Accepted Manuscript Not Copyedited: The Effect of Mid-Plane Guide Vanes in A Biplane Wells Turbine

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

Journal of Fluids Engineering.

Received May 08, 2018;


Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

The Effect of Mid-Plane Guide Vanes in a


Biplane Wells Turbine
Tapas K. Das, first author
Wave Energy and Fluids Engineering Lab (WEFEL), IIT Madras

d
Ocean Engineering Dept., IIT Madras, Chennai-600036, India

ite
mech.tapas@gmail.com

ed
Abdus Samad, second author1
Wave Energy and Fluids Engineering Lab (WEFEL), IIT Madras

py
Ocean Engineering Dept., IIT Madras, Chennai-600036, India
samad@iitm.ac.in

Co
ABSTRACT

ot
tN
Guide vanes improve the performance of a turbine in terms of efficiency, torque or operating range. In this
rip

work, a concept of different orientations of guide vanes in between a two-row biplane Wells turbine was

introduced and analyzed for the performance improvement. The fluid flow was simulated numerically with
sc

a commercial software Ansys CFX 16.1. The Reynolds-averaged Navier-Stokes equations with the k-ω
nu

turbulence closure model were solved for different designs and flow conditions. For the base model, the
Ma

results from simulation and experiments are in close agreement. Among the designs considered, the

configuration, where, the blades are in one line (zero circumferential angle between blades of two plane)
ed

and the mid-plane guide vane has concave side to the leading edge of the blade, performed relatively better.

However, the performance was still less compared to the base model. The reason behind the reduction in
pt

performance from the base model is attributed to the blockage of flow and the change of flow path occurring
ce

due to the presence of the mid-plane guide vanes. The flow analysis of different cases and comparison with
Ac

the base model is presented in the current study.

Keywords: Well turbine, biplane, guide vane, OWC, CFD

1
Associate Professor, Ocean Engineering Dept., IIT Madras

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

INTRODUCTION

Oscillating Water Column (OWC) has been a popular and extensively studied

device for wave energy extraction resulting in large numbers of the prototype developed

d
to date. [1]. The device works with the help of pressure difference created inside the

ite
chamber by the ocean waves. A turbine is used to convert the pressure energy into kinetic

ed
energy, which is then used by a generator to produce electricity. Both impulse and

py
reaction turbines have been used in OWC. The Wells turbine is one type of reaction

Co
turbine that has been used widely in OWC. For the same geometric parameters and

similar operating conditions, the Wells turbine has a higher efficiency[2] and requires less

ot
tN
maintenance to operate[3] compared to impulse turbines in both onshore or offshore

conditions.
rip

Raghunathan [4] analytically studied the different parameters of Wells turbine


sc

and came up with limiting values of some important parameters such as solidity, hub to
nu

tip ratio, aspect ratio, and tip clearance. Several different blade geometries of the Wells
Ma

turbine were tested both experimentally and numerically to improve the performance of
ed

the turbine. Blade sweep [5,6], blade skew [7] have shown marginal improvement in

performance from the baseline design. Variation in thickness from hub to tip increases
pt
ce

the turbine performance significantly [8,9]. All such studies on blade geometry were
Ac

carried out on monoplane turbine, and it was reported that a large quantity of kinetic

energy is lost at the downstream of the rotor blade.

To recover such kinetic energy, two methods were suggested by Raghunathan [4]:

use of guide vanes at the exit of the rotor blade or use of another set of rotor blade in a

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

plane downstream of the main rotor blades. It has been proved both experimentally and

numerically that use of double row of GVs results in considerable increase in efficiency

[10]. Takao et al. [11] experimentally investigated the effect of the gap between the rotor

blade and the GVs. The starting characteristics, as well as running characteristics, are

d
ite
significantly enhanced by the use of GVs. The best efficiency occurs for an optimum gap

ed
between the rotor blade and GVs. Halder and Samad [12] numerically studied different

GV angles and concludes that the GV angle depends on the solidity of the turbine. All the

py
studies on GV mentioned above were carried out using 2D GV for which the stagger angle

Co
of GV remains constant throughout the length of the blade. Takao et al. [13] and

ot
Setoguchi et al.[14] studied the effect of 3D GV for further performance enhancement of
tN
Wells turbine. The experimental studies show that the 3D GV has better performance
rip

characteristics compared to 2D GVs.


sc

While use of GV is a way to recover the exit kinetic energy, adding one more rotor
nu

blade downstream of the main blade also helps in recovering exit kinetic energy and
Ma

improvement of performance. A comparison of different types of Wells turbine [15]

shows that the biplane Wells turbine has better performance characteristics compared to
ed

other types of Wells turbines. The aerodynamics of Biplane Wells turbine has been
pt

studied extensively by Raghunathan [16] and the performance is predicted using inlet and
ce

outlet GV. The hysteresis characteristic of the monoplane Wells turbine is studied
Ac

numerically in [17]. The origin of the hysteresis is attributed to the compressibility effect

of the OWC system. Mamun et al.[18] and Kinoue et al.[19] investigated the hysteresis of

biplane Wells turbine and compared it with monoplane Wells turbine. For a lower angle

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

of attack, both monoplane and biplane turbine show similar hysteresis loops. As the angle

of attack increases, the dynamic stall phenomenon is noticed and the torque as well

efficiency drops drastically. This is attributed to the flow separation near the hub and the

suction surface of the upstream blade.

d
ite
The main parameters affecting the performance of a Biplane Wells Turbine (BWT)

ed
are the gap-to-chord ratio, blade setting angle, solidity and the circumferential angle

between the rotor blades of two plane. Kaneko et al.[20] experimentally investigated the

py
effects of these parameters on BWT and recommends the blade setting angle between 2-

Co
4o and gap to chord ratio 1.5. The results also show that the hysteresis characteristic is

ot
more sensitive to blade solidity compared to the blade setting angle and gap-to-chord
tN
ratio. Shaaban [21] carried out a numerical analysis of BWT, which shows that the
rip

downstream rotor contributes to only 10-30% of the total torque. The numerical results
sc

suggested improving the downstream rotor performance in future works. Arlitt et al. [22]
nu

introduced a concept of mid-plane guide vane for BWT to reduce the exit kinetic energy
Ma

loss from the downstream blade. A numerical study of BWT with a generator in between

rotating planes [23] shows the aerodynamic behavior and the flow interaction between
ed

rotating and stationary stages. Alves [24] carried out an experimental and numerical
pt

analysis of BWT with GVs in the mid-plane. The performance characteristics were
ce

compared with the numerical and experimental result. However, no detail flow analysis
Ac

in the mid- plane of the turbine is provided in this work.

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

The purpose of the present work is to study the flow phenomenon when GVs are

placed in the mid-plane of a BWT. The interaction of the flow with the GVs is studied

numerically and the performance characteristics are compared with turbine without GV.

d
ite
GEOMETRIC CONFIGURATIONS

ed
Four different geometries of BWT with mid-plane GV are considered here for

analysis and compared with two reference geometries without GV. The reference

py
geometries, Ref-1 and Ref-2, are shown in Fig 1. Both the turbine has four blades in each

Co
plane having an overall solidity of 0.64 (0.32 per plane) and has symmetric NACA0015

ot
blade profile. The Ref-1 has blades in a staggered way; the circumferential angle between
tN
the blades is 45o. The Ref-2 has blades in a straight line with no circumferential angle
rip

between them. The dimensions of various parts are presented in Table 1. The geometries
sc

are varied based on the position of GVs. Figure 2 shows the four different cases. For Case
nu

A and B, the rotor blades are in one line in two planes without any circumferential angle
Ma

between them. For case C and D, the rotor blades have a circumferential angle of 45 o in

each plane. The gap between each plane is fixed at 1.5c for all the cases.
ed

The spacing between GVs is an important factor to consider for selecting the
pt

number of GVs. The spacing between the GVs defines the load acting on them and the
ce

frictional loss due to surface area. If there are large spacing, the load on the blade will be
Ac

too high. For smaller spacing, the frictional loss is very high due to an increased surface

area. Hence there exists an optimum value of blade spacing which provides reasonably

good blade loading and optimum frictional loss as well. Zweifel’s blade loading criteria

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

(Zweifel coefficient) is based on non-dimensional force on the cascade. The Zweifel

coefficient is given by Eq. (1)

2F
z (1)
V22c

d
This in terms of blade inlet and exit angles, Eq. (2)

ite
s
z  2 cos2   tan   tan   (2)

ed
c

py
Zweifel [25] suggested a value of 0.8 for coefficient z for optimal blade loading and

Co
frictional loss. Thus Zweifel criteria can be used to determine the number of stator blades

provided other values are known or assumed properly.

ot
The lift interference factor, defined as the ratio of lift coefficient of a blade in
tN
cascade to that of an isolated blade, is very important while choosing the gap between
rip

blades. Weinig [26] used potential flow theory to analytically determine the lift
sc

interference factor for a linear cascade of infinitesimally flat blades. The interference
nu

factor based on Weinig’s theory for isolated aerofoil is given by Eq. (3)
Ma

CL
K0  (3)
2 sin  
ed

For axial turbine with 90o stagger angle, like in the case of Wells turbine, the
pt

interference factor is given by Eq. (4)


ce

2 s  c
K0  tan   (4)
Ac

c  2 s

This shows that the interference factor does not depend on the blade angle and

only governs by the ratio of blade pitch to blade chord. For this study, the number and

profile of the GVs are chosen based on Zweifel criterion and Weinig theory. The GVs are

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

placed in the mid-plane with equal distance from each blade. The GVs have a circular arc

profile with a chord length of 90 mm and camber angle 60o. The stagger angle of the GV

is kept at 0o to maintain the symmetry of the bidirectional turbine. The gap between the

blade and the GV is 48 mm.

d
ite
ed
NUMERICAL SOLUTION APPROACH

py
The main steps of any CFD analysis involve the creation of the computational

Co
domain, meshing the domain and finally using a turbulence model to solve the governing

ot
equations associated with the problem. The main challenge for creating the
tN
computational domain is to find the optimum size of the domain so that the
rip

computational cost is minimized and at the same time, the accuracy of the numerical
sc

simulation is not compromised. A detailed CFD analysis of monoplane Wells turbine with
nu

steady flow is given in [27]. The boundary conditions and use of different turbulence
Ma

models for monoplane Wells turbine to capture incipient and deep stall condition are

discussed in detail by the authors. A similar numerical approach is used in the present
ed

study with biplane Wells turbine having mid-plane guide vanes. The computational
pt

domain and mesh of the reference blade for present study are shown in Fig. 3(a). As the
ce

turbine is symmetric about its rotational axis, only one-fourth of the whole turbine is
Ac

taken as computational domain. The Ref-1 geometry is used for the validation with

available experimental results [28]. The domain consists of one blade from each plane

and is five chord length long in the upstream side and ten chord length long in the

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

downstream side. Figure 3(b) shows the computational domain with GVs in mid-plane.

While the rotor blades have a rotational speed, the GVs are fixed. The domain interface

between rotor blade and GV is modeled using frozen rotor approach. In frozen rotor

approach, the relative position of the components across the domain interface remains

d
ite
fixed. The flow variables are derived near the boundary of rotating domain and it is

ed
transferred directly to the neighboring fixed domain.

Mesh generation is a key component in the simulation environment. One of the

py
main aspects of meshing is to reduce the discretization error that depends on the quality

Co
of the mesh generated. The computational domain in the present study was discretized

ot
using unstructured mesh (Fig. 3c & 3d). A fine prism layer was created around the blade
tN
to capture the boundary layer effect on flow. The height of the first layer from the wall of
rip

the rotor blade was 1.2x10-5 m. This height was chosen based on y+ <1. Total 20 layers
sc

were used with an exponential growth ratio of 1.2. The total height of the prism layer
nu

around the blade was 2.24 mm. The Reynolds number was approximately 6.5x105 and
Ma

6.6x105 for flow coefficient 0.225 and 0.250 respectively. The important mesh parameters

calculated from CFD post (running in double precision): maximum edge length ratio 136.1,
ed

minimum face angle 10.72o, maximum face angle 137.57o and maximum element volume
pt

ratio 17.8.
ce

The different cases of BWT with GVs were numerically studied by solving RANS
Ac

equations for steady incompressible flow in a coupled, finite volume based solver ANSYS

CFX v16.1, which solves the hydrodynamic equations (for u, v, w, p) as a single system[29].

High-resolution advection scheme of second-order accuracy is implemented to calculate

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

the advection terms in discrete finite-volume equations. The turbulence is modeled using

SST k-ω model, which is a low Reynolds number model. Here, SST equation combines the

k-ω model for near wall flow and the k-ε model for free stream flow. The advantage of

SST k- ω model over other models is its ability to capture the near wall flow as well as the

d
ite
free stream turbulence. A study [30] of different turbulence models for flow over

ed
symmetrical NACA airfoil shows that the SST k-ω model is the most suitable one. As the

Wells turbine rotor also consists of symmetrical NACA profiles, the SST k- ω model was

py
chosen for further numerical analysis of the BWT.

Co
The fluid (air) in the computational domain was incompressible, and a steady flow

ot
was considered for analysis. The inlet had a steady velocity profile and ambient pressure
tN
at the outlet. The periodic boundary condition was imposed on the circumferential sides,
rip

and the surface of rotor blade had no-slip boundary condition. Both the rotors rotated
sc

with the same rotational speed of 2500 rpm, and this was fixed throughout the analysis.
nu

The boundary conditions applied to the computational domain are detailed in Table 2.
Ma

The Grid Convergence Index (GCI) is calculated based on Richardson extrapolation

to calculate the discretization error due to computational grids. The detailed procedure
ed

of GCI calculation is provided by Roeche [31] and Celik and Karatekin [32]. The same has
pt

been followed here and the values are tabulated in Table 3. Three different grid systems
ce

(N1=10.495x106 cells, N2=4.406x106 cells, N3=2.413x106 cells) were taken for calculation
Ac

of the torque coefficient of the reference geometry. The grid refinement factors are 1.33

and 1.22 as mentioned in Table 3. The calculated values of torque coefficient have an

21 32
oscillatory convergence. The extrapolated relative errors, 𝑒𝑒𝑥𝑡 and 𝑒𝑒𝑥𝑡 , are 0.064% and

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

21
0.260% respectively. The GCI obtained with the finest grid, 𝐺𝐶𝐼𝑓𝑖𝑛𝑒 , is 0.076%. Whereas

21
GCI obtained with the coarsest grid, 𝐺𝐶𝐼𝑐𝑜𝑎𝑟𝑠𝑒 , is 0.322%. As both the GCI is quite small,

the further calculations are carried out with the coarsest grid (N3=2.413x106 cells) to save

computational cost.

d
ite
Due to the computational requirement, the numerical simulation was done using

ed
Virgo super cluster having a total computational power of 97 TFlops. The Virgo

supercluster consists of IBM System x iDataPlex dx360 M4 highly optimized servers for

py
high-performance computing. Using this high-performance computing system, the

Co
average time for each simulation were 10 ~ 12 hours.

ot
VELOCITY DIAGRAM:
tN
Most of the works reported on Wells turbine with GVs are for monoplane Wells
rip

turbine having GVs at the inlet and the outlet of the rotor. In case of Wells turbine with
sc

an inlet GV, the angle of attack near hub is greater than that of the case without GVs [13].
nu

The velocity diagram of a BWT with and without mid-plane GV is shown in figure 4.
Ma

Positions before and after the rotor blade are mentioned as 0,1,2,3. The relative velocity

(W1) and incidence angle (α0) at the inlet of the upstream blade is same for both the
ed

turbines. Airflow leaves the upstream blade with a velocity V 1 and angle β1. If there is no
pt

GV in the mid-plane, the airflow hits the downstream blade with relative velocity (W1) at
ce

an incidence angle (α1). However, if GVs are introduced in the mid-plane, the direction
Ac

(β2) of absolute velocity (V2) changes. As a result, the incidence angle (α2) for the relative

velocity (W2) for the downstream blade is higher than it is without GV (αw/o). So, the use

10

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

of GV in the mid-plane makes the angle of attack higher for the downstream blade

compared to the angle of attack without the GV.

RESULTS AND DISCUSSION

Based on the previous literature, the Wells turbine performance is expressed

d
ite
using four non-dimensional parameters:

ed
The torque coefficient (Eq. 5):

py
T
CT  (5)
 2 rt5

Co
The pressure drop coefficient (Eq. 6):

ot
P0
P0*  (6)
 2 rt 2
tN
The efficiency (Eq. 7):
rip

T
 (7)
sc

P0Q
nu

Flow coefficient (Eq. 8):


Ma

va
 (8)
ut
ed

The torque coefficient, pressure coefficient, and efficiency are compared with
pt

respect to the flow coefficient. As the turbine rotational speed is fixed, the rotor tip speed
ce

is also fixed. The flow coefficient was varied by varying the inlet flow velocity. For all
Ac

different cases, numerical analysis was carried out at different points starting from a low

flow coefficient (  =0.075) to a high flow coefficient (  =0.30).

11

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

The computational result of Ref-1 is validated with the available experimental

results [28] (Figure 5). The numerical results are in close agreement with the experimental

results for  < 0.225. An average deviation of ~5% from the experimental result is

observed for pressure coefficient, torque coefficient, and efficiency for  < 0.225.

d
ite
However, at a high flow coefficient, the numerical result has a higher deviation from the

ed
available experimental results. This is the region of dynamic stall, where the efficiency

py
falls drastically mainly because of the flow separation near the trailing edge of the blade.

Co
One of the reasons for this mismatch at high flow coefficient is the inability of RANS model

to solve beyond the stall region [33]. The numerical result of the present work is

substantiated by the numerical result of Shaaban [21].


ot
tN
Figure 6 shows the comparison of performance for four different cases with the
rip

two reference geometries. The torque coefficient for the reference geometries is much
sc

higher compared to the geometries including GVs. However, the trend of the torque
nu

coefficient is same for all the geometries. For both the reference geometries and Case A
Ma

and C, the maximum torque coefficient is observed at  = 0.25. For  = 0.225, the angle

of attack at the tip is 12.68o and for  = 0.250, it is 14.03o. Till  = 0.225, flow remains
ed

attached to the blade, no phenomenon of flow separation is observed. The angle of attack
pt

is also a function of the span of the blade, and it increases from blade tip to the hub. In
ce

the region of incipient stall [27], 0.225<  <0.250, the flow starts to separate from hub
Ac

due to high angle of attack, but still remains attached near the tip. At  = 0.250, the flow

is completely detached from the blade throughout the span, and the fluid elements near

the blade can no longer proceed in forward direction. This leads to the phenomenon of

12

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

dynamic stall at this flow coefficient, and the torque coefficient decreases drastically. For

Case B and D, the stall occurs near  = 0.225.

The pressure difference between inlet and outlet of the rotor is defined as the

pressure coefficient. The pressure coefficient varies linearly with the flow coefficient till

d
ite
the point of stall. After stall, pressure coefficient drops significantly for all the cases

ed
including the reference geometries. Similar to the torque coefficient, the pressure

py
coefficient values are also very high for reference geometries compared to the four cases

with GVs.

Co
Efficiency of the rotor is a function of both torque and pressure drop across the

ot
rotor. As a result, although the torque and pressure drop values are less for the cases with
tN
GVs, the efficiency of the reference geometries and the modified geometries (including
rip

GVs) are in good agreement. It can be seen that for all the turbine geometries, with and
sc

without GV (except Case C), the efficiency is almost the same for a range of flow
nu

coefficients. This range of flow coefficient (0.125 to 0.250) can be defined as the optimum
Ma

operating range for all the geometries considered here. Within this range, the rotor

efficiency does not vary much with the change of the inlet velocity. It can be inferred from
ed

the comparisons that the inclusion of mid-plane GVs does not improve the torque or
pt

pressure drop across the rotor. The reason for this can be found out by further looking at
ce

the flow characteristics on the rotor blades.


Ac

Figure 7 shows the streamlines on the suction surface of the blades at flow

coefficient 0.225. It can be seen that for all four cases (Case A-D) the suction surface of

the upstream blade has similar velocity streamlines. For both the reference turbines, the

13

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

flow on the suction surface on upstream blade follows the blade till mid-chord length and

then a reverse flow occurs resulting in flow separation near the trailing edge. For four

cases with GVs, the flow gets separated very near to the leading edge of the upstream

blade. The flow separation starts at hub and it progresses to tip. However, for the

d
ite
downstream blades, the streamlines show a different pattern. For the reference blades,

ed
the flow is attached with the blade suction surface up to the trailing edge. Only small

vortices can be noticed near the hub at the suction surface (Ref-2). For Case A and D, flow

py
separation occurs at the downstream blade suction surface almost at the mid-chord of

Co
the blade. In Case B, vortices are generated near the hub of the rotor blade and it

ot
propagates towards the blade tip. For Case C, flow circulation occurs near the leading
tN
edge at blade tip region. This leads to the drop in torque and efficiency for this case (refer
rip

Fig.6). The inclusion of mid plane GVs affects the flow path for all the different cases. Mid
sc

plane GVs increase the angle of attack at downstream blade for all flow coefficient
nu

compared to the reference geometry. This results in the creation of vortices, early flow
Ma

separation and drop in performance.

The effect of mid plane GVs on the rotor blades can be understood by comparing
ed

the velocity contours with reference geometry. Figure 8 shows the velocity contours of
pt

the reference blades and the four different cases at 75% blade span at flow coefficient
ce

0.225. For all the geometries, the flow hits the leading edge of the upstream blade with
Ac

relative velocity. So the velocity contours around the upstream blade are almost similar

for all the cases. As the flow hits the leading edge, the stagnation point at the leading

edge of the pressure side has zero velocity. A low-pressure zone is created near the

14

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

leading edge of the suction surface which can be seen with the high-velocity contours.

For reference geometries, as the flow passes through the upstream blades, the flow

moves without any obstruction and hits the downstream blade leading edge with relative

velocity. A similar velocity contour pattern is observed in the downstream blades,

d
ite
although the sizes of the contours are bigger compared to upstream blades. For Ref-1,

ed
the area of velocity contours for velocity up to 160 m/s is 13.2 x 10-5 m2 near the suction

surface leading edge of the downstream blade and 2.9 x 10-5 m2 near suction surface

py
leading edge of the upstream blade. The similar values for Ref-2 are 9.52 x 10-5 m2 and

Co
2.59 x 10-5 m2 respectively. The velocity near the leading edge of suction surface reaches

ot
close to the incompressible limit (0.4~0.5Ma), however, not near the transonic shock
tN
limit. The area of such high velocity contour is very small. The discretization of mass flow
rip

terms in mass conservation equation in ANSYS CFX is made as implicit as possible to take
sc

care of this compressibility effect[29]. The main difference between the reference
nu

geometry and others (with GV) can be seen in the downstream blades. With the inclusion
Ma

of GVs in the mid plane, the flow gets obstructed after the upstream blade and a

significant difference is observed in the downstream blade contours. The angle of attack
ed

and relative velocity changes due to the GVs. While the high velocity contours are
pt

concentrated only near the suction side leading edge for reference geometries, such
ce

contours are more spread out in the downstream region for the turbines with GVs.
Ac

Figure 9 shows the tangential vorticity contours at the mid chord plane for two

different flow coefficient before (0.225) and after (0.275) stall. Vorticity is the microscopic

measurement of rotation of fluid element at any point. It defines the strength of local

15

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

circulation in any flow field. For low flow coefficient, the suction surface of upstream

blade has high strength of vorticity contour near the tip surface of all geometries. This is

due to the tip leakage vortex generated from the tip gap. For downstream blade, high

strength of vorticity contours can be observed throughout the span of the blade along

d
ite
with the tip region. The vorticity contours along the span of the blade arise due to the

ed
abrupt high velocity in the downstream section as described in Fig 9. For post stall

condition at higher flow coefficient, the high strength vorticity contours are distributed

py
throughout the blade span for both upstream and downstream blades. As a result, the

Co
flow cannot follow the rotor blade walls till the trailing edge and separation occurs near

ot
the mid chord length of the blade. This results in sudden drop of torque and efficiency
tN
after the stall point.
rip

The generation and leakage aspect of vorticity around the blade surface can be
sc

better understood by the isosurfaces of Q-criterion, where Q stands for the second
nu

invariant of velocity gradient tensor. Q-criterion identify vortices as connected fluid


Ma

region where vorticity magnitude is more than strain rate magnitude [34]. Figure 10

shows the isosurfaces of Q-criterion differentiated by velocity magnitude at the suction


ed

surface of the blades for two different flow coefficient. At low flow coefficient (  = 0.225),
pt

for both Ref-1 and Ref-2, the vortices are seen only near the trailing edge (trailing edge
ce

vortices) and near the tip section (tip leakage vortices). For other cases, at  = 0.225, the
Ac

vortices start to generate much before the trailing edge as the angle of the attack

increases due to the presence of mid plane GVs. At higher flow coefficient,  = 0.275, the

tip vortices are more prominent, and starts from the leading edge. On the suction side of

16

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

blade, the isosurfaces reveal separation of the boundary layer. At lower flow coefficient,

for turbines without GV, such separation occurs only near trailing edge. However, for

turbines with GV, the vortices with higher velocity is generated near the leading edge and

distributed throughout the span of the blade. At higher flow coefficient, after the stall

d
ite
point, the vortices are bigger in size and randomly scattered along the suction surface of

ed
the blades.

Figure 11 shows the streamline distribution at 75% blade span for two different

py
flow coefficients. The flow pattern can be explained from the streamlines around the

Co
rotor blade and the GVs. For GVs with concave side to the leading edge (case A and case

ot
C) a vortex is generated near the leading edge of GV and it proceeds towards the center
tN
of the GV. Such vortex are generated due to the presence of GV in mid plane. Instead of
rip

helping the flow to follow the GVs, here the GVs act as a blockage to the flow. For case B
sc

and case D, where the GVs concave side is towards the trailing edge, the vortex is
nu

generated in the center and towards the trailing edge of GV respectively. Such orientation
Ma

of GVs (case B & D) changes the angle of attack for downstream blade. As a result the flow

separation occurs near the trailing edge of downstream blade for case B which is absent
ed

in case A.
pt

At higher flow coefficient, the angle of attack is high for both upstream and
ce

downstream blade. As a result, flow separation can be observed near the leading edge on
Ac

the suction side of upstream blade for all the cases. As the GVs are stationary, the vortex

formation near the GVs remains almost similar for both the flow coefficients. At higher

flow coefficient, the flow separation in downstream blades occur very near to the leading

17

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

edge. This is visible for all the cases. This is due to the higher angle of attack as well as due

to the presence of GVs. Such early flow separation leads to the drop in performance for

the turbine.

In contrast to the conventional turbines, where the GVs are placed according to

d
ite
the blade inlet angle, in this case the GVs are placed based on the symmetry of the

ed
geometry due to bi directional nature of the turbine. As a result, the mid-plane GVs do

not assist in reducing the angle of attack for downstream blades. Rather the mid-plane

py
GVs cause abrupt flow separation and vortex generation near the surface of GVs. As a

Co
result, use of mid-plane GV does not show any improvement in the performance

ot
compared to the base geometry.
tN
Figure 12 shows the blade loading curve at 50% of blade span at flow coefficient
rip

0.225. The blade loading curve shows the pressure distribution from leading edge to
sc

trailing edge of the rotor blade. For any aerofoil, the pressure distribution around surface
nu

is a qualitative measure of the lift force generated by the aerofoil. The tangential force
Ma

generated by the rotor blade is a function of the lift and drag force and the angle of attack.

So the pressure distribution around the Wells turbine blade can give a measure of the
ed

torque generated by the blade. It can be seen that, for reference blades, the pressure
pt

distribution is similar for both upstream and downstream blades. This suggests that the
ce

total torque generated is shared equally by the upstream and downstream blade. The
Ac

pressure distribution for the four cases is less than the two reference geometries. This

suggests that that the total torque generated is less for the cases with GVs compared to

the reference geometry. Also, for the cases with GV, the downstream blade pressure

18

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

coefficient values are less compared to the upstream blade. This is due to the obstruction

in flow caused by the GVs in mid plane. For Case C, the pressure coefficient on upstream

blade is very less. The pressure coefficient is close to zero or negative in the pressure

surface. This leads to a drop in performance for this case.

d
ite
CONCLUSION

ed
Four different geometries of biplane Wells turbine with mid plane guide vane are

analyzed numerically and compared with two reference geometries without guide vanes.

py
The numerical analysis is carried out in commercial CFD code ANSYS. The flow

Co
characteristics around the blade and the mid plane is discussed in detail. The guide vanes

ot
in the mid plane increase the angle of attack for the downstream blades. Thus the
tN
incipient and dynamic stall points shifts towards lower flow coefficient for downstream
rip

blade. For reference turbines and Case A and C, the stall occurs at flow coefficient 0.250.
sc

Whereas, for cases B and D, the stall occurs before this point. As a result, Biplane turbine
nu

with mid-plane guide vane have less torque compared to the reference turbines. This
Ma

leads to overall decrease of turbine performance. Among the cases studied, Case A has

better performance than other three cases. The guide vanes in the mid plane creates
ed

obstruction to the flow path. Velocity of flow past guide vane is highly irregular and
pt

random in nature. For the cases with guide vanes, the flow separation area on suction
ce

side of the blade and the size of the vortices near tip are more compared to the reference
Ac

cases. One main reason for poor performance of guide vanes is the zero stagger angle

which is a constraint for bidirectional airflow. Biplane Well turbine with mid plane guide

vane having variable stagger angle can improve the performance of the turbine.

19

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

NOMENCLATURE

α blade inlet angle, degree

d
β blade exit angle, degree

ite
c chord length, mm

ed
CL lift coefficient

py
CT torque coefficient

Co
F tangential force, N

ot
G gap between rotor and stator blades, mm
tN
Ko lift interference factor
rip

 flow coefficient
sc

ρ density of fluid medium, kg/m3


nu

Ma Mach no.
Ma

Q volume flow rate, m3/sec

s cascade pitch, mm
ed

T total torque, N-m


pt
ce

u rotor velocity, m/s


Ac

v fluid inlet velocity, m/s

V fluid exit velocity, m/s

W relative velocity of fluid, m/s

20

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

ω rotational speed, rpm

θ circumferential angle between blades, degree

z Zweifel coefficient

∆𝑃0 pressure drop across rotor, Pa

d
ite
∆𝑃0∗ pressure drop coefficient

ed
py
SUBSCRIPTS

Co
0,1,2,3 meridional position in the

biplane turbine

ot
tN
a air

b rotor blade
rip

g guide vane
sc

t rotor tip
nu
Ma

ACRONYMS:

BWT Biplane Wells Turbine


ed

GCI Grid Convergence Index


pt
ce

GV Guide Vane
Ac

OWC Oscillating Water Column

RANS Reynold’s Averaged Navier Stokes

21

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

ACKNOWLEDGEMENT

Comments by the two anonymous reviewers are gratefully acknowledged.

d
ite
ed
py
Co
ot
tN
rip
sc
nu
Ma
ed
pt
ce
Ac

22

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

REFERENCES

[1] Falcao, A. F. D. O., and Henriques, J. C. C., 2016, “Oscillating-Water-Column Wave


Energy Converters and Air Turbines: A Review,” Renewable Energy, 85, pp. 1391–
1424.

d
[2] Thakker, A., Frawley, P., Khaleeq, H. B., and Bajeet, E. S., 2001, “Comparison of

ite
0.6m Impulse and Wells Turbines for Wave Energy Conversion under Similar
Conditions,” Proc. 11th International Offshore and Polar Engineering Conference,

ed
Norway, pp. 630–633.
[3] Shehata, A. S., Xiao, Q., Saqr, K. M., and Alexander, D., 2017, “Wells Turbine for
Wave Energy Conversion: A Review,” International Journal of Energy Research,

py
41(1), pp. 6–38.
[4] Raghunathan, S., 1995, “The Wells Air Turbine for Wave Energy Conversion,”

Co
Progress in Aerospace Sciences, 31(4), pp. 335–386.
[5] Suzuki, M., and Arakawa, C., 2008, “Influence of Blade Profiles on Flow around
Wells Turbine,” International Journal of Fluid Machinery and Systems, 1(1), pp.

ot
148–154.
tN
[6] Kim, T. H., Setoguchi, T., Kaneko, K., and Raghunathan, S., 2002, “Numerical
Investigation on the Effect of Blade Sweep on the Performance of Wells Turbine,”
Renewable Energy, 25(2), pp. 235–248.
rip

[7] Starzmann, R., and Carolus, T., 2014, “Effect of Blade Skew Strategies on the
Operating Range and Aeroacoustic Performance of the Wells Turbine,” Journal of
sc

Turbomachinery, 136(011003-1), pp. 1–11.


[8] Govardhan, M., and Chauhan, V. S., 2007, “Numerical Studies On Performance
nu

Improvement Of Self-Rectifying Air Turbine For Wave Energy Conversion,”


Engineering Applications of Computational Fluid Mechanics, 1(1), pp. 57–70.
Ma

[9] Soltanmohamadi, R., and Lakzian, E., 2016, “Improved Design of Wells Turbine for
Wave Energy Conversion Using Entropy Generation,” Meccanica, 51(8), pp. 1713–
1722.
ed

[10] Gato, L. M. C., and Falcao, A. F. D. O., 1990, “Performance of the Wells Turbine with
a Double Row of Guide Vanes,” JSME International Journal, 33(2), pp. 265–271.
[11] Takao, M., Setoguchi, T., and Kaneko, K., 1996, “Performance of Wells Turbine with
pt

Guide Vanes for Wave Energy Conversion,” Journal of Thermal Science, 5(2), pp.
ce

82–87.
[12] Halder, P., and Samad, A., 2014, “Effect of Guide Vane Angle on Wells Turbine
Ac

Performance,” Proc. of the ASME 2014 Gas Turbine India Conference,New Delhi,
India, pp. 1–7.
[13] Takao, M., Setoguchi, T., Kim, T. H., Kaneko, K., and Inoue, M., 2001, “The
Performance of a Wells Turbine with 3D Guide Vanes,” International Journal of
Offshore and Polar Engineering, 11(1), pp. 72–76.
[14] Setoguchi, T., Santhakumar, S., Takao, M., Kim, T. H., and Kaneko, K., 2001, “Effect
of Guide Vane Shape on the Performance of a Wells Turbine,” Renewable Energy,

23

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

23, pp. 1–15.


[15] Setoguchi, T., Takao, M., and Kaneko, K., 2000, “A Comparison of Performances of
Turbines for Wave Power Conversion.,” International Journal of Rotating
Machinery, 6(2), pp. 129–134.
[16] Raghunathan, S., 1993, “The Prediction of Performance of Biplane Wells Turbine,”
Proc. 3rd International Offshore and Polar Engineering Conference, Singapore, pp.
167–175.

d
[17] Ghisu, T., Puddu, P., and Cambuli, F., 2016, “Physical Explanation of the Hysteresis

ite
in Wells Turbines: A Critical Reconsideration,” ASME J. Fluids Eng., 138(11), pp.
111105-111105–9.

ed
[18] Mamun, M., Kinoue, Y., Setoguchi, T., Kim, T. H., Kaneko, K., and Inoue, M., 2004,
“Hysteretic Flow Characteristics of Biplane Wells Turbine,” Ocean Engineering,
31(11–12), pp. 1423–1435.

py
[19] Kinoue, Y., Kim, T. H., Setoguchi, T., Mohammad, M., Kaneko, K., and Inoue, M.,
2004, “Hysteretic Characteristics of Monoplane and Biplane Wells Turbine for

Co
Wave Power Conversion,” Energy Conversion and Management, 45(9–10), pp.
1617–1629.
[20] Kaneko, K., Setoguchi, T., Hamakawa, H., and Inoue, M., 1991, “Biplane Axial

ot
Turbine for Wave Power Generator,” International Journal of Offshore and Polar
tN
Engineering, 1(2), pp. 122–128.
[21] Shaaban, S., 2012, “Insight Analysis of Biplane Wells Turbine Performance,” Energy
Conversion and Management, 59, pp. 50–57.
rip

[22] Arlitt, R., Banzhaf, H.-U., Starzmann, R., and Biskup, F., 2009, “Air Turbine for a
Wave Power Station, Patent No. WO2009089902.”
sc

[23] Arlitt, R. G. H., Tease, K., Starzmann, R., and Lees, J., 2007, “Dynamic System
Modeling of an Oscillating Water Column Wave Power Plant Based on
nu

Characteristic Curves Obtained by Computational Fluid Dynamics to Enhance


Engineered Reliability,” Proc. 7th European Wave and Tidal Energy Conference,
Ma

Portugal, pp. 1–8.


[24] Alves, J. S., 2013, “Experimental and CFD Analysis of a Biplane Wells Turbine for
Wave Energy Harnessing,” Master Thesis, KTH Royal Institute of Technology.
ed

[25] Zweifel, O., 1945, “The Spacing of Turbo-Machine Blading, Especially with Large
Angular Deflection,” Brown Boveri Review, 32, p. 12.
[26] Weinig FS, 1964, Section B: Theory of Two-Dimensional Flow through Cascades. In:
pt

Aerodynamics of Turbines and Compressors, Princeton university Press.


ce

[27] Torresi, M., Camporeale, S. M., and Pascazio, G., 2009, “Detailed CFD Analysis of
the Steady Flow in a Wells Turbine Under Incipient and Deep Stall Conditions,”
Ac

ASME J. Fluids Eng., 131(7), pp. 071103-1–17.


[28] Gato, L. M. ., and Curran, R., 1996, “Performance of the Biplane Wells Turbine,”
Transaction of the ASME, 118(August 1996), pp. 210–215.
[29] Ansys, 2014, “ANSYS CFX-Solver Theory Guide-Release 15.0,” ANSYS, Inc.,
Canonsburg, Pennsylvania.
[30] Douvi C. Eleni, 2012, “Evaluation of the Turbulence Models for the Simulation of
the Flow over a National Advisory Committee for Aeronautics (NACA) 0012 Airfoil,”

24

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Journal of Mechanical Engineering Research, 4(3), pp. 100–111.


[31] Roache, P. J., 1998, “Verification of Codes and Calculations,” AIAA Journal, 36(5),
pp. 696–702.
[32] Celik, I., and Karatekin, O., 1997, “Numerical Experiments on Application of
Richardson Extrapolation with Nonuniform Grids,” Journal of Fluids Engineering,
119(3), p. 584.
[33] Gareev, A., 2011, “Analysis of Variable Pitch Air Turbines for Oscillating Water

d
Column (OWC) Wave Energy Converters,” Ph.D Thesis, University of Wollongong.

ite
[34] Hunt, J. C. R., Wray, A. A., and Moin, P., 1988, “Eddies, Streams, and Convergence
Zones in Turbulent Flows,” Proc. Summer Program, Center for Turbulence Research,

ed
p. 19.

py
Co
ot
tN
rip
sc
nu
Ma
ed
pt
ce
Ac

25

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Figure Captions List

Fig. 1 Ref-1 and Ref-2 BWT

d
Fig. 2 Schematic representation of different geometric cases for analysis

ite
(dimensions in mm)

ed
Fig. 3 Computational domain and mesh

py
Fig. 4 Velocity diagram for BWT with and without GV

Co
Fig. 5 Validation with experimental result

Fig. 6 Comparison of different cases with reference


ot
tN
Fig. 7 Streamlines on suction surface of blade at  = 0.225
rip

Fig. 8 Velocity contour at 75% blade span for  = 0.225


sc

Fig. 9 Tangential vorticity contours at mid chord for different flow coefficient
nu

Fig. 10 Q-criterion isosurface colored by velocity magnitude at different flow


Ma

coefficient
ed

Fig. 11 Streamlines at 75% blade span for different flow coefficient


pt

Fig. 12 Blade loading curve at 50% blade span (  = 0.225)


ce
Ac

26

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

d
ite
ed
py
Fig.1 Ref-1 and Ref-2 BWT

Co
ot
tN
rip
sc
nu
Ma
ed
pt
ce
Ac

27

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

d
ite
ed
py
Co
ot
tN
rip
sc
nu
Ma
ed
pt
ce

Fig.2 Schematic representation of different geometric cases for analysis (dimensions in


Ac

mm)

28

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Pressure
Turbine rotation direction
outlet
1.4C Periodic
5C

d
Velocity 10C
inlet

ite
Downstream blade
Upstream blade

ed
a) Computational domain without GV
Rotating domain

py
Co
Stationary domain

ot
Rotating domain tN
rip

Stationary domain
b) Computational domain with GV
sc
nu
Ma

LE TE
ed
pt

c) Mesh around the rotor blade


ce
Ac

d) Mesh around GVs

Fig.3 Computational domain and mesh

29

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

d
ite
ed
py
Co
ot
tN
rip

Fig. 4 Velocity diagram for BWT with and without GV


sc
nu
Ma
ed
pt
ce
Ac

30

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

0.7 Present result


Numerical, Shaaban (2012)
0.6 Experimental, Gato & Curran(1996)
0.5
0.4

P0*
0.3

d
0.2

ite
0.1
0.05 0.10 0.15 0.20 0.25 0.30 0.35

ed

py
0.16
Present result
Numerical, Shaaban (2012)

Co
0.12 Experimental, Gato & Curran(1996)

0.08
CT

ot
tN
0.04

0.00
rip

0.05 0.10 0.15 0.20 0.25 0.30 0.35



sc

Present result
0.8
Numerical, Shaaban (2012)
nu

0.7 Experimental, Gato & Curran(1996)


0.6
Ma

0.5

0.4
ed

0.3
0.2
pt

0.1
0.05 0.10 0.15 0.20 0.25 0.30 0.35
ce


Fig 5. Validation with experimental result
Ac

31

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

0.7
Ref-1 Ref-2 Case A
0.6 Case B Case C Case D

0.5
0.4

P0*
0.3

d
0.2

ite
0.1

ed
0.05 0.10 0.15 0.20 0.25 0.30

py
0.16

Co
Ref-1 Ref-2 Case A
Case B Case C Case D
0.12

0.08
ot
CT

tN
0.04
rip

0.00
sc

0.05 0.10 0.15 0.20 0.25 0.30



nu

0.8
Ma

Ref-1 Ref-2 Case A


0.7 Case B Case C Case D
0.6
ed

0.5

0.4
pt

0.3
ce

0.2
0.1
Ac

0.05 0.10 0.15 0.20 0.25 0.30


Fig 6: Comparison of different cases with reference

32

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

TE LE TE LE
Ref-1

Ref-2
LE TE
TE LE

d
ite

Case A
Case C

ed
py
Co
Case D

Case B
ot
tN
Downstream Upstream Downstream Upstream
rip

Fig. 7 Streamlines on suction surface of blade at  = 0.225


sc
nu
Ma
ed
pt
ce
Ac

33

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Ref-1

Ref-2
d
ite
ed
py
Case A
Case C

Co
ot
tN
rip
Air flow

sc
Case D

Case B
nu
Ma
ed
pt

Fig 8. Velocity contour at 75% blade span for  = 0.225


ce
Ac

34

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Upstream Downstream Upstream Downstream


Ref-1

Air Flow

Blade
PS SS

d
Case A

ite
ed
High intensity
vortices only

py
near tip

Case B

Co
ot
tN
Case C
rip

High intensity
sc

vortices near
midplane
nu

Case D
Ma
ed

 = 0.225  = 0.275
pt
ce

Fig 9. Tangential vorticity contours at mid chord for different flow coefficient
Ac

35

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

 = 0.225  = 0.275
Downstream Upstream Downstream Upstream
Ref-2 TE LE

PS
PS SS
SS

d
ite
Case A

ed
py
Co
Case B

Downstream Upstream
ot
Downstream
tN
Upstream
LE TE
LE
Ref-1

TE
rip
sc

SS SS
nu
Case C

Ma
ed
Case D

pt
ce
Ac

Fig 10. Q-criterion isosurface colored by velocity magnitude at different flow coefficient

36

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

d
ite
Case A,  = 0.225 Case A,  = 0.275 Case B,  = 0.225 Case B,  = 0.275

ed
py
Co
Air Flow

ot
tN
Case C,  = 0.225 Case C ,  = 0.275
rip
sc
nu
Ma
ed

Case D,  = 0.225 Case D ,  = 0.275


pt
ce

Fig 11. Streamlines at 75% blade span for different flow coefficient
Ac

37

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

2 2
PS PS
0 0

-2 -2

Cp
Cp

-4 -4

d
SS
SS
-6 -6

ite
Ref-1 Ref-2 Case A Ref-1 Ref-2 Case A
Case B Case C Case D Case B Case C Case D
-8 -8

ed
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
X/C X/C

py
Upstream blade Downstream blade

Co
Fig.12 Blade loading curve at 50% blade span (  = 0.225)

ot
tN
rip
sc
nu
Ma
ed
pt
ce
Ac

38

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Table Caption List

Table 1 Geometric parameters

Table 2 Boundary Conditions

d
Table 3 Estimation of uncertainty due to discretization

ite
ed
py
Co
ot
tN
rip
sc
nu
Ma
ed
pt
ce
Ac

39

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Table 1: geometric parameters

Blade profile type NACA0015


Blade chord length 125 mm
8 (4 in each

Rotor Blade
Number of blades
plane)
Solidity at mean 0.64 (0.32 per

d
radius plane)

ite
Tip radius 294 mm
Hub to tip ratio 0.68

ed
Tip clearance 1 mm
Profile type Circular arc

py
Cord length 90 mm
Number of GVs 20

Co
GV

GV thickness 2 mm
Solidity at mean
1.16
radius

ot
Stagger angle tN 0o
rip
sc
nu
Ma
ed
pt
ce
Ac

40

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Table 2: Boundary conditions

Parameter Description
Flow domain Two blades
(One blade/
plane)

d
Fluid (air) Incompressible

ite
Mesh Unstructured
Interface Periodic

ed
Outlet Pressure
Residual 1x10-6

py
criteria
No. of 2000

Co
iterations

ot
tN
rip
sc
nu
Ma
ed
pt
ce
Ac

41

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

Table 3: Estimation of uncertainty due to discretization

Parameters Values
Number of cells N 1/ N 2/ 10.495x106/
N3 4.406x106/

d
2.413x106

ite
Grid refinement factor 𝑟21 1.33
𝑟32 1.22

ed
Computed torque ∅1 0.09261
coefficients (CT) at φ =0.225 ∅2 0.09250

py
corresponding to N1, N2, N3 ∅3 0.09275
Apparent order p 3.6

Co
Extrapolated values ∅21
𝑒𝑥𝑡 0.09267
∅32
𝑒𝑥𝑡 0.09226
Approximate relative error 𝑒𝑎21 0.0011

ot
𝑒𝑎32 0.0027
21
Extrapolated relative error 𝑒𝑒𝑥𝑡 0.064%
tN
32
𝑒𝑒𝑥𝑡 0.260%
21
Grid convergence index 𝐺𝐶𝐼𝑓𝑖𝑛𝑒 0.076%
rip

32
𝐺𝐶𝐼𝑐𝑜𝑎𝑟𝑠𝑒 0.322%
sc
nu
Ma
ed
pt
ce
Ac

42

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Fluids Engineering. Received May 08, 2018;
Accepted manuscript posted September 28, 2018. doi:10.1115/1.4041600
Copyright (c) 2018 by ASME
Journal of Fluids Engineering

d
ite
ed
py
Co
ot
tN
rip
sc
nu
Ma
ed
pt
ce
Ac

43

Downloaded From: http://fluidsengineering.asmedigitalcollection.asme.org/ on 10/04/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like