Está en la página 1de 19

Twenty-Ninth Symposium on Naval Hydrodynamics

Gothenburg, Sweden, 26-31 August 2012

Quasi-Steady Two-Quadrant Open Water Tests for


the Wageningen Propeller C- and D-Series
Jie Dang, Joris Brouwer, Ren Bosman and Christiaan Pouw
(MARIN, 2 Haagsteeg, 6708PM Wageningen, The Netherlands)
m, ma
n

ABSTRACT
The Maritime Research Institute Netherlands (MARIN)
has recently started a Joint Industry Project (JIP) on
controllable pitch propeller (CPP) series called the Wageningen
Propeller C- and D-series, after the successful development of
the famous Wageningen B-series which are used by designers
and engineers worldwide.
The B-series comprise the open water characteristics of
conventional fixed pitch propellers (FPPs) designed for
merchant ships with various numbers of blades and blade area
ratios at different pitch. For several of these propellers, also the
four-quadrant characteristics were published by MARIN in the
sixties and seventies of the last century.
Today many ships are equipped with CPPs. Also used
widely are the ducted CPPs, both for ships and offshore
structures. The off-design performance of the CPPs is not only
of importance for ships powering performance, but also for e.g.
dynamic positioning and manoeuvring of those vessels. Due to
a lack of systematic information for the CPPs in such cases, the
B-series data are often used instead, both for the estimation in
an early design stage and also as the final data delivered for
specific new CPP designs, simply because there is no other
systematic data available rather than the B-series data.
However, the characteristics of CPPs differ substantially from
those of FPPs. There is a high demand for developing CPP
series with full off-design information - the complete twoquadrant open water characteristics at all possible pitch
settings.
In order to reduce the cost, a quasi-steady propeller open
water test technique has been developed and thoroughly studied
under support of this JIP, which reduced the tank test time by a
factor of 8 to 10. This method ensures the affordability of the
tests for the C- and D-series, and therefore the whole JIP.
In addition to the propeller thrust and torque, the propeller
blade spindle torque is also provided as systematic data in
propeller series for the first time.
NOMENCLATURE

hydrodynamic pitch angle [o]


D
propeller diameter [m]; drag [N]
I, Ia
mass moment and added mass moment of
inertia [kgm2]
k
reduced frequency[-]; Fourier harmonics [-]

P
Q
Qblade
R
T
Va

mass and added mass [kg]


propeller shaft rotational rate [1/s]
angular acceleration [1/s2]
propeller pitch [m]
propeller shaft torque [Nm]
propeller blade spindle torque [Nm]
propeller radius [m]
propeller thrust [N]
test run period [s]
propeller advance speed [m/s]
linear acceleration [m/s2]

INTRODUCTION
The Maritime Research Institute Netherlands (MARIN),
former Netherlands Ship Model Basin (N.S.M.B.), started to
develop the well-known Wageningen Propeller B-series right
from the establishment of this institute in 1932 (Kuiper 1992).
The first series were published by van Lammeren (1936) and
Troost (1938 and 1940), followed by a long period of further
developments and expansions of the series over more than 40
years. A major review of the available data was given by van
Lammeren et al (1969 and 1970). The B-series had been further
extended to 6 and 7 bladed propellers in the 1970s. Totally, 20
series with more than 120 propellers were tested over that
period.
Systematic series have also been developed for ducted
propellers since 1954 (van Manen 1954). A major amount of
data of the Ka-series were published by Oosterveld (1970). In
the meantime, other systematic propeller series were also
developed worldwide, such as the Taylor, Gawn and MAU
series. However, none of the these series is so extensive as the
B-series which have found widespread applications.
Besides that the propeller characteristics (the thrust and the
torque) of the series in design operation conditions have been
made available by model tests between J=0 and KT=0, fourquadrant open water characteristics of some of the propellers in
the B-series and in the Ka ducted propellers series were also
made available in the 1980s (MARIN report 1984) for offdesign conditions. Table 1 provides an overview of the
propellers in the B-series where their 4-quadrant open water
characteristics are available. For the Ka-series, only Ka4-70
propellers in 19A and 37 ducts have been published.

Table 1 Overview of B-series with four-quadrant open


water characteristics (pitch ratio P/D of the propellers are
listed in the table).
AE/A0 [%]
Z=3

40

55

Z=4

1.0

1.0

Z=5
Z=6
Z=7

65
1.0

70

75

80

0.5, 0.6, 0.8


1.0, 1.2, 1.4

85

100

1.0

1.0

1.0
1.0
1.0

Different from fixed pitch propellers (FPPs), controllable


pitch propellers (CPPs) are well-known for their advantage for
full power utilization at any circumstances: accelerating and
stopping; rapid manoeuvring; dynamic positioning (DP); etc.
For these reasons, CPP are widely used for multi-purpose
vessels where their propulsors are often used in off-design
conditions.
In order to predict the performance of a CPP in off-design
conditions, people have to either carry out dedicated and
expensive measurements for a specific propeller design, such as
often done for navy vessels (Hampton 1980, Queen 1981), or
rely on the estimated values from the existing four-quadrant
open water data from the B-series (Roddy et al 2006), which
were primarily designed for merchant ships with FPP blade
forms. Scarce information is available in the public domain for
the complete two-quadrant open water characteristics of CPPs,
especially when the propeller blades are deflected away from
its design pitch (Yazaki 1962, Chu et al 1979). In the
Wageningen series book (Kuiper 1992), off-design information
is only available for two CPPs in ahead and astern conditions,
one with a design pitch ratio of zero and the other of one.
With the booming business in oil exploration in recent
years, accurate prediction of the off-design performance of a
propulsor becomes more important than ever, especially in the
requirements for DP operations. Dedicated tests for each
propeller design is unaffordable for most of the projects, while
the existing limited information is far than enough. There is a
strong demand on developing new contemporary CPP series
with complete information of their off-design performance.
In addition to these, a CPP blade has a completely
different blade form as an FPP. This is because more practical
issues need to be considered for a CPP, such as: that the blades
must be able to pass each other from positive pitch to negative
pitch; that the blade has to sit on the blade foot between bolt
holes; that the blade overhang at the blade root is not preferable
to prevent stress concentration; that the blade tip must not touch
the inner side of a duct at any deflected pitch angles for the
ducted CPPs; etc. Within all of these, one of the important and
unique thing is the blade spindle torque of CPPs (Pronk 1980),
where very limited information can be found (Chu et al 1979,
Ito et al 1984, Jessup et al 2009, Koushan 2011). To the
knowledge of the authors, there is also no CPP series with
systematic information on the propeller blade spindle torque at
all possible blade pitch settings (from full positive pitch to full
negative pitch and over the complete two quadrants).

With the strong demands from the industries and by taking


into account the fact that CFD calculations (Chen and Stern
1999) are not yet accurate enough and needs still to be
validated against model test results, from the beginning of 2011
MARIN started to consider, together with the universities and
the industries, the possibilities of developing a new CPP series.
In September 2011, a Jointed Industry Project (JIP) was
officially launched, which is called the Wageningen Propeller
C- and D-series for both open and ducted CPPs. Here the C
stands for controllable and the D stands for ducted.
Conducting propeller series tests for the complete two
quadrants, especially at different pitch settings for CPPs
(typically more than 10 pitch settings are needed between full
positive and full negative pitch), is not affordable at the present
economic situation. New test technology needs to be developed
in order to reduce the cost significantly. This is made possible
by the rapid development of sensor technology in the past
decades, which makes dynamic measurement possible at higher
frequencies with rapid response. This leads to the idea of a
quasi-steady test technique for propeller open water
characteristics.
Quasi-steady test techniques have been already used for the
propulsion tests in the towing tank of MARIN for some years
(Holtrop and Hooijmans 2002, Verhulst and Hooijmans 2011).
However, quasi-steady propeller open water test has never been
explored in the past. Under support of the Wageningen C- and
D-series JIP, a pilot study has been successfully carried out,
which proves that the quasi-steady test results are as accurate as
the conventional steady test results, while reduces the test time
by a factor of 8 to 10.
The results of the pilot study is presented in this paper,
with detailed discussions on the test set-up, the sensors, the test
procedure, the data analysis and the results, including also the
uncertainty analysis.
FACILITY, TEST SET-UP AND INSTRUMENTATION
The model tests for the study have been carried out in the
Deep Water Towing Tank (DT) of MARIN, which measured
250m long, 10.5m wide and 5.5m deep. The detailed
description of the facility can be found on MARIN web site.
An open water test set-up, as sketched below in Figure 1,
has been used for the present study. This test set-up has a very
slender POD body and a thin strut. The propeller shaft is driven
by a toothed-belt through the hollow strut, connecting the
propeller shaft to the electric motor shaft above on the towing
carriage.

Figure 1 A sketch of the open water test set-up with test cap.

Dedicated sensors have been designed and manufactured to


measure the propeller thrust, the torque on the shaft and also
the blade spindle torque on one blade the key blade. The
blade transducer is special designed and capable to measure the
spindle torque on the key blade with a negligible disturbance of
the thrust and torque forces which provide bending moments on
the transducer at the same time. All the sensors are shown in
Figure 2.
After the placement of the strain gauges and soldering
lacquer threads for the blade transducer, the transducer is
coated with a special watertight coating which stays flexible to
avoid hysteresis and creep. When this process is done, the
transducer is tested in water for three days to check if the
transducer is still watertight. After three days, the insulation
value should be more than 500 M otherwise the
measurements can be disturbed.
The signals from the sensors are transmitted by cables
through the hollow shaft to the other end of the test set-up and
are sent to the carriage through slip-rings. This open water test
set-up is equipped with only an eight-channel slip ring set.
Normally this is enough for the thrust and torque measurements
- four channels for the excitation voltage of the two strain
gauge bridges and four channels for the signals. For the
present three sensors, the excitation voltage is shared. The other
six channels are used for the signals of the three sensors.

anodized surface, in order to reduce the influence of the mass


and mass moment of inertia on the measurements. Two photos
of the propeller with the blades fitted to the instrumented hub
are shown in Figure 3 where the key blade is bolted to the blade
sensor which can be easily identified on the photo by the gaps
between the blade foot and the hub (the blade facing the reader
in the photo on the left is the key blade).

Figure 3 MARIN stock propeller blades No. 7216R fitted to


the measuring hub with blade sensor, at design pitch.
The inside space of the hub has been fully used to
accommodate the blade spindle torque sensor and no shaft hole
with keyway is able to be made through the hub. The propeller
is hence mounted on one end of the hub by flange directly on
the thrust and torque sensors without any friction of sealing and
bearings. In order to have control on the accuracy of the blades,
the mounting method and the CPP pitch setting, the propeller
has been optically scanned at its design pitch. The results are
compared to the theoretical geometry and the deviations are
shown in Figure 4 and 5.
The major deviations are seen as a kind of small inclination
due to the structure of the two halves of the hub. The pitch at
0.7R is judged as accurate enough and attention needs to be put
on the adjustment of the pitch settings between test runs.

Figure 2 Propeller shaft thrust and torque sensors and the


blade spindle torque sensor
A four-bladed controllable pitch propeller model No.
7216R from the stock of MARIN has been chosen for the
present study. The propeller model is made in such a way that
three of the blades are directly clamped by the two-halves of
the hub with four bolts (see Figure 2) while the key blade is
bolted to one end of the blade spindle torque sensor and the
other end of the sensor is clamped also by the two halves of the
hub. By loosening the bolts on the hub, the pitch of each blade
can be adjusted and set to different pitch settings, including the
key blade. This is usually done on the MARINs pitch
adjustment table.
This selected stock propeller is a typical controllable pitch
propeller (CPP) with contemporary blade design for high power
density and high speed vessels with comfort requirements. Both
the propeller blades and the hub are made of aluminium with

Figure 4 Deviations (in mm) of the stock propeller from its


theoretical geometry, pressure side, results of optical scan.

At propeller off-design conditions, the propeller


hydrodynamic pitch angle is often used, instead of the
advance ratio J, to define the operation condition of the blades,

Under this definition, a complete set of two-quadrant open


water characteristics of a controllable pitch propeller covers the
range -90o +90o.
A quasi-steady open water test is, in principle, an unsteady
model test by continuously varying the advance speed and/or
the rotational rate in order to obtain the steady state
performance of a propeller. For the present study, we have
proposed the following four test runs in order to cover the
complete two quadrants, as numbered in Table 2.

Figure 5 Deviations (in mm) of the stock propeller from its


theoretical geometry, suction side, results of optical scan.
The test setup after mounting the propeller is shown by the
photo in Figure 6. During the tests, the shaft is immersed under
the water surface with a distance according to ITTC (2008)
standard procedure.

Figure 6 The instrumented test set-up, connected to the


towing carriage before immersing into the water.
QUASI-STEADY TEST PROCEDURE & ASSUMPTIONS
In a conventional propeller open water test from J=0 to
KT=0, the propeller shaft rotational rate is often kept constant
while the advance speed of the propeller varies, as
recommended by the ITTC (2008). During propeller fourquadrant open water tests, both the advance speed and the shaft
rotational rate have to vary and change directions, because only
a finite towing speed of the carriage can be achieved. However,
most controllable pitch propellers will never rotate reversely,
except for some special applications e.g. a CPP connected to a
diesel-electric drive system. This practice has been also used
here during the model tests, where only one rotational direction
(positive rotational direction) has been tested. Therefore, only
two-quadrant (the first and the fourth quadrant) open water
characteristics have been studied and discussed in this paper for
the quasi-steady test technique. Extending this technique to the
full four-quadrant tests should be straightforward.

Table 2 Quasi-steady test runs for the complete 2-quadrant


open water characteristics of a controllable pitch propeller.
run shaft rotational rate advance speed
range
1
2
3
4

constant +nmax
0 to +nmax to 0
constant +nmax
0 to +nmax to 0

0 to +Va max to 0
constant +Va max
0 to -Va max to 0
constant -Va max

0 o to ~+30o to 0 o
+90o to ~+30o to +90o
0 o to ~-30o to 0 o
-90o to ~ -30o to -90o

This proposal makes it possible to test the complete twoquadrant open water characteristics of a propeller in only 4 test
runs, using 2 runs by varying the towing speed of the carriage
and 2 runs by varying the shaft rotational rate.
From the first two runs - No. 1 and No. 2, the results in the
first quadrant for from 0 to +90 degrees can be obtained.
From the last two runs - No. 3 and No. 4, the results in the
fourth quadrant for from 0 to -90 degrees can be obtained.
Two forms of variations of the carriage (advance) speed
and the propeller rotational rate have been considered and
thoroughly investigated in the present study. They are the
sinusoidal variations and the trapezoidal variations as sketched
in Figure 7.

Figure 7 Two forms of variations sinusoidal & trapezoidal.


The advantage of the sinusoidal variations is that all of its
higher derivatives are smooth functions of time. However, it
has a high rate of change at its two shoulders. The trapezoidal
variations have the advantage of a () constant rate of change
for the speed or the rotational rate over the whole range of one
test run, but may suffer from the non-continuity of their
derivatives at the beginning and end, and also in the top region.

For the first quadrant (test runs No. 1 and No. 2), the
towing carriage is travelling in the normal towing direction,
which we call the positive direction as shown in the sketch in
Figure 8.

+n
+Va

In order to make the quasi-steady open water test technique


valid, the following assumptions have been made. The basis of
the assumptions will be discussed in the following sections in
more detail. The test results will further prove the validity of
these assumptions.
An open or ducted propeller consists basically of lifting
surfaces (the blades and the duct). When varying the advance
ratio of the propeller, the angle of attack of the flow to the
propeller blades varies, resulting in a change of the strength of
the bounded vortex . The fact is that if the varying of the shaft
rotational rate or the towing speed is infinitely slow, meaning
that,

Figure 8 Sketch of test set-up for the first quadrant tests.


For the fourth quadrant (test runs No. 3 and No. 4), we
have studied the following two possibilities. One is shown in
the sketch in Figure 9 with exactly the same set-up used for the
first quadrant test but towed by the carriage in the reverse
direction. The advantage of this method is that the whole set-up
remains the same as for the first quadrant, except for the towing
direction of the carriage. The drawback is that the flow goes
first over the open water test POD housing and strut before it
reaches the propeller. The influence of the wake from the strut
needs to be studied carefully.

+n
-Va

Figure 9 Sketch of test set-up for the fourth quadrant tests


option 1
The other way of carrying out the fourth quadrant tests is
to reversely install the propeller on the shaft (in fact, to fit each
blade with 180o deflection angle, see Figure 6) and to reverse
the shaft rotational direction too, as shown in the sketch in
Figure 10. By doing so, the propeller is in the upstream of the
POD and the strut. The drawback is that the test cap is in the
downstream of the propeller slip stream. The drag of the test
cap is not easily subtracted from the measured thrust on the
propeller shaft in order to obtain the pure propeller blade thrust
from the measurements. In addition, reversely-fitting the
propeller will result in more uncertainties to the system and the
test results. It also costs extra preparation time.
reversely fitted propeller
(each blade deflects 180o)

-n
+Va

Figure 10 Sketch of test set-up for the fourth quadrant tests


option 2

then,

This means,
unsteady

steady.

Practically, the variation of the shaft rotational rate and the


towing speed of the carriage cannot be infinite slow. So we
need the following assumption with regard to the unsteadiness
of the flow.
Assumption I:
The variation of the shaft rotational rate and
the variation of the towing speed is so slow that the
hysteresis effect due to the unsteadiness of the flow is very
limited. Therefore averaging the increasing and decreasing
(Va and n) parts of the test results at the same value
represents the steady state test result at this value,
assuming the deviations are linear.
The viscous effects of the fluid should also be considered
during the tests. Around the propeller design condition, the
correct simulation of the viscous effects is ensured by a high
shaft rotational rate so that the Reynolds number of the
propeller blade is higher than a critical value, as we often do for
conventional propeller open water tests. At off-design
conditions, due to the high angle of attack and also the
separation of the flow on the blade, Reynolds effects become
less dominant, and thus a somewhat lower shaft rotational rate
would be allowed. The extreme situations are at +90o and -90o
of the hydrodynamic inflow pitch angle , where the rotational
rate of the shaft has to be zero.
The biggest influence of the viscous effects on the quasisteady test technique can be the hysteresis effect due to flow
separation and re-attachment at off-design conditions. The flow
separation and re-attachment usually do not occur at the same
advance ratio during increasing compared to decreasing the
towing speed or the shaft rotational rate. These effects can be
clearly seen in the later sections of this paper with the test
results. To make the quasi-steady test technique valid, we need

the following additional assumption with regard to the flow


separation at off-design conditions.
Assumption II: In the instable regime where the flow
separates and re-attaches, the average of the increasing and
decreasing (Va and n) parts of the test results at the same
value represents the steady state test result at this value in
this regime.
HYSTERESIS EFFECTS
Three hysteresis effects have been identified during a
quasi-steady propeller open water test. They are the hysteresis
effect due to the mass and mass moment of inertia of the
propeller, including also the added mass effects; the hysteresis
effect due to the unsteady hydrodynamic flow around the
propeller, mainly the vortex shedding to the propeller wake,
both spanwise and also chordwise; and the hysteresis effect due
to flow separation and re-attachment.
Mass and mass moment of inertia
Since the way to increase the carriage speed or the shaft
rotational rate in the increasing part is the same as in the
decreasing part, both for sinusoidal and also for trapezoidal
variations, the rate of change (acceleration or deceleration) are
exactly the same, but with a different sign. The hysteresis
effects of this part cancels perfectly with each other. However,
large hysteresis effect should be prevented. This can be
achieved by using light materials for the propeller, such as
aluminium.
This part of the hysteresis effect is seen as the additional
thrust and torque on the shaft resulting from the acceleration
and deceleration,

where m and I are the mass and the mass moment of inertia,
respectively, where subscript p denotes the propeller and a
denotes the added mass effect. For stock propeller No. 7216R
and by using the maximum acceleration and deceleration
during the present tests, these additional thrust and torque are
estimated and listed in Table 3 for indication.
Table 3 Thrust and torque levels due to acceleration and
deceleration mass and mass moment of inertia effects.
Indicative values for Propeller 7216R
unit
~ 0.05
N
T
~ 0.01
Nm
Q
Compared to the hydrodynamic thrust and torque levels of
this model propeller during the tests, these values are rather
small although it is not negligible. Even if a bronze propeller is
tested by using the present test technique, no large hysteresis

effect is expected. Attention may need to be paid when a quasisteady test technique is used for open water tests of the total
unit performance of an azimuth thruster or a podded propulsor
(POD). Due to the total mass and the added mass of a thruster
or a POD, the hysteresis effect can be much larger.
Unsteady flow
Since an open propeller or a ducted propeller consists of
mainly lifting surfaces, the hysteresis effect due to unsteadiness
of the flow comes mainly from the memory effect of the wake
system. The hydrodynamic unsteadiness is governed by the
Strouhal number and can be expressed as the reduced frequency
for the present study,

where
is the period of one test run with sinusoidal or
trapezoidal variations.
For the present model tests with stock propeller No. 7216R
and with the achievable longest period of each test run in the
DT of MARIN, the reduced frequency k is around 0.0003 to
0.0030. These values are regarded as very small which will not
result in any significant unsteady force and moment to the
system that will finally be measured by the sensors.
Flow separation and reattachment
The largest hysteresis effect can be expected from the flow
separation and reattachment at off-design conditions because
the hydrodynamic forces and moments differ significantly for
flows with and without separation.
It is difficult to quantify the influence of the flow
separation and reattachment. However, experience from model
tests of e.g. oscillating fins (fin stabilizer) and unsteady
azimuthing tests of thrusters or POD shows that the average
values of the hydrodynamic forces and moments does represent
the steady state test values. Observation and analysis of the raw
data signals of steady tests also show that a strong oscillating
flow in the regime of separation and reattachment results in the
measured forces and moments jumping between the values of
flows with and without separation.
This effect will be clearly shown later in the section on the
test results of this paper (see Figure 23), which occurs mainly
in the area close to = +90o and around = -10 o to -60 o at
design pitch setting. When the pitch is deflected, the area will
also be shifted.
PROPERTY OF SENSORS AND CENTRIFUGAL FORCE
Natural frequency and sensor properties
Good sensor properties are important factors to guarantee
the quality of the measurements. Two requirements are often
working against each other - the static accuracy and the
dynamic response. For accurate static measurements, sensors
need to be elastic enough. However, the sensors must also be

stiff enough so that the natural frequency of the system is high


enough to ensure that the sensors do measure physical
phenomena.
A sketch is shown in Figure 11 for a typical response of a
sensor system to the impact source which is shown as a step
function in the small graphs. Both the amplifying effect in the
area close to its natural frequency fn of the system and the
damping effect in the high frequency range should be avoided.
Source
Response

Value

Value

Source
Response

X
Response (log)

Time t

X
Time t

Figure 12 Computer model for finite element analysis of the


natural frequency, with assumed added mass of water.

Source
Response

Value

key blade

fn
Frequency f

Time t

Figure 11 Sketch of sensor properties on the measured


values.
A sensor system is a mass-spring-damping system
(Hagesteijn et al 2012). The natural frequency of the system is
not only determined by the sensors itself but also the propeller
mass and mass moment of inertia, including the effect of added
mass. To determine accurately the effect of added mass of a
propeller is difficult. An approximation has been made by
adding a ball of water to the blades as shown in Figure 12. In
hindsight the added mass effect is likely over-exaggerated by
this method providing a conservative estimation on the natural
frequency of the system.
Some examples of the blade deformations are shown in
Figure 13 and Figure 14. The calculated natural frequencies of
the system for the first mode are listed in Table 4.
Since the shaft rotational rate is around 900RPM, meaning
15Hz, these natural frequencies of the propeller and its blades
are considered high enough to obtain reliable results.

Figure 13 Blade deformation in propeller thrust direction


(at 104.9 Hz).

key blade

Table 4 Calculated natural frequencies of the sensors with


MARIN stock propeller 7216R (made of aluminium)

Forces / moments
Propeller shaft thrust
Propeller shaft torque
Blade spindle torque

Natural frequencies (first mode)


104.9 Hz
108.4 Hz
251.4 Hz

Figure 14 Blade deformation in blade spindle torque


direction (at 251.4 Hz).

Blade spindle torque caused by centrifugal force


In order to remove the spindle torque induced by blade
centrifugal force and obtain the pure hydrodynamic torque, one
test has been carried out in air for each pitch setting of the
propeller by slowly varying the shaft rotational rate. The
measured results are filtered and fitted with a quadratic curve.
Figure 15 shows the spindle torque correction for the
centrifugal force. Because it is a low-pass frequency filter of
10Hz, the gravity effect is clearly seen in the low frequency
range from the shaft rotational rate of 0 to 600RPM, although it
is much smaller than the effect of the centrifugal force. This is
partly because of the light material, aluminium, which is used
for manufacturing the blades.
These measurements are also checked on every pitch
settings by using the CAD model of the blade. Good
agreements have been found.

Figure 15 Spindle torque measured in the air (centrifugal


force induced blade spindle torque) and the correction,
filtered at 10Hz.

Figure 17 Development of the wake from the strut (first


plot) to the propeller disc (last plot), where Lpp is the total
shaft length of the open water test set-up.
TEST SET-UP WAKE FLOW AND HUB CORRECTIONS
Figure 16 Flow velocity field on the central plane of the
open water test set-up, towed reversely without operating
propeller.

Wake of the test-setup and CFD calculations


For the fourth quadrant with test option 1, as proposed and
shown in Figure 9, the test set-up will be towed in the reverse
direction. The influence of the POD and the strut of the open

water test set-up has been studied by using CFD calculations.


MARINs in-house RANS code PARNASSOS has been used
for these calculations. All calculations are carried out at -4m/s.
The results are shown in Figure 16 and Figure 17.
The strongest influence is on the central plan of the test setup behind the strut in the peak of the wake. Figure 16 shows the
development of the peak wake flow from the strut when there is
no propeller in operation. This is rather similar to the nominal
wake of a ship. It is seen that the wake peak reduces rather
rapidly after the end of the strut. At about 2 to 3 chord lengths
of the strut, the flow deficit is reduced to smaller than 10% of
the nominal velocity. In addition, the peak is very narrow as
shown in the plots in Figure 17, starting from the strut to the
propeller disc.
On the propeller disc, the flow deficit is less than 5% in the
narrow wake peak. Taking into consideration of the small over
speed resulted by the displacement effect of the shaft and the
hub, the mean advance velocity on the propeller disc is not
really affected by the strut. This concludes that no speed
correction is necessary to be applied to all of the reverse towing
tests with the present test set-up.
Test cap corrections
The hydrodynamic drag on the test cap must be subtracted
since there is no sensor to measure the force on the cap
separately. The drag is determined by a pre-defined drag
coefficient,

where S is the frontal area of the test cap


.
To determine the drag coefficient, tests have been carried
out by using a dummy hub with the test cap at both positive and
negative carriage speeds and at positive and negative shaft
rotational rates.
It is known that the drag coefficient changes with the
Reynolds numbers for difference towing speeds. It is found that
the drag coefficient for a negative carriage speed is higher than
that for a positive carriage speed, which are both not sensitive
to the shaft rotational rate. This can be explained by the fact
that the flow at the end of the cap is separated during the
reverse towing tests, which results in more pressure drag on the
test cap than that in the normal towing direction. Practically, a
constant drag coefficient is applied for the whole speed range.
For the present test set-up with the test cap, a value of 0.13 has
been found for the CD when towing in the normal direction.
During the reverse towing with operating propeller, the
situation becomes more complicated. When a propeller has a
positive pitch setting and rotates in the positive direction, the
propeller slipstream is working against the inflow, leaving the
test cap in the dead water area of the blocked flow behind the
propeller. Even if the propeller blade is set to negative pitch at
0.7R, it is often the case that the blade pitch at the root remains
positive.
In order to simulate the flow and to find out the influence
of the test cap, an additional CFD calculation has been carried

out by applying a negative thrust to the flow with an actuator


disc model. The results are shown in Figure 18.
Indeed, the flow behind the propeller is fully blocked by
the operating propeller, leaving a separation zone where the
complete test cap is inside. The calculated results also show
that the total drag force on the test cap is very small.
Based on the investigations above, it is decided to apply
the following drag coefficients for the test cap corrections as
listed in Table 4, when reverse towing is used.

Figure 18 Flow velocity field on the central plane of the


open water test set-up, towed reversely with propeller
operating against the flow.
Table 4 Test cap corrections used for the data analysis
Towing condition
Drag coefficient CD
positive carriage speed
0.13
negative carriage speed
0.00

DATA ACQUISITION, REDUCTION & PRESENTATION


The measured propeller shaft thrust and torque, and the
blade spindle torque, are non-dimensionalized by the relative
velocity at 0.7R radius defined as,

with the propeller thrust coefficient defined as,

the propeller torque coefficient defined as,

and the blade spindle torque coefficient defined as,

where, the positive directions of the propeller shaft thrust,


torque and the blade spindle torque are shown in Figure 19. The
positive blade spindle torque is defined as the direction that
tends to drive the propeller to a larger pitch.
During each open water test run, 5 channels of signals have
been sampled and recorded. These are the speed of the carriage
Va, the shaft rotational rate n, the propeller shaft thrust T, the
propeller shaft torque Q and the blade spindle torque Qblade. All
channels have been sampled up to a frequency of 1kHz. Some
selected raw data samples are shown in Figure 20 to Figure 22.

fitted with one of the following Fourier series, respectively, and


the Fourier series coefficients have been determined up to 30
harmonics.

(12)

Figure 19 Definition of positive directions for the thrust,


torque and the blade spindle torque.
Figure 20 shows examples of the sampled carriage towing
speed variations and the shaft rotational rate variations,
following the predefined sinusoidal form variations. It can be
seen that both the towing carriage speed and the shaft rotational
rate can follow the sinusoidal curve quite well.
In the first quadrant when the propeller blade is operating
around its design point, no severe flow separation is expected.
The measured data show rather stable values as can be seen in
the examples in Figure 21 for the thrust and the spindle torque.
When the propeller operates in the fourth-quadrant, flow
separation and reattachment occurs, resulting in a strong
oscillating flow. This is found in the sampled thrust and spindle
torque data, as examples show in Figure 22.
As the present study is not aimed at the dynamic response
of a propeller and its shafting system where higher frequencies
play an important role, it is decided to filter the raw data by a
low-pass filter with an upper-bound frequency at 10 Hz, which
is lower than the shaft frequency in the majority of time in
order to remove the possible noise coming from the bearing and
the toothed belt.
After the filtering of the raw data, the data was further
grouped and averaged over each degree of the hydrodynamic
pitch angle , forming a set of discrete data sets of 181
elements from -90 to +90 degrees. Thereafter, each set of the
data the propeller thrust coefficients, the propeller torque
coefficients and the blade spindle torque coefficients has been

Figure 20 Typical examples of the sampled carriage speed


(top) and shaft rotational rate (bottom), Va max = 4m/s, nmax =
1100RPM, sinusoidal variations.

10

Q [Nm]

Q [Nm]

Figure 21 Typical examples of the sampled thrust (top) and


blade spindle torque (bottom) during test run No. 1, at
design pitch, n=900RPM, sinusoidal variations.

Figure 22 Typical examples of the sampled thrust (top) and


blade spindle torque (bottom) during test run No. 3, at
design pitch, n=750RPM, sinusoidal variations.

RESULTS AND DISCUSSIONS

With the sinusoidal variations, both for the advance speed


and also for the shaft rotational rate, 4 test runs have been
carried out according to the procedure described in Table 2 for
the stock propeller No. 7126R at the design pitch, where the
reverse towing for the fourth quadrant has been used. The
sampled results are filtered at 10Hz, calculated into coefficients
as defined by Equation 9, 10 and 11 and their curves plotted in
Figure 23.
Also plotted in this figure are the conventional steady test
results by dots (circles, squares and triangles) with their 95%
occurrence intervals indicated by the +s. A 95% occurrence
interval is the interval where 95% of the sampled signals are
within this interval, while 5% are outliers. Also shown in this
figure are the dark and light coloured curves for each thrust or
torque, representing the accelerating (dark) and decelerating
(light) parts of the sinusoidal variations, respectively.
Hysteresis effect can be clearly seen between the dark and light
coloured curves for each thrust and torque. This occurs mainly
in the area close to +90o and the area around -10o to -60o for
this pitch setting.

To verify the quasi-steady propeller open water test


technique, two-quadrant open water tests for MARINs stock
propeller No. 7216R have been carried out, both by the
conventional method as well as by the quasi-steady method.
Two typical pitch settings have been investigated, being the
design pitch setting and a negative pitch setting by deflecting
all blades with -35o from their design pitch.
For the conventional tests, the thrust, torque and the blade
spindle torque have been measured from -90o to +90o with a
step of 10o.
For the quasi-steady tests, both sinusoidal and trapezoidal
variations of the speed and the shaft rotational rates have been
studied in order to investigate the sensitivity of the results to the
test methods. Also done are the tests for the fourth-quadrant
with option 2 (Figure 10) where the propeller is reversely fitted
to the shaft and the shaft is rotating in the negative direction.
Sinusoidal variation at design pitch

11

Also seen in Figure 23 is that the width of the 95% interval


from the conventional steady open water tests at discreet points
is rather similar to the fluctuation in amplitude of the measured
thrust or torque during the quasi-steady tests. In area where
most of the fluctuations have been measured by the quasisteady tests, the widths of the 95% intervals of the conventional
steady test are also larger.

By fitting the measured data with a Fourier series as given


in Equation 12, the complete two-quadrant propeller open water
characteristics are expressed by Fourier coefficients up to 30
harmonics. The re-generated thrust and torque coefficients from
the Fourier series are plotted in Figure 24, together with the
steady measurement points with their 95% occurrence intervals.
It is seen that the quasi-steady test results match the
conventional steady test results very well.

2
50CQBlade
10CQProp

1.5

CTProp

50CQBlade/10CQProp/CTProp

0.5

-0.5

-1

-1.5

-2
-90

-75

-60

-45

-30

-15

0
Beta [deg]

15

30

45

60

75

90

Figure 23 Comparison of the filtered raw data from quasi-steady tests to the steady test results with their 95% occurrence
intervals, sinusoidal variations, at design pitch.
One of the most important observations of the results is
that only strong hysteresis effects have been found in the area
where flow separation occurs. This proves that the hysteresis
effects from both the mass, the mass moment of inertia and the
unsteadiness of the flow are very small, as analyzed in the
previous sections. In addition, the average of the test results in
the accelerating and decelerating parts of the tests, even in the
area where the flow separates and reattaches, equals to the
steady test results. These prove that Assumption I and II made
at the beginning of this paper are reasonable and valid
assumptions. The same observations were also found for the
other pitch settings, for the trapezoidal variations and for the
propeller reversely-fitted tests for the fourth quadrant.
Sinusoidal variations at negative pitch setting
Also investigated in the present study is the quasi-steady
tests for negative pitch setting by deflecting the blade pitch
angle by -35o from its design pitch angle. It should be

mentioned that at this pitch setting, although the pitch at 0.7R is


negative, the pitch at the blade root remains slightly positive.
The filtered quasi-steady test measurements are plotted in
Figure 25 together with the steady test results with their 95%
occurrence intervals. The dark and light coloured curves for
each thrust and torque represents the accelerating (dark) and the
decelerating (light) parts of the sinusoidal variations,
respectively. At this test condition, large fluctuations of the test
results are only seen for the blade spindle torque in the area
between -40o to -90o. No strong hysteresis effect has been
found in the whole range of the two-quadrant open water
characteristics.
After the Fourier fitting of the filtered raw data, the thrust
and torque coefficients are regenerated from the Fourier series
and plotted together with the conventional steady test results in
Figure 26. The quasi-steady test results agree perfectly with the
conventional steady test results.

12

2
50CQBlade
10CQProp

1.5

CTProp

50CQBlade/10CQProp/CTProp

0.5

-0.5

-1

-1.5

-2
-90

-75

-60

-45

-30

-15

0
Beta [deg]

15

30

45

60

75

90

Figure 24 Comparison of Fourier series fitted curves to steady test results with their 95% occurrence intervals, sinusoidal
variations, at design pitch.
2
50CQBlade
10CQProp

1.5

CTProp

50CQBlade/10CQProp/CTProp

0.5

-0.5

-1

-1.5

-2
-90

-75

-60

-45

-30

-15

0
Beta [deg]

15

30

45

60

75

90

Figure 25 Comparison of filtered raw data from quasi-steady test to steady test results with their 95% occurrence intervals,
sinusoidal variations, pitch deflected -35o from design pitch.

13

2
50CQBlade
10CQProp

1.5

CTProp

50CQBlade/10CQProp/CTProp

0.5

-0.5

-1

-1.5

-2
-90

-75

-60

-45

-30

-15

0
Beta [deg]

15

30

45

60

75

90

Figure 26 Comparison of Fourier series fitted curves to steady test results with their 95% occurrence intervals, sinusoidal
variations, pitch deflected -35o from design pitch.

2
50CQBlade
10CQProp

1.5

CTProp

50CQBlade/10CQProp/CTProp

0.5

-0.5

-1

-1.5

-2
-90

-75

-60

-45

-30

-15

0
Beta [deg]

15

30

45

60

75

90

Figure 27 Comparison of filtered raw data from quasi-steady test to steady test results with their 95% occurrence intervals,
sinusoidal variations, at design pitch, propeller reversely fitted to the hub.

14

The reverse towing method for the fourth quadrant open


water characteristics, as shown in Figure 9, has been used for
the fourth quadrant tests of the propeller at this negative pitch
setting.
Propeller reversely fitted for 4th quadrant
As also proposed for the fourth quadrant open water tests option 2 (Figure 10), fitting the propeller reversely to the shaft
and rotating the shaft in opposite direction can be used too. This
method has been tried, by keeping the hub as it is while rotating
each blade 180o on the CPP hub, as shown by the photo in
Figure 6. The test results are plotted in Figure 27.
When comparing Figure 27 to Figure 23 in the fourth
quadrant, a clear difference can be seen. This difference is seen
also at bollard condition although it is rather small. This
difference is believed to be caused by the test cap which is in
the slipstream of the propeller when the propeller is reversely
fitted. The flow in the slipstream of a propeller is so
complicated that the drag and torque on the test cap cannot
easily and accurately be subtracted. In addition, deflecting the
propeller blades by turning each blade 180 o may result in
additional uncertainties to the pitch setting of the propeller.
In general, reversely fitting the propeller on the shaft is not
advised unless the drag and torque on the test cap is able to be
measured by a separate sensor independently.

Trapezoidal variations
Until now, only tests with sinusoidal variations have been
discussed in detail. However, during the study, the same
amount of tests with trapezoidal variations have been carried
out as well. The test results at the design pitch setting are
shown in Figure 28 while the test results at the negative pitch
setting by deflecting the blades with -35o are plotted in Figure
29, together with the conventional steady test results.
By comparing Figure 28 and Figure 29 to Figure 23 and
Figure 25, respectively, it is seen that the test results are very
close to each other. Further investigations show that the only
deviations occur in the region where the derivatives of the
variations are not continuous. However, the deviations are
within the uncertainties of the test itself.
This concludes that the quasi-steady open water test is not
very sensitive to the variation in form of the towing carriage
speed and of the shaft rotational rate. Making a perfect
variation of the towing carriage speed or the shaft rotational
rate is therefore not necessary.
However, in order to prevent discontinuity of the variations
and their derivatives, smooth variations, e.g. sinusoidal
variations, are recommended rather than a method such as the
trapezoidal method.

2
50CQBlade
10CQProp

1.5

CTProp

50CQBlade/10CQProp/CTProp

0.5

-0.5

-1

-1.5

-2
-90

-75

-60

-45

-30

-15

0
Beta [deg]

15

30

45

60

75

90

Figure 28 Comparison of filtered raw data from quasi-steady test to steady test results with their 95% occurrence intervals,
trapezoidal variations, at design pitch.

15

UNCERTAINTY ANALYSIS
In order to indicate the uncertainty involved in the
measurements performed, so-called occurrence intervals are
presented in most of the figures around the steady
measurements by + symbols. These intervals do not represent
the confidence of the mean value itself, but are a very good
indication of the stability of the flow and therefore provide an
indication on the validity of the presented values in various
regimes.
The interval denoted by the + symbols is the 95% occurrence
interval. This means that during a steady measurement of
approximately 10 seconds on model scale, 95% of all samples
are within this interval. Prior to establishing the 95%
occurrence interval, all signals are filtered using a 10 Hz model
scale low pass filter. This is done to subtract noise created by
bearings, the carriage and drive belt of the set-up. The signals
which are left after filtering contain only relatively low frequent

components, which are assumed to be a result of hydrodynamic


forces only, except for the blade spindle torque where the
centrifugal force effect will be subtracted later on.
Signals treated in this manner provide a lot of information.
For example in Figure 23, the regime from = 0o to 60 o shows
both a very smooth quasi-steady signal and a very small
occurrence interval. This regime spans from bollard pull to the
normal working area and beyond. The smooth lines and the
very small occurrence intervals tell us that the flow over the
propeller is very stable.
Starting from = 60o, a small hysteresis effect can be
identified in the quasi-steady measurement lines which starts to
grow rapidly beyond = 80o. The light coloured curves
indicate increasing during the test and the dark coloured
curves indicate decreasing during the test. The hysteresis
above = 80o is very likely due to the flow separation and
reattachment, locally over the blades of the propeller.

2
50CQBlade
10CQProp

1.5

CTProp

50CQBlade/10CQProp/CTProp

0.5

-0.5

-1

-1.5

-2
-90

-75

-60

-45

-30

-15

0
Beta [deg]

15

30

45

60

75

90

Figure 29 Comparison of filtered raw data from quasi-steady test to steady test results with their 95% occurrence intervals,
trapezoidal variations, pitch deflected -35o from design pitch.
Below = 0o a small hysteresis effect seems visible as
well, but below = -10o the flow becomes rapidly unstable.
The quasi-steady measurements show very large fluctuations in
this area, especially around = -30o. In this region the advance
velocity is reversed while the rotation rates are relatively high,
resulting in flow reversal and large turbulence. Again the quasisteady measurements show a very good agreement with the
occurrence intervals of the steady tests.
In order to judge the accuracy of the results further, the
Fourier solution of the quasi-steady results has been compared

to the 95% confidence interval resulting from variance analysis


of the conventional steady test results. Variance analysis uses
the auto covariance of a signal to determine how likely that a
found mean value is in the neighbourhood of the actual mean
value. The actual mean value can never be found since the
measurement time for that needs to be infinite, neglecting other
sources of uncertainties (Bendat and Piersol 2010).
For now it is stated that the measured signal itself is
assumed to be true (ignoring calibration uncertainty sources
etc.) and the task is to find the true mean. To do this, the

16

variance of the mean is used to calculate a 95% confidence


interval for the calculated means. Comparing these values to
the Fourier fitted values gives Figure 30 and Figure 31 for two

different regimes, where the confidence intervals are shown by


diamonds in the figures.

50CQBlade
10CQProp

CTProp

50CQBlade/10CQProp/CTProp

0.8

0.6

0.4

0.2

0
-45

-30

-15

Beta [deg]

Figure 30 Comparison of Fourier components versus 95% occurrence (+) and 95% confidence intervals () for turbulent
regime.
-0.2

50CQBlade
10CQProp

-0.4

CTProp

50CQBlade/10CQProp/CTProp

-0.6

-0.8

-1

-1.2

-1.4

-1.6

60

75
Beta [deg]

90

Figure 31 Comparison of Fourier components versus 95% occurrence (+) and 95% confidence intervals () for turbulent
regime.

17

In Figure 30, the turbulent region from = -45 to 0 deg is


presented. The large turbulent and low frequent motions cause
a relatively big uncertainty of the mean value. In general, an
increased measurement time would lead to a decreased
confidence interval. Since the steady measurements take place
for about 10 seconds and the quasi steady measurements sweep
through these points in a similar or slightly faster pace, one
might expect the quasi-steady measurements to be
approximately within these 95% confidence intervals or slightly
outside. As can be observed, this is the case, especially in the
most turbulent regions.
It has to be noted that the calculated value of variance of
the mean itself is very inaccurate in itself. Only changing the
calculation method (i.e. how the auto covariance is determined)
might alter the confidence interval by a factor 2 or 3 depending
on random factors. Variance of the mean by itself shows a lot of
scatter, only a large amount of these values indicate a reliable
result. In other words, a few intervals missed by the quasi-static
measurements might be random scatter rather than a bad result
from the quasi-steady measurement method.
However, in Figure 31 it can be seen that the points around
= 80o do miss the calculated interval. Here it seems no
random scatter, but a systematic deficiency occurs. The Fourier
fit likely has too few components to correctly fit the sharp bend
the actual data contains at is 80o.
The 95% occurrence intervals have shown very good
agreement with the raw data. In all regions the steady
measurements show an equally large 95% occurrence interval
comparable to the local fluctuation and/or the hysteresis found
during the quasi-steady tests. This indicates that the quasisteady measurement method contains the same information as
the steady measurements. On top of that, the raw data could be
used to judge what physical phenomena are occurring, random
turbulent events or hysteresis.
In summary, confidence intervals have been used to judge
the quality of the Fourier fit of the results. It is found that the fit
agrees very well with the 95% confidence intervals except
when the original data seems to contain sharp bends, which are
filtered out by the Fourier fitting, not the quasi-steady method.
CONCLUSIONS
A quasi-steady open water test technique has been
successfully and thoroughly investigated in the present study
for two-quadrant open water tests of a controllable pitch
propeller at various pitch settings. The method has been proven
to be a reliable and accurate test technique when proper
measures have been taken, such as the weight of the propeller,
the sensors properties, etc. The quasi-steady test technique
provides the same accuracy for the open water characteristics of
a propeller as the conventional steady test technique. In
addition, this method reduces the tank test time dramatically by
reducing the number of test runs.
For a typical two-quadrant open water test of a propeller,
only four test runs are required by using the quasi-steady test
technique. Compared to the conventional steady tests, with a

step of each 5o on the hydrodynamic pitch angle , it reduces


the tank test time by a factor of 8 to 10. This makes it possible
for the complete two-quadrant open water tests of the
controllable pitch propellers of the Wageningen C- and D-series
for a typical set of 12 pitch settings of each propeller, but still
within affordable cost.
In addition to this, the sampled signals in the normal
operational range from J=0 to KT=0 by using quasi-steady test
technique show very small fluctuations and almost no
hysteresis effect. The results are identical to that from the
conventional steady test method. This implies that the quasisteady open water test technique may replace the conventional
steady open water test technique in the near future for all
propeller open water tests. This may include also bronze
propellers, thrusters, PODs, pump jets, surface piercing
propellers, etc. as far as careful measures have been taken.
The study further shows that the test results are not so
sensitive to the form of the variations of the towing speed and
the shaft rotational rate. This means that perfectly following a
predefined variation form for the towing speed and the shaft
rotational rate is not really necessary. However, a variation
form such as a sinusoidal form is highly recommended due to
the continuity of all its higher derivatives.
ACKNOWLEDGMENTS
The authors are grateful for the valuable support from all
participants to the Joint Industry Project (JIP) the Wageningen
Propeller C- and D-series.
REFERENCES
Bendat J.S. and Piersol A.G. (2010). Random Data:
Analysis and Measurement Procedures, fourth edition, Johan
Wiley & Sons, Inc., Hoboken, New Jersey.
Chen B. and Stern F. (1999). Computational Fluid
Dynamics of Four-Quadrant Marine-Propulsor Flow, Journal
of Ship Research, Vol. 43, No.4, Dec. 1999, pp.218-228.
Chu C., Chan Z.L., She Y.S. and Yuan V.Z. (1979). The 3bladed JD-CPP series Part 1, Proceedings of the 4th LIPS
Propeller Symposium, Drunen, The Netherlands.
Hagesteijn G., Brouwer J. and Bosman R. (2012).
Development of a Six-Component Blade Load Measurement
Test Setup for Propeller-Ice Impact, Proceedings of the ASME
31st International Conference on Ocean, Offshore and Arctic
Engineering OMAE2012-84192, Rio de Janeiro, Brazil.
Hampton G.A. (1980). Four Quadrant Open Water
Characteristics of Controllable Pitch Propeller 4739 Designed
for LSD-41 (Model 5367), DTNSRDC/SPD-0049-12.
Holtrop J. and Hooijmans P. (2002). Quasi-Steady Model
Experiments on Hybrid Propulsion Arrangements, Group
discussion A.1: New Experimental Techniques and Facilities,
Proceedings of the 23rd ITTC Volume III, pp. 751-752.
Ito M., Yamasaki S., Oku M., Koizuka H., Tamashima M.,
and Ogura M. (1984). An Experimental Study of Flow Around
CPP Blade (3rd Report): Measurement of CPP Blade Spindle
Torque, Journal of the Kansai Society of Naval Architects, No.
192, pp.81-91.

18

ITTC (2008). ITTC Recommended Procedures and


Guidelines: Testing and Extrapolation Methods Propulsion,
Propulsor Open Water Test, ITTC document No. 7.5-02-0302.1 revision 02, pp.1-10.
Jessup S., Donnelly M., McClintock I. and Carpenter S.
(2009). Measurements of Controllable Pitch Propeller Blade
Loads under Cavitating Conditions, Proceedings of the First
International Symposium on Marine Propulsors, Trondheim,
Norway, June.
Kuiper G. (1992). The Wageningen Propeller Series,
MARIN Publication 92-001, published on the occasion of its
60th anniversary, Wageningen, the Netherlands.
Koushan K. Spence S and Savio L. (2011). Ventilated
Propeller Blade Loadings and Spindle Moment of a Thruster in
Calm Water and Waves, Proceedings of Second International
Symposium on Marine Propulsors, Hamburg, Germany.
van Lammeren W.P.A. (1936). Resultaten van
Systematische Proeven met Vrij-varende 4-bladige Schroeven,
type A4.40, Het Schip 18, No. 12 pp. 140-144, N.S.M.B.
publication No. 21.
van Lammeren W.P.A., van Manen J.D. and Oosterveld
M.W.C. (1969). The Wageningen B-screw Series,
Transactions of SNAME, Vol. 77, pp. 269-317.
van Lammeren W.P.A., van Manen J.D. and Oosterveld
M.W.C. (1970). The Wageningen B-screw Series, Schip en
Werf, No. 5, pp. 88-103 and No. 6 pp. 115-124.
van Manen J.D. (1954) Open Water Test Series with
Propellers in Nozzle, International Shipbuilding Progress, Vol.
1.
MARIN (1984) Vier_kwadrant Vrijvarende-SchoefKaracteristieken voor B-series Schroeven, Fourier-Reeks
Ontwikkeling en Operationeel Gebruik.
Oosterveld M.W.C. (1970) Wake Adapted Ducted
Propellers, Thesis of Technical University Delft, N.S.M.B.
Publication No.. 345.
Pronk C. (1980). Blade Spindle Torque and Off-Design
Behaviour of Controllable Pitch Propellers, Dissertation to the
Technical University Delft, The Netherlands.
Queen C.G. (1981). Four Quadrant Open Water
Characteristics of Controllable Pitch Propeller 4837 Designed
for MCM (Model 5401), DTNSRDC/SPD-0983-04.
Roddy R.F., Hess D.E. and Faller W. (2006). Neural
Network Predictions of the 4-Quadrant Wageningen Propeller
Series, NSWCCD-50-TR-2006/004, April, West Bethesda,
Maryland.
Troost L. (1938). Open water Test Series with Modern
Propeller Forms, Transactions of North East Coast Institute of
Engineers and Shipbuilders, pp. 321, N.S.M.B. publication No.
33.
Troost L. (1940) Open water Test Series with Modern
Propeller Forms, Part 2, Transactions of North East Coast
Institute of Engineers and Shipbuilders, pp. 91, N.S.M.B.
publication No. 42.
Verhulst M. and Hooijmans P. (2010). Quasi-Steady
Model Propulsion Measurements for Complex Propulsion
Systems, Proceedings of the 11th International Symposium on

Practical Design of Ships and Other Floating Structures,


pp.695-702, Rio de Janeiro, Brazil.
Yazaki A. (1962). Design Diagrams of Four-Bladed
Controllable-Pitch Propellers, Journal of Zosen Kyokai, Vol.
112, November.

19

También podría gustarte