numericalpredictionsofcavitatingflowaroundmodelscale ... · 2019. 7. 31. · pptc propeller. the...

12
Hindawi Publishing Corporation International Journal of Rotating Machinery Volume 2012, Article ID 618180, 11 pages doi:10.1155/2012/618180 Research Article Numerical Predictions of Cavitating Flow around Model Scale Propellers by CFD and Advanced Model Calibration Mitja Morgut and Enrico Nobile Dipartimento di Ingegneria e Architettura, Universit` a degli Studi Trieste, P.le Europa 1, 34127 Trieste, Italy Correspondence should be addressed to Mitja Morgut, [email protected] Received 3 February 2012; Accepted 5 June 2012 Academic Editor: Jules W. Lindau Copyright © 2012 M. Morgut and E. Nobile. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. The numerical predictions of the cavitating flow around two model scale propellers in uniform inflow are presented and discussed. The simulations are carried out using a commercial CFD solver. The homogeneous model is used and the influence of three widespread mass transfer models, on the accuracy of the numerical predictions, is evaluated. The mass transfer models in question share the common feature of employing empirical coecients to adjust mass transfer rate from water to vapour and back, which can aect the stability and accuracy of the predictions. Thus, for a fair and congruent comparison, the empirical coecients of the dierent mass transfer models are first properly calibrated using an optimization strategy. The numerical results obtained, with the three dierent calibrated mass transfer models, are very similar to each other for two selected model scale propellers. Nevertheless, a tendency to overestimate the cavity extension is observed, and consequently the thrust, in the most severe operational conditions, is not properly predicted. 1. Introduction In the field of marine applications, and in the particular case of marine propellers, the onset of cavitation is, in general, associated with negative design implications such as thrust reduction, noise, vibration, and erosion. In the last decades, also owing to the steady increasing of the computer performances, in order to improve the design process, several CFD (Computational Fluid Dynamics) methods have been developed for the prediction of the cavitation appearance and the estimation of its eects. In this study, we evaluated the capabilities of the homo- geneous, that is, one-fluid model implemented in the ANSYS CFX 12 (for brevity CFX hereafter) commercial CFD solver, for the prediction of cavitating flow around model scale propellers working in uniform inflow. This model treats the cavitating flow as a mixture of two fluids behaving as a single one. The set of the governing equations is composed by the (volume) continuity and momentum equations for the mixture, plus a transport equation for the water volume fraction. The mass transfer rate due to cavitation is regulated by the same source term appearing in the (volume) continuity and volume fraction equations. This source term, in CFX, using the default setting is modelled by employing the mass transfer model originally proposed by Zwart et al. [1] (Zwart for brevity). However, in literature several other mass transfer models are available (see [2]). Thus, in order to improve the reliability of the simulations besides the Zwart model we employed also two other widespread mass transfer models: the model originally proposed by Kunz et al. [3] (for brevity Kunz) and the model originally proposed by Singhal et al. [4] also known as Full Cavitation Model (FCM hereafter). The Kunz and FCM models were added to the CFX solver using CEL (CFX Expression Language). It is fundamental to clarify that the considered mass transfer models share the common feature of employing empirical coecients to adjust the mass transfer rate due to condensation and evaporation processes, and their values can significantly aect both the stability and accuracy of the numerical predictions. Thus, in order to correctly evaluate the influence of the three dierent mass transfer models on the stability and accuracy of the numerical predictions, their empirical coecients were first properly and congruently calibrated using an optimization strategy [5]. The models

Upload: others

Post on 13-Apr-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

Hindawi Publishing CorporationInternational Journal of Rotating MachineryVolume 2012, Article ID 618180, 11 pagesdoi:10.1155/2012/618180

Research Article

Numerical Predictions of Cavitating Flow around Model ScalePropellers by CFD and Advanced Model Calibration

Mitja Morgut and Enrico Nobile

Dipartimento di Ingegneria e Architettura, Universita degli Studi Trieste, P.le Europa 1, 34127 Trieste, Italy

Correspondence should be addressed to Mitja Morgut, [email protected]

Received 3 February 2012; Accepted 5 June 2012

Academic Editor: Jules W. Lindau

Copyright © 2012 M. Morgut and E. Nobile. This is an open access article distributed under the Creative Commons AttributionLicense, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properlycited.

The numerical predictions of the cavitating flow around two model scale propellers in uniform inflow are presented and discussed.The simulations are carried out using a commercial CFD solver. The homogeneous model is used and the influence of threewidespread mass transfer models, on the accuracy of the numerical predictions, is evaluated. The mass transfer models in questionshare the common feature of employing empirical coefficients to adjust mass transfer rate from water to vapour and back, whichcan affect the stability and accuracy of the predictions. Thus, for a fair and congruent comparison, the empirical coefficients of thedifferent mass transfer models are first properly calibrated using an optimization strategy. The numerical results obtained, with thethree different calibrated mass transfer models, are very similar to each other for two selected model scale propellers. Nevertheless,a tendency to overestimate the cavity extension is observed, and consequently the thrust, in the most severe operational conditions,is not properly predicted.

1. Introduction

In the field of marine applications, and in the particularcase of marine propellers, the onset of cavitation is, ingeneral, associated with negative design implications such asthrust reduction, noise, vibration, and erosion. In the lastdecades, also owing to the steady increasing of the computerperformances, in order to improve the design process, severalCFD (Computational Fluid Dynamics) methods have beendeveloped for the prediction of the cavitation appearanceand the estimation of its effects.

In this study, we evaluated the capabilities of the homo-geneous, that is, one-fluid model implemented in the ANSYSCFX 12 (for brevity CFX hereafter) commercial CFD solver,for the prediction of cavitating flow around model scalepropellers working in uniform inflow. This model treats thecavitating flow as a mixture of two fluids behaving as asingle one. The set of the governing equations is composedby the (volume) continuity and momentum equationsfor the mixture, plus a transport equation for the watervolume fraction. The mass transfer rate due to cavitation isregulated by the same source term appearing in the (volume)

continuity and volume fraction equations. This source term,in CFX, using the default setting is modelled by employingthe mass transfer model originally proposed by Zwart et al.[1] (Zwart for brevity). However, in literature several othermass transfer models are available (see [2]). Thus, in orderto improve the reliability of the simulations besides the Zwartmodel we employed also two other widespread mass transfermodels: the model originally proposed by Kunz et al. [3](for brevity Kunz) and the model originally proposed bySinghal et al. [4] also known as Full Cavitation Model (FCMhereafter). The Kunz and FCM models were added to theCFX solver using CEL (CFX Expression Language).

It is fundamental to clarify that the considered masstransfer models share the common feature of employingempirical coefficients to adjust the mass transfer rate dueto condensation and evaporation processes, and their valuescan significantly affect both the stability and accuracy of thenumerical predictions. Thus, in order to correctly evaluatethe influence of the three different mass transfer models onthe stability and accuracy of the numerical predictions, theirempirical coefficients were first properly and congruentlycalibrated using an optimization strategy [5]. The models

Page 2: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

2 International Journal of Rotating Machinery

were calibrated considering the two-dimensional stable sheetcavity flow around a NACA-66(MOD) hydrofoil. The cali-brated mass transfer models were then used to investigate thecavitating flow around two model scale propellers recognizedas international benchmarks, that is, E779A propeller andPPTC propeller.

The numerical results were compared with the avail-able experimental data. For both propellers, three selectedoperational conditions, corresponding to the appearance ofpartial and/or tip vortex cavitation, were considered [6, 7].For such conditions, the numerical results obtained with thedifferent calibrated mass transfer models were very similar toeach other and compared well with the experimental data,even though the numerical cavitation patterns were slightlyoverestimated. However, in the case of E779A propellerwhere also the more severe operational conditions weresimulated, the thrust breakdown was not properly predicted.

In the following, the homogeneous (one-fluid) modelused in this study is presented first, followed by thedescription of the optimization strategy used to calibrate thedifferent mass transfer models. Then the results obtainedconsidering the cavitating flow around the two differentmodel scale propellers are provided. Eventually, some con-cluding remarks are given.

2. Mathematical Model

Cavitating flows can be modelled using several methods.An excellent review of different methods is provided forinstance by Koop [8]. In this work, we used the homogeneoustransport equation-based model [9, 10], described in thefollowing.

2.1. Governing Equations. As already indicated the cavitatingflow is modelled as a mixture of two species, that is, vapourand liquid behaving as a single one, where both phases sharethe same velocity as well as pressure fields.

In the case of the RANS (Reynolds-averaged NavierStokes) approach to turbulence, employing the eddy viscositymodels and assuming both liquid and vapour phases incom-pressible, turbulent cavitating flows can be described by thefollowing set of governing equations:

∇ ·U = m

(1ρL− 1

ρV

),

∂(ρU)

∂t+∇ · (ρUU

) = −∇P +∇

·[(μ + μt

)(∇U + (∇U)T)]

+ S,

∂γ

∂t+∇ · (γU

) = m

ρL,

(1)

which are, in order, the volume continuity and the momen-tum equation for the liquid-vapour mixture and the volumefraction equation for the liquid phase. In the above equa-tions, ρL (kg/m3) and ρV (kg/m3) are the liquid and vapour

densities, respectively. U (m/s) is the averaged velocity andP (Pa) the averaged pressure. S represents the additionalmomentum sources (e.g., the Coriolis and centrifugal forcesin the rotating frame of reference). γ is the water volumefraction which is related to the vapour volume fraction αthrough the volume fraction constraint:

γ + α = 1. (2)

The mixture density ρ (kg/m3) and laminar dynamic viscos-ity μ (kg/ms) are evaluated according to

ρ = γρL +(1− γ

)ρV ,

μ = γμL +(1− γ

)μV ,

(3)

where μL (kg/ms) and μV (kg/ms) are the liquid and vapourdynamic viscosities, respectively. m (kg/m3s) represents theinterphase transfer rate due to cavitation. In this study, m wasmodelled using alternatively the three different mass transfermodels presented in the next section.

Finally, in order to close the system of the govern-ing equations, the mixture turbulent viscosity μt (kg/ms)was evaluated using the eddy viscosity turbulence modelsdeveloped for the single phase flow. In particular, the two-equations standard k-ε and SST (shear stress transport)turbulence models were employed. For the description of theabove turbulence models we refer to [10–12].

2.2. Mass Transfer Models. The mass transfer modelsdescribe/regulate the interphase mass transfer rate due tocavitation. In the last two decades, several authors proposeddifferent mass transfer models.

In the following, we provide a brief description of threedifferent mass transfer models employed in this study, wherethe interphase mass transfer rate due to cavitation wasassumed positive if directed from vapour to water. Thesemodels, except for the full cavitation model, are simplifiedversions of the original formulations, where the contributionof the dissolved gasses was omitted.

2.2.1. Zwart Model. The Zwart model is the native CFX masstransfer model. It is based on the simplified Rayleigh-Plessetequation for bubble dynamics [13]:

m =

⎧⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎩

−Fe 3rnuc(1− α)ρVRB

√23PV − P

ρLif P < PV

Fc3αρVRB

√23P − PVρL

if P > PV .

(4)

In the above equations, PV is the vapour pressure, rnuc isthe nucleation site volume fraction, RB is the radius of anucleation site, and Fe and Fc are two empirical calibrationcoefficients for the evaporation and condensation processes,respectively. In CFX, the above mentioned coefficients, bydefault, are set as follows: rnuc = 5.0 × 10−4, RB = 2.0 ×10−6 m, Fe = 50, and Fc = 0.01.

Moreover, the above equations show that expressionsfor condensation and evaporation are not symmetric.

Page 3: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

International Journal of Rotating Machinery 3

In particular, in the expression for evaporation, α is replacedby rnuc(1−α) to take into account that, as the vapour volumefraction increases, the nucleation site density must decreaseaccordingly.

2.2.2. Full Cavitation Model. The mass transfer modelproposed by Singhal et al. [4], originally known as fullcavitation model, is currently employed in some commercialCFD codes, that is, FLUENT [14] and PUMPLINX [15]. Thismodel is also based on the reduced form of the Rayleigh-Plesset equation for bubble dynamics, and its formulationstates as follows:

m =

⎧⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎩

−Ce

√k

TρLρV

√23PV − P

ρL

(1− fV

)if P < PV

Cc

√k

TρLρL

√23P − PVρL

fV if P > PV ,

(5)

where fV is the vapour mass fraction, k (m2/s2) is theturbulent kinetic energy, T (N/m) is the surface tension, andCe = 0.02 and Cc = 0.01 are two empirical calibrationcoefficients.

It is important to note that in this work, for convenience,we did not use the original formulation of the model, butthe formulation derived by Huuva [16] in which the vapourmass fraction, fV , is replaced by the vapour volume fractionα.

2.2.3. Kunz Model. The Kunz mass transfer model is basedon the work of Merkle et al. [17] and currently is one ofthe mass transfer models implemented in the OpenFOAMlibrary [18]. In this model, unlike the above mentionedmodels, the mass transfer is based on two different strategiesfor creation m+ and destruction m− of liquid. The transfor-mation of liquid to vapour is calculated as being proportionalto the amount by which the pressure is below the vapourpressure. The transformation of vapour to liquid, otherwise,is based on a third-order polynomial function of volumefraction, γ. The specific mass transfer rate is defined as m =m+ + m−:

m+ = CprodρVγ2(1− γ

)t∞

,

m− = CdestρVγ min[0,P − PV ](1/2ρLU2∞

)t∞

.

(6)

In the above equations, U∞ (m/s) is the free-stream velocityand t∞ = L/U∞ is the mean flow time scale, where L is thecharacteristic length scale. Cdest and Cprod are two empiricalcoefficients. In the original formulation Cdest = 100, Cprod =100.

3. Calibration of Mass Transfer Models

As described in the former section, in the homogeneoustransport equation-based model, a particular mass transfer

model which regulates the mass transfer rate from liquid tovapour and back is needed in order to simulate cavitatingflow.

In this study, we employed the Zwart model, the FCMmodel, and the Kunz model. As already indicated in the pre-vious section, these models employ empirical coefficients totune the models of condensation and evaporation processes,that in turn can influence the accuracy and stability of thenumerical predictions.

Thus, in order to ensure, for all three different masstransfer models, the stability and a good level of accuracyof the predictions, the empirical coefficients of the threedifferent mass transfer models were properly calibrated usingan optimization strategy. The entire calibration process wasdriven by the modeFRONTIER 4.2 optimization systemand is described in the following. We remind that mode-FRONTIER [19] is a general integration and multiobjectiveoptimization platform, which can drive a variety of CAEpackages and provide statistical and visualization tools. Itis commonly used for functional and shape optimizationof systems and devices, for example, see [20]. It is worthto say that with the developed strategy, the empiricalcoefficients were tuned considering the two-dimensionalsheet cavity flow around the NACA66(MOD) hydrofoil [21].In this manner, due to the lower computational costs, wecould explore, in a reasonable computational time and withlimited computational resources, a significant number ofcoefficients’ combinations. Consequently, we had also theopportunity to verify if the coefficients calibrated for thehydrofoil case, could have a general character, and could besuccessfully applied to the propeller flow problem.

3.1. The Idea and Logic of the Optimization Strategy. In thedeveloped optimization strategy, the three different masstransfer models were properly calibrated by searching thevalues of the empirical coefficients which minimized theobjective function f . The objective function represented thedifferences between the numerical and experimental pressuredistributions on the suction side of a NACA66(MOD)hydrofoil evaluated at an angle of attack AoA = 4◦ and forthree different cavitating flow regimes. More precisely, theobjective function f was expressed as follows:

f =∑σ

N∑i=1

∣∣∣CPi − CPi,Exp

∣∣∣; σ = 1.00, 0.91, 0.84, (7)

where CPi and CPi,Exp were the numerical and experimentalvalues of the pressure coefficient, taken at N = 12 locationson the suction side of the hydrofoil, and σ was the cavitationnumber defined as follows:

σ = PREF − PV(1/2)ρLU2∞

, (8)

where PREF (Pa) was the reference pressure and U∞ (m/s) thefree-stream velocity.

In order to find the values which minimized f , twodifferent optimization algorithms were run in sequence.The design space was firstly explored using the MOGA-II, a multiobjective genetic optimization algorithm available

Page 4: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

4 International Journal of Rotating Machinery

in modeFRONTIER 4.2 [19]. The MOGA-II was run forten generations starting from an initial DOE (design ofexperiments) of ten designs randomly generated. In orderto refine the solution, the Simplex [22], single objectiveoptimization algorithm was used, starting from the threebest solutions obtained using MOGA-II. No significantimprovements were observed using the Simplex algorithm.

Figure 1 shows the logic of the optimization strategy,where the optimizer block stays for both the MOGA-II andSimplex algorithms, and where (X1, X2) represent the coupleof empirical coefficients for all three different mass transfermodels.

The numerical setup adopted to run the simulations,within the optimization process, is briefly described in thefollowing. Further details can be found in [5].

3.2. Simulation Setup. During the optimization process, thesimulations were carried out using the following strategy.

The cavitating flow around the hydrofoil was simu-lated on the rectangular domain shown in Figure 2. Thesimulations were carried out in 2D and assuming steady-state conditions. For turbulence closure, the standard k-ε turbulence model, in combination with scalable wallfunctions [10], was employed. The following boundaryconditions were applied. On solid surfaces (top, bottom, andNACA66(MOD)), the no-slip condition was set. On outletboundary, a fixed static pressure, PREF = 202,650 Pa wasimposed and on the side faces the symmetry condition wasenforced. On inlet boundary, the values of the free-streamvelocity components and turbulence quantities were fixed.Water and vapour volume fractions were set equal to 1 and0, respectively. In order to match the experimental setup,during the numerical simulations the same Reynolds numberwas used. Since the water kinematic viscosity was assumedνL = 8.92 × 10−7 m2/s, the free-stream velocity was set toU∞ = 12.2 (m/s). Assuming a turbulence intensity of 1%, theturbulent kinetic energy k and the turbulent dissipation rateε were set equal to k = 0.0223 m2/s2, ε = 0.1837 m2/s3 onthe inlet boundary. The water density was set equal to ρL =997 kg/m3 and the maximum water-vapour density ratio waslimited to ρL/ρV = 1000 in order to ensure solver stability. Allthe simulations were performed on a hexahedral grid with58700 nodes which proved to give mesh independent results[5], for fully wetted flow. Figure 3 shows the computationalmesh around the hydrofoil generated with the ANSYS-ICEMmesh generator (ICEM hereafter). The average value of y+

evaluated on the solid surfaces of the hydrofoil was equal to28. y+ was defined as y+ = μτ y/ν, where μτ = (τw/ρ)1/2 is thefriction velocity, y is the normal distance from the wall, andτw is the wall shear stress.

3.3. Calibration Results. In this study, the evaporation Fe andcondensation Fc coefficients of the Zwart model were tunedwithin the following ranges: 30 ≤ Fe ≤ 500, 5.0 × 10−4 ≤Fc ≤ 8.0× 10−2. The best result was found with Fe = 300 andFc = 0.03.

For FCM, the values of the evaporation coefficient Ce

and of the condensation coefficient Cc were tuned within

Generate initial set of designs by DOE

Solve cavitating flows

Compute objective,

function f(X1, X2)

Stopcondition

Set current design, couple (X1, X2)

Postprocessing

True

False

modeFRONTIER

Optimizer

CFD solver

Figure 1: Logic of the optimization strategy.

−3.5c 0 5.5c−2.5c

0

2.5c

Outlet

Top

Bottom

NACA66(MOD)Inlet

Figure 2: NACA66(MOD), shape of the computational domain.

Figure 3: Mesh around the NACA66(MOD) hydrofoil.

Page 5: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

International Journal of Rotating Machinery 5

Table 1: Default and calibrated coefficients for the different masstransfer models.

Zwart FCM Kunz

Fe Fc Ce Cc Cdest Cprod

Default 50 0.01 0.02 0.01 100 100

Calibrated 300 0.03 0.40 2.3E − 04 4100 455

the following ranges: 0.01 ≤ Ce ≤ 1, 1.0 × 10−6 ≤ Cc ≤1.0 × 10−2. The best solution was found with Ce = 0.40 andCc = 2.30× 10−4.

For Kunz, the designs space was defined as follows: 100 ≤Cdest ≤ 5000, 10 ≤ Cprod ≤ 1000. The best solution wasachieved with Cprod = 455 and Cdest = 4100. In the Kunzmodel, in order to set the mean flow time scale t∞ = L/U∞,the chord of the hydrofoil was chosen as a length scale, L,following [16].

For the sake of completeness, Table 1 highlights that theempirical coefficients which resulted from the calibrationprocess were quite different compared to the default ones.The differences between the calibrated and default values ofthe coefficients were lower for the Zwart model, probablybecause being this one the native mass transfer model ofCFX, the suggested empirical values have been already foundto work well for a wide range of applications.

Figure 4 shows, for instance for σ = 0.91, how the cavitiespredicted with the noncalibrated mass transfer models had alower vapour content and were shorter than those obtainedwith the calibrated models. These differences were morepronounced for the FCM and Kunz mass transfer models.As a matter of fact considering the suction side pressuredistributions depicted in Figure 5, it is possible to note thatwith the noncalibrated mass transfer models, and especiallywith the Kunz and FCM mass transfer models, the numericalcavities were significantly underpredicted compared to theexperimental ones, for all the three different cavitating flowregimes, that is, σ = 1.00, 0.91, and 0.84. On the otherhand, from the cavitation patterns presented in Figure 4, andthe suction side pressure distributions given in Figure 6, it ispossible to appreciate that the numerical results provided bythe three different calibrated mass transfer models were veryclose to each other and in line with the experimental data.

Finally, let us clarify that before being applied to thepredictions of the three-dimensional cavitating flow aroundmodel scale propellers, the three different calibrated masstransfer models were first assessed considering further two-dimensional sheet cavity flow regimes around a hydrofoil [5].

4. Model Scale Propellers

4.1. Test Cases. In this study, the model scale propellersE779A and PPTC, depicted in Figure 7, were considered.

The E779A propeller is a four-bladed, fixed-pitch, andlow-skew propeller, designed in 1959 with a diameter D= 0.2272 m. Since 1997, it has been used in experimentalactivities performed by INSEAN (Istituto Nazionale di Studied Esperienze di Architettura Navale) aimed at providing

Table 2: Distances of the boundaries/surfaces from the propellermid plane in axial direction for inlet, outlet, front, aft, and from thepropeller rotation axis centreline in radial direction for outer andtop.

Propeller Inlet Outlet Outer Front Aft Top

E779A 3.40D 5.35D 5.00D 0.40D 0.35D 0.70D

PPTC 2.30D 5.30D 5.00D 0.41D 0.31D 0.60D

a thorough characterization of marine propeller hydrody-namics and hydroacoustics over a wide range of operationalconditions. It has been widely used for the validation of CFDcodes (e.g., see [23–25]).

The PPTC (Potsdam Propeller Test Case) is a five-bladed,controllable pitch propeller having a diameter D = 0.250 mwhich was used as a blind test case at the 2011 Workshopon Cavitation and Propeller Performance, [26]. A significantamount of the data covering propeller’s geometry, openwater tests, velocity field measurements, and cavitation testsis currently available in [27]. The experimental data wererecorded in the towing tank and cavitation tunnel of the SVA(Potsdam Model Basin) [7, 28, 29].

4.2. Solution Strategy. The propellers were assumed to workin a uniform inflow and thus only one passage blade wasmodelled for computational convenience. Figures 8 and 9show the shapes of the computational domains used in thecase of the E779A propeller and PPTC propeller, respectively.The domains’ dimensions are listed in Table 2.

Both domains were subdivided in two regions, that is,Rotating and Fixed, and the following boundary conditionswere applied: on inlet boundary, the free-stream velocitycomponents and a turbulence level of 1% were set. Onoutlet boundary, a fixed value of the static pressure wasimposed. On the periodic boundaries (sides of the domain),the rotational periodicity was ensured. On solid surfacesthe no-slip boundary condition was applied, and on outerboundary, the slip condition was set. Since the propellerrotation was simulated using the MRF (Multiple ReferenceFrame) approach in the stationary region called Fixed, thegoverning equations were solved into a fixed frame ofreference, while into a rotating region called Rotating, thegoverning equations were solved into a rotating frame ofreference. For the discretization of the advective terms, thehigh resolution scheme was used [10].

For turbulence closure, the two-equation SST turbulencemodel was used in combination with the automatic walltreatment [10, 30].

For a given value of the advance coefficient J = U/nD,a particular cavitating flow regime was set according tothe cavitation number σn defined, for the propeller case, asfollows:

σn = PREF − PV0.5ρL(nD)2 , (9)

where n (rps) is the propeller rotational speed and PV (Pa) isthe vapour pressure. In this study, we assumed PREF = POutlet.

Page 6: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

6 International Journal of Rotating Machinery

Zwart-default

FCM-default

Kunz-default

0 0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Vapour volume fraction

(a)

Zwart-calibrated

FCM-calibrated

Kunz-calibrated

0 0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Vapour volume fraction

(b)

Figure 4: Sheet cavities over the NACA66(MOD) hydrofoil at AoA = 4◦, Re = 2 × 106, σ = 0.91, computed using noncalibrated (a) andcalibrated (b) mass transfer models.

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

1.2

X/C

Noncalibrated mass transfer models

ZwartFCMKunz

σ = 1 exp.σ = 0.91 exp.σ = 0.84 exp.

−Cp

Figure 5: NACA66(MOD) hydrofoil at AoA = 4o, Re = 2 × 106.Suction side pressure distributions computed using noncalibratedZwart, FCM, and Kunz mass transfer models.

For all the different cavitating flow regimes, the numer-ical predictions were carried out using alternatively all thethree different calibrated mass transfer models.

4.3. Meshing. For both propellers, the meshes of the twodifferent domain regions, that is, Fixed and Rotating, were

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

1.2

ZwartFCMKunz

Calibrated mass transfer models

σ = 1 exp.σ = 0.91 exp.σ = 0.84 exp.

X/C

−Cp

Figure 6: NACA66(MOD) hydrofoil at AoA = 4o, Re = 2 ×106. Suction side pressure distributions computed using calibratedZwart, FCM, and Kunz mass transfer models.

generated independently from each other. They were gener-ated with ICEM without trying to ensure a 1 : 1 matchingof the nodes at the interfaces (the common surfaces of thetwo different domain regions). The different domain regionswere subsequently joined in CFX using the GGI (GeneralizedGrid Interfaces) capabilities of the solver. From Table 3, it

Page 7: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

International Journal of Rotating Machinery 7

(a) (b)

Figure 7: Propellers’ CAD models. View looking downstream. E779A propeller (a), PPTC propeller (b). The propellers’ models are notshown using the same scale.

Outer

Fixed

Front

Inlet

Outlet

Top

Aft

Figure 8: E779A, computational domain.

Outer

FixedFront

Inlet

OutletTopRotating

Aft

Figure 9: PPTC, computational domain.

is possible to note that for both propellers the Rotatingregion was discretized by a hexa-structured mesh which wasgenerated, as suggested in [31, 32], by decomposing thatdomain region in a large number of blocks. The resolutionand also the quality of the cells were set through a propernode distribution on the blocks edges.

Table 3: Computational grids.

PropellerNodes at rotating Nodes at fixed

Total nodesHexa Hexa Hybrid

E779A 644655 163041 807696

PPTC 1838655 275680 2114335

As far as the discretization of the Fixed region isconcerned, in the case of the E779A propeller a hybrid-unstructured mesh was used, while in the case of the PPTCpropeller a hexa-structured mesh was employed. In thecase of E779A propeller, we preferred to use a hybrid-unstructured mesh for Fixed, because we observed that usingthe hexastructured approach the block decomposition led tohighly distorted elements in the twisted region over Rotating.

It is important to clarify that both mesh arrangementsused here proved to guarantee mesh independent results informer studies [33, 34], and the average value of y+ measuredon blade surfaces was approximatively equal to 38 for E779Apropeller and 32 for PPTC propeller. Figure 10 illustrates theblade surface meshes of both propellers.

4.4. Results and Discussion. Before discussing the results, itis important to point out that in all the simulations themaximum density ratio was limited to ρL/ρV = 1000 inorder to guarantee solver stability. Moreover, in the caseof the Kunz model the value of t∞ was set according to

the operational conditions as t∞ = C0.7R/√U2 + (2πn0.7R)2,

where C0.7R was the propeller blade chord at 70% of propellerradius R (m) and U (m/s) the incoming free-stream velocity.

The numerical results were quantitatively comparedconsidering the global values of the problem represented bythe thrust KT and torque KQ coefficients defined as follows:

KT = T

ρLn2D4,

KQ = Q

ρLn2D5,

(10)

where T (N) and Q (Nm) were the propeller thrust andtorque, n (rps) was the propeller rotational speed and D (m)the propeller diameter.

Page 8: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

8 International Journal of Rotating Machinery

(a) (b)

Figure 10: Suction side surface meshes, E779A propeller (a), PPTC propeller (b).

Zwart

Zwart

Zwart

FCM

FCM

FCM

Kunz

Kunz

Kunz

Exp. J = 0.71, σn = 1.763

Exp. J = 0.77, σn = 1.783

Exp. J = 0.83, σn = 2.063

Figure 11: Comparison of the cavitation patterns for three different operational conditions for the E779A propeller. The numerical cavitationpatterns, obtained using the three different mass transfer models, are depicted as isosurfaces of vapour volume fraction α = 0.5.

The relative percentage errors of the computed KT , KQ

values were defined as follows:

ΔKT (%) = KT ,CFD − KT ,EXP

KT ,EXP· 100,

ΔKQ (%) = KQ,CFD − KQ,EXP

KQ,EXP· 100,

(11)

where KT ,CFD, KQ,CFD were the numerical values and KT ,EXP,KQ,EXP, the experimental values.

For a qualitative comparison, also sketches of cavitationpatterns were considered.

Next, the results obtained for the E779A propeller arepresented first, followed by those obtained for the PPTCpropeller.

4.4.1. E779A Propeller. The cavitating flow predictions werecarried out following the experimental setup suggested byINSEAN [35] and reported in [6, 23].

In Figure 11, the cavitation patterns, predicted forthree particular operational conditions and using the three

Page 9: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

International Journal of Rotating Machinery 9

0 0.5 1 1.5 2 2.5 3 3.50.1

0.15

0.2

0.25

0.3

J = 0.71

J = 0.077

J = 0.83

ExpZwart

FCMKunz

KT

σn

Figure 12: Influence of the cavitation number on the thrustcoefficient for the E779A propeller.

0 0.5 1 1.5 2 2.5 3 3.50.2

0.3

0.4

0.5

J = 0.71

J = 0.77

J = 0.83

ExpZwart

FCMKunz

σn

10 KQ

Figure 13: Influence of the cavitation number on the thrustcoefficient for the E779A propeller.

different mass transfer models, are qualitatively compared.From Figure 11, it is possible to note that the cavitationpatterns predicted with the three different mass transfermodels were very similar to each other. Even though thecavitation patterns were slightly overestimated, from Table 4it is possible to appreciate that for these operational condi-tions the predicted values of thrust and torque coefficientscompared very well with the experimental data.

Figures 12 and 13 show that also covering a wider rangeof operational conditions, the results obtained using thedifferent mass transfer models were similar. However, theeffect of cavitation number on thrust and torque was notproperly reproduced leading to the premature prediction ofthe thrust breakdown.

This discrepancy between the numerical results andexperimental measurements could be probably related to theoverestimation of the cavitation pattern (see Figure 11).

Table 4: Relative percentage errors for the E779A propeller.

J σnΔKT (%) ΔKQ (%)

Zwart FCM Kunz Zwart FCM Kunz

0.71 1.763 −1.01 −2.27 −0.61 −1.63 −3.07 −1.60

0.77 1.783 2.11 1.46 2.35 −2.59 −3.18 −2.52

0.83 2.063 2.77 2.51 2.85 −2.33 −2.52 −2.35

4.4.2. PPTC Propeller. The numerical predictions were car-ried out following the instruction kindly provided by thesmp’11 workshop organizers and thus the cavitating flowpredictions were carried out according to the thrust identity(KT noncavitating flow conditions) for J = 1.019, J =1.269, and J = 1.408. In our case, in order to fulfil thisrequisite within an error tolerance of 3% in the case of thesmaller J value, that is, J = 1.019, the inflow velocity hadto be slightly reduced leading to J = 1.016. Table 5 collectsthe KT values predicted for the three different operationalconditions and those obtained experimentally. Figure 14shows that the cavitation patterns obtained using the threedifferent calibrated mass transfer models were, in general,very similar to each other and compared very well with theexperimental visualizations for the considered operationalconditions, that is, (J = 1.019, σn = 2.024), (J = 1.269,σn = 1.424), and (J = 1.408, σn = 2.000). From Table 6,it is possible to note that the values of the thrust andtorque coefficients predicted using alternatively the threedifferent calibrated mass transfer models were surprisinglyvery close to each other and in excellent agreement with theexperimental data for J = 1.019, σn = 2.024, even though theshape of the cavitation pattern was properly reproduced onlywith the FCM. Regarding other operational conditions, thedifferences between the predicted and experimental valueswere slightly more pronounced for the Zwart model at (J =1.269, σn = 1.424) and for the FCM model at (J = 1.408,σn = 2.000).

5. Concluding Remarks

In this study, a well-known commercial CFD solver, thatis, ANSYS-CFX 12, was used to predict the cavitating flowaround model scale propellers working in uniform inflow.The simulations were carried out using the homogeneousmodel, and the influence of three widespread mass transfermodels, on the accuracy of the numerical predictions, wasevaluated. The considered mass transfer models share thecommon feature of employing empirical coefficients toadjust the mass transfer rate due to the evaporation andcondensation processes, which can affect both accuracy andstability of the numerical predictions. Thus, for a fair andcongruent comparison the empirical coefficients of the threemass transfer models were first properly calibrated usingan optimization strategy driven by the modeFRONTIER 4.2optimization framework. The models were calibrated on thebasis of the two-dimensional stable sheet cavity flow arounda hydrofoil.

Page 10: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

10 International Journal of Rotating Machinery

Zwart

Zwart

Zwart

FCM

FCM

FCM

Kunz

Kunz

Kunz

0.95

0.95

0.95

0.9

0.9

0.9

0.8

0.8

0.8

0.7

0.7

0.7

0.5

0.5

0.5

Suction side

Suction side

Pressure side

Exp. J = 1.019, σn = 2.024

Exp. J = 1.269, σn = 1.424

Exp. J = 1.408, σn = 2

Figure 14: Comparison of the cavitation patterns for three different operational conditions for the PPTC propeller. The numerical cavitationpatterns, obtained using the three different mass transfer models, are depicted as isosurfaces of vapour volume fraction α = 0.5.

Table 5: Fulfil of thrust identity (KT non-cavitating) for PPTCpropeller.

CFD Exp.

J KT J KT

1.016 0.382 1.019 0.387

1.269 0.240 1.269 0.245

1.408 0.166 1.408 0.167

Table 6: Relative percentage errors for the PPTC propeller.

J σnΔKT (%) ΔKQ (%)

Zwart FCM Kunz Zwart FCM Kunz

1.019 2.024 0.13 0.40 0.67 −1.63 −1.11 −1.22

1.269 1.424 −5.04 −1.65 1.74 −4.15 −0.82 3.45

1.408 2.000 −2.35 −4.55 −2.35 −1.84 −2.45 −2.45

The calibrated models were then used to predict thecavitating flow around the model scale propellers E779A andPPTC.

No significant differences were observed in the sim-ulations performed using alternatively the three differentcalibrated mass transfer models. For selected operationalconditions, corresponding to the appearance of partial andtip-vortex cavitation, the numerical results compared well

with the experimental data, even though the tendency toslightly overestimate the cavity extension was observed.Unfortunately, considering other more severe operationalconditions, the numerical results showed significant discrep-ancies with the experimental data. Further investigations onthis aspect are in progress.

Acknowledgments

This study was performed in the context of the projectOpenSHIP, supported by Regione FVG—POR FESR 2007–2013 Obiettivo competitivita regionale e occupazione. Theauthors wish to thank INSEAN, and in particular Dr.Francesco Salvatore, for the assistance and for providing thegeometry and the experimental measurements for the E779Apropeller.

References

[1] P. J. Zwart, A.G. Gerber, and T. Belamri, “A two-phasemodel for predicting cavitation dynamics,” in Proceedings ofthe International Conference on Multiphase Flow (ICMF ’04),Yokohama, Japan, 2004.

[2] S. Frikha, O. Coutier-Delgosha, and J. A. Astolfi, “Influence ofthe cavitation model on the simulation of cloud cavitation on2D foil section,” International Journal of Rotating Machinery,vol. 2008, Article ID 146234, 12 pages, 2008.

Page 11: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

International Journal of Rotating Machinery 11

[3] R. F. Kunz, D. A. Boger, D. R. Stinebring et al., “A pre-conditioned Navier-Stokes method for two-phase flows withapplication to cavitation prediction,” Computers and Fluids,vol. 29, no. 8, pp. 849–875, 2000.

[4] A. K. Singhal, M. M. Athavale, H. Li, and Y. Jiang, “Mathemat-ical basis and validation of the full cavitation model,” Journalof Fluids Engineering, vol. 124, no. 3, pp. 617–624, 2002.

[5] M. Morgut, E. Nobile, and I. Bilus, “Comparison of masstransfer models for the numerical prediction of sheet cavita-tion around a hydrofoil,” International Journal of MultiphaseFlow, vol. 37, no. 6, pp. 620–626, 2011.

[6] F. Pereira, F. Salvatore, and F. D. Felice, “Measurement andmodeling of propeller cavitation in uniform inflow,” Journalof Fluids Engineering, vol. 126, no. 4, pp. 671–679, 2004.

[7] H. J. Heinke, “Potsdam propeller test case (PPTC), cavitationtests with the model propeller,” Tech. Rep. VP1304, SVA(Potsdam Model Basin), 2011.

[8] A. H. Koop, Numerical simulation of unsteady three-dimen-sional sheet cavitation [Ph.D. thesis], University of Twente,2008.

[9] Y. A. Bouziad, Phisical modelling of leading edge cavitation:computational methodologies and application to hydraulicmachinery [Ph.D. thesis], Ecole Polytechnique Federale deLausanne, 2005.

[10] ANSYS, ANSYS CFX-Solver Theory Guide, Release 12.0, 2009.

[11] H. K. Versteeg and W. Malalasekera, An Introduction toComputational Fluid Dynamics, the Finite Volume Method,Pearson Education Limited, 2nd edition, 2007.

[12] J. Blazek, Computational Fluid Dynamics: Principles andApplications, Elsevier, New York, NY, USA, 2nd edition, 2005.

[13] C. E. Brennen, Fundamentals of Multiphase Flows, CambridgeUniversity Press, New York, NY, USA, 2005.

[14] http://www.ansys.com.

[15] http://www.simerics.com.

[16] T. Huuva, Large eddy simulation of cavitating and non-cavitating flow [Ph.D. thesis], Department of Shipping andMarine Technology, Chalmers University of Technology,Gothenburg, Sweden, 2008.

[17] C. L. Merkle, J. Feng, and P. E. O. Buelow, “Computationalmodeling of the dynamics of sheet cavitation,” in Proceedingsof the 3rd International Symposium on Cavitation, vol. 2,Grenoble, France, 1998.

[18] http://www.openfoam.com.

[19] ESTECO, modeFRONTIER 4 User Manual, 2009.

[20] A. Clarich, G. Mosetti, V. Pediroda, and C. Poloni, “Applica-tion of evolutionary algorithms and statistical analysis in thenumerical optimization of an axial compressor,” InternationalJournal of Rotating Machinery, vol. 2, pp. 143–151, 2005.

[21] Y. T. Shen and P E. Dimotakis, “The influence of surfacecavitation on hydrodynamic forces,” in Proceedings of the 22ndAmerican Towing Tank Conference (ATTC ’89), pp. 44–453, St.Johns, Canada, 1989.

[22] S. S. Rao, Engineering Optimization—Theory and Practice,John Wiley & Sons, New York, NY, USA, 1996.

[23] F. Salvatore H. Streckwall and T. van Terwisga, “Propellercavitation modelling by CFD-results from the VIRTUE 2008,”in Proceedings of the 1st International Symposium on Cavitation(SMP ’09), pp. 362–371, Trondheim, Norway, June 2009.

[24] G. N. V. B. Vaz, Modelling of sheet cavitation on hydrofoilsand marine propellers using boundary element methods [Ph.D.thesis], Universidade Tecnica de Lisboa, 2005.

[25] R. E. Bensow and G. Bark, “Implicit LES predictions of thecavitating flow on a propeller,” Journal of Fluids Engineering,vol. 132, no. 4, Article ID 041302, 2010.

[26] http://www.marinepropulsors.com.[27] http://www.sva-potsdam.de/pptc.

[28] U. H. Barkmann, “Potsdam propeller test case (PPTC), openwater tests with the model propeller,” Tech. Rep. VP1304, SVA(Potsdam Model Basin), 2011.

[29] K. P. Mach, “Potsdam propeller test case (PPTC), LDV velocitymeasurements with the model propeller,” Tech. Rep. VP1304,SVA (Potsdam Model Basin), 2011.

[30] F. R. Menter, “Two-equation eddy-viscosity turbulence modelsfor engineering applications,” AIAA Journal, vol. 32, no. 8, pp.1598–1605, 1994.

[31] N. Berchiche and C.E. Janson, “Grid influence on the propelleropen-water performance and flow field,” Ship TechnologyResearch, vol. 55, pp. 87–96, 2008.

[32] M. Abdel-Maksoud, F. Menter, and H. Wuttke, “Viscousflow simulations for conventional and high-skew marinepropellers,” Ship Technology Research, vol. 45, no. 2, pp. 64–71,1998.

[33] M. Morgut and E. Nobile, “Influence of grid type and tur-bulence model on the numerical prediction of the flowaround marine propellers working in uniform inflow,” OceanEngineering, vol. 42, pp. 26–34, 2012.

[34] M. Morgut and E. Nobile, “Numerical predictions of thecavitating and non-cavitating flow around the model scalepropeller pptc,” in Proceedings of the Workshop on Cavitationand Propeller Performance, the 2nd International Symposium onMarine Propulsors (SMP ’11), Hamburg, Germany, 2011.

[35] F. Salvatore, Private communication, 2010.

Page 12: NumericalPredictionsofCavitatingFlowaroundModelScale ... · 2019. 7. 31. · PPTC propeller. The numerical results were compared with the avail-able experimental data. For both propellers,

International Journal of

AerospaceEngineeringHindawi Publishing Corporationhttp://www.hindawi.com Volume 2010

RoboticsJournal of

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Active and Passive Electronic Components

Control Scienceand Engineering

Journal of

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

International Journal of

RotatingMachinery

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Journal ofEngineeringVolume 2014

Submit your manuscripts athttp://www.hindawi.com

VLSI Design

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Shock and Vibration

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Civil EngineeringAdvances in

Acoustics and VibrationAdvances in

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Electrical and Computer Engineering

Journal of

Advances inOptoElectronics

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

The Scientific World JournalHindawi Publishing Corporation http://www.hindawi.com Volume 2014

SensorsJournal of

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Modelling & Simulation in EngineeringHindawi Publishing Corporation http://www.hindawi.com Volume 2014

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Chemical EngineeringInternational Journal of Antennas and

Propagation

International Journal of

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

Navigation and Observation

International Journal of

Hindawi Publishing Corporationhttp://www.hindawi.com Volume 2014

DistributedSensor Networks

International Journal of