You are on page 1of 20

Ocean Engineering xxx (2017) 1–20

Contents lists available at ScienceDirect

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

Design of ducted propeller nozzles through a RANSE-based


optimization approach
Stefano Gaggero a, *, Diego Villa a, Giorgio Tani a, Michele Viviani a, Daniele Bertetta b
a
University of Genoa, Department of Electrical, Electronic, Telecommunications Engineering and Naval Architecture, Via Montallegro, 1, 16145 Genoa, Italy
b
Fincantieri S.p.A. Naval Vessel Business Unit, Naval Architecture Dept. (MM-ARC), Genoa, Italy

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

Keywords: Marine propellers design requirements are always more pressing and the application of unusual propulsive
Ducted propellers configurations, like ducted propellers with decelerating nozzles, may represent a valuable alternative to fulfill
Accelerating nozzles stringent design constraints. Accelerating duct configurations were realized mainly to increase the propeller ef-
Decelerating nozzles ficiency in the case of highly-loaded functioning. The use of decelerating nozzles sustains the postponing of the
Optimization cavitating phenomena that, in turn, reflects into a reduction of vibrations and radiated noise. The design of
RANSE
decelerating nozzle, unfortunately, is still challenging. The complex interaction between the propeller and the
OpenFOAM
nozzle, both in terms of global flow feature and local (tip located) phenomena, is not yet fully understood. No
extensive systematic series, as in the case of accelerating configurations, are available and the design still relies on
few measurements and data. On the other hand, viscous flow solvers appear as reliable and accurate tools for the
prediction of complex flow fields and their application for the calculation of ducted propeller performance and
nozzle flow was almost successful. Hence, using CFD as a part of a design procedure based on optimization, by
combining a parametric description of the geometry, a RANSE solver (OpenFOAM) and a genetic type algorithm
(the modeFrontier optimization environment), is the obvious step towards an even more reliable ducted propeller
design. An actuator disk model is adopted to include efficiently the influence of the propeller on the flow around
the duct; this allows avoiding the weighting of the computational effort that is necessary for the calculations of the
thousands of geometries needed for the indirect design by optimization. Design improvements, in model scale, are
measured by comparing, by means of dedicated fully resolved RANSE calculations, the performance of the
optimized geometries with those of conventional shapes available in literature. For both nozzle typologies,
dedicated shapes reducing the risk of cavitation and increasing the delivered thrust are obtained, showing the
opportunity of customized nozzle design out of usual systematic series. In addition, by analyzing the results of the
optimization histories, appropriate design criteria are derived for both accelerating and decelerating nozzle
shapes.

1. Introduction systems. Their ability to reduce side effects like induced pressure pulses
and vibrations by operating the propellers in more uniform inflow, with
Marine propellers design requirements are, nowadays, always more minimum optimal diameters and far from the hull, has been proven over
pressing. Not only maximum efficiency but also comfort and environ- the years.
mental demands and regulations have to be satisfied. The application of Ducted propellers represent an additional, valid, answer to current
unusual propulsive configurations may represent a valuable alternative design requirements. In the case of vessels for which requirement of high
to fulfill these constraints. Contra-rotating and tip loaded blades (CLT thrust is critical in the low-speed operation range, or when the screw is
and Kappel like geometries) were used to improve propulsive efficiency, limited in diameter, the use of accelerating nozzles is widely documented
together with Energy Saving Devices, like pre- and post-swirl stators, in literature. With accelerating nozzles, the duct increases the flow rate
PBCF or Mewis Ducts. Combinations of these devices mounted on PODs through the propeller, which consequently operates at a more favorable
(i.e. through the EU project TRIPOD, Sanchez-Caja et al., 2013) were loading. The nozzle by itself produces a certain positive thrust.
exploited to maximize the effects and to provide flexible propulsive This conceptual idea dates back to Stipa (1932) and Kort (1934) but

* Corresponding author.
E-mail address: stefano.gaggero@unige.it (S. Gaggero).

https://doi.org/10.1016/j.oceaneng.2017.09.037
Received 14 February 2017; Received in revised form 21 July 2017; Accepted 24 September 2017
Available online xxxx
0029-8018/© 2017 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Gaggero, S., et al., Design of ducted propeller nozzles through a RANSE-based optimization approach, Ocean
Engineering (2017), https://doi.org/10.1016/j.oceaneng.2017.09.037
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

probably the works by Van Manen (1954, 1957, 1962), Van Manen and Recently Bontempo et al. (2014, 2016) proposed, again in the
Superina (1959) and Van Manen and Oosterveld (1966) represent some framework of potential flow based theories, some analyses of deceler-
of the most accurate reviews of ducted propellers performance and ating duct shapes with momentum theory and semi-analytical actuator
design guidelines. By combining even simple theoretical considerations disk models to account for radial distributions of forces, including
and systematic measurements, they provided a rather wide overview of tangential ones, and derive relationships between propeller load and
the influence of some geometrical parameters of nozzles on efficiency, nozzle performance. All these calculations, however, neglect again the
correlating non-dimensional parameters and performance in case of both role of viscosity. This, instead, significantly influences duct and propeller
highly- and lightly-loaded propellers. They identified the nozzle profile forces by interacting with the flow at the duct/propeller tip gap giving
number 19A as the optimal compromise, having performance in towing rise to typical tip leakage vortexes, as demonstrated by Gaggero et al.
and pushing conditions not appreciably inferior to those of considerably (2013), Haimov et al. (2011), Sanchez-Caja et al. (2008),
longer nozzles. This shape, fifty years later, is still the default choice in Abdek-Maksoud and Heinke (2002), Hoekstra (2006), Baltazar et al.
the case of accelerating ducted propellers applied to a variety of boats (2013) also in the simpler case of accelerating nozzles.
and vessels (tugboats, towboats, and supply vessels) which require In the context of accelerating duct shapes, instead, significant de-
increased propulsive efficiency, especially near the bollard velopments and applications can be recognized. Since the work of Kerwin
pull condition. et al. (1987) and Hughes (1993, 1997), several models have been pro-
In the same works, also the potentialities of flow decelerating type of posed to deal partially with the design and particularly to the analysis of
nozzle were pointed out. By reducing the flow rate to the impeller, a local accelerating ducted propellers. Boundary Element Methods were pro-
increase of the static pressure is achieved: this may be effective in posed to efficiently account for vortical wake alignment (Baltazar et al.,
retardation of propeller cavitation phenomena at a cost, unfortunately, of 2015; Kinnas et al., 2012a), influence of duct boundary layer on tip
efficiency reduction and of an additional drag represented by the nega- propeller loading (Baltazar et al., 2012) or presence of the blunt trailing
tive thrust delivered by the duct itself. An improvement of the cavitation edge (Kinnas et al., 2009, 2012b; Baltazar et al., 2015), typical of
characteristics of the propeller can be obtained, thus, only if the gain in accelerating configurations but critical for the application of BEM. Good
static pressure exceeds the unfavorable effect of the increased screw predictions of propulsive forces were achieved using RANSE, which
loading, necessary to compensate the duct drag. In addition, the risk of encouraged direct full-scale applications or the definition, through nu-
duct cavitation (at leading edge or midchord, depending on the propeller merical analyses, of scaling rules (Bhattacharyya et al., 2016a, 2016b).
loading and the camber of the nozzle) represents another issue, which Simultaneously the combined use of RANSE and simplified propeller
has to be accounted for when evaluating the usefulness of the deceler- models (i.e. actuator disks) highlighted the usefulness of such calcula-
ating configuration. tions for a preliminary design of the duct. Zondervan et al. (2006),
The design of decelerating nozzles consequently seems to better Hoekstra (2006), Tamura et al. (2010), for instance, successfully applied
comply with the constraints circa radiated noise and vibrations strictly an actuator disk to investigate the influence of the propeller on the flow
related to cavitation suppression. However, it is still challenging and around the nozzle and to derive simple correlations circa induced ve-
poorly addressed in the literature. Despite the wide employment of this locities useful in the design stage. Kinnas et al. (2013) and Bosschers et al.
propeller configuration (i.e. fast carriers and navy vessels for which side (2015) further extended the computational model by including the pro-
effects reduction is significantly more important than propulsive effi- peller action through distributed momentum sources (i.e. body forces) on
ciency), there are several design and analysis issues still open, starting Euler or RANSE solvers (as done for self-propulsion analyses by Villa
from the downsides previously evidenced. The complex interaction be- et al., 2011). This inclusion efficiently solved the interactions (separation
tween the propeller and the nozzle, both in terms of global flow feature of the flow on the inner duct surface or thickening of the viscous
and local, tip located, phenomena, is not yet fully understood. No sys- boundary layer) between propeller and duct improving, at a relatively
tematic series, widely diffused in the case of accelerating configurations, low computational cost, the reliability of both potential based (for the
are available; thus, the design still relies on few measurements and data. propeller) and viscous based (for the duct) solvers. Only propellers inside
Oosterveld (1970) carried out some measurements with decelerating Nozzle 19A were, however, analyzed without any attempt to improve the
ducts where the influence of some geometrical features of the duct was overall propulsive performance by modifying the duct shape.
investigated. The analyses were however limited to a single propeller Considering the need of a reliable and efficient design tool for both
geometry (a finite-chord Kaplan propeller, model K5-100) operating in- accelerating and decelerating propeller nozzles, the scarcity of data and
side the duct. Only qualitative trends can thus be derived for the appli- design guidelines, and the satisfactory results achieved by numerical
cation of those nozzles in the cases of unloaded (zero chord at the tip) analyses, in the present work optimization is exploited to investigate the
propellers. From a theoretical/numerical point of view, the availability of possible enhancement of nozzles performance provided by the applica-
literature data is even scarcer. Apart classical momentum theories tion of high-fidelity numerical tools embedded into an automatic process.
(Oosterveld, 1970; Van Manen and Oosterveld, 1966) applied to decel- Thousands of different geometries, defined by a parametric description,
erating duct geometries, only a few applications are available. Abdel-- need to be tested, using viscous approaches to account for the relevant
Maksoud et al. (2010) developed a design and analysis (Yu et al., 2013) phenomena characterizing “thick” ducts and verified against appropriate
method for multi-component propulsors. Çelik et al., 2011 investigated design criteria (specifically derived for both accelerating and deceler-
the optimum duct geometry for passenger ships. Gaggero et al. (2012, ating configurations). This leads to the definition of an optimal geometry
2013) developed a design approach based on classical lifting line/lifting that, due to the multi-objective nature of design, is, of course, a balance of
surface approaches further refined by using Boundary Element Methods contrasting purposes. Only thank to the favorable ratio between accuracy
and optimization algorithms to improve the performance of a propeller and computational efficiency of simplified representations of the influ-
surrounded by a decelerating nozzle. Results, confirmed by dedicated ence of the propeller on the nozzle performance, such the ones repre-
RANSE calculations and model scale measurements at the cavitation sented by actuator disks, the process is possible in a reasonable
tunnel, concerned however only the well-established problem of pro- computational time compatible with the usual design routine. Actuator
peller blades design, using given nozzle shapes. The inclusion of the disk models, already proposed with success for the analysis of acceler-
nozzle shape in the optimization process, indeed, would have been ating nozzles performance, are at first validated on the basis of the
possible. However, it would have excessively stressed the applicability of measurements carried out at the cavitation tunnel of the University of
BEM over its inherent limitation (thick geometries, subjected to flow Genoa. Then, they are employed also for the analysis of decelerating
separation), leading to nozzle designs significantly conditioned by the ducts and, through optimization, for the improvement of the perfor-
scarce reliability of the computational approach in dealing with complex mance of reference nozzle geometries. The first purpose of this paper is to
viscous flows. demonstrate the reliability of the proposed design approach. Then, by

2
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

using the amount of numerical data provided by the systematic analyses nozzles and propellers by themselves: an optimal duct (depending on its
inherent to any optimization process, it is possible to identify correlations accelerating or decelerating nature) to be adopted for the iterative defi-
and trends among parameters and performance, providing useful nition of an optimal propeller for those given conditions of flow. The use
guidelines in the peculiar case of decelerating configurations. of actuator disk models (as proposed in this work) to include the pro-
The successful application of the design by optimization of model peller into the calculations of the duct performance, indeed, does not
scale nozzles by using the simpler actuator disk model would represent a permit any type of interactions on propeller forces being these ones
significant step towards a more reliable design approach. Once combined prescribed and independent by the performance (and the resulting
with any potential based design approach for the propeller (a simpler induced inflow) of the nozzle. Only by a blade redesign, it would be
lifting line or a BEM, as proposed in Gaggero et al., 2012, 2013) the possible to have a sort of feedback and to design the entire propulsive
hybrid BEM/RANSE would extend the reliability of current geometry- system for a given total thrust. Similarly, by using a lifting line or a more
optimization algorithms for the design of the entire propulsive system accurate BEM nested into the design loop, it would be possible to include
directly in full scale: RANSE methods would account for flow separation the changes of the propeller loading distribution due to the different
phenomena on the “thick” duct, whereas BEM will be used for prediction inflow of a modified nozzle shape in the optimization process of the
of forces on the “thin” propeller blades where separation is less probable. nozzle. This approach, of course, would still not consider the constraint
on the total delivered thrust. In both cases, being duct and propeller
2. Designs of ducts through optimization performances dependent on each other (loading distribution changes
inflow which, in turn, changes loading distribution), additional inner
The design of propeller nozzles, despite the promising results ach- iterations would be required for any given nozzle shape. By assuming an
ieved with numerical calculations, is mainly based on systematic series. unchanged propeller thrust (i.e. unchanged actuator disk intensity),
Van Manen (1962), Van Manen and Oosterveld (1966) derived simplified instead, the analysis of the optimal geometries has to be restricted to
correlations and identified key points for both accelerating and decel- those nozzle shapes that delivering the thrust of the reference geometry
erating shapes considering efficiency, the risk of flow separation and (i.e. not changing the working condition), improve performance in terms
critical cavitation indexes for both the propellers and the surrounding of inner static pressure. This inherently implies that also nozzle induced
nozzles. In the case of accelerating duct geometries, a substantial increase velocities to the propeller are assumed not to be reasonably affected by a
of efficiency was obtained at higher screw loading with relatively long modified geometry that, having to provide the same thrust of the refer-
nozzles that instead have to be the shortest possible in lightly loaded ence shape, should also be significantly similar in terms of induction. Any
conditions to limit the nozzle drag which is substantial in reducing the other case (i.e. an increase of the duct thrust at unchanged inner static
total propulsive efficiency. Similar analyses were proposed to investigate pressure or heavily changed velocity field to the propeller) has to be
the influence of maximum thickness and maximum camber over length regarded as the first step towards the redesign of propeller blades. In this
ratio and equivalent nose tail angle on the pressure on the outer duct case, the propeller design will take advantage of the unloading which is
surface, controlling flow separation as well. Guidelines for decelerating allowed by the higher thrust delivered by the newly designed duct.
geometries were even scarcer. A favorable increase of the static pressure A list of possible design objectives that could be exploited consists,
at the impeller was evidenced only at low propeller thrust coefficients consequently, in:
with shorter nozzle length to minimize drag loss and the consequent
reduction of efficiency.  For accelerating duct shapes, conceived for highly loaded
These key points turn, naturally, into the opposite objectives of a functioning:
design through numerical optimization. For an accelerating nozzle, ○ Maximization of the nozzle thrust close to bollard pull condition
which is usually designed to operate close to bollard pull, the maximi- ○ Minimization of the nozzle drag at higher advance coefficients,
zation of the duct thrust at a given propeller loading conflicts with the when the ducted propeller provides the propulsive thrust for other
need of reasonably high values of inner static pressure to contrast cavi- operating conditions of the ship (e.g. cruising speed for tugs; other
tation inception. Most of the thrust delivered by the duct depends by the possible “mixed conditions” like those typical of a take home
suction on the inner surface. In bollard pull conditions, a certain risk of propulsor, which needs higher thrust density than a conventional
cavitation could be accepted (until the propeller does not suffer from one, with the necessity to operate at a certain speed)
excessive thrust breakdown). In the case of functioning at higher advance ○ Maximization of the inner static pressure in any operating
coefficients, a geometry that allows simultaneously to maximize the duct conditions
thrust and to obtain reasonable values of static pressure against cavita- ○ Avoidance of nozzle cavitation
tion inception would be preferred. On the contrary, for a decelerating  For decelerating duct shapes:
duct propeller operating at relatively high advance coefficients, maxi- ○ Minimization of the nozzle drag at the design condition
mization of the static pressure with minimum increased duct drag turns ○ Maximization of the inner static pressure at the design condition
into the design objective. Avoiding simultaneously the risk of cavitation ○ Avoidance of nozzle cavitation
on the outer surface of the nozzle in very critical working conditions has
to be considered as well. In parallel, the risk of flow separation has to be A design by optimization, definitely, can be explained as an automatic
accounted. This has been monitored experimentally and addressed in the “trial and error” procedure that, by analyzing hundreds (thousands) of
traditional and simplified design momentum theories with criteria on the different geometrical configurations, iteratively changes their shape to
maximum sectional lift coefficient of the nozzle profile (Van Manen achieve a Pareto convergence of the designs towards the satisfaction of
(1962), Van Manen and Oosterveld (1966)). The analysis of the duct the outlined objectives. For this purpose, an efficient tool for the char-
forces via viscous calculations inherently includes the possible detri- acterization of flow features and nozzle performance by including the
mental effects of flow separation on duct performance. propeller action in the simplified form of an actuator disk (Sections 3.1
A real design would require accounting for the interactions taking and 3.2) is necessary. A parametric description of the geometry to
place between the duct and the propeller to exploit the maximum from manage automatically shape variations in accordance with given con-
both the designs: a nozzle providing the maximum thrust and the highest straints (Section 3.3) is required as well. The need to handle opposite
possible level of inner static pressure and a propeller optimized (effi- objectives requires the use of multi-objective optimization algorithms
ciency, thrust and cavitation, depending on the application) for that that, in present calculations, are of genetic type. By exploiting the anal-
resulting flow. Such design is, of course, possible at a considerable ogy of natural selection, subsequent generations are created by
computational cost by analyzing the propulsive system as a whole. combining the best characteristics of each pre-computed combinations
However, it could take advantage of preliminary, independent, designs of via crossover and mutation. These algorithms, moreover, allow to

3
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

avoiding local minima, with respect to which computationally efficient formulation exploits “wedge type” boundary conditions for purely
gradient-based methods are more prone, by the inclusion of a certain axisymmetric calculations. For 2-dimensional axisymmetric cases, the
randomness in the selection of the characters of any subsequent gener- geometry is specified as a wedge of small angle (<5 ) with a thickness of
ations allowing more unrestrained analyses. This feature was considered 1 cell, running along the plane of symmetry, straddling one of the co-
important for the analysis of trends and of the influence of combinations ordinate planes, as shown in Fig. 1b. With this arrangement, only mo-
of parameters. The entire process has been coded into the modeFrontier mentum sources in the axial directions are allowed and all the flow
(Esteco, 2016) optimization environment. features different from those on the meridian plane (i.e. slipstream
rotation) are neglected. The choice of the most appropriate computa-
3. Numerical methods tional domain depends on the influence of the slipstream rotation on duct
forces. The analyses by Hoekstra (2006) showed a certain importance of
3.1. Flow solver and computational domain slipstream which was calculated using a formulation able to account for
the tangential effect also by using only an equivalent two-dimensional
Detailed analyses of duct performance cannot be achieved by simply meridian plane. In the present case, the need of computationally effi-
using inviscid flow models. Flow separation, except by using heavily cient calculations would encourage, on the contrary, the use of the effi-
approximated formulations, is beyond potential flow calculations. Blunt cient setup consisting in the 2-dimensional wedge formulation, thus
trailing edges, in the particular case of accelerating ducts, pose significant neglecting the 3-dimensional nature of the flow. The preliminary analysis
difficulties for the application of the Kutta condition. Local modification proposed for the reference geometries in Section 4 would definitely
of the duct shape (Bosschers et al., 2015) or inclusion of transpiration clarify the influence of slipstream in the case of the calculations carried
sources have been proposed but the necessary calibration of such models with OpenFOAM, allowing for the selection of the most appropriate
is far from the accuracy needed even in comparative analyses as those setup. As proposed by Hoekstra (2006), also the propeller hub has been
represented by design by optimization. Viscous models are definitely included in the calculations. Its presence, in fact, has a certain positive
needed and RANSE approximations with suitable turbulence models can influence on the thrust delivered by the duct: momentum sources are
adequately answer all the specific issues related to design procedures. distributed over a small area, inducing higher velocities that change the
In this study, OpenFOAM (Weller et al., 1998; the OpenFOAM angle of attack of the nozzle and the resulting pressure distribution.
Foundation, 2017) has been employed. OpenFOAM is a collection of li- Details of the hub shape are also important. The use of semi-infinite (aft
braries suited for the solution of partial differential equations using a or forward) rather than infinitely long shafts has a different influence on
Finite Volume approach and cell-centered collocated variables with flow streamlines affecting the performance of the nozzle. In present
face-based implementation to allow for arbitrary cell shapes. calculations, the focus is on the definition of a procedure for nozzles
Pre-processing, solvers, and post-processing tools are fully scriptable, design rather than on the exact characterization of the flow. Conse-
which is of fundamental importance when a completely automatic pro- quently, the simplest adoption of an infinitely long shaft was deemed
cess has to be setup. sufficient to account at least for its principal influence. Similarly, a slip
For this particular application, where steady flow with respect to a boundary condition, without the need of appropriate prism layers, was
blade fixed reference frame is expected, the simpleFOAM solver with the considered sufficient to catch its influence on the flow to the nozzle. For
SST k-ω turbulence model (Menter et al., 2003) has been used. Continuity both cyclic and wedge calculations, anyhow, the computational domain
and RANSE equations are solved in a segregated form, with a SIMPLE consists of an angular sector (25 when periodic boundary conditions are
based pressure-velocity correction and under-relaxation to achieve employed, 5 in the case of the “wedge” formulation) with the inlet
steady state solutions. Second order accurate schemes in space have been placed 2 propeller diameters in front of the nozzle and the outlet 5
preferred for momentum and turbulence equations. propeller diameters aft the nozzle. The external boundary lies on a cy-
The use of actuator disks instead of actual propeller blades results in a lindrical surface 4 propeller diameters from the propeller shaft. A sum-
fully axisymmetric flow solution. Therefore, the computational domain mary of the boundary conditions for both the configurations is reported
can be reduced to a blade sector smaller than that of a blade passage. in Table 1.
In OpenFOAM, there are two possible formulations to exploit the
rotational symmetry, which however condition the type of flow that can
be modeled. The first consists in a 3D angular sector of sufficient 3.2. Propeller model and grid arrangement
amplitude (Fig. 1a), arranged with periodic (cyclic) interfaces allowing
the inclusion of the propeller slipstream rotation by means of momentum The influence of the propeller on the flow around nozzles has been
sources representative of the forces acting on the propeller blade in the modeled by using an actuator disk. The actuator disk can be seen, defi-
tangential direction. Periodicity requires the use of a certain number of nitely, as an infinite bladed propeller. The load of the propeller is
cells in the tangential direction (as in Section 3.2) to take properly into distributed over the entire area of the propeller, neglecting the real
account the tangential component of the momentum. The second material nature of a finite number of blades. Even if more detailed rep-
resentations of blade forces are available (Bosschers et al., 2015), by

Fig. 1. Computational domains for the analysis of ducted propeller performance. (a) Cyclic type boundary conditions. (b) Wedge type boundary conditions.

4
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Table 1
Summary of boundary conditions.

Patch Velocity Pressure ω k


∂p Fixed (depending on U), Fixed (depending on U),
Inlet Fixed
∂n ¼0
(depending on J) 1% turbulence intensity 1% turbulence intensity
∂U p¼0 ∂ω ∂k
∂n ¼ 0 ∂n ¼ 0 ∂n ¼ 0
Outlet
∂U p¼0 ∂ω ∂k
External ∂n¼0 ¼0
∂n ¼0
∂n
∂p
Nozzle No-slip
∂n ¼0 Wall Function Wall Function
Shaft/hub Slip Slip Slip Slip
Interfaces Wedge/Cyclic Wedge/Cyclic Wedge/Cyclic Wedge/Cyclic

using local momentum sources, their application for the characterization direction) could have a certain influence on duct performance. Blade
of flow in the nozzle design process seems redundant. Only the radial forces are largely due to suction on the back side, while the pressure jump
distribution of load is, in fact, really influential. The tip loading has a due to a momentum source is evenly distributed between over- and
positive influence on duct forces; this is largely confirmed by the pref- under-pressure. The unsymmetrical distribution of pressure exerted by
erence, in the case of accelerating nozzle shape, of Kaplan-like propeller the propeller blades could have been taken into considerations by
geometries, with a finite chord at tip fostering the development of higher employing a sequence of actuator disks of different magnitude aiming at
forces at outer radii. Unloaded tips introduce a substantial risk of flow reproducing a sort of chord-wise distribution of load in order to partially
separation on the inner surface of the duct (which can be further enlarged address the uneven pressure distribution between sides of the blade. This
by the simplified model represented by the actuator disk). This effect can arrangement, definitely needed when the interactions between pressure
be mitigated by dedicated nozzle shapes, as those that for instance may fields are of primary importance (i.e. for the evaluation of the thrust
be derived from an integrated design approach as the one proposed in deduction during self-propulsion calculations) has, however, a certain
this paper. computational cost that seems unnecessary in this phase.
Hoekstra (2006) proposed a simple three-parameter model to Grids were arranged accordingly to the computational domain
describe the propeller loading and to allow for variations of the radial selected for the analyses. A structured multi-block layout has been
position of the maximum thrust. Similar approaches are currently preferred to describe in details the most important features of the noz-
available in many commercial CFD software and often consider radial zles. In the case of accelerating duct configuration, essentially an O-grid
variations of the loading in accordance with Goldstein optimum criteria mesh has been arranged to have better resolution at the blunt trailing
(Goldstein, 1929). edge. C-type grids were adopted, instead, in the case of decelerating
In present calculations, instead, the radial distribution of forces (axial geometries to manage the sharp trailing edge of the nozzle. An example
and tangential, depending on the type of simulation and on the compu- of the meshes is shown in Fig. 3. Both the arrangements for cyclic and
tational domain, as in Section 4) have been derived from Boundary wedge type boundary conditions are presented. The meridian sections
Element Method calculations. They are subsequently loaded into RANSE show clearly the multi-block nature of the meshes, which were arranged
by using interpolating tables and polynomials through a dedicated li- with the OpenFOAM blockMesh utility. Meshes, in the case of the wedge
brary specifically developed to extend the computational functionalities arrangement, consist of a one cell thick layer of structured hexahedral
of OpenFOAM. For both the accelerating and the decelerating cases, the elements with appropriate grading to comply with wall treatment
optimization has been carried out starting from reference geometries modeling through wall functions. In the case of 3-dimensional calcula-
(Gaggero et al., 2012) designed and tested at the cavitation tunnel. tions with periodic boundary conditions, exactly the same cell arrange-
Exactly these geometries were used to obtain a numerical estimation of ment is used and the angular amplitude of the sector is filled with 25
the loading distributions and to define the design points in correspon- layers of cells.
dence to which the design of the new nozzles has been carried out. The The influence of the grid resolution in the case of the wedge type
same set of data served also for the preliminary validation of the boundary condition (that, as highlighted in Section 4, will be used for the
computational model to be employed in the optimization process. optimization activity) was investigated. In order to carry out this inves-
The proposed actuator disk is a finite thickness disk set equal to the tigation, a reference computational mesh (i.e. Medium) and two other
average longitudinal extension “cut out” by the propeller (Fig. 2). No meshes, one with reduced (Coarse) and one with increased resolution
investigations circa the influence of the disk thickness have been carried (Fine) arranged using a refinement factor of 2 have been considered.
out and the simplest possible model has been used in the light of Richardson extrapolation was applied to derive, in principle, extrapo-
computational efficiency. As discussed by Hoekstra (2006), also the lated values at zero grid spacing from a series of lower-order discrete
chord-wise load distribution (i.e. the distribution of forces in the axial values. Values of the nozzle thrust coefficient

Fig. 2. Actuator disk representations into the computational domain (cyclic or wedge). Colors represent the strength of the momentum sources per unit volume. (a) Cyclic type boundary
conditions. (b) Wedge type boundary conditions. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

5
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 3. Multi-block Structured mesh arrangements around typical accelerating and decelerating nozzles. (a) Accelerating type nozzle, cyclic b.c. (b) Decelerating type nozzle, cyclic b.c. (c)
Accelerating type nozzle, meridian plane, wedge b.c. (d) Decelerating type nozzle, meridian plane, wedge b.c.

(CT Nozzle ¼ TNozzle =0:5ρπR2prop: V 2 ) were analyzed at equivalent propeller and thickness distributions. In present calculations, however, a slightly
advance coefficients of J ¼ 0.8 (accelerating duct) and of J ¼ 103.4% of different approach has been adopted. In order to avoid unfeasible shapes
Jref. (decelerating duct). For both geometries, the convergence trend is (i.e. maintain the inner duct surface in correspondence to the propeller
particularly good. Especially for the accelerating duct configuration, the on a cylindrical surface of a given radius) it was preferred to describe
medium mesh arrangement allows for negligible errors with respect to with B-Spline curves directly the inner and the outer contour of the
the extrapolated value (Table 2). Also for the decelerating configuration, nozzle to facilitate the application of the geometrical constraints (Ver-
the difference is well below 1%. Particularly low values of the Grid nengo et al., 2016). The control points defining the B-Spline curves, as
Convergence Index (Roache, 1994), in addition, justify the use of the shown in Fig. 5, are free to move with some “compatibility” constraints.
Medium mesh arrangement in the light of computationally efficient and At leading edge (and at trailing edge, in the case of the sharp closure of
reliable calculations throughout the design by optimization. Finally, as decelerating nozzles) the control points of back and face sides have to
per Fig. 4, also the iterative convergence of both analyses with this grid move as one, and only in the radial direction. The control points that
setup is good and consequently 800 iterations were selected for all the handle the central part of the inner surface have to be altered, instead,
calculations of the optimization process. only axially to keep it rectilinear at propeller location.
With these assumptions, the description of the accelerating duct
consists of 11 free parameters while the decelerating nozzle has 7 degrees
3.3. Nozzles parametric description of freedom. In both cases, a scaling factor in the axial direction is added to
change anisotropically the length/diameter ratio. The ranges of variation
The parametric description of nozzles is rather simple. Nozzles are of the parameters are summarized in Table 3. Choices, also based on
definitely hydrofoils that can be handled, analogously to what already previous experiences, are motivated by the need of a trade-off between
done in the case of propellers (Gaggero et al., 2016a, 2016b) and wings allowed modifications (the widest possible) and fairness of the geometry,
(Olivucci and Gaggero, 2016), by using B-Spline curves for both camber which, in turn, is facilitated in any case by the use of B-Spline curves.
However, for design purposes this description is not intuitive and usual
Table 2 parameters, like the angle of attack of the nozzle, the maximum camber
Discretization errors for the nozzles. Wedge type boundary condition.
and thickness (plus their chord-wise position) or entrance and exit radius
Cells Accelerating Nozzle Decelerating Nozzle can be easily derived and monitored throughout the design process.
CT Nozzle CT Nozzle/CT ref.

Coarse 4400 0.1887 0.01948 4. Validation of the simplified propeller model


Medium 17,600 0.2381 0.01858
Fine 70,400 0.2379 0.01846
The validation of the computational tools (actuator disk model,
Extrapolated 0.2379 0.01844 wedge or cyclic type boundary conditions) has been carried out consid-
Approximated 0.10% 0.66% ering two ducted propellers for which measurements of forces and flow
relative error (Fine are available (Gaggero et al., 2012, 2013). The first one is an accelerating
– Medium meshes) duct propeller designed for a take home propulsor. It is equipped with a
Extrapolated relative 0.10% 0.76%
error for Medium
conventional Nozzle 19A. Rather than exclusively favor bollard pull ef-
Mesh ficiency, the design was carried out to increase the propulsive thrust of a
GCI 0.00% 0.13% relatively small propeller operating at a certain speed. The second one is

6
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 4. Iterative convergence (residuals and forces) of the solver. (a) Accelerating type nozzle, L1 residuals convergence. (b) Decelerating type nozzle, L1 residuals convergence. (c)
Accelerating type nozzle, thrust coefficient convergence. (d) Decelerating type nozzle, thrust coefficient convergence.

Fig. 5. Parametric description of nozzles shape. Accelerating duct case.

a decelerating duct propeller developed to postpone cavitation for ap- geometries, propellers are unloaded at the tip and have a small chord,
plications in which reduction of noise (or better, retardation of cavita- which made the selection of the optimal nozzle shape problematic for the
tion) is important. In both cases, differently from usual Kaplan like reasons discussed in the introduction. By using a Boundary Element

7
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Table 3
B-Spline control points on the local nozzle reference system, the range of variability and the compatibility conditions.

Accelerating Nozzles Decelerating Nozzles

x/L y/L x/L y/L

P1 inner 0 (fixed) 0.15–0.20 P1 inner 0 (fixed) 0.05–0.1


P2 inner 0 (fixed) 0.06–0.14 P2 inner 0 (fixed) 0–0.02
P3 inner 0.1–0.25 0 (fixed) P3 inner 0.3–0.55 0 (fixed)
P4 inner 0.329 (fixed) 0 (fixed) P4 inner 0.7–0.85 0 (fixed)
P5 inner 0.65–0.85 0 (fixed) P5 inner 1 (fixed) 0.08–0
P6 inner 1 (fixed) 0–0.04

P1 outer 0 (fixed) y/L P1 inner P1 outer 0 (fixed) y/L P1 inner


P2 outer 0 (fixed) 0.21–0.25 P2 outer 0 (fixed) 0.14–0.24
P3 outer 0.07–0.3 0.16–0.23 P3 outer 0.4–0.8 0.16–0.24
P4 outer 0.5–0.8 0.08–0.14 P4 outer 1 (fixed) y/L P5 inner
P5 outer 1 (fixed) 0.05–0.08

Length scaling 0.75–1.5 Length scaling 0.75–1.5

TNozzle
Method, the radial distributions (Fig. 6) of axial and tangential forces on CT Nozzle ¼1 (2)
2
ρπRprop: V
2 2
the propeller blades were derived and momentum sources were assigned
over the actuator disks coherently. Predicted forces exerted by the noz-
CT prop:
zles in the case of both cyclic type (for which also tangential forces have τ¼ (3)
been included in the analysis) and wedge type boundary conditions were CT prop: þ CT Nozzle

compared to those measured during experiments. Results, as shown in


Figs. 7 and 8, are summarized in two ways. As usual, thrust coefficient TNozzle:
KT ¼ (4)
values of propeller and nozzle are compared in terms of thrust ratio τ in ρN 2 D4
order to verify that the relative contribution of the duct to the total thrust
Results, for both the nozzle geometries, are in good accordance with
increases with increasing propeller loading. Coefficients, respectively CT
experimental data. The agreement between calculations and measure-
prop and CT Nozzle, are non-dimensional with respect to the inflow velocity
ments is good, especially from medium to heavy loading conditions. The
V and the area of the propeller disk. Also thrust coefficient KT, assuming
inclusion of the tangential forces as an additional momentum source
the revolution rate (N) of the propeller during experiments and its
distribution, when cyclic boundary conditions are adopted, has a favor-
diameter as reference to derive non-dimensional quantities, is used for
able effect on the numerical calculations. As pointed out in Hoekstra
comparison purposes.
(2006), the additional pressure drop in the swirling slipstream due to the
Tprop: modeling of tangential velocities increases the pressure difference be-
CT prop: ¼ 1 (1) tween the inner and the outer duct surface, raising the duct forces to
2
ρπR2prop: V 2
values closer to measurements. This effect is particularly evident at

Fig. 6. Axial and tangential force distributions. Reference stands for the propeller loading as computed by the Boundary Element Method for the given blade geometry. Loaded and
Unloaded stand for tip modified distributions used to assess the sensitivity of the design process to local changes of load. (a) Accelerating ducted propeller. (b) Decelerating duc-
ted propeller.

8
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 7. Comparison of measured and computed nozzle thrust. Accelerating duct. (a) Comparison of the nozzle over propeller thrust ratio (τ) as a function of the propeller thrust coefficient
CT prop. (b) Comparison of thrust coefficients (KT) as a function of the propeller advance coefficient J.

Fig. 8. Comparison of measured and computed nozzle thrust. Decelerating duct. For confidentiality reasons, values are given non-dimensional with respect to a reference operative
condition. (a) Comparison of the nozzle over propeller thrust ratio (τ) as a function of the propeller thrust coefficient CT prop. (b) Comparison of thrust coefficients (KT) as a function of the
propeller advance coefficient J.

higher propeller loadings. In lightly loaded conditions, differences be- along its camber line for a propeller thrust coefficient CT prop. of about 4.
tween computations and model tests are larger. The inclusion of On the contrary, in lightly loaded conditions, the risk of a separation
tangential velocities in the numerical simulation does not contribute bubble on the outer side is clear, completely in agreement with the design
significantly to improve the correlation with experiments by modifying of the nozzle, devoted to highly loaded functioning. For the decelerating
the pressures. Especially in the case of the accelerating duct geometry, nozzle shape, it is possible to appreciate the increase of the inner pressure
the thrust of the nozzle is significantly overestimated. This difference close to the design point (CT prop./CT prop. ref. ¼ 0.10, J ¼ 103.4% of Jref.),
may be explained by considering that the flow predicted by the actuator together with streamlines approaching the nozzle at its ideal angle
disk models is axisymmetric and neglects the tip leakage vortex; this can of attack.
alter the streamlines path in the circumferential direction and produce a It is worth noting that calculations have been carried out with a sort of
negative interaction with the duct blunt trailing edge (Bosschers “GAP” model. The flow passing through the gap between the propeller tip
et al., 2015). and the inner duct surface is, indeed, strongly influenced by viscosity.
Accelerating duct geometries, however, are primarily designed to Simplified representations have been proposed to include these effects in
operate at high propeller loading. The discrepancy at very high advances, Boundary Element Methods calculations. It is the case of Hughes (1993),
not contemplated in present design activity, can be consequently who proposed to “fill” the gap between the propeller tip and the inner
accepted, as well as the approximations introduced by the wedge type surface of the duct with an additional strip of chord-wise panels. The
boundary conditions setup. The comparative nature of a design by strength of dipoles and sources distributed over these panels was pre-
optimization, indeed, relies more on the ability to identify relative dif- scribed through a transpiration velocity in accordance with empirical
ferences among designs, rather than on the absolute accuracy of the coefficients to account for the influence of viscosity on the flow rate
calculations. The computational efficiency that can be exploited with the (reduced with respect to pure potential approximations) through the gap.
wedge type boundary conditions largely compensates for the slightly less Calculations with momentum sources by Hoekstra (2006) partially
accurate results. showed the presence of a recirculation zone in correspondence to the
Streamline patterns and inner pressure fields propeller plane for relatively highly load conditions, with non-reattached
(CP ¼ 2ðp  pref : Þ=ðρV 2 Þ) for the two nozzle geometries are shown in streamlines as the tip clearance increases or the loading at the tip is
Fig. 9. The position of the leading edge stagnation point is evident, as reduced. The extension of this zone was, however, overpredicted with
well as how streamlines change with propeller loading. Streamlines, as respect to the few experimental measurements available. Bosschers et al.
already evidenced in Hoekstra (2006), approach Nozzle 19A almost (2015) further analyzed this phenomenon and concluded that the

9
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 9. Streamlines and inner pressure distribution at different loading conditions. Accelerating (Nozzle 19A, left) and decelerating (right) ducts. For confidentiality reasons, the
reference decelerating nozzle shape cannot be disclosed. (a) CT prop. ¼ 12.0 (J ¼ 0.3) (b) CT prop. ¼ 4.0 (J ¼ 0.5) (c) CT prop. ¼ 0.35 (J ¼ 1.1) (d) CT prop./CT ref. ¼ 0.41 (J ¼ 60% of Jref.)
(e) CT prop./CT ref. ¼ 0.11 (J ¼ 100% of Jref.) (f) CT prop./CT ref. ¼ 0.02 (J ¼ 140% of Jref.).

presence of a recirculation zone is due to the loading of the propeller that required also in the case of viscous calculations with momentum sources.
generates an upstream velocity in the gap where no momentum sources Such model could be easily realized, perfectly in analogy with the one
are present. Momentum sources, even in the case of viscous calculations, proposed for BEM calculations, by placing momentum sources of reduced
cannot account for the influence of a “physical” surface as that of the intensity in the clearance between the propeller tip and the inner surface
propeller blades. Body forces are almost “transparent” for the flow that is of the nozzle. Exactly as the propeller forces modeled through the radi-
solely accelerated and not deviated by the “obstacle” represented by the ally varying actuator disk, also the “GAP” flow model is applied uni-
blades. The circumferential averaged distribution of forces realized by formly in the circumferential direction, averaging its influence all over
the actuator disk further prevents any asymmetry of the flow, avoiding the ring across the propeller tip, consistently with the modeling of the
the generation of the tip leakage vortex that could mitigate the risk of propeller with an actuator disk. In addition, this approach is the only
separation. Hence, as proposed by Bosschers et al. (2015), a gap model is possible with the wedge type computational domain. More complex

Fig. 10. Influence of the “GAP” model on separation on the inner surface of Nozzle 19A. CT prop. ¼ 4.0 (J ¼ 0.5). (a) Without the “GAP” model, CT Nozzle ¼ 1.07 (CT Nozzle Exp. ¼ 1.587). (b)
With the “GAP” model, CT Nozzle ¼ 1.58 (CT Nozzle Exp. ¼ 1.587).

10
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Table 4
Functioning conditions for the accelerating duct design.

Condition 1 Condition 2
(i.e. bollard pull) (higher advance)

CT prop. 12 1.23
J (equivalent propeller 0.3 0.8
advance coefficient)

combinations. The optimization workflow lets the initial population


evolve for 15 generations; this, in the light of preliminary calculations
and the relatively low number of free parameters, was considered suffi-
cient to achieve convergence. Results of the analyses are summarized in
Sections 5.1 and 5.2.

5.1. Accelerating ducts


Fig. 11. Pareto diagram of the accelerating duct design by optimization. Correlation be-
tween nozzle thrust and average inner pressure at the higher advance coefficient The outcomes of the design by optimization of the accelerating duct
(J ¼ 0.80, CT prop ¼ 1.23).
are summarized in the Pareto diagrams of Figs. 11 and 12. Among the
design activities taken into consideration, the one involving the accel-
propeller representations, like the local distributions of momentum erating duct is the more complex. It considers, indeed, multiple working
sources resembling each single blade adopted in Villa et al. (2011), conditions, as outlined in Table 4. The nozzle is designed to provide the
would have allowed for the use of more sophisticated “GAP” approxi- maximum thrust close to bollard pull condition that in numerical cal-
mations, which could be based, for instance, on the exact location of culations was approximated with a relatively low propeller advance co-
blade tips. As shown by Bosschers et al. (2015), however, even the in- efficient to avoid the complications related to zero inflow. At the same
clusion of such a simplified “GAP” flow model dramatically helps in time, minimization of the nozzle drag at higher advance coefficients
avoiding recirculation. The “GAP” model provides a flow in the trailing (J ¼ 0.8 in the present case) was required, together with maximization of
part of the duct that is closer to the average flow obtained from fully inner static pressure at both functioning. The static pressure was moni-
resolved RANSE calculations. The use of the Boundary Element Method tored through the average pressure on a propeller disc in front of the
(that includes the simplified model by Hughes) to define the radial dis- actuator disc at x/D ¼ 0.15 from the propeller plane, as in Fig. 5. The
tribution of forces, finally, provides naturally the strength of these propeller loading is that of Fig. 6a. Compared to the parametric load
additional momentum sources. distributions of Hoekstra (2006) that resembled tip loaded Kaplan like
An example of application is shown in Fig. 10. Without the “GAP” propeller force distributions due to the finite chord at the tip, the present
model a serious recirculation zone, resulting in a significant change distribution of force is particularly unloaded. This particular feature, that
(reduction) of the thrust delivered by the nozzle, can be evidenced in was preferred for the nature of the application, further reinforces the
correspondence to the propeller disc. Streamlines instead are aligned need for the “GAP” model (even though simplified) to limit the risk of
with the duct inner surface and force is better in agreement with mea- flow separation on the inner of the nozzle evidenced by Bosschers
surements when momentum sources fill the clearance between the pro- et al. (2015).
peller and the nozzle. Results of the optimization process show certain margins of im-
provements, which can be verified by considering some relevant geom-
5. Optimization of nozzle shapes etries among those belonging to the Pareto Frontier. As outlined in the
discussion, without redesigning the propeller, the recommended optimal
The design of nozzles through optimization was achieved starting nozzles are those operating at constant thrust coefficient in order not to
from two reference geometries available at the cavitation tunnel of the change the design point of the propulsion system. The almost constant
University of Genoa. In both cases, the design consists in the analysis of nozzle thrust allows also minimizing the influence of the nozzle on the
an initial population of 150 members. This population was selected by propeller inflow and, in turn, on its performance. Depending on the
using the quasi-random Sobol sequencing, resulting in a sufficiently importance of bollard pull with respect to routinely operative conditions
uniform distribution in the design space of the possible geometrical at higher advance coefficient, there are multiple possible choices. In the

Fig. 12. Pareto diagrams of the accelerating duct design by optimization. Correlation between nozzle thrusts (left) and average inner pressures (right) computed in bollard pull and at the
higher advance coefficient conditions. (a) Correlation of thrusts. (b) Correlation of pressures.

11
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Table 5
Geometrical characteristics and performance of the selected designs for the accelerating duct case (reference points and camber line definition slightly different from those proposed for the
reference Nozzle 19A by Van Manen and Oosterveld, 1966).

Nozzle 19A ID 2172 ID 1987 ID 1736 ID 1835 ID 1056

L/D 0.5000 0.4075 0.3875 0.4825 0.4625 0.6525


Rinlet/D 0.5913 0.5860 0.5885 0.5905 0.5765 0.5850
Routlet/D 0.5218 0.5238 0.5245 0.5248 0.5235 0.5298
cambermax/L 0.05421 0.05007 0.06600 0.06872 0.05164 0.04264
thicknessmax/L 0.16452 0.21127 0.20288 0.14951 0.15582 0.11475
x max camber/L 0.2910 0.3437 0.3527 0.3617 0.4576 0.3288
x max thickness/L 0.2768 0.3171 0.3288 0.2437 0.2656 0.2910
Incidence angle [ ] 7.900 8.661 9.384 7.785 6.569 4.862

CT Nozzle (J ¼ 0.3, CT prop. ¼ 12) 6.731 6.725 5.829 7.370 7.080 7.510
CT Nozzle (J ¼ 0.8, CT prop. ¼ 1.23) 0.238 0.236 0.237 0.237 0.267 0.282
CP (J ¼ 0.3, x/D ¼ 0.15) 9.363 8.368 7.268 9.866 9.239 11.105
CP (J ¼ 0.8, x/D ¼ 0.15) 1.019 0.861 0.815 1.096 1.010 1.278

present design, where the reference propeller/nozzle system was necessary for these achievements are summarized in Table 5 while in
designed for take home purposes (i.e. provide a reasonably high thrust Figs. 13 and 14 the computed pressure and velocity fields are shown for
with a small propeller at a relatively high advance coefficient) it was the selected geometries at both the advance coefficients (i.e. propel-
considered mandatory, at first, to improve the performance at the higher ler loadings).
advance coefficient. In order to limit cavitation phenomena, moreover, The presentation of the results in the Pareto diagrams shows some
the increase of the inner nozzle pressure in this condition is another correlations among objectives. Obviously, there is a relation between
important objective of the design that turns into the second most nozzle thrusts at different advance coefficients (propeller loadings): a
important aspect for the selection of an optimal geometry. Among the geometry that maximizes bollard pull performance improbably will
Pareto geometries, ID 2172 is the optimal balance between the opposite provide poor performance in terms of thrust at different operative con-
design requirements. Providing a constant thrust at the higher advance ditions. At the higher advance coefficient, results are a bit more scattered
coefficient, ID 2172 increases, at that condition, the inner static pressure and there is a certain degree of freedom, exploited for the selections of
at its maximum without worsening the performance at bollard pull the candidates for the detailed analyses of Figs. 13 and 14. The correla-
(constant thrust) in correspondence to which a certain improvement in tion between inner pressures is even stronger and almost linear. By
terms inner pressure can be appreciated too. Other geometries presup- analyzing the performance of all the geometries tested during the opti-
pose trade-offs among objectives. At constant nozzle delivered thrust for mization process, some important correlations among performance and
J ¼ 0.80, ID 1987 provides the maximum increase of static pressure geometrical parameters can be furthermore discussed. Data are collected
(appreciable also at bollard pull) at the cost of a sensible reduction of in the charts of Figs. 15 and 16 where the Student t-test was used to es-
bollard pull thrust (Fig. 12). On the opposite, ID 1736 maximizes bollard timate the influence (and the significance, i.e. the trustworthiness of the
pull performance at constant thrust for J ¼ 0.80 with a detrimental in- devised relationship) of the input variables (the design parameters) on
fluence on cavitation avoidance. By accepting the redesign of the pro- the outputs (Esteco, 2016). By combining variances, mean values and
peller, ID 1835 exploits the maximum thrust from the propeller without number of variables in the lower and in the upper part of the design
any significant influence on pressure fields while ID 1056 is devoted only space, essentially the t-test provides a measure of the strength of the
to the maximization of nozzle delivered thrust regardless the inner relationship between the outputs and the inputs, that is summarized in
pressure field. Detailed descriptions of the geometrical modifications the pie chart of Fig. 15. Another output of the analysis is the kind of

Fig. 13. Streamlines and inner pressure distribution for the selected optimal geometries. Bollard pull condition (J ¼ 0.3, CT prop. ¼ 12). (a) Reference Nozzle 19A. (b) ID 1056. (c) ID 1835.
(d) ID 1736. (e) ID 1987. (f) ID 2172.

12
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 14. Streamlines and inner pressure distribution for the selected optimal geometries. Higher advance coefficient condition (J ¼ 0.8, CT prop. ¼ 1.23). (a) Reference Nozzle 19A. (b) ID
1056. (c) ID 1835. (d) ID 1736. (e) ID 1987. (f) ID 2172.

Fig. 15. Overall Student t-test for design objectives of the accelerating duct.

relationship (direct or inverse) among inputs and outputs, highlighted by indeed, is probably not large enough. In particular, it is not equally
the bar chart of Fig. 16. Even if the number of elements over which the distributed over the design space to account evenly for the dependencies
statistical analysis is carried out is rather limited, some observations circa and the relationships existing between free parameters and objectives.
the observed influences can be drawn, also in the light of the very low The use of a sampling of the design space based on results of an opti-
values of significance obtained by the t-test (the lower the significance, mization activity, in addition, directs the candidates for the t-test towards
the higher the trustworthiness of the effects, direct or inverse, between the geometries and the choice of the parameters that favor the objectives
inputs and outputs). of the design. The candidates in the upper and in the lower part of the
The length of the nozzle, as well as the maximum nozzle thickness, design space, in turn, could be not equally divided, affecting the accuracy
have a significant importance on design objectives. Higher values of of the test, whose reliability has to be considered in relation to this choice
thrust, especially close to bollard pull, are achieved with longer nozzles. of design space and range of variations of the parameters. In this sense,
The direct effect of nozzle length in correspondence to lightly loaded some contrasting results of the analysis could be explained. For instance,
propeller functioning is lower, exactly as the trends evidenced by Van the opposite influence of the angle of attack with respect to the inlet
Manen and Oosterveld (1966). Increased thrust is directly influenced by radius (that in some way is dependent on a combination of the angle of
the decrease in thickness and of incidence angle. The exit radius (defined attack itself and the nozzle length) on the average pressure, as well as the
in Fig. 5) has a non-negligible influence on delivered thrust while the role role of the maximum thickness, could be ascribed to an insufficient and
of the camber (value and position) is marginal, as already evidenced by not perfectly appropriate choice of the elements for the t-test.
the analyses carried out in the case of conventional Nozzles 18, 19 and In the absence of dedicated experimental analyses, an indirect vali-
20. When looking at the maximization of the inner pressure, these vari- dation of the outcomes of the design by optimization has been proposed
ables have exactly an opposite influence. The improvement of cavitating by using 3D RANSE calculations of the selected geometries. OpenFOAM
performance favors shorter but thicker nozzles, higher angles of attack calculations were carried out based on the numerical setup for propeller
and lower exit radii, encouraging configurations and trends that some- performance evaluation devised in Gaggero and Villa (2017), using a
how tend towards those typical of the decelerating nozzles of Section 5.2. polyhedral mesh between 1.4 and 1.6 Million cells per blade depending
This demonstrates the complexity of the design and the need of a on the geometry under investigation. The reliability of the proposed
trade-off between contrasting objectives also in the light of constraints calculations has been preliminary verified in the case of the reference
(regarding geometry or structural strength) that, presently, have not been Nozzle 19A arrangement by comparing, in Table 6, the computed results
considered. This analysis and the similar one proposed in next Section for with the available experimental measurements. The overall agreement at
the decelerating nozzle case has not to be considered, however, fully both the advance coefficients under investigation is good. The propeller
exhaustive. The number of configurations accounted for the t-test, thrust is slightly underestimated and torque is reasonably predicted. The

13
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 16. Effect size of each geometrical parameter on the design objectives. Accelerating duct. (a) Effect on CP (J ¼ 0.3, CT prop. ¼ 12). (b) Effect on CP (J ¼ 0.8, CT prop. ¼ 1.23). (c) Effect on
CT Nozzle (J ¼ 0.3, CT prop. ¼ 12). (d) Effect on CT Nozzle (J ¼ 0.8, CT prop. ¼ 1.23).

Table 6
Comparison between measured and predicted performance. Accelerating ducted propeller inside Nozzle 19A.

Exp. KT Prop. Δ% Exp. 10KQ Δ% Exp. KT Nozzle Δ%


RANSE RANSE RANSE

J ¼ 0.3 0.4269 0.4168 2.36% 0.9867 0.9862 0.05% 0.2561 0.2629 þ2.64%
J ¼ 0.8 0.3091 0.3046 1.46% 0.7712 0.7826 þ1.48% 0.0414 0.0557 þ34.44%

thrust delivered by the nozzle is in good agreement with measurements three-dimensional interaction of the flow may justify these differences. It
close to the bollard pull condition (J ¼ 0.30). At higher J, the differences is remarkable, however, that the highlighted differences do not invali-
are amplified by the low reference value with respect to which small date the results of the design. With the exception of ID 1736 and ID 1835
absolute discrepancies turn into high percentage errors; these, however, (which do not have the same ranking at J ¼ 0.30), the trend and the
appear practically negligible when looking at the total (propeller plus ranking evidenced at the lower advance coefficient by the simplified
nozzle) thrust. Close to bollard pull, the difference is less than 1% and optimization process are confirmed. The same ID 1736, ID 1835 plus ID
calculations show an underpredicted total thrust of about 0.5%. At 1056 nozzle geometries are those that provide the highest increase of the
J ¼ 0.80 the overestimation is about 2.8%, perfectly in line with the delivered thrust at bollard pull which, however, turns into a worsening
assumption of calculations. (reduction) of the propeller performance (thrust), with obvious conse-
The five geometries identified by means of the optimization were quences discussed in the following on the inner pressure distributions. At
analyzed (Fig. 17) with the same numerical setup of the reference Nozzle J ¼ 0.80, nozzles ID 2172, ID 1736 and ID 1987 deliver almost the same
19A configuration. The percentage differences with respect to the nu- thrust of the reference configuration, as expected by the optimization
merical results of the reference geometry, both in terms of delivered (Fig. 11). Only ID 1056 does not confirm the results achieved with the
thrust and average pressure coefficient in front of the propeller are actuator disk. This geometry should have been the one delivering the
collected in Table 7. The complexity of the three-dimensional flow higher thrust at both the advance coefficients with the poorer perfor-
around a ducted propeller, already evidenced in the case of the issues mance in terms of average pressure distribution to the propeller. While
related to the gap and the risk of flow separation, is sensibly simplified the substantial worsening of the performance in terms of inner pressure
with the adoption of actuator disk models. Hence the comparison is distribution is confirmed, this geometry is not the one providing the
significant only for what regards trends rather than in terms of absolute highest increase of nozzle thrust at J ¼ 0.80 with respect to the reference
value of improvements of one geometry with respect to the others. In this Nozzle 19A. In this respect, ID 1835 performs better. Also the increase of
sense, the results collected with fully resolved RANSE calculations are inner pressure, foreseen by the optimization, is generally confirmed by
satisfactory. Close to bollard pull condition, fully resolved calculations the fully resolved RANSE calculations. The improvements of ID 1987 and
underestimate the increase (decrease) of nozzle thrust with respect to the ID 2172 at both advances are clear and in line with the results of the
calculations with the simplified actuator disk. A different propeller optimization. ID 1736, instead, has an opposite trend with respect to the
loading at bollard pull (not accounted during optimization) and the optimization process. With the actuator disk, this geometry should have

14
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 17. Pressure coefficient distribution on propeller and duct at J ¼ 0.8. Fully resolved RANSE calculations. Accelerating duct design. Pressure coefficient based on propeller rate or
revolution (CpN ¼ 2ðp  pref : Þ=ðρN 2 D2 Þ). (a) Reference Nozzle 19A. (b) ID 2172. (c) ID 1987. (d) ID 1736. (e) ID 1835. (f) ID 1056.

inflow as a consequence of sensibly modified nozzle shaped) predicted by


Table 7 the fully resolved RANSE calculations and not accounted by the fixed
Optimized accelerating nozzle performance. Differences with respect to the Nozzle 19A
configuration.
body forces distribution adopted throughout the optimization process.
For those two configurations (but also for ID 1056, as discussed previ-
ID 2172 ID 1987 ID 1736 ID 1835 ID 1056
ously) the predicted propeller thrust is lower than that of the propeller
ΔT Nozzle, 0.08% 13.39% 9.51% 5.19% 11.58% with the Nozzle 19A (about 4% for both the geometries at J ¼ 0.8).
Optimization,
Lower propeller loading turns immediately into a lower decrease of
J ¼ 0.30
ΔT Nozzle, RANSE, 0.92% 2.71% 2.07% 2.32% 2.50%
pressure on the suction side and the consequences on the average pres-
J ¼ 0.30 sure in front of the propeller are obvious. On the other hand, geometries
(ID 2172 and 1987) selected to provide the same nozzle thrust at J ¼ 0.80
ΔT Nozzle, 0.94% 0.40% 0.58% 12.20% 18.35%
Optimization, without improvements at bollard pull (both the analyses shows a
J ¼ 0.80 reduction of the delivered thrust in this condition) are those that avoid
ΔT Nozzle, RANSE, 0.52% 1.01% 2.54% 16.12% 5.91% any significant modification of the propeller functioning. At both oper-
J ¼ 0.80
ating points, fully resolved RANSE calculations predict propeller per-
ΔCP, Optimization, 10.62% 22.37% 5.37% 1.33% 18.61% formance with a maximum of 1% variation with respect to the reference
J ¼ 0.30 configuration with Nozzle 19A, proving the assumption of the design at
ΔCP, RANSE, 17.72% 24.07% 2.56% 8.01% 11.29%
constant nozzle thrust.
J ¼ 0.30

ΔCP, Optimization, 15.44% 20.04% 7.55% 0.84% 25.41%


J ¼ 0.80 5.2. Decelerating ducts
ΔCP, RANSE, 27.78% 35.65% 2.84% 10.60% 23.15%
J ¼ 0.80
The optimization of the decelerating duct has been carried out by
considering the propeller loadings of Fig. 6b. As partially discussed in the
provided a reduction of the inner pressure (7.55% with respect to the introduction, the loading distribution has a certain influence on the
average pressure inside Nozzle 19A at J ¼ 0.80) while resolved RANSE nozzle performance that could be exploited during the design. The
calculations suggest an increase (þ2.84%) at the same advance coeffi- reference propeller was designed, as in the case of the accelerating duct,
cient. Similarly, ID 1835 provides increases substantially higher than to postpone as much as possible cavitation inception (Gaggero et al.,
those evidenced in the optimization process. These differences may be 2012), which of course affects the tip at first. Through optimization,
ascribed to the different functioning of the propeller (due to a modified therefore, also the influence of propeller tip loading on nozzle shape was
investigated as a first step in the direction of a combined propeller/nozzle
Table 8 design. Any redesign of the propeller, in fact, turns into a different load
Functioning conditions for the decelerating duct design and influence of tip loading on the distribution and it is important to evaluate which could be the influence
computed performance of the reference nozzle shape. of such modifications on the nozzle performance. An unloaded tip blade
Load BEM Tip Loaded Tip Unloaded obviously improves the inception speed of the propeller. The possible
CT prop./CT prop. ref. 0.10 0.10 0.10 reduction of the nozzle thrust or of the inner static pressure, however,
J (equivalent propeller 103.4% of J ref. 103.4% of J ref. 103.4% of J ref. could nullify the positive influence of tip unloading. On the contrary, the
advance coefficient) negative effects related to tip loading (i.e. anticipated cavitation incep-
CT Nozzle/CT prop. ref. 0.0182 0.0186 0.0170 tion) could be mitigated by an increase of the inner pressure or by the
CP at x/D ¼ 0.15 0.035 0.041 0.018
possibility to operate the propeller at a most favorable point thanks, for

15
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

reduction of the pressure. This behavior, within the simplification of


circumferentially averaged body forces, can be explained by the change
of the flow around the duct, induced by the different loading distribution.
With the tip loaded, the angle of attack of the duct increases, and the
stagnation point moves towards the inner surface of the nozzle. This local
increase of pressure overbalances its reduction due to the higher load (i.e.
higher suction) at the tip. The shape of the reference nozzle, with the
usual converging geometry at the trailing edge, inherently reduces the
risk of flow separation observed in the case of accelerating duct in lightly
loading conditions. Also the need of the “GAP” model should be dis-
cussed in the case of decelerating duct geometries in the light of the
different behavior of the flow at the inner nozzle surface. The satisfactory
results collected during the validation analyses, which were computed in
any case (accelerating and decelerating) with momentum sources also in
the gap region, authorize the application of the same model also for the
optimization activities.
Optimizations have been performed at the propeller design advance
Fig. 18. Pareto diagram of the decelerating duct design by optimization. Reference pro-
peller loading distribution.
coefficient, in correspondence to the same propeller loading condition
used for the preliminary assessment of the influence of tip. The design
was carried out by requiring the maximization of the inner pressure (also
in this case the average pressure extracted on a plane 0.15D in front of the
propeller disc) and the minimization of the nozzle drag. The results of the
calculations with different loading distributions have been collected
through Pareto diagrams among which the one for the reference distri-
bution is shown in Fig. 18. Trends are very similar regardless the loading
distribution and the Pareto frontiers show an almost linear correlation
between thrust and inner pressure.
The comparison of the Pareto frontiers (shown in Fig. 19) for the three
cases under investigation is more significant. At constant nozzle thrust
(drag, in this case), that is the condition that minimizes the need of a
propeller redesign (needed only if induced velocities to the propeller are
significantly changed), the very low overpressure with respect to undis-
turbed flow in front of the disk of the reference geometry is more than
doubled. At constant inner pressure, the drag of the nozzle can be
reduced by about 20%, which may represent a certain margin for the
Fig. 19. Comparison of the Pareto frontier for different propeller loading distributions.
redesign of a propeller in a more favorable, for cavitation inception,
unloaded working condition.
Results of the optimization activities are somehow in contrast with
instance, to a reduction of the nozzle drag. The tip loaded distribution has
the preliminary verification of the influence of tip loading. From the
been realized by simply increasing the value of the delivered propeller
analysis of the reference geometry in Table 8, the unloading at tip proved
forces only on the outer actuator disk sections. The unloaded tip distri-
to be effective in reducing the nozzle drag with a slight worsening (i.e.
bution, instead, couples the decreasing of the forces at outer radii with
reduction) of the inner pressure. Nozzle designs obtained through opti-
the shift of the position of the maximum of the distribution towards the
mization in the case of the unloaded tip show, instead, a simultaneous
center of the disk.
improvement of both the objectives (reduction of drag and increase of
A preliminary analysis of the influence of the propeller loading dis-
inner pressure) with respect to designs obtained with the reference and
tributions (at constant total propeller loading) on the reference nozzle
the tip loaded propeller load distributions. Designs obtained with the
performance is reported in Table 8. Calculations have been carried out
unloaded tip are overall better both at any nozzle thrust (providing the
for the propeller design condition and are summarized in the same table:
highest inner pressure) and at any inner pressure level (providing the
loading the tip slightly increases the average static pressure in front of the
lower drag). The selected optimal shapes have the main geometrical
actuator disk at the cost of a negligibly higher drag. Unloading, on the
characteristics collected in Table 9 while non-dimensional pressure fields
contrary, reduces the drag of the nozzle at the cost of a more significant

Table 9
Geometrical characteristics and performance of the selected designs for the decelerating duct case (CT prop./CT prop. ref. ¼ 0.10, J ¼ 103.4% of Jref.).

Reference Load Loaded Tip Unloaded Tip

ID 2201 ID 2192 ID 1930 ID 1914 ID 2231 ID 1948

L/D 0.4400 0.3850 0.4250 0.4375 0.3950 0.3850


Rinlet/D 0.5370 0.5493 0.5500 0.5500 0.5383 0.5455
Routlet/D 0.4738 0.4800 0.4758 0.4798 0.4743 0.4785
cambermax/L 0.03930 0.03775 0.02997 0.02594 0.04315 0.03957
thicknessmax/L 0.12513 0.17058 0.15210 0.14653 0.13693 0.16360
x max camber/L 0.5830 0.6292 0.6231 0.6322 0.6047 0.5892
x max thickness/L 0.2437 0.2740 0.2198 0.2276 0.2518 0.2601
Incidence angle [ ] 8.180 10.197 9.910 9.122 9.203 9.872

CT Nozzle/CT prop. ref. 0.0182 0.0150 0.0182 0.0154 0.0180 0.0148


CP 0.086 0.034 0.091 0.041 0.093 0.037

16
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 20. Streamlines and inner pressure distribution for the selected optimal geometries with the reference propeller loading distribution (J ¼ 103.4% of Jref., CT prop./CT prop. ref. ¼ 0.10).
(a) Reference Geometry. (b) ID 2201. (c) ID 2192.

Fig. 21. Streamlines and inner pressure distribution for the selected optimal geometries with the propeller tip loaded distribution (J ¼ 103.4% of Jref., CT prop./CT prop. ref. ¼ 0.10). (a)
Reference Geometry. (b) ID 1930. (c) ID 1914.

Fig. 22. Streamlines and inner pressure distribution for the selected optimal geometries with the propeller tip unloaded distribution (J ¼ 103.4% of Jref., CT prop./CT prop. ref. ¼ 0.10). (a)
Reference Geometry. (b) ID 2231. (c) ID 1948.

and streamlines on the meridian plane are presented in Figs. 20–22. For of accelerating duct geometries, the objectives are influenced by the
geometries improving the increase of pressure at the propeller plane, it is parameters in opposite ways. The exit radius, the position of the
possible to observe stagnation points slightly moved to the inner part of maximum thickness, the maximum camber (and its position) and the
the nozzle. The suction peak on the outer duct surface is not sensibly incidence angle turn into the most influencing geometrical characteris-
affected by the geometrical modifications and the risk of sheet cavitation tics. The nozzle length, as for the discussion circa the selected optimal
on the nozzle, even though not directly monitored throughout the opti- shapes, marginally influences performance improvement.
mization process, is not appreciably enlarged for the selected geometries. Exactly as done in the case of the accelerating nozzle case, a valida-
The substantial identity of performance turns into very close shapes tion of the results of the design by optimization was carried out by means
regardless the distribution of forces of the actuator disk: geometries with of fully resolved RANSE analyses. Only the reference propeller loading
higher inner pressure are generally thinner and with a smaller exit sec- distribution (i.e. geometries ID 2192 and ID 2201) was analyzed, being
tion. Also the length is decreased to avoid high losses due to friction but, the tip loaded and unloaded distributions only case studies without a real
with respect to the analyses of the parameters influence in the case of propeller geometry available. In the case of the reference geometry
accelerating nozzles, its importance is scaled down. All the optimal ge- (Table 10), numerical calculations with OpenFOAM reasonably predict
ometries have a nozzle length substantially lower than the reference the propeller performance, with results in line with those already pro-
geometry, regardless the objective that is addressed. posed by Gaggero et al. (2012, 2013) using a different RANSE solver
A thorough estimation of the relations existing between geometrical (StarCCMþ). The overprediction of the nozzle drag, when looking at the
modifications and objectives is outlined in Figs. 23 and 24. Also for the percentage differences, is again amplified by the very low reference
decelerating nozzle design, the Student t-test, with the same limitations value. Its influence on the total propulsive thrust (numerically under-
highlighted for the accelerating duct case, has been used to verify the estimated by less than 0.7%) is, however, negligible.
relationships and their nature, direct of inverse, between modifications of The results of the optimization based on the simplified model of the
main parameters outlined in Fig. 5 and effects on objectives. Due to the actuator disk are, also in the case of the decelerating nozzle configura-
very similar design outcomes with different choices of the propeller tion, confirmed by the fully resolved RANSE analyses (Fig. 25) with few
loading distribution, only the reference case is considered. As in the case discrepancies worth of discussion (Table 11). ID 2201 provides a nozzle

17
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Fig. 23. Overall Student t-test for design objectives of the decelerating duct. Reference propeller loading distribution.

Fig. 24. Effect size of each geometrical parameter on the design objectives. Reference propeller loading distribution. (a) Effect on CP. (b) Effect on CT Nozzle.

actuator disk) on the flow field, as discussed in Section 3.2. A propeller


Table 10
Comparison between measured and predicted performance. Decelerating ducted propeller
produces an asymmetrical pressure jump: the suction side is commonly
inside the reference geometry. characterized by a reduction of the pressure that is two or three times
higher (in magnitude) with respect to the pressure increase on the
ΔKT Prop. Δ10KQ ΔKT Nozzle
pressure side of the blade (Gaggero et al., 2017). On the contrary, a
J ¼ 103.4% of Jref. 3.19% 6.14% 24.34%
momentum source always generates a symmetrical pressure jump equally
distributed between decrease of pressure in its front and increase in its
thrust slightly higher than the reference geometry but reasonably in line aft. This means that with the actuator disk the reduction of pressure in
with the outcomes of the optimization, which identified this geometry as front of the propeller is underestimated, resulting in the higher (with
the one providing the maximum increase of inner pressure at constant respect to fully resolved propellers with RANSE) inner pressure seen
nozzle thrust (Fig. 19). The increase of the inner pressure, indeed, is throughout the optimization process.
substantial (about þ30%) but far from the predictions with the actuator The relative (with respect to the reference geometry performance)
disk which foreseen a more than doubled inner pressure. Reasons for this increase of nozzle thrust of ID 2192 is similarly predicted by both the
discrepancy can be found in the influence of momentum sources (i.e. the approaches. Fully resolved RANSE calculations, however, identify a

Fig. 25. Pressure coefficient distribution on propeller and duct. Fully resolved RANSE calculations. Decelerating duct design. Pressure coefficient based on propeller rate or revolution
(CpN ¼ 2ðp  pref : Þ=ðρN 2 D2 Þ). (a) Reference Geometry. (b) ID 2201. (c) ID 2192.

18
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

Table 11 by maximizing the delivered thrust of the nozzle and, in the end, by
Optimized decelerating nozzle performance. Differences with respect to the reference redesigning the propeller in the more favorable (unloaded) functioning
geometry.
condition allowed by the increase of the nozzle thrust.
ID 2201 ID 2192 A first step towards the combined design of the whole propulsive
ΔT Nozzle, Optimization, J ¼ 103.4% of Jref. 0.17% þ17.57% system was proposed in the case of the decelerating nozzle. The adoption
ΔT Nozzle, RANSE, J ¼ 103.4% of Jref. þ3.90% þ16.59% of the actuator disk modeling avoids any possible interactions from the
ΔCP, Optimization, J ¼ 103.4% of Jref. 146% 3% propeller point of view, which instead, would be necessary to account for
ΔCP, RANSE, J ¼ 103.4% of Jref. 27.55% 8.29%
the modification of the inflow to the propeller due, at first, to the mod-
ifications of the nozzle shape and vice versa. For the decelerating duct,
certain increase of the inner pressure, not evidenced in the case of the consequently, three simultaneous designs were carried out with different
actuator disk (ID 2192 was selected as the geometry providing the propeller loading distributions at constant delivered thrust. A change in
maximum increase of nozzle thrust at almost constant pressure). This the inflow turns immediately, indeed, into a modification of sectional
discrepancy, again, can be explained by the different working condition angles of attack. A preliminary assessment of the influence of a different
of the propeller. Nozzle ID 2192 is the shortest and has the highest propeller on duct performance (or the investigation on how the duct has
incidence angle. The inflow to the propeller is such that the propeller to be modified to take advantage of a particular loading distribution) was
delivered thrust results lower with respect to reference geometry by possible by simply modifying the momentum source distribution. Also in
about 4.2%. Exactly as discussed in the case of the accelerating nozzle this case, the outcomes of the design by optimization were encouraging.
geometries, unloading the propeller reduces the suction on the back side It was demonstrated that a dedicated nozzle design, as in the previous
of the blade with an obvious influence on the average inner pressure in case, supports simultaneously the increase of the inner pressure and the
front of the propeller. reduction of drag. Moreover, the opportunity to exploit even small
These results, together with the analyses carried out for the acceler- modification of the radial load distribution of the propeller to further
ating nozzle configurations, mainly confirm the robustness of the opti- increase the propulsive performance was verified, showing the benefits
mization approach. In parallel, however, they suggest once more the of a custom design based on current requisites. In addition, in both the
need of a feedback to the propeller loading/actuator disk momentum accelerating and the decelerating case, useful correlations between usual
distribution inside the optimization loop. In addition, a selection of an design parameters and nozzle performance were identified to support
optimal candidate based on a posteriori fully resolved RANSE analyses on design choices.
a reasonable set of geometries selected through the (simplified) optimi- Dedicated fully resolved RANSE calculations mainly confirmed the
zation procedure, as already proposed in the case of conventional highly outcomes of the optimization for both the configurations. Trends, rather
loaded propellers of Gaggero et al. (2017b) could furthermore refine than absolute values, and, in general, the ranking of the analyzed optimal
the choice. geometries, were verified, demonstrating the feasibility of the proposed
design by optimization for both accelerating and decelerating nozzles.
6. Conclusions These analyses, however, confirmed some of the limitations discussed
throughout the paper, of the adopted simplified models and the necessity
In this paper, an optimization approach for the design of the accel- of dedicated investigations on few critical points was evidenced. The role
erating and decelerating nozzles for ducted propeller applications has of the tip loading in case of accelerating duct geometries, especially for
been proposed. The design workflow consists of a parametric description what regards the flow features on the nozzle inner surface, would require
of nozzle shape, in a fully scriptable mesh generation tool based on the dedicated analyses even if the use of highly tip unloaded propellers
OpenFOAM blockMesh library, in the efficient RANSE solver simpleFOAM would nullify the scope of these propulsors, devised for heavy loading
and in an automatic post-processing of the data together combined functioning. Within the simplifications of the actuator disk model, opti-
through the modeFrontier optimization environment. A simplified rep- mization would provide also in this case useful design guidelines and the
resentation of the propeller action on the flow field (an actuator disk with opportunity to customize the nozzle shape. RANSE calculations, more-
radially varying momentum sources) was used to achieve the necessary over, clearly shown a certain degradation of the propeller performance
computational efficiency required by a design through optimization with specific nozzle shapes (those delivering the highest thrust) in both
involving RANSE calculations. A preliminary validation confirmed the accelerating and decelerating configurations, which obviously could not
reliability of the computational model with all its simplifications (wedge be evidenced by a fixed momentum source distribution.
type calculations, GAP model). A redesign of the propeller geometry in the actual inflow for any
Traditionally, accelerating nozzles are designed mainly for highly analyzed nozzle geometry (turning into an updated body forces distri-
(tip) loaded propellers while for decelerating type shapes the availability bution) or a modification of the body forces radial distribution based on
of design guidelines is even more scarce. The use of tip unloaded pro- the performance of the given propeller subjected to a modified inflows
pellers, motivated by increasingly important side effects, furthermore appear to be the first unavoidable steps for even more reliable and per-
stresses the design requirements of nozzles which may suffer from flow forming designs. The propeller could take advantage of a more favorable
separation and worsening of performance. The proposed design by inflow, of a higher static pressure and of a global unloading of the blade:
optimization meets these requirements: to provide a reliable tool in order these achievements are due to the different design point allowed by the
to customize the nozzle shape based on different objectives and con- increased thrust (or reduced drag) provided by the optimal nozzle. A
straints, allowing for the full exploitation of the potentialities of ducted feedback mechanism in the optimization loop, indeed, would be defi-
propellers also in the case of non-usual propeller functioning conditions nitely necessary in order to account for the mutual influence between the
and loadings. propeller and the newly designed ducts to carry out a combined design
In the specific case of the accelerating duct, it was possible to identify for constant delivered thrust. The simpler procedure could consist in a
a set of trade-offs configurations for the multiple operative conditions computationally efficient lifting line design tool embedded into an inner
simultaneously addressed in the optimization. Improving the cavity calculation loop. For any nozzle shape, a certain number of iterations
inception speed appeared as a possible objective, without any detri- allows defining the optimal propeller loading distribution dependent by
mental effects on thrust at both bollard pull and at the higher advance the nozzle inflow to the propeller, in turn, influenced by the performance
coefficient selected to resemble the routinely functioning of the propel- of the propeller itself. Dedicated fully resolved RANSE calculations,
ler. Similarly, the optimization procedure demonstrated the possibility to extended to a wider set of candidates could be employed, finally, as a post
improve the propulsive efficiency at unvaried cavitation inception speed processing of the optimization results for the definitive choice of the
optimal configuration. More wisely, RANSE calculations could be

19
S. Gaggero et al. Ocean Engineering xxx (2017) 1–20

adopted in order to continuously correct results throughout the design Hughes, M.J., 1993. Analysis of Multi-component Ducted Propulsors in Unsteady Flow.
PhD. Thesis. Massachusetts Institute of Technology, USA.
process, increasing the reliability of the proposed procedure, leading to
Hughes, M.J., 1997. Implementation of a special procedure for modeling the tip clearance
the simultaneous design of the nozzle and of the propeller by employing flow in a panel method for ducted propellers. In: Proceedings of the Propellers/
the well-established procedure for propeller optimization including the Shafting ’97 Symposium, Virginia Beach, Va, USA.
systematic variation of the duct driven by high-fidelity viscous Kerwin, J.E., Kinnas, S.A., Lee, J.-T., Shih, W.-Z., 1987. A surface panel method for the
hydrodynamic analysis of ducted propellers. Trans. Soc. Nav. Archit. Mar. Eng. 95.
calculations. Kinnas, S.A., Chang, S.-H., Yu, Y.-H., He, L., 2009. A hybrid viscous/potential flow
method for the prediction of the performance of podded and ducted propellers. In:
References Proceedings of Propeller/Shafting ’09 Symposium. Virginia Beach, VA, USA.
Kinnas, S.A., Yu, X., Tian, Y., 2012a. Prediction of propeller performance under high
loading conditions with viscous/inviscid interaction and a full wake alignment
Abdel-Maksoud, M., Heinke, H.J., 2002, July. Scale effects on ducted propellers. In: model. In: Proceedings of 29th Symposium on Naval Hydrodynamics, Sweden.
Proceedings of the 24th Symposium on Naval Hydrodynamics, Fukuoka, Japan, Kinnas, S.A., Jeon, C.-H., Tian, Y., 2012b. A hybrid viscous/potential flow method for the
pp. 744–759. prediction of the wetted and cavitating performance of ducted propellers. In: SNAME
Abdel-Maksoud, M., Steden, M., Hundemer, J., 2010, September. Design of a multi- Propellers/Shafting Symposium 2012, VA, USA.
component propulsor. In: Proceedings of 28th Symposium on Naval Hydrodynamics, Kinnas, S.A., Jeon, C.H., Purohit, J., Tian, Y., 2013. Prediction of the unsteady cavitating
Pasadena, pp. 12–17. performance of ducted propellers subject to an inclined inflow. In: Proceedings of the
Baltazar, J., Falc~
ao de Campos, J.A.C., Bosschers, J., 2012. Open-water thrust and torque Third International Symposium on Marine Propulsors (SMP2013), Launceston,
predictions of a ducted propeller system with a panel method. Int. J. Rotating Mach. Tasmania, Australia.
2012 https://doi.org/10.1155/2012/474785. Article ID 474785, 11 pages. Kort, L., 1934. Der Neue Dusenschrauben-antrieb. Werft, Reederei und Hafen.
Baltazar, J., Rijpkema, D., Falcao de Campos, J.A.C., Bosschers, J., 2013, May. Menter, F.R., Kuntz, M., Langtry, R., 2003. Ten years of industrial experience with the SST
A comparison of panel method and RANS calculations for a ducted propeller system turbulence model. Turbul. heat mass Transf. 4 (1), 625–632.
in open-water. In: Proceedings of the Third International Symposium on Marine Olivucci, P., Gaggero, S., 2016. A framework for the design by optimization of hydrofoils
Propulsors (SMP2013), Launceston, Tasmania, Australia. under cavitating conditions. In: Proceedings of the 7th European Congress on
Baltazar, J., Falcao de Campos, J.A.C., Bosschers, J., 2015. Potential flow modelling of Computational Methods in Applied Sciences and Engineering, ECCOMAS Congress
ducted propellers with a panel method. In: Proceedings of the Fourth International 2016. Crete, Greece.
Symposium on Marine Propulsors (SMP2015). Austin, Texas, USA. Oosterveld, M.W.C., 1970. Wake Adapted Ducted Propellers. Tech. rep. 345. Netherlands
Bhattacharyya, A., Krasilnikov, V., Steen, S., 2016a. Scale effects on open water Ship Model Basin, Wageningen, Netherlands.
characteristics of a controllable pitch propeller working within different duct designs. OpenFOAM Foundation, 2017. OpenFOAM 4.1.0. The OpenFOAM Foundation. www.
Ocean. Eng. 112, 226–242. openfoam.org.
Bhattacharyya, A., Krasilnikov, V., Steen, S., 2016b. A CFD-based scaling approach for Roache, P.J., 1994. Perspective: a method for uniform reporting of grid refinement
ducted propellers. Ocean. Eng. 123, 116–130. studies. ASME J. Fluids Eng. 116.
Bontempo, R., Cardone, M., Manna, M., Vorraro, G., 2014. Ducted propeller flow analysis Sanchez-Caja, A., Pylkk€anen, J.V., Sipil€a, T.P., 2008, October. Simulation of the
by means of a generalized actuator disk model. Energy Procedia 45, 1107–1115. incompressible viscous flow around ducted propellers with rudders using a RANSE
Bontempo, R., Cardone, M., Manna, M., 2016. Performance analysis of ducted marine solver. In: In 27th Symposium on Naval Hydrodynamics.
propellers. Part I–Decelerating duct. Appl. Ocean Res. 58, 322–330. Sanchez-Caja, A., Perez-Sobrino, M., Quereda, R., Nijland, M., Veikonheimo, T., Gonzalez-
Bosschers, J., Willemsen, C., Peddle, A., Rijpkema, D., 2015. Analysis of ducted propellers Adalid, J., Saisto, I., Uriarte, A., 2013. Combination of pod, CLT and CRP propulsion
by combining potential flow and RANS methods. In: Proceedings of the Fourth for improving ship efficiency: the TRIPOD project. In: Proceedings of Third
International Symposium on Marine Propulsors (SMP2015), Austin, Texas, USA. International Symposium on Marine Propulsors (SMP2013), Launceston, Tasmania,
Çelik, F., Dogrul, A., Arõkan, Y., 2011. Investigation of the optimum duct Geometry for A Australia.
Passenger ferry. In: Proceedings of the 9th Symposium on High Speed Marine Stipa, L., 1932. Experiments with intubed propellers. L’Aerotecnica 1932, 923–953.
Vehicles. Translated by D. M. Miner, NACA TM 655.
Esteco, S.R.L., 2016. ModeFRONTIER 2016 User's Manual. Tamura, Y., Nanke, Y., Matsuura, M., Taketani, T., Kimura, K., Ishii, N., 2010.
Gaggero, S., Villa, D., 2017. Steady cavitating propeller performance by using Development of a high performance ducted propeller. In: Proc. 21st Int. Tug Salvage
OpenFOAM, StarCCMþ and a boundary element method. Proc. Inst. Mech. Eng. Part Convent. Exhibit. (ITS), pp. 1–8.
M J. Eng. Marit. Environ. 231 (2), 411–440. https://doi.org/10.1177/ Van Manen, J.D., 1954. Open Water test series with propellers in nozzles. Int. Shipbuild.
1475090216644280. Prog. 1 (3).
Gaggero, S., Rizzo, C.M., Tani, G., Viviani, M., 2012. EFD and CFD design and analysis of Van Manen, J.D., 1957. Recent research on propellers in nozzles. Int. Shipbuild. Prog. 4
a propeller in decelerating duct. Int. J. Rotating Mach. 2012 https://doi.org/ (36).
10.1155/2012/823831. Article ID 823831, 15 pages. Van Manen, J.D., 1962, May. Effect of Radial Load Distribution on the Performance of
Gaggero, S., Rizzo, C.M., Tani, G., Viviani, M., 2013, May. Design, analysis and Shrouded Propellers. Paper 7, Transactions of the Royal Institution of Naval
experimental characterization of a propeller in decelerating duct. In: Proceedings of Architects.
the Third International Symposium on Marine Propulsors (SMP2013), Launceston, Van Manen, J.D., Oosterveld, M.W.C., 1966. Analysis of ducted propeller design. Trans.
Tasmania, Australia. SNAME 74, 522–561.
Gaggero, S., Gonzalez-Adalid, J., Sobrino, M.P., 2016a. Design of contracted and tip Van Manen, J.D., Superina, A., 1959. The design of screw propellers in nozzles. Int.
loaded propellers by using boundary element methods and optimization algorithms. Shipbuild. Prog. 6 (55), 95–113.
Appl. Ocean Res. 55, 102–129. Vernengo, G., Bonfiglio, L., Gaggero, S., Brizzolara, S., 2016. Physics-based design by
Gaggero, S., Gonzalez-Adalid, J., Sobrino, M.P., 2016b. Design and analysis of a new optimization of unconventional supercavitating hydrofoils. J. Ship Res. 60 (4),
generation of CLT propellers. Appl. Ocean Res. 59, 424–450. 187–202.
Gaggero, S., Villa, D., Viviani, M., 2017. An extensive analysis of numerical ship self- Villa, D., Gaggero, S., Brizzolara, S., 2011, March. Simulation of ship in self propulsion
propulsion prediction via a coupled BEM/RANS approach. Appl. Ocean Res. 66, with different CFD methods: from actuator disk to potential flow/RANS coupled
55–78. solvers. In: Proceedings of International Conference-developments in Marine CFD
Gaggero, S., Tani, G., Villa, D., Viviani, M., Ausonio, P., Travi, P., Bizzarri, G., Serra, F., RINACFD2011, London England, pp. 1–12.
2017b. Efficient and multi-objective cavitating propeller optimization: an application Weller, H.G., Tabor, G., Jasak, H., Fureby, C., 1998. A tensorial approach to
to a high-speed craft. Appl. Ocean Res. 64, 31–57. computational continuum mechanics using object-oriented techniques. Comput.
Goldstein, S., 1929. On the vortex theory of screw propellers. In: Proceedings of the Royal Phys. 12 (6), 620–631.
Society of London, Series a, vol. 123, pp. 440–465, 1929. Yu, L., Greve, M., Druckenbrod, M., Abdel-Maksoud, M., 2013. Numerical analysis of
Haimov, H., Vicario, J., Del Corral, J., 2011, June. Ranse code application for ducted and ducted propeller performance under open water test condition. J. Mar. Sci. Technol.
endplate propellers in open water. In: Proceedings of the Second International 18 (3), 381–394.
Symposium on Marine Propulsors. Zondervan, G.J., Hoekstra, M., Holtrop, J., 2006. Flow analysis, design and testing of
Hoekstra, M., 2006. A RANS-based analysis tool for ducted propeller systems in open ducted propellers. In: Proc., Propellers/Shafting 2006 Symposium, Virginia, USA.
water condition. Int. Shipbuild. Prog. 53 (3), 205–227.

20

You might also like