0% found this document useful (0 votes)
42 views13 pages

Eslamdoost 2018

Shimeney

Uploaded by

Alecrates Tundag
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
42 views13 pages

Eslamdoost 2018

Shimeney

Uploaded by

Alecrates Tundag
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 13

Ocean Engineering 152 (2018) 100–112

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

Analysis of the thrust deduction in waterjet propulsion – The Froude


number dependence
Arash Eslamdoost *, Lars Larsson, Rickard Bensow
Department of Mechanics and Maritime Sciences, Chalmers University of Technology, Gothenburg, Sweden

A R T I C L E I N F O A B S T R A C T

Keywords: The definition of thrust deduction in waterjet propulsion is different from that of a propeller driven hull and
Waterjet propulsion cannot be interpreted in the same way. A particularly interesting feature of the waterjet thrust deduction is the
Thrust deduction large variation with Froude number. This is well known from experience, but has never been fully explained. The
Net thrust objective of this paper is to use CFD to address the reasons for these large variations. To this end, the thrust
Gross thrust deduction fraction is split into resistance increment fraction and jet thrust deduction fraction. The former is due to
Sinkage and trim
the self-propelled hull resistance change in comparison with the bare hull resistance and the latter is due to the
Resistance
difference between gross and net thrust. This split reveals that the main reason for the thrust deduction variation
is the hull resistance change. Analysis of the resistance increment in different speed ranges is performed by
studying the hydrostatic and hydrodynamic pressure changes on the hull as well as the friction change due to the
waterjet system. Despite the negative thrust deduction fraction in the intermediate speed range there is no
indication of a resistance reduction compared to that of the bare hull at these speeds.

1. Introduction the wake momentum losses and the loss of the planing surface at the
intake opening as the most significant ones. Coop reports the largest
The waterjet propulsion concept is based on increasing the flow head measured negative thrust deduction fractions (up to 8%) around the
through a pump inside a ducting channel and then discharging the hump speed. He states that the ITTC suggested method for towing the
ingested flow with a higher momentum flux behind. The most common hull along the shaft line is the most conservative approach, which yields
waterjet intake design is flush to the hull bottom, however, depending on the highest bare hull resistance.
the craft type the intake geometry can be ram or scoop shaped (Kruppa et In his doctoral thesis, van Terwisga (1996) showed that the thrust
al., 1968). Depending on the specific speed of the waterjet pump, a mixed deduction fraction of a waterjet propelled hull varies considerably with
flow or an axial flow pump can be employed in the unit (White, 2008). speed and may even obtain negative values in an intermediate speed
Within the speed range of 30–60 knots, the major waterjet manufacturers range. Through a set of analytical, numerical and experimental studies,
suggest this propulsion method over conventional propellers. he found a difference between gross thrust and net thrust1 especially
Since the waterjet unit is embedded in the hull, the interaction effects around ship speeds where the transom is not fully cleared. He states that
between the waterjet system and the hull are quite different from those of the difference is practically zero for higher speeds and therefore, the
conventional propellers. In contrast with the propeller/hull interaction difference between gross thrust and bare hull resistance is a good mea-
effects, which are rather well defined, there are still unanswered ques- sure of the resistance increment of the hull due to the waterjet-induced
tions regarding the waterjet/hull interaction. Three major research flow. Through an uncertainty analysis of propulsion tests, he shows
studies in this field are the doctoral thesis by Coop (1995), van Terwisga that the error made in the flow rate measurement in power estimation
(1996) and Bulten (2006). Coop investigated the interaction between increases with decreasing jet velocity to ship velocity ratio. He divides
waterjet and hull using model scale and full scale measurements, as well the effects causing the resistance change of a self-propelled hull, into a
as empirical and analytical methods. He discusses the possible mecha- global sinkage/trim effect and a local effect on the flow around the intake
nisms contributing to the overall interaction effect and lists the waterjet due to the ingested flow. Then he states that if the assumption of inde-
momentum forces causing lift and moment about the centre of gravity, pendence between the changes due to the local and the global flow is

* Corresponding author.
E-mail address: [email protected] (A. Eslamdoost).
1
Definition of gross thrust and net thrust are given in Section 2.

https://doi.org/10.1016/j.oceaneng.2018.01.037
Received 6 August 2016; Received in revised form 9 October 2017; Accepted 7 January 2018

0029-8018/© 2018 Elsevier Ltd. All rights reserved.


A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Nomenclatures SP self-propelled hull


t thrust deduction fraction (waterjet) ½
A control surface Tg gross thrust vector component in x-directon ½N
B hull beam [m] tj jet system thrust deduction fraction ½
BH bare hull Tnet net thrust vector component in x-directon ½N
BHSP BH with the same sinkage and trim as SP tp thrust deduction fraction (propeller) ½
cp local pressure coefficient ½– tr resistance increment fraction ½–
CT hull resistance coefficient [] u velocity vector ½m=s
De exit drag [N] u0 undisturbed velocity ½m=s
Di intake drag [N] x, y, z Cartesian earth fixed coordinates
Fn Froude number ½– ΔR resistance increment ½N
Fpx pump force per unit mass in x-directon ½N ρ density of fluid ½kg=m3 
i; j; k tensor indices denoting the ordinates σ hull sinkage ½m
Lpp length between perpendiculars ½m σx x-component of pressure and shear stresses ½Pa
N normal unit vector τ hull trim angle ½degrees
Q volume flow rate ½m3 =3 η transverse distance from hull symmetry plane [m]
Ra rope force ½N ξ vertical distance from transom edge [m]
Rbh bare hull resistance ½N

true, the change in resistance may be estimated from a Taylor series in including power estimation, e.g. (Kandasamy et al., 2010, 2011; Takai et
the global and local flow parameters, keeping only linear terms. Based on al., 2011). However, the detailed waterjet/hull interaction effects have
the measurements of a powerboat propelled by a single waterjet unit, he not been the focus of these studies.
concludes that the trim angle is the most important parameter for Eslamdoost et al. (2013) presented a method based on potential flow
analyzing the resistance increment of the hull. theory for studying the waterjet driven hull flow. They also identified the
Bulten (2006) studied the flow inside waterjet propulsion systems parameters which contribute to the waterjet/hull interaction in (Eslam-
employing Computational Fluid Dynamics (CFD) tools. The integration of doost et al., 2014). In their third consequent paper (Eslamdoost et al.,
axial force component on the waterjet unit, as well as a simplified version 2016), these authors investigated the difference between the net thrust
of the integral momentum balance equation, were applied to calculate and gross thrust of a waterjet unit through a RANS study. They show that
the waterjet thrust. Bulten reports a clear deviation between the results of the net thrust and the gross thrusts are not the same and deviate from
these methods for higher ship speeds. He also computed a large vertical each other, especially in the speed range where the nozzle is submerged
force in the same speed range. According to these findings, Bulten con- in the transom wave.
cludes that the method based on the momentum balance for the waterjet Thus, many researchers have contributed to the understanding of the
control volume is incorrect. possibly because of the influence of the hull waterjet/hull interaction effects but yet there is a knowledge gap in
in the vicinity of the waterjet inlet and partly because of neglecting the explanation of the reasons for the large variation of the thrust deduction
contributions of the pressure distribution acting on the streamtube.2 fraction with speed, as well as of the negative thrust deduction values
The ITTC High-Speed Marine Vehicle Committee (ITTC 17, 1987) and sometimes reported in the intermediate speed range. The goal of the
the ITTC Specialist Committee on Waterjets (ITTC 21, 1996; ITTC 22 present paper is to study the thrust deduction fraction of a waterjet
1998; ITTC 23 2002; ITTC 24 2005) have proposed a test procedure for propelled hull in a wide speed range and to clarify the reasons for the
investigation of waterjet/hull interaction which was modified during the large variation of this fraction.
years. A measurement campaign was also conducted by the ITTC In the following, first, the definitions required for the analysis of the
Specialist Committee on Waterjets on a semi-displacement hull with two waterjet system are introduced in section 2. Then, the hull geometry and
sets of axial flow waterjets. The results of this study, which was per- the towing tank measured sinkage and trim are presented in section 3 and
formed by many different institutes, show a large scatter for the thrust section 4, respectively. A brief review of the computational technique for
deduction fraction. This highlights the importance of the waterjet flow the waterjet flow simulation is given in section 5 and in section 6 the
rate measurement. components which contribute to the thrust deduction fraction in different
The Office of Naval Research (Rispin, 2007) carried out a compre- speed ranges are discussed in detail. Finally, several conclusions are
hensive set of measurements on a demi-hull with a pair of waterjet units. drawn.
The difference of boundary layer thickness due to scale change was taken
into account in the data scaling procedure. Consequently, although the 2. Theory
thrust deduction fraction in the intermediate speed range was positive at
model scale, taking this correction into account resulted in a negative The thrust deduction fraction is the relation between the bare hull
thrust deduction fraction. This conclusion raises the question of whether resistance and the thrust required to propel a vessel. For a conventional
or not the thrust deduction is dependent on scaling. propeller, the net thrust is employed to define the thrust deduction
Kamal et al. (2015) carried out an experimental comparative study of fraction as follows,
the powering of waterjet and screw propeller for a medium speed
wave-piercing catamaran. They predict the thrust deduction fraction to Rbh  Ra
tp ¼ 1  ; (1)
be very low for the waterjet propelled hull (0.1 to 0.35) but no certain Tnet
reason for the small values is provided.
where Rbh and Tnet are the bare hull resistance and the net thrust of the
Several studies employing Reynolds Averaged Navier-Stokes (RANS)
propeller/s. Since the frictional resistance of the hull at model scale, due
methods have been carried out modeling waterjet-propelled hulls and
to lower Reynolds number, is larger than the full scale frictional resis-
tance a towing force is applied to the hull during the self-propulsion test.
This force is called the rope force and is shown with Ra .
2
See Fig. 1 and read the corresponding text for the definition of the streamtube.

101
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

In contrast to a propeller, for which the net thrust is transmitted to the


hull through its shaft line, the net thrust of a waterjet unit is transmitted
to the hull through the shaft line and the ducting channel. The integrated
geometry of the waterjet unit and hull makes it tricky to measure the net
thrust directly. In order to skip the impracticality of measuring the net
thrust, the ITTC High Speed Vehicle Committee (ITTC 17, 1987) sug-
gested to use gross thrust, Tg , which is the momentum flux change
through the waterjet control volume defined in Fig. 1.
In the control volume shown in this figure, surface 2 is part of the
waterjet inflow streamtube. No flow passes through this surface. The
capture area, surface 1, is the cross section of the streamtube located “far
enough in front of the intake ramp tangency point, before inlet losses
occur” (ITTC 21, 1996) and as a practical solution, the ITTC Specialist
Committee on the Validation of Waterjet Test Procedures recommends
this section to be placed one inlet length forward of the ramp tangency
point (ITTC 24, 2005). The internal material boundaries of the waterjet
system that extends to surface 1 and surface 2 is denoted with surface 6;
Surface 7 is the boundary area of the pump control volume and surface 8 Fig. 2. Three-dimensional presentation of surfaces 1,2, 4 and 8.
represents the nozzle exit area. The cross section of the jet, where the
mean static pressure equals the atmospheric pressure, is called Considering the material boundaries of the waterjet system, “the net
vena-contracta and is indicated with surface 9. The vena-contracta has !
thrust, T net , is defined as the force vector acting upon the material
the minimum jet cross sectional area. The only surface shown in Fig. 1
boundaries of the waterjet system, directly passing the force through to
that belongs neither to the waterjet control volume nor the hull is surface
the hull” (van Terwisga, 1996). Like the gross thrust in this paper, the
4. This surface is enclosed by the stagnation line on the intake lip (cut
horizontal component of the net thrust vector will be called net thrust,
water), and the line where the channel merges with the flat bottom.
Tnet , and is defined as,
Surfaces 1, 2 and 4 are flow dependent and can vary one operation
condition to another. The actual shapes of these surfaces are shown in Tnet ¼  ∬ σ x dA  ∭ ρFpx dV; (4)
Fig. 2. A4 þA6 V7

The horizontal component of the gross thrust vector can be obtained


by writing the horizontal momentum flux balance over the control vol- where σ x is the x-component of the surface stresses (pressure and shear
ume shown in Fig. 1, stresses) and Fpx is the x-component of the pump force per unit mass.
As mentioned earlier, it is a tricky task to measure the net thrust of a
Tg ¼  ∫ ρux ðuk nk ÞdA; (2) waterjet unit but with the aid of CFD it is possible to obtain the terms
A1 þA8
shown in Equation (4). Then the hull resistance increment fraction based
on the net thrust of the waterjet unit would be as follows,
where ρ is the density of the fluid, u is the velocity vector and n is the unit
vector normal to the control surface pointing outward of the control Rbh  Ra
volume. The Einstein notation is used in Equation (2) where index k tr ¼ 1  : (5)
Tnet
ranges over 1, 2 and 3.
According to the definitions presented in Equation (3) and Equation
The total thrust deduction fraction of a waterjet propelled craft, then
(5), the difference between the total thrust deduction fraction and the
can be based on the gross thrust as follows,
resistance increment fraction is the jet system thrust deduction fraction,
Rbh  Ra tj , plus a second order term including tr and tj (Eslamdoost, 2014),
t ¼1 : (3)
Tg
t  tr ¼ tj  tr tj ; (6)
A collection of thrust deduction fraction data for different hulls has
been obtained through self-propulsion tests at MARIN for a wide variety where,
of hulls with length to beam ratio varying from 3 to 15 (see Fig. 9). The
overall trend of this collection shows that t has large positive values in Tnet
tj ¼ 1  : (7)
the low speed range, small negative values in the intermediate speed Tg
range and then approaches small positive values in the higher speed The jet system thrust deduction fraction represents the difference
range. In this paper, we will study the reasons for the variation of the between gross thrust and net thrust. A negative tj implies that net thrust is
thrust deduction fraction in different speed ranges. To be able to proceed larger than gross thrust and vice versa.
with this investigation some more definitions are explained in the The objective of the current paper is to explain the physics of the
following part. waterjet-hull interaction. In order to achieve this goal a Reynolds Aver-
aged Navier-Stokes (RANS) solver is employed to compute the compo-
nents of the thrust deduction fraction for a high-speed hull. The relative
importance of tr and tj is studied and conclusions are drawn on the rea-
sons for the variation of the thrust deduction fraction from displacement
speeds to planing speeds including the intermediate speed range, where
the thrust deduction fraction is negative.

3. Hull and waterjet geometry

The hull geometry used in this study is a test case designed at SSPA.
Fig. 1. Section cut through the waterjet ducting system (Eslamdoost The hull was 2.3 m long and for the self-propulsion test it was equipped
et al., 2016). with a mixed-flow waterjet pump and an intake designed by Rolls-Royce.

102
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Fig. 3. Hull geometry and positioning of the


waterjet unit.

Fig. 4. Measured fore draft, σ f and aft draft, σ a of the bare hull and self-
propelled hull. Negative draft means that the hull has moved downwards.

Fig. 7. Presentation of the mesh used for the self-propulsion simulation.

filled with water to create the same initial conditions for the bare hull and
self-propelled one (Brown, 2013).

4. Measurements

Resistance and self-propulsion tests were carried out at SSPA to


provide data for validation of the numerical method and to determine the
Fig. 5. Measured sinkage of the bare hull and the self-propelled hull at sinkage and trim under all conditions. To increase the accuracy of the
mid-ship. numerical computations the measured sinkage and trim was employed in
the computations of the present paper.
The measured aft and fore draft of the bare hull (BH) and self-
propelled hull (SP) are plotted in Fig. 4. Comparing the aft draft of the
self-propelled hull with that of the bare hull, it is seen that the transom
submergence of the waterjet driven hull is larger than the bare hull
transom submergence for the entire measured speed range. The sinkage
and trim values of BH and SP can be calculated using the fore and aft
drafts. Fig. 5 shows the sinkage of the hulls at mid-ship. The waterjet-
propelled hull sinks deeper due to the action of the waterjet unit
(Eslamdoost et al., 2013). The measured trim angle of the bare hull and
self-propelled hull is plotted in Fig. 6. At the intermediate and higher
speed ranges (Fn > 0:6) there is a bow down trimming moment, which
causes the self-propelled hull to trim less than the bare hull. Since the aft
Fig. 6. Measured trim angle of the bare hull and the self-propelled hull. part of the waterjet-driven hull submerges deeper at the lower speed
ranges, the trim angle of the self-propelled hull becomes larger than that
The hull geometry and the positioning of the waterjet unit on the hull are of the bare hull.
shown in Fig. 3. Note that the unit was mounted on the hull also in the
resistance tests, but then the intake opening was covered and the unit was

103
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Fig. 8. Grid distribution in the aft part of the hull (left) and inside the ducting channel (right).

with the use of a very thin region.


In order to eliminate the uncertainties of sinkage and trim measured
values have been used in the computations.
It is assumed that the flow around the hull is symmetric and therefore
just the half the hull model is used in the simulations. The computational
domain is extended 1.5 LPP in front of the hull and 4.5 LPP behind it. The
lower boundary of the domain is 2.5 LPP below the undisturbed free-
surface and the upper 1.5 LPP above the undisturbed free-surface level.
The width of the computational domain is 2.5 LPP. Inlet velocity
boundary conditions are used for all far field boundaries except the
downstream boundary, where pressure outlet is applied. A systematic
variation of the domain size showed that the pressure and frictional
resistance components varied less than 1%, when the domain size was
enlarged by a factor of 3. This ensures the results are not depending on
the domain size.
The mesh used for the simulations consists of unstructured surface
Fig. 9. Collected thrust deduction fraction (van Terwisga, 1996) and the mesh, prism layers and core mesh. Prism layers are employed to resolve
computed thrust deduction fraction (current work, full line). the boundary layer, and hexahedral cells outside the boundary layer grid
form the core mesh. The hexahedral cells are aligned with the direction of
5. Computational method the undisturbed flow, see Fig. 7. The surface mesh and the symmetry-
plane mesh in the aft part of the hull are shown in Fig. 8. Anisotropic
The numerical simulations presented in this paper are carried out mesh refinement in three different levels are used close to the undis-
employing the code STAR-CCMþ 8.02 that uses a Finite Volume method turbed free-surface, and a finer mesh is employed around the waterjet
to solve the Reynolds-Averaged Navier-Stokes (RANS) equations and the intake, as well as around the discharged jet, to capture the details of the
continuity equation in integral form. The realizable k-ε turbulence model flow in these regions. The mesh size and mesh distribution around the
with wall function is used to solve the turbulence effect on the mean flow. hull are kept almost the same between the waterjet-propelled hull and
The tracing of the free-surface is done using the Volume of Fluid method the bare hull for the consistency of the simulations.
(VOF) in combination with the High-Resolution Interface Capturing A systematic grid refinement study was carried out for the bare hull
(HRIC) scheme to discretise the convective term of the volume fraction and self-propelled hull grids to obtain a suitable cell size and hence total
transport equation (Muzaferija and Peric, 1999). This scheme is suitable grid number, see (Eslamdoost et al., 2015) and (Eslamdoost et al., 2016),
for tracking sharp interphase of an immiscible phase mixture and re- respectively. The formal verification of the resistance and thrust co-
solves the free-surface within typically one cell. efficients is presented in these two papers. The total number of cells for
Since the flow features inside the waterjet pump is not the focus of the bare hull and the self-propelled hull were 3:7E6 and 5:7E6, respec-
this study, a body force model is employed instead of modeling the actual tively. Based on the least squares root method by Eça and Hoekstra
geometry of the waterjet pump to reduce the computation costs. Thus, (2014), a numerical uncertainty of 2.3% was computed for both grids.
the impeller blades and the stator guide vanes are removed. In this The bare hull resistance, the self-propulsion volumetric flow rate and
approach, as far as the body force induced head rise through the pump is the gross thrust are validated against the measured values in (Eslamdoost
equal to that of the actual pump, the average flow would be the same as if
the actual pump geometry was simulated. The body force magnitude is
equal to the resistance of the entire system including the hull and the
ducting channel minus the rope force applied to the hull during the self-
propulsion tests for unloading the water jet. In the computations, the
body force is inserted as an axial source term into the momentum
equation and is distributed uniformly inside the volume which surrounds
the impeller. A thinner or thicker region can be used for distribution of
the body force but it should be kept in mind that the volumetric integral
of the body force inside this region should balance the other forces acting
on the entire system. In order to achieve a more stable numerical
convergence, the thickness of this region should be such that the flow
accelerates gradually instead of an abrupt acceleration that may happen Fig. 10. Computed thrust deduction fraction, t, and resistance increment
fraction, tr .

104
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

et al., 2016). The average discrepancy between the computed and the contribution of each of them to the resistance increment may be esti-
measured bare hull resistance in the Froude number range of 0.2–1.6 is mated. The combined effect of sinkage and trim can be called global flow
less than 4%. The computed flow rate and the gross thrust are pattern change and the contribution of the flow change around the intake
under-predicted by less than 3% and 8% in comparison to the measured can be referred to as local flow change. The effect of the global flow
values, respectively. change on the resistance can be obtained by introducing a bare hull
which has the same sinkage and trim as the waterjet-propelled hull. Also,
6. Analysis of the thrust deduction the comparison of the resistance between this new hull with the self-
propelled hull enables to identify the explicit effect of the waterjet
In order to compute the thrust deduction fraction, as shown in induced flow (local effects) on the hull resistance increment. In the
Equation (3), it is required to compute the bare hull resistance and the following, the notation BH and SP are used for identification of bare hull
gross thrust, for which the momentum flux through the capture area and and self-propelled hull, respectively, and the notation BHSP is used for
nozzle exit are needed. The procedure for computing the capture area bare hull with the same sinkage and trim as the self-propelled hull. The
geometry is explained in (Eslamdoost et al., 2016), where the computed way that the global and local effects are obtained is schematically shown
bare hull resistance and the gross thrust for the Froude number range in Fig. 12.
between 0.2 and 1.4 and are also presented. Based on these computed For the following discussion three Froude numbers have been
values, the thrust deduction fraction is plotted in Fig. 9 along with selected, each one representing an interesting speed range in Fig. 10. The
collected thrust deduction fractions shown in Fig. 2. The trend seen in the Froude numbers are 0.2, 0.8 and 1.4. Based on the resistance components
computed t is practically the same as in the collected values. presented in Fig. 11 and the global and local flows resistance increments
The resistance increment fraction, tr , can be obtained from Equation defined in Fig. 12, results for the three selected Froude numbers are
(5). In Fig. 10 the t and tr curves are plotted together. There is an almost discussed in the following sections.
constant shift between t and tr , at all Froude numbers but the smallest
one. The reasons why the shift is larger here is explained in (Eslamdoost 6.1.1. Lower speed range (0:20 < Fn < 0:47)
et al., 2016). All features of the t variation with speed – the high values at The resistance increment fraction curve can be divided into two
low speed, the rapid drop near the Froude number where the transom distinct regions in the low speed range as shown in Fig. 10. In the speed
clears, the low values in the intermediate speed range and the increase at range between Froude number 0.2 and 0.43 the tr curve slope is negative
higher speeds – are caused by the variation of tr. It is the variation of the but is not as steep as the part of the curve between Froude numbers 0.43
self-propelled hull resistance that causes the variation of the thrust deduction and 0.47. In fact, these Froude numbers are the transom clearance Froude
fraction. Therefore, the following discussion will focus on tr. numbers for the self-propelled hull and the bare hull, respectively. This
indicates that the resistance increment is directly associated with the
6.1. Resistance increment critical Froude numbers of the hulls where the transom becomes dry. The
resistance increment in these two regions and the influence of the
The components of the resistance are obtained by splitting the total transom clearance on tr are discussed in the following.
resistance into the resistance of the transom and the resistance of the rest
of the hull, which in the following discussion is mentioned as ‘hull’. The 6.1.1.1. 0:20 < Fn < 0:43. The resistance increment fraction has large
waterjet unit is excluded from the self-propelled hull but the intake and positive values in this speed range and decreases with Froude number at
exit openings are on the hull and the transom. The resistance of the hull an almost constant rate until the transom of the waterjet driven hull
and the transom individually are split into the pressure resistance and the clears water at Froude number 0.43. Fig. 13 shows the hull and the
frictional resistance, where the pressure resistance in turn is split into the transom resistance increment components at Froude number 0.2 due to
hydrostatic and hydrodynamic resistance. This split is shown in a flow- the global, local and total effects in a schematic form. The flags in this
chart form in Fig. 11. figure represent the resistance increment in Newtons due to each resis-
Assuming that sinkage, trim and the local flow changes independently tance component separately. A backward pointing flag indicates an
influence the resistance change, the resistance increment of the hull may increased resistance component and vice versa. In the following the
be estimated from a linear expansion in a Taylor series and expressed as reasons behind the magnitude and sign of each component are discussed.
follows (van Terwisga, 1996), Global effects: Starting with the study of the global effects, it is seen
that the hydrostatic resistance of the transom decreases (1:0 N) while it
∂R ∂R ∂R increases on the hull (þ1:1 N). The different signs of the hydrostatic
ΔRðQ; σ ; τÞ ¼ ðQ0 ; σ 0 ; τ0 Þdσ þ ðQ0 ; σ 0 ; τ0 Þd τ þ ðQ0 ; σ 0 ; τ0 ÞdQ;
∂σ ∂τ ∂Q resistance increment on the transom and the hull are due to the normal
(8) vectors of these surfaces which have their axial components in opposite
directions. The main reason for the hydrostatic resistance change at this
where R is the hull resistance and Q, σ and τ are the flow rate through the
speed is the deeper sinkage of BHSP compared to BH (1:5 mm at mid-ship,
ducting channel, the hull sinkage and the hull trim angle, respectively.
see Fig. 5), which increases the projected area of the hull and transom
Index 0 denotes the bare hull.
below the undisturbed free-surface. Since the hull does not trim at this
The bare hull equilibrium position is the reference point, about which
speed, the increase in the projected area of the transom and hull should
the Taylor expansion is made. Obtaining the partial derivatives of resis-
be the same. However, it is seen that the hydrostatic resistance reduction
tance with respect to sinkage, trim and flow rate individually, the

Fig. 11. The resistance components


flowchart.

105
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Fig. 12. Global and local flow definitions.

Fig. 13. Schematic presentation of the hull and the transom


resistance increment components (in N) at Fn ¼ 0:20. The
bare hull resistance at this speed is 5:1 N.

due to the pressure on the transom is smaller than the increase on the the hydrodynamic pressure resistance of the transom slightly increases
hull. This is caused by the lower water level at the transom for BHSP. The (þ0:1 N) due to the lower hydrodynamic pressure acting on the transom.
slightly deeper transom edge causes a reduction in hydrodynamic pres- Despite the deeper sinkage of BHSP, the hydrodynamic resistance of the
sure when the flow passes across the edge. This reduces the hydrody- hull does not change. At this speed, the hull sinks by 1:5 mm at mid-ship
namic pressure in the whole, massively separated flow behind the and the corresponding displacement change is 2%. As a rough estimate,
transom, causing the water level to drop. Hence a smaller area for the the resistance may be assumed proportional to the displacement.
hydrostatic pressure to act on. Therefore, there would be 2% increase in the hydrodynamic resistance of
Since there is a minor increase in the transom submergence of BHSP, the hull at this speed, but since the hydrodynamic resistance of the hull is

Fig. 14. Hydrodynamic pressure coefficient contours and streamlines behind the BHSP transom at Fn ¼ 0:20 (left: top view and right: side view).

106
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Fig. 15. Hydrodynamic pressure coefficient contours and streamlines behind the SP transom at Fn ¼ 0:20 (left: top view and right: side view).

so small from the beginning (þ0:6 N), the anticipated 2% increase is All effects: A direct comparison of SP and BH resistance components is
negligible. also made in Fig. 13. Interestingly, the sum of the resistance components
The frictional resistance of the transom is zero, since the transom is due to the global and the local effects are equal to the components ob-
practically at right angles to the x-axis. The change between the wetted tained directly by comparing the resistance change between BH and SP.
surface areas of BHSP and BH is very small and therefore the frictional This finding supports the hypothesis that the resistance increment of the
resistance increment of the hull due to the global effects is also zero. hull is a linear function of sinkage/trim and flow rate (Equation (7)). The
Local effects: To study the local flow effects, the SP and of BHSP sum of all the resistance components at this speed becomes rather large
resistance components are compared. Since these hulls have same compared to the bare hull resistance (5.1 N). This is the reason for the
sinkage/trim the only cause, which can change the hydrostatic force large positive t and tr in the lower speed range.
component on the transom, can be the waterjet system. An important The total resistance of the self-propelled hull compared to the bare
effect that influences the pressure distribution on the transom, as well as hull increases by 58.8%, of which the local effects contribute 54.9%.
the pressure in the vicinity of the transom edge, is the jet entrainment. Only 3.9% is due to global effects. The resistance increment fraction tr is
Therefore, before starting the analysis of the resistance changes due to very large: 0.37, of which 0.34 is due to local effects.
local effects, the physics of the jet entrainment and its influence on the
hydrodynamic and hydrostatic pressures are discussed. 6.1.1.2. 0:43 < Fn < 0:47. In the speed range to be discussed in this
The hydrodynamic pressure coefficient on BHSP and SP as well as the section, tr drops abruptly from large positive values to values close to
streamlines behind the transoms of these hulls are shown in Figs. 14 and zero. In this study, it is discovered that the sudden change occurs in the
15, respectively. As seen in Fig. 15, the discharged jet entrains the dead speed range between the critical Froude number (for transom clearance)
water behind the transom. Therefore, the flow at the transom edge bends of the self-propelled hull and that of the bare hull. In the following, first
towards the transom hollow faster to replace the entrained water. The the possible causes that may result in the earlier transom clearance of the
faster turn of the flow means that the curvature of the streamlines leaving waterjet driven hull are discussed and afterwards the reasons for the
the transom edge is larger compared to the curvature of the bare hull sharp drop of the tr value are explained.
streamlines. The larger curvature of the streamlines means that the hy- Both measurements and computations show that the critical Froude
drodynamic pressure is lower at the transom edge (on the average Cp is number, where the transom becomes dry, differs between the towed hull
about 0.03 lower for SP at the transom edge, see Eslamdoost (2014)) and and the waterjet propelled one. Table 1 shows the measured and
thereby also on the entire submerged transom. This leads to a lower computed critical Froude numbers for both hulls. Note that the ranges
water level at the transom, and thereby to a smaller hydrostatic pressure shown for the measurements and computations are given by the step in
force. Froude number. The step was smaller in the experiments, hence a smaller
Having explained the effect of jet entrainment on the hydrodynamic range. Although, the computed wetted surface area of the hull is very
and hydrostatic pressure distributions on the transom, we will now similar to the towing tank result, the critical Froude number is under-
proceed with the discussion of the resistance increments due to the local predicted in the computations compared to the measured ones which
effects. can also be seen from the computed bare hull free-surface in comparison
The reduction of the hydrostatic pressure force on the transom due to to the towing tank results (Fig. 16). According to both the measurements
jet entrainment causes a slight increase in hydrostatic resistance, þ 0.3 and computations the transom of the waterjet driven hull clears earlier
N. A larger contributor is the missing area of the nozzle exit, which than the bare hull, as seen in Fig. 16. While the bare hull transom is still
contributes þ0:8 N out of the total increase of þ1:1 N (see Fig. 13). wet (foamy region behind the transom) the waterjet propelled hull has
The hydrostatic resistance of the hull due to the local effects also cleared water (glassy surface behind the transom). The evolution of the
increases by þ0:5 N, mostly caused by the missing surface covering the computed transom wetted surface area for the bare hull and the waterjet-
intake opening. Since this surface is facing backwards at this low Froude propelled hull is plotted in Fig. 17.
number, removing the surface will cause an increased hydrostatic resis- There are several reasons why the self-propelled hull has an earlier
tance. transom clearance than the bare hull, but it cannot be due to transom
As explained above, the jet causes a decreased hydrodynamic pres- draft changes, since SP has a larger transom draft than BH. Deeper
sure on the transom, which yields an increase in resistance by þ1.1 N. transom submergence causes a later transom clearance. In order to show
The hydrodynamic pressure resistance of the hull also increases as a that the earlier clearance of SP is caused by local effects a set of simu-
result of the local effects (þ0:2 N). The reason is not the missing surface lations was carried out for SP and BHSP. The results from this set of
at the intake opening, which has a negligible contribution to this effect.
Rather, it is the induced flow over the hull caused by the intake suction
Table 1
that gives this small resistance increase. Computed and measured critical Froude numbers.
The frictional resistance increment of the transom due to local effects
Measured Computed
is zero, since the transom is practically at right angles to the x-axis. This
resistance component on the hull decreases (0:1 N) due to the missing Bare Hull 0:494 < Fn  0:498 0:470 < Fn  0:475
Self-propelled Hull 0:430 < Fn  0:434 0:415 < Fn  0:434
area of the intake-opening.

107
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Fig. 16. Towing tank pictures (left) and computed


(right) flow pattern around the bare hull (upper) and
self-propelled hull (lower) at Fn ¼ 0:478.

Fig. 17. Transom wetted surface area of the bare hull (first row) and waterjet driven hull (second row) for various Froude numbers.

Fig. 20. Axial velocity variations on three vertical lines with different trans-
verse distance from the symmetry plane of the hull at the transom edge at
Fn ¼ 0:434. ξ and η are the vertical distance from the transom edge and
transverse distance from the symmetry plane of the hull, respectively.
Fig. 18. Wave pattern around BHSP (lower half) and SP (upper half)
at Fn ¼ 0:434.
whereas no such recirculating zone can be observed in the self-propelled
case. We believe this to be an effect of the high momentum jet flow which
simulations are presented in the following. pushes the dead-water flow behind the transom downstream and elimi-
Fig. 18 shows the contour plot of SP and BHSP wave height at nates the breaking waves filling the hollow behind the transom at Froude
Fn ¼ 0:434. The wave patterns of these two cases are very different numbers just below the critical one.
behind the transom. The aft-part streamlines on the symmetry plane can The second reason for the delay in transom clearance of SP is the
be compared in Fig. 19. Despite the equal sinkage and trim for both cases, boundary layer development behind the waterjet intake. Low momentum
the bare hull transom is wet and transom of the self-propelled hull is dry. boundary layer flow is sucked into the waterjet unit and then expelled
A recirculating zone is noted behind the transom of the bare hull,

Fig. 19. Flow pattern in the aft of BHSP (left) and SP


(right) on the symmetry plane at Fn ¼ 0:434. The
dashed lines show the free-surface and the full lines
show the streamlines.

108
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Fig. 21. Wave cuts at y=Lpp ¼ 0:243 for BHSP and SP at Fn ¼ 0:434. The
coordinate system has its origin at the zero speed bow waterline, x points
backwards, y to starboard and z vertically upwards. The transom is located at
x=Lpp ¼ 1 and the capture area is located at x=Lpp ¼ 0:71.

out far beyond the transom. The boundary layer developing on the bot-
tom behind the intake is thus thinner than for the bare hull. Fig. 20
shows the axial velocity at the transom edge on three vertical lines
positioned in different transverse distances from the hull symmetry plane
for both the bare hull and the waterjet driven hull. The axial velocity at
the transom edge is larger for SP compared to BHSP. The largest differ-
ence between these hulls is seen at the transom centreline. This means Fig. 23. Schematic presentation of the hull and the transom resistance
that the flow leaving the waterjet propelled hull transom edge has a increment components at Fn ¼ 0:80. The bare hull resistance at this speed
higher momentum and therefore the stern wave is less prone to breaking. is 110:4 N.
A third effect may be envisaged considering Fig. 21, which compares
the wave cut at y=Lpp ¼ 0:243 for BHSP and SP at the same Froude
number. Due to the suction of the waterjet unit the wave height is low- number where the two curves meet. This explains the abrupt drop in the
ered on the sides of the hull in the aft-part (starting near the capture area resistance increment fraction curve of Fig. 10.
section, x=Lpp ¼ 0:71) and therefore part of the flow, which could fill in
the hollow behind the transom from the sides, is removed. 6.1.2. Intermediate speed range (0:47 < Fn < 1:00)
These three effects result in an earlier transom clearance of the As seen in Fig. 10, the resistance increment fraction is close to zero in
waterjet-driven hull in comparison to the bare hull. The effect of the the intermediate speed range, 0:47 < Fn < 1:00. The speed in this range
different critical Froude numbers on the resistance increment is discussed is beyond the critical Froude numbers of the bare hull and self-propelled
in the following section. hulls and therefore there is no hydrostatic or hydrodynamic force acting
During the resistance tests of the hull used in this study, it was on the transom. It is the change of forces acting on the hull that
discovered that there is an abrupt peak in the resistance coefficient of the contribute to the resistance increment, see Fig. 23.
bare hull at the critical Froude number. The reasons for the peak were Global effects: The global contribution to the hydrostatic pressure
investigated using CFD and presented in (Eslamdoost et al., 2015). resistance increment of the hull is 2:2 N. The deeper sinkage of BHSP
Further studies of the self-propelled hull revealed the existence of the compared to BH (6 mm at mid-ship, see Fig. 5) increases the hydrostatic
same peak in the resistance coefficient of the waterjet-driven hull, as pressure acting on the hull; but despite such an increase, the smaller trim
appears in Fig. 22. Since the self-propelled hull has a lower critical angle of BHSP (0:2 less) positions the hull such that the imbalance be-
Froude number the peak is shifted to the left. It is this shift that causes the tween the hydrostatic pressure force acting on the fore and aft parts of the
rapid drop in tr . hull diminished. This is the reason for the reduced hydrostatic pressure
Replacing resistance by CT in Equation (5) yields, resistance due to the global effects.
The hydrodynamic pressure resistance increment due to the global
CTbh effects is þ7:8 N. The reason is the deeper sinkage, which gives an
tr ¼ 1  :
CTsp increased displacement and thereby a higher hydrodynamic pressure on
the bottom.
Employing this equation in combination with the data given in Fig. Comparing BHSP and BH, it is seen that the increased sinkage and
22, it is it is seen that tr is large and positive at the critical Froude number decreased trim angle result in increased wetted surface area of the hull
of the self-propelled hull thereafter dropping to zero at the Froude and therefore the frictional resistance of BHSP increases (þ0:6 N).
Local effects: The local effects cause a reduction in the hydrostatic
pressure resistance of the hull (5:1 N). The missing intake opening area
is the major part of this reduction (4:9 N). The trim angle of the hull at
Froude number 0:8 is about 3:8 (see Fig. 6). This positions the intake-
opening surface such that the hydrostatic force acting on this surface
has an axial component pointing backward. The hydrostatic pressure
resistance of SP decreases by removal of this surface. The rest of the
hydrostatic pressure resistance increment due to the local effect (0:2 N)
is caused by the intake-induced flow, which alters the waves close to the
aft part of the hull.
The local effects contribution to the hydrodynamic pressure resis-
tance increment influences the resistance by 1:1 N because of the
missing intake surface. Subtracting the hydrodynamic pressure resistance
Fig. 22. Resistance coefficients of the bare hull and the self-propelled hull.

109
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Local effects: Removing the contribution of the hydrostatic pressure


force acting on the intake-opening will decrease the resistance. This is the
reason for the negative hydrostatic pressure resistance increment at this
speed (5:0 N).
Like in the previous speed range, the surface normal of the intake-
opening at Fn ¼ 1:4 is pointing forward. Since the hydrodynamic pres-
sure is negative in the vicinity of the opening, the removal of a surface
from this region results in an increased hydrodynamic pressure resis-
tance. But only a trivial proportion of the total reduction of the hydro-
dynamic pressure resistance by the local effects is caused due to the
missing intake-opening area (0:1 N out of 2:0 N). This means that the
local effects alter the hydrodynamic pressure resistance of the hull
mainly through the intake-induced flow, which changes the waves
around the hull. The altered pressure distribution around the intake
opening on SP is noticeable comparing to BHSP.
The missing intake-opening surface causes a decrease in the wetted
surface area of the hull by 6:3% resulting in 4:3 N decrease in the
frictional resistance of the hull.
Total effects: The total resistance increment is obtained by comparing
the resistances of SP and BH. As seen in Fig. 24, like for the other speed
ranges, the total resistance increment is equal to the increment which is
obtained through the superposition of the global and local effects. The
sum of all resistance increment components vary such that the total
resistance increases in the higher speed range. Among all the compo-
Fig. 24. Schematic presentation of the hull and the transom resistance nents, the hydrodynamic pressure resistance of the hull is the dominating
increment components at Fn ¼ 1:4. The bare hull resistance at this speed
one.
is 141:5 N.

of SP and BHSP and integrating the hydrodynamic pressure on the area of 6.2. Difference between t and tr
the intake-opening of BHSP results in the same number. This can be
considered as a confirmation that the intake-opening is the only impor- So far, the causes of thrust deduction fraction variations in different
tant factor which influences the hydrodynamic pressure resistance of the speed ranges have been discussed through analyzing the resistance
hull. increment fraction, which dictates the trend of the t-curve (Fig. 10). The
The missing surface of the intake-opening decreases the frictional gap between the t and tr curves is caused by the difference between the
resistance of SP (1:2 N). waterjet net thrust and gross thrust and as shown in Equation (6) is equal
Total effects: Like in the lower speed range, in the intermediate speed to tj  tr tj . The difference between the waterjet net thrust and gross thrust
range the sum of the resistance increment components due to the global is discussed in Eslamdoost et al. (2016), where the difference is divided
and local effects are equal to the directly computed resistance increment into two main components, intake drag, Di , and exit drag, De . Intake drag
components. At this operating Froude number, the hull hydrostatic, hy- is the x-component of the force acting on the capture area and the
drodynamic and frictional resistances vary such that the sum of all these stream-tube and exit drag is the x-component of the force acting on the
variations becomes almost zero, and that is the reason for obtaining near nozzle exit (see Fig. 2). The jet thrust deduction fraction, tj , is obtained
zero resistance increment fraction in the intermediate operating speed from the difference between the gross and net thrust divided by the gross
range. thrust (implicit formula),

Tg  Tnet
6.1.3. High speed range (1:0 < Fn < 1:4) tj ¼ : (9)
Tg
The resistance increment fraction gradually increases and has positive
values in this speed range (Fig. 10). The BH, SP and BHSP resistance It can also be obtained from the sum of the intake and exit drag
increment components at Fn ¼ 1:4 are given in Fig. 24. Each of these (explicit formula),
resistance increment components is discussed in the following
subsections. Di þ De
tj ¼ : (10)
Global effects: The sinkage of BHSP is larger compared to BH (11 mm Tg
deeper at mid-ship, see Fig. 5) and due to the bow down trimming The jet thrust deduction fraction using both the implicit and explicit
moment of the waterjet the hull trims less (0:2 , see Fig. 6). The com- formulas is shown in Fig. 25. Despite the fact that the implicit tj is
bination of these two effects results in an increased hydrostatic pressure
on the hull. Due to the trim angle of BHSP at this speed ( 3:9 ), the
x-component of the surface normal points forwards and therefore the
resultant force caused by the increased hydrostatic pressure on the hull
will point backwards resulting in a slight increase of the hull hydrostatic
pressure resistance (þ0:8 N).
As stated in the previous paragraph, BHSP sinkage is larger compared
to BH which increases the displacement. The increased hydrodynamic
pressure resistance due to the global effects (þ14:9 N) is a result of the
increased displacement.
The wetted surface area of the BHSP increases by almost 5.9% because
of the changed sinkage and trim of the hull. This results in þ2:3 N in-
crease in the frictional resistance of the hull. Fig. 25. Jet thrust deduction fraction (Eslamdoost et al., 2016).

110
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

almost constant and has values around zero. The hull has rather large
trim angles in this speed range ( 4 bow up), so the hydrostatic pressure
force acting on the intake opening will have a horizontal component
backwards. This surface is not present in self-propulsion and therefore
the hydrostatic pressure resistance is reduced. There is however an in-
crease in the hydrodynamic pressure resistance of the hull caused by the
increased displacement (sinkage). The net result is a resistance increment
close to zero.
In the high speed range the resistance increment fraction gradually
increases. Like in the intermediate speed range there is a decrease in the
hydrostatic pressure resistance due to the missing intake-opening surface
Fig. 26. Thrust deduction fraction computed from three different methods. and an increase in the hydrodynamic resistance due to increased sinkage.
However, the latter effect gradually becomes more important, which
obtained from subtraction of two large numbers and the explicit one is causes the increase in the total resistance.
obtained from summation of two small numbers, both approaches predict Finally, it should be emphasized that the negative thrust deduction
almost the same value. Except in the low speed range, where the nozzle fraction in the intermediate speed range is not due to reduced hull
exit is not ventilated, tj is small. The resistance increment fraction is also resistance. In fact, the waterjet-propelled hull resistance remains rather
small in this speed range (Fig. 10) and therefore the second order term in unaffected compared to the bare hull resistance. The major reason for
Equation (6) becomes negligible, as assumed by The ITTC Waterjet negative thrust deduction is the jet system thrust deduction fraction, tj .
Specialist Committee (ITTC 24, 2005). However, this is not a valid
assumption if the waterjet is operating in the low speed range, where the Acknowledgment
second order term becomes as important as the other terms in Equation
(6). The research presented has been sponsored by Rolls-Royce Marine
We have all the components on the right hand side of Equation (6), through the University Technology Centre at Chalmers. The measure-
and therefore the total thrust deduction fraction obtained through this ment data presented were obtained from SSPA. Computing resources
equation can be compared with t computed directly from its definition were provided by C3SE, Chalmers Centre for Computational Science and
(Equation (3)). The results of these two approaches, as well as the case Engineering.
which the second order term neglected are shown in Fig. 26. The thrust
deduction fraction computed from these different approaches are in a References
good agreement with each other. This verifies the approach that we have
Brown, Matz, 2013. Model tests of a complete waterjet-hull system. In: SSPA Report
used for decomposing the components of thrust deduction fraction. Also, RE40084817-16-00-b.
the importance of taking the second order term in Equation (6) into ac- Bulten, Norbert Willem Herman, 2006. Numerical Analysis of a Waterjet Propulsion
count in the low speed range is clearly seen in this figure. Disregarding System. PhD Thesis. Eindhoven University of Technology.
Coop, Hamish George, 1995. Investigation of Hull-Waterjet Interaction Effects. PhD
this term will result in under-prediction of t at lower speeds. Thesis. Department of Mechanical Engineering, University of Canterbury.
Eça, Luis, Hoekstra, Martin, 2014. A procedure for the estimation of the numerical
uncertainty of CFD calculations based on grid refinement studies. J. Comput. Phys.
7. Conclusions 262, 104–130.
Eslamdoost, Arash, 2014. The Hydrodynamics of Waterjet/Hull Interaction. PhD Thesis.
The motivation of this work was to understand the reasons behind the Shipping and Marine Technology, Chalmers University of Technology.
Eslamdoost, Arash, Larsson, Lars, Bensow, Rickard, 2013. A pressure jump method for
varying thrust deduction fraction at different speed ranges. Thus, the
modeling waterjet/hull interaction. Ocean. Eng. 88, 120–130.
thrust deduction was investigated in detail over a large Froude number Eslamdoost, Arash, Larsson, Lars, Bensow, Rickard, 2014. Waterjet propulsion and thrust
range. It was shown that the main contribution to the thrust deduction deduction. J. Ship Res. 58 (4), 1–15.
fraction is the resistance increment fraction, whose variation can be Eslamdoost, Arash, Larsson, Lars, Bensow, Rickard, 2015. On transom clearance. Ocean.
Eng. 99, 55–62.
explained by analyzing the resistance components for the low, interme- Eslamdoost, Arash, Larsson, Lars, Bensow, Rickard, 2016. Net and gross thrust in waterjet
diate and high speed ranges. propulsion. J. Ship Res. 60 (2), 78–91.
The resistance increment due to the global flow (i.e. sinkage and trim ITTC 17, 1987. Report of the High-speed Marine Vehicle Committee. The Society of Naval
Arctitects of Japan, Kobe, Japan.
variations) and local flow (waterjet induced flow) were computed inde- ITTC 21, 1996. The Specialist Committee on Waterjets: Final Report and
pendently. Interestingly, this study showed that the global and local ef- Recommendations to the 21st ITTC.
fects independently influence the total resistance increment. ITTC 22, 1998. The Specialist Committee on Waterjets Final Report and
Recommendations to the 22nd ITTC.
In the low speed range large positive thrust deduction fraction values ITTC 23, 2002. The Specialist Committee on Validation of Waterjet Test Procedures: Final
are computed. This is due to the increase in the hydrodynamic pressure Report and Recommendations to the 23rd ITTC.
resistance of the transom (larger suction) as well as the increase in the ITTC 24, 2005. The Specialist Committee on Validation of Waterjet Test Procedures: Final
Report and Recommendations to the 24th ITTC.
hydrostatic pressure resistance of the hull. The former is caused by the jet Kandasamy, Manivannan, Georgiev, Svetlozar, Milanov, Evgeni, Stern, Frederick, 2011.
entrainment which reduces the hydrodynamic pressure at the transom Numerical and experimental evaluation of waterjet propelled delft catamarans. In:
and the latter is due to the increase transom submergence. 11th International Conference on Fast Sea Transportation (FAST 2011). Honolulu,
Hawaii, USA.
The transoms of the bare hull and the self-propelled hull clear the
Kandasamy, Manivannan, Keat Ooi, Seng, Carrica, Pablo, Stern, Frederick, 2010. Integral
water near the hump speed. The latter clears at a lower Froude number force/moment waterjet model for CFD simulations. J. Fluid Eng. 132 (10), 101103.
mainly due to the jet entrainment. There is a peak in the resistance co- https://doi.org/10.1115/1.4002573.

Kruppa, C., Brandt, H., Ostergaard, C., 1968. Wasserstrahlantriebe Für
efficient curve at the critical Froude number. The resistance difference
Hochgeschwindigkeitsfahrzeuge. In: Jahrbuch Der STG, vol. 62, pp. 228–258.
between the hulls is the largest at the peak where the waterjet-driven hull Kamal, Mustaffa, Iwan, Konrad Zürcher, Bose, Neil, Binns, Jonathan, Chai, Shuhong,
transom becomes dry. Then the difference diminishes rapidly with speed Davidson, Gary, 2015. Powering for medium speed wave-piercing catamarans
until the bare hull transom clears the water. This causes the rapid drop of comparing waterjet and screw propeller performance using model testing. In: Fourth
International Symposium on Marine Propulsors. Austin, Texas, USA.
the thrust deduction fraction observed near the hump speed.
In the intermediate speed range the resistance increment fraction is

111
A. Eslamdoost et al. Ocean Engineering 152 (2018) 100–112

Muzaferija, Samir, Peric, Milovan, 1999. Computation of free surface flows using speed ship. J. Mar. Sci. Technol. 16 (4), 434–447. https://doi.org/10.1007/s00773-
interface-tracking and Interface- capturing methods. In: Nonlinear Water Wave 011-0138-x.
Interaction. WIT Press, Southampton, pp. 59–100. van Terwisga, Tom, 1996. Waterjet-Hull Interaction. PhD Thesis. Delft Technical
Rispin, Paul, 2007. Waterjet Self-propulsion Model Test for Application to a High-speed University.
Sealift Ship, 6300 State University Drive, Suite 220, Long Beach, CA 90815. White, Frank M., 2008. Fluid Mechanics, Sixth. McGraw-Hill.
Takai, Tomohiro, Kandasamy, Manivannan, Stern, Frederick, 2011. Verification and
validation study of URANS simulations for an axial waterjet propelled large high-

112

You might also like