You are on page 1of 100

Marco Ferrando Motor Yacht Design

UNIVERSITY OF GENOVA

Master of Science course in

Yacht Design

Lecture Notes:

Motor Yacht Design

Prof. Marco FERRANDO

Genova, December 2013

rev. March 25th 2014 pag. 1 di 100


Marco Ferrando Motor Yacht Design

1  Dynamics ............................................................................................................................. 3 
1.1  Hull typologies ............................................................................................................... 3 
1.2  Planing hulls................................................................................................................... 8 
1.2.1  Definitions .............................................................................................................. 8 
1.2.2  Planing hull resistance: The Savitsky method ...................................................... 10 
1.2.2.1  Planing hull geometry ...................................................................................... 10 
1.2.2.2  Planing hull lift ................................................................................................ 15 
1.2.2.3  Planing hull resistance ..................................................................................... 18 
1.2.2.4  Planing hull centre of pressure......................................................................... 21 
1.2.2.5  Solution of the equilibrium equations.............................................................. 23 
1.2.2.6  Extension of the Savitsky method ................................................................... 26 
1.2.3  Planing hull towing tank tests ............................................................................... 27 
1.2.4  Methods for the resistance prevision .................................................................... 31 
1.2.4.1  Systematic Series ............................................................................................. 32 
1.2.4.2  Statistical or numerical methods ...................................................................... 50 
1.2.5  Other components of the resistance ...................................................................... 52 
1.2.5.1  Spray Resistance .............................................................................................. 52 
1.2.5.2  Air Resistance .................................................................................................. 60 
1.2.5.3  Appendage Resistance ..................................................................................... 60 
1.2.6  Porpoising ............................................................................................................. 64 
1.2.7  Spray Rails ............................................................................................................ 67 
1.2.8  Trim Control ......................................................................................................... 69 
1.2.8.1  Design options ................................................................................................. 69 
1.2.8.2  Active devices .................................................................................................. 74 
1.2.9  Surface piercing propellers ................................................................................... 80 
1.2.9.1  Mode of operation............................................................................................ 81 
1.2.9.2  Controlling parameters .................................................................................... 84 
1.2.9.3  Design issues ................................................................................................... 94 
1.2.9.4  Conclusions...................................................................................................... 96 
1.2.10  References............................................................................................................. 97 
1.3  Sea trials ..................................................................................................................... 100 

rev. March 25th 2014 pag. 2 di 100


Marco Ferrando Motor Yacht Design

1 Dynamics
The displacing and planing hull resistance is dealt with using different approaches. This results
from the planing hull extremely different behaviour with respect to that of the displacing hulls in
spite of it follows the same physical laws ruling a body movement at the interface between the wa-
ter and atmosphere

1.1 Hull typologies


Displacing hulls
We call displacing those hulls that, as the speed changes, maintain more or less unchanged the
hull volume. Just to give a general information we can say that the displacing hulls can found them-
selves in one of the following cases:
 displacing regime, Fr < 0.39 ( Fr  0.9 ): at these speeds the hull is completely dis-
placing, that is its weight is completely sustained by the buoyancy.
 semi-planing regime, 0.39< Fr <0.89 ( 0.9  Fr  1.0 ): in this speed range the dynam-
ic lift begins to have some effects, but has still a modest entity.
The hulls which must run in displacement regime are normally characterized by the following
features:
 small water plane entrance angle corresponding to the designed waterline
 convex buttocks, particularly in the stern area
 narrow stern (waterlines)
 convex bilge, normally with constant radius for a great part of the length.

A typical example of a hull designed to run in displacement regime is shown in Figure 1.1, and
the relevant lines plan is displayed in Figure 1.2.

Figure 1.1: Displacing hull form

rev. March 25th 2014 pag. 3 di 100


Marco Ferrando Motor Yacht Design

Figure 1.2: displacing hull lines plan

The hulls which must run in semi-planing regime are on the contrary characterised by:
 Thin entrance body
 Rectilinear buttocks slightly raising sternward
 Wide and partially submerged transom stern
 Round bilge for the whole hull length or round at the bow and sharpened sternward
 Rectilinear V-shaped sections at the bow

Figure 1.3: Semi-planing hull form

Figure 1.3 represents a hull destined to run in semi-planing regime, and the relevant lines plan
is shown in Figure 1.4.

rev. March 25th 2014 pag. 4 di 100


Marco Ferrando Motor Yacht Design

Figure 1.4: Semi-planing hull lines plan

Planing hulls
Planing hulls are those hulls that as the speed increases, under the effect of the dynamic pres-
sure which develops on the bottom, undergo to a remarkable reduction of the hull volume.This phe-
nomenon is shown in Figure 1.5, representing the different equilibrium positions as the speed in-
creases.

Figure 1.5: Equilibrium positions of a planing hull at different speeds


Just to give a general information we can say that the planing hulls can found themselves in
one of the following cases:
 pre-planing regime, 0.39< Fr <1.4 ( 1.0  Fr  3.0 ): the hull weight quote sustained
by the dynamic lift progressively increases while the hydrostatic component gradually
decreases because the hull volume decreases due to the progressive development of the
dynamic pressure on the bottom.
 planing regime , Fr > 1.4 ( Fr  3.0 ):the hull weight is supported mainly or almost
completely by the hydrodynamic lift.

rev. March 25th 2014 pag. 5 di 100


Marco Ferrando Motor Yacht Design

Figure 1.6: Planing hull form


The hulls designed to operate in planing regime are normally characterized by the following
features:
 Complete lack of convex surfaces (excepting for the bow that in any case results out of
the water in planing conditions);
 Sharp corner at the intersection between bottom and sides
 Wide transom stern well immersed, joined to the sides and the bottom with sharp cor-
ners to assure the full separation of the flow and guarantee the complete transom venti-
lation;
 Rectilinear and horizontal buttocks in the stern area;
 V-shaped sections with deadrise angle growing towards the bow.

A typical example of hull designed to operate in the planing regime is illustrated in Figure 1.6
and the relevant lines plan is displayed in Figure 1.7

Figure 1.7: Planing hull lines plan

rev. March 25th 2014 pag. 6 di 100


Marco Ferrando Motor Yacht Design

This classification is in any case qualitative since the beginning of the planing regime does not
exclusively depend on the speed of the craft but also on the longitudinal position of the centre of
gravity, denoted with XCG or LCG
With reference to Figure 1.5 the speeds of 10 and 20 knots correspond to the pre-planing re-
gime; real planing begins at 30 knots and it can be noted that also at this regime the hull continues
to decrease its submerged volume as the speed increases

Remarks on the resistance


Figure 1.8 represents the curves of the RT/ ratio (resistance on displacement) for displacing,
semi-planing and planing hulls.

Figure 1.8: Comparison of Resistance curves of different hull types


Looking at the figure it is possible to observe that in the Froude number range up to 0.39 the
displacing hull is that offering the best performances; beyond this limit the RT/ ratio takes rapidly,
for the displacement hull, really higher values than those typical of the semi-planing hull. A similar
behavior can be noted for the semi-planing hull, when it is brought to values of the Froude number
higher than 0.89 for which the planing hull is that having the lower value of the RT/ ratio.
It is then evident that the hull shape choice must be suitable to the speed regime in which the
craft will be used.

rev. March 25th 2014 pag. 7 di 100


Marco Ferrando Motor Yacht Design

1.2 Planing hulls


As concern the planing hulls in theory the resistance could be divided into the same compo-
nents already described in the case of the displacement hulls; particularly it is possible to make the
following qualitative distinction:
 pre-planing regime, 0.39< Fr <1.4 ( 1.0  Fr  3.0 ): at this regime the resistance due
to the breaking waves and the resistance of the eddies disappear; the wave resistance
decreases with the speed (because of the progressive raising of the hull from the water)
whilst the friction resistance increases with the square of the same speed.
 planing regime , Fr > 1.4 ( Fr  3.0 ) the hull weight is almost completely supported
by the dynamic lift; the wave resistance is negligible, the friction resistance becomes
the preponderant component followed by the induced resistance, the spray resistance
and that of the appendages.
Really for this hull typology is necessary to consider the hydrodynamic effects developing on
the hull bottom, which contribute to balance the hull weight provoking a decrease of the immersed
hull volume as the speed increases.
For the displacement hull we have:
W  S   (1)
while for the planing hull case equation (1) transforms into:
W   V   L V  (2)

where L denotes the vertical component of the dynamic pressure resultant acting on the hull bot-
tom.It increases at the V speed increasing, as a result, to maintain the equilibrium in the vertical di-
rection, the immersed hull volume must decrease.
In the field of the planing hulls, where the quantity of the experimental data concerning the re-
sistance is remarkably lower than in the case of the displacement hulls, it is very common the use of
the Daniel Savitsky scheme, who was the first to deal in a comprehensive way of the planing hull
resistance prediction.
1.2.1 Definitions
Since the planing hull, unlike from the displacing ones, emerges from the water as the speed
increases, the quantities used to describe the displacing hulls as the length of the waterline, the wa-
terline breadth etc. lose in this case significance since the waterline modifies as the speed changes.

Figure 1.9: Projection of the Bottom


Usually the planing hull is described according to the properties of the bottom surface. Since
planing hulls are characterized by a sharp knuckle, named chine, the bottom surface is easily de-
fined as the hull surface that lies below the chine. Particularly the hull bottom projection on a hori-
zontal plane is considered, as shown in Figure 1.9

rev. March 25th 2014 pag. 8 di 100


Marco Ferrando Motor Yacht Design

The area of this projection is denoted with the AP symbol. This projection is more over used to
define the LP length, called chine projection length, and the BPC, BPX and BPT breadths called bottom
breadth (in generic longitudinal position), maximum bottom breadth and transom breadth respec-
tively. The above defined dimensions are shown in Figure 1.10.

Figure 1.10: Projection of the bottom and related dimensions


It is moreover used also the average bottom width, defined by the relation:
AP
BPA 
LP
Finally the bottom loading coefficient is used, defined as follows:
AP
2
 3

The David Taylor Model Basin developed a method to represent the most important shape
characteristics of a V bottom hull using three "Shape characteristic curves"; these three curves show
in non-dimensional way the longitudinal distribution of the BPC width, of the  deadrise angle and of
the average buttock height with respect to the base line.

Figure 1.11: Shape characteristic curves of a V-bottom hull

The average buttock is defined as that buttock BPA/4 far from the symmetry plane. An example
of this description is shown in Figure 1.11. It should be noted that the average buttock heights are

rev. March 25th 2014 pag. 9 di 100


Marco Ferrando Motor Yacht Design

measured from a line, parallel to the base line, tangent to the average buttock at the aft end of the
hull bottom.
1.2.2 Planing hull resistance: The Savitsky method
The planing hull physical model conceived by Savitsky [1.2.10.1.1] is based on the study of the
planing hull equilibrium condition under the effect of the forces, shown in Figure 1.12, which are:
 W Craft weight
 N Resultant of the pressure acting on the hull bottom
 DF Frictional resistance due to the water flowing on the hull bottom
 T Propeller thrust
The hull must be in completely developed planing conditions, that is that the water leaves the
hull at the chine, without touching the side and the transom is dry.
According to Saunders [1.2.10.1.2] the submerged transom ventilation occurs if:
V
Frh   45 (3)
gh
where h denotes the transom immersion; according to Savitsky and Brown [1.2.10.1.3] for the plan-
ing hulls the ventilation takes place if:
V
CV   0.5 (4)
gb
where b denotes the hull breadth.

Figure 1.12: Physical model of the planing hull

1.2.2.1 Planing hull geometry


The considered hull is prismatic, that is having breadth b and deadrise angle  constant along
the whole length, excepting for a small forward part which remains in any case out of the water.
The said hull is studied in full planing condition, that is with the transom ventilated and dry sides, in
given immersion, speed and trim conditions. This situation is represented in Figure 1.13

rev. March 25th 2014 pag. 10 di 100


Marco Ferrando Motor Yacht Design

Figure 1.13: Geometry of the planing hull

rev. March 25th 2014 pag. 11 di 100


Marco Ferrando Motor Yacht Design

To deal with the planing hulls it is necessary to note that the hull movement through the water
perturbs the water free surface. This one rises giving origin to the first wave train diverging from
the hull. In this way the wetted hull bottom surface is no more that denoted by the undisturbed wa-
terline, shown in Figure 1.13by the polygon OQ’R’R”Q”, but that delimited by OP’R’R”P”; the tri-
angular areas OP’Q’ e OP”Q” are then also wetted and contribute to form that which is called "Bot-
tom Pressure Area" or simply “Pressure Area”.
The pressure area remains delimited at the bow by the line OP, called "Spray root line". This
denomination takes its origin from a particular phenomenon, the spray production, characterizing
the high speed hulls. The spray is a biphasic fluid composed by air and very small water droplets.
To describe this phenomenon consider the simple case of a planing hull with a flat bottom.
Represented in Figure 1.14
Observing the picture we can note the rising of the fluid surface, due to the forward wave, and
the thin region, at forward of this last, really concerned by the spray. In the spray area, the flow
speed is consistent with the hull advancing speed; in other words the spray is flung forward.

Figure 1.14: Planing flat plate

Figure 1.14 is the reproduction of a picture contained in the original work of Savitsky and ex-
ploits the non dimensional average wetted length concept.
LM
 (5)
b
where LM denotes the average wetted length of the hull and b the bottom width in correspondence of
the chine. Undoubtedly the average wetted length concept has not significance in case of flat bot-
tom, but it takes importance in the case of V bottom craft (0) as it will be shown later.
Considering again Figure 1.14 we can note as the actual wetted length, denoted with b is
greater than the wetted length which would result from the intersection of the hull with the undis-
turbed water surface.
Figure 1.15 shows the same situation considered in the previous picture but it describes as the
dynamic pressure ends practically in correspondence of the spray root. The spray thickness, denoted
with , is defined as the distance between the undisturbed water surface and the current line ending
in the stagnation point.
It is then evident how it is important to determinate the real area of the bottom surface under
pressure. This problem has been experimentally faced for as concern the planing hulls with flat bot-
tom, for which two relations have obtained linking  to ’:
  1.6  0.302  0    1 (6)

     0.30 1     4 (7)

rev. March 25th 2014 pag. 12 di 100


Marco Ferrando Motor Yacht Design

These relations have been derived through regression procedure applied to the experimental da-
ta and they are valid for:
2    24
4
0.6  CV  25.0

where
V
CV  (8)
gb

Figure 1.15: Pressure distribution on a planing flat plate

If we consider now the hulls with 0 we can observe that the stagnation line, which is the lo-
cus of stagnation points at the changing of the distance from the keel line, is no more orthogonal to
the keel line as in the case of the flat bottom hull. As a result, also the spray root line is no more or-
thogonal to the keel line, as shown by the OP, OP’ and OP” segments in Figure 1.13. It results that
the average wetted length can be obtained by the average of the keel wetted length LK and the chine
wetted length LC
With reference to Figure 1.16, we can observe that while the LK length can be easily deter-
mined:
d
LK  (9)
sin 
it is not the same for the LC length; to determine this last the quantity L1 it is used.
To obtain a relation between the b1 theoretic wetted width and the b1e effective one, Savitsky
made use of the work of Wagner [1.2.10.1.4] from which it is obtained:

b1e  b1 (10)
2

rev. March 25th 2014 pag. 13 di 100


Marco Ferrando Motor Yacht Design

Figure 1.16
Let us consider again Figure 1.16; the PP” and P’P” segments length in the longitudinal view
can be obtained observing the A-A section:
PP  b1e tan 
PP  b1 tan 
observing then that in the section A-A we have:
b
b1e 
2
it is immediate to obtain
b
PP  tan 
2
and remembering equation (10), we obtain:
2 2b b
PP  b1e tan   tan   tan 
  2 
If we consider now the OP’P” square triangle, we can obtain:
b tan 
L1  (11)
 tan 
which is valid for any  and  combination if CV>2. For the validity field in the case in which the
speed coefficient is smaller than 2 see the original work of Savitsky.
The average wetted length LM can be obtained:
L1 d b tan 
LM  LK   
2 sin  2 tan 
from which we obtain the non-dimensional average wetted length:

rev. March 25th 2014 pag. 14 di 100


Marco Ferrando Motor Yacht Design

d b tan 

L
  M  sin  2 tan  (12)
b b

1.2.2.2 Planing hull lift


It is opportune to underline at first that the lift, denoted with L, is the force developing on a sur-
face in relative movement with a fluid, in orthogonal direction to the speed with which the craft
moves with respect to the undisturbed fluid. In the case of the planing hull, then, the lift is a force
having vertical direction. Savitsky, in its work, assumes that the craft weight is totally balanced by
the lift. Obviously the lift is one of the components of the N resultant force from the pressures act-
ing on the hull bottom, as shown in Figure 1.17 for the case of a planing surface with =0 in
movement in an ideal fluid. The second component of the N resultant force is a resistance which is
denoted with DP as it will be shown later.

Figure 1.17: Components of the pressure resultant vector


Savitsky obtained an empiric equation for the lift coefficient of a surface having =0, using a
wide experimental data collection with which, by regression techniques, he could obtain the coeffi-
cients of the equations that he formulated.
To formulate the lift equation Savitsky used the analogy between the lifting hull surface and a
wing. Therefore it is necessary to remember that the lift coefficient relative to a planing surface can
be divided into two components, since the pressure acting on the planing hull bottom has a hydro-
static component and a hydrodynamic one. The analogy with a wing, limits to the sole dynamic lift
component, while for the hydrostatic part a different approach is used.
For wings having high aspect ratio (ratio between wing span and chord), where the flow goes
mainly in the chord direction, the dynamic lift coefficient is proportional to the incidence angle; in
the case of wings with low aspect ratio the flow takes a predominantly transversal direction and the
lift coefficient is proportional to the square of the incidence angle. In the case of the planing surface
the average wetted length  can be considered as the inverse ratio with respect to the aspect ratio of
a wing.
It is therefore possible to assume that the dynamic lift coefficient of the planing surface, which
has intermediate aspect ratio values, can be expressed in the following way:

rev. March 25th 2014 pag. 15 di 100


Marco Ferrando Motor Yacht Design

C Ld  A  B 2 (13)

Limiting the analysis to the  value range characteristic of the planing crafts, the second term
of the second member of equation (13) becomes a small correction to the first term and the (13) can
be reformulated in terms of 1.1 in the following way:
CLd   1.1 f  , CV  (14)

The analysis carried out by Sottorf [1.2.10.1.5] on experimental data concerning high speed
planing surfaces, for which the hydrostatic contribute to the lift is negligible, showed that, for a
fixed  trim angle, the lift coefficient dynamic component changed proportionally to 1\2. According
to this assumption equation (14) becomes:

CLd  c 1.1  (15)

The hydrostatic lift, denoted by Savitsky with Lb, is determined referring to the hull volume de-
fined by the undisturbed waterline. Remembering the equation (7) it is possible to obtain the corre-
sponding wetted keel length
     0.30

Figure 1.18: Schematization of the hull volume


The area of the right triangle which constitutes the trace of the hull volume on the side is given
by the relation
1
   0,30  b    0,30  b tan 
2
Multiplying this area by the width b we obtain the hull volume which if multiplied for the spe-
cific weight    g gives the hydrostatic lift Lb:

rev. March 25th 2014 pag. 16 di 100


Marco Ferrando Motor Yacht Design

1
 gb3    0.30  tan 
2
Lb  (16)
2
The hydrostatic lift coefficient can be obtained dividing both the members of equation (16) by
the quantity 1  b 2V 2 obtaining:
2
gb
2 
  0.30  tan 
2
C Lb 
V
Making reference to the equation (8), accepting that, from the moment that the  angle always
assumes small values, tan  can be approximated by 1.1 and assuming finally to approximate (-
0.30)2 with the quantity Dn it is possible to obtain:
D n 1.1
CLb   (17)
CV2

The lift coefficient for the planing surface with =0, denoted with CL0, can so be obtained
summing the equations (15) e (17):
 D n 
CL 0   c 1 2  2  1.1 (18)
 CV 

The values of the constants c, D and n has been obtained applying the relation (18) to a huge
quantity of experimental data as shown in [1.2.10.1.6] and the final result is:
 0.0055 5 2  1.1
CL 0   0.012 1 2  2  (19)
 CV 
where  is expressed in degrees.

Figure 1.19: Effect of deadrise on the Lift

rev. March 25th 2014 pag. 17 di 100


Marco Ferrando Motor Yacht Design

Equation (19), having been obtained by analyzing the results of the experimental tests, shows
the following applicability field:
0.6  CV  13.0
2    15
4
It is interesting to note that even for the planing hull physical model conceived by Savitsky it
takes place the situation for which at the low speeds, that is in the field 0.6<CV<1.0, a hull immer-
sion greater than the static one can be recorded. This effect is quite similar to the “sinkage” which
can concern the displacing hulls at the low speeds.

Let us consider now the hull with 0; according to simple geometric assumptions it is possi-
ble to argue that the dynamic lift should be lower and decreases as  increases. Observing indeed
Figure 1.19 displaying two parts of transversal section, it is possible to note that, at parity of inten-
sity of the vector resultant from the pressures N, its component in vertical direction, that is the lift,
decreases as  increases.
To formulate an equation for the dynamic lift coefficient relevant to a planing surface having
0 experimental data concerning flat and V surfaces at parity of CV,  and  have been compared.
It has been noted that said coefficient could be represented by the equation:
CLβ  CL0  0.0065 CL0.60
0 (20)

where  is expressed in degrees.


Bearing in mind that, according to the Savitsky theory assuming the dynamic lift equal to the
craft weight, we have also:
W
CLβ  (21)
1  b 2V 2
2

1.2.2.3 Planing hull resistance


The approach used by Savitsky for the planing hull resistance is very different from the classic
one concerning the displacing hulls. Savitsky, indeed, considers only two components of the re-
sistance; the frictional resistance and the pressure resistance.
The main assumption at the base of the Savitsky’s approach is that the lift L which should be
developed by the hull must be equal as module to the weight W of the craft.
In the case that the fluid is considered ideal, that is in absence of friction, the resultant N of the
pressure on the hull bottom must have a vertical L component with an intensity equal to W. In this
case, shown in Figure 1.17, the resistance is composed by the horizontal component DP of the resul-
tant vector N. This is a pressure resistance being a component of the resultant of the pressures on
the bottom and it is denoted with the symbol DP (Drag Pressure). Remembering the fundamental as-
sumption L=W this can be written in the following way:
DP  W tan  (22)
Let us consider now the case of real fluid represented in Figure 1.20; on the bottom of the hull
acts the friction, due to the relative motion between the hull surface and a viscous fluid, which will
be indicated by the symbol DF (Drag Friction)

rev. March 25th 2014 pag. 18 di 100


Marco Ferrando Motor Yacht Design

In this case the vertical component of the resultant vector from the pressures must have an in-
tensity equal to the sum of the craft weight and the DFV vertical component of the frictional drag.
This gives origin to a further horizontal component of the resultant vector from the pressures, hav-
ing an intensity L1 tan . But being L1= DFV and DFV =DF sin  the pressure resistance RP will be
given by:
RP  W tan   DF sin  tan 

The friction resistance shall be expressed by:


RF  DF cos 

Figure 1.20: Total resistance components


Observing Figure 1.20 we note that
DF
DF sin  tan   DF cos 
cos
therefore the total planing hull resistance can be written in the form:
DF
RT  W tan   (23)
cos
For as concern DF it has been demonstrated [1.2.10.1.6] that it can be obtained by the relation:
1 b2 2
DF  CF  V1 (24)
2 cos 
where:
 V1 = mean velocity on the hull bottom
 CF Shoenherr friction coefficient (ATTC’47)

rev. March 25th 2014 pag. 19 di 100


Marco Ferrando Motor Yacht Design

Since for the planing hull the pressure on the bottom is greater than the hydrostatic pressure,
the average speed on the bottom V1 must result, according to the Bernoulli theorem, lower than the
speed V of the craft. To determine the value of the mean bottom velocity Savitsky [1.2.10.1.8] de-
veloped the following methodology: we consider the dynamic lift coefficient of the flat surface,
given by the first term of the second member of equation (19)
1
CL0d  0.012 2 1.1 (25)
the coefficient relative to the case of V hull can be obtained applying the structure of equation (20)
to the previous one obtaining:
CL d  CL0d  0.0065 CL0.60
0d (26)

by means of this equation it is possible to obtain the dynamic lift; allowing, at its turn, to obtain the
mean dynamic pressure on the bottom
Ld
pd  (27)
b cos
2

1 2 2 1
pd  CL d b V (28)
2 b cos
2

Applying the Bernoulli theorem to two points of the same immersion, one in the undisturbed
current and one in contact with the hull, denoted with the subscript 1, we obtain:
1 1
p  V 2  p1  V12
2 2
which can be written again in the form:
1 1
V12  V 2   p1  p 
2 2
Remembering that the difference between the two pressures is equal to the dynamic pressure
we can write:
1
V1  2 pd  2
 1   (29)
V  V 2 
At this point, with the value of V1 is possible to calculate the Reynolds number
V1b
Re  (30)

to obtain the CF and, then, use this last to determinate the DF.

rev. March 25th 2014 pag. 20 di 100


Marco Ferrando Motor Yacht Design

Figure 1.21: Optimum trim definition


Figure 1.21 displays a diagram reproduced by the original publication of Savitsky, displaying
the ratio RT  as a function of the trim angle  for the planing surfaces having a width equal to
3.048m. Observing the said diagram it is evident that there is a trim angle, for each value of the 
angle, at which a minimum of the resistance of the planing hull corresponds; the value of this opti-
mum trim angle is a function of  and it is about 4°.

1.2.2.4 Planing hull centre of pressure


The longitudinal position of the centre of pressure of the hull bottom is determined with refer-
ence to the components of the same pressure. Dealing with the lift, we saw that the pressure acting
on the hull bottom can be thought as the sum of two contributions, the hydrostatic and the dynamic
ones. As result the lift coefficient has been formulated as sum of the hydrostatic and hydrodynamic
contributions, see equations (15), (17), (18) and (19):
CL 0  CL 0D  CL 0B (31)
1
CL 0D  0.012 2 1.1
0.0055 5 2 1.1 (32)
CL 0B  
CV2

rev. March 25th 2014 pag. 21 di 100


Marco Ferrando Motor Yacht Design

For as concerns the hydrostatic component of the lift, we can assume that, considering the
shape of the hull volume shown in Figure 1.18, this is applied to a point placed at lB=LM/3 forward
of the transom.

Figure 1.22: Vectors used for the determination of the position of the center of pressures
For as concern the dynamic lift component the planing hull is assimilated to a wing and, by
analogy with what occurs for the aerodynamic lift where the lift takes place at ¾ of chord starting
from the wing trailing edge, we assume that the hydrodynamic component of the lift is applied to a
point located at lD=3/4LM forward of the transom.
Calculating the first order moments of the system of forces represented in Figure 1.22 we ob-
tain:
CL 0lP  CL 0D lD  CL 0BlB

Dividing both members by CL0lM we obtain:


lP CL 0D lD CL 0B lB
 
lM C L 0 lM C L 0 lM

and remembering that C L 0D  CL 0  CL 0B , we obtain:

lP CL 0  CL 0B lD CL 0B lB
 
lM CL 0 lM C L 0 l M
from which it is possible to obtain:

lP lD CL 0B  lD  lB 
   
lM lM C L 0  lM 

that, by convenience, it is better to write again in the form:

l P l D   l D  l B  lM 
   (33)
lM lM  CL 0 CL 0B 

rev. March 25th 2014 pag. 22 di 100


Marco Ferrando Motor Yacht Design

Remembering the equation (31) the ratio CL 0 CL 0B can be written in the form:

CL 0 CL 0D
 1
CL 0B CL 0B
that by means of the relations (32) takes the form:
1
CL 0 0.012 2
 1
CL 0B 0.0055 5 2
CV2
which becomes:
CL 0 2.182CV2
 1
CL 0B 2
Remembering the assumptions made about lD and lB, the equation (33) can be now be rewritten
in the form:
 
lP  0.75  0.3333 
 0.75   2 
lM  2.182CV  1 
 2 
which becomes:
 
l  1 
CP  P  0.75   2  (34)
lM  5.21 CV  2.39 
 2 

As we can note the equation (34) does not contain neither  nor  according to the previous
studies [1.2.10.1.6] which had demonstrated that the centre of pressure position is practically inde-
pendent from the values of  and.

1.2.2.5 Solution of the equilibrium equations


We are now able, referring to Figure 1.12, to write the following equilibrium equations system
composed by the translation equilibrium equations in vertical and horizontal direction and by the
rotation equilibrium equation around the centre of gravity:
W  N cos   T sin(   )  DF sin  (35)

T cos(   )  DF cos   N sin  (36)

Nc  DF a  Tf  0 (37)

The problem data are: W, , , b, XCG ZCG and the levers a, c and f. Savitsky suggests a numeri-
cal solution of the equation system, assuming the trim angle  from which all the quantities in the
system depend on.
The system solution is more easy if we consider the equilibrium equation to the translation
along the keel line:
T cos    sin   DF (38)

rev. March 25th 2014 pag. 23 di 100


Marco Ferrando Motor Yacht Design

from which, assuming cos   1 , we obtain:


T   sin   DF (39)

Replacing the T expression in the (35), we obtain:


  N cos   (  sin   DF ) sin(   )  DF sin 
  N cos    sin  sin(   )  DF sin(   )  DF sin 
Assuming that DF is smaller than the other concerned forces it is possible to say that the quanti-
ty DF sin(   )  DF sin  is negligible and write:
  N cos    sin  sin(   )
from which it is possible to obtain:
 1  sin  sin(   )
N (40)
cos 
Replacing now the expressions of T and N into (37) we obtain:
 1  sin  sin(   ) c 
  f sin    DF  a  f   0 (41)
 cos  
Once obtained the values of , c and DF satisfying equation (41) the boat is in equilibrium and
it is possible evaluate its performances.
In Table 1.2-1 it is displayed a calculation example in which we assume the values of  which
would allow to evaluate the result of the second member of equation (41) and to find the trim angle
interval where the change of sign occurs. The final solution can be obtained once obtained the trim
angle verifying the rotation equilibrium equation (41); in this case we have operated by interpola-
tion.
Table 1.2-1
Quantity Equation Value Units Result
W datum 267 kN
XCG datum 8.839 m
ZCG datum 0.61 m
b datum 4.267 m
 datum 10 °
V datum 40 kn
f datum 0.152 m
 datum 4 °
CV (8) 3.18 =
CL (21) 6.75E-02 =
CL0 (20) 0.082 =
 assumed 1 2 3 ° 2.20
 (19) 6.21 3.77 2.52 = 3.46
V1 (29) 20.541 20.470 20.367 m/s 20.452
Re (30) 4.58E+08 2.77E+08 1.84E+08 = 2.54E+08
CF ATTC’47 1.69E-03 1.80E-03 1.90E-03 = 1.82E-03
CF+CF CF=0.0004 2.09E-03 2.20E-03 2.30E-03 = 2.22E-03

rev. March 25th 2014 pag. 24 di 100


Marco Ferrando Motor Yacht Design

DF (24) 51847 32975 22808 kN 30.511


CP (34) 4.840E-01 5.861E-01 6.566E-01 =
c -3.977 -0.586 1.786 m
a 0.422 0.422 0.422 m
RT (23) kN 40.801
PE kW 839.9
m (41) -103.2698 -147.079 475.605 kNm 0

This relatively simple approach has the defect that it cannot be applied during the preliminary
phases of a craft design since that, in this circumstance, the quantities , ZCG and the levers a and f
are not precisely known or are absolutely unknown.
To get around to this situation Savitsky proposed a simplification of his method based on a fur-
ther assumption that all the concerned forces pass through the boat centre of gravity.
According to this assumption equation (37) is identically verified and the system composed by
equations (35) and (36) gives:
N   cos (42)
and produces the transformation of the equation (34) into the following:
  
  1  
CP  b  lP  LCG  0.75   2   b (43)
 C
 5.21 2  2.39  
V

 
   

Table 1.2-2 displays a calculation example based on this simplified procedure. In many cases,
as in that shown in the two calculation examples, the simplifying assumption is not far from the re-
ality and the two procedures give very similar results.

Table 1.2-2
Quantity Equation Value Units
 datum 266.99 kN
XCG datum 8.839 m
b datum 4.267 m
 datum 10 °
V datum 40 kn
CV (8) 3.18
CL (21) 6.75E-02
CL0 (20) 8.20E-02
 (43) 3.43
 (19) 2.23
V1 (29) 20.45 m/s
Re (30) 2.51E+08
CF ATTC’47 1.83E-03
DF (24) 30.23 kN
RT (23) 40.63 kN
PE RTV 836.14 kW

rev. March 25th 2014 pag. 25 di 100


Marco Ferrando Motor Yacht Design

1.2.2.6 Extension of the Savitsky method


The presence of the characteristic “Hump” in the resistance curve of planing hulls, in corre-
spondence of the pre-planing regime, makes absolutely fundamental to know the behaviour of the
hull in this regime to be able to guarantee a good matching of the propeller-hull system with the en-
gine.
To solve this problem Blount e Fox [1.2.10.1.9] developed what they had defined a “Engineer-
ing Method” based on the Savitsky methodology for the fully planing condition.
The method consists in the application of a multiplicative m correction to the resistance value
obtained with the Method of Savitsky applied for speed corresponding to the pre-planing regime.
The correction is defined of engineering character since it has been obtained with mathematical-
statistical methods applied to towing tank model tests results and to the calculations with the
Method of Savitsky of the corresponding conditions. By comparing the experimental data with the
calculations the authors have elaborated the following expression of the m correction factor:
1.45
X  2 Fr  0.85 X  3 Fr 0.85
m  0.98  2  CG  e  3  CG e (44)
 BPX   BPX 
The correction is conceived so that to annul itself when it is not yet necessary and depends ex-
clusively on the ratio between the longitudinal abscissa of the centre of gravity and the maximum
breadth at the chine X CG BPX as well as on the volumetric Froude number.
Figure 1.23 displays the m correction factor trend with the variation of the parameters affecting
it.

Figure 1.23: Blount & Fox correction factor m


The use of the m correction factor extends the validity of the Savitsky method, which is so ap-
plicable starting from Fr  1.0 .
Blount and Fox have moreover investigated on the possibility to use the Savitsky Method even
for non prismatic hulls, which were the majority in those years.

rev. March 25th 2014 pag. 26 di 100


Marco Ferrando Motor Yacht Design

The investigation concerned the breadth and the deadrise angle to input into the Savitsky
Method as representative of a prismatic hull. For as concern the breadth the authors concluded that
the chine breadth which better represents the non prismatic boat is BPX, while for the deadrise angle
they did not attained definitive conclusions and suggest to use the deadrise angle measured at half
of the length LP.
With the time passing the planing hull geometry approached itself a lot to the prismatic hull as-
sumed by Savitsky and the correction of Blount and Fox revealed itself as excessive. For this rea-
son, in a following publication [1.2.10.1.20], Donald Blount suggested to half the correction result-
ing from the original formulation. Therefore we have:

m 
 m  1 (45)
2
where with m it is denoted the value given by equation (44).
XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX
1.2.3 Planing hull towing tank tests
High Speed Marine Vehicles are for this purpose defined to be vessels with a design speed cor-
responding to:
Fr  0.45
(46)
Fr  1.18
The testing of resistance of HSMV is in many respects very similar to testing the resistance of con-
ventional displacement ships. They require special care during the design and executions of towing
tank tests since three peculiarities differentiate them from displacing hulls:
 Dynamic lift and trim have a greater influence,
 air resistance is not negligible and can affect the trim of the craft,
 scale effect on lifting surfaces and appendages can generate problems.

The primary focus of these comments is on semi-displacement mono-hulls and catamarans as


well as planing hulls.

Models
It should be noted that compared with conventional displacement ship models, many HSMVs
require special attention to minimizing the model weight. This is especially the case for models that
are going to be used for propulsion tests.
It also is recommended that the model be equipped with a superstructure with the same basic
shape and main dimensions as that of the ship. Adequate grid reference lines must be applied for es-
timating the dynamic wetted area.
The application of a boundary layer turbulence stimulation is recommended when the Reynolds
number is less than 5×106 based on mean or effective wetted length. For models tested solely at
higher Reynolds numbers, turbulence stimulation might be omitted.
The use of trip wires is not recommended on high speed models due to the risk of air suction.
For vessels with significant change in running attitude with speed, great care must be taken in the
placement of the turbulence stimulation. Test runs must be carried out if there are doubts about the
placement.
The resistance of appendages is often an important and difficult question for HSMVs but the fol-
lowing basic approximate rule is offered: Appendages not used for producing lift or altering the
trim could be left off the model and the computed resistance of these appendages added in the ex-

rev. March 25th 2014 pag. 27 di 100


Marco Ferrando Motor Yacht Design

trapolation to full scale. Appendages required for the propulsion test (if such a test is to be carried
out) must be present. For small models it is advisable to leave out appendages following the above
rule in order to avoid problems with laminar separation. For large models it can be beneficial to in-
clude appendages, at least the ones located in the wake affected area in the aft part of the model.
Turbulence stimulation is recommended for appendages protruding out of the boundary layer of the
model.
The size of HSMV appendages is often too small to obtain a Reynolds number of 5×106. In such
cases, turbulence stimulation on the appendages might be a reasonable solution.

Installation
The application of the tow force should be such that it resembles the direction of the propulsion
force as closely as possible. This is in order to avoid artificial trim effects due to the tow force. The
preferred way of doing this is to tow in the elongation and the direction of the propeller shaft.

Figure 1.24: Towing arrangement

If this cannot be accomplished, then the artificial trim moments introduced by the towing should
be corrected for by an appropriate shift in the XCG. An alternative is to test the model fixed to the
carriage in a range of different heave and trim values.
The resistance is taken as the horizontal component of the applied tow force.
Guides may be fitted to prevent the model from yawing or swaying: these should not restrain
the model in any other direction of movement, nor be able to impose any force or moment on the
model which would cause it to roll or heel. The arrangement of any such guides that include sliding
or rolling contacts should introduce the least possible friction forces. The model should be posi-
tioned in a way that it is in the centerline of the tank and parallel to the tank walls.
If any instruments carried on the model are linked to the carriage by flexible cables, great care
should be taken to ensure that the cables do not impose any force on the model in the running con-
dition; in practice the cables should therefore hang vertically from the carriage. Care should also be
taken to balance any instruments that must have attachments to both the model and the carriage (e.g.
mechanical trim recorders).
Sinkage fore and aft may be measured with mechanical guides, potentiometers, encoders or with
remote (laser or ultrasonic) distance meters; the running trim is then calculated from the measured

rev. March 25th 2014 pag. 28 di 100


Marco Ferrando Motor Yacht Design

running sinkage fore and aft. Alternatively, the running trim may be measured directly using an an-
gular measuring device with the measurement of the sinkage at one point.

Analysis of model scale results


Resistance RTM measured in the resistance tests is expressed in the non-dimensional form
RTM
CTM  1
2  M S0MVM2
It should be noted that the observed running wetted surface area will normally be used for
HSMVs. It is however practical to use a constant value of wetted surface for non-
dimensionalization, so in the formula above, S0 is a nominal wetted surface, for instance the value
for zero speed, while S is used as symbol for the running wetted surface.

The residual resistance coefficient CR is calculated without the use of a form factor k:
CR  CTM  CFM  S M S0M  CAAM  CAppM

where CFM is derived from the ITTC1957 correlation line for the model, CAAM is the model wind
resistance coefficient, and CAppM is the model appendage resistance coefficient (if appendages are
present and their resistance scaled separately). CAppM can be found by calculation or from the differ-
ence in resistance by testing with and without appendages.
The CR or CT curve is the best basis for judging if a sufficient number of test points have been
obtained in order to define humps and hollows. The model resistance curve should be faired in or-
der to facilitate reliable interpolation to obtain the resistance at the required speeds. The smoothing
should be carried out with care in order not to remove humps and hollows.

Extrapolation to full scale


The total resistance coefficient of a HSMV is
CTS  CR  CFS  SS S 0S  CAAS  CAppS  CA

where
CFS is the frictional resistance coefficient of the ship according to the ITTC-1957 model-ship
correlation line
CR is the residual resistance coefficient obtained by the analysis of the model test results.

CAA, is the air resistance coefficient


A VA2 AV CD
CAA 
 V 2 S0
The equation for air resistance coefficient CAA is used for both model (CAAM) and full scale
(CAAS). For model scale the wind velocity VA might be different from the through water velocity V
due to carriage displacement effects. In addition, the wind area AV and the drag coefficient CD
might be different in model and full scale. However, if VA=V, and AV and CD are considered to be
equal in model and full scale, then the wind resistance might be left out of the extrapolation process.

CAppS is the appendage resistance coefficient of the ship. It can be found by calculations, using the
same method as for finding CAppM but at full scale Re. If CAppM is determined by testing with
and without appendages, then CAppS should be obtained from extrapolation of CAppM using an
acceptable friction line.

rev. March 25th 2014 pag. 29 di 100


Marco Ferrando Motor Yacht Design

CA is the model-ship correlation allowance.


The full scale ship resistance is then
RTS  1
2 S VS2 S 0  2 CTS

Form factor
The use of the 1978 powering performance procedure implies the use of a form factor k. Par-
ticular problems arise with estimates of (1+k) for HSMVs as low speed tests are not normally relia-
ble or sufficient. Many HSMVs employ transom sterns, leading to a confused flow aft of the tran-
som at low speeds and wetted surface area generally changes with speed, resulting in a change in
true (1+k) with speed. For this reason it is currently recommended that, for consistency and for the
time being, form factors for HSMVs with transom sterns continue to be assumed (1+k) = 1.0. With
respect to form factors, SWATH, which is normally not really a high speed ship, and where the
form factors can be calculated, is an exception.

Model-Ship Correlation
The proposed extrapolation method requires an established model - ship correlation for each
type HSMV. It is not possible to give general guidance to what this correlation factor should be, but
is left instead to each facility to establish its own correlation factor. The extrapolation method
adopted should be documented clearly in the test report.

Air Resistance
This is an important area to address for the testing of HSMVs. However, given the differences
in physical characteristics of each facility it is not possible to propose a particular testing method
that will provide identical results in each facility. Factors such as the size of the carriage and perme-
ability of its structure are difficult to quantify but can significantly affect the flow of air around the
model as the carriage travels down the tank.
The speed at which air resistance becomes significant varies with the vehicle type. If it is de-
cided that air resistance is insignificant for a particular HSMV model test, the justification for that
decision should be documented in the test report.
When air resistance is considered to be significant, wind tunnel tests provide the best source of
information for the superstructure since the model can be tested at higher Reynolds number.
Before making air resistance corrections for the model hull it is important to measure the actual
airspeed beneath the carriage, in the area the model will be tested. These measurements can be
made without the model in place if the model cross section is small compared with the cross section
of the air space housing the tank. Air speed measurements should be made over the speed range of
interest with the carriage configured as it will be when tests are conducted. The air speed measure-
ments and physical features of the above-water portion of the model should be well documented in
the test report so that users of the test data can make their own estimates of air effects if they wish.
When estimates of air resistance are made by staff members at the test facility, the method used, in-
cluding details such as frontal cross section area and drag coefficient should be documented in the
test report. Drag coefficients typically range from 0.3- 1.0. Since HSMVs such as planing boats are
extremely sensitive to trim, estimates of the effects of aerodynamic forces on trim should be made
and documented in the same manner as for air resistance.
The recommended method of accounting for aerodynamic effects on trim, which are not prop-
erly taken into account on the model, is to calculate the difference in bow-up or bow-down moment
between the model and full-scale vehicle by assuming centres of aerodynamic pressure and hydro-
dynamic pressure. These forces are then balanced against the towing force and the resulting mo-
ment converted to an effective shift in longitudinal centre of gravity.

rev. March 25th 2014 pag. 30 di 100


Marco Ferrando Motor Yacht Design

Appendage Effects
It is important to make adequate corrections for appendage effects on HSMV model test results.
Two methods are commonly used to account for appendage effects:
(i) Testing the bare hull and then separately accounting for the lift and drag of individual com-
ponents using analytical methods. This method doesn’t account for hull-appendage interac-
tion effects, and should only be used in case only a very small model can be tested.
(ii) Testing the hull with and without appendages and extrapolating the values based on the local
Reynolds number of each component.
(iii) A less time-consuming, but also less accurate method is to test the hull with appendages
only, and then to calculate the scale effect of the appendages (CAppM - CAppS) by considering
the local Re of each appendage.
Testing both with and without appendages has the advantage of providing more information for
extrapolating the test data using different methods. Trim moments caused by appendage forces not
correctly represented in the experiment should be accounted for using equivalent shifts in centre of
gravity location and displacement. If these corrections are made after the tests are completed, the
results can be obtained by interpolating between results from tests with different centre of gravity
locations.

Wetted Area Estimation


In cases where the wetted surface area varies significantly with speed, which is quite frequent
with HSMVs, the running wetted surface area (WSA) should be estimated for each different speed.
Possible methods include:
 visual observations from outside the model
 visual observations from inside the model
 above water photography or video
 underwater photography or video
 insoluble paint techniques
 water soluble paint techniques
 electrical wetting probes
Surface tension may have an effect on WSA, as discussed in some detail in the proceedings of
the 18th ITTC (1987). Surface tension leads to a different form of spray between model and full
scale, the model spray appearing like a sheet of water rather than droplets as at full scale. For this
reason, separation of the spray sheet at model scale is delayed and the WSA tends to become rela-
tively larger with decreasing model size and model speed. Minimisation of scale effects due to sur-
face tension can be achieved by using larger models, higher speeds, and the fitting of model spray
rails which correctly simulate full scale rails and which can aid to the correct determination of
WSA.
When making estimates of WSA and wetted length, a distinction is made between the area cov-
ered by spray and that covered by solid water. It is common practice to disregard the viscous drag
of spray-covered areas and to account for only the viscous drag of the area wetted by solid flow.
This practice is questionable but the flow in the spray region is extremely complex and no alterna-
tive practices are known.

XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX

1.2.4 Methods for the resistance prevision


This chapter will deal with the most common methods used for the hull resistance prevision
without turning to the towing tank tests.
Even if it is universally recognized that, to determine the hull resistance the towing tank test
still remain the best and more reliable method, the first steps of a hull design are carried out using

rev. March 25th 2014 pag. 31 di 100


Marco Ferrando Motor Yacht Design

not equally precise systems but which have the advantage to give a reasonably precise estimate of
the resistance without taking long times and the high costs of towing tank tests.
In the design initial steps, indeed, it is necessary to evaluate the resistance characteristics of
different possible solutions and later the design is more and more retouched and corrected as more
precise data are available; this method involves the execution of several resistance estimates which
must be carried out in short times and which, even when the time factor is not a problem, would
lead to prohibitive financing charges if done using towing tank tests.
The towing tank tests are generally carried out in the final design phase, when the solution to
take is well identified, while in the previous steps we make use of other systems for the resistance
prevision.
These prevision systems can be based on tests in the towing tank related to similar hulls, sys-
tematic series on hull families, statistical methods, and numerical methods with a theoretical or
semi-empirical nature.

1.2.4.1 Systematic Series

1.2.4.1.1 The EMB 50 Series


The EMB 50 series collects the results of towing tank tests carried out on 20 models, derived
by a single parent hull, designed at the “United States Experimental Model Basin” and originally
called U.S.E.M.B. 50 Series. The original work has been published on October 1941, but it has been
later revised and published on March 1949 [1.2.10.1.10].
The models were 40 inches long and they had been built so that to have values of   0.01L 
3

equal to 40, 80, 120, 160 and values of B/T equal to 4, 6, 8, 11, 15. Figure 1.25 displays the dimen-
sions of the series parent hull, having   0.01L  equal to 110 e B/T equal to 5.3; Figure 1.26
3

shows a reasonably accurate reproduction, of the parent hull lines plan.


Table 1.2-3 displays the dimensions of all models of the series, while Table 1.2-4 shows the
coefficients which were used to obtain the geometry of the different series models from the parent
hull.
For all models the towing point has been located at half length (corresponding to station 5) and
at ½ inch over the design waterline. The tests have been carried out at the design displacement, and
at two 10% and 20% higher displacements than the design one.

rev. March 25th 2014 pag. 32 di 100


Marco Ferrando Motor Yacht Design

Figure 1.25: EMB 50 Series, particulars and dimensions of the parent hull form

rev. March 25th 2014 pag. 33 di 100


Marco Ferrando Motor Yacht Design

Figure 1.26: EMB 50 Series – approximate lines plan of the parent hull

rev. March 25th 2014 pag. 34 di 100


Marco Ferrando Motor Yacht Design

Table 1.2-3: EMB 50 Series – particulars and dimensions of the models


 3
Displace-
 L 
  ment B/H 4 6 8 11 15
 100 
pounds

Model 2727 2728 2729 2730 2731


40 3.24 Beam 4.160 5.827 6.725 7.890 9.21
Draft 1.190 0.971 0.841 0.717 0.61
Model 2732 2733 2734 2735 2736
80 6.48 Beam 6.730 8.240 9.518 5160 13.040
Draft 1.682 1.373 1.190 1.014 0.869
Model 2737 2738 2739 2740 2741
120 9.73 Beam 8.240 10.100 11.650 13.660 15.960
Draft 2.060 1.683 1.457 1.242 1.064
Model 2742 2743 2744 2745 2746
160 12.97 Beam 9.515 11.660 13.60 15.780 18.440
Draft 2.379 1.943 1.683 1.435 1.229
Values of beam and draft in inches. Beam of LWL, and draft to the rabbet. Model Length,
40 inches.

Table 1.2-4: EMB 50 Series – variations with respect to the parent hull
Enlargement Enlargement
Model Model
Beam Draught Beam Draught
2727 0.4444 0.5765 2737 0.7694 0.9989
2728 0.5441 0.4705 2738 0,9431 0.8160
2729 0.6280 0.4075 2739 l.0880 0.7063
2730 0.7365 0.3475 2740 1.2760 0.6022
2731 0.8600 0.2976 2741 1.4900 0.5158
2732 0.6282 0.8160 2742 0.8884 1.1535
2733 0.7695 0.6655 2743 1.0884 0.9420
2734 0.8887 0.5770 2744 1.2570 0.81 59
2735 1.0420 0.4916 2745 1.4740 0.6958
2736 1.2177 0.4213 2746 1.7220 0.5958

The test results have been given by means of contours representing RTM/M as a function of
B/T e   0.01L  with constant static trim angle  and relative speed V L ; an example is shown
3

in Figure 1.27.

rev. March 25th 2014 pag. 35 di 100


Marco Ferrando Motor Yacht Design

Figure 1.27: EMB 50 Series – Example of Resistance Diagram


This kind of representation offers the advantage of not being anchored to a particular formula-
tion of the friction coefficient, but it leaves the user free to chose the most suitable formula.
To transfer the resistance data in full scale the authors supply suitable graphs, whose an exam-
ple is shown in Figure 1.28, displaying contours representing the running model wetted surface,
them also as a function of B/T and   0.01L  with constant static trim angle  and relative speed
3

V L , through which it is possible to go back to the model friction resistance and then to the re-
siduary resistance.

Figure 1.28: EMB 50 Series – Example of Wetted Surface Diagram

Moreover graphs have been supplied displaying contours representing the value of the running
trim angle, them also as a function of B/T and   0.01L  with constant static trim angle  and rela-
3

tive speed V L . An example of these graphs is shown in Figure 1.29.

rev. March 25th 2014 pag. 36 di 100


Marco Ferrando Motor Yacht Design

Figure 1.29: EMB 50 Series – Example of Dynamic Trim Diagram

Finally graphs have been supplied displaying, always as a function of B/T and   0.01L  at
3

constant static trim angle , the longitudinal position of the centre of gravity as shown in Figure
1.30.

Figure 1.30: EMB 50 Series – Example of Center of Gravity position Diagram

The EMB50 series represents the first example of systematic series of planing hulls and it has
been useful for a period of time; in spite of it today its importance is secondary since shows a series
of disadvantages which limit its usefulness:
 For some test conditions the flow regime was not completely turbulent
 The parameters used in the series are typical of the displacing hulls, but not much suit-
able in case of planing hulls
 The hull form became obsolete for that concerns the design aesthetic side.

rev. March 25th 2014 pag. 37 di 100


Marco Ferrando Motor Yacht Design

1.2.4.1.2 Series 62
The 62 Series [1.2.10.1.11] has been developed in the United States at the David Taylor Model
Basin to obtain more modern hull shapes with respect to the 50 Series ones. The preliminary work,
aiming to detect a hull to use as the series parent hull, consisted in the analysis of the tests per-
formed before on planing hulls and led to identify the following characteristics of the parent hull:
 considerable deadrise angle at stern,
 constant deadrise angle in the hull aft part, so that to obtain a non twisted bottom pres-
sure area at high speed,
 tapered stern, in order to have a breadth at stern of about 65% of the maximum chine
breadth,
 convex bow section.
The result of this preliminary work led to identify a hull shape which resulted the best, in terms
of resistance, with respect to the hull tested before care of that institution. The shape of this hull is
represented by its body plan displayed in Figure 1.31.

Figure 1.31: 62 Series – preliminary parent hull


A characteristic of the so identified hull shape was that to have an almost completely develop-
able surface. Therefore it was decided to slightly modify the shape so that to obtain a hull with
wholly developable surface. The result of this last modification had been the parent hull of the 65
series, whose lines plan is represented in Figure 1.32.
The series consists of 5 models, which have been described using the symbols introduced in
paragraph 1.2.1. The authors considered as fundamental parameters for the planing hull resistance:
the ratio length on width, the ratio among the hull dimensions and its weight as well as the centre of
gravity longitudinal position. Consequently the parameters LP BPX , AP  2 3 and the distance of CG
from the AP centroid expressed in percentage of LP have been systematically varied. Table 1.2-5
displays the main characteristics of the series models.
It should be noted that the extreme models of the series are out of the generally accepted pro-
portion range for the crafts, but one of the systematic series purpose is that to investigate even be-
yond the usual.
The lines plans of the five hulls are practically identical, excepting for some slight changes in
the aft side of the hulls having LP BPX equal to 2.00 and 3.06. This results from the fact that the
other four hulls of the series have been derived from the parent hull maintaining the shape of the or-
dinates and modifying the interval between the ordinates and the body plan dimensions (then in af-
finity) to obtain the different values of the ratio LP BPX .

rev. March 25th 2014 pag. 38 di 100


Marco Ferrando Motor Yacht Design

Figure 1.32: 62 Series – parent hull Lines Plan

rev. March 25th 2014 pag. 39 di 100


Marco Ferrando Motor Yacht Design

Figure 1.33: 62 Series – example of presentation of results


The hull having LP BPX  3.06 has proportions similar to those of the smaller pleasure craft;
they are generally powered by outboard or stern drives, this brings the centre of gravity in quite
back position and the stern shape consequent to the change scheme above illustrated would have
given rise to a too advanced pressure centre. To get round this situation the stern of the two hulls
having the lowest values of LP BPX has been arbitrarily enlarged.
All the models have been equipped with spray strips and have been tested in a speed range cor-
responding to volumetric Froude numbers interval from about 2.0 up to 6.0. The values of AP  2 3
achieved during the tests have been 5.5, 7.0 and 8.5. In the lower part of the speed range tests have
been carried out with AP  2 3 = 4. As concerns the CG position, even this one has been systemati-
cally changed placing it at 0, 4, 8 and 12 % LP astern of the centroid of AP. Each model has been
tested in 16 different conditions for a total of 80 tests.

rev. March 25th 2014 pag. 40 di 100


Marco Ferrando Motor Yacht Design

Table 1.2-5: 62 Series – models’ particulars


Model 4665 4666 4667-1 4668 4669
2
AP ft 6.469 9.715 12.800 9.518 7.479
LP ft 3.912 5.987 8.000 8.000 8.000
BPA ft 1.654 1.623 1.600 1.190 0.935
BPX ft 1.956 1.956 1.956 1.455 1.143
BPT ft 1.565 1.386 1.250 0.934 0.734
LP / BPA - 2.365 3.690 5.000 6.720 8.560
LP / BPX - 2.000 3.060 4.090 5.500 7.000
BPX / BPA - 1.1800 1.210 1.220 1.220 1.220
BPT / BPX - 0.800 0.710 0.640 0.640 0.640
Centroid of % LP
47.500 48.200 48.800 48.800 48.800
AP from transom
iE deg 58.000 49.000 46.000 39.000 37.000

Moreover each model has been tested in the standard condition established by David Taylor
Model Basin for planing hulls, providing AP  2 3 = 7 and the center of gravity at 6 %LP astern of
the centroid of AP. This standard condition is shown in [1.2.10.1.12].
The series supplies the test conditions and the tests results in tabular format, so that to allow
the most general possible use of the results. Moreover curves have been supplied of the ratio total
resistance on displacement of the hulls brought to displacements of 10˙000 and 100˙000 lb. An ex-
ample of the result representations is shown in Figure 1.33.

1.2.4.1.3 Dutch Series 62


In the period of time following the publication of the 62 Series the use of planing craft in ex-
posed sea areas increased considerably and it has been extended to pilot’s boats, coast patrolling
boats, working boats for the off-shore industry and small military boats. For all this kinds of craft
seakeeping is very important and, notoriously, to improve the seakeeping of a planing hull it is nec-
essary to increase the deadrise angle, but this remarkably affects the resistance.
The Dutch 62 series [1.2.10.1.13], published in 1982 by Keuning and Gerritsma, has been de-
veloped with the purpose to quantify the influence of the deadrise angle on the resistance. At this
purpose 5 models have been conceived similar to those used for the 62 series in all the details ex-
cept for the deadrise angle which has been brought from 12° to 25°.
These models guaranteed the same variations of the geometric parameters used in the 62 series,
even if the original test speed range 0.5  Fr  6.0 has been reduced to 0.75  Fr  3.00 because
of the limitations of the test facility.
Table 1.2-6 displays the comparison among the main geometric characteristics of the parent
hulls of the two series.
As we have said the parent hull of the Dutch 62 Series has been obtained from the 62 Series
parent hull; in this process a remarkable effort has been done so that the new hull would result the
most possible similar to the 62 Series one in order to guarantee that the differences in resistance
values could be undoubtedly be ascribed to the deadrise angle variation.
The results of this work have been:
 the chine length LP has been kept identical (scale ratio 1:1.624)

rev. March 25th 2014 pag. 41 di 100


Marco Ferrando Motor Yacht Design

 the average chine width BPA and the vertical projection of the chine have been main-
tained identical on scale
 the deck vertical projection has been maintained identical on scale
 the keel line has been kept identical on scale up to the ordinate 16 and from here up to
the bow it has been modified to obtain the same length on scale
 the transom inclination has been kept identical
 the transom  angle has been brought from 12.5° to 25.0° and the portion of the hull
length with constant  has been kept identical
 the hull of the new parent hull is wholly composed by developable surfaces as the par-
ent hull of the 62 series.
Figure 1.34 displays the superposed body plans of the 62 and Dutch 62 parent hulls, while Fig-
ure 1.35 shows the lines plan of the Dutch 62 Series parent hull.

Table 1.2-6: Parent hull particulars of 62 and Dutch 62 Series


Model 4667-1 188
 ° 12.5 25.0
LP m 2.436 1.500
BPA m 0.487 0.300
BPX m 0.596 0.367
BPT m 0.381 0.235
LP / BPA - 5.000 5.000
LP / BPX - 4.090 4.087
BPX / BPA - 1.220 1.220
BPT / BPX - 0.640 0.640
Centroid % LP
48.800 48.800
of AP from transom

Figure 1.34: Comparison of parent hull body plans of 62 and Dutch 62 series

rev. March 25th 2014 pag. 42 di 100


Marco Ferrando Motor Yacht Design

Figure 1.35: Dutch 62 Series parent hull Lines Plan

rev. March 25th 2014 pag. 43 di 100


Marco Ferrando Motor Yacht Design

Table 1.2-7: Dutch 62 Series – models’ particulars


Model 186 187 188 189 190
2
AP m 0.42967 0.42770 0.45000 0.33470 0.26280
LP m 1.0000 1.2500 1.5000 1.5000 1.5000
BPA m 0.42967 0.34216 0.30000 0.22300 0.17520
BPX m 0.5000 0.4080 0.3670 0.2730 0.2140
BPT m 0.4000 0.2900 0.2350 0.1750 0.1374
LP / BPA - 2.3720 3.6530 5.000 6.7260 8.5600
LP / BPX - 2.0000 3.0640 4.0870 5.4940 7.0100
BPX / BPA - 1.1637 1.1920 1.2200 1.2200 1.2200
BPT / BPX - 0.800 0.7108 0.6400 0.6400 0.6420
Centroid % LP
47.1130 47.8792 48.800 48.800 48.800
of AP from transom

Even in this case the hulls having the lowest values of the LP BPX ratio have been modified, in
the stern region, with respect to the shape which would be supplied by the by affinity process of
transformation. The details of the Dutch 62 series 5 models are displayed in Table 1.2-7.

For as concern the result presentation it was tried to maximally favor the possibility of compar-
ison with the 62 series. The Shoenherr friction resistance coefficient has been used with CA=0 and
the curves of the ratio resistance on displacement of the hulls brought to 45˙000 and 450˙000 N dis-
placements have been supplied. The series supplies the test conditions and the tests results in tabular
format, so that to allow the most general possible use of the results.
In 1993 an extension of the series has been published concerning four models having =30°
[1.2.10.1.14]. The geometric details of these last are detailed in Table 1.2-8.

Table 1.2-8: Dutch 62 Series – particulars of models with  = 30°


AP m2 0.3843 0.4499 0.3346 0.2627
LP m 1.250 1.500 1.500 1.500
BPA m 0.300 0.300 0.223 0.175
BPX m 0.367 0.367 0.273 0.214
BPT m 0.260 0.235 0.175 0.137
LP / BPA - 4.170 5.000 6.726 8.571
LP / BPX - 3.410 4.090 5.500 7.000
BPX / BPA - 1.220 1.220 1.220 1.220
BPT / BPX - 0.710 0.640 0.640 0.640
Centroid of AP % LP from transom 47.90 48.80 48.60 48.60

Even in this case the parent hull shown in Figure 1.36, has been obtained from the 62 Series
parent hull, with the same criteria used to obtain the hull of the 188 model having  = 25°. Like-
wise, the hull having the smaller value of LP / BPX has been modified in the stern region with respect
to the shape that it would be obtained with the transformation in affinity of the parent hull.

rev. March 25th 2014 pag. 44 di 100


Marco Ferrando Motor Yacht Design

Figure 1.36: Dutch 62 Series  = 30° parent hull Lines Plan

rev. March 25th 2014 pag. 45 di 100


Marco Ferrando Motor Yacht Design

Figure 1.37 displays the superimposed body plans of the 62 and Dutch 62 series having T
equal to 12.5, 25 and 30° respectively.

Figure 1.37: Comparison of parent hull body plans of 62 and Dutch 62 series with B = 25° and 30°

1.2.4.1.4 Series 65
The 65 Series [1.2.10.1.15] has been developed in the first 70s in the United States, with the
purpose to find useful data for the hydrofoil study. The results have been later re-elaborated
[1.2.10.1.16] so that to allow the use also for planing hulls.
The series consists of 7 models, developed for hydrofoils with airplane configuration foils and
called 65 A Series and other 9 models for hydrofoils with foils in canard configuration and called
65 B Series. For each of the configurations a systematic variation has been performed of the ratios
L/B, B/T as well as of the  angle through the transformation by affinity of the parent hull.

Table 1.2-9: 65 Series – models characteristics

rev. March 25th 2014 pag. 46 di 100


Marco Ferrando Motor Yacht Design

Figure 1.38: 65A Series - model 4966-1

rev. March 25th 2014 pag. 47 di 100


Marco Ferrando Motor Yacht Design

Figure 1.39: 65B Series – model 5184

rev. March 25th 2014 pag. 48 di 100


Marco Ferrando Motor Yacht Design

Table 1.2-9 displays the main characteristics of the series models, while Figure 1.38 and Figure
1.39 display the lines plans of the subseries A and B models.
The test results have been presented in [1.2.10.1.16], in terms of RT/ ratio as a function of Fr
relatively to a displacement of 100000 lb (45.36 t). An example of this result presentation is shown
in Figure 1.40 for the 65A Series and in Figure 1.41 for the 65B Series.

Figure 1.40: 65A Series – example of results presentation

rev. March 25th 2014 pag. 49 di 100


Marco Ferrando Motor Yacht Design

Figure 1.41: 65B Series – example of results presentation

1.2.4.1.5 USCG Series


XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX

1.2.4.2 Statistical or numerical methods


The methods which will be explained in this paragraph are also based on tests performed at the
towing tank but, unlike from the systematic experiences, they take into consideration the highest
number possible of non related hulls to obtain correlations among the performances and the geomet-
ric parameters affecting them.

rev. March 25th 2014 pag. 50 di 100


Marco Ferrando Motor Yacht Design

These correlations are generally obtained using statistical regression methods, therefore their
reliability, strongly depends on the data used to derive them; considering that it is not always possi-
ble to accede to a great number of data, concerning fast boats, such to guarantee a good reliability of
the regressions, these methods are generally used to obtain first approximation previsions or to
compare different design solutions.
Therefore these methods are useful once the user has evaluated their reliability and known their
limits.

1.2.4.2.1 Second VTT method


This method is illustrated in [1.2.10.1.18], and it is known also as VTT method, which is the
abbreviation of the naval research centre of Tietotie in Finland.
The method, which is valid for hard chine hulls working in the range Fr= 1.8  3.3, has been
elaborated analysing, through the regression technique, the test data relevant to 13 model tests two
of them suitably realized.
The analysis through regression has been applied to the towing tank test results expanded to the
standard boat dimension having  = 100000 lb (45.36 t); during the analysis have been identified as
significant the parameters listed hereunder along with their variation range.
L/1/3 = 4.49  6.81

L/B = 2.73  5.43

B/T = 3.75  7.54

AT/AX = 0.43  0.995


where:
 L length of the waterline,
 B maximum width of the waterline,
 T mean draught
  hull volume
 AT immersed transom area,
 AX immersed area of the maximum area section
The equation resulting from the analysis is the following:
 RT  1 3 6

 W    Ai Pi  Fr  Ai Pi Fr  Ai Pi
2

W 100000 lb i 0 i 2 i 4

valid for speeds such to produce Froude numbers, relevant to the still ship hull volume, in-
cluded in the range Fr = l.8  3.3.
The coefficients Ai and parameters Pi values are the following:

•i Ai Pi
0 -0.03546471 1
1 +0.00129099  / T3
2 +0.51603410  1/3/L
3 -0.00010596 (L/T) 2
4 -0.00090300 (L/ 1/3)2
5 +0.00017501 (L/ 1/3)3

rev. March 25th 2014 pag. 51 di 100


Marco Ferrando Motor Yacht Design

6 -0.02784726 (B/L)(AT/AX)

To calculate the RT/W value relevant to the a craft having W  100000 lb, the formula dis-
played hereunder can be used, which is similar to that used by Mercier and Savitsky:

 RT   RT  S Fr2
      CF  CF 100000  CA 
2 3
2
 W corr.  W 100000
in which:
(RT/W)corr. is the value to determine,
(RT/W)100000 is the value supplied by the regression,
CF is the friction resistance coefficient calculated for the concerned hull with the
formula ITTC’57 in which the Reynolds number has the classical expression
Re = VL/;
CA is the hull roughness allowance which can vary between 0.0002 and 0.0008
according to the hull and its surface dimensions, the authors advise to adopt
0.00025,
CF 100000 is the standard ship frictional resistance coefficient calculated with the
ITTC’57 formula in which the number of Reynolds has the following expres-
sion:

L W*
Fr
3 
1

Re 

having indicated with W* the displacement of 100000 lb expressed in coherent units with .
The authors believe that using the Mercier and Savitsky method in the range Fr = l.0  1.8
and that now exposed in the range Fr = l.8  3.3 a good curve of RT = f(Fr) can be obtained in the
whole range Fr = 1.0  3.3 corresponding to the pre-planing and the planing regimes.

1.2.5 Other components of the resistance


Naturally, beside to the bare hull resistance, it is necessary to consider, as for the displacing
hulls, other resistance components; hereunder the main components shall be discussed.

1.2.5.1 Spray Resistance


As we have many times had the occasion to note at the paragraph 1.2.2.1, and particularly in
Figure 1.14 and in Figure 1.15 for as concern the case having =0, the bottom pressure area ends at
the bow where the area concerned by the spray begins.

rev. March 25th 2014 pag. 52 di 100


Marco Ferrando Motor Yacht Design

Figure 1.42: Spray production on a planing hull model

Figure 1.43: Spray on a fishing boat


Even in the case of hull with 0 we observe the spray production, as shown in Figure 1.42
concerning a towing tank test of a planing hull model and in Figure 1.43 regarding an actual boat; in
the first case the spray appears as a water sheet, because during the test the equality of Weber num-
bers of the model and the full scale hull has not been guaranteed.
Since even in the bottom area concerned by the spray we have relative motion between the hull
and the water, it is obvious to think that even this region contributes to the hull resistance.
The first problem to face is then to determinate the area wetted by the spray, delimited by the
Spray Root line, which is often confused with the stagnation line given that these two lines are very
close at the trim angles typical of the planing hulls, and by the Spray Edge line. The Spray Area is
shown in Figure 1.44.

rev. March 25th 2014 pag. 53 di 100


Marco Ferrando Motor Yacht Design

Figure 1.44: Pressure area and Spray area on a planing hull


A first formal discussion of this problem is due to Savitsky [1.2.10.1.1]. Recently the same au-
thor has published a revised approach to the topic [1.2.10.1.19].
The stagnation line position with respect to the hull line is identified by the angle A in the bot-
tom plane and, on the  plane passing through the keel line and rotated of  with respect to the hull
bottom, by the angle  between the projection of the stagnation line and the keel line, as shown in
Figure 1.45.

Figure 1.45: Position of the Stagnation Line with respect to the Keel Line

Observing the mentioned figure we note that:


b
2
cos 
tan  
LK  LC

Remembering that LK  LC  L1 , equation (11) allows us to write:

 tan  1
tan   (47)
2 tan  cos 
As concern  we have:

rev. March 25th 2014 pag. 54 di 100


Marco Ferrando Motor Yacht Design

 tan 
tan   (48)
2 tan 
tan 
tan   (49)
cos 
Having identified with C the length of the stagnation line and with c that of its projection on 
and again with reference to Figure 1.45 we obtain:
b2
c (50)
sin 
and also:
b2
cos 
C (51)
sin 

Figure 1.46: Position of the Spray Edge Line with respect to the Keel Line
Analogously, with reference to Figure 1.46, also the angle that the Spray Edge forms with the
Keel Line can be defined;  on the plane  and Θ on the hull bottom, and we obtain:
tan 
tan   (52)
cos 

rev. March 25th 2014 pag. 55 di 100


Marco Ferrando Motor Yacht Design

As concern the flow direction in the spray area it has been observed that the direction of the
fluid leaving the hull is practically the reflection, with respect to the stagnation line of the entrance
direction, as shown in Figure 1.47.

Figure 1.47: Projection of the spray area on the longitudinal plane


The angle Θ that the spray edge forms with the Keel Line is then related with the angle with
which the spray is practically reflected by the stagnation line. The original Savitsky paper
[1.2.10.1.1] supplies a very complex procedure for the Θ determination while the most recent work
of Savitsky [1.2.10.1.10] considers acceptable the approximation of the flow reflection on the stag-
nation line.
Reasoning, by commodity, on the projection of the phenomenon on the  plane, represented in
Figure 1.48, we can obtain:
  2 (53)

Figure 1.48: Projection on plane  of the spray area


As a consequence we can affirm that the area wetted by the spray, represented by the shaded
area in Figure 1.48, is an isosceles triangle having as base the stagnation line and angle width at the
base equal to .
Identifying with AS the projection on the plane  of the spray wetted hull total surface, we can
affirm that the value of AS is given by the double of the shaded area in Figure 1.48. The latter can
be obtained once known the height h relevant to the isosceles triangle base; remembering that the
stagnation line length had been identified with c we obtain:
c
h tan 
2
and using equation (50) we obtain:
b tan 
h
4 sin 

rev. March 25th 2014 pag. 56 di 100


Marco Ferrando Motor Yacht Design

Using then equation (48) the previous relation can be rewritten in the form:
 b tan 
h
8sin  tan 
which allows to obtain:
As ch  b 2 tan 
 
2 2 32sin 2  tan 
from which we finally obtain:
 b2 tan 
As 
16sin  tan 
2

Using then the equation (48), rewritten in the form:


2
tan   tan  tan 

we can simplify the formula which expresses AS obtaining:
b2
As  (54)
4sin 2
The spray wetted hull area surface value AS is obtained from its projection AS value through
the expression:
As
As 
cos 
leading to:
b2
As  (55)
4sin 2 cos 
It is well worth to remember that this approach is extremely simplified with respect to that pre-
sented by Savitsky in his original work, considering that it is assumed that the spray is reflected by
the spray root line instead of the stagnation line. This approximation is therefore acceptable since
the two lines are modestly distant and this produces a slight overestimation of the spray wetted hull
surface; this overestimation is undoubtedly included in the general method approximation.
Equation (55) holds for realistic values of , that is for 0    35 . Remembering indeed
equation (48) we can observe that the angle  becomes 90° if =0 that is that the stagnation line is
orthogonal to the keel line; in these conditions the equation (55) gives AS   . This can be ex-
plained considering that the spray is completely reflected forward and, as consequence, the area
wetted by it is indeterminate.
As concern the shape of the area wetted by the spray and the spray speed direction it is possible
to observe that they are governed by the angle  which, according to the equation (48) is a function
of  and .
After having obtained the geometric definition of the hull area concerned by the spray, it is
possible to obtain the magnitude of the force FS due to the relative motion between the spray and
the hull.

rev. March 25th 2014 pag. 57 di 100


Marco Ferrando Motor Yacht Design

Figure 1.49: Spray resistance as a function of 

Assuming the spray speed equal to the speed V of the undisturbed current, the resulting FS of
the spray viscous actions can be written in the form:
1
FS  CF  ASV 2
2
and, remembering equation (55), transformed in the:

1 b2
FS  CF V 2 (56)
2 4sin 2 cos 
The resulting force FS lays in the hull bottom and forms the angle  with the Keel, being paral-
lel to the Spray Edge line. The spray resistance RS can be defined as the component of FS on a par-
allel plane to the undisturbed water plane and directed afterward; it can be obtained then through the
formula:
RS  FS cos  cos  (57)

Since the trim angle of the typical planing hulls is small (usually lower than 6°) we have cos 
1, therefore this term can be neglected without performing great approximations. Equation (57) can
so be rewritten in the form:

rev. March 25th 2014 pag. 58 di 100


Marco Ferrando Motor Yacht Design

1 b 2 cos 
RS  CF V 2 (58)
2 4sin 2 cos 
Remembering equations (52), (53) e (48) we note that as the  decreases the angles  and 
decrease at their turn producing an increase of RS, as shown in Figure 1.49.
A combination exists of angles  and  giving rise to a resulting force FS orthogonal to the
Keel Line. In this case, in spite it exists a relative motion between spray and keel, the resistance RS
is null.
To simplify equation (58), Savitsky proposed to introduce the quantity , which we could de-
fine equivalent spray wetted length, so defined:
cos 
  (59)
4sin 2 cos 
Using this quantity, equation (58) can be re-formulated as follows:
1
RS  CF  b 2 V 2 (60)
2

Figure 1.50: Equivalent spray wetted length


As it can be deduced by the previous discussion the quantity  is a function of  and  and its
trend is shown in Figure 1.50. Since Savitsky approach is only valid for the cases in which  is 
90°, that is for the cases in which the spray speed does not have components in the direction of the
hull speed V, it supplies only values of RS  0 and that, considering equation (60), can only occur
for   0.
As concern the friction resistance coefficient CF to use for the spray resistance calculation,
since in full scale the motion in the spray region can be undoubtedly assumed as in turbulent re-
gime, it is possible to use both the ATTC’47 and the ITTC’57 formulations.
In any case the Reynolds number to use for the coefficient calculation must be obtained using
the mean spray wetted length LWS defined as follows:

rev. March 25th 2014 pag. 59 di 100


Marco Ferrando Motor Yacht Design

b
1 2
LWS  (61)
2 sin 2 cos 
which gives a value equal to half of the spray edge line length being, as we have seen, the direction
of the spray speed with respect to the hull parallel to the spray edge line.

1.2.5.2 Air Resistance


The air resistance can be, for planing hulls, a significant component of the total resistance; also
in this case the air resistance can be evaluated through the formula:
RAA   12 air ATV 2 (62)

In the case of planing hulls, since the hull fore side is completely emerged, the transversal pro-
jection area of the craft will include the whole hull.
As concern the coefficient , Blount [1.2.10.1.20] suggests to assume a value included in the
range 0.5    0.8, while Savitsky [1.2.10.1.19] reports that the experience of the Davidson Labo-
ratory suggests that, for normal bow shapes we usually have   0.7.

1.2.5.3 Appendage Resistance


For planing hulls, the hull appendages are usually the rudder and the shaft lines, one of which
is represented in Figure 1.51.

Figure 1.51: Typical shaft line configuration

An approximated discussion of planing hull appendages resistance has been published by


Blount and Fox [1.2.10.1.9], who have proposed a simple formula supplying a value of the append-
age resistance for a planing hull with conventional double shaft line configuration.
 1 
RAPP  RBH   1 (63)
 A 
where with RAPP is denoted the appendage resistance, with RBH the bare hull resistance and with A
a coefficient given by the following equation:
1
A  (64)
0.005 Fr2  1.05

rev. March 25th 2014 pag. 60 di 100


Marco Ferrando Motor Yacht Design

This expression is based on a collection of model tests performed with and without appendages
for traditional twin-shaft configuration (shafts, struts, rudders): it can be used as first estimate of the
appendage resistance during the design preliminary phase, when the appendage geometry, required
by the more complex formulas for the resistance evaluation, has not yet been completely defined.
It should be noted that this method evaluates only the resistance increase produced by the ap-
pendages, completely neglecting the effect that their presence can produce on the boat trim.

Hadler [1.2.10.1.21] proposed formulas for the main appendage types of the planing hulls. In
this case each of the considered appendages is treated separately and a formula is proposed for the
evaluation of its effect on the boat.
Skeg
The skeg is a hull portion that sometimes is installed under very flat hulls to assure a suitable
value of the lateral plane area and improve the course stability qualities of the boat, as shown in
Figure 1.52.
Strictly speaking the skeg is a part of the hull and, normally, it is not considered a hull append-
age, but if planing hull resistance is evaluated with the Savitsky method, this last does not consider
the presence of an eventual skeg and then its resistance must be separately added.

Figure 1.52: Example of hull with Skeg


The skeg resistance can be evaluated with the following formula:
1
DSK  CF  SSKVM2
2
where:
 CF  friction resistance coefficient, based on the number of Reynolds calculated with
the average skeg length.
 SSK  skeg wetted surface
 VM  mean speed on the hull bottom

As concern the properly said appendages they can be divided in cylindrical-shaped components
and those having a wing shape. With reference to Figure 1.53, the components A, B and C are of
the cylindrical type, while D and E have a wing shape.

rev. March 25th 2014 pag. 61 di 100


Marco Ferrando Motor Yacht Design

Figure 1.53: Different types of appendages


Cylindrical components
The appendage cylindrical components, typically shafts, stern tubes etc, produce lift and resis-
tance. These forces can be determined assuming that the flow investing them is parallel to the hull
bottom.
The Reynolds number of these components must be calculated using the diameter instead of
the length. Assuming that it is:
Vd
Re   5.5105

resistance and lift of the cylindrical component can be evaluated with the following equations:

1
D
2

 ldV 2 1.1sin 3   CF  (65)

1
L
2

 ldV 2 1.1sin 2 cos   (66)

where:
l = component length
d = component diameter
 = angle that the component axis forms with the keel line.
As concern the resistance, equation (65) can be rewritten as follows:
1 1
D  1.1  ld sin  V sin    CF  2 rlV 2
2

2 2
underlining as the resistance is composed by two components, represented by the two terms of the
second equation member.
The first of them represents a pressure resistance. Indeed the quantity V sin  is the component
of the speed vector normally directed to the considered appendages while ld sin  represents the ap-
pendage projection in normal direction to the speed vector; the pressure resistance coefficient is rep-
resented by the quantity 1.1.
The second terms quantifies the appendage friction resistance.

rev. March 25th 2014 pag. 62 di 100


Marco Ferrando Motor Yacht Design

As concern the lift, equation (66) can be reformulated as follows:


1
L  1.1  ld cos  V sin  
2

2
where V sin  represents the speed vector component normal to where V sin  represents the normal
speed vector component at the appendage while ld cos  represents the appendage projection in di-
rection parallel to the speed vector; the lift coefficient is represented by the quantity 1.1 the append-
age while ld cos  represents the appendage projection in direction parallel to the speed vector; the
lift coefficient is represented by the quantity 1.1.

Wing shaped components


The wing shape component resistance, provided that they are not ventilated, can be expressed
by:

1  t t 
4

D  SV 2CF 1  2  60    (67)
2  c  c  

where:
S = appendage wetted surface
CF = Schoenherr friction resistance coefficient ( if Re  5105 )
t
= Thickness/ chord ratio of the appendage foil section.
c
Strut palms
In some cases, it can occurs that the struts are fixed to the hull through projecting palms, as
shown in Figure 1.54; even if immersed in the boundary layer these elements can give rise to re-
markable resistance.

Figure 1.54: Example of strut palm

rev. March 25th 2014 pag. 63 di 100


Marco Ferrando Motor Yacht Design

The strut palm resistance can be expressed by:

h 1
D  0.75CDP 3  whV 2 (68)
 2
where:
h = plate height under the hull surface
w = plate width, in normal direction to the flow
 = boundary layer thickness, to assume  0.016 X P
XP = distance between the stagnation line and the plate leading edge.

If the plate is rectangular with the rounded edges one can assume:
CDP  0.65

1.2.6 Porpoising
The Porpoising phenomenon consists in a boat oscillation in the longitudinal plane consisting
of a combination of heave and pitch with constant or increasing amplitude; this movement mani-
fests in planing conditions even when the boat runs with a perfectly calm sea and it is the result of
an instability in the longitudinal plane.
Conventionally we think to be in presence of porpoising if the oscillation amplitude is greater
or equal to one degree.
This phenomenon can lead to boat structural damage because of the very high pressures devel-
oping on the bottom due to the impacts; it can also be the cause of the boat capsizing in longitudinal
direction if the bow meets a wave in correspondence of a low value of the trim angle. This longitu-
dinal unsteadiness has been responsible of many serious accidents.
This phenomenon has been studied by some authors [1.2.10.1.22, 1.2.10.1.23, 1.2.10.1.24] who
detected the main parameters affecting it.

rev. March 25th 2014 pag. 64 di 100


Marco Ferrando Motor Yacht Design

Figure 1.55: Steady and unsteady planing scheme

The Day & Haag experimental work [1.2.10.1.22], more recently confirmed by Celano
[1.2.10.1.24], shows that for a given value of the deadrise angle a specific relation exists linking the
lift coefficient CL to the trim angle  beyond which the instability phenomenon occurs.
In his original work Savitsky [1.2.10.1.1], refers to the Day & Haag work and proposes a
graphic, reproduced in Figure 1.56, with which can be determined if the boat will be subjected to
porpoising.
The use of this figure is obviously limited to the deadrise angles for which the porpoising limit
curves are represented which are 0°, 10° and 20°.

rev. March 25th 2014 pag. 65 di 100


Marco Ferrando Motor Yacht Design

Figure 1.56: Porpoising limits proposed by Savitsky

As we can note, for a given lift coefficient the porpoising occurrence is linked to an excessive
trim angle.
The problem has been faced also by Angeli [1.2.10.1.25] who proposed an empiric formulation
of the critical trim angle, beyond which the porpoising occurs, which considers also the effect of the
non prismatic character of the hull.

0.75
 K 
C   2  (69)
 Fr XYF 
where:
  B      T   X 
0.25

K  106  85  T   1  0.2    
  BPX        Y 
X
X  CG
3
1

B
Y  PX
3
1

C Lβ
F
CL 0

In this formulation the non prismatic hull is considered through the quantities BT and T re-
spectively representing the chine breadth and the deadrise angle in correspondence of the transom.

rev. March 25th 2014 pag. 66 di 100


Marco Ferrando Motor Yacht Design

Moreover we note that, among the parameters ruling the phenomenon, also the longitudinal po-
sition of the centre of gravity and the maximum breadth at the chine have been introduced which
were not provided in the previous studies.
1.2.7 Spray Rails
To prevent the spray flow over the hull, use is made of the so called “Spray Rails” or “Spray
Strips”. [1.2.10.1.26].
A spray strip is a relatively narrow strip, of small cross-section, attached to the hull for the pur-
pose of controlling or diverting spray and reduce the wetted area. Their effect is also to increase the
lift.
The lift increase is due to the fact that the after part of the strips is a surface having almost 0°
deadrise angle and contribute to the lift with a high CL.

Figure 1.57: Transversal view of spray Strips

Spray rails also affect both yaw and roll characteristics of the hull. The build-up of pressure
along the vertical faces of the inner side of a turn will increase the course stability of the hull, while
the increase in incidence on the bottom faces of the depressed side will render the craft less prone to
rolling.
Ideally the boat should make a coordinated turn, as airplanes do. A coordinated turn is one in
which a person's weight is felt to be directly down in the seat, or perpendicular to the deck; to ob-
tain this the craft should bank toward the inner side of the turn so that the tangent of the bank angle
is equal to the number of lateral g's that the boat is pulling in the turn.
With reference to Figure 1.58, and having indicated the gravity acceleration with g, the centrif-
ugal acceleration due to the turn with a and the resultant acceleration with r, it is clear that a coordi-
nated turn will be obtained if the tangent of the angle x of the deck is equal to a/g. In this condition
the resultant acceleration is normal to the deck and people do not need to brace themselves to main-
tain their equilibrium during the turn.
A V hull naturally tends to bank toward the inner side of the turn, but this natural tendency can
be negated by an improper disposition of spray strips that, among the other effects, have the capa-
bility to increase the roll stability.

rev. March 25th 2014 pag. 67 di 100


Marco Ferrando Motor Yacht Design

Figure 1.58: Coordinated turn

According to Renato “Sonny” Levi [1.2.10.1.27], the total 0° deadrise breadth of the spray
rails, which would also include the chine flats, should amount to between 20 and 25 % of the chine
beam aft. This area is usually split up into 2 or 3 rails plus a flat per side, though sometimes up to 4
rails have been applied.

Figure 1.59: Example of multiple spray strips

rev. March 25th 2014 pag. 68 di 100


Marco Ferrando Motor Yacht Design

Spray strips have been positioned in a variety of ways from following the contours of the wa-
terlines to those of the buttocks. Locating them in either of these extreme ways can cause undesira-
ble effects.

Figure 1.60: Longitudinal view of a spray strip

The former will tend to cause tripping (yawing) with poor lift qualities and the latter can induce
wetness since there is little spray suppressing effect. Running the rails parallel to the chine seems on
the whole to give the best results.
Whilst these strakes are indispensable to improve the all-around characteristics of a V hull they
do make the boat hard riding, more stable in her course, and stiffer in banking, and so should not be
overdone.
A further important point is that some deadrise should be worked into the rails in the fore body.
This will deflect the spray aft, reduce the hardness and avoid the possibility of tripping at high
speed.
1.2.8 Trim Control
1.2.8.1 Design options
In his original work [1.2.10.1.1] Savitsky demonstrates that an optimum trim angle exist for
which the ratio RT  has its minimum value; this optimum trim angle is approximately 4 degrees
with slight variations depending on the deadrise angle. It is obvious that the designer should try to
obtain such optimum trim for his vessel at the design speed.
In this paragraph some design options are illustrated that can be used by the designer to control
the trim of his boat during the design phase in order to achieve the optimum trim.
For the sake of simplicity let’s assume that a tentative solution has been identified for a pris-
matic hull form, but its running trim at the design speed is unsatisfactory and needs to be corrected.
Obviously, different options are available either for increasing or decreasing the trim angle.

TRIM INCREASING ACTIONS


Increasing the deadrise angle 
We have seen in equation (20) that the lift coefficient of a planing surface decreases as an in-
verse function of the deadrise angle.
Increasing the deadrise angle will then produce a reduction of the lift and the boat will conse-
quently be more immersed to obtain the equilibrium between the lift and the weight. This greater
immersion will in turn produce an increase of the length of the pressure area toward the bow that
will cause the movement of the center of pressure toward the bow. In this way the boat with the
greater deadrise angle will find its equilibrium position at a higher trim angle.

rev. March 25th 2014 pag. 69 di 100


Marco Ferrando Motor Yacht Design

Increasing the L/B ratio


Imagining a small change of dimensions, one can assume that it can be done without changing
the unit weight. In this perspective let’s assume a small decrease in the unit breadth that will pro-
duce a corresponding increase of the value of the L/B ratio.
To achieve the same amount of pressure area as the wider boat, the narrower one will require a
longer pressure area, that will have a center of pressure located more forward. This will result in a
new equilibrium position at a higher trim angle with respect to the one that characterized the wider
boat.

Tapering
Reducing the hull breadth toward the stern brings forward the center of pressure, resulting in a
higher equilibrium trim angle.

Adding spray strips


Adding spray strips has the effect of increasing the trim angle, as can be observed by the results
of towing tank tests illustrated in [1.2.10.1.26].
This can be explained if we think of the hull region where the spray is located, the bow region.
Here the spray strips deflect the spray away from the hull as illustrated in Figure 1.61.

Figure 1.61: Spray strip action


The hull exerts a force on the current to deflect it and, as a consequence, a corresponding force
is applied to the hull. This force has a vertical component that contributes to the equilibrium of
moments producing an increase of the trim angle with respect to the same hull without the spray
strips.

rev. March 25th 2014 pag. 70 di 100


Marco Ferrando Motor Yacht Design

Using convex sections

Figure 1.62: Effect of using convex sections


Convex sections have, for the greatest part of the length of the pressure area, a greater width;
for this reason the pressure distribution is shifted forward, as illustrated in Figure 1.62. Furthermore
this effect is amplified by the fact that the fore sections have a smaller deadrise angle, comparing
with those of the straight sections, and this produces more lift. As a result the center of pressure is
located more forward resulting in an increased equilibrium trim angle.

Using rocker

Figure 1.63: Effect of using rocker


The shape of the buttocks illustrated in Figure 1.63 is called “Rocker” since it resembles a
rocker-chair.
The use of so curved buttocks affects the local value of the trim angle: the fore portion of the
rocker hull has a greater incidence than the corresponding straight and this, according to equation
(19), gives more lift. On the other hand the aft part of the rocker hull has lower values of the inci-
dence angle producing the opposite effect.

rev. March 25th 2014 pag. 71 di 100


Marco Ferrando Motor Yacht Design

The total result is a pressure distribution that is lower in the aft region and higher in the fore
part, comparing with that of the straight solution, resulting in an increased equilibrium trim angle
due to the more forward location of the center of pressure.

TRIM DECREASING ACTIONS


Decreasing the deadrise angle 
This action has the opposite effect of increasing the deadrise angle; an increase of the lift. The
boat will consequently be less immersed to obtain the equilibrium between the lift and the weight.
This reduced immersion will in turn produce a shortening of the length of the pressure area that will
cause the movement of the center of pressure toward the stern. In this way the boat with the smaller
deadrise angle will find its equilibrium position at a lower trim angle

Reducing the deadrise angle astern


According to equation (20), reducing the deadrise angle toward the stern of the hull, as illus-
trated in Figure 1.64, causes an increase in the lift generated by the after part of the hull.
This will cause the center of pressure to move aft, resulting in a reduced equilibrium trim an-
gle.

Figure 1.64: Deadrise reduction astern

Reducing the L/B ratio


This action has the opposite effect of increasing the L/B ratio.

rev. March 25th 2014 pag. 72 di 100


Marco Ferrando Motor Yacht Design

Using concave sections

Figure 1.65: Effect of using concave sections


Concave sections have, for the greatest part of the length of the pressure area, a smaller width;
for this reason the pressure distribution is shifted afterward, as illustrated in Figure 1.65. Further-
more this effect is amplified by the fact that the fore sections have a higher deadrise angle, compar-
ing with those of the straight sections, and this produces less lift. As a result the center of pressure is
located more aft resulting in a reduced equilibrium trim angle.

Using hook

Figure 1.66: Effect of using hook

The shape of the buttocks illustrated in Figure 1.66 is called “Hook”.


The use of so curved buttocks affects the local value of the trim angle: the fore portion of the
hook hull has a smaller incidence than the corresponding straight and this, according to equation
(19), gives less lift. On the other hand the aft part of the rocker hull has higher values of the inci-
dence angle producing the opposite effect.

rev. March 25th 2014 pag. 73 di 100


Marco Ferrando Motor Yacht Design

The total result is a pressure distribution that is higher in the aft region and lower in the fore
part, comparing with that of the straight solution, resulting in an reduced equilibrium trim angle due
to the more rear location of the center of pressure.

GUIDE FOR THE ALTERATION OF THE BUTTOCKS SHAPE

Figure 1.67: Guide for the choice of buttocks curvature


Aggiungere descrizione geometrica della zona da modificare
XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX

1.2.8.2 Active devices


As we have seen, a vessel is designed, and possibly optimized, for the design speed. If the lat-
ter assumes high values the position of the Center of Gravity should be quite aft in order to ensure a
good trim. Unfortunately this position of the Center of Gravity produces very high trim angles at the
hump speed, causing the hump of the resistance curve to become very pronounced due to the pres-
sure resistance.
To overcome this situation a couple of solution are available, that allow the reduction of the
trim angle if it is so needed. Both are controllable devices that can be used when the need arises and
then stowed in neutral position when they are no more required.

Trim tabs

rev. March 25th 2014 pag. 74 di 100


Marco Ferrando Motor Yacht Design

Trim tabs are in every aspect analogous to that used on airplanes. They are movable surfaces,
usually hinged on the transom, that can be moved to decrease the trim angle of the vessel, their ef-
fect being proportional to the angle of rotation.

Figure 1.68: Example of trim tabs installation


Figure 1.68 gives an example of transom installation; Trim Tabs can also be installed under the
hull flush with the hull bottom, but in this case the installation is much more complicated.
Figure 1.69 illustrates the main components of a trim tab.

Figure 1.69: Parts of a trim tab


The first thorough discussion about trim tabs is due to Brown [1.2.10.1.26], subsequently the
matter was again considered in [1.2.10.1.30].
In both of the mentioned papers equations are given to predict the performances of trim tabs as
a function of the geometric characteristics of the device; these characteristics are defined as follows:

rev. March 25th 2014 pag. 75 di 100


Marco Ferrando Motor Yacht Design

cT Tab chord [m]


T  cT B Non dimensional tab chord
ST Tab span [m] (measured in horizontal plane)
  ST B Non dimensional tab span
 Tab deflection angle [degrees]
LT Tab Lift [N]
DT Tab Drag [N]
MT Tab hydrodynamic moment [Nm] (about the tab trailing edge)

It must be noted that the tab span used into the performance prediction equations is measured
in an horizontal plane, as illustrated in Figure 1.70, and differs from the span showed in Figure 1.69,
the difference being a function of the cosine of the deadrise angle cos  .
Remembering this difference, the performances of the trim tab can be obtained by the follow-
ing equations:
LT
C LT   0.046T (70)
1
2  B 2V 2
DT
CDT   0.0052CLT     (71)
1
2  B 2V 2
MT
CMT   CLT 0.6  T 1     (72)
2 B V
1 3 2

These equations hold in the following range:


0  T  0.1
0    15
0    10
2  CV  7

where  is the non dimensional mean wetted length of the hull.


Looking at equation (70) one can note that the tab lift coefficient is proportional to the product
T , that in turn can be seen as the product of the tab area T times the deflection angle . In
this way the designer can choose to produce lift either by using a big area and a small deflection an-
gle or the opposite.

rev. March 25th 2014 pag. 76 di 100


Marco Ferrando Motor Yacht Design

Figure 1.70: Geometric quantities used for trim tab performance prediction

Equation (71) shows that the most efficient way to generate lift with a trim tab is using a big
area and a small deflection angle, since the trim tab drag is proportional to the trim angle  and the
deflection angle .
As regards equation (72) it can be seen that the lift lever:
0.6B  T B 1   
that can be rewritten in the form:
cT c
B  0.6 B   T B
B B
allows to understand how the lift produced by the trim tab lies on the hull. This is due to the fact
that the pressure contribution of the trim tab affects also the pressure distribution on the hull, as il-
lustrated in Figure 1.71. This places the resultant of the trim tab pressure contribution on the hull ra-
ther than on the trim tab itself, as one could have guessed at first glance.

rev. March 25th 2014 pag. 77 di 100


Marco Ferrando Motor Yacht Design

Figure 1.71: Trim Tab pressure contribution

Interceptors
Interceptors are movable surfaces, usually mounted in a plane parallel to the transom, that can
be lowered into the flow below the hull to decrease the trim angle of the vessel, their effect being
proportional to their protrusion below the hull.

Figure 1.72: Example of interceptors installation

Interceptors, when in protruded position, block the path of the flow along the hull and cause its
stagnation. This stagnation produces a very high pressure peak, though very narrow as illustrated in
Figure 1.74, that generates a bow down moment and reduces the trim of the vessel.

rev. March 25th 2014 pag. 78 di 100


Marco Ferrando Motor Yacht Design

Figure 1.73: Example of protruded interceptor

Figure 1.74: Interceptor pressure contribution

Literature about interceptors is practically absent, save an article by Dawson and Blount
[1.2.10.1.31] in which a way to assess their performances is offered.

rev. March 25th 2014 pag. 79 di 100


Marco Ferrando Motor Yacht Design

Dawson and Blount actually propose a method to calculate the protrusion depth d required to
produce the same trimming effect of an equivalent trim tab.
Once the trim tab dimensions and deflection angle are obtained to achieve the desired effect on
the vessel, the interceptor characteristics for equivalent effect can be calculated. In the proposed
method it is assumed that trim tabs and interceptors have the same span.
The process starts having determined that a couple of trim tabs having span ST, chord cT and
deflected by an angle T provide the adequate trim for the vessel.

Figure 1.75: Interceptor equivalence


The proposed method provide an equation that relates the trim tab deflection angle T and the
equivalent geometric angle for interceptor I, illustrated Figure 1.75, as follows:
 I  0.175 T  0.0154 T2 (73)

At this point the interceptor protrusion d can be calculated by means of the relation
d  cT tan  I (74)

Unfortunately the authors do not mention the interceptor resistance, but a rough calculation can
be done assuming that the protruded interceptor surface is loaded with a linear pressure distribution
starting with the stagnation pressure in correspondence of the hull bottom and ending with the at-
mospheric pressure at the interceptor lower edge.
The result of this crude calculation shows that the interceptor exhibits more or less the same re-
sistance as the trim tab.

1.2.9 Surface piercing propellers


An ever increasing number of recreational powerboats are now equipped with the so-called
Surface Piercing Propellers (SPPs). Notwithstanding their wide diffusion only a little technical lit-
erature on the matter is available. This is probably due to the fact that this kind of propeller was
mainly employed and studied for racing and for naval applications. In both these instances, whoever

rev. March 25th 2014 pag. 80 di 100


Marco Ferrando Motor Yacht Design

has some knowledge on the matter keeps it restricted for obvious reasons. As a consequence, the
peculiarities in the behavior of this kind of propulsion system are not widely known.
An SPP is actually a quite normal screw propeller. Compared with a traditional propulsion sys-
tem its main feature is that it is intended to operate with only its lower half in the water.
This mode of operation can appear very strange to those used to conventional propellers, and
indeed it is. Trying to figure out the reason for such a layout can be misleading if one looks only at
the propeller, forgetting the rest of the boat. Actually, the main reason for adopting SPPs is the con-
siderable reduction in form and frictional resistance that can be obtained by raising half of the pro-
peller out of the water.

Figure 1.76: Typical SPP layout


Figure 1.76 illustrates one of the most common layouts for an SPP. It can be observed that the
unit is placed through the transom of the boat. Assuming that the water breaks out of the transom
along the prolongation of the bottom of the hull, the propeller has only its lower half in the water. In
this way most of the shafting is also out of the water, allowing for a dramatic reduction of append-
age resistance.
Operating partially submerged, the propeller itself does not improve its performances, though
these are not overly spoiled. The boat can generally achieve a greater speed because of the reduced
resistance.
The amount of appendage drag reduction depends upon the speed of the craft. For very high
speeds, i.e. over 35 knots, the appendage resistance for a conventional twin shaft arrangement can
easily exceed 30% of total resistance. For these applications SPPs become an appealing alternative.
In addition surface piercing propellers are less sensitive to cavitation and this contributes to their in-
creasing popularity.

1.2.9.1 Mode of operation


Contrary to what happens with conventional shaft arrangements, SPPs operate at the interface
between air and water, like hulls do. For this reason some peculiar phenomena have to be taken into
account to describe the “modus operandi” of this kind of propulsion system.
Figure 1.77 illustrates an SPP arrangement, where hT is the maximum blade tip immersion.

rev. March 25th 2014 pag. 81 di 100


Marco Ferrando Motor Yacht Design

Figure 1.77: SPP immersion


First of all one has to consider that each blade, during a propeller revolution, pierces the water
surface twice. When a blade enters the water it draws some air below the water level, the air quanti-
ty depending on the propeller loading. On exiting from the water a blade carries along some water
which is then thrown away producing the classic spray tail that characterizes this kind of propeller.
At advance coefficients near the zero-thrust value, the amount of air that follows the blade un-
derwater is quite negligible, while as J decreases the air quantity increases, along with the depth to
which air is drawn.
In this first stage of the operating profile the air is confined at the trailing edge of the blade and
this regime is called base vented.
In the base vented mode the blade itself is still fully wetted and is followed by an air cavity
which extends itself from the free surface to the trailing edge. As the blade proceeds in its revolu-
tion the cavity follows until a balance is reached between the buoyancy of the air and the suction of
the blade. After this point the cavity stops its descent in the water and the blade proceeds quite nor-
mally.
In the meantime the tip vortex of the blade is also ventilated and this phenomenon looks like
tip vortex cavitation, the main difference being that cavities are vented to the atmosphere. This situ-
ation is illustrated in Figure 1.78.

Figure 1.78: Base vented operation


As the propeller loading rises the suction on the back of the blade draws more and more air
down as the blades are immersed. The cavity increases its volume and generally remains confined

rev. March 25th 2014 pag. 82 di 100


Marco Ferrando Motor Yacht Design

to the trailing edge. Some streaks of air spring forward at radial positions corresponding to the max-
imum load of the blade. This is the partially vented mode of operation, illustrated in Figure 1.79.

Figure 1.79: Partially vented operation


Figure 1.80 illustrates the typical shape of the KT versus J curve of a SPP: As it can be seen the
behavior of SPPs is quite different from that of a conventional propeller. In the mentioned figure the
numbers identify different regions of operation, where particular phenomena occurs that explain the
shape of the curve.
Region 1 of Figure 1.80 illustrates that in both of the above mentioned regimes the shape of the
KT versus J curve is almost the same as for a conventional propeller; the same is true for KQ.

Figure 1.80: Typical KT curve for a SPP


A further reduction of the advance coefficients brings the propeller into an unstable condition
called transition.
Here the air cavity becomes highly unstable. The back of the blade changes continuously be-
tween fully wetted, air streaked and completely dry conditions.

rev. March 25th 2014 pag. 83 di 100


Marco Ferrando Motor Yacht Design

This phenomenon is not restricted to a single value of the advance coefficient, but rather it
spans over a certain range of J. In this region KT and KQ values are not unique for a given advance.
Actually their value can change, depending on the condition of the back of the blades. This condi-
tion is represented by region 2 of Figure 1.80.
For advance coefficients below the transition range the blade is fully vented, and the volume of
the air cavity increases as an inverse function of the advance coefficient. This is the so-called fully
vented condition. Here the shape of thrust and torque curves departs completely from that of fully
submerged propellers. When an SPP is fully vented, mainly the face of the blades produces the
thrust. Actually the contribution of the suction is lost, because the back of the blades is vented to the
atmosphere.
In this mode of operation the air cavities attached to the back of the blades have a considerable
thickness, which keeps increasing as J is lowered. In this way the interaction of the blades is very
strong and the propeller is affected by a considerable cascade effect which limits both thrust and
torque. This phenomenon explains the decreasing of KT and KQ as J decreases, and it can be ob-
served in region 3 of Figure 1.80.
At lower advance coefficients another phenomenon appears, which furthermore limits the ca-
pability of the propeller to produce thrust and absorb torque. This is the inflow retardation. The air
cavities are so big as to block the flow of water between the blades, decreasing the mass flow
through the propeller. This, in accordance with the momentum theory, reduces the thrust that the
propeller can produce. Region 4 of Figure 1.80 illustrates this situation.
Finally, at extremely low values of the advance coefficient (see region 5 of Figure 1.80) air
demand for cavity ventilation can be choked by the huge amount of spray. In this way the pressure
inside the cavities is lower than atmospheric and this increases somewhat the thrust of the propeller.

1.2.9.2 Controlling parameters


As partially submerged propellers operate at the interface between air and water, two additional
quantities must be taken into account along with the parameters usually employed to describe the
behavior of conventional propellers. The first one is the immersion coefficient IT, the second one is
the Weber number We. The thrust and torque coefficients are a function of the following parame-
ters:
 P A 
K  f  z, , E , J , Re,  , I T ,  , We, Fr 
 D AO 
The influence of the number of blades, pitch ratio, expanded area ratio and advance coefficient
on the behavior of an SPP is much the same as in the case of a fully immersed propeller. The same
is true for the Reynolds Number. The immersion coefficient IT is the ratio between the maximum
blade tip immersion and the propeller diameter and the value of this coefficient indicates how much
of the propeller is working under water in nominal conditions. It is defined as follows:
hT
IT  (75)
D
The Weber number is a ratio between inertial and surface tension forces and it has the follow-
ing structure:
V2 L
We  (76)

This number is significant when inertial and surface tension forces are of comparable magni-
tude;  is the kinematic capillarity of the fluid.

rev. March 25th 2014 pag. 84 di 100


Marco Ferrando Motor Yacht Design

Influence of 
The angle between the shaft line and the incoming flow, indicated by , plays a more im-
portant role with surface piercing propellers which experience considerable forces in the propeller
plane. This kind of force has the same origin as in the case of fully submerged inclined propellers,
but here, due to the “extreme” asymmetry of the flow field, their magnitude is greater.

Figure 1.81: Disk plane forces coefficients


The resultant force in the propeller plane can be projected in two orthogonal directions. These
directions coincide with the traces of two orthogonal planes passing through the propeller axis. The
first one is vertical, the second horizontal. The component lying in the vertical plane is indicated by
the symbol FN; the relevant non-dimensional coefficient being named KN; the symbols FH and KH
will be used to indicate the horizontal component of the resultant force in the propeller plane and
the relevant non-dimensional coefficient. Figure 1.81 shows an example of such disk plane force
coefficients.

Figure 1.82: SPP vertical plane forces


As for inclined immersed propellers all the forces have to be combined in order to obtain the
horizontal resultant thrust TX and the resultant vertical force TZ. According to ITTC these forces are
obtained from the following equations:

rev. March 25th 2014 pag. 85 di 100


Marco Ferrando Motor Yacht Design

TX  T cos   FN sin 
(77)
TZ  T sin   FN cos 

where T is the axial thrust. Figure 1.82 illustrates the vertical plane forces components, while Fig-
ure 1.83 shows the result of such force composition.

Figure 1.83: Total resultant forces


One important point must be made when dealing with the interpretation of open water test re-
sults. Generally, open water tests on SPPs are carried out in towing tanks or cavitation tunnels with
horizontal water surfaces.
In most of real applications the craft in front of the propeller has a certain amount of deadrise 
(see Figure 1.84). Since the flow breaks out of the transom along the bottom, the free surface of the
water that flows toward the propeller is not horizontal but inclined of the angle .

Figure 1.84: SPPs stern layout


Thus curves like KN and KH in Figure 1.81 are representative of a planing craft with 0 degrees
deadrise angle. The same is true for KTX and KTZ of Figure 1.83.
To obtain true values of horizontal and vertical forces the following equations [1.2.10.1.32]
have to be used:
F ' N  FH sin   FN cos 
(78)
F ' H  FH cos   FN sin 
The effect of the deadrise angle can be observed looking at Figure 1.85

rev. March 25th 2014 pag. 86 di 100


Marco Ferrando Motor Yacht Design

Figure 1.85: Effect of the deadrise angle on the disk plane forces
Accordingly, TX and TZ must be calculated using F’N instead of FN, producing the results illus-
trated in Figure 1.86

Figure 1.86: Effect of the deadrise angle on total resultant forces


Focusing the interest on the capability of the propeller of producing thrust, even a considerable
change in shaft inclination produces a minor effect.

Figure 1.87: Effect of the shaft inclination on the axial thrust


Figure 1.87 illustrates the effect of changing  from 4 to 8 degrees. In order to avoid the bias
imposed on the advance coefficient by the shaft inclination, another formulation is used, J which
is defined by the equation:

rev. March 25th 2014 pag. 87 di 100


Marco Ferrando Motor Yacht Design

VA cos
J  (79)
nD

Influence of IT
It is obvious that the immersion of a SPP strongly affects its behavior. The non-dimensional
coefficient representing the immersion of the propeller is the immersion coefficient IT, defined in
equation (75). Figure 1.88 illustrates the maximum tip immersion.

Figure 1.88: Maximum tip immersion

The nominal operating condition for a SPP is IT=0.50, but sometimes a different immersion is
preferred. Accordingly, it is very important to know the impact of the immersion coefficient on the
propeller performance.
Thrust and torque increase along with the immersion coefficient. Figure 1.89 illustrates the
changes in KT of a SPP as a function of IT. From a designer's point of view this is surely the most
suited presentation, for it allows an easy calculation of the propeller thrust.

Figure 1.89: Influence of IT on KT

Nevertheless KT and KQ do not allow a quick understanding of the real behavior of a SPP. Ac-
tually, these coefficients are obtained dividing thrust and torque by the diameter. In case of a par-
tially submerged propeller T and Q are sensitive to a change in the propeller immersion while the
diameter is a constant. This explains the considerable variation of KT and KQ with IT for a given J.

rev. March 25th 2014 pag. 88 di 100


Marco Ferrando Motor Yacht Design

In this way one cannot assess if the immersion coefficient affects the thrust production of the
immersed part of the propeller. Accordingly, Hadler and Hecker [1.2.10.1.33] suggested a modified
version of the thrust and torque coefficients:
T Q
CT '  CQ ' 
1 1
 AO ' VA 2  D AO ' VA 2
2 2
in which AO’ is the actual submerged disk area. This is obtained calculating the area of the pro-
peller disk that lies below the undisturbed water surface, as shown in Figure 1.90.

Figure 1.90: Actual immersed area.


They also presented their test results in graphics illustrating J and O as a function of C’T. This
kind of presentation pre-vents the description of the behavior of the propeller at low values of the
advance coefficient because in this zone the thrust and torque coefficients approach infinity.
Following this, it appears more useful to define [1.2.10.1.34] two dimensionless coefficients
having the same structure as KT and KQ in which the actual immersed area AO’ is used, namely:
T Q
K 'T  K 'Q 
 n D 2 AO '
2
 n D3 AO '
2

Figure 1.91: Influence of IT on K’T

rev. March 25th 2014 pag. 89 di 100


Marco Ferrando Motor Yacht Design

In Figure 1.91 the same test results of Figure 1.89 are illustrated, as an example of the use of
K’T. The same procedure can be applied to KN and KH.
The use of AO’ has been criticized by Kruppa [1.2.10.1.35] because, due to the possible rise of
the water level as a consequence of flow retardation, it is not clear which is the actual immersed
disk area if the advance coefficient is below the transition limit. From Figure 1.91 it can be inferred,
however, that the behavior of the propeller is substantially unaffected by the immersion ratio pro-
vided that the operating point falls in the partially ventilated region.

Influence of 
The cavitation index affects the performance of SPPs, because surface piercing propellers can
also suffer from cavitation.
Cavitation on a SPP is actually possible only when the blade surface is wetted. Following this,
it is clear that for advance coefficients lower than the transition value, the development of back cav-
itation is prevented by the presence of the air bubble which covers almost the whole blade. So the
possibility of the occurrence of back cavitation is confined in the region where the blades are par-
tially ventilated, namely in the partially vented mode of operation (see region 1 of Figure 1.80).
However cavitation is much less important than with conventional propellers. This can be fully
understood by recalling the mode of operation of a SPP. In fact in the partially ventilated mode the
radial sections that should carry the greater load in a conventional propeller are generally ventilated,
as illustrated in Figure 1.79, preventing the onset of vapor bubbles. Accordingly, bubble cavitation
in the wetted portions of the blades will eventually develop at very low values of cavitation index.
Face cavitation could also develop at high values of the advance coefficient, where the outer
radial sections experience a negative angle of attack. Again this is not the case for a SPP, because
those sections are generally face vented by streak-like bubbles drawing air through the water sur-
face.

Figure 1.92: Influence of  on SPPs performances.


Unfortunately cavitation of SPPs is rarely dealt with in the available literature. Brandt
[1.2.10.1.36] performed the first study regarding the influence of  on the performances of SPPs.
Figure 1.92, from[1.2.10.1.37], offers an example of that influence.
Olofsson [1.2.10.1.38] also gives some considerations on the effect of cavitation on SPP’s per-
formances.

rev. March 25th 2014 pag. 90 di 100


Marco Ferrando Motor Yacht Design

Influence of We
As is well known the Weber number is a ratio between inertial and surface tension forces. Its
influence can be easily foreseen for an SPP, which continually pierces the water surface.
Shiba developed the first study on the influence of We in his comprehensive work on air draw-
ing of marine propellers [1.2.10.1.39]. Shiba’s work was devoted to the study of merchant ship pro-
pellers in semi-submerged condition, but his findings also hold in the case of SPPs.
The Weber number has the following structure:
V2 L
We 

where  is the kinematic capillarity. Shiba introduced a particular kind of Weber number,
which he judged more adequate to his needs:
n 2 D3
We' 

According to Shiba, surface tension plays its role when the propeller is about to be fully venti-
lated. Actually complete ventilation is a rather sudden phenomenon that can be correlated to a cer-
tain value of J called the critical advance coefficient JCR. The critical advance coefficient can rough-
ly be located in the middle of the transition region and the sudden drop of KT, KQ, KN and KH identi-
fy its position (see Figure 1.80 and Figure 1.81).
Indeed Shiba found a correlation between We’ and JCR for a single propeller. Figure 1.93 re-
produces this correlation showing that the influence of the Weber number almost disappears for We’
> 180. In this perspective model tests can be used to predict the behavior of full-scale propellers if
they are performed at values of We’ greater than 180.

Figure 1.93: Shiba's We'-JCR correlation


Shiba’s kind of Weber number does not contain the immersion of the propeller. If one tries to
build a diagram like that in Figure 1.93 containing data of an SPP at different immersions the result
is Figure 1.94. This figure summarizes experiments on E9401 model that were carried out in the
cavitation tunnel of the University of Genoa and in the towing tank at the University of Naples. The
data points fall in two distinct sets because the test conditions in the two facilities were kept as far
apart as possible in order to obtain the widest range for the Weber number.

rev. March 25th 2014 pag. 91 di 100


Marco Ferrando Motor Yacht Design

Figure 1.94: Effect of IT change on We’-JCR correlation.


In each set different points have the same We’ but different critical advance coefficients. These
points pertain to different immersions of the propeller. Following this consideration a new kind of
Weber’s number was devised [1.2.10.1.34] containing the maximum immersion of the propeller tip:
n 2 D 2 hT n 2 D3 IT
We ''  
 
’’
If the same data of Figure 1.95 are plotted against We again a correlation can be found as Fig-
ure 1.96 illustrates [1.2.10.1.34]. Unfortunately in the facilities where E9401 propeller was tested it
was impossible to attain values of We’’ greater than 190 and so there is not sufficient evidence to
suggest 190 as a new minimum value of the Weber number for scaling purposes.

Figure 1.95: Example of the use of We’’.


This impression is reinforced if one looks at Figure 1.96 [1.2.10.1.34], where test results of
three different propellers are illustrated. The main characteristics of the propellers are summarized
in Table 1.2-10.

Table 1.2-10: Geometric characteristics of propeller models

rev. March 25th 2014 pag. 92 di 100


Marco Ferrando Motor Yacht Design

Model Type D [m] P/D AE/AO


E9401 SPP with cupped sections 0.25 1.2 0.80
E042 Wageningen B-Screw Series 0.18 1.4 0.95
E9501 SPP with Diamond Back sections 0.25 1.2 0.68

Figure 1.96: We’’ - JCR correlation for different SPPs.


Looking at this figure one can assume that each propeller has its own minimum Weber num-
ber.
It is then evident that further investigations are required in order to ascertain the real influence
of the Weber number upon the performances of SPPs.

Influence of Fr
The Froude number is usually included among the parameters that govern the behavior of a
propeller as a result of dimensional analysis. Actually its influence is negligible for deeply im-
mersed propellers and so the equality of model and ship Froude numbers is never achieved.
As every Naval Architect could guess the Froude number does influence the operation of a sur-
face piercing propeller.
The first comments on the role of Fr are again due to Shiba [1.2.10.1.39]. In his investigation
he pointed out that gravity affects the shape of the air cavity through the Bernoulli equation that can
be enforced at the boundary between water and the atmosphere vented cavity.
Accordingly, Shiba showed in his paper that the influence of:
nD
FrD 
gD

vanished when the air cavity approached its ultimate form, i.e. for Froude numbers greater than 3.
Hadler and Hecker [1.2.10.1.40] indirectly concurred on this hypothesis. Actually they stated
that the maximum pressure differential obtainable for each radial section of an SPP is h, h being
the maximum depth of the section. Accordingly, they derived a sort of cavitation index in the fol-
lowing form:

rev. March 25th 2014 pag. 93 di 100


Marco Ferrando Motor Yacht Design

h
h 
U2
1
2
where U is the total inflow velocity to the propeller blade section. This equation can be transformed
as follows:
2gh 2
h  2

U Frh 2
Furthermore they suggested the existence of a limiting value of h below which no further
change could be obtained in the propeller performances.
In addition Hadler and Hecker noted that at the low advance coefficients the propeller and its
hub were generating a wave train which appeared to modify the submergence.
Also Brandt [1.2.10.1.41] commented on this matter stating that the Froude number, calculated with
the immersion hs of the shaft:
V
Frhs 
g hs
has a definite influence on model tests if its values are less than 4.
Finally, Olofsson too acknowledges an influence of FrD but he sets to 4 the limiting value be-
yond which the influence disappears.
Summarizing, it is widely recognized that the Froude number does affect the behavior of SPPs,
and all authors have suggested the existence of a threshold value that limits this influence. The same
general agreement has not been reached on the minimum value that must be attained to avoid scal-
ing problems. This is probably due also to the different kind of Froude numbers that have been
used.

1.2.9.3 Design issues


Setting aside all of the open questions regarding performance scaling, the surface piercing pro-
peller appears to have reached a sufficient degree of reliability to offer a valid and appealing alter-
native to conventional propulsive means, especially for very fast crafts.
Being rather an extreme propeller, special care must be used when employing the SPP. In this
perspective a few design issues are worth of some attention.

Start-up
Due to the reduction of the suction side contribution to the total thrust, surface piercing propel-
lers usually have a rather high pitch ratio; this can be source of troubles at start-up.
When the boat is at rest or at low speed, the propeller operates fully submerged and then the
active area is doubled. In this condition much more torque is required to spin the propeller and
sometimes the engine cannot provide enough torque to accelerate the boat.
Figure 1.97 illustrates the characteristics curves of a SPP for IT=0.5 and for IT=2.0.
From the figure one can see that at lower advance coefficients the torque requirement in fully
submerged condition is about ten times that in surface piercing condition.
To avoid start-up problems two solutions are currently adopted. Firstly, reduction of propeller
submergence is employed in articulated drive installations. Secondly, artificial ventilation of the
back of the blades can be obtained by means of passive aeration pipes or directing engine exhaust
into the water in front of the propeller.

rev. March 25th 2014 pag. 94 di 100


Marco Ferrando Motor Yacht Design

Figure 1.97: Comparison between fully and partially immersed conditions.

Transom ventilation
A similar immersion problem may arise when dealing with two speed craft. Actually the pro-
peller is generally chosen to fulfill top speed requirements, then the designer checks the propeller at
cruising speed.
In some instances the cruising speed is so low that the transom remains wetted and the propel-
ler is almost fully sub-merged. This again produces abnormal torque absorption and prevents the
boat from maintaining the cruising speed

Backing performances
Figure 1.98 shows some of the most common foil sections that are employed in SPPs.

Figure 1.98: Most common blade sections for SPP use.


Though the first section is patented and designed in view of improving reverse performances, nei-
ther of these provides backing thrust comparable with that of a conventional section. For this reason
final users of SPPs must be cautioned about maneuvering with great care, especially if accustomed
to conventional propellers.

rev. March 25th 2014 pag. 95 di 100


Marco Ferrando Motor Yacht Design

Fatigue and loads


SPPs are probably the most challenging propellers from a structural point of view, because of
the extreme asymmetry of their operation.
It is well known that every conventional propeller experiences variations in blade loading dur-
ing a revolution, as a con-sequence of the wake field. This problem is well known and designers
have the tools to deal with it.
A new aspect of asymmetry in blade loading arose when considerably inclined shaft lines were
introduced in fast pleasure boats and hydrofoils. Highly inclined shafts introduced fatigue as a de-
sign item due to the oscillatory nature of the load. This problem is more extreme for SPPs because a
blade is actually unloaded for a considerable part of its revolution. This feature leads to a pulsating
load that is far worse with regard to fatigue.
Figure 1.99 shows a qualitative trend of the blade load of an SPP.

Figure 1.99: Qualitative trend of blade load.


The peak load corresponding to the entry of the blade into the water is due to two causes. The
first is the impact of the blade with the water during the entry phase; the second is the maximum
angle of attack which occurs at the 90° angular position of the blades on an inclined propeller. For
these reasons the propeller itself has to be designed bearing in mind the severe fatigue loading.
Consequently shafts also suffer from fatigue due to the alternating flexure and torsion that are im-
posed by the propeller. Furthermore the considerable disk plane forces, that characterize SPPs, re-
quire a careful design of the bearings.
Summarizing, a considerable effort must be devoted to the structural design both of the propel-
ler and the shafting in order to achieve a safe and reliable propulsion unit.

1.2.9.4 Conclusions
This overview on surface piercing propellers may have suggested the idea that this kind of pro-
pulsion is still immature and skeptics will probably lean toward some more classic propulsion de-
vice.
From a scientific point of view this feeling is well justified because what we know does not
completely explain the operation of SPPs. In addition, the influence of the various parameters has
not been thoroughly assessed. But every scientist knows of not knowing.
From a practical point of view, the widening diffusion of surface propulsion and the growing
interest on this matter, suggest that the moment has come for using such a propulsive mean. Fur-

rev. March 25th 2014 pag. 96 di 100


Marco Ferrando Motor Yacht Design

thermore one must remember that every propulsion system has been immature at the beginning, but
wide usage has allowed the eventual ripening.
Summarizing we can draw the following conclusions:
potential benefits
 the SPP concept is very promising and some applications have proved that this propulsor
could be superior at very high speed;
 the efficiency of SPPs are comparable to those of conventional fully submerged propellers;
 the appendage resistance saving can be substantial at high speed;
drawbacks
 performance scaling needs further effort in order to ascertain the real influence of the vari-
ous parameters, namely the Froude number and the Weber number;
 the influence of geometric parameters has not been completely studied. In particular blade
rake and skew have so far been neglected;
It is hoped that the increasing use of SPPs will boost the research efforts in this field leading to
further exciting research results.

1.2.10 References
1.2.10.1.1 Savitsky D. “Hydrodynamic Design of Planing Hulls”, Marine Technology, October 1964.

1.2.10.1.2 Saunders H.E., “Hydrodynamics in Ship Design”, vol. II, SNAME, 1957

1.2.10.1.3 Savitsky D. Brown P.W., “Procedures for Hydrodynamic Evaluation of Planing Hulls in Smooth and
Rough Water”, Marine Technology, Vol. 14, No. 4, October 1976

1.2.10.1.4 Wagner H., "The Phenomena of Impact and Planing on Water," NACA translation 1366, ZAMM, August
1932.

1.2.10.1.5 Sottorf W., "Experiments With Planing Surfaces," NACA TM 661, 1932, and NACA TM 739, 1934.

1.2.10.1.6 Savitsky D., Neidinger J. W., "Wetted Area and Center of Pressure of Planing Surfaces at Very Low Speed
Coefficients," Stevens Institute of Technology, Davidson Laboratory, Report No. 493, July 1954.

1.2.10.1.7 Korvin-Kroukovsky B. V., Savitsky D. and Lehman W. F., "Wetted Area and Center of Pressure of Planing
Surfaces," Stevens Institute of Technology, Davidson Laboratory Report No. 360, August 1949.

1.2.10.1.8 Savitsky D. and Ross E., "Turbulence Stimulation in the Boundary Layer of Planing Surfaces," Stevens
Institute of Technology, Davidson Laboratory Report 44, August 1952.

1.2.10.1.9 Blount D., Fox D.,”Small Craft Power Predictions“, Marine Technology. Vol. 13, No. 1, Jan 1976, pp. 14-
45

1.2.10.1.10 Davidson K. and Saurez A., “Tests of Twenty Related Models of V-Bottom Motor Boat, E.M.B. Series 50,"
David Taylor Model Basin Report 170, December 1948.

1.2.10.1.11 Clement, Eugene P. and Blount, Donald L. “Resistance Tests of a Systematic Series of Planing Hull
Forms” SNAME Transactions, Vol. 71, 1963

1.2.10.1.12 Clement, Eugene P, "How to use the SNAME Small Craft Data Sheets for Design and for Resistance Pre-
diction," SNAME Technical and Research Bulletin No. I-23.

1.2.10.1.13 Keuning J. A. and Gerritsma J., "Resistance, Tests of a Series of Planing Hull Forms with 25 Degrees
Dead Rise Angle", International Shipbuilding Progress, Sept. 1982 No. 337.

rev. March 25th 2014 pag. 97 di 100


Marco Ferrando Motor Yacht Design

1.2.10.1.14 Keuning J. A., Gerritsma J. and van Terwisga P. F., "Resistance, Tests of a Series of Planing Hull Forms
with 30 Degrees Dead Rise Angle and a Calculation Model Based on This and Similar Systematic Series",
International Shipbuilding Progress 40, 1993, No. 424.

1.2.10.1.15 Holling, H. D. and Hubble, E. N., "Model Resistance Data of Series 65 Hull Forms, Applicable to Hydro-
foils and Planing Craft," Naval Ship Research and Development Center Report 4721 (May 1974).

1.2.10.1.16 Hadler J. B., Hubble E. N., Holling N. O., “Resistance Characteristics of a Systematic Series of Planing
Hull Forms - Series 65”, SNAME Chesapeake Section, May 1974

1.2.10.1.17 Metcalf B. J., Faul L., Bumiller E, Slutsky J., “Resistance Tests of a Systematic Series of U.S. Coast Guard
Planing Hulls, NSWCCD-50-TR-2005/063, December 2005.

1.2.10.1.18 Lahtiharju E., Karppinen T., Hellevaara M.,. Aitta T, “Resistance and Seakeeping Characteristics of Fast
Transom Stern Hulls with Systematically Varied Form”, SNAME Transactions, vol 99, 1991.

1.2.10.1.19 Savitsky D., DeLorme M. F., Datla R., "Inclusion of “Whisker Spray” Drag in Performance Prediction
Method for High-Speed Planing Hulls", Technical Report SIT-DL-06-9-2845, Davidson Laboratory-
Stevens Institute of Technology, March, 2006.

1.2.10.1.20 Blount D., Bartee R., "Design of Propulsion Systems for High-Speed Craft", Marine Technology. Vol. 34,
No. 4, Oct. 1997, pp. 276-292

1.2.10.1.21 Hadler, J. B., " The Prediction of Power Performance of Planing Craft", SNAME Transactions, Vol. 74,
1966.

1.2.10.1.22 Day, J. P., Haag, R. J.” Planing Boat Porpoising.”, Webb Institute of Naval Architecture, New York, 1952.

1.2.10.1.23 Blount, D. L. "Dynamic Stability of Planing Boats.", Marine Technology. Vol 29. 1992.

1.2.10.1.24 Celano, T. “The Prediction of Porpoising Inception for Modern Planing Craft”, U.S. Naval Academy, An-
napolis, MD, Trident Report No. 254, 1998.

1.2.10.1.25 Angeli, J.C., “Evaluation of the Trim of a Planing Boat at Inception of Porpoising”, SNAME, Spring Meet-
ing, Paper D, April 1973.

1.2.10.1.26 Clement, E. P., "Effects of Longitudinal Bottom Spray Strips on Planing Boat Resistance," David Taylor
Model Basin Report 1818 (Feb 1964).

1.2.10.1.27 Levi, R. S., “Planing Craft Design and Performance”, 10th WEGEMT School on "High speed and pleasure
crafts" October 21 - 26 1985, Santa Margherita Ligure.

1.2.10.1.28 Planing Boat Resistance," David Taylor Model Basin Report 1818 (Feb 1964).

1.2.10.1.29 Brown, P. W. “An experimental and theoretical study of planing surfaces with trim flaps”, Report SIT-DL-
71-1463, Stevens Institute of Technology, April 1971.

1.2.10.1.30 Savitsky D., Brown, P. W. “Procedures for hydrodynamic evaluation of planing hulls”, Report SIT-DL-75-
1859, Stevens Institute of Technology, November 1975.

1.2.10.1.31 Dawson D., Blount D. L., “Trim Control” Professional Boatbuilder, No. 75, February-March 2002.

1.2.10.1.32 Rose, J.C. Kruppa C.F. & Koushan K., “Surface Piercing Propellers – Propeller/Hull Interaction”,
FAST’93, Vol. 1, pp. 867-881. Yokohama: Society of Naval Architects of Japan, 1993.

1.2.10.1.33 Hadler J.B. & Hecker R. “Performance of partially submerged propellers”. In Proceedings, Seventh Sym-
posium on Naval Hydrodynamics (ed. R.D.Cooper & S.W.. Doroff), pp. 1449-1493. Arlington: Office of
Naval Research – Department of the Navy, 1968.

rev. March 25th 2014 pag. 98 di 100


Marco Ferrando Motor Yacht Design

1.2.10.1.34 Ferrando, M. & Scamardella, “A. Surface Piercing Propellers: Testing Methodologies, Result Analysis
and Comments on Open Water Characteristics.” In Proceedings Small Craft Marine Engineering Resis-
tance & Propulsion Symposium, pp.5-1 – 5-27. Ypsilanti: University of Michigan, 1996.

1.2.10.1.35 Kruppa C.F. “Testing of partially submerged propellers”, in Proceedings, 13TH ITTC Report of Cavita-
tion Committee, Appendix V. Berlin, 1972.

1.2.10.1.36 Brandt, H. “Modellversuche mit Schiffpropellern an der Wasseroberflache”, Schiff und Hafen, vol.24 No.
5, pp. 415-422, 1973.

1.2.10.1.37 Rose, J.C. & Kruppa C.F. “Surface Piercing Propellers - Methodical Series Model Test Results”, In
FAST’91 (ed. K.O. Holden et al.), Vol. 2, pp. 1129-1147. Trondheim: Tapir Publishers, 1991.

1.2.10.1.38 Olofsson, N. “Force and Flow Characteristics of a Partially Submerged Propeller”, Doctoral Thesis.
Goteborg: Chalmers University of Technology Department of Naval Architecture and Ocean Engineering,
1996.

1.2.10.1.39 Shiba H., “Air-Drawing of Marine Propellers”, Transportation Technical Research Institute, Report No. 9.
Tokyo: the Unyu-Gijutsu Kenkyujo, 1953.

1.2.10.1.40 Hadler J.B. & Hecker R,. “Performance of partially submerged propellers”, In Proceedings, Seventh Sym-
posium on Naval Hydrodynamics (ed. R.D.Cooper & S.W.. Doroff), pp. 1449-1493. Arlington: Office of
Naval Research – Department of the Navy, 1968.

1.2.10.1.41 Brandt, H., “Modellversuche mit Schiffpropellern an der Wasseroberflache”, Schiff und Hafen, vol.24 No.
5, pp. 415-422, 1973.

rev. March 25th 2014 pag. 99 di 100


Marco Ferrando Motor Yacht Design

1.3 Sea trials

rev. March 25th 2014 pag. 100 di 100

You might also like