You are on page 1of 39

Ocean Engineering 157 (2018) 262–300

Contents lists available at ScienceDirect

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

Effect of passive flow control on the aerodynamic performance, entropy


generation and aeroacoustic noise of axial turbines for wave
energy extractor
Ahmed S. Shehata a, b, *, Qing Xiao a, Mohamed A. Kotb b, Mohamed M. Selim b, c, A.H. Elbatran b,
Day Alexander a
a
Department of Naval Architecture, Ocean and Marine Engineering, University of Strathclyde, Glasgow G4 0LZ, UK
b
College of Engineering and Technology, Arab Academy for Science Technology and Maritime Transport, P.O. 1029, AbuQir, Alexandria, Egypt
c
Material Science and Engineering, University of Alabama at Birmingham, Birmingham, AL, USA

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

Keywords: Wells turbine is the simplest type of an axial flow self-rectifying air turbine that can be used in conjunction with
Oscillating flow Oscillating Water Column (OWC) system in the extraction of ocean wave energy. It has been noticed that this
Wave turbine turbine is subjected to early stall. As a consequence, several attempts for improving the energy extraction per-
Passive flow control formance of Wells turbine within the stall regime have been investigated. One of these attempts was using an
Entropy generation inclined slot as a passive flow control to obtain a delayed stall. In the following study, the impact of varying the
Aeroacoustic noise
angle for the slot on the performance of Wells turbine in the stall regime was investigated. Furthermore, the first
Egyptian coasts
law of thermodynamics and the entropy analysis has been used to examine the effect of the slot angle on the
entropy generation features around the turbine blade. Moreover, Investigation of slot angle effect on the aero-
dynamics noise emission from Wells turbine airfoil during the normal operation and the stall regime is covered in
this study. The blade of turbine with optimum angle of slot was investigated using the OWC based on actual data
from the Egyptian Northern Coast. It was found that the optimum slot angle is 10 clockwise which results in 3%
improvement in the torque coefficient before the stall and 15% after the stall as compared to the 0 slot.
Otherwise, it gives a lower global entropy generation rate than the 0 slot by 4% before the stall and 3% after the
stall. Furthermore, using airfoil of blade turbine with a slot resulted in a reduction of aeroacoustic noise by
21.2% at the stall regime under oscillating flow conditions.

1. Introduction Suzuki and Tagori, 1985; Yukihisa Washio et al., 1985; Folley et al.,
2006) which is a remarkable subject that needs to be explored and
The main challenge facing the wave energy extraction devices with improved. In case of large angles of attack and that means large flow
oscillating water column system is to find an economical and efficient rates, the separation area will be increased on the blade due to the
means of converting bidirectional flow from the waves motion to uni- separate boundary layers, which decrease the value of torque coefficient
directional rotary motion for driving electrical generators (Falcao and and thus the efficiency. The reason behind that, is the decrease in the lift
P.A.P., 1999; Falcao, 2004; Torres et al., 2016), as seen in Fig. 1 A. A value and also a big increase in the drag value. From references (Shehata
self-rectifying air turbine such as Wells turbine, see Fig. 1 B (Shaaban, et al., 2017a; Okuhara et al., 2013; Mamun, 2006) it can be concluded
2012, 2016; Kinoue et al., 2004; Setoguchi and Takao, 2006; Torres et al., that Wells turbine can produce energy at relatively low air flow rate,
2018; Halder et al., 2017a), can be used to extract the energy from the when other types of turbines con not (Okuhara et al., 2013; Setoguchi
oscillating air column (Shehata et al, 2017a, 2017b; Boccotti, 2007a, et al., 2001b; Liu et al., 2016). Furthermore, the aerodynamic efficiency
2007b; Halder et al., 2017b; Scarpetta et al., 2017). Wells turbine suffers increases with the increase of the angle of attack or the flow coefficient
low performance problem at stall condition (Shehata et al., 2017a; Wang up to an appointed value, after that it decreases. Thus, most of the pre-
et al., 2012; Hitoshi Hotta, 1985; Masahiro Inoue et al., 1985; Masami vious studies (Shehata et al., 2017a; Shehata, 2017) aimed to improve

* Corresponding author. Department of Naval Architecture, Ocean and Marine Engineering, University of Strathclyde, Glasgow G4 0LZ, UK.
E-mail address: a_samir@aast.edu (A.S. Shehata).

https://doi.org/10.1016/j.oceaneng.2018.03.053
Received 12 December 2017; Received in revised form 28 February 2018; Accepted 20 March 2018

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


A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Nomenclature and abbreviations To Reservoir temperature (K)


tsin Time period for sinusoidal wave ¼ 1=f (second)
A Total blade area (m2)
ui Reynolds Averaged velocity component in i direction (m/s)
C Blade chord (m)
V Volume of a computation cell (m3)
CD Drag force coefficient
Va Instantaneous Velocity (m/s)
CL Lift force coefficient
Vam highest speed of axial direction (m/s)
CT Torque coefficient
Vo Initial velocity for computation (m/s)
D Fluid domain 
Dss Suction slot diameter (m) W Net-work transfer rate (W)

f The cycle frequency (Hz) W rev Reversible work (W)
FD In-line force acting on cylinder per unit length (gf) αSS Angle of suction slot (Degree)
G Filter function ηF Efficiency in first law of thermodynamics
Ke Specific kinetic energy (J/kg) ηS Second law efficiency
Lss Suction slot location from leading edge in chord percentage μ Viscosity (Kg/m.s)
% μt Turbulent viscosity (N.s/m2)
K Turbulent kinetic energy (Joules) ρ Density (Kg/m3)
Δp Pressure difference across the turbine (N/m2) ∅ Flow coefficient
Rm Mean rotor radius (m) ω Rotor angular speed (rad/sec)
Sgen Local entropy generation rate (W/m2K) ð  ρu0 i u0 j Þ Reynolds stress tensor
SG Global entropy generation rate (W/K) CFD Computational Fluid Dynamics
Sij Mean strain rate (1/s) NACA National Advisory Committee for Aeronautics
St Thermal entropy generation rate (W/m2K) OWC Oscillating Water Column
SV Viscous entropy generation rate (W/m2K)

Fig. 1. OWC energy converters A) An illustration of the principle of operation of OWC system, where the wave motion is used to drive a turbine through the oscillation
of air column B) Typical structure of Wells turbine rotor.

the torque coefficient (the turbine output) and improve the turbine using numerical computer simulation (quasi-steady analysis), it was
performance during the stall condition. observed that the three dimensional case has superior characteristics in
The following references (Raghunathan, 1995a; Dixon, 1998; Shel- the starting and running characteristics. At high pressure values, usually
dahl and Klimas, 1981) noted that the delay of stall beginning contributes a multi plane Wells turbine systems are used. Such configuration avoids
to improved Wells turbine performance. Employing guide vanes on the the need for guide vanes and, therefore the turbine would require less
rotor's hub can achieve this delay (Raghunathan, 1995a, 1995b; Brito-- maintenance and repairs (Raghunathan, 1995a). The performance of a
Melo et al., 2002). It was found that adding guide vanes to a multi-plane biplane Wells turbine is based on the gap between the two turbines as
turbine increases the efficiency by approximately 20% compared to the described in (Raghunathan, 1995a). It was recommended that a
one without guide vanes. Setoguchi et al. (2001a) carried out a com- gap-to-chord ratio between the two turbines was 1.0. References
parison between Wells turbines having two and three dimensional guide (Thakker and Abdulhadi, 2007, 2008) investigated experimentally the
vanes. By performing a steady flow testing of Wells turbine model, and effects of rotor solidity and blade profile on hysteretic behavior of Wells

263
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 2. Factors affecting determination of the slot velocity direction for the pressure distribution and velocity vector direction A) accelerating flow at compression
cycle B) decelerating flow at suction cycle.

turbine operating under bi-directional airflow. It is noticed that reducing Exergy analysis has been performed in (Koroneos et al., 2003) using
rotor solidity leads to decreasing the size of hysteretic loop of pressure the CFD for biplane Wells turbines (Shaaban, 2012; Mohamed et al.,
coefficient. In order to investigate the relation between the hysteretic 2011) in steady state where the rotor of the upstream flow has a design
behavior of Wells turbine and the rotor solidity, setting angles, and blade point of 0.82 s law efficiency, however the downstream rotor has second
thickness, Computational Fluid Dynamics (CFD) simulations were pro- law efficiency equals 0.61. The entropy generation rate, due to viscous
ceed for the flow field around a single blade of a Wells turbine by the dissipation, around different two dimension airfoil sections for Wells
following references (Kim et al., 2002a; Setoguchi et al., 2003a). turbine blade was studied in (Shehata et al, 2014, 2016). It was observed
Therefore, Torresi et al. (2008) studied in details the flow-field charac- that, when Reynolds number was increased from 6  104 till 1  105 the
teristics through a high solidity (σ ¼ 0.64) Wells turbine using blades total entropy generation rate increased accordingly more than double for
with constant chord NACA0015 profiles. For high flow rate values, a tested airfoils. Although, Reynolds number was increased further to
radical shift of the mass flow through the turbine due to the cascade 2  105, the total entropy generation rate showed values ranging from
effect (Raghunathan et al., 1981) occurs. This leads to flow separation at 0.25 reduction to 0.2 increase compared to corresponding values at
the outer radii. Reynolds number of 1  105. For the four different airfoils, the efficiency
The estimation of blade sweeps for the Wells turbine have been in compression cycle is higher than that in suction cycle at angle of attack
conducted numerically by (Kim et al., 2002b) and experimentally with equals to 2 . On contrast, the efficiency for suction cycle was more than
quasi-steady analysis by (Setoguchi et al., 2003b). As a result, it was the compression cycle with the increase of angle of attack. The re-
concluded that the Wells turbines performance was affected by the blade searchers in references (Shehata et al, 2014, 2016) suggested that a
sweep area. Paderi M. et al. (Paderi and Puddu, 2013), carried out the possible existence of critical Reynolds number at which viscous entropy
experimental characterization of Wells turbine with NACA0015 profiles takes least values. A comparison of the entropy generation characteristics
submitted to a bi-directional flow, the results also are presented and between a Wells turbine with a variable chord design and another one
analyzed the study mentioned in (Paderi and Puddu, 2013). For various with a constant chord was investigated in (Soltanmohamadi and Lakzian,
test conditions, the maximum efficiency of the turbine (from 30% to 2015). The detailed results demonstrated that static pressure difference
43%) takes place for the flow coefficient values between 0.19 and 0.21. A around new blade is increased and that the value of entropy generation
modified Wells turbine with setting angle has been examined by proto- throughout the total running range is decreased also by average 26.02%.
type testing and numerical modeling in (Setoguchi et al., 2003c). It was Second law analysis of a Wells turbine was numerically performed using
proven that the modified turbine using blades with a specific setting the Open FOAM in (Mahboubidoust and Ramiar, 2017) to express its
angle is superior to the normal Wells turbine, and that the optimum optimal performance and show how irreversibility factors lead to exergy
setting angle was 2 for compression velocity amplitude to suction ve- destruction and also second law efficiency reduction. The dielectric
locity ratios of 0.6 and 0.8, in both with guide vanes and without guide barrier discharge (DBD) plasma actuator is applied in the trailing edge of
vanes configurations. the turbine blade in order to improve performance and reduce

264
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

condition. At most, a 63.37% increase in torque coefficient and 72.8%


increase in efficiency are achieved by the variable thickness blade in the
deep stall condition. The majority of the researchers, who investigated
the performance of various airfoil designs and various operational con-
ditions, were analyzing the problem based only on the first law of ther-
modynamic parameters. However, it is essential to take into account the
second law of thermodynamic in order to form a thorough understand-
ing, since it was proven to show highly hopeful results in many similar
systems, such as gas turbine in (Alklaibi et al., 2016; Yucer, 2016), and
wind turbine in (Pope et al., 2010; Baskut et al., 2010, 2011; Redha et al.,
2011).
Techniques were developed to manipulate the boundary layer, either to
increase the lift or decrease the drag, they are classified under the general
heading of boundary layer control or flow control (Shehata et al., 2017c;
Ismail and Vijayaraghavan, 2015). Methods of flow control to achieve
separation postponement, lift enhancement and drag reduction have been
considered in these studies. Such studies have demonstrated that adding
suction slot could modify the pressure distribution over an airfoil surface
and have an essential impact on both the lift and drag coefficients (Yousefi
et al., 2014; Chapin and Benard, 2015; Schatz et al., 2007; Chawla et al.,
2014; Fernandez et al., 2013; Volino et al., 2011). Also, many studies have
been conducted to draw on flow control techniques. Prandtl, 1904
(Schlichting, 1968), was the first scientist who used boundary layer suction
on a cylindrical surface to delay the separation of boundary layer. The oldest
recognized experimental studies on boundary layer suction for wings were
conducted in the late 1930 and the 1940 (Richards and Burge, 1943; Walker
and Raymer, 1946; Braslow, 1999). Huang et al. (2004) investigated both
suction and blowing flow control mechanism on an airfoil with NACA 0012
section profile. When jet location and angle of attack were combined,
perpendicular suction at the leading edge increased lift coefficient more
than other suction situations. Furthermore, the downstream locations with
tangential blowing have the maximum increase rate in the lift coefficient.
Rosas in (Rosas, 2005) was numerically investigated flow separation con-
trol through oscillatory fluid injection, in which lift coefficient was
increased. The authors in (Akcayoz and Tuncer, 2009) studied the optimi-
zation of synthetic jet parameters on a NACA 0015 airfoil to increase the lift
and decrease the drag in different angles of attack. It can be concluded that
the optimum location for the jet moved toward the leading edge, in addition
the optimum angle for the jet increased with the increment of angle of
attack. The CFD method has been widely used to study the technique of
boundary layer control. Many flow control studies by CFD approaches (Kim
and Kim, 2009; Rumsey and Nishino, 2011; Yagiz et al., 2012) have been
conducted to study the blowing and suction jets effects on the aerodynamic
Fig. 3. Boundary conditions A) The sinusoidal wave boundary condition, which
represents a regular oscillating water column. B) Dimensions of whole compu- performance of blade or wing airfoils.
tational domain and location of airfoil. C) The near views of slot mesh. Aiming at minimizing the overall environmental impact of oscillating
water column technology requires a new effort to reduce the aeroacoustic
noise associated with Wells turbine operation (Starzmann and Carolus,
2013a). A new blade design methodology for a Wells turbine with
aerodynamic load. It can be noted that by applying plasma the average skewed blades is investigated by (Starzmann and Carolus, 2013a, 2013b)
increase in torque coefficient was about 39.36% and the average to be very successful for reducing noise and vibrations. The flow gener-
decrease in lift coefficient was about 30.53%. It was also found that the ated sound in normal operation (unstalled) was decreased up to 3 dB by
plasma be applied in the flow coefficients almost equal to ones that cause optimal backward/forward blade skew. The vast majority of the re-
the stall, in which case 39.56% increment is observed in the first law searchers and investigations focused on the aerodynamic noise predic-
efficiency. However, the second law efficiency increases about 39.16% tion for wind turbine (Maizi et al., 2018; Botha et al., 2017; Wasala et al.,
without considering viscous dissipation term and decreases about 2015; Shaltout et al., 2015; Kaviani and Nejat, 2017; Solís-Gallego et al.,
64.63% with considering the viscous dissipation term. The effects of 2018) and around an airfoil (Rumpfkeil, 2017; Gea-Aguilera et al., 2017;
blade thickness on the performance of a Wells turbine are investigated Wang et al., 2017; Avallone et al., 2017; Shen et al., 2017; Miotto et al.,
based on aerodynamic and entropy generation analysis in (Nazeryan and 2017; Siozos-Rousoulis et al., 2017). However, no study until now study
Lakzian, 2018). Two kinds of blade profiles are being investigated, a the aeroacoustic noise prediction or reduction of Wells turbine at stall
constant thickness blade and a variable thickness blade. The computation and near-stall conditions under oscillating flow conditions, even, the only
is performed by solving the 3D steady incompressible Reynolds-averaged research group for the aerodynamic noise prediction of Wells turbine
Navier–Stokes equations. The results show the interaction between tip mentioned be in (Starzmann and Carolus, 2013a, 2013b), assumed that
leakage vortex and suction surface of the blade is substantially reduced Wells turbine work under unidirectional flow with steady state condition.
by using the variable thickness blade. The results reveal that entropy One of the best locations for applying the Wells turbine energy
generation seems to give an advantageous effect of reducing the sepa- extractor as OWC system is the Egyptian Northern Coast. The Egyptian
ration at the tip section of the variable thickness blade in the deep stall Coast is considered the most energetic coast of the Southern

265
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 4. The validation results A) Measured torque coefficient from reference Torresi (2007b) and calculated torque coefficient from present CFD. B) Measured un-
steady in-line force FD from reference (Nomura et al., 2003), (angle of attack ¼ 0 ) and FD calculated from the present CFD.

Mediterranean Basin, lying between the Nile Delta and the Libyan bor- and other locations. Subsequently, the potential wave energy can be
ders with a potential of above 6.8 kW/m wave power in winter season revealed and exploited (Zodiatis et al., 2014). Otherwise, sea states with
and 3.4 kW/m in summer season (Mørk et al., 2010; Barstow et al., 2009) the wave heights more than 5 m are not very effective for the annual
and the wave energy of about 36003 kWh/m. The most active sea states power (Ayat, 2013).
have wave heights ranging from 1 to 4 m and wave periods between 4 The objective of the present work is to demonstrate the performance
and 8 s. The sea wave in Egypt is comparatively low but on the other hand of Wells turbine at stall and near-stall conditions which can be effectively
stable and that are the significant differences between the sea in Egypt improved by using passive flow control technique such as blowing and

Table 1
The error percentage between measured torque coefficient from reference Torresi (2007b) and calculated torque coefficient from CFD under unsteady flow with
non-oscillating velocity.
Torque Coefficient Angle of attack (Degree)

8.7 10.1 10.6 11.3 11.7 12.304 13.6 14.4

Experimental 0.04881 0.06305 0.07119 0.08068 0.08746 0.0922 0.08136 0.07254


CFD 0.050918 0.066887 0.07264 0.07928 0.08564 0.09103 0.08304 0.067587
Error % 4 6 2 2 2 1 2 7

266
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 2
The error percentage between measured FD from reference (Nomura et al., 2003) and calculated FD from CFD under unsteady flow with sinusoidal inlet velocity.
FD (gf) Time (Second)

14.02 14.1 14.12 14.2 14.3 14.34 14.4 14.5 14.6 14.7 14.8 14.9 15

Frequency 2 Hz
Experimental 3.4 7.7 9.9 14.4 12.9 3.4 4.11 2.3 7.5 14.7 10.7 3.9 2.7
CFD 3.8 7.8 9.8 14.5 12.5 3.5 3.4 2.7 7.8 14.9 10.9 3.8 2.4
Error % 11 1 1 1 4 1 17 17 4 1 2 2 11
Frequency 1 Hz
Experimental 4.5 6.9 12.6 14.1 14.3 13 10.2 7.8 4.7 2.8 2.3 2.5 3
CFD 4.6 7.2 12.6 13.1 14.3 13.2 10.3 8.6 4.5 2.7 2.2 2.6 3.3
Error % 2 4 0 7 0 1 1 10 4 1 4 4 10

The finite-volume, in FLUENT, discretization itself implicitly provides


the filtering operation (Mamun, 2006):

1  
ϕðxÞ ¼ ∫ ϕ x' dx' ; x' 2 V (2)
V V

Where the volume of a computational cell is define by V. G (x, x'), the


filter function, implied here is by

  1=V for x' 2 V
G x; x' ¼ (3)
0 otherwise

The LES model will be applied to essentially incompressible flows.


Filtering the incompressible Navier-Stokes equations, one obtains
Fig. 5. Airfoil diagram with suction slot has angle. (Dahlstrom, 2003)

∂ρ ∂ρui
þ ¼0 (4)
suction slot with different angles. This is achieved by conducting CFD
∂t ∂xi
based first and second law analysis for the Wells turbine airfoil, without  
∂ui ∂ui   ∂ ∂u ∂ρ ∂τij
and with slot which has different angles, under oscillating and non- þ uj ¼ μ i  þ (5)
oscillating flow conditions. A typical slot is attached to the airfoil pro- ∂t ∂xj ∂xj ∂xj ρ ∂x i ρ ∂xj
file section, normal to the chord length, and due to the pressure differ-
Where, τij is the sub-grid-scale stress equals to:
ence between the two surfaces, a suction effect occurs which delays the
stall. Hence, there is no need to generate any specific active suction or τij ¼ ρui uj  ρui uj (6)
blowing within the airfoil or the slot. During the compression cycle, this
slot pulls the flow from the high pressure area (lower surface) and blows The sub-grid-scale stresses producing from the filtering process are
it to the low pressure area (upper surface). On the other hand, during the unknown and need to modeling. The most of sub-grid-scale models are
suction cycle, the slot draws the flow from the upper surface (high eddy viscosity models of the next form (Moin et al., 1991):
pressure) and blows it to the lower surface (low pressure) (Shehata et al.,
1
2017d), see Fig. 2A) and B). The slot will be known as a suction slot in the τij  τkk σ ij ¼ 2μt Sij (7)
3
results which were presented henceforward. Moreover, another objective
of this study is assessing the effect of slot angle on aerodynamic noise Where Sij is the rate-of-strain tensor for the resolved scale and it is
emitted from Wells turbine airfoil under oscillating and non-oscillating
defined by:
flow conditions. According to the literature, this is fairly a first study to
 
use the slot with different angles in Oscillating Water System and Wells 1 ∂ui ∂uj
turbine design. Furthermore, this is also the first study to investigate the Sij ¼ þ (8)
2 ∂xj ∂xi
effect of a slot with different angles in aerodynamic performance, entropy
generation and aeroacoustic noise of Wells turbine airfoil at the stall and and μt represents the sub-grid-scale turbulent viscosity, which the
near-stall conditions under oscillating flow conditions. Smagorinsky-Lilly model is applied for it (DK., 1992). The majority of
sub-grid-scale models for Smagorinsky-Lilly model was presented by
2. Mathematical formulations and numerical methodology Smagorinsky (Hinze, 1975) and was further developed by Lilly (Launder
and Spalding, 1972). In the Smagorinsky-Lilly model, the eddy viscosity
The Large Eddy Simulation (LES) are used by the governing equations is modeled by:
to obtain by filtering the time-dependent Navier-Stokes equations. The  
filtering procedure filters out eddies whose scales are smaller than the μt ¼ ρL2s S (9)
filter width or grid spacing used in the computations. The resulting
equations thus decide the large eddies dynamics. A filtered variable, Where, the mixing length for sub-grid-scale models is define by Ls and
qffiffiffiffiffiffiffiffiffiffiffiffi
 
which denoted by an over-bar, is defined by (SB., 2000): S ¼ 2Sij Sij . The Ls iscomputed using:
   
ϕ ðxÞ ¼ ∫ FD ϕ x' G x; x' dx' (1)  
Ls ¼ min kd; Cs V 1=3 (10)
Where FD is the fluid domain, and the filter function which defines the
Where k ¼ 0:42, d is the distance to the closest wall, Cs is the Smagor-
scale of the resolved eddies is represented by G.

267
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 3
The suction slot with different position angles at NACA0015.

insky constant, and V is the volume of the computational cell. Lilly Va


α ¼ tan1 (13)
derived a value of 0.23 for Cs from homogeneous isotropic turbulence. ω Rm
However, this value was found to cause excessive damping of large-scale
fluctuations in the transitional flows or in the existence of mean shear. and the torque is defined as:
Due to the turbulence flow at all domain, a dynamic SGS model was not 1  2 
necessary in the LES models. Therefore, Cs ¼ 0.1 has been found to yield Torque ¼ ρ Va þ ðωRm Þ2 ARm CT (14)
2
the superior results for a vast range of flows (Kinoue et al., 2004; Mamun,
2006; Mamun et al., 2004). The first law of thermodynamics efficiency (ηF ) is determined by:
The drag and lift coefficients CD and CL are determined from FLUENT
Torque* ω
(post processing software). For each angle of attack, the average value for ηF ¼ (15)
ΔP*Q
drag and lift coefficients was used to determine one value for torque
coefficient. Thereafter, the torque coefficient can then be expressed as The transport equations that govern these models can be found in
(Sheldahl and Klimas, 1981; Curran et al., 1998; Whittaker et al., 1997): turbulent flow texts such as (Hirsch, 2007). The second law of thermo-

dynamic defines the network transfer rate W as (Bejan, 1996):
CT ¼ ðCL sin α  CD cosαÞ (11)
 
The flow coefficient ϕ based on both axial and tangential velocties of W rev  W ¼ To Sgen (16)
the rotor is given by
Which, it has been known as the Gouy–Stodola theorem (Stodola, 1910).
Va In the absence of chemical reactions and phase change, the irrevers-
ϕ¼ (12)
ω*Rm ible entropy generation could be expressed in terms of the derivatives of
local flow parameters. In viscous flows, there are two main dissipative
Where the α angle of attack equal to
mechanisms:1) the strain-originated dissipation which corresponds to

268
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 4
The value of improvement in torque coefficient for suction slot with different position angles under non-oscillating ve-
locity.

the viscous entropy generation and 2) the thermal dissipation which


Where Ke ¼ 12V 2
correspond to the thermal entropy generation (Iandoli, 2005). Thus, it
From the above equations, several conclusion can be noted. Firstly,
can be written as,
the torque coefficient (CT ) relates to the first law efficiency (ηF ). Sec-
Sgen ¼ SV þ Sth (17) ondly, the increase in torque coefficient (CT ) leads to increase in the first
law efficiency (ηF ). Thirdly, the global entropy generation rate (SG ) re-
For the present case study (isothermal incompressible flow), the lates to the second law and efficiency (ηS ). Finally, the decrease in the
thermal dissipation term is insignificant. Therefore, the local viscous ir- global entropy generation rate (SG ) leads to an increase in the second law
reversibilities could be stated as: efficiency (ηS ).
μ Regarding acoustic simulation, the Lighthill formulation (Lighthill,
SV ¼ φ (18) 1952, 1954) for the aeroacoustic noise was the first approach, where the
To
Navier–Stokes equations were rearranged into an inhomogeneous wave
where φ is the viscous dissipation term, which could be expressed in two equation, shown in equation (23).
dimensional Cartesian Coordinates as (Iandoli, 2005):
∂ 2 P' ∂2 Qij
"   2 #    r 2 P' ¼ (23)
∂u 2 ∂v ∂u ∂v 2 C 0 ∂t
2 2
∂xi ∂xj
φ¼2 þ þ þ (19)
∂x ∂y ∂y ∂x
Where c0 is the ambient speed of the sound, P' ¼ C02 ρ and Qij is Lighthill's
Equations (18) and (19) were used to create the user defined function stress tensor. The left hand side describes the propagation of the acoustic
(UDF) file, which is used to calculate the local entropy generation form wave in both the spatial and temporal domains, otherwise, the right hand
the software (FLUENT). After that, the global entropy generation rate is side represents the source term.
expressed as: Ffowcs-Williams and Hawkings (1969) modified and extended
Lighthill's equation to treat the problem of sound generated by a body in
SG ¼ ∬ SV dydx (20) arbitrary motion in both frequency and time domains. The
xy
Ffowcs-Williams and Hawkings (FW–H) equation has two extra source
SG is calculated by integral the global value using the FLUENT software. terms which represent monopole and dipole sources, shown in equation
The Exergy is giving as, (Bejan, 1995): (24):

Exergy ¼ Ke þ SG (21) ∂ 2 P' ∂2




 r 2 P' ¼ Qij Hðf Þ  Pij nj þ ρui ðun  vn Þ δðf Þ þ f
C 0 ∂t
2 2
∂xi ∂xj ∂xi ∂t
and efficiency of the second law can be written as (Pope et al., 2010):  ½ρ0 vn þ ρðun  vn Þδðf Þg
Ke (24)
ηS ¼ (22)
Exergy
and

269
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 6. The effect of suction slot with different angle at the stall angle (13.6 Degree) and f ¼ 0.167 Hz A) Torque coefficient. B) The global entropy generation rate.

" # " #
Qij ¼ ρui uj þ Pij  C02 ðρ  ρ0 Þδij (25) 1 Lr Lr  L
4π P'L ðx; tÞ ¼ ∫ ds þ ∫ ds
c0 f ¼0 rð1  Mr Þ2 f ¼0 r ð1  Mr Þ
2 2

Where, Pij is the compressive stress tensor and equal to P' δij , ui is the fluid "
ret
#
ret

velocity in the i direction, vi is the surface velocity in the i direction, ni is 1 Lr frMr þ c0 ðMr  M 2 Þg
þ ∫ ds (27)
the unit vector in the i direction, vn is the surface velocity normal to the f ¼0 c0 r2 ð1  Mr Þ3 ret
surface ðf ¼ 0Þ, un is the fluid velocity normal to the surface ðf ¼ 0Þ, Hðf Þ
is the Heaviside delta function, δðf Þ is the Dirac delta function, f is the M is local Mach number, Mr is the Mach number of a point on the moving
mathematical surface of the moving body and P' is the acoustic pressure surface, r is the distance to the observer. The subscript ret denotes that the
at the far field and equal to P'T þ P'L . The solution at a receiver location integrals are computed at the corresponding retarded times, the dot
can be calculated analytically using generalized function theory and the above a variable represents the source time derivative of that variable.
free space Green's function. The Sound Pressure Level (SPL) in dB, see equation (28), is obtained
The thickness contribution P'T is given by (Wasala et al., 2015; Wil- with respect to the reference acoustic pressure Pref ¼ 2  105 Pa
liams and Hawkings, 1969),   
SPL ¼ 20 log10 P' Pref (28)
2 3
ρ0 u_n þ un
4π P'T ðx; tÞ ¼ ∫ 4 2
5 ds 3. CFD verification and validation result
f ¼0 rð1  Mr Þ
ret
" # In this study, The CFD verification and validation results for the
ρ0 un frMr þ c0 ðMr  M 2 Þg
þ ∫ ds (26) mentioned models are presented. Furthermore, this section proposes a
f ¼0 r2 ð1  Mr Þ3 ret detailed description of the turbulence models used in stall condition, an
expression of the discretization technique used and the boundary con-
The loading contribution P'L is given by (Kaviani and Nejat, 2017;
ditions for the current work.
Williams and Hawkings, 1969),

270
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 5
The value of global entropy generation rate for suction slot with different position angles under non-oscillating velocity.

3.1. Computational model and boundary conditions data from the site (Mørk et al., 2010; Barstow et al., 2009). Reynolds
numbers up to 2  104 are created due to the sinusoidal wave condition
Two-dimensional numerical simulation models for NACA 0015 air- based on the reference (Torresi et al., 2009). The dimensions of the whole
foils section were built and then validated by experimental tests under computational domain, location of airfoil and near views of slot mesh are
unsteady flow conditions with constant velocity, as well as, under un- shown in Fig. 3B) and C).
steady flow with sinusoidal inlet velocity. GAMBIT code was used for
discretizing the computational domain to Cartesian structured finite 3.2. Grid sensitivity test (verification)
volume cells. In ANSYS FLUENT, the Green-Gauss cell based evaluation
method is used for calculating the gradient terms, which could be applied In order to ensure that the numerical results are minimally dependent
to the application of such boundary condition types (Starzmann and on the grid size, several grids had been tested to determine the grid cells
Carolus, 2013b; Mamun et al., 2004; Mohamed and Shaaban, 2013; required number to create a dependable model. Four grids with different
Torresi et al., 2009). Different interpolation schemes were tested to mesh sizes ranging from 112603 up to 446889 cells are studied for per-
transfer cell center values to cell face values for the variables of flow field. forming grid sensitivity test. It can be noticed that the grid with mesh size
In addition, several convergence tests have been performed. The second of 312951 cells gives good results within reasonable computation time
order upwind (Smagorinsky, 1963) interpolation scheme was used in this (more details about this result in (Shehata et al., 2014; Shehata et al.,
work. Where, there was no noticeable difference in the results when the 2017d)). Thus, it was chosen to conduct the investigative and analysis of
third order MUSCL scheme was used in the current cases. More than that, the present work.
the third order MUSCL scheme resulted in high oscillatory residuals
during the solution, in some cases.
3.3. Validation of the CFD model
The axial inlet flow velocity of Wells turbine boundary condition is
modeled as a sinusoidal wave in current study. Hence, time-dependent
Large Eddy Simulation (LES) model was used to model the flow
inlet boundary conditions were created. The following function was
around NACA0015 airfoil to give the better agreement results with
used to apply the inlet boundary condition (see Fig. 3 A).
experimental data taken from (Torresi et al., 2009) and (Nomura et al.,
Va ¼ Vo þ Vam sinð2π ftsin Þ (29) 2003). According to the literature survey in (Dahlstrom, 2003; Kawai and
Asada, 2013; Richez et al., 2007; Alferez et al., 2013; Kim et al., 2015;
Where a time period, tsin , of 6 s was used in this study based on the AlMutairi et al., 2015; Armenio et al., 2010; Hitiwadi et al., 2013;
conducted literature survey (Kinoue et al, 2003, 2004; Setoguchi et al., Bromby, 2012), LES model gave excellent results when they are used to
2003a; Mamun et al., 2004). Time step is set as 0.0089 s in order to satisfy simulate the airfoil in stall condition.
CFL (Courant Friedrichs Lewy) (De Moura and Carlos, 2013) condition Although, LES is definitely 3D model, there have been numerous
equal to 1. For the Egyptian coasts boundary condition the tsin is equal to successful endeavors to be applied in 2D engineering applications (e.g.,
4, 6 and 8 s (f equal to 0.25,0.167 and 0.125 Hz), considering the real flow over obstacles (Skyllingstad and Wijesekera, 2004), hump (Avdis
et al., 2009), block (Cheng and Porte-Agel, 2013), airfoils (Hitiwadi et al.,

271
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 7. Hysteretic behavior comparisons between the optimum angles of suction slot with sinusoidal inlet velocity and f ¼ 0.167 Hz A) 55 anticlockwise B) 65
anticlockwise C) 35 clockwise.

272
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 8. Hysteretic behavior comparisons between the optimum angles of suction slot with sinusoidal inlet velocity and f ¼ 0.167 Hz A) 15 anticlockwise B)
10 clockwise.

273
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 9. Comparison between the instantaneous torque coefficients for the optimum angles of suction slot with sinusoidal inlet velocity and f ¼ 0.167 Hz A) based on
Force analysis B) based on velocity analysis.

274
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 6
The value of improvement in torque coefficient for NACA0015 with suction slot at different position angles under sinu-
soidal inlet velocity.

2013; Tenaud and Phuoc, 1997) and Hills (Chaudhariet al, 2014)). oscillating velocities are almost exact sinusoidal waves.
Moreover, numerous 2D models applications including the problems For unsteady flow with constant velocity, Fig. 4 A shows a very good
dealing with dam-break (Ozg€ € okmen et al., 2007), mechanism of agreement between the experimental data for the torque coefficient from
pollutant (Michioka et al., 2010; Chung and Liu, 2013), heat transfer reference (Torresi et al., 2007a, 2007b, 2009) and the calculated torque
(Horvata and Jure Marnb, 2001; Matos et al., 1999) used LES as a tur- coefficients at Reynolds number equals 2  105 from CFD results. It can
bulence model. In this work two sets of experimental data from refer- be noted that the CFD model has close stall condition value as the
ences were used to validate the numerical model. First experimental data experimental data from reference. Table 1 shows the comparison be-
from references (Torresi et al., 2007a, 2007b, 2009) is used to validate tween those results and the percentage of error. Furthermore, for an
the stall condition. Details of the first validation case, where Wells tur- unsteady flow with sinusoidal inlet velocity, Fig. 4 B shows good
bine prototype under investigation is characterized by the following agreement between experimental data for the drag force from reference
parameters: hub radius is 101 mm and 155 mm for the tip radius. (Nomura et al., 2003) and the calculated drag force from CFD at two
NACA0015 blade profile section with 74 mm chord length, and number different frequencies (1 Hz and 2 Hz). It can be seen from Fig. 4 B that the
of blades are 7. Furthermore, the hub-to-tip ratio is 0.65 and the solidity CFD model has approximately the same characteristics of oscillating flow
is 0.64, with 0.05 uncertainty in the measurements. Second experimental velocity as the experimental data from reference. The error percentage
data from reference (Nomura et al., 2003) is adopted to simulate and for the two frequencies is also shown in Table 2.
validate the sinusoidal wave inlet flow unsteady velocity. In the second
validation case, the unsteady forces (FD ) exerted on a square cylinder 4. Results and discussion
under oscillating flow with nonzero mean velocity were computed and
compared to experimental data. In the experiment, the oscillating air A single suction slot with certain diameter ðDss Þ representing 0.1%
flow was generated by an AC servomotor wind tunnel. The generated (Shehata et al., 2017d)of the blade chord and located at distance ðLss Þ

275
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 10. Comparison between the average torque coefficients for the optimum angles of suction slot with sinusoidal inlet velocity and f ¼ 0.167 Hz A) based on Force
analysis B) based on velocity analysis.

276
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 7
The value of global entropy generation rate for NACA0015 with suction slot at different position angles under sinusoidal
inlet velocity.

equal to 50% of the blade chord with different angles for the suction slot condition to decide which angle provides an optimum value of CT and
positions ðαss Þ (anticlockwise and clockwise) was created, with a shape SG . Third, investigate the effect of a slot with different angles in aero-
of NACA0015 from reference (Torresi et al., 2007a, 2007b, 2009), see acoustic noise of Wells turbine airfoil at the stall and near-stall conditions
Fig. 5. The angles for the suction slot were varied to get an optimum value under oscillating and non-oscillating flow conditions. Finally a compar-
of CT and SG . Table 3 highlights the different test cases that were ative analysis with different sinusoidal wave frequencies depend on
investigated in this work. The angle of suction slots at upper and lower operating conditions from Egyptian Northern Coast was investigated as
surface was defined with ðαss Þ of upper surface only in the analysis and well.
results which were presented hereafter. First, test cases under unsteady
flow with non-oscillating velocity were investigated in order to obtain an
indication about the well-performing suction slot angles. Second, these 4.1. Different suction slot angles under non-oscillating inlet velocity
well-performing angles were investigated under sinusoidal wave
Fig. 6 A shows the CT values of suction slot with different angles. It

277
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 11. Comparison between the second law efficiency for the compression cycle for NACA0015 with suction slots at different angles “—without suction slot” “ —
αSS ¼ 0 ” (f ¼ 0.167 Hz) A) 11.3 B) 11.7 C) 12.3 D) 13.6 E) 14.4.

278
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 12. Path-line colored by mean vorticity magnitude for sinusoidal flow around NACA0015 without suction slot and with the optimum angles of suction slot under
sinusoidal inlet velocity at the stall angle (13.6 Degree) and f ¼ 0.167 Hz A) 2.92 m/s -maximum velocity B) 1.8 m/s -decelerating flow.

279
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

can be noticed that the suction slot with a non-zero angle gives a higher 4.2. Optimum suction slot angles based on first and second law analysis
CT than that with 0 . For instance, the suction slots with αss equal to 35 under sinusoidal inlet velocity
clockwise, 65 anticlockwise and 55 anticlockwise give CT higher than
that with 0 by 22%, 20% and 19%, respectively. Table 4 lists CT values From the previous section, it was noted that three values of αss (35
for the investigated αss angles. The SG values of suction slot with clockwise and 65, 55 anticlockwise) provide the highest values of CT
different angles are shown in Fig. 6 B. It can be concluded that, the (60%, 58%, and 57%) at the stall regime. In addition, two values of αss
suction slot with αss equal to 15 anticlockwise, 10 clockwise and 10 (15 anticlockwise and 10 clockwise) have shown to provide the lowest
anticlockwise gives SG lower than that with 0 by 13%, 13% and 3%, SG value (8%) at the stall regime. In this section, the optimum αss for
respectively, see Table 5 for more details. The slot increases CT and de- single suction slot at the middle of the airfoil was investigated based on
lays the stall angle, furthermore, at some αss not equal to zero produce both CT and SG values under sinusoidal wave condition. Figs. 7 and 8
higher improvement in CT than αss equal to zero due to the improvement illustrate the hysteretic behavior due to the reciprocating flow.
in the flow layers around the aerofoil. On the other hand, SG value de- Fig. 7 A) shows the comparison between the suction slot with 0 and
pends on the velocity gradient and the increases in velocity magnitude that with 55 anticlockwise. It can be noted that both the accelerating
around the aerofoil lead to increase in SG . Furthermore, at some αss not flow and decelerating flow of suction slot with αss equal to 55 anti-
equal to zero, the change in velocity gradient was lower than αss equal to clockwise have higher CT than that with suction slot angel αss equal to
zero due to the improvement in the flow layers around the aerofoil. More 0 by 6% (accelerating) and 5% (decelerating). Furthermore, the accel-
details about the flow layers and pressure distribution around the aerofoil erating flow for αss equal to 65 anticlockwise has a lower CT than that
will be shown in upcoming sections. for αss equal to 0 by 7%, see Fig. 7 B), while for the decelerating flow it
provides a 3% higher CT than that for αss equal to 0 . From the

Fig. 13. The pressure distribution around NACA0015 without slot and with the optimum angles of suction slot under sinusoidal velocity and f ¼ 0.167 Hz A)
Contours of pressure coefficient B) pressure coefficient at the upper and lower surface.

280
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

comparison in Fig. 7C) between the suction slot with 0 and that with 35 than that of the suction slot with αss equal to 0 by 4% (accelerating) and
clockwise, it can be noted that the CT is the same as an average value in 9% (decelerating).
accelerating flow but for the decelerating flow the 35 clockwise has Fig. 9 shows the instantaneous torque coefficient at compression cycle
higher CT than the 0 by 7%. Fig. 8 A) shows that the accelerating flow for different suction slot angles, these values are at angle of attack of
for suction slot with αss equal to 15 anticlockwise has a lower CT than 13.6 .It can be concluded that all five angles in Fig. 9A) and B) have
that of the suction slot with αss equal to 0 by 2%, while the deceler- higher peak value of CT than the suction slot with αSS equal to 0 and also
ating flow has a higher CT than the suction slot with αss equal to 0 by as average for the compression cycle. For more details about other angles
4%. Finally, Fig. 8 B) shows that both the accelerating flow and decel- of attack see Table 6.Fig. 10 shows the average value of torque coefficient
erating flow of suction slot with αss equal to 10 clockwise has higher CT during the compression cycle. It can be noted that the suction slot with

Fig. 13. (continued).

281
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 14. The Contour at the stall angle (13.6 Degree) for sinusoidal flow around NACA0015 without suction slot and with the optimum angles of suction slot at
maximum velocity 2.92 m/s under sinusoidal velocity and f ¼ 0.167 Hz A) Velocity magnitude B) Global entropy generation rate.

282
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 15. The Sound pressure level in dB for NACA0015 under non-oscillating velocity at a far field receiver located at 35 chord with different slot angle and different
angles of attack A) 11.3 B) 12.3 C) 14.4 .

283
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 16. The Sound pressure level in dB for NACA0015 under non-oscillating velocity at a far field receiver located at 128 chord with different slot angle and different
angles of attack A) 11.7 B) 13.6 .

284
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 8
The overall sound pressure level in dB for NACA0015 under non-oscillating velocity at a far field receivers with different
slot angle and different angles of attack.

285
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 17. The Sound pressure level in dB for NACA0015 under oscillating velocity at a far field receiver located at 35 chord with different slot angle and different angles
of attack A) 11.3 B) 12.3 C) 13.6 .

286
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 17. (continued).

αss equal to 10 clockwise has the highest improvement in CT value second law efficiency than the NACA0015 without suction slot by 0.5%
before the stall by 21%. Furthermore, both the suction slot with αss equal before the stall (Fig. 11 A, B and C) and 0.2% after the stall (Fig. 11D and
to 55 anticlockwise and the suction slot with 10 clockwise have the E) as an average value. Furthermore, it has the same improvement in
highest improvement in CT value after the stall by 44%, see Table 6. second law efficiency from suction slot with αss equal to 0 as an average
Table 7 highlights the comparison between the values of SG for value before the stall (Fig. 11 A, B and C) and after the stall (Fig. 11D and
NACA0015 with suction slot at different position angles under sinusoidal E).
inlet velocity. It can be concluded that both the suction slots with αss The most important reason behind the improvement in the cases for
equal to 15 anticlockwise and the suction slot with αss equal to 10 the αss not equal to 0 more than the cases for the αss equal to 0 due to the
clockwise have SG lower than the suction slot with αss equal to 0 . flow layers and pressure distribution around the aerofoil. Therefore,
Where, the suction slot with αss equal to 10 clockwise has the lowest Figs. 12–14 were highlighting the flow layers and pressure distribution
increase in SG value before the stall by 20%. Furthermore, the suction around the aerofoil at different αss values.
slot with αss equal to 15 anticlockwise has the lowest increase in SG The effect of suction slot on the separation layers around the trailing
value after the stall by 18%. edge area can be noted at Fig. 12. Where, the mean velocity magnitude
Fig. 11shows the comparison between the second law efficiency of path lines around the NACA0015 without and with suction slot for
NACA0015 without and with suction slot which have different angles. different αss were presented. These values were at the instantaneous
The comparison was provided as an average value for the compression velocities of 2.92 m/s (Figs. 12A) and 1.8 m/s (Fig. 12B) for the decel-
stage with various angles of attack. The increase in SG (Table 7) leads to erating flow and at angle of attack of 13.6 (stall angle). Also, the
decrease in second law efficiency in most cases than that without suction improvement effect of suction slot on separation layer increased in the
slots. Otherwise, the suction slot with αss equal to 15 anticlockwise and second half of the cycle (deceleration flow) because the separation region
10 clockwise give a higher second law efficiency than the NACA0015 around the end of the blade increased, especially at the decelerating flow
without suction slot and with suction slot have αss equal to 0 . Where, the at Fig. 12B).
suction slot with αss equal to 15 anticlockwise provides a higher second From Fig. 13, it can be noted that the low pressure zones around the
law efficiency than the NACA0015 without suction slot by 0.4% before aerofoil, especially at the trailing edge area, were reduced by adding the
the stall (Fig. 11A, B and C) and 0.7% after the stall (Fig. 11D and E) as an suction slot. The slot with angle not equal to zero gives better result from
average value. Also it provides a higher second law efficiency than suc- that with zero degree, for example, 35 (clockwise) degrees or 10
tion slot with αss equal to 0 by 0.5% after the stall (Fig. 11D and E) but (clockwise) degrees. The slot with 35 or 10 were reduced the low
the same value as average before the stall (Fig. 11 A, B and C). On the pressure area around the aerofoil and also the difference between the
other hand, the suction slot with αss equal to 10 clockwise has higher pressure at the upper and lower surface higher than the slot with zero

287
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 18. The Sound pressure level in dB for NACA0015 under oscillating velocity at a far field receiver located at 128 chord with different slot angle and different
angles of attack A) 11.7 B) 13.6 C) 14.4 .

288
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 18. (continued).

degree. Therefore, the slot with 35 or 10 have more improvement in the obtained of sound pressure level are investigate and analyze. Figs. 15 and
separation layer and the torque coefficient than the slot with zero degree. 16 show the sound pressure level in dB for NACA0015 under non-
The flow structures over the NACA0015 aerofoil without and with oscillating velocity at two far field receivers with different slot angles
suction slot has different slot angles in oscillating flow are shown in and different angles of attack. It can be noted that, the sound pressure
Fig. 14. Fig. 14shows the contours of velocity magnitude (Fig. 14A) and level (SPL) has the highest value at the stall angle (13.6 ). Furthermore,
entropy (Fig. 14B) at maximum velocity 2.92 m/s and 13.6 angle of the NACA0015 with αSS not equal to zero have direct effect on the value
attack (stall angle). The improvement on flow structures was clear, due to of SPL, however, αSS equal to 65 anticlockwise decreased from SPL by
the suction slot, when comparing the NACA 0015 airfoil section without 12.7% before the stall and by 11.6% after the stall regime at first
and with slot. Mostly, in the separated layer regime that located on the receiver. The value of SPL was decreased, with αSS equal to 65 anti-
end of aerofoil. The suction slot effects directly on the flow structures that clockwise, by 20.1% before the stall and by 21% after the stall regime
located on the end of blade, and then leads to an improvement in the at second receiver. Under non-oscillating velocity condition, NACA0015
separation regime. However, adding a suction slot shows a bad effect on with a αSS not equal to zero are preferable than that equal to zero, except
the entropy generation rate, wherever higher entropy generation values the αSS equal to 35 clockwise at 14.4 . However, the maximum reduc-
were obtained for all suction slots cases. The suction slot with αss equal to tion create due to αSS equal to 65 anticlockwise by 16.6% at 11.7 at a
55 anticlockwise generates the highest value of entropy with an increase far field receiver located at 35 chord, on the other hand, the αSS equal to
of 63% than the NACA0015 without suction slot case. In addition the 65 anticlockwise give the maximum reduction at a far field receiver
lowest value for aerofoil with suction slot was obtained with αss equal to located at 128 chord by 26.4% at 11.7 , see Table 8 for more details
15 anticlockwise, with an increase of 18% only than the NACA0015 about the overall sound pressure level in dB for NACA0015 under non-
without suction slot case. From Fig. 12A and B it can be concluded that oscillating velocity with different angles of attack.
the attached slot with angle not equal to zero to the aerofoil lead to in- The sound pressure levels for NACA0015 under oscillating velocity at
crease in velocity magnitude around the aerofoil, furthermore, it also a far field receivers located at 35 chord and 128 chord with different slot
lead to increase in the entropy generation in Fig. 14 B where the entropy angles and different angles of attack were shown in Figs. 17 and 18.
value depends on the velocity gradient, see equation (19). Compared with the non-oscillating velocity, the SPL increase due to the
oscillating velocity condition at all angles of attack. It can be concluded
4.3. Aeroacoustic noise analysis that, the stall regime has the highest value of (SPL) at 13.6 and 14.4 .
Moreover, the NACA0015 with αSS equal to 35 clockwise has the lowest
The Ffowcs-Williams and Hawkings (FW–H) acoustic analogy is used value of SPL at 11.3 by 4.1% at first receiver and 19% at second
to calculate the acoustic pressure at a far field receiver located at 35 receiver. The αSS equal to 10 clockwise has the lowest value of SPL with
chord and 128 chord, perpendicular to the chord and the results that angle of attack 11.7 by 5.2% at the receiver located 35 chord and by

289
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 9
The overall sound pressure level in dB for NACA0015 under oscillating velocity at a far field receivers with different slot
angle and different angles of attack.

17% at the receiver located 35 chord. The lowest value of SPL at 12.3 the 10 clockwise has higher CT than the 0 by 1% (Fig. 19 A). The
given by αSS equal to 55 anticlockwise, which it was decreased by suction slot with αSS equal to 10 clockwise has a higher peak value of CT
5.2% at first receiver and 12.3% at second receiver. At stall regime, than the suction slot with αSS equal to 0 at angle of attack of 13.6 . On
the αSS equal to 10 clockwise has the lowest value of SPL with 13.6 and the other hand, as an average value for the compression cycle, the αSS
14.4 at the two receivers. Where, the SPL was decreased by 7.3% at equal to 10 clockwise has a lower value of CT than the αSS equal to 0 by
receiver one and by 21.2% at receiver two after the stall regime. Table 9 3% (Fig. 19B) and C)). For more details about other angles of attack for
show for more details about the overall sound pressure level in dB for sinusoidal wave with time period equal to 4 s, see Table 10. It can be
NACA0015 under oscillating velocity with different angles of attack. noted that from Fig. 20A) and B) and C), the suction slot with αss equal to
Finally, the maximum reduction create due to αSS equal to 10 clockwise 10 clockwise has higher improvement in CT value before the stall than
by 8% at 13.6 at a far field receiver located at 35 chord, on the other the aerofoil without slot by 7%. However, the suction slot with αss equal
hand, the αSS equal to 10 clockwise give the maximum reduction at a far to 0 gives higher improvement in CT value after the stall by degrees
field receiver located at 128 chord by 23.2% at 14.4 . 10% than the aerofoil without slot. See Table 10 for more details about
sinusoidal wave with time period equal to 8 s.
It can be concluded that from Table 11, the suction slot with αss equal
4.4. Different frequencies effect
to 10 clockwise gives SG value lower than the suction slot with αss equal
to 0 at different time period before and after the stall, except at the time
It was concluded from previous section that the optimum αSS is 10
period equal to 4 s, where the suction slot with αss equal to 0 has the
clockwise. Since this suction slot angle gives the highest CT before and
lowest increase in SG value after the stall by 12%.
after the stall (Table 6). On the other hand, it gives a lower SG than other
The increase in SG (Table 11) leads to the decrease in second law
angles before the stall and also it gives lower SG than 0 angle after the
efficiency in most cases than that without suction slots, see Fig. 21.
stall (Table 7). So, this optimum αSS is studied using the OWC technique
Where, the suction slot with αss equal to 10 clockwise under sinusoidal
depend on the real data from the Egyptian site under different fre-
wave with time period equal to 4 s has lower second law efficiency than
quencies (f equal to0.25 and 0.125 Hz) and time periods (4 and 8 s).
the NACA0015 without suction slot by 0.3% before the stall (Fig. 21 A, B
The CT for both αSS equal to 0 and 10 clockwise have the same
and C) and after the stall (Fig. 21D and E) as an average value.
value (as an average) at the accelerating flow but for decelerating flow

290
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 19. The NACA0015 without and with suction slot at optimum angle under sinusoidal wave with time period 4 s (0.25 Hz) “—without suction slot — αSS ¼ 0 ”,“
— αSS ¼ 10 clockwise A) The hysteretic behavior B) The instantaneous torque coefficient C) The average torque coefficient.

Furthermore, it also has the highest second law efficiency than suction ciency than the NACA0015 without suction slot by 0.6% before the stall
slot with αss equal to 0 as an average value before the stall (Fig. 21 A, B (Fig. 21 A, B and C) and after the stall (Fig. 21D and E) as an average
and C) by 0.1% and lower than it after the stall (Fig. 21D and E) by value. Furthermore, it has the highest second law efficiency when
0.2%.On the other hand, the suction slot with αss equal to 10 clockwise compared to suction slot with αss equal to 0 as an average value before
under sinusoidal wave with time period equal to 6 s has a higher second the stall (Fig. 21 A, B and C) by 0.1% and lower than it after the stall
law efficiency than the NACA0015 without suction slot by 0.5% before (Fig. 21D and E) by 0.6%.
the stall (Fig. 21 A, B and C) and 0.2% after the stall (Fig. 21D and E) as an As it was mentioned above, the flow layers and pressure distribution
average value. Also, the 10 clockwise has the same improvement in around the aerofoil are the most important reason behind the improve-
second law efficiency compare with the 0 as an average value before the ment in the cases with αss not equal to 0. Therefore, Figs. 22 and 23 were
stall (Fig. 21 A, B and C) and after the stall (Fig. 21D and E). highlighting the flow layers and pressure distribution around the aerofoil
Finally, the suction slot with αss equal to 10 clockwise under sinu- at different conditions, such as tsin and αss .
soidal wave with time period equal to 8 s gives a lower second law effi- The separation layer at the end of blade was effected by the suction slot

291
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 10
Comparison between the torque coefficients values at different time periods under sinusoidal inlet velocity.

292
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 20. The NACA0015 without and with suction slot at optimum angle under sinusoidal wave with time period 8 s (0.125 Hz) “—without suction slot — αSS ¼ 0 ”,“
— αSS ¼ 10 clockwise A) The hysteretic behavior B) The instantaneous torque coefficient C) The average torque coefficient.

293
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Table 11
Comparison between the global entropy generation rate values at different time periods under sinusoidal inlet velocity.

(Fig. 22 A) can be noted at the different time periods from the path lines 5. Conclusions
colored by mean velocity magnitude around the NACA0015 at 1.8 m/s
velocity for the decelerating flow and at angle of attack of 13.6 (stall Several cases were solved to determine the optimum angle for single
angle). More than that, both the low pressure zones and the difference suction slot at the middle of two-dimensional airfoil. The optimum value
between the upper and lower surface were decreased especially for the slot has been decided based on two criteria: 1) maximizing the obtained
with αss equal to 10 clockwise at different time periods, see Fig. 22B and torque coefficient value to maximize the first law of thermodynamics
C. It can be noted that from Fig. 23 the suction slots have a negative effect efficiency, and 2) minimizing the generated entropy value to maximize
on the entropy generation at the different time periods. However, the the second law of thermodynamics efficiency. For this purpose, the en-
suction slot with αss equal to 10 clockwise has the lower SG value with an tropy generation minimization method was used to obtain the local en-
increase of only 10% before the stall and 14% after the stall than the tropy viscosity predictions of the different cases. Furthermore, the effect
NACA0015 without suction slot by under sinusoidal wave with time of a single suction slot with angle equal and not equal to zero attached to
period equal to 4 s. In addition, it has the lowest value under sinusoidal airfoil section on the aeroacoustic noise at the far field. The aerodynamic
wave with time period equal to 8 s with an increase than the NACA0015 noise under different operating conditions, (such as, before and after the
without suction slot by 16% before the stall and 18% after the stall. stall regime and non-oscillating and oscillating velocity) and different

294
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 21. The second law efficiency at the compression cycle for the NACA0015 without and with suction slot at optimum angle with sinusoidal velocity and tsin equal
to 4 sec (0.25 Hz), 6sec (0.167 Hz), 8sec (0.125 Hz) “ ■ without suction slot ■ αSS ¼ 0 ”,“ ■ αSS ¼ 10 clockwise A) 11.3. B) 11.7. C) 12.3. D) 13.6. E) 14.4.

295
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 22. Flow structure around the NACA0015 and tsin equal to 4 sec (0.25 Hz), 6sec (0.167 Hz), 8sec (0.125 Hz) A) Path-line coloured by mean velocity magnitude B)
Contour of pressure coefficient C) pressure distribution at upper and lower surface.

296
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 22. (continued).

design parameters, (such as, airfoil section without slot and with slot the suction slot with angle to progress the performance at the stall con-
have angle equal and not equal to zero) was analyze and investigate. dition with minimize the increase in entropy generation. Otherwise, it
Then, the comparative analysis based on actual location data relevant to may not be effective. Moreover than that, the aerodynamics noise can be
northern coast of Egypt was applied using the airfoil with optimum angle reduction by a single slot with angle not equal to zero. The suction slot
of suction slot. The two-dimensional incompressible unsteady flow was with angle equal to 10 clockwise is decreased the aeroacoustic noise at
used to simulate these cases under different conditions. the stall regime by 7.3% and by 21.2% at a far field receiver located at
The modeling results show that optimum angle for suction slot is 10 35 chord and 128 chord respectively.
clockwise. This angel gives the highest improvement in the torque co- To conclude, future study and research should concentrate on prog-
efficient by 21% before the stall and 44% after the stall. These values are ress the efficiency of first law with a minimize entropy generation, by
higher than the suction slot with 0 by 3% before the stall and 15%. On using numerical algorism (Selim et al., 2017) and experimental labora-
the other hand, this angel gives the lowest global entropy generation rate tory studies, to enhance the overall Wells turbine performance under
than the suction slot with 0 by 4% before the stall and 3%.The airfoils flow control method. In addition, the passive flow control using slot with
with optimum locations for multi suction slots under conditions based on angle given very promising result for Wells turbine, therefore, it be worth
Egyptian northern coast with different frequency were investigated. The to investigate its effect on other turbine such as water turbine (Elbatran
suction slot with angle equal to 10 clockwise is mostly providing a et al., 2017). Moreover, there are many parameters that can be used
higher torque coefficient and a lower global entropy generation rate than future study to reduction the aerodynamics noise such as the number of
the suction slot with angle equal to 0 at different time periods (4, 6 and suction slots, the distance between suction slots, the location of suction
8 s) before and after the stall. The delay of stall condition lead to the slots and the angle of suction slots. Furthermore, the operating conditions
improvement in the torque coefficient after the stall. The suction slot for the Egyptian northern coast are very appropriate for the OWC system
increases the torque coefficient and delays the stall angle which further with a wave energy extractor such as Wells turbine. Therefore, it is very
leads to an increase the efficiency of first law. Otherwise, it decreases the important that to consider to the potential of wave energy in Egypt as the
efficiency of second law. For-that, if the turbine will be under the passive path to minimize fossil fuel usage.
flow control by using the suction slot, it is strongly recommended to use

297
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Fig. 23. The sinusoidal flow around NACA0015 without slot and with the optimum angle of suction slot at decelerating flow and tsin equal to 4 sec (0.25 Hz), 6sec
(0.167 Hz), 8sec (0.125 Hz) A) Path-line coloured by mean velocity magnitude B) Contour of global entropy generation rate.

Acknowledgements Boccotti, P., 2007. Caisson breakwaters embodying an OWC with a small opening—Part I:
Theory. Ocean. Eng. 34 (5–6), 806–819.
Botha, J.D.M., Shahroki, A., Rice, H., 2017. An implementation of an aeroacoustic
The authors would like to acknowledge the support provided by the prediction model for broadband noise from a vertical axis wind turbine using a CFD
Department of Naval Architecture, Ocean and Marine Engineering at informed methodology. J. Sound Vib. 410, 389–415.
Strathclyde University, UK and the Marine Engineering Department at Braslow, A.L., 1999. A history of suction type laminar flow control with emphasis on
flight research, NASA History Division. Monogr. Aerosp. Hist. 13.
Arab Academy for Science, Technology and Maritime Transport. Brito-Melo, A., Gato, L.M.C., Sarmento, A.J.N.A., 2002. Analysis of Wells turbine design
parameters by numerical simulation of the OWC performance. Ocean. Eng. 29,
References 1463–1477.
Bromby, D.Y.a.W., 2012. Large-eddy simulation of unsteady separation over a pitching
airfoil at high Reynolds number. In: Seventh International Conference on
Akcayoz, E., Tuncer, I.H., 2009. Numerical investigation of flow control over an airfoil Computational Fluid Dynamics (ICCFD7). Big Island, Hawaii.
using synthetic jets and its optimization. In: International Aerospace Conference, Chapin, V.G., Benard, E., 2015. Active control of a stalled airfoil through steady or
Turkey. unsteady actuation jets. J. Fluids Eng. 137 (9), 091103.
Alferez, N., Mary, I., Lamballais, E., 2013. Study of stall development around an airfoil by Chaudhari, Ashvinkumar, et al., 2014. Large Eddy Simulation of Boundary-layer Flows
means of high fidelity large eddy simulation. Flow, Turbul. Combust. 91 (3), over Two-dimensional Hills, in Industrial Mathematics at ECMI 2012. 2012, Springer
623–641. International Publishing Switzerland, pp. 211–218.
Alklaibi, A.M., Khan, M.N., Khan, W.A., 2016. Thermodynamic analysis of gas turbine Chawla, J.S., et al., 2014. Efficiency improvement study for small wind turbines through
with air bottoming cycle. Energy 107, 603–611. flow control. Sustain. Energy Technol. Assessments 7, 195–208.
AlMutairi, J., AlQadi, I., ElJack, E., 2015. In: Large Eddy Simulation of a NACA-0012 Cheng, W.-C., Porte-Agel, F., 2013. Evaluation of subgrid-scale models in large-eddy
Airfoil Near Stall, 20, pp. 389–395. simulation of flow past a two-dimensional block. Int. J. Heat Fluid Flow 44, 301–311.
Armenio, V., Geurts, B., Fr€ohlich, J., 2010. In: Large Eddy Simulation of Flow Around an Chung, T.N.H., Liu, C.-H., 2013. On the mechanism of air pollutant removal in two-
Airfoil Near Stall, 13, pp. 541–545. dimensional idealized street canyons: a large-eddy simulation approach. Boundary-
Avallone, F., van der Velden, W.C.P., Ragni, D., 2017. Benefits of curved serrations on Layer Meteorol. 148 (1), 241–253.
broadband trailing-edge noise reduction. J. Sound Vib. 400, 167–177. Curran, R., et al., 1998. Performance prediction of contrarotating wells turbines for wave
Avdis, A., Lardeau, S., Leschziner, M., 2009. Large eddy simulation of separated flow over energy converter design. J. Energy Eng. 124 (2), 35–53.
a two-dimensional hump with and without control by means of a synthetic slot-jet. Dahlstrom, S., 2003. Large eddy simulation of the flow around a high-lift airfoil. In:
Flow, Turbul. Combust. 83 (3), 343–370. Department of Thermo and Fluid Dynamics. CHALMERS UNIVERSITY OF
Ayat, B., 2013. Wave power atlas of eastern mediterranean and aegean seas. Energy 54, TECHNOLOGY, Goteborg, Sweden, p. 62.
251–262. Dixon, S.L., 1998. Fluid Mechanics, Thermodynamics of Turbomachinery. Pergamon
Barstow, S., et al., 2009. WorldWaves wave energy resource assessments from the deep Press Ltd.
ocean to the coast. In: The 8th European Wave and Tidal Energy Conference. DK, L., 1992. A proposed modification of the Germano subgrid-scale closure method.
Proceedings of the 8th European Wave and Tidal Energy Conference: Uppsala, Phys. Fluids A 4 (3), 633–635.
Sweden. Elbatran, A.H., Ahmed, Y.M., Shehata, A.S., 2017. Performance study of ducted nozzle
Baskut, O., Ozgener, O., Ozgener, L., 2010. Effects of meteorological variables on Savonius water turbine, comparison with conventional Savonius turbine. Energy 134,
exergetic efficiency of wind turbine power plants. Renew. Sustain. Energy Rev. 14 566–584.
(9), 3237–3241. Falcao, A.F.d., 2004. Stochastic Modelling in wave power-equipment optimisation:
Baskut, O., Ozgener, O., Ozgener, L., 2011. Second law analysis of wind turbine power maximum energy production versus maximum profit. Ocean. Eng. 31 (2004),
plants: cesme, izmir example. Energy 36 (5), 2535–2542. 1407–1421.
Bejan, A., 1995. Entropy Generation Minimization: the Method of Thermodynamic Falcao, A.F.d.O.J., P.A.P., 1999. OWC wave energy devices with air flow control. Ocean.
Optimization of Finite-size Systems and Finite-time Processes. Taylor & Francis. Eng. 26 (12), 1275–1295.
Bejan, A., 1996. Entropy generation minimization- the new thermodynamics of finite-size Fernandez, E., Kumar, R., Alvi, F., 2013. Separation control on a low-pressure turbine
devices and finite-time processes. Appl. Phys. Rev. 79 (3), 1191–1218. blade using microjets. J. Propuls. Power 29 (4), 867–881.
Boccotti, P., 2007. Comparison between a U-OWC and a conventional OWC. Ocean. Eng.
34, 799–805.

298
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Folley, M., Curran, R., Whittaker, T., 2006. Comparison of LIMPET contra-rotating wells De Moura, C.A.K., Carlos, S., 2013. The Courant–friedrichs–lewy (CFL) Condition: 80
turbine with theoretical and model test predictions. Ocean. Eng. 33 (8–9), Years after its Discovery, 1 ed. Birkh€auser Basel, Boston.
1056–1069. Nazeryan, M., Lakzian, E., 2018. Detailed entropy generation analysis of a Wells turbine
Gea-Aguilera, F., Gill, J., Zhang, X., 2017. Synthetic turbulence methods for using the variation of the blade thickness. Energy 143, 385–405.
computational aeroacoustic simulations of leading edge noise. Comput. Fluids 157, Nomura, T., et al., 2003. Aerodynamic forces on a square cylinder in oscillating flow with
240–252. mean velocity. J. Wind Eng. Ind. Aerod. 91, 199–208.
Halder, P., Samad, A., Th evenin, D., 2017. Improved design of a Wells turbine for higher Okuhara, S., et al., 2013. Wells turbine for wave energy conversion —improvement of the
operating range. Renew. Energy 106, 122–134. performance by means of impulse turbine for Bi-Directional flow. Open J. Fluid Dyn.
Halder, P., Rhee, S.H., Samad, A., 2017. Numerical optimization of Wells turbine for wave 03 (02), 36–41.
energy extraction. Int. J. Nav. Archit. Ocean Eng. 9 (1), 11–24. € okmen, T.M., et al., 2007. Large eddy simulation of stratified mixing in two-
Ozg€
Hinze, J.O., 1975. Turbulence. McGraw-Hill Publishing Co, , New York. dimensional dam-break problem in a rectangular enclosed domain. Ocean. Model. 16
Hirsch, C., 2007. Numerical Computation of Internal and External Flows: the (1–2), 106–140.
Fundamentals of Computational Fluid Dynamics. Elsevier Science. Paderi, M., Puddu, P., 2013. Experimental investigation in a Wells turbine under bi-
Hitiwadi, M., et al., 2013. Large eddy simulations of 2D and open-tip airfoils using voxel directional flow. Renew. Energy 57, 570–576.
meshes. Procedia Eng. 61, 32–39. Pope, K., Dincer, I., Naterer, G.F., 2010. Energy and exergy efficiency comparison of
Hitoshi Hotta, Y.W., 1985. A study on the matching between the air turbine and phase horizontal and vertical Axis wind turbines. Renew. Energy 35 (9), 2102–2113.
control for the OWC wave power generator. Ocean. Eng. 12 (6), 585–586. Raghunathan, S., 1995. The wells air turbine for wave energy conversion. Prog. Aerosp.
Horvata, Andrej, Jure Marnb, I.K., 2001. Two dimensional large eddy simulation of Sci. 31, 335–386.
turbulent natural convection due to internal heat generation. Int. J. heat mass Transf. Raghunathan, S., 1995. A methodology for Wells turbine design for wave energy
44 (21), 3985–3995. conversion. ARCHIVE: proceedings of the Institution of Mechanical Engineers. Part A
Huang, L., Huang, P.G., LeBeau, R.P., 2004. Numerical study of blowing and suction J. Power Energy 1990-1996 209 (31), 221–232 (vols 204-210).
control mechanism on NACA0012 airfoil. J. Aircr. 41 (5), 1005–1013. Raghunathan, S., Tan, C.P., Wells, N.A.J., 1981. Wind tunnel tests on airfoils in tandem
Iandoli, C.L., 2005. 3-D numerical calculation of the local entropy generation rates in a cascade. AIAA J. 19 (11), 1490–1492.
radial compressor stage. Int. J. Thermodyn. 8, 83–94. Redha, A.M., Dincer, I., Gadalla, M., 2011. Thermodynamic performance assessment of
Ismail, M.F., Vijayaraghavan, K., 2015. The effects of aerofoil profile modification on a wind energy systems: an application. Energy 36 (7), 4002–4010.
vertical axis wind turbine performance. Energy 80, 20–31. Richards, E.J., Burge, C.H., 1943. An airfoil designed to give laminar flow over the surface
Kaviani, H.R., Nejat, A., 2017. Aerodynamic noise prediction of a MW-class HAWT using with boundary layer suction. Aeronaut. Res. Counc. R&M 2263.
shear wind profile. J. Wind Eng. Ind. Aerod. 168, 164–176. Richez, F., et al., 2007. Zonal RANS/LES coupling simulation of a transitional and
Kawai, S., Asada, K., 2013. Wall-modeled large-eddy simulation of high Reynolds number separated flow around an airfoil near stall. Theor. Comput. Fluid Dyn. 22 (3–4),
flow around an airfoil near stall condition. Comput. Fluids 85, 105–113. 305–315.
Kim, S.H., Kim, C., 2009. Separation control on NACA23012 using synthetic jet. Aerosp. Rosas, C.R., 2005. Numerical Simulation of Flow Separation Control by Oscillatort Fluid
Sci. Technol. 13 (4), 172–182. Injection. A&M University, Texas.
Kim, T.H., et al., 2002. Hysteretic characteristics of wells turbine for wave power Rumpfkeil, M.P., 2017. Using steady flow analysis for noise predictions. Comput. Fluids
conversion. In: The Twelfth International Offshore and Polar Engineering Conference. 154, 347–357.
The International Society of Offshore and Polar Engineers, Kitakyushu, Japan, Rumsey, C.L., Nishino, T., 2011. Numerical study comparing RANS and LES approaches
pp. 687–693. on a circulation control airfoil. Int. J. Heat Fluid Flow 32 (5), 847–864.
Kim, T.H., et al., 2002. Numerical investigation on the effect of blade sweep on the SB, P., 2000. Turbulent Flows. Cambridge University Press.
performance of Wells turbine. Renew. Energy 25, 235–248. Scarpetta, F., et al., 2017. CFD simulation of the unsteady flow in an Oscillating Water
Kim, Y., Castro, I.P., Xie, Z.T., 2015. In: Large-eddy Simulations for Wind Turbine Blade: Column: comparison between numerical and experimental results for a small scale
Dynamic Stall and Rotational Augmentation, 20, pp. 369–375. experimental device. In: Proceedings of the 12th European Wave and Tidal Energy
Kinoue, Y., et al., 2003. Mechanism of hysteretic characteristics of wells turbine for wave Conference, pp. 988(001)–988(007). Cork, Ireland.
power conversion. J. Fluids Eng. 125 (2), 302. Schatz, M., Günther, B., Thiele, F., 2007. Computational investigation of separation
Kinoue, Y., et al., 2004. Hysteretic characteristics of monoplane and biplane wells turbine control for high-lift airfoil flows. In: King, P.D.R. (Ed.), Active Flow Control, 95,
for wave power conversion. Energy Convers. Manag. 45 (9–10), 1617–1629. pp. 173–189. Berlin, Germany.
Koroneos, Christopher, Spachos, Thomas, Moussiopoulos, N., 2003. Exergy analysis of Schlichting, H., 1968. Boundary Layer Theory. McGraw-Hill, New York, USA,
renewable energy sources. Renew. Energy 28 (2003), 295–310. pp. 347–362.
Launder, B.E., Spalding, D.B., 1972. Lectures in Mathematical Models of Turbulence. Selim, M.M., Koomullil, R.P., Shehata, A.S., 2017. Incremental approach for radial basis
Academic Press, London, England. functions mesh deformation with greedy algorithm. J. Comput. Phys. 340, 556–574.
Lighthill, M.J., 1952. On sound generated aerodynamically I. General theory. Proc. R. Soc. Setoguchi, T., Takao, M., 2006. Current status of self rectifying air turbines for wave
a Math. Phys. Eng. Sci. 211 (1107), 564–587. energy conversion. Energy Convers. Manag. 47 (15–16), 2382–2396.
Lighthill, M.J., 1954. Sound generated aerodynamically.II.Turbulence as a source of Setoguchi, T., et al., 2001a. Effect of guide vane shape on the performance of a wells
sound. Proc. R. Soc. a Math. Phys. Eng. Sci. 222 (1148), 1–32. turbine. Renew. Energy 23, 1–15.
Liu, Z., et al., 2016. Numerical study on a modified impulse turbine for OWC wave energy Setoguchi, T., et al., 2003. Hysteretic characteristics of wells turbine for wave power
conversion. Ocean. Eng. 111, 533–542. conversion. Renew. Energy 28 (13), 2113–2127.
Mahboubidoust, A., Ramiar, A., 2017. Investigation of DBD plasma actuator effect on the Setoguchi, T., et al., 2003. Effect of rotor geometry on the performance of wells turbine.
aerodynamic and thermodynamic performance of high solidity Wells turbine. Renew. In: The Thirteenth International Offshore and Polar Engineering Conference. The
Energy 112, 347–364. International Society of Offshore and Polar Engineers, Honolulu, Hawaii, USA,
Maizi, M., et al., 2018. Noise reduction of a horizontal wind turbine using different blade pp. 374–381.
shapes. Renew. Energy 117, 242–256. Setoguchi, T., et al., 2003. A modified Wells turbine for wave energy conversion. Renew.
Mamun, M., 2006. The study on the hysteretic characteristics of the wells turbine in a Energy 28, 79–91.
deep stall condition. In: Energy and Material Science Graduate School of Science and Setoguchi, T., S.S., Maeda, H., Takao, M., Kaneko, K., 2001b. A review of impulse turbine
Engineering. Saga University, Japan, p. 141. for wave energy conversion. Renew. Energy 23 (2), 261–292.
Mamun, M., et al., 2004. Hysteretic flow characteristics of biplane wells turbine. Ocean. Shaaban, S., 2012. Insight analysis of biplane wells turbine performance. Energy Convers.
Eng. 31 (11–12), 1423–1435. Manag. 59, 50–57.
Masahiro Inoue, K.K., Setoguchi, Toshiaki, Shimamoto, Katsumi, 1985. On the starting Shaaban, S., 2016. Aero-economical optimization of Wells turbine rotor geometry. Energy
and quasi-steady characteristics of wells turbine under oscillating flow condition. Convers. Manag. 126, 20–31.
Ocean. Eng. 12 (6), 563. Shaltout, M.L., et al., 2015. Tradeoff analysis of energy harvesting and noise emission for
Masami Suzuki, C.A., Tagori, Tetsuo, 1985. Fundamental studies on oscillating water distributed wind turbines. Sustain. Energy Technol. Assessments 10, 12–21.
column wave power generator with wells turbine. Ocean. Eng. 12 (6), 565. Shehata, A.S., 2017. Investigation and improvement of wells turbine performance- fluid
Matos, A.d., Pinho, F.A.A., Silveira!Neto, A., 1999. Large-eddy simulation of turbulent analysis & 2nd law of thermodynamics study. In: Department of Naval Architecture,
~ow over a two! dimensional cavity with temperature ~uctuations. Int. J. Heat Mass Ocean and Marine Engineering. University of Strathclyde, Glasgow G4 0LZ, U.K,
Transf. 42, 3848. p. 291.
Michioka, T., et al., 2010. Large-eddy simulation for the mechanism of pollutant removal Shehata, A.S., et al., 2014. Entropy generation due to viscous dissipation around a wells
from a two-dimensional street canyon. Boundary-Layer Meteorol. 138 (2), 195–213. turbine blade: a preliminary numerical study. Energy Procedia 50, 808–816.
Miotto, R.F., Wolf, W.R., de Santana, L.D., 2017. Numerical computation of aeroacoustic Shehata, A.S., et al., 2016. Performance analysis of wells turbine blades using the entropy
transfer functions for realistic airfoils. J. Sound Vib. 407, 253–270. generation minimization method. Renew. Energy 86, 1123–1133.
Mohamed, M.H., Shaaban, S., 2013. Optimization of blade pitch angle of an axial turbine Shehata, A.S., et al., 2017. Wells turbine for wave energy conversion: a review. Int. J.
used for wave energy conversion. Energy 56, 229–239. energy Res. 41 (1), 6–38.
Mohamed, M.H., et al., 2011. Multi-objective optimization of the airfoil shape of wells Shehata, A.S., et al., 2017. Comparative analysis of different wave turbine designs based
turbine used for wave energy conversion. Energy 36 (1), 438–446. on conditions relevant to northern coast of Egypt. Energy 120, 450–467.
Moin, P.,S.K., Cabot, W., Lee, S., 1991. A dynamic subgrid-scale model for compressible Shehata, A.S., et al., 2017. Enhancement of performance of wave turbine during stall
turbulence and scalar transport. Phys. Fluids A 3 (11), 2746–2757. using passive flow control: first and second law analysis. Renew. Energy 113,
Mørk, G., et al., 2010. Assessing the global wave energy potential. In: 29th International 369–392.
Conference on Ocean, Offshore Mechanics and Arctic Engineering. ASME, Shanghai, Shehata, A.S., et al., 2017. Passive flow control for aerodynamic performance
China. enhancement of airfoil with its application in Wells turbine – under oscillating flow
condition. Ocean. Eng. 136, 31–53.

299
A.S. Shehata et al. Ocean Engineering 157 (2018) 262–300

Sheldahl, R.E., Klimas, P.C., 1981. Aerodynamic characteristics of seven symmetrical Torresi, M., Camporeale, S., Pascazio, G., 2007a. Performance of a small prototype of a
airfoil sections through 180-degree angle of attack for use in aerodynamic analysis of high solidity wells turbine. In: Seventh European Conference on Turbomachinery
vertical Axis wind turbines. In: Sandia National Laboratories Energy Report, p. 118 Fluid Dynamics and Thermodynamics. Athens, Greece.
the United States of America. Torresi, M., Camporeale, S., Pascazio, G., 2007b. Experimental and numerical
Shen, X., et al., 2017. Surface curvature effects on the tonal noise performance of a low investigation on the performance of a wells turbine prototype. In: Seventh European
Reynolds number aerofoil. Appl. Acoust. 125, 34–40. Wave and Tidal Energy Conference. Porto, Portugal.
Siozos-Rousoulis, L., Lacor, C., Ghorbaniasl, G., 2017. A flow control technique for noise Torresi, M., et al., 2008. Accurate numerical simulation of a high solidity Wells turbine.
reduction of a rod-airfoil configuration. J. Fluids Struct. 69, 293–307. Renew. Energy 33 (4), 735–747.
Skyllingstad, E.D., Wijesekera, H.W., 2004. Large-eddy simulation of flow over two- Torresi, M., Camporeale, S.M., Pascazio, G., 2009. Detailed CFD analysis of the steady
dimensional obstacles: high drag states and mixing. J. Phys. Oceanogr. 34, 94–112. flow in a wells turbine under incipient and deep stall conditions. J. Fluids Eng. 131
Smagorinsky, J., 1963. General circulation experiments with the primitive equations. I. (7), 071103.
The basic experiment. Mon. Wea. Rev. 91, 99–164. Volino, R.J., Kartuzova, O., Ibrahim, M.B., 2011. Separation control on a very high lift
Solís-Gallego, I., et al., 2018. LES-based numerical prediction of the trailing edge noise in low pressure turbine airfoil using pulsed vortex generator jets. J. Turbomach. 133 (4),
a small wind turbine airfoil at different angles of attack. Renew. Energy 120, 041021.
241–254. Walker, S.W., Raymer, W.G., 1946. Wind tunnel test on the 30 percent symmetrical griffth
Soltanmohamadi, R., Lakzian, E., 2015. Improved design of wells turbine for wave energy aerofoil with ejection of air. Aeronaut. Res. Counc. R&M 2475.
conversion using entropy generation. Meccanica 51 (8), 1713–1722. Springer Wang, S., et al., 2012. Turbulence modeling of deep dynamic stall at relatively low
Netherlands. Reynolds number. J. Fluids Struct. 33, 191–209.
Starzmann, R., Carolus, T., 2013. Effect of blade skew strategies on the operating range Wang, J., et al., 2017. Numerical study on reduction of aerodynamic noise around an
and aeroacoustic performance of the wells turbine. J. Turbomach. 136 (1), 011003. airfoil with biomimetic structures. J. Sound Vib. 394, 46–58.
Starzmann, R., Carolus, T., 2013. Model-based selection of full-scale wells turbines for Wasala, S.H., et al., 2015. Aeroacoustic noise prediction for wind turbines using Large
ocean wave energy conversion and prediction of their aerodynamic and acoustic Eddy Simulation. J. Wind Eng. Ind. Aerod. 145, 17–29.
performances. Proc. Inst. Mech. Eng. Part A J. Power Energy 228 (1), 2–16. Whittaker, T.J.T., Stewart, T.P., Curran, R., 1997. Design synthesis of oscillating water
Stodola, A., 1910. Steam and Gas Turbines. McGraw-Hill, New York. column wave energy converters: performance matching. Proc. Inst. Mech. Eng. Part A
Tenaud, Christian, Phuoc, L.T., 1997. Large eddy simulation of unsteady, compressible, J. Power Energy 211 (6), 489–505.
separated flow around NACA 0012 airfoil. In: Fifteenth International Conference on Williams, J.E.F., Hawkings, D.L., 1969. Theory relating to the noise of rotating machinery.
Numerical Methods in Fluid Dynamics, pp. 424–429. J. Sound Vib. 10 (1), 10–21.
Thakker, A., Abdulhadi, R., 2007. Effect of blade profile on the performance of wells Yagiz, B., Kandil, O., Pehlivanoglu, Y.V., 2012. Drag minimization using active and
turbine under unidirectional sinusoidal and real sea flow conditions. Int. J. Rotating passive flow control techniques. Aerosp. Sci. Technol. 17 (1), 21–31.
Mach. 2007, 1–9. Yousefi, K., Saleh, R., Zahedi, P., 2014. Numerical study of blowing and suction slot
Thakker, A., Abdulhadi, R., 2008. The performance of wells turbine under Bi-Directional geometry optimization on NACA 0012 airfoil. J. Mech. Sci. Technol. 28 (4),
airflow. Renew. Energy 33 (11), 2467–2474. 1297–1310.
Torres, F.R., Teixeira, P.R.F., Didier, E., 2016. Study of the turbine power output of an Yucer, C.T., 2016. Thermodynamic analysis of the part load performance for a small scale
oscillating water column device by using a hydrodynamic – aerodynamic coupled gas turbine jet engine by using exergy analysis method. Energy 111, 251–259.
model. Ocean. Eng. 125, 147–154. Yukihisa Washio, H.H., Miyazaki, Takeaki, Masuda, Yoshio, 1985. Full-scale performance
Torres, F.R., Teixeira, P.R.F., Didier, E., 2018. A methodology to determine the optimal tests on tandem wells turbine. Ocean. Eng. 12 (6), 564.
size of a wells turbine in an oscillating water column device by using coupled hydro- Zodiatis, G., et al., 2014. Wave energy potential in the eastern mediterranean levantine
aerodynamic models. Renew. Energy 121, 9–18. basin. An integrated 10-year study. Renew. Energy 69, 311–323.

300

You might also like