Está en la página 1de 19

MANOEUVRING SIMULATION

OF SHIPS WITH AZIMUTHAL PROPULSION UNITS


Stefano Brizzolara
University of Genova, Dept. of Naval Architecture and Marine Technologies
Via Montallegro, 1 16145 Genova, Italy
Email: Brizzolara@dinav.unige.it

ABSTRACT:
The paper presents the main details and results of the new module recently
incorporated in a numerical ship manoeuvring simulator to deal with ships having
azimuthal propulsion units. The complex problem of the hydrodynamic forces acting
on an azimuthal thruster in oblique flow has been modelled with a practical effect
decomposition approach: the propeller forces and torque in inclined flow, the
hydrodynamic forces on the strut and nacelle and the interaction effect of the
pulling propeller slip-stream on strut hydrodynamic behaviour. The new modules of
the simulator are also discussed in view of their numerical implementation into an
modern simulation environment.
The simulator is applied to an existing a ro-ro passenger ferry with twin
conventional shaft lines and propellers, whose main propulsion has been ideally
converted into one using two azimuthal thrusters of about 11 MW each: a typical
contemporary design case. The simulated ship performance in terms of standard
manoeuvres, i.e. turning circles and zig zag, are compared with the full scale
results of the original ship having a conventional shaft line arrangement and rudder
layout. Interesting conclusions are drawn at the end from the comparison of the
results obtained by the simulation and those known for the existing ship at full
scale.

1 INTRODUCTION
The simulation of ships with azimuthal propulsion units is a very actual and
challenging task. In fact, while several azimuthal propulsion units have been
installed in different commercial and cruise ships since many years by now, it is
still difficult for a naval architect to find even a simplified prediction method to
evaluate the manoeuvring performance of a ship with such thrusters. The
reason of this lack of studies is not known and, perhaps, is hidden under the
initially limited number of applications of these type of propulsion, mainly
confined to specialised ships types, such as working boats, tugs, small ferries,
etc. Under the term azimuthal propulsion unit it is intended to comprise any
type of mechanical rudder-propellers, usually able to deliver power up to 5-6
MW or the larger electrical or hybrid mechanical-electrical POD units which can
reach input power of 20 MW and more. In spite of the particular design of each
different concept, the common characteristics of these modern propulsion
systems is that they are able to direct the propeller thrust in the complete 360
range on a horizontal (or sub-horizontal) plane, by rotating the propeller and
thruster body around a (sub-) vertical axis.
Nowadays, the world wide diffusion of azimuthal propulsion devices has very
much increased and with them the interest on their performance especially for
application for the main propulsion into larger ships, having bigger installed
power. Various research institutions worldwide, then, have started to develop
simplified theoretical models to simulate the hydrodynamic actions of these type
of propulsion devices or are actually studying the details of some specific
hydrodynamic problem. It is true, in fact, that many ship manoeuvring simulators
around the world can nowadays consider ships with azimuthal propulsion units,
but very seldom the theoretical/mathematical models have been released in the
open literature.
In this context, it was decided to extend the ship manoeuvring simulator
developed at the University of Genova, to consider this new interesting and
modern propulsion typology. The simulator was initially developed as an add-on
module of the DINAV ship propulsion system simulator [1], with the aim to study
the transients of the main diesel engines due to ship evolutions. After some
successful applications [1] of the complete dynamic simulator, the ship
manoeuvring module was evolved in such a way to be independently used as a
stand alone tool for preliminary or final design evaluation, just for manoeuvring
simulation with a less sophisticated engine simulation module. The correlation
with available full scale tests on contemporary commercial and navy ships is in
general very satisfactory [2]. So the manoeuvring simulator was further
extended in order to use the most recent formulations for the calculation of the
hull hydrodynamic derivatives with any combinations of conventional shaft line
arrangements with FP or CP propeller and rudders type [3].
The implementation of the numerical simulator is done in MATLABSIMULINK environment, a state of the art standard for the simulation of
dynamic systems, which permits a number of conceptual and structural
advantages in the writing of the code, against other less dedicated
environments.

CONVENTIONAL PROPULSION

------------------------------------------------------AZIMUTHAL PROPULSION

Figure 1 High level functional diagram of the Conventional Ship Manoeuvring Simulator
(above) and the its variation in case of Electrical Azimuthal Propulsion Units (below).

3 BASICS OF THE MANOEUVRING SIMULATOR MODEL


The present paragraphs intends to briefly introduce the basic differential
equation used to simulate the ship dynamic in the horizontal plane. As appears
by the comparison of the high level flow charts of the two simulators in Figure 1,
it is clear that the numerical method which describes the hydrodynamic of the
hull during evolution is maintained unaltered between the ship simulator with a
conventional propulsion system and that with an azimuthal propulsion system.
The main changes, in fact, regard the propulsion modules which will be
described in more detail in the following sub-paragraphs. The azimuthal
propulsor units are incorporated in the simulator as a series of alternative
modules, instead of the conventional propellers and rudders. In fact, with
reference to the
high level
diagram of Figure 1, the
azimuthal propulsion unit blocks
are
equivalent
to
the
conventional propulsion blocks
from the interface point of view
with the ship dynamic system.
So the modules can be easily
interchanged in the simulator,
especially in the case of main
diesel engines (and mechanical
azimuthal thruster). A minor
further modification is needed
only in case of electrical driven
units for what regards the
control logic of the electric
motor.
The conventional definition of Figure 2 Reference systems and positive
velocities, forces and moments directions for forces, moments and velocities
and their positive directions components
used in the simulator model are
given in the sketch of Figure 2.
The system of differential equation solved in the ship hydrodynamic block is
that pertaining to a planar motion of a rigid body with 3 degrees of freedom:
surge, sway ad yaw, expressed in the non inertial (body fixed) {x,y,z} reference
system of Figure 2. translated in formulae:

m u vr xG r 2 = X

m ( v ur + xG r ) = Y
I r + mx ( v + r u ) = N
G
z

(1)

In (1) X, Y are the longitudinal and transversal component of the forces


acting on the ship and N is the yaw moment induced by these forces. The
instantaneous surge and sway ship speed are indicated by u and v,
respectively, while r is the yaw speed. Dots means derivatives with respect to

time, so u for instance is the surge acceleration. The forces and moment acting
on the ship can be approximately defined as a linear superposition of various
components, i.e.:

X = X H + X HA + X P + X R + X W

Y = YH + YHA + YP + YR + YW
N = N + N + N + N + N
H
HA
P
R
W

(2)

where the subscripts indicate the portion of the origin of such forces or moment.
So the H stays for hull, HA for hull appendages, P for propellers, R for rudders
(or thruster body, in case of azimuthal propulsors), W for wind and/or waves
and other external forces (towing forces, etc.).
Following the usual development, the hydrodynamic forces acting on the
underwater ship hull are though as dependent only on the instantaneous speed
and acceleration components of the ship motion and are expanded in a Talylor
series around an initial condition (usually the steady rectilinear ship motion).
Under this assumption, indicating with subscripts the partial derivative of the
forces or moment with respect to the indicated ship motion variable, the
following standard expression is used in the simulator for describing the
hydrodynamic forces acting on the hull:

X = X u R (u ) + ( X Y )vr Y rr
u
vr
v
r
H
YH = Yv v + Yr r + Yv v + Yuv uv + Yur ur + Yv v v v + Yr r r r + Yv r v r

N H = N v v + N r r + N v v + N r r + N uv uv + N ur ur + N r r r r + N rrv rrv + N vvr vvr

(3)

In which the hull resistance component R(u) is evaluated, when better data
are not available, with the Holtrop method. The hull hydrodynamic derivatives
which appears in (3) are usually evaluated using Clarke formulae for the linear
and added mass terms and Inoue formulae for the dissipative terms, as detailed
in [3], on the base of the main hull dimensional ratio and block coefficient. In
case these derivatives are known from model tests or identification procedures,
they can be directly given as input to the program, up to the fifth order.
In case of conventional propulsion system configuration, the other force
contributions, such as those of the FP or CP propeller and plain, horn or flapped
rudders and of the different appendages, are estimated in the simulator with a
different approaches better described in [3].
In case of azimuthal propulsion system, the forces developed by the thrusters
are modelled by a new method which is specifically described in the next two
sub-paragraphs, in the most common case of a pod with pulling propeller.
The simplification adopted, in this case, is to isolate the problem of the
forward propeller working in an oblique flow from that of the aft strut and nacelle
working in the propeller slip stream.
A simplified one way interaction effect of propeller flow on the strut and
nacelle hydrodynamic forces is considered as described in the following. No
other interaction effect, such as for instance the influence of the aft thruster
body on the propeller hydrodynamics, is considered in the present method.

3.1 Propeller working in an oblique flow


The theoretical description of forces developed by a propeller working in
oblique flow with high angles of attack is not an easy task: it is much more
difficult than the simulation of a propeller working in off-design conditions. Some
Perhaps the only tool capable of describing the dynamic characteristics of a
propeller in oblique flow is the CFD, with an adequate turbulence model to
adequately capture the separated flow on the blades at high inflow angles.
Moreover, in addition to the strong viscous effects, the physical phenomena,
even for a propeller working in open water, are highly non stationary. Indeed
this problem still represents a frontier of the CFD research in the naval field.
It is clear, then, that for the purpose of the more general and complex
problem of the simulation of ship manoeuvres, the dynamics of a propeller in
oblique flow needs a simplification. Some attempt to simplify the problem using
the equivalent section theory were made in the past [4] showing a generally
good correlation with experimental results, but only limited to small angle of
attacks (below 20 degrees). A similar approach is that recently proposed by
Woodward et al. [5] [6]. Using the equivalent blade section theory, he reduces
the problem of the propeller in oblique flow onto that of a propeller working in
head or astern flow. In this way without need of ad hoc tests, he can exploit the
systematic open water tests of propellers in the four quadrants, as for instance
those of the Wageningen B-series.
X

Tp

TTOT

Lp

v-KrrXp

VE

u (1u w)

Figure 3 Forces diagram for a pulling propeller of an azimuthal thrusters rotated with a rudder
angle and subjected to inflow speed VE induced by ship motions

In our approach, use is made of the results of a systematic series of propeller


tested in oblique flows [7]. In this series several propellers of the Wageningen
B-series with different pitch ratios, ranging from P/D=0.8 to P/D=1.2 and a
given expanded area ratio AE/AO=0.55, were tested at different attack angles
with respect to the uniform inflow with a step of 15 degrees, in the complete
azimuthal range 0 deg to 360 deg. Results are given in terms of propeller
absorbed torque coefficient KQ and developed force coefficients: in the
longitudinal (KT) and lateral (KL) directions. The component of the thrust are
defined with respect to the incoming flow direction forming an effective angle of
attack E with the propeller shaft axis, so the lateral component correspond to
the lift obtain from the fin effect of the propeller.
The definition of the effective inflow angle to the propeller as well as the
definition of the thrust components is well capture in Figure 3. The magnitude of

the incoming flow VE on the propeller, allows for the ship advance sway and
yaw speeds, as well as for the effective wake fraction (1-w) and is expressed
by the following relation:

VE =

[u (1 w) ] + [v K
2

r xP ]

(3)

This incoming flow is thought as uniform onto the propeller disk and is
inclined of an angle with respect to the ship longitudinal axis. The effective
oblique angle E of the incoming flow onto the propeller, rotated by the
azimuthal angle , is then found by:

E = ,

= atan (v K r r x P u (1 w) )

(4)

In both (2) and (3), a correction coefficient Kr can be adopted to reduce the
effect of yaw rate onset flow on the propeller and the hull longitudinal wake
fraction (1-w) can be expressed as a function of the ship instantaneous yaw
angle , typically through a cosine function.
Due to the fin effect of the propeller working in this oblique flow, the total
produced thrust is, in general, not aligned with the incoming flow, but rotated by
an angle measured from the geometric thruster azimuthal angle . The
deviation angle of the thrust is known when its longitudinal and lateral
components are known as a function of the propeller working conditions, i.e.:

= atan(K L K T )

(5)

Figure 4 Deflection angle of the thrust with respect to the Wageningen propeller axis, as a
function of the inflow relative angle of attack E, for different advance coefficient J.

An example of the behaviour of this angle is derived from the results of the
systematic model tests in [7] and presented in Figure 4, in the case of the
Wageningen propeller B-4-55 with P/D=1.2. For the interpretation of the graph,
the advance coefficient as a function of the magnitude of the incoming flow
speed as J = Va nD , so does not depend on the azimuthal propeller angle

relative to the flow E. From Figure 4 it can be noted that he thrust deflection
angle depends very much on advance coefficient and has an sinusoidal like
behaviour with respect to the inflow angle of attack to the propeller. The
oscillations noted for J=0 evidence the systematic error which affects the
experimental results, most probably induced by the interaction effects between
the propeller and the carrying arm used for the tests, not correctly depurated.
The propeller developed forces thus derived can be finally project on the ship
longitudinal and transversal direction to be added in the external force
calculation (2) and used in the general equation of motions (1), according the
following expression:

X P = TP cos LP sin

YP = TP sin + LP cos

(6)

For the shaft line dynamics the information about the torque requested by the
propeller in the resulting inclined flow is derived from the torque coefficient KQ
derived by interpolation of the mentioned model tests results.
3.2 Forces on the thruster body
For the thruster configuration considered in this study, i.e. with single pulling
propeller, the inflow to the thruster body is largely conditioned by the the
propeller slipstream. Following a similar approach as that proposed Woodward
in [6], which seems to give satisfactory results [5], the rotational speed induced
by the propeller in its wake is neglected, while it is considered the acceleration
induced by the delivered thrust.
+

VS

VE
aVE

X
D

VS

TTOT
Y

Figure 5 Kinematics scheme used for the calculation of the inflow velocity to a thruster body
(left) and definition of the direction of the developed hydrodynamic forces (right).

The kinematics of the calculated inflow to the thruster body is sketched in


Figure 5. The propeller acceleration is thought to act on the incoming stream
flow VE but, for the conservation of momentum, the direction of the propeller
induced velocity is assumed parallel and opposite to the direction of the total
developed thrust. So, indicating with aVE the propeller induced velocity forming
an angle with the propeller axis, the resultant flow speed Vs on the thruster

body is approximated by the composition of these two flow components. In


formulae:

VS2 = [VE + aVE cos( + E )] + [aVE sin ( + E )]


2

(7)

The flow is diverted of an angle with respect to the inflow speed VE defined
in (3), which is calculated with the following formula:

aVE sin ( + E )

VE + aVE cos( + E )

= arctan

(8)

In (7) ad (8) the acceleration factor a caused by the propeller on the


incoming flow, can be calculated using the axial momentum theory:

a = KM

( 1+ C

(9)

in which, KM is a correction factor that depends on the distance between the


propeller and the strut mean chord, while the thrust coefficient is calculated
using the total propeller thrust, i.e.:

CT =

K T2 + K L2
P
P

1
J2

(10)

Once the equivalent uniform incident speed on the thruster body Vs is known
in intensity and direction, the assembly of strut and nacelle is assimilated to an
equivalent lifting surface. As per Figure 6, the equivalent lifting surface has the
same root chord of the propulsor strut, a tip chord approximately equal to the
nacelle length and a span measured from the shaft axis to the strut root profile.
The empirical formulae of Whicker & Felner, developed for trapezoidal rudders
having four digit symmetrical profile with low aspect ratios and moderate sweep
angles, are used to evaluate the lift and drag coefficients of this equivalent lifting
body in the , namely:
C L
=

0.9 2

2
57.3cos
+
4
+
1
.
8

cos 4

, C L ( ) =

C L
+ C D0 sin 2 cos

(11)

C L2
C D ( ) = C D0 +
e

at small angles (linear behaviour of CL)

(12a)

C D ( ) = 0.9 C DC sin

at higher angles (non linear part of CL curve)

(12b)

In which is the sweep angle of he equivalent lifting body, is the angle of


attack corresponding to the previously found angle ; e is the Osvald
coefficient (it depends on the planform shape of the lifting body), CD0 is the drag
coefficient of the 2D profile and the effective aspect ratio is evaluated
according with the following formula:

=k

Sm
+
Cm

A1
R
A2
Cm

(13)

The effective aspect ratio takes into account of the boundary effect induced
by the hull through the correction coefficient k (k=2 for thruster without gap
between the hull surface and strut) and of the boundary effect induced by the
nacelle which acts almost like an end plate, actually increasing the strut
effective aspect ratio. In (13) R is the nacelle radius, A1 and A2 are the two
planar areas highlighted in Figure 6; Sm and Cm are respectively the mean span
and the mean geometric chord of the lifting surface. It is suggested to measure
the mean span Sm from the axis of the nacelle to the strut root profile.

Figure 6 Geometric variables used for the calculation of strut and nacelle forces

In (12b) the cross flow drag coefficient CDc is estimated case by case from
experience. In this study has been assumed equal to 1.4. After the initial
growing trend defined by (11), the lift coefficient vs. angle of attack curve is
interrupted at the stall angle and from that point brought to zero. An example of
calculated lift and drag coefficient as a function of the angle of attack is
presented in Figure 7.
1.5

CL, CD _

1.0

CD
CL nonlin

0.5

0.0
0
-0.5

30

60

90

120

150

180

a.o.a [deg]

-1.0

Figure 7 Estimated lift and drag coefficient for an Azimuthal thruster body as a function of the
angle of attack of the relative inflow.

Once the hydrodynamic force coefficient are known at all incident angles, the
lift and drag forces developed by each thruster body are calculated with the
following formulations:

2
L = C L ( E ) 2 Cm S m VS

D = C ( ) 1 C S V 2
D
E
m m S

(14)

and these forces oriented according to the relative inflow speed to the
thruster are projected onto the ship fixed longitudinal an transversal axes to
finally obtain force and moments in this non-inertial reference frame:
X R = L sin ( + ) D cos( + )

YR = L cos ( + ) D sin ( + )

(15)

In case of two or more thruster, these calculations are performed


independently for each propulsion unit.

4 NUMERICAL IMPLEMENTATION
The non linear differential system of equations (1) can be integrated after
some mathematical manipulation. First (1) is rewritten isolating on the left hand
side the terms in phase with motion accelerations:

m11u = X * + m rv + xG r 2 = f X (u, v, r , , n,...)

*
m 22 v + m 23 r = Y mur = f Y (u, v, r , , n,...)

*
m32 v + m33 r = N mxG ur = f N (u , v, r , , n,...)

(16)

in which the generalized mass and inertia terms are defined as follows:
m11 = m X u

m22 = m Yv

m23 = mxG Yr

m32 = mxG N v

m33 = I z N r

(17)

The terms appearing on the right hand side of (16) do not depend on the
acceleration variable, in general depending all the other free variables, such as
ship motion instantaneous velocities u,v,r, propeller rotational speed n, thruster
azimuthal angle . To simplify the writing, the right hand side terms of (16) are
named fX, fY, fZ, for the surge, sway and yaw equations respectively.
A further linear combination of the second and third equations in (16) permits
to obtain the following equivalent differential system of equations, in which the
three ship motions accelerations have been isolated on the left hand side of the
relative dynamic equilibrium equation.

u = f X / m11

v = (m33 f Y m23 f N ) / (m23 m32 m22 m33 )


r = (m f m f ) / (m m m m )
32 Y
22 N
23 32
22 33

(18)

The system of differential equations (18) can be easily integrated using a


simple finite difference scheme. Usually even a first order scheme is sufficiently
accurate: we use a fifth order variable step Dormand-Prince algorithm. The
numerical simulator has been written in Matlab-Simulink a modern scientific

environment dedicated to the simulation of dynamic systems of any complexity.


The high level functional diagram of the simulator is presented in Figure 8, in
which the main calculation blocks have been identified with the following
progressive numbers:
hydrodynamic forces acting on the hull
delivered thrust and requested torque of the azimuthal propellers
(separated for port and starboard for two units as in this case)
hydrodynamic forces on the strut and nacelle of each azimuthal prop. unit
integrator blocks of the three acceleration components u, v, r
calculation of rhs of system (18) that gives the three motions accelerations
integration of ship motion speed for calculating the ship trajectory
pod electric motor and regulator block
instantaneous ship motion speed
rudder angle control (automated for standard ship manoeuvres)
The lines connecting different blocks mean an exchange of data between
calculation modules; they transmit the instantaneous values of the share
variables between blocks. So for instance the right hand sides of system (18) is
given as input to the integration block , which supply in output the
instantaneous ship motion velocities. These velocities are then further
integrated into to derive the ship motions and to calculate the trajectory
followed by the origin of the body fixed reference system. Ship motions are also
input parameters to blocks , and which calculates the instantaneous
values of hydrodynamic forces and moments developed by the hull and by its
propulsion units. The signals lines between the propeller and strut blocks
are needed to account for the influence of the propeller slipstream on the
hydrodynamic forces on the body of the azimuthal propulsor.
Each main block of the high level diagram in Figure 8 can contain other sub
level calculation blocks up to a virtually infinite complexity. Figure 9 and Figure
10, for instance, present the inside structure of the blocks used to calculate the
azimuthal propellers and the thruster bodies hydrodynamic characteristics,
respectively.
The two rounded blocks represent the input variables to the block: the input
is represented by the three motions speeds u,v,r and the rotational speed of the
electric motor N and the azimuthal angle of the unit .
The top left part of Figure 9 is dedicated to the calculation of the
instantaneous hull wake factor and applies it to the instantaneous ship forward
speed u as well as the total inflow speed VE according (3). The central part of
the diagram of Figure 9 is dedicated to the calculation of the instantaneous
propeller force (and torque) coefficients, KT, KL, KQ, by interpolation of the
tabulated results of systematic series of model tests. The azimuthal propeller
characteristics are interpolated using the effective azimuthal E using (4)
calculated in the blocks on the left, the instantaneous advance coefficient J and
the given pitch ratio. Finally the three blocks on the right perform a
dimensionalisation of the calculated forces and torque coefficient and project
them on the ship reference system according (6).

A similar analysis can be made on the block diagram of Figure 10 for the
thruster body. It is interesting to note that the calculation of the hydrodynamic
forces is split in two parts: a portion of the thruster housing that is influenced by
the wake of the propeller and the other which is not influenced by the
accelerated and deflected slipstream of the propeller.
Traiettoria Nave

u, v , r [m/s rad/s]

Traiettoria Nave

u, v , r

Out-psi

Velocit Istantanea

In-u,v,r

1/s
Integra udot

Out-u,v ,r
Out - V
In-u,v ,r
Out-du',dv ',dr'

u, v , r

1/s

udot, v dot, rdot

em

Integra vdot

Velocit Istantanea

1/s

Integra rdot

Equazioni del Moto

In1

Forze Idrodinamiche di carena

Plotta Forze e Mom Carena


du', dv ', dr'

In-u',v ',r'
Out-FXh,FYh,Nh

V [m/s]

FXh, FYh, Nh

In - V

f(u)

udot

Forze Idrodinamiche di Carena


f(u)

FXtot,FYtot,Ntot

v dot

FXtot,FYtot,Ntot,V
V [m/s]

psi

f(u)

rdot

In1

Plotta Forze e Mom Timoni

Forces on POD Struts

u, v , r [m/s rad/s]

Gain1

In-a, omega, VE
Out-XRp,YRp
In-delta_e, ddelta

f(u)

Gain

Calcolo FXR

Forces PS Strut
XRp,YRp,XRs,YRs [kN]

FXr, Fy r, Nr

f(u)
In-a, omega, VE

V [m/s]

Calcolo FYR

Out-XRp,YRp
In-delta_e, ddelta

In1

f(u)

Motor and Regulator

Plotta Forze e Mom Eliche

Calcolo NR

Forces SB Strut1

Out-N

In-Q

Regulator

Helm
Command
Out- a_p, omega_p, VE
In-Np, delta
In-psi

Out-d

a_p, omega_p, VE

Out-delta_p,ddeltap

d [rad]

Out-Tp-Ttp

Timone_EVU
In-u,v ,r

Out-Qp

Tp,Ttp [kN]
Qp [kNm]

PS POD

f(u)

PODs - Thrust, Torque

Calcolo FXp
Tp,Ttp,Ts,Tts [kN]
Tp,Ttp,Ts,Tts, V
V [m/s]

f(u)

FXp, FYp, Np

Calcolo FYp

Out- a_p, omega_p, VE


In-Np, delta

f(u)
Out-delta_p,ddeltap

Calcolo Np
Out-Tp-Ttp
In-u,v ,r
Out-Qp

PS POD1

n0
Out-N

n0 [Hz]

Nota:
Tt: spinta trasv . eliche con v erso positiv o esterno nav e;
Yp: distanza centro disco elica-piano simmetria nav e;
Xp: distanza centro disco eliche-origine assi nav e

In-Q

Regulator1

Figure 8 High level diagram of the devised Simulink model for the simulation of a ship with
twin azimuthal thrusters

Figure 9 Functional block diagram used to simulate delivered thrust and requested torque by
the azimuthal thrusters propeller during ship manoeuvres

Figure 10 Functional block diagram used to simulate hydrodynamic forces of the thrusters
body (strut and nacelle) during ship manoeuvres.

5 APPLICATION TO AN EXISTING CONVENTIONAL RO-PAX FERRY


A first application of the manoeuvring simulator has been made on an
existing ship already modelled with conventional propulsion layout [2]. The ship
is a conventional medium speed ro-pax ferry, operating since some years in the
Mediterranean Sea (top view of Figure 11).

Figure 11 Longitudinal view of the Ro-Pax ship assumed as test case. Top: original propulsion
system with twin CP propellers. Bottom: simulated alternative with 2 podded propulsors.

This ship has a length between perpendiculars LBP=176, a moulded breadth


BM=28m and a design draft TD=6.6m, with a carrying capacity of about 2250
passenger and 1000 cars or 2200 lane meters for trucks. In the original version,
she has two conventional shaft lines and CP propellers, each one driven by two
diesel engines each with 7450 kW MCR. At 85% of MCR, the existing ship is
able to reach a speed of about 24 knots. The alternative propulsion layout
features two electrical azimuthal propulsion units, each one driven by 11 MW
electric motor. This solution, as sketched in the bottom part of Figure 11, is
already compatible with the ship hull stern shape, which require only a minor
modification (flattening of the extreme aft portion of the buttocks). Also the
central skeg profile needs to be slightly modified with respect to the original
solution, but the total area is nearly unmodified. With respect to these minor
modifications, no alteration of the hydrodynamic derivatives have been
considered for the hull.
For the existing ship the results of full scale manoeuvring trials are available:
mainly the turning circles and zig-zag tests. The same manoeuvres have been
simulated, in this study, for the above mentioned alterative azimuthal propulsion
system configuration, at the same initial ship speed and same rudder
(azimuthal) angles.
Table 1 resumes the main manoeuvring characteristics of the simulated ship
with podded propulsion against those registered for the real ship at full scale.
Figure 12 shows the simulated trajectory and drift angle followed by the
alternative ship during the turning circle manoeuvre. Figure 14 presents a
standard time history of the azimuth angle of the propulsion units together with
the ship bow angle, simulated during the 10-10 zig-zag manoeuvre.

In general, from the comparison of results, the ship with azimuthal propulsion
units appears to have more initial steering capacity (she is more reactive to
thrusters rotation) than the existing ship, with conventional shaft lines and
rudders. This is, in fact, demonstrated by the smaller overshooting time and
angle obtained during the simulated Z manoeuvre with respect to original
configuration as well as by the lower advance and transfer distances obtained
at the initial phase of the turning circle.
At the same time, the ship with podded propulsion seems to be more
directionally stable than the original. This can be argued from the comparison of
the overshoot angles which do not increase from the first to the third
manoeuvres for the ship with pods, while the original ship registered a
considerable increase. In addition also the final diameter is higher for the ship
with pods.
Table 1 Synthesis of the results of turning circle and Z manoeuvres for the original ship with
conventional shaft lines arrangement and simulations for the same ship with podded propulsion
Manoeuvre

Initial Speed

Helm

Advance

Transfer

Tactical Diam.

Final Diam.

[knots]

[deg]

[m]

[m]

[m]

[m]

Turning Circ.

21.6

35 sb

453

118

303

243

Turning Circ.

21.6

35 ps

434

145

334

265

Simulated

21.6

35

280

200

350

290

Manovra

Initial Speed

Helm

Mean
Period

Angle / Time
1 Overshoot

Angle / Time
2 Overshoot

Angle / Time
3 Overshoot

[s]

[deg] - [s]

[deg] - [s]

[deg] - [s]

Zig-Zag 10-10

[nodi]

21.6

10 ps

132

14.5 20.0

16.0 20.0

24.8 28.0

Simulated

21.6

10

75

4.0 9.0

4.5 9.0

4.8 9.0

Figure 12 Simulated trajectory of the ship during turning circle evolution, with initial speed
21.6 knots. And rudder angle 35 degrees.

Interesting is also the analysis of the forces developed by the azimuthal


thruster during the turning circle manoeuvre. In the time histories of Figure 12,
the component of the forces relative to the ship reference system have been
separated between that produced by the azimuthal propellers and that by the
thruster bodies. The manoeuvre starts from a condition of uniform rectilinear
motion at the given speed f 21.6 knots. Then at the time T=50s, the azimuthal
units starts to be rotated up to 35 degrees. From the calculation it appears that
the transversal force of the thrusters body is initially comparable, as intensity,
with that of the propellers, but during the stabilization of the manoeuvre this
force component is reduced by nearly one half. The side force of pulling
propellers, instead, remain almost unvaried during the manoeuvre.

a)

b)
Figure 13 Longitudinal and lateral force components developed by the thrusters bodies (a)
and azimuthal propellers (b), during the ship turning circle manoeuvre of Figure 4.

Overshoot time
Overshoot angle

Figure 14 Simulated 10-1 zig-zag manoeuvre at 21.6 knots initial speed

6 CONCLUSIONS AND FURTHER PROSPECTS


The paper presented a practical theoretical model for the manoeuvring
simulation of ships with azimuthal propulsion units. The hydrodynamic forces
produced by the azimuthal thrusters have been divided into its main
components given by the azimuthal propellers and the thruster body. The
calculation of the forces produced by propellers in inclined flow is based on a
systematic series of model tests on propellers. This approach has the
advantage to eliminate the approximations introduced the different methods
used in other simulators, but is inherently confined within the limited range of
propeller characteristics varied in the available series of model tests. In this
respect, the very next task will be that of comparing the simplified model
proposed of Woodward et al. [5] to calculate forces produced by propeller in
oblique flow, with the mentioned model test results. This would permit an
important extension of the generality and applicability of the present simulator.
The numerical simulator written on the base of the method described in the
paper has been also described in the paper, in its principal computational
characteristics. The simulator has been successfully applied to the case of an
existing ro-pax ferry, ideally converted to a podded propulsion with two
azimuthal units. From the comparison of the simulated results with those of the
existing ship at full scale it appears that the alternative design with pods has a
considerably higher initial turning ability, than the existing one, and it also
posses a higher directional stability. This last finding is probably due to the large
dimension of the thrusters bodies in comparison of the existing ship rudders and
the fact that the hull lines and central skeg were not substantially altered
between the two designs.
It would be interesting, in the future, to validate the results obtained by this
simulator with full scale data of a ship with azimuthal propulsion.
Next enhancements of the presented simulator will hopefully be:
- the extension of the model to the case of azimuthal propulsors with
pushing propeller. This would enlarge the field of application of the
simulator to the many smaller vessels with mechanical rudder propellers.
In this case a different approach for the forces on the strut and nacelle is
needed.
- The extension of the model to the case of azimuthal propulsion units with
two propellers in tandem: the Schottel Twin Propeller concept. This
would require a more sophisticated model of the inflow on to the aft
propeller as modified by the forward propeller and intermediate strut and
nacelle. A large number of ships are sailing with these kind of thrusters
and model/full scale tests should be available for comparison.

7 AKNOWLEDGEMENTS
A truthful thank to Mr. Stefan Kaul of Schottel, for his kind availability in
supporting this independent research by releasing data about azimuthal
propeller tests in oblique inflow. Without his precious contribution this work
would not be possible.

REFERENCES
[1] BENVENUTO G., BRIZZOLARA S., FIGARI M. (2001) Simulation of The Propulsion
System Behaviour during Ship Standard Manoeuvres, Proc. of Int. Conf. on
Practical Design of Ships and Offshore Structures PRADS 2001, Sept. 2001,
Shanghai, Elservier ed., Vol. I, pp.657-663.
[2] BENVENUTO G., BRIZZOLARA S., CARRERA G. (2003) Ship Propulsion Numerical
Simulator: Validation of the manoeuvrability Module, Proc. of NAV 2003, Int.
Conf. on ships and shipping research, June 2003, Palermo (IT).
[3] BRIZZOLARA S. (2003) La Simulazione di Manovra delle Navi di Superficie: Cenni
Teorici ed Esempi di Applicazione del Simulatore DINAV , Proc. of MIMOS 2003,
III Italian Scientific Conference on Simulation, Nov. 2003, Torino (IT), (in Italian).
[4] CASSELLA P. (1971) On the propeller in yaw: a comparison between some results
obtained with the equivalent section theory and the experiments, Tecnica Italiana,.
201, August 1971, (in Italian).
[5] WOODWARD M. D., ATLAR M., CLARKE D. (2005) Comparison of Stopping modes
for Pod-driven Ships by Simulation Based on Model Testing , Proc. of IMECHE,
Engineering for the Marine Environment, Vol. 129, Part M.
[6] WOODWARD M. D., CLARKE D., ATLAR M. (2003) On the Manoeuvring
Predictions of Pod Driven Ships, Proc. of Marine Simulation and Ship
Manoeuvrability, MARSIM 2003, Vol. 2, paper 7.
[7] KAUL S. (2004) Regressionsanalyse fr die Berechnung auf Basis von
Polynomkoeffizienten, daten aus VBD-Bericht Nr. 707 vom 1974, from
SCHOTTEL database of propeller tests, private comunication with the author.

También podría gustarte