You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/337524092

A Critical Examination of the Hysteresis in Wells Turbines Using CFD and


Lumped Parameter Models

Conference Paper · June 2019


DOI: 10.1115/OMAE2019-96518

CITATIONS READS

0 12

6 authors, including:

Tiziano Ghisu Francesco Cambuli


Università degli studi di Cagliari Università degli studi di Cagliari
74 PUBLICATIONS   629 CITATIONS    32 PUBLICATIONS   378 CITATIONS   

SEE PROFILE SEE PROFILE

Irene Virdis Mario Carta


Università degli studi di Cagliari Università degli studi di Cagliari
10 PUBLICATIONS   44 CITATIONS    5 PUBLICATIONS   18 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Uncertainty Quantification and Robust Design View project

Experimental and Numerical Analysis of Wave Energy Conversion Systems View project

All content following this page was uploaded by Tiziano Ghisu on 07 June 2021.

The user has requested enhancement of the downloaded file.


Fabio Licheri1
Department of Mechanical, Chemical and
Materials Engineering,
University of Cagliari,
Cagliari 09123, Italy
e-mail: fabio.licheri@unica.it A Comparison of Different
Francesco Cambuli
Assistant Professor
Approaches to Estimate the
Department of Mechanical,
Chemical and Materials Engineering, Efficiency of Wells Turbines

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


University of Cagliari,
Cagliari 09123, Italy Wells turbines are among the most interesting power takeoff devices used in oscillating
e-mail: cambuli.f@unica.it water column (OWC) systems for the conversion of ocean-wave energy into electrical
energy. Several configurations have been studied during the last decades, both experi-
Pierpaolo Puddu mentally and numerically. Different methodologies have been proposed to estimate the
Professor efficiency of this turbine, as well as different approaches to evaluate the intermediate
Department of Mechanical, quantities required. Recent works have evaluated the so-called second-law efficiency of a
Chemical and Materials Engineering, Wells turbine, and compared it to the more often used first-law efficiency. In this study,
University of Cagliari, theoretical analyses and numerical simulations have been used to demonstrate how these
Cagliari 09123, Italy two efficiency measures should lead to equivalent values, given the low pressure ratio of
e-mail: pierpaolo.puddu@dimcm.unica.it the machine. In numerical simulations, small discrepancies can exist, but they are due to
the difficulty of ensuring entropy conservation on complex three-dimensional meshes.
Tiziano Ghisu The efficiencies of different rotor geometries are analyzed based on the proposed meas-
Associate Professor ures, and the main sources of loss are identified. [DOI: 10.1115/1.4049686]
Department of Mechanical,
Chemical and Materials Engineering,
University of Cagliari,
Cagliari 09123, Italy
e-mail: t.ghisu@unica.it

1 Introduction
The growing demand for renewable energy has drawn the atten-
tion to the strong potential of ocean-wave energy [1]. Among the
different technologies that have been studied and tested for the
conversion of ocean-wave energy into electrical energy, oscillat-
ing water column (OWC) systems represent a reliable and simple
solution [2]. Figure 1(a) presents a schematic of an OWC system,
composed of two main units: an open chamber, partially sub-
merged under the sea free surface, where the water movement
induces an alternative movement of a column of air, and a turbine
driven by the air flow. The periodic inversion of the air flow in the
OWC chamber requires a system that is capable of maintaining
the same direction of rotation regardless of the direction of the air
flow. The Wells turbine, patented by Wells in 1970s [3], is charac-
terized by a symmetrical blade profile staggered at 90 deg with
respect to the axis of rotation, and it represents a solution of self-
rectifying air turbine (Fig. 1(b)), which ensures similar perform-
ance during outflow (air flowing out of the chamber) and inflow
(air flowing into the chamber).
Owing to their simplicity of construction and reliability, Wells
turbines have been widely studied as power-take-offs in OWC
systems, both experimentally [4–6] and numerically [7–9]. In
recent years, a number of authors [10–13] have analyzed the tur-
bine performance from a second-law point of view, trying to esti-
mate the turbine efficiency based on an exergy balance,
evaluating the entropy produced by the air flow through the
machine.

1
Corresponding author.
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received July 14, 2020; final manuscript
received December 27, 2020; published online February 9, 2021. Assoc. Editor: Fig. 1 Working principle of a OWC system and a Wells turbine:
Arindam Banerjee. (a) OWC system and (b) Wells turbine

Journal of Fluids Engineering Copyright V


C 2021 by ASME MAY 2021, Vol. 143 / 051205-1
One of the limitations of Wells turbines comes from its sym- Several authors [24,25,28] use the formulation in Eq. (2), as it
metrical profile, which leads to a limited lift at low incidence contains parameters that can be more easily measured in experi-
angles and the occurrence of stall at high flow coefficients [14], ments and it is representative of an aerodynamic efficiency of the
which reduces the operating range of the machine. A number of turbine rotor
papers focused on the improvement of the Wells turbine perform-
ance not only by extending its operating range [15–17], but also TX T =r X1
by increasing the efficiency and output torque. Most of the pro- gad ¼ ¼   ¼ (4)
Dp Q Dpp r 2  r 2 Va =Xr Z /
posed performance improvements are based on a first-law tip hub
approach [18–20]; more recently, some authors investigated new
design solutions by looking at the entropy production inside the where X and Z are axial and tangential forces.
machine. Shehata et al. [10] investigated the entropy generation of Other authors [5,6,8,23,29] use Eq. (3). It is the ratio between
several two-dimensional blade profiles frequently used in Wells the rotor useful power and the power for an ideal (isentropic) pro-

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


turbines in unsteady flow conditions. Nazeryan and Lakzian cess between the same initial and final total pressures (i.e., the
[12,21] compared the entropy generated by a rotor with constant available power), and it represents a total-to-total isentropic
and one with variable thickness blades, showing a reduction in efficiency [27].
entropy production and therefore increased performance in the lat- If the exhaust kinetic energy is entirely wasted, as in case of the
ter. Shehata et al. [22] showed the effect of passive flow control Wells turbine [26,30], the total-to-static isentropic efficiency defi-
on the onset of stall for a symmetrical two-dimensional profile nition is more appropriate than the total-to-total definition
and the implications on the second-law efficiency. reported in Eq. (3). Considering the subscripts 1 and 2 for the con-
In this paper, the common definition for the efficiency of a ditions upstream and downstream the rotor, respectively, the total-
Wells turbine [8,23–25] is compared with a different evaluation to-static efficiency reads as follows [26]:
that has been recently introduced and is based on a second law
analysis [10–12,21,22]. A theoretical derivation and numerical TX
gts ¼ (5)
results from computational fluid dynamics (CFD) simulations are ðpt1  p2 ÞQ
used to demonstrate how the two definitions lead to equivalent
results, under the assumptions used for this type of turbomachines. Even though this formulation is more representative of the
Different models for turbulence closure have been compared, not energy conversion process in a Wells turbine, it has (seldom) been
only with respect to their global parameters predictions (torque adopted both in experimental and numerical analyses and there-
and pressure drop across the rotor) but also looking at the local fore it is not used in this paper.
flow quantities calculations (i.e., the entropy production). Five dif-
ferent rotor geometries have been compared, highlighting the
2.2 Second-Law Analysis. In some recent works [10,12,13,
effect of blade thickness and rotor solidity on turbine performance
21,22], the efficiency presented in Eq. (3) has been referred to as
and entropy production.
first-law efficiency, in contrast to a second-law efficiency derived
This paper is organized as follows: Sec. 2 presents the defini-
from an exergy analysis.
tions of first- and second-law efficiencies for Wells turbines.
Considering a steady-flow and adiabatic process, the exergy
Section 3 describes how the intermediate quantities required to
balance for an open system can be written as follows [31]:
evaluate the turbine efficiency can be calculated from CFD
results. Section 4 introduces the numerical approach adopted in
E_ x;in  E_ x;out ¼ W_ þ Tt;ref S_ G (6)
CFD simulations. Section 5 presents the results obtained under
different operating conditions, for several turbulent models, while
Sec. 6 analyses the performance of a number of rotor geometries. where the net exergy flux per unit time (E_ x;in  E_ x;out ) is equal to
Finally, Sec. 7 draws the conclusions of this work. the sum of the rate of exergy due to the useful work (W_ ), and the
lost exergy per unit time (Tt;ref S_ G ). S_ G represents the entropy gen-
eration rate inside the control volume and Tt;ref is a reference tem-
2 Wells Turbine Efficiency perature. A second-law efficiency can be defined as the ratio
2.1 First-Law Efficiency. Performance of Wells turbines is between the useful work and the net exergy flux, i.e.,
usually presented [26], in terms of the following nondimensional
parameters: the flow coefficient /, the torque coefficient T  , and W_ W_ TX
gII ¼ ¼ ¼ (7)
static or total pressure drop coefficients, p and pt , respectively, E_ x;in  E_ x;out W_ þ Tt;ref S_ G T x þ Tt;ref S_ G
Va T Dp Dpt
/¼ T¼ p ¼ pt ¼ (1)
Xrtip qX2 rtip
5
qX2 rtip
2 qX2 rtip
2
2.3 Linking First- and Second-Law Efficiency. For a machine
evolving an ideal gas, the Gibbs’ equation integrated between the
where Va is the (spatially averaged) axial flow velocity in the tur- initial and final total conditions reads
bine duct, q is the air density (assumed constant and equal to the
ambient density), X is the angular velocity of the rotor, rtip is its ð2 ð2 ð2
tip radius, T is the torque, p and pt are the static and total pres- dpt
dht ¼ þ Tt ds (8)
sures, and D represents the difference between conditions at the 1 1 qt 1
two sides of the rotor.
The efficiency of a turbine is defined as the ratio between the The term on the left-hand side represents the actual work of the
useful output work and the available energy of the turbine. Two machine, w, the first term on the right-hand side is the polytropic
different formulations can be found in the scientific literature for work, wpol, and the last term represents the lost work, wlost. Under
Wells turbines, both entailing an assumption of incompressible the assumption of constant heat values and for a turbine
flow [27]
ð2
TX jW_ j
gad ¼ (2) jwj ¼ ¼  dht ¼ cp ðTt1  Tt2 Þ (9)
Dp Q m_ 1

ð2
TX jW_ pol j dpt
gtt ¼ (3) jwpol j ¼ ¼ (10)
Dpt Q m_ 1 qt

051205-2 / Vol. 143, MAY 2021 Transactions of the ASME


In terms of powers, observing that for low speed flows qt can
be approximated with the static density, q, one has
ð2
dpt m_
jW_ pol j ¼ m_ jwpol j ¼ m_  ðpt1  pt2 Þ (18)
1 qt qt
ð2
W_ lost ¼ m_ wlost ¼ m_ Tt ds  mT
_ t;ref ðs2  s1 Þ (19)
1

ð2
jW_ j ¼ m_ jwj ¼ m_ dht ¼ mc
_ p ðTt1  Tt2 Þ ¼ T X (20)
1

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


In addition, the lost work can be written in terms of the entropy
generated inside the domain, following Gauss’ divergence theo-
rem [32]:
ð
_ t;ref ðs2  s1 Þ ¼ Tt;ref
mT qr dV ¼ Tt;ref S_ G (21)
CV

where r ¼ dsV =dt represents the entropy generation rate per unit
mass, and CV a control volume enclosing the turbine.
Putting together Eqs. (18), (19), and (20) with the definitions of
Fig. 2 Representation of specific works on the Tt–s plane first- and second-law efficiencies, Eqs. (3) and (7), it follows that:

TX TX
ð2 gII ¼  ¼ gtt (22)
W_ lost T X þ Tt;ref S_ G Dpt Q
wlost ¼ ¼ Tt ds (11)
m_ 1
Equation (22) is a direct consequence of the equivalence of
jwj ¼ jwpol j  wlost (12) polytropic and isentropic work for a low pressure ratio process. In
light of the growing interest in evaluating the performance of
We can also define an isentropic work, referring to a transfor- Wells turbines with the second-law efficiency, the validity of this
mation at constant entropy between the initial and final total approximation will be verified in this paper with several turbu-
pressures lence models and for different Wells turbine rotors.
ð 2is
jW_ is j 3 Entropy Calculation From Computational Fluid
jwis j ¼ ¼ dht ¼ cp ðTt1  Tt2;is Þ (13)
m_ 1 Dynamics (RANS) Simulations
The quantities required to estimate the efficiency of a Wells tur-
Figure 2 reports graphically the difference between polytropic,
bine can be calculated using numerical simulations, i.e., by solv-
isentropic, actual, and lost specific works for a machine evolving
ing the governing Navier–Stokes (NS) equations numerically.
a compressible flow.
These quantities include the turbine torque T , the static and total
The isentropic work represents the maximum work that a tur-
pressure drops Dp and Dpt , the volumetric flow rate Q, and the
bine can exchange with the fluid (i.e., the available energy per
entropy generation rate S_ G .
unit time), while the polytropic work is the sum of actual work
As a direct numerical simulation of the NS equations for most
and lost work (due to friction). In a turbine, the polytropic work is
flows of industrial interest is still beyond the capability of current
always greater than the isentropic work, with the difference being
computers, the most common and practical approach is the solu-
the heating work gain [27]
tion of the Reynolds–averaged Navier–Stokes (RANS) equations,
jwpol j ¼ jwj þ wlost > jwis j (14) which are derived by replacing the instantaneous flow variables in
the NS with the sum of a mean value and a fluctuating component
Conversely, the isentropic efficiency gis ¼ jwj=jwis j is always with zero mean value (u ¼ u þ u0 ).
greater than the polytropic efficiency gpol ¼ jwj=jwpol j. The latter For a compressible flow, assuming negligible heat transfer as
is more representative of the inefficiencies in turbomachine, as it common in turbomachinery applications, the RANS equations of
does not take into account the heating work gain and only consid- conservation of mass, momentum, and energy read as follows
ers the lost work. [33]:
For low pressure ratio machines (as in the case of Wells tur- 8
bine) ðTt1  Tt2;is Þ  Tt1 , and the following approximations can > @q  
>
> þ r  qV ¼ 0
be made >
>
>
> @t
>
>
wpol  wis ) gpol  gis (15) <  
@qV
þ r  qV  V þ qV0  V0 ¼ rp þ r  P (23)
ð2 > @t
>
>
>
Tt ds  Tt;ref ðs2  s1 Þ (16) >
> @qh
>
>   Dp
1 >
: þ r  qV h ¼ þ P : rV þ P0 : rV0
@t Dt
provided that Tt;ref is chosen appropriately (i.e., in the range of
temperatures involved in the transformation). Under these where V is the velocity vector, h is the static enthalpy, and P is
assumptions the deviatoric stress tensor, that for a Newtonian flow can be
expressed as the sum of the contributions due to the strain rate ten-
jwj  jwpol j  Tt;ref ðs2  s1 Þ (17) sor rS V and the volumetric tensor ðr  VÞI, each one multiplied

Journal of Fluids Engineering MAY 2021, Vol. 143 / 051205-3


by a constant, i.e., the molecular viscosity l and the bulk viscosity neglecting turbulent convection and diffusion across the bounda-
k, respectively, ries, and integrating in an appropriate control volume CV
ð ð
P ¼ lðrV þ rVT Þ þ kðr VÞI ¼ 2lrS V þ kðrVÞI (24) @
S_ G  q sV  ndA þ q sdV (30)
A @t CV
In Eq. (23), the prime symbol denotes fluctuating quantities,
and the overbar denotes time-averaged flow quantities. Density In Eq. (30), A represents the boundary of the control volume
fluctuations, i.e., q0 , have been neglected as they start to affect tur- CV, and s the specific entropy which can be evaluated under the
bulence around a Mach number of 1 [34]. In the presence of sig- assumption of ideal gas, as recalled in Sec. 2.3.
nificant density fluctuations, the Favre-averaged NS equations can
be used instead [35].
The terms which represent the effects of turbulent fluctuations 3.1 Evaluation of Entropy Production With Different Tur-

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


on the mean flow are not resolved by the RANS approach and bulence Models. As mentioned in Sec. 3, turbulent quantities are
need to be modeled to allow the solution of the system of equa- not directly available in RANS approaches and need to be mod-
tions. The turbine torque T can then be calculated by integrating eled. Different turbulence closure methods are available in CFD
pressure and viscous stresses on the turbine blade, while the pres- solvers, the most common being the ones based on the so-called
sure drops require the evaluation of the (mass-flow) averaged Boussinesq’s hypothesis, which assumes a linear dependency
pressure on surfaces appropriately defined upstream and down- between Reynolds’ stress and strain tensors (Eq. (28)). The most
stream of the turbine. The evaluation of the entropy generation famous ones are k  e and k  x models, which derive the turbu-
rate S_ G requires more attention, as an entropy equation is not gen- lent viscosity lT based on two additional partial differential equa-
erally solved in CFD programs, as it would make the system of tions, for turbulent kinetic energy (k) and its rate of dissipation
equations in Eq. (23) overdetermined. An entropy equation can be (e), and for k and its specific dissipation rate (x), respectively. Of
derived by linking Gibbs’ relation with the momentum and energy particular interest for this work is the equation of conservation for
equations, as described in Refs. [32] and [36] the turbulent kinetic energy, which is reported in the following
equation:
     
@q s  
þ r  qVs þ r  qV0 s0 ¼ qrV;mf þ q rV;T (25) @qk 0 0 0 0 1 0 0 0
@t þ r  qkV ¼ r  p V þ P  V  qV  V  V
@t 2
where the right-hand side represents the entropy production rate  P0 : rV0  qV0  V0 : rS V (31)
per unit mass due to fluid flow: in particular, rV;mf is the contribu-
tion due to the mean flow and rV;T is the one due to turbulent fluc- where the last two terms on the right-hand side of the equation
tuations. The two terms on the right-hand side of Eq. (25) are represent the rates of dissipation and production of k, respectively.
defined as follows, neglecting the effect of temperature fluctua- The former, referred to with the symbol e, is present also in the
tions on viscous entropy production as in Refs. [36] and [37]: energy equation (Eq. (23)), as the dissipated turbulent kinetic
energy is transformed into heat. The production of k can be
1 expressed following the Boussinesq’s hypothesis in Eq. (28):
q rV;mf ¼ ðP : rV Þ (26)
T
P0 : rV0 ¼ qe (32)
1 
q rV;T ¼ P0 : rV0 (27) 2 2
T qV0  V0 : rS V ¼ 2lT ðrS V Þ  qkðr  V Þ (33)
3
Only the terms containing mean quantities are solved (and
hence available) in a RANS approach. All terms involving fluctu- Using Eq. (32), the viscous dissipation in Eqs. (26) and (27) can
be calculated as follows:
ating quantities (qV0  V0 and P0 : rV0 ) need to be modeled.
The most common approaches are linear eddy viscosity models, 1 
based on the so-called Boussinesq’s hypothesis T rV ¼ T rV;mf þ T rV;T ¼ P : rV þ P0 : rV0
q
  2  
T 1 2 2
qV0  V0 ¼ PR ¼ lT rV þ rV  qkI ¼ 2lðr V Þ þ kðr  V Þ þ qe
S
(34)
3 q
S 2
¼ 2lT r V  qkI (28) Many authors [38–40], assuming a local equilibrium between
3
turbulent entropy production and dissipation, i.e.,
where PR is the Reynolds’ stress tensor, lT is the turbulent viscos- qe ¼ 2lT ðrS VÞ2  23 qkðr  VÞ, introduce the following
ity, and k is the turbulent kinetic energy per unit mass approximation:
(k ¼ 12 ðqu02  qv02  qw02 Þ). The term lt is usually modeled   
adding additional transport equations. The quantity ðP0 : rV0 Þ 1 S Þ2 2
T rV ¼ ð
2ðl þ lT Þ r V þ kr  V  qk r  V ð Þ (35)
represents the dissipation of turbulent kinetic energy into heat and q 3
is usually referred to with the symbol e. Finally, the global entropy
generation per unit time (S_ G ) can be estimated by integrating the The last two expressions (Eqs. (34) and (35)) are alternative
viscous dissipation in the domain of interest methods for calculating the entropy production due to viscous dis-
sipation per unit mass, and the choice between the two methodolo-
ð gies depends on the selected turbulence closure model:
S_ G ¼ qðrV;mf þ rV;T ÞdV (29)
CV (1) In k  e and k  x models, both approaches can be adopted
because in the former e is directly available, while in the
This approach for the calculation of entropy production is latter, it can be calculated as a function of the turbulent
referred to as a direct method in Ref. [36]. Alternatively, the same kinetic energy k and specific dissipation rate x, i.e.,
authors suggested the use of an indirect method, which is derived e ¼ xkb , where b is a model constant which depends on
from the equation of conservation of entropy (Eq. (25)), the specific implementation. An important consideration

051205-4 / Vol. 143, MAY 2021 Transactions of the ASME


Table 1 Geometric and operating data

Rotor tip diameter (mm) 300


Rotor hub diameter (mm) 210
Tip clearance (mm) 1
Chord length (mm) 90
Sweep ratio 0.417
Number of blades 5, 6, 7
Blade profiles NACA0012, NACA0015,
NACA0020
Solidity at tip radius 0.48, 0.57, 0.67
Rotational speed (rpm) 2500
Operating frequency (s1) 1/6
Nondimensional frequency ðpfcÞ=U 1:2  103

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


Reynolds’ number 2  105 (based on rotor tip radius and
chord length)

has been proposed in Ref. [39], where the authors note that
the second approach (Eq. (34)) is more reliable, because in
RANS approaches, e is only used as an intermediate quan-
tity to calculate lt and PR , which interact with the mean
flow through the momentum and energy equations.
(2) In Spalart–Allmaras (S–A) models only the second Fig. 3 Computational domain and mesh for simulations: (a)
approach can be used, as the method solves directly a trans- computational domain and (b) blade and hub mesh
port equation for kinematic eddy viscosity ~ , which is
related to lT through the following expressions:
[9,13,21,48–51]. The OWC chamber is not included in the domain
ð~ = Þ3 (as done, for example, in Refs. [52–54]), because its main effect is
lT ¼ q~
 fv1 fv1 ¼ (36) to cause a delay between the movement of the water level in the
ð~ = Þ3 þ C3v1
chamber and the mass-flow in the turbine duct, without significant
where fv1 is the viscous damping function. In addition to the stand- modifications to the turbine performance.
ard formulation of the k  e model, the realizable (REAL) k  e Uniform inlet boundary conditions have been used for velocity,
formulation has also been considered. It is a newer implementa- total temperature, and turbulent quantities. The velocity has been
tion [41] that differs from the original for a new formulation of set to a value, fixed in steady (fixed-/) simulations and sinusoi-
the turbulent viscosity and a new transport equation for the dissi- dally variable in time in dynamic simulations, chosen to obtain
R
pation rate e. The FLUENTV User’s Guide [42] states that the k  e the required value or range of flow coefficients (see Eq. (1)).
REAL model provides superior performance for flows involving Figure 4 shows a typical variation of flow coefficient /, where Tw
rotation, boundary layers under strong adverse pressure gradients, represents the wave period. The inlet total temperature has been
separation, and recirculation. set to 288 K and the turbulent quantities are calculated by the
solver based on the values set for turbulent intensity and length
scale (2% and 7% of the blade height, as suggested in Ref. [42]).
4 Methodology A uniform value of static pressure has been specified at the outlet,
The turbine geometry and operating conditions simulated in while periodic boundary conditions have been used at the two
this work are the ones presented in the experimental work from sides of the passage of the computational domain (Fig. 3). Inlet
Setoguchi et al. [43]. The main details are summarized in Table 1. and outlet are inverted for negative values of the flow coefficient.
The domain for the numerical simulations is reported in A multiblock structured grid has been used to discretize the vol-
Fig. 3(a): it is a straight duct representing a single blade passage ume, with a C-grid around the blade able to capture the boundary
of the turbine, with periodic boundary conditions at the two sides. layer flow and an H-grid in the rest of domain, see Fig. 3(b).
Simulating a single passage of a tubomachinery’s blade row (or Four different turbulence closure models have been compared:
even multiple single passages from different blade rows, with an the k  x shear stress transport (SST), the standard (STD) and
appropriate treatment of the interrow interface) is a common prac- REAL k  e and the Spalart–Allmaras (S–A) models. Numerical
tice in turbomachinery RANS simulations [44–46] when the simulations have been conducted using the commercial CFD soft-
hypothesis of periodic flow with respect to the blade pitch is valid, ware ANSYS FLUENT 17.0. The SIMPLEC algorithm has been used
i.e., in the absence of flow structures larger than the blade pitch. for the pressure–velocity coupling, a second-order centered
This approximation has been often used in Wells turbine simula-
tions [8,9,21,47]. A comparison between the simulation of the
Wells turbine’s full rotor and of a single passage was conducted
by the authors, showing a maximum difference in the performance
coefficients of less than 0.1%. The interaction between stationary
and moving parts has been modeled using a “frozen rotor
approach,” also referred to as the “multiple reference frame”
model in the FLUENT User’s Guide [42]. The model can be used for
turbomachinery applications in which rotor–stator interaction is
relatively weak, and the flow is relatively uncomplicated at the
interface between the moving and stationary zones. This seems to
be a good approximation for the present case study, where the
interaction between moving parts is only at the hub and casing of
the duct where the turbine is housed. The same model has been Fig. 4 Inlet boundary condition for dynamic simulations, in
used in the majority of previous CFD analyses of Wells turbines terms of flow coefficient /

Journal of Fluids Engineering MAY 2021, Vol. 143 / 051205-5


scheme for pressure and viscous terms and a second-order upwind
scheme for convective terms. Multiple reference frames have
been adopted to simulate the interaction between stationary and
rotating volumes. Steady (fixed-/) simulations have been run
with a time-dependent approach (and constant boundary condi-
tion) with a time-step of 104 until convergence of the monitored
quantities (torque and pressure drops) were obtained (within 0.1%
in the last 1000 time steps), while time-dependent simulations
were run for three wave periods, with a time-step sufficient to
obtain results independent from the temporal discretization.
Results from the last two periods differed always by less than
0.1%; hence, periodic convergence has been considered achieved.

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


A first-order implicit temporal discretization approach was used, Fig. 5 Grid convergence of nondimensional parameters eval-
with five subiterations per time-step. The default settings of ANSYS uated with five grid sizes: (a) torque coefficient and (b) pressure
FLUENT have been maintained: at every subiteration, two sweeps of drop coefficient
the algebraic multigrid with a maximum of 40 levels are allowed,
with the Gauss–Seidel smoother. An explicit convergence crite- Table 2 Grid convergence estimation using grids
rion was not set in time-dependent simulations, but rather the B, C, D
number of subiterations was kept fixed, and the time-step size
reduced until results were independent from its value. Maximum P T
residuals were found to be of the order of 104 for appropriate N1 ¼ NB 631,200
values of the time-step. Increasing the number of subiterations N2 ¼ NC 1,836,480
would have achieved the same effect, at least to a certain time- N3 ¼ ND 5,360,165
step size, as shown in Ref. [55]. r12 1.43
All simulations were run on dual processor 3.30 GHz 8 core Intel r23 1.43
Xeon E5-2667 v2 CPUs. Each time-step required about 3 s, and S1 =Sext 0.9874 0.9963
complete periodic simulations required as long as three months. S2 =Sext 0.9950 1.0006
S3 =Sext 0.9980 0.9994
p 2.57 3.62
4.1 Verification and Validation. The choice of the spatial GCI1 1.048% 0.180%
discretization has been made following a grid convergence study, GCI2 0.420% 0.029%
with the k  x model for turbulence closure, following the guide- GCI3 0.168% 0.029%
lines given in the JFE Editorials Policy Statement [56] and UStern — 0.220%
expanded in several articles [57,58]. These are based on the use of
the Richardson extrapolation, which involves the solution of the (each grid has about 2.9 times the cells of the previous one), the
numerical problem on three grids with increasing size. Defining central being the one that has been used for the rest of the
h1, h2, and h3 the representative grid dimensions (these can be cal- calculations.
culated as ðV=NÞ1=3 , with V the volume of the fluid domain, and N It is clear that none of the grids presents a large variation with
the number of cells), and S1, S2, and S3 the solutions obtained with respect to the finest one (the maximum deviation is less than 1.5%
the three grids for a quantity of interest, the apparent order of con- on the coarsest grid for the pressure drop coefficient). The grid
vergence can be calculated as follows: used in this study, which has about 2  106 cells, produces results
! that are extremely close to the ones obtained with the finest grid
p
1 r12 s (17  106 cells), producing errors of less than 0.3% and 0.05% of
p¼ j logðje12 =e23 jÞj þ log p (37)
logðr23 Þ r23  s the most accurate prediction, for p and T  , respectively. It is
interesting to note how, while the pressure drop coefficient exhib-
where s ¼ signðe12 =e23 Þ; eij ¼ ðSi  Sj Þ is the difference between its a monotonic convergence, the torque coefficients’ convergence
the solution obtained with two grids of different sizes, and rij ¼ hi =hj is oscillatory. Tables 2 and 3 present the evaluation of apparent
the ratio of grid representative sizes. If s > 0 the convergence is convergence error and evaluation uncertainties, for two sets of
monotonic, while s < 0 might be an indication of oscillatory con- grid triplets (B-C-D and C-D-E, respectively). The apparent con-
vergence. A negative value for p might be an indicator of diver- vergence order calculated with the first set of grids is similar to
gence (or that the grids do not lie within the asymptotic range). the theoretical order of the method, which suggests that this grids
The asymptotic value can be estimated as follows: are in the asymptotic convergence range. The value obtained with
p
r23 S3  S2
Sext ¼ p (38) Table 3 Grid convergence estimation using grids
r23 1 C, D, E

Regarding the uncertainty linked to the evaluation, Celik et al. P T


[58] suggest using a grid convergence index GCI defined as
follows: N1 ¼ NC 1,836,480
N2 ¼ ND 5,360,165


Sext  Sj
N3 ¼ NE 17,376,540
GCIj ¼ 1:25


(39) r12 1.43
Sext
r23 1.43
S1 =Sext 0.9964 1.0006
while [57], in the presence of oscillatory convergence, suggest S2 =Sext 0.9995 0.9994
using half the range of variability in the data S3 =Sext 0.9999 1.0003
p 5.37 0.69
1 GCI1 0.062% 0.244%
UStern ¼ jmaxðSj Þ  minðSj Þj (40)
2 GCI2 0.009% 0.244%
GCI3 0.001% 0.133%
Figure 5 presents the convergence of p and T  for a flow coef- UStern — 0.061
ficient / ¼ 0:16, using five grids with different number of cells

051205-6 / Vol. 143, MAY 2021 Transactions of the ASME


the last set of grids suggests that mesh independence has being of turbulent kinetic energy (as explained in Sec. 3), and the results
reached. have been expressed in nondimensional form, as follows:
The validation has been conducted on the selected grid, i.e.,  
grid C, which has nondimensional wall distance (yþ) below unity. Ð ðrS V Þ2 þ kðr  V Þ2 þ qe dV
CV
2l
Figure 6 reports the results obtained for the nondimensional coeffi-
_ ¼
KS;d (41)
cients of torque and pressure drop for steady calculations at different 1 3 5
2 qin X rtip
flow coefficients, for the turbine with the highest number of blades
(z ¼ 7, i.e., with a rotor tip solidity equal to 0.67) and NACA0020 h   i
profile, where the largest gradients of flow quantities are expected, in
Ð 2 ðl þ l T ÞðrS V Þ2 þ kr  V  2 qk ðr  V Þ dV
CV 3
comparison with experimental data. The latter present the famous _ ¼
KS;g 1 3 5
hysteretic loop due to the capacitive effect of the OWC chamber 2 qin X rtip
[52,54,59]. Numerical results lie in the middle of the hysteresis loop, (42)

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


therefore attesting the appropriateness of the numerical results.
Figure 7 reports the effect of the time-step size on the torque where KS;d_ is the nondimensional entropy production rate in the
and the entropy generation coefficients. Five time-step values have control volume, evaluated using the turbulent kinetic energy dissi-
been compared running 3 periods of oscillation, with an amplitude pation defined in Eq. (34), and KS;g
_ is the nondimensional entropy
for the sinusoidal inlet velocity appropriate to produce a maximum production rate in the control volume, evaluated using the turbu-
flow coefficient / of 0.23, as in the experiment [43]. Only the last lent kinetic energy generation defined in Eq. (35). It should be
period is reported, for brevity. The nondimensional entropy genera- noted that in the presence of an incompressible flow, the second
tion has been evaluated using both the dissipation and the generation term in the round brackets of Eq. (41) and the last term in the
square brackets of Eq. (42) are null. In the current analyses, given
the low relative Mach number in the vicinity of the blade (about
0.3) these terms never accounted for more than 0.02% of the total
entropy production. The authors preferred to treat the flow as
compressible for generality, as the derivation in the current form
can be applied also for higher Mach number machines [60,61].
The analysis highlights how time steps larger than
Tw =16; 000  4  104 s produce spurious phase errors [56],
which manifest in a false delay, previously erroneously inter-
preted as an aerodynamic hysteresis of the turbine [52,54,59,62].
This effect is reported in Fig. 7, where it is evident how hysteretic
effects disappear when sufficiently small time-step sizes are used
(below Tw =16; 000  4  104 s). Reducing the time-step size
Fig. 6 Grid (C) validation with respect to the experimental data has a similar effect to increasing the number of subiterations: [55]
(dotted line): (a) non-dimensional torque and (b) non-dimen- have shown how it is the total number of subiterations per cycle
sional static pressure drop (i.e., number of time steps per cycle times number of subiterations
per time-step) that influences the temporal convergence. For the
selected grid and working conditions, the maximum value of the
Courant number (CFL) is proportional to the time-step size and
ranges from 6400 (for dt ¼ Tw =1000) to 25 (for
dt ¼ Tw =256; 000). Performance during inflow (negative flow
coefficients) and outflow (positive flow coefficients) does not
present significant differences. The absence of dynamic effects is
in agreement with the large literature on oscillating lifting surfa-
ces [63–65], where significantly larger nondimensional frequen-
cies are required to produce an appreciable hysteresis [66,67].
A similar analysis, Fig. 8, has been conducted in the presence
of a temporal profile for the inlet velocity with a larger amplitude
(/max ¼ 0:325), sufficient to lead to blade stall. The results of the
analysis show how an even smaller temporal discretization is
required to achieve results independent from the time-step (about
1:25  105 s), with a corresponding value of the maximum CFL
of about 12.5. In the presence of stall, a small hysteretic loop is
present, which is caused by the fact that the boundary layer reat-
taches to the blade surface for a flow coefficient (and therefore an
angle of attack) smaller than the one leading to stall during the
acceleration phase. This phenomenon is not necessarily linked to
the dynamic operating conditions, as a static stall hysteresis is docu-
mented for many lifting surfaces [68,69], while significantly larger
nondimensional frequencies are required to produce appreciable
effects on the performance [64,65]. In any case, the value of the
performance parameters after reattachment, during deceleration, is
indistinguishable from the one attained during acceleration.

Fig. 7 Verification of the temporal discretization using five differ- 5 Results—Comparison Among Different Turbulence
ent time-step sizes with a maximum flow coefficient /max 5 0.23: (a) Models
non-dimensional torque history (final period), (b) Non-dimensional
torque as a function of the flow coefficient, (c) non-dimensional In this paragraph, simulations with different turbulence models
entropy generation (final period), and (d) non-dimensional entropy are reported for the same rotor geometry, i.e., the rotor with six
generation as a function of the flow coefficient blades (rotor tip solidity equal to 0.57) and NACA0015 blade

Journal of Fluids Engineering MAY 2021, Vol. 143 / 051205-7


Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021
Fig. 8 Verification of the temporal discretization using five dif- Fig. 9 Nondimensional performance of the rotor with
ferent time-step sizes with a maximum flow coefficient NACA0015 blade profile and z 5 6 as a function of flow coeffi-
/max 5 0.325: (a) non-dimensional torque history (final period), cient / estimated with different turbulence closure models: (a)
(b) non-dimensional torque as a function of the flow coefficient, non-dimensional torque, (b) non-dimensional total pressure
(c) non-dimensional entropy generation (final period), and (d) drop, (c) non-dimensional static pressure drop, and (d) non-
non-dimensional entropy generation as a function of the flow dimensional entropy production
coefficient

profile. Two periodic operating conditions have been simulated:


one with a maximum flow coefficient /max ¼ 0:23, not sufficient
to produce stall, and one with a /max ¼ 0:345, enough to cause
deep stall of the rotor at least with some turbulence models.

5.1 Simulations With /max 50:23. Figure 9 reports the non-


dimensional parameters commonly adopted in Wells turbines
characterization with respect to the flow coefficient, as calculated
in Eq. (1). Positive values of the flow coefficient refer to the out-
flow, while negative ones refer to the inflow phase.
Using different turbulence models does not affect the prediction
of the rotor performance and their estimated trends are substan-
tially overlapping (see Fig. 9) and matching the experimental data
as reported in Fig. 6. A small overprediction of the pressure coef- Fig. 10 Mean nondimensional entropy production rate eval-
ficients, i.e., pt and p , can be observed in the simulations made uated in a smaller (a) and larger (b) control volume
with the k  e models, while the comparison of k  x and S–A
does not highlight differences. The torque is well predicted by all
models, and no significant differences can be observed. dissipation of turbulent kinetic energy leads to the same results
On the contrary, the nondimensional entropy generation, only if the calculation is made in a large enough control volume
Fig. 9(d), shows different values depending on the turbulence (control volume B), while an appreciable difference exists when
model selected. In particular, k  e models estimate a higher the boundaries of the control volume are too close to the rotor
entropy production with respect to the other models. In order to (control volume A), mainly due to the neglection of the wake,
better understand these large differences, the entropy generation where the turbulent energy dissipation is larger than its produc-
calculations have been reported in Fig. 10: the mean values over a tion. In the Spalart–Allmaras model, as explained in Sec. 3.1, only
cycle of both KS;d _ and KS;g_ (defined in Eqs. (41) and (42)) are the production of turbulent kinetic energy is available.
reported, for two different control volumes, a smaller one going Even more interesting it is to observe the differences between
from half a chord upstream to half a chord downstream of the the turbulence models selected for this study. The k  e models
rotor (A), and a larger one enclosing all computational domain predict an entropy generation rate significantly larger than the
(B) (eight chords upstream and downstream of the rotor), as indi- k  x model, by about 30%, in the larger control volume. The
cated in Fig. 3(a). The results from control volume A are reported Spalart–Allmaras prediction is lower, but still larger than the one
in Fig. 10(a), the results from control volume B are reported in given by the k  x model, by about 15%. This is due to the dif-
Fig. 10(b). ferent formulation of the turbulent kinetic energy production and
From the results presented in Fig. 10, the evaluation of the destruction terms, which are strictly related to the entropy pro-
(nondimensional) entropy production using the generation and duction, as shown in Sec. 3. The k  x SST models, in

051205-8 / Vol. 143, MAY 2021 Transactions of the ASME


Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021
Fig. 11 Total exergy calculations averaged on a cycle for differ-
ent turbulence closure models

particular, use a low-Reynolds correction in the boundary layer,


which is particularly effective in the presence of low Reynolds Fig. 12 Close-up view of the nondimensional entropy produc-
number flows, as in this case [70,71]. This has an effect both not tion at several blade span positions for the maximum flow coef-
only on the amount of entropy generated but also on the capabil- ficient / 5 0.23 calculated using k–x SST (left) and k–e STD
ity of the model to predict flow separation and stall at low Reyn- (right) models: (a) 20% span, (b) 50% span, and (c) 80% span
olds numbers.
Figure 11 reports the total exergy in a cycle of period Tw,
defined in Eq. (6), using three approaches, which, as explained in
Sec. 2.3, should in theory lead to very similar results
ð
EP ¼ Dpt Q dt
T
ðw
EE ¼ ðmc
_ p DTt  mT_ t;ref DsÞdt (43)
T
ðw
ES ¼ ðT X þ Tt;ref S_ G Þdt
Tw

The last two methods are the indirect and direct approaches
described by Herwig [36]. The exergy calculation ES has been Fig. 13 Rotor efficiencies averaged on cycle and calculated with
evaluated both considering the generation, ES;g , and dissipation, different turbulence closure models: (a) Non-dimensional torque,
ES;d , of turbulent kinetic energy (Eqs. (41) and (42)). (b) non-dimensional total pressure drop, (c) non-dimensional
It is interesting to note how the traditional approach (EP) and static pressure drop, and (d) non-dimensional entropy production
the indirect method lead to same results, and this is a confirmation
of the validity of the assumptions made in Sec. 2.3
more evident near the suction side of the blade and near the trail-
(Dpt Q  mc _ p DTt  mT
_ t;ref Ds). On the contrary, the direct method
ing edge, at all spanwise positions from hub to tip.
(which requires the integration of the entropy production in the
The above results have been used to evaluate the turbine effi-
computational domain) based on the turbulent kinetic energy pro-
ciencies defined in Sec. 2. The values reported in Fig. 13 are a
duction (ES;g ) leads to an overestimation of the available energy.
direct consequence of the results in Figs. 9–11. The aerodynamic
This difference is not too significant for the k  x model (about
efficiency gad is lower than the total-to-total efficiency gtt, as the
2%), larger for the other turbulence models, and especially for the
static pressure drop that appears in its denominator is larger than
k  e STD (about 9%). This result is in line with the differences
the total pressure drop that is used to calculate the first-law effi-
encountered when estimating the entropy generation in other
ciency (the former includes the exit dynamic head, which in a
applications: differences as large as 15% are not uncommon [72],
Wells turbine is lost). The second-law efficiency (which theoreti-
and are due to the fact that CFD software do not solve the entropy
cally should be approximately equal to the latter) has a very simi-
equation, which therefore can be not strictly satisfied due to
lar value only when the denominator is calculated using the
numerical errors, as explained in Ref. [73]. Lower discrepancies
indirect method, while it is lower when the direct method is
between the direct method and the other ones can be obtained
selected. A further difference exists depending on whether the
when using the turbulent kinetic energy dissipation (ES;d ). The
entropy generation S_ G is calculated using the dissipation or the
overestimation still remains high for k  e models, but it almost
production of turbulent kinetic energy, being the latter slightly
disappears for the k  x SST model.
larger than the former (see Fig. 10). These differences are smaller
The larger estimation of the available energy from
for the k  x model than for Spalart–Allmaras and k  e models.
Spalart–Allmaras and k  e models is linked to the overestimation
The second-law efficiency calculated using the turbulent kinetic
of local entropy generation (see Fig. 10), which is due to the dif-
energy dissipation with the k  x SST model is remarkably close
ferent treatment of the boundary layer region, assumed fully tur-
to the first-law efficiency.
bulent [70]. On the contrary, the k  x SST model adopts the
standard k  e model only away from the walls and an improved
formulation within the boundary layer [71], where viscous effects 5.2 Simulations With /max 5 0:345. Figure 14 reports the
predominate over turbulent ones. performance parameters of the Wells turbine with an operating
This difference in boundary layer treatment among k  x SST (sinusoidal) cycle with a maximum flow coefficient sufficient to
and k  e STD model is highlighted in Fig. 12, that clearly shows produce deep stall conditions.
higher intensity of nondimensional entropy production in the When comparing the curves in Fig. 14, the k  e model in its
boundary layer region for the k  e model. This overestimation is standard formulation is unable to predict the turbine stall and the

Journal of Fluids Engineering MAY 2021, Vol. 143 / 051205-9


Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021
Fig. 14 Non-dimensional performance of the rotor with
NACA0015 blade profile and z 5 6 as a function of flow coeffi-
cient / estimated with different turbulence closure models: (a)
non-dimensional torque, (b) non-dimensional total pressure
drop, (c) non-dimensional static pressure drop, and (d) non-
dimensional entropy production

performance is always increasing with the flow coefficient (in the


range considered here, i.e., / ¼ 0:345 0:345). The k  e
REAL model predicts the presence of light stall for a / just above
0.3. This is due to the difficulty of k  e models to correctly pre-
dict the separation of the boundary layer for low Reynolds number
flows, in particular in the k  e STD formulation where the bound-
ary layer is considered as fully turbulent [70]. k  x SST and
Spalart–Allmaras model predict the occurrence of stall (at / equal
to 0.28 and 0.32, respectively). After stall, the torque coefficient
drops dramatically, and the entropy generation increases corre-
spondingly. A high frequency oscillation in all performance Fig. 15 Isosurfaces of Q criterion (Q 5 2.25 3 106 s22) colored
parameters can be observed. During deceleration, the reattach- by non-dimensional entropy production: (a) / 5 0.19, (b) / 5
ment of the boundary layer, which corresponds to the exit from 0.23, (c) / 5 0.30, and (d) / 5 0.345
the stalled conditions, happens for a lower flow coefficient (equal
to 0.2 and 0.22 for k  x SST and Spalart–Allmaras models).
This leads to the presence of a hysteretic loop, which does not 6 Results—Comparisons Among Different Rotor
extend to the clean part of the curves, i.e., after reattachment. This
hysteresis is not necessarily caused by dynamic effects, as a static Geometries
stall hysteresis is well documented for many lifting surfaces Five rotor geometries have been compared, with three rotor-tip
[68,69,74,75]. solidities (0.48, 0.57, 0.67) and three blade thicknesses (NACA
In order to better understand the difference between k  x and 0012, NACA 0015, NACA 0020 profiles). All the calculations
k  e models predictions, Fig. 15 shows the vortical structures have been performed using k  x SST model under dynamic
around the blade calculated with the two models. Isosurfaces of (sinusoidal) flow conditions.
Q-criterion, colored by the nondimensional entropy production, Starting from the global performance comparisons as in Sec. 5,
are reported. Figs. 16 and 17 report the well-known nondimensional parameters
The comparisons in Fig. 15 show that the k  e model underes- for the rotors with the same number of blades and with equal
timates the large vortex located near the suction-side (SS) of the blade profile, respectively.
blade, as well as the large area of reversed flow near the trailing In Fig. 16, no significant modifications in performance can be
edge. The tip vortex grows for larger flow-coefficient values until noticed for a different blade profile thicknesses, while Fig. 17
it appears destroyed and a roll-up vortex can be observed on the highlights the strong effect of the rotor solidity on both torque and
same blade side near the trailing edge. The vortical structures pressure coefficient. This is due to the blockage effect exerted by
located near the trailing edge when the blade is stalled are again the rotor on the air flow, directly dependent on the number of
smaller in k  e predictions than in k  x ones. Both models blades [47,76,77].
show the entropy production growing with the flow-coefficient, as Figure 18 compares the averaged nondimensional entropy pro-
expected, with larger values being predicted by the k  e model duction for different rotor geometries. The results are in agree-
especially near the blade surface, as previous observed in Fig. 12. ment with the considerations drawn from the performance

051205-10 / Vol. 143, MAY 2021 Transactions of the ASME


Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021
Fig. 18 Averaged non-dimensional entropy production rate
evaluated near the rotor for different rotor geometries

Fig. 16 Non-dimensional performance of the rotor with z 5 6


and different blade profiles as a function of flow coefficient /:
(a) non-dimensional torque, (b) non-dimensional total pressure
drop, (c) non-dimensional static pressure drop, and (d) non-
dimensional entropy production

Fig. 19 Non-dimensional entropy production at several blade


span positions for the maximum flow coefficient / 5 0.23 and
for rotor with z 5 5 (left), z 5 6 (center), and z 5 7 (right): (a) 20%
chord, (b) 50% span, and (c) 80% span

Fig. 17 Non-dimensional performance of the rotor with


NACA0020 blade profile and different number of blades as a
function of flow coefficient /: (a) non-dimensional torque, (b)
non-dimensional total pressure drop, (c) non-dimensional
static pressure drop, and (d) non-dimensional entropy
production

parameters comparisons, i.e., it is possible to state that the change


in blade-profile thickness does not modify substantially the aver-
aged rotor losses over a cycle and this is confirmed by a very simi-
lar entropy production for all three rotors with equal solidity. On Fig. 20 Isosurfaces of Q criterion (Q 5 2.25 3 106 s22) colored
the contrary, when the rotor solidity increases, the entropy produc- by nondimensional entropy production for different flow condi-
tion also increases and this fact is again related to the blockage tions and for rotor with z 5 5 (left), z 5 6 (center), and z 5 7
effect exerted by the rotor on the flow. (right): (a) / 5 0.19 and (b) / 5 0.23

Journal of Fluids Engineering MAY 2021, Vol. 143 / 051205-11


Finally, the mean rotor efficiencies reported in Fig. 22 allow to
make a better comparison among the different geometries. The
rotors with different blade thicknesses do not present significant
variations in efficiency for the selected rotor solidity, i.e., 0.57: a
small drop in efficiency is reported for the rotor with the largest
thickness, as already reported in the experiments [43]. On the con-
trary, the solidity has a higher impact on the efficiency, and the
two lower solidities represent the best solutions under these flow
conditions. These results are comparable with the ones obtained in
the experiments of Refs. [43] and [76].

7 Conclusions

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


This work attempts to clarify the underlying assumptions, as
Fig. 21 Energy calculations averaged on a cycle for different
well as the methodologies available for estimation of the effi-
rotor geometries: (a) rotors with z 5 6 and (b) rotors with ciency of the Wells turbine, with particular focus on the solution
NACA0020 blade profile of RANS equations. In particular, it is shown that, under the
assumptions typical for this type of turbomachinery, the so-called
first- and second-law efficiencies give almost the same values.
While the former requires the evaluation of the pressure drop
across the machine, the latter can be calculated either by evaluat-
ing the entropy rise across the machine (the so-called indirect
method [36]) or by integrating the local entropy generation rate in
the domain (the direct method [36]). These two measures are theo-
retically equivalent, but some discrepancies can arise in CFD sim-
ulations due to numerical errors, as entropy is not necessarily
conserved in CFD solvers [73].
The results of this work can be summarized as follows:
(1) The first- and second-law efficiencies are equivalent meas-
ures for the efficiency of Wells turbines, at least when the
assumption of incompressibility is valid.
(2) The available power Dpt Q is approximately equivalent to
the net exergy W_ þ Tt;ref S_ G . This is theoretically valid for a
machine evolving an incompressible flow, and has been
Fig. 22 Rotor efficiencies averaged on a cycle for all the tested verified in RANS simulations for a low speed Wells tur-
geometries: (a) rotors with z 5 6 and (b) rotors with NACA0020 bine, for several turbulence models and geometries.
blade profile (3) In numerical simulations, the net entropy flux is not strictly
equal to the entropy generation rate inside the domain
(href S_ gen 6¼ mT
_ t;ref Ds).
The effect of the rotor solidity on the entropy production can be (4) The error arising from the previous point is significantly
better observed in Fig. 19, where the local entropy production smaller with the k  x SST model than with k  e (stand-
field around the blade is shown at different spanwise positions. ard and realizable) models.
The contour plots allow to make a qualitative analysis of the (5) The entropy generation rate can be estimated using either
losses along the blade span. In particular, while low solidity rotors the production or dissipation of turbulent kinetic energy,
(z ¼ 5, 6) show similar amounts of entropy production at all span- and the difference between the two measures is small pro-
wise positions, the highest solidity rotor (z ¼ 7) experiences higher vided that the calculation is made in a large enough control
losses along the whole blade span. This is more evident in the hub volume.
region where the blades are the closest (i.e., the solidity assumes (6) The (local) entropy generation rate provides a useful
its highest value) and the blockage effect determines higher method for identifying the main sources of loss.
losses, as it can be determined by comparing the vortical struc- (7) The blade thickness has a minimal effect on the efficiency
tures in Fig. 20 for two different flow conditions. In fact, near the of the turbine under consideration. On the contrary, the effi-
tip region, the entropy generation field around the three rotors ciency is significantly affected by the solidity of the
looks more similar than for lower spanwise positions (Fig. 19), machine.
except for a more pronounced disruption of the vortical structures (8) Theoretical derivations and numerical analyses have been
(Fig. 20) downstream the suction side. developed for a compressible flow in order to make the
Figure 21 reports the energy production for the different geo- methodology applicable also to other problems. For the
metries simulated, calculated as in Eq. (43). The available energy problems analyzed in this work, compressibility effects
calculated with the traditional and the indirect methods (EP and never accounted for more than 0.02% of the total entropy
EE) lead to very close predictions. The differences are larger when generation.
the available energy is evaluated using the direct method, i.e., by
integrating the production or destruction of turbulent kinetic
energy inside the domain: it is interesting to note how this last
estimation is always larger than the other two, except for the case Nomenclature
of the rotor with the highest solidity. It is reasonable to assume Dimensional Properties
larger numerical errors in the prediction of entropy generation for
the rotor with the largest solidity, where secondary flow structures c ¼ blade chord (m)
are larger and more complex due to the smaller flow passage and E ¼ energy (kg m2 s2)
higher pressure difference across the blade. The difference is in E_ x ¼ exergy per unit time (kg m2 s3)
any case well below the values reported in other similar studies f ¼ frequency (s1)
[36,39]. h ¼ enthalpy (kg m2 s2)

051205-12 / Vol. 143, MAY 2021 Transactions of the ASME


k¼ turbulent kinetic energy (m2 s2) CV ¼ control volume
s¼ specific entropy (m2 s2 K1) LE ¼ leading edge
S_ G ¼ entropy generation rate (kg m2 s3 K1) NS ¼ Navier–Stokes
p¼ pressure (kg m1 s2) OWC ¼ oscillating water column
Q¼ volumetric flow rate (m3 s1) PS ¼ pressure side
r¼ turbine radius (m) RANS ¼ Reynolds-averaged Navier–Stokes
t¼ time (s) REAL ¼ realizable
T¼ temperature (K) S–A ¼ Spalart–Allmaras
T ¼ torque (kg m2 s2) SS ¼ suction side
Tw ¼ wave cycle period (s) SST ¼ shear–stress transport
U¼ blade speed (m s1) STD ¼ standard
V¼ velocity (m s1) TE ¼ trailing edge

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


V¼ volume (m3)
W_ ¼ work (kg m2 s3) References
X¼ turbine tangential force (kg m s2) [1] Pelc, R., and Fujita, R., 2002, “Renewable Energy From the Ocean,” Mar. Pol-
Z¼ turbine axial force (kg m s2) icy, 26(6), pp. 471–479.
e¼ rate of dissipation of k (m2 s3) [2] Falc~ao, A., 2010, “Wave Energy Utilization: A Review of the Technologies,”
Renewable Sustainable Energy Rev., 14(3), pp. 899–918.
k¼ thermal conductivity (kg m s3 K1) [3] Wells, A., 1976, “Fluid Driven Rotary Transducer,” Patent No. 1595700A.
k¼ volume viscosity (kg m1 s1) [4] Curran, R., and Gato, L. M. C., 1997, “The Energy Conversion Performance of
l¼ dynamic viscosity (kg m1 s1) Several Types of Wells Turbine Designs,” Proc. Inst. Mech. Eng., Part A,
¼ kinematic viscosity (m2 s1) 211(2), pp. 133–145.
[5] Thakker, A., and Abdulhadi, R., 2008, “The Performance of Wells Turbine
P¼ stress tensor (kg m1 s2) Under bi-Directional Airflow,” Renewable Energy, 33(11), pp. 2467–2474.
q¼ air density (kg m3) [6] Paderi, M., and Puddu, P., 2013, “Experimental Investigation in a Wells Tur-
r¼ entropy generation rate per unit mass (m2 s3 K1) bine Under Bi-Directional Flow,” Renewable Energy, 57, pp. 570–576.
x¼ specific dissipation rate (s1) [7] Dhanasekaran, T., and Govardhan, M., 2005, “Computational Analysis of Per-
formance and Flow Investigation on Wells Turbine for Wave Energy Con-
X¼ angular rotational frequency (s1) version,” Renewable Energy, 30(14), pp. 2129–2147.
rS ¼ sum of gradient and gradient transposed (m1) [8] Setoguchi, T., Kinoue, Y., Kim, T., Kaneko, K., and Inoue, M., 2003,
“Hysteretic Characteristics of Wells Turbine for Wave Power Conversion,”
Non-dimensional Properties Renewable Energy, 28(13), pp. 2113–2127.
[9] Torresi, M., Camporeale, S., and Pascazio, G., 2009, “Detailed CFD Analysis
fv1 ¼ viscous damping function of the Steady Flow in a Wells Turbine Under Incipient and Deep Stall Con-
H ¼ efficiency ditions,” ASME J. Fluids Eng., 131(7), p. 071103.
I ¼ identity tensor [10] Shehata, A., Saqr, K., Xiao, Q., Shehadeh, M., and Day, A., 2016,
“Performance Analysis of Wells Turbine Blades Using the Entropy Generation
KS_ ¼ entropy generation/dissipation rate Minimization Method,” Renewable Energy, 86, pp. 1123–1133.
Re ¼ Reynolds number [11] Shehata, A., Xiao, Q., El-Shaib, M., Sharara, A., and Alexander, D.,
p ¼ static pressure drop coefficient 2017, “Comparative Analysis of Different Wave Turbine Designs Based on
pt ¼ total pressure drop coefficient Conditions Relevant to Northern Coast of Egypt,” Energy, 120, pp.
T  ¼ torque coefficient 450–467.
[12] Soltanmohamadi, R., and Lakzian, E., 2016, “Improved Design of Wells Tur-
z ¼ number of blades bine for Wave Energy Conversion Using Entropy Generation,” Meccanica,
/ ¼ flow coefficient 51(8), pp. 1713–1722.
[13] Shaaban, S., 2012, “Insight Analysis of Biplane Wells Turbine Performance,”
Energy Convers. Manage., 59, pp. 50–57.
Subscripts and Superscripts [14] Gratton, T., Ghisu, T., Parks, G., Cambuli, F., and Puddu, P., 2018,
“Optimization of Blade Profiles for the Wells Turbine,” Ocean Eng., 169, pp.
a ¼ axial 202–214.
ad ¼ aerodynamic [15] Kim, T., Setoguchi, T., Takao, M., Kaneko, K., and Santhakumar, S., 2002,
d ¼ dissipation “Study of Turbine With Self-Pitch-Controlled Blades for Wave Energy Con-
g ¼ production version,” Int. J. Therm. Sci., 41(1), pp. 101–107.
[16] Starzmann, R., and Carolus, T., 2014, “Effect of Blade Skew Strategies on the
hub ¼ turbine hub Operating Range and Aeroacoustic Performance of the Wells Turbine,” ASME
II ¼ second-law approach J. Turbomach., 136(1), p. 011003.
in ¼ input [17] Licheri, F., Climan, A., Puddu, P., Cambuli, F., and Ghisu, T., 2018,
ind ¼ indirect “Numerical Study of a Wells Turbine With Variable Pitch Rotor Blades,”
Energy Procedia, 148, pp. 511–518.
is ¼ isentropic [18] Mohamed, M. H., Janiga, G., Pap, E., and Thevenin, D., 2011, “Multi-
lost ¼ lost Objective Optimization of the Airfoil Shape of Wells Turbine Used for Wave
m ¼ mid radius Energy Conversion,” Energy, 36(1), pp. 438–446.
mf ¼ mean flow [19] Mohamed, M., and Shaaban, S., 2013, “Optimization of Blade Pitch Angle of
an Axial Turbine Used for Wave Energy Conversion,” Energy, 56, pp.
out ¼ output 229–239.
pol ¼ polytropic [20] Halder, P., Rhee, S. H., and Samad, A., 2017, “Numerical Optimization of
R ¼ Reynolds’ Wells Turbine for Wave Energy Extraction,” Int. J. Nav. Arch. Ocean Eng.,
ref ¼ reference 9(1), pp. 11–24.
[21] Nazeryan, M., and Lakzian, E., 2018, “Detailed Entropy Generation Analysis
t ¼ total quantity of a Wells Turbine Using the Variation of the Blade Thickness,” Energy, 143,
T ¼ turbulent pp. 385–405.
tip ¼ blade tip [22] Shehata, A., Xiao, Q., Selim, M., Elbatran, A., and Alexander, D., 2017,
ts ¼ total-to-static “Enhancement of Performance of Wave Turbine During Stall Using Passive
Flow Control: First and Second Law Analysis,” Renewable Energy, 113, pp.
tt ¼ total-to-total 369–392.
V ¼ due to fluid flow [23] Suzuki, M., Arakawa, C., and Tagori, T., 1984, “Fundamental Studies on Wells
0
¼ fluctuating component Turbine for Wave Power Generator (1st Report, the Effect of Solidity, and Self-
ð Þ ¼ mean value Starting),” Bull. JSME, 27(231), pp. 1925–1931.
[24] Gato, L., and de O. Falc~ao, A., 1988, “Aerodynamics of the Wells Turbine,”
Acronyms Int. J. Mech. Sci., 30(6), pp. 383–395.
[25] Folley, M., Curran, R., and Whittaker, T., 2006, “Comparison of Limpet
CFD ¼ computational fluid dynamics Contra-Rotating Wells Turbine With Theoretical and Model Test Predictions,”
CFL ¼ Courant–Friedrichs–Lewy number Ocean Eng., 33(8–9), pp. 1056–1069.

Journal of Fluids Engineering MAY 2021, Vol. 143 / 051205-13


[26] Raghunathan, S., Tan, C. P., and Ombaka, O. O., 1985, “Performance of the [54] Ghisu, T., Puddu, P., Cambuli, F., and Virdis, I., 2017, “On the Hysteretic
Wells Self-Rectifying Air Turbine,” Aeronaut. J., 89, pp. 369–379. Behaviour of Wells Turbines,” Energy Procedia, 126, pp. 706–713.
[27] Dixon, S. L., and Hall, C. A., 2010, Fluid Mechanics and Thermodynamics of [55] Ghisu, T., Cambuli, F., Puddu, P., Virdis, I., Carta, M., and Licheri, F., 2020,
Turbomachinery, 6th ed., Elsevier, Burlington, MA. “A Critical Examination of the Hysteresis in Wells Turbines Using Computa-
[28] Raghunathan, S., Tan, C., and Wells, N., 1982, “Theory and Performance of a tional Fluid Dynamics and Lumped Parameter Models,” ASME J. Offshore
Wells Turbine,” J. Energy, 6(2), pp. 157–160. Mech. Arct. Eng., 142(5), p. 052001.
[29] Gato, L. M. C., and Curran, R., 1996, “Performance of the Biplane Wells [56] Roache, P. J., Ghia, K. N., and White, F. M., 1986, “Editorial Policy Statement
Turbine,” ASME J. Offshore Mech. Arct. Eng., 118(3), pp. 210–215. on the Control of Numerical Accuracy,” ASME J. Fluids Eng., 108(1), p. 2.
[30] Gato, L. M. C., and Falc~ao, A. F. D. O., 1984, “On the Theory of the Wells [57] Stern, F., Wilson, R. V., Coleman, H. W., and Paterson, E. G., 2001,
Turbine,” ASME J. Eng. Gas Turbines Power, 106(3), pp. 628–633. “Comprehensive Approach to Verification and Validation of CFD
[31] Bejan, A., 2006, Advanced Engineering Thermodynamics, 3rd ed., Wiley, Simulations—Part 1: Methodology and Procedures,” ASME J. Fluids Eng.,
Hoboken, NJ. 123(4), pp. 793–802.
[32] Ghisu, T., Cambuli, F., Puddu, P., Mandas, N., Seshadri, P., and Parks, G. T., [58] Celik, I. B., Ghia, U., Roache, P. J., Freitas, C. J., Coleman, H., and Raad, P. E.,
2018, “Numerical Evaluation of Entropy Generation in Isolated Airfoils and 2008, “Procedure for Estimation and Reporting of Uncertainty Due to Discreti-
Wells Turbines,” Meccanica, 53(14), pp. 3437–3456. zation in CFD Applications,” ASME J. Fluids Eng., 130(7), p. 078001.

Downloaded from http://asmedigitalcollection.asme.org/fluidsengineering/article-pdf/143/5/051205/6632792/fe_143_05_051205.pdf by Universita Di Cagliari user on 10 February 2021


[33] Ferziger, J. H., and Peric, M., 2002, Computational Methods for Fluid Dynam- [59] Ghisu, T., Cambuli, F., Puddu, P., Virdis, I., and Carta, M., 2019, “Discussion
ics, Springer-Verlag, Berlin. on “Unsteady RANS Simulations of Wells Turbine Under Transient Flow Con-
[34] Versteeg, H. K., and Malalasekera, W., 2007, An Introduction to Computational ditions” by Hu and Li,” ASME J. Offshore Mech. Arct. Eng., 141(4),
Fluid Dynamics: The Finite Volume Method, Pearson Education Limited, Har- p. 045501.
low, UK. [60] Brito-Melo, A., Neuman, F., and Sarmento, A. J. N. A., 2008, “Full-Scale Data
[35] Anderson, D. A., Tannehill, J. C., and Pletcher, R. H., 1984, Computational Assessment in OWC Pico Plant,” Proceedings of the 17th International Journal
Fluid Mechanics and Heat Transfer, Taylor & Francis, Boca Raton, FL. of Offshore and Polar Engineering, Lisbon, Portugal, July, Vol. 18, pp. 27–34.
[36] Herwig, H., and Kock, F., 2006, “Direct and Indirect Methods of Calculating [61] Garrido, A. J., Otaola, E., Garrido, I., Lekube, J., Maseda, F. J., Liria, P., and
Entropy Generation Rates in Turbulent Convective Heat Transfer Problems,” Mad Er, J., 2015, “Mathematical Modeling of Oscillating Water Columns
Heat Mass Transfer, 43(3), pp. 207–215. Wave-Structure Interaction in Ocean Energy Plants,” Math. Probl. Eng., 2015,
[37] Moore, J., and Moore, J. G., 1983, “Entropy Production Rates From Viscous pp. 1–11.
Flow Calculations—Part I: A Turbulent Boundary Layer Flow,” ASME Paper [62] Ghisu, T., Puddu, P., Cambuli, F., Mandas, N., Seshadri, P., and Parks, G.,
No. 83-GT-70. 2018, “Discussion on “Performance Analysis of Wells Turbine Blades Using
[38] Iandoli, C. L., Sciubba, E., and Zeoli, N., 2008, “The Computation of the the Entropy Generation Minimization Method” by Shehata, A. S., Saqr, K. M.,
Entropy Generation Rate for Turbomachinery Design Applications: Some The- Xiao, Q., Shahadeh, M. F. and Day, A,” Renewable Energy, 118, pp. 386–392.
oretical Remarks and Practical Examples,” Int. J. Energy Technol. Policy, 6(1/ [63] Carr, L. W., McAlister, K. W., and McCroskey, W. J., 1977, “Analysis of the
2), pp. 64–95. Development of Dynamic Stall Based on Oscillating Airfoil Experiment,”
[39] Jin, Y., and Herwig, H., 2015, “Turbulent Flow in Rough Wall Channels: Vali- NASA AMES Research Center, Washington, DC, Report No. D-8382.
dation of Rans Models,” Comput. Fluids, 122, pp. 34–46. [64] Ericsson, L. E., and Reding, J. P., 1988, “Fluid Mechanics of Dynamic Stall—
[40] Asinari, P., Fasano, M., and Chiavazzo, E., 2016, “A Kinetic Perspective on Part 1: Unsteady Flow Concepts,” J. Fluids Struct., 2(1), pp. 1–33.
ke Turbulence Model and Corresponding Entropy Production,” Entropy, [65] Barakos, G. N., and Drikakis, D., 2003, “Computational Study of Unsteady Tur-
18(4), p. 121. bulent Flows Around Oscillating and Ramping Aerofoils,” Int. J. Numer. Meth-
[41] Shih, T.-H., Liou, W. W., Shabbir, A., Yang, Z., and Zhu, J., 1995, “A New k-e ods Fluids, 42(2), pp. 163–186.
Eddy Viscosity Model for High Reynolds Number Turbulent Flows,” Comput. [66] Cambuli, F., Ghisu, T., Virdis, I., and Puddu, P., 2019, “Dynamic Interaction
Fluids, 24(3), pp. 227–238. Between OWC System and Wells Turbine: A Comparison Between CFD and
[42] ANSYS, Inc., 2013, “ANSYS Fluent User’s Guide,” 15.0 ed., Southpointe/ Lumped Parameter Model Approaches,” Ocean Eng., 191, p. 106459.
ANSYS, Canonsburg, PA. [67] Ghisu, T., Cambuli, F., Puddu, P., Virdis, I., Carta, M., and Licheri, F., 2020,
[43] Setoguchi, T., Takao, M., and Kaneko, K., 1998, “Hysteresis on Wells Turbine “A Lumped Parameter Model to Explain the Cause of the Hysteresis in OWC-
Characteristics in Reciprocating Flow,” Int. J. Rotating Mach., 4(1), pp. 17–24. Wells Turbine Systems for Wave Energy Conversion,” Appl. Ocean Res., 94,
[44] Wilkinson, M. B., van der Spuy, J., and von Backstr€ om, T. W., 2018, p. 101994.
“Performance Testing of an Axial Flow Fan Designed for Air-Cooled Heat [68] Mittal, S., and Saxena, P., 2000, “Prediction of Hysteresis Associated With the
Exchanger Applications,” ASME Paper No. GT2018-75964. Static Stall of an Airfoil,” AIAA J., 38(5), pp. 933–935.
[45] Sun, T., Petrie-Repar, P., Vogt, D. M., and Hou, A., 2019, “Detached-Eddy [69] Sarlak, H., Fr Re, A., Mikkelsen, R., and Sørensen, J., 2018, “Experimental
Simulation Applied to Aeroelastic Stability Analysis in a Last-Stage Steam Tur- Investigation of Static Stall Hysteresis and 3-Dimensional Flow Structures for
bine Blade,” ASME J. Turbomach., 141(9), p. 091002. an NREL S826 Wing Section of Finite Span,” Energies, 11(6), p. 1418.
[46] Hill, D. J., and Defoe, J. J., 2020, “Scaling of Incidence Variations With Inlet [70] Menter, F., 1992, “Performance of Popular Turbulence Models for Attached
Distortion for a Transonic Axial Compressor,” ASME J. Turbomach., 142(2), and Separated Adverse Pressure Gradient Flows,” AIAA J., 30(8), pp.
p. 021003. 2066–2072.
[47] Watterson, J., and Raghunathan, S., 1998, “Computed Effects of Solidity on [71] Menter, F., 1994, “Two-Equation Eddy-Viscosity Turbulence Models for Engi-
Wells Turbine Performance,” JSME Int. J., Ser. B, 41(1), pp. 177–183. neering Applications,” AIAA J., 32(8), pp. 1598–1605.
[48] Hashem, I., Abdel Hameed, H., and Mohamed, M., 2018, “An Axial Turbine in [72] Li, D., Wang, H., Qin, Y., Han, L., Wei, X., and Qin, D., 2017, “Entropy Pro-
an Innovative Oscillating Water Column (OWC) Device for Sea-Wave Energy duction Analysis of Hysteresis Characteristic of a Pump-Turbine Model,”
Conversion,” Ocean Eng., 164, pp. 536–562. Energy Convers. Manage., 149, pp. 175–191.
[49] Das, T. K., and Samad, A., 2019, “The Effect of Midplane Guide Vanes in a [73] Fidkowski, K. J., Ceze, M. A., and Roe, P. L., 2012, “Entropy-Based Drag-
Biplane Wells Turbine,” ASME J. Fluids Eng., 141(5), p. 051107. Error Estimation and Mesh Adaptation in Two Dimensions,” J. Aircr., 49(5),
[50] Kumar, P. M., and Samad, A., 2019, “Introducing Gurney Flap to Wells Tur- pp. 1485–1496.
bine Blade and Performance Analysis With Openfoam,” Ocean Eng., 187, [74] Mizoguchi, M., Kajikawa, Y., and Itoh, H., 2014, “Static Stall Hysteresis of
p. 106212. Low-Aspect-Ratio Wings,” AIAA Paper No. 2014–2014.
[51] Valizadeh, R., Abbaspour, M., and Rahni, M. T., 2020, “A Low Cost Hydroki- [75] Modarres, R., Peters, D., and Gaskill, J., 2016, “Dynamic Stall Model With Cir-
netic Wells Turbine System for Oceanic Surface Waves Energy Harvesting,” culation Pulse and Static Hysteresis for NACA 0012 and VR-12 Airfoils,” J.
Renewable Energy, 156, pp. 610–623. Am. Helicopter Soc., 61(3), pp. 1–10.
[52] Ghisu, T., Puddu, P., and Cambuli, F., 2016, “Physical Explanation of the Hys- [76] Thakker, A., and Abdulhadi, R., 2007, “Effect of Blade Profile on the Perform-
teresis in Wells Turbines: A Critical Reconsideration,” ASME J. Fluids Eng., ance of Wells Turbine Under Unidirectional Sinusoidal and Real Sea Flow
138(11), p. 111105. Conditions,” Int. J. Rotating Mach., 2007, pp. 1–9.
[53] Ghisu, T., Puddu, P., and Cambuli, F., 2017, “A Detailed Analysis of the [77] Torresi, M., Camporeale, S., Strippoli, P., and Pascazio, G., 2008, “Accurate
Unsteady Flow Within a Wells Turbine,” Proc. Inst. Mech. Eng., Part A, Numerical Simulation of a High Solidity Wells Turbine,” Renewable Energy,
231(3), pp. 197–214. 33(4), pp. 735–747.

051205-14 / Vol. 143, MAY 2021 Transactions of the ASME

View publication stats

You might also like