Figure 18
indicates. Possible reasons for this are erroneous predictions of pitch and heave
and possible larger-than-estimated drag, due to temporary ow separation in
the unsteady uid motion.
6. COMPUTATION OF LIFT AND INDUCED
DRAG OF SAILS
Numerical methods for calculating the lift and induced drag of sails are pre-
dominantly inviscid vortex lattice methods. The rst was done 30 years ago
(Milgram1968a,b). The sails were represented by vortex lattices and the wakes
were at. The lower boundary beneath the sails was approximated as a at plane
at a user-specied location below the sails. The procedure was aimed at the
inverse design problem for which pressure distributions were specied and the
shapes needed to attain them were computed. A number of proprietary de-
velopments using a similar approach, and possibly incorporating some of the
improvements described below, were done by several researchers for commer-
cial use, but details are not published.
The next signicant published advancement was the work of Greeley et al
(1989). They solved the analysis problem iteratively with the vortex wakes of
the sails convected along the horizontal components of streamlines at each step.
In addition, the sails and their images were heeled with respect to the image
plane. The resulting computer program has been broadly used to evaluate lift
and induced drag of sail shapes.
The best use of a sail analysis program is for comparing performance of
candidate designs. To do that, the form and friction drag of the sails must be
included and the resulting total sail forces used in a VPP. Adescription and block
diagram for such a procedure, which proved useful in sail shape development,
is given by Milgram (1993).
Ramsey (1996) has taken the next step in the development of numerical hy-
drodynamics of sails by including the aerodynamics of the above-water portion
of the hull. The sails are represented as vortex lattices with convected wake
geometries determined by iteration. The hull is represented by source panels,
as in the appendage numerical method of Section 5.4. If a sail is sealed to
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 649
Figure 19 Geometry and panelization for example sail computations. The view shown is from
the windward side with the boat heeled 20 degrees, with the viewpoint elevated and forward of the
midship.
the deck of the hull, its root circulation is carried on a continuation of the sail
through the deck down to the water plane. Figure 19 shows the hull, sails,
and panelization on an IACC conguration for which Ramsey (1996) has done
computations. The sailing conditions for this example are a wind speed of
4.52 m/s at a height of 10 m, a boat speed of 4.51 m/s, a heel angle of 20
degrees, a true wind angle of 40 degrees from ahead, and a boundary condition
of an impermeable free surface. The true wind velocity has a logarithmic prole
with a roughness height of 0.001 m for heights greater than 0.5 m, with a linear
reduction below this height to a speed at the free surface that is half the speed
at 0.5 m to avoid the actual steep gradient very close to the free surface. The
camber and twist distributions are chosen to be consistent with good sails in
practice. Computations were carried out for the sails without a hull, for the sails
and the hull together with spaces between the bottoms of the sails and the deck,
and for the sails and hull with the bottom of the jib sealed to the deck. Table 1
shows the results for the forward and side forces. The existence of the hull
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
650 MILGRAM
Table 1 Sail forces
a
Hull
Force None No jib seal With jib seal
Forward
Hull 14 39
Jib 2345 2470 3084
Main 1580 1615 1664
Total 3925 4071 4787
Side
Hull 717 490
Jib 6809 7080 8184
Main 8549 8648 8678
Total 15358 16445 17352
a
Units are Newtons.
Figure 20 Circulation distributions on the sails.
adds signicantly to the forces, particularly on the jib, which is closest to it. This
is particularly true when the bottom of the jib is sealed to the deck. Figure 20
shows the spanwise circulation distributions on the sails for each of these
cases. It demonstrates the inuence on the mainsail of the jib, which is of lesser
span.
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 651
7. CONCLUSIONS
Knowledge about the uid mechanics of sailing vessels continues to increase,
and the application of this knowledge to improving designs and design proce-
dures is becoming more universal. We can see where the eld is going, and
when this vision is combined with engineering and scientic judgement, it is
not difcult to predict what some of the important developments over the next
several years are likely to be.
For calm-water hydrodynamics, numerical methods, which include heel, will
become commonplace. Amajor step will have been achieved when viscous and
wavemaking resistances, along with their interactions, are computed together
with good accuracy. Both RANS and coupled outer-ow/boundary layer meth-
ods show promise, and only time will tell which, if either, can actually do the
job. The level of importance of separated hull vortices will be determined, and
if they are important, researchers will work at ways to determine their strengths
and locations.
In the determination of unsteady vessel motions and the associated added
resistance, detailed and accurate methods for predicting the pitch and heave-
damping caused by the sails could be developed very soon. Quantifying the
effects of large waves on the added resistance will probably take longer, largely
because difcult and expensive experiments will probably be part of the process.
The effects of the boundary layers on sail aerodynamics will almost certainly
be incorporated in the numerical methods within a very few years. The degree
to which these methods can accurately deal with the wake of the mast that then
streams across the sail attached to it remains to be seen.
Visit the Annual Reviews home page at
http://www.AnnualReviews.org.
Literature Cited
Campbell IMC. 1997. Optimisation of a sailing
rig using wind tunnel data. Proc. Chesapeake
Sailing Symp., Soc. Naval Archit. Mar. Eng.,
13th, pp. 4963
Caponnetto M. 1993. A review on Il Moro di
Venezia design. Proc. Chesapeake Sailing
Symp., Soc. Naval Archit. Mar. Eng., 11th,
pp. 10712
Dawson CM. 1977. A practical computer
method for solving ship-wave problems.
Proc. Int. Conf. Numer. Ship Hydrodyn., 2nd,
Univ. Calif., Berkeley, pp. 3038
Drela M. 1989. Xfoil: an analysis and design
system for low reynolds number airfoils. In
Lecture Notes in Engineering. Low Reynolds
Number Aerodynamics, Vol. 54., pp. 112.
New York: Springer-Verlag
Drela M, Giles MB. 1987. Viscous-inviscid
analysis of transonic and low Reynolds num-
ber airfoils. AIAA J. 25(10):134755
Euerle SE, Greeley DS. 1993. Towards a ratio-
nal upwind sail force model for vpps. Proc.
Chesapeake Sailing Symp., Soc. Naval Ar-
chit. Mar. Eng., 11th, pp. 7586
Farmer J, Martenelli L, Jameson A. 1995.
Yacht97: a fully viscous nonlinear free-
surface analysis tool for IACC yacht design.
Proc. Chesapeake Sailing Symp., Soc. Naval
Archit. Mar. Eng., 12th, pp. 15770
Gerritsma J, Beukelman W. 1972. Analysis of
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
652 MILGRAM
the resistance increase inwaves of a fast cargo
ship. Int. Shipbuild. Prog. 19:28593
Gerritsma J, Keuning JA. 1988. Speed loss in
waves. Proc. Symp. Dev. Interest Yacht Des.
Yacht Build., 10th, Amsterdam, pp. 84111
Gerritsma J, Moeyes G. 1973. The seakeeping
performance and steering properties of sail-
ing yachts. Proc. Symp. Dev. Interest Yacht
Archit., 3rd, Amsterdam, pp. 10734
Glauert H. 1959. Airfoil and Airscrew Theory.
London: Cambridge Univ. Press
Greeley DS, Cross-Whiter JH. 1989. Design
and performance of sailboat keels. Mar. Tech-
nol. 26(4):26081
Greeley DS, Kirkman KL, Drew AL, Cross-
Whiter JH. 1989. Scientic sail shape design.
Proc. Chesapeake Sailing Symp., Soc. Naval
Archit. Mar. Eng., 9th, pp. 3380
Herman JS. 1988. A sail force dynamometer:
design and data handling. Masters thesis.
MIT, Cambridge, MA. 51 pp.
Hess JL. 1972. Calculation of potential ow
about arbitrary three-dimensional lifting bod-
ies. Tech. Rep. MDC-J5679-01, Douglas
Aircr. Co.
Hoerner SF. 1965. Fluid Dynamic Drag. Mid-
land Park, NJ: Hoerner
Katz J, Plotkin A. 1991. LowSpeed Aerodynam-
ics. New York: McGraw-Hill
Kerwin JE. 1975. Avelocity prediction program
for ocean racing yachts. Tech. Rep. 75-17,
MIT, Cambridge, MA
Kerwin JE. 1994. The MIT marine hydrody-
namics water tunnela 53
rd
anniversary cel-
ebration. Mar. Technol. 31:18394
Kinnas SA. 1991. Hydrofoil lift and drag from
momentum integrations. Tech. Rep. 91-4,
MIT, Cambridge, MA
Kirkman KL, Pedrick DR. 1974. Scale effects
in sailing yacht hydrodynamic testing. Trans.
Soc. Naval Archit. Mar. Eng., pp. 77125
Larsson L. 1987. Numerical predictions of the
ow and resistance components of sailing
yachts. Proc. Conf. Yachting Technol., Univ.
West. Aust. Perth, pp. 2633
Larsson L. 1990. Scientic methods in yacht
design. Annu. Rev. Fluid Mech. 22:349
85
Larsson L, Broberg L, Kim KJ, Zhang DH.
1989. New viscous and inviscid cfd tech-
niques for ship ows. Proc. Conf. Numer. Hy-
drodyn., 5th, Hiroshima, Japan, pp. 185210
Lauchle GC, Gurney GB. 1984. Laminar
boundary-layer transition on a heated under-
water body. J. Fluid Mech. 144:79101
Lee JT. 1987. A potential based panel method
for the analysis of marine propellers. PhD
thesis. MIT, Cambridge, MA. 150 pp.
Liebeck RH, Camacho PP. 1985. Airfoil design
at low Reynolds number with constrained
pitching moment. Conf. Low Reynolds Num-
ber Airfoil Aerodyn., UNDAS-CP-77B123,
Univ. Notre Dame, pp. 2751
Mack LM. 1977. Transition prediction and lin-
ear stability theory. In Laminar-Turbulent
Transition, AGARD-CP-224, 1:122. Lon-
don: Hartford House
Mantzaris DA. Evaluationof the performance of
sailing vessels based on towing tank and sail-
force data. Masters thesis. MIT, Cambridge,
MA. 95 pp.
Masuyama Y, Fukasawa T. 1997. Full scale
measurement of sail force and the valida-
tion of numerical calculation method. Proc.
Chesapeake Sailing Symp., Soc. Naval Ar-
chit. Mar. Eng., 13th, pp. 2336
Milewski W. 1997. Three dimensional viscous
ow computations using the integral bound-
ary layer equations simultaneously coupled
with a low order panel method. PhD thesis.
MIT, Cambridge, MA. 108 pp.
Milgram JH. 1968a. The aerodynamics of sails.
Proc. Symp. Naval Hydrodyn., 7th, Rome, pp.
1397434
Milgram JH. 1968b. The analytical design of
yacht sails. Trans. Soc. Naval Archit. Mar.
Eng., pp. 11860.
Milgram JH. 1971. Section data for thin, highly
cambered airfoils in incompressible ow.
Contract. Rep. CR-1767, NASA
Milgram JH. 1993. Naval archicture technol-
ogy used in winning the 1992 Americas Cup
match. Trans. Soc. Naval Archit. Mar. Eng.
101:399436
MilgramJH. 1996. Hydrodynamics inadvanced
sailing vessel design. Proc. Symp. Naval Hy-
drodyn., 21st, Univ. Trondheim, Norway, pp.
7498
Milgram JH, Frimm FC. 1993. Elements of re-
sistance of IACC yachts. Proc. Chesapeake
Sailing Symp., Soc. Naval Archit. Mar. Eng.,
11th, pp. 22329
Milgram JH, Peters DB, Eckhouse DN. 1993.
Modeling IACC sail forces by combining
measurements with cfd. Proc. Chesapeake
Sailing Symp., Soc. Naval Archit. Mar. Eng.,
11th, pp. 6573
Miyata H. 1996. Time marching cfd simula-
tion for moving boundary problems. Proc.
Symp. Naval Hydrodyn., 21st, Univ. Trond-
heim, Norway, pp. 121
Morino L, Gennaretti M. 1992. Boundary
integral equation methods for aerodynam-
ics. In Computational Nonlinear Mechanics
in Aerospace Engineering, ed. SN Atluri,
146:279315. Washington, DC: AIAA
Mughal BH. 1992. A calculation method for the
three-dimensional boundary-layer equations
in integral form. Masters thesis. MIT, Cam-
bridge, MA. 74 pp.
Nakos DE, Kring D, Sclavounos PD. 1993.
Rankine panel methods for transient free sur-
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

P1: KKK
November 11, 1997 10:39 Annual Reviews AR049-20
SAILING VESSEL DESIGN 653
face ows. Proc. Int. Conf. Numer. Ship Hy-
drodyn., 6th, Univ. Iowa, pp. 61334
Nakos DE, Sclavounos PD. 1990. Ship motions
by a three dimensional rankine panel method.
Proc. Symp. Naval Hydrodyn., 18th, Univ.
Michigan, Ann Arbor, pp. 2140
Nishida BA. Fully simultaneous coupling of
the full potential equation and the integral
boundary layer equations in three dimen-
sions. PhD thesis. MIT Cambridge, MA. 126
pp.
Parsons BL, Pallard R. 1997. The institute for
marine dynamics model yacht dynamometer.
Proc. Chesapeake Sailing Symp., Soc. Naval
Archit. Mar. Eng., 13th
Peters DB. 1992. Determination of force coef-
cients for racingyacht sails usingactual force
measurements. Masters thesis. MIT, Cam-
bridge, MA. 54 pp.
Prohaska CW. 1966. A simple method for the
evaluation of the form factor and the low
speed wave resistance. Proc. Int. Towing Tank
Conf., 11th, pp. 6566
Ramsey W. 1996. Numerical methods for ows
around lifting bodies with vortex wake rollup.
PhD thesis. MIT, Cambridge, MA. 217 pp.
Raven HC. 1994. Invsicid cfd codes applied
in sailing yacht design. Ship Technol. Res.
Schiffstech. 41(3):16071
Rosen BS, Laiosa JP. 1997. Splash nonlinear
and unsteady free surface analysis code for
grand prix yacht design. Proc. Chesapeake
Sailing Symp., Soc. Naval Archit. Mar. Eng.,
13th, pp. 21125
Rosen BS, Laiosa JP, Davis JH, Stavetski D.
1993. Splash free-surface code methodol-
ogy for hydrodynamic design and analysis
of IACC yachts. Proc. Chesapeake Sailing
Symp., Soc. Naval Archit. Mar. Eng., 11th,
pp. 3549
Salvesen N, Tuck EO, Faltinsen O. 1970. Ship
motions and sea loads. Trans. Soc. Naval Ar-
chit. Mar. Eng. 78:25887
Sclavounos PD. 1995. Computation of ship
wave interactions. InAdvances in Marine Hy-
drodynamics, 4:177231. Southampton, UK:
Comput. Mech. Pubs.
Sclavounos PD, Huang Y. 1997. Rudder wing-
lets on sailing yachts. Mar. Tech. 34(3):211
32
Sclavounos PD, Nakos DE. 1993. Seakeep-
ing and added resistance of IACC yachts
by a three-dimensional panel method. Proc.
Chesapeake Sailing Symp., Soc. Naval Ar-
chit. Mar. Eng., 11th, pp. 24759
Skinner GT. 1982. Sailing vessel dyuamics:
investigations into aero-hydrodynamic cou-
pling. Masters thesis. MIT, Cambridge, MA.
135 pp.
Thwaites B. 1960. Incompressible Aerodynam-
ics. Oxford: Oxford Univ. Press
Van Oossanen P. 1993. Predicting the speed of
sailing yachts. Trans. Soc. Naval Archit. Mar.
Eng. 101:33997
A
n
n
u
.

R
e
v
.

F
l
u
i
d

M
e
c
h
.

1
9
9
8
.
3
0
:
6
1
3
-
6
5
3
.

D
o
w
n
l
o
a
d
e
d

f
r
o
m

w
w
w
.
a
n
n
u
a
l
r
e
v
i
e
w
s
.
o
r
g
b
y

P
o
l
i
t
e
c
n
i
c
o

d
i

M
i
l
a
n
o

-

B
i
b
.

C
e
n
t
r
a
l
e

d
i

I
n
g
e
g
n
e
r
i
a

o
n

1
1
/
2
3
/
1
1
.

F
o
r

p
e
r
s
o
n
a
l

u
s
e

o
n
l
y
.

Você também pode gostar