Sunteți pe pagina 1din 14

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1, pp.

57-70 (2007)

NUMERICAL STUDIES ON PERFORMANCE IMPROVEMENT OF


SELF-RECTIFYING AIR TURBINE FOR WAVE ENERGY CONVERSION
M. Govardhan* and V. S. Chauhan
Department of Mechanical Engineering
Indian Institute of Technology Madras, Chennai 600 036, India
*Email: gova@iitm.ac.in (Corresponding author)
ABSTRACT: Wells turbine is a self-rectifying air turbine capable of converting pneumatic power of the periodically
reversing air stream in Oscillating Water Column (OWC) into mechanical energy. The Wells turbine has inherent
disadvantages; lower efficiency, poorer starting characteristics, higher axial force and lower tangential force in
comparison with conventional turbines. Guide vanes before and after the rotor suggest a means to improve the
tangential force, hence its efficiency. In the present computational investigations, the performance of the Wells turbine
is predicted for constant chord (CONC) rotor, variable chord (VARC) rotor and variable chord rotor using guide vanes
on upstream and downstream side of rotor. The predicted values of pressure drop with flow coefficient shows almost a
linear variation and are in agreement with the experimental results. Power coefficient obtained from the VARC rotor
with and without guide vanes is more than the CONC rotor at all flow coefficients. Due to recovery of rotor exit kinetic
energy by the downstream guide vanes, the pressure drop across the turbine has increased, resulting in higher energy
transfer and consequently higher turbine efficiency in VARC rotor.
Keywords: self rectifying air turbine, constant chord rotor, variable chord rotor, guide vanes, power coefficient,
pressure coefficient, efficiency, stall margin

1. INTRODUCTION
Wave energy is an alternate form of energy, which
is pollutant free and in near future it is likely to be
economically viable. The most frequently proposed
principle for use of wave energy is that of
Oscillating Water Column (OWC). The OWC
device is a rectangular box closed on all sides
except for a submerged opening facing the waves at
front. The water column in this device oscillates
due to wave action and produces periodically
reversing (bi-directional) stream of air.
Wells turbine invented by A.A Wells is a selfrectifying airflow turbine. The turbine is capable of
converting pneumatic power available in OWC into
mechanical energy. It is an axial flow turbine with
untwisted rotor blades of symmetrical airfoil section
set radially at 900 angle of stagger. The turbine
blading is symmetrical with respect to the direction
of flow. Consequently the turbine rotates in same
direction and produces power regardless of which
way the air is flowing. The inherent drawback of
this turbine is that the tangential force is too small
compared to axial component. The axial force is
transmitted as axial thrust requiring proper bearing,
whereas tangential force is the useful component
reflected as torque of the turbine. With the result,

Wells turbine suffers from poor starting and


accelerating characteristics. Guide vanes before and
after the rotor suggest a means to improve the
tangential force, hence efficiency and starting
characteristics.
There are several reports, which give information
on performance of the Wells turbine focusing on
starting and running characteristics (Raghunathan et
al. (1981 a, 1990), Raghnathan and Tan (1983,
1985)). Watterson and Raghunathan (1998) have
studied the effect of solidity on Wells turbine,
pressure drop, torque and efficiency and concluded
that CFD can predict the performance of the Wells
turbine with reasonable accuracy. Takao et al.
(2001) and Thakker and Hourigan (2004) studied
the effect of high solidity on the Wells turbine and
suggested that presence of guide vanes can improve
the turbine efficiency.
There are few reports on computational analysis of
Wells turbine, especially on comparative merits of
different types of turbine geometries (Kim and
Raghunathan (2002), Takao et al. (2006)).
Therefore the present research is undertaken with
the objective to investigate computationally the
flow patterns in Wells turbine and compare the
performance of Wells turbine with different types

Received: 21 Nov. 2006; Revised: 6 Jan. 2007; Accepted: 19 Jan. 2007


57

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

incidence angle along the blade height compared


CONC rotor.
If the same profile of NACA 0021 were to be used
from hub to tip for the VARC rotor, it would have
resulted in thick blade section at the tip and thin
blade section at the hub resulting in excessive
centrifugal forces on the blade to be resisted by thin
sections at the hub. This is not acceptable from
mechanical strength considerations. To match this
requirement, it was decided to use a thicker profile
at the hub (NACA 0020) and thinner profile
(NACA 0010) at the tip. The details of blade
profiles and other geometric parameters are given in
Table1. For guide vanes, blade sections were
created using circular arc profiles. Both inlet guide
vanes (IGV) and downstream guide vanes (DGV)
were fixed at the distance of 80 mm from the rotor.
Thus VARC rotor is a new rotor but still works
under Wells turbine principles. Both CONC and
VARC rotors are subjected to same inlet flow
conditions. As both of these rotors are power
modules, a comparison is made with regards to
power coefficient, pressure coefficient, efficiency
and operating range.

of rotors. The data is compared with the


experimental results of Swaminathan (1990).

2. METHODOLOGY
A commercial code FLUENT 6.1 is used for the
numerical analysis. The steady incompressible 3-D
Navier-Stokes equations were descretized by finite
volume method. Two rotors, namely, constant
chord (CONC) rotor, variable chord (VARC) rotor
are analysed. Raghunathan et al. (1981 a) presented
constant chord Wells turbine characteristics in nondimensional form and concluded that higher
blockage solidity favours the acceleration of the
turbine from rest to operating condition and thicker
profiles like NACA 0021 are suitable from the point
of view of good operating range. Blockage solidity
is defined as the ratio of the area occupied by the
blades in the radial plane to the free annulus area.
Raghunathan and Tan (1983, 1985) studied the
influence of blade profiles for efficient running of
the Wells turbine. The preferred geometry arrived
by them are; blockage solidity = 0.6, hub to tip ratio
= 0.6 and blade profile NACA 0021. In the present
investigations, blockage solidity of 0.6 and hub to
tip ratio of 0.62 were maintained. Though Wells
turbine with constant chord exhibits excellent
starting characteristics, it suffers from flow
separation in the hub region. Experiments by
Raghunathan et al. (1981b) on Wells turbine
cascade with NACA 0021 profile and chord to pitch
ratio of 0.683 indicated that the stalling would
occur at an incidence angle of 110. As axial velocity
is constant from hub to tip and peripheral velocity
increases with radius, the angle of incidence will be
larger near the hub region compared to the tip
region. With the result Wells turbine with constant
chord rotors exhibits flow separation in the hub
region leading to strong wakes and subsequent
mixing of the flow at the downstream side of the
rotor.
At higher flow rates, the separated zone increases
covering from hub region to almost tip region. This
is the main reason for poor performance of constant
chord rotors. This is because the rotors with
constant chord blades have same thickness of
profiles from hub to tip and have very close interval
near the hub region with incidence angles above the
stalling limit. To overcome the drawback a new
rotor with blade chord varying from hub to tip is
investigated which will have more favourable

Table 1. Geometric details of CONC and VARC


rotors
CONC
VARC
Guide
Rotor
Rotor
Vanes
Tip Diameter, 263 mm
263 mm
263 mm
Dt

58

Hub Diameter, 163 mm


Dh

163 mm

163 mm

Hub/Tip Ratio 0.62

0.62

0.62

Chord

66.7 mm

84 mm at 120 mm
Hub
135.5 mm
at tip

Number of
Blades

Solidity
(chord/blade
space)

0.58 at
mean
radius

0.58

1.25 at
mean
raduis

Profile

NACA
0021

NACA
0020 to
NACA
0010

Circular
arc

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

Mesh generation for all geometries was done using


pre-processor GAMBIT 2.0. For mesh generation
of CONC and VARC configurations, whole domain
was divided into three portions; first portion is
upstream of the blade, second is middle portion that
includes blades and the thirds portion is
downstream of the blades. As velocity gradients
near the blade surface, hub and tip wall regions are
always high; a very fine meshing is required in
these regions. In order to get fine meshes, special
size functions were attached to the blade and wall
surfaces. Size function allows to control the size of
the mesh intervals for edges and mesh elements for
faces or volumes.
The following special size
functions were used for meshing the geometry;
meshed size function and curvature size function.
Meshed size function specifies the maximum mesh
element edge length as a function of distance from a
given source entity. To define the meshed size
function, a growth rate of 1.2 was specified. A
growth rate of 1.2 results in 20% increase in mesh
element edge length with each succeeding layer of
elements. This option eliminates the need for
calculating the viscous boundary layer as more
elements are positioned in the vicinity of solid walls.
The growth rate of 1.2 was found to be more than
adequate. Curvature size functions are particularly
useful for highly curved blade surfaces, hub and tip
walls. With the help of this option, GAMBIT
reduces the size of the elements in the highly
curved regions like leading and trailing edges
whereas in the broad centre region, the curvature
function requirement is satisfied by larger elements.
For grid independence study, comparison of result
was made in terms of static pressure coefficient.
The values of size functions were varied that
increased the degree of fineness of the mesh. For
CONC configuration, mesh was tested for 280,000
cells, 310,000 cells and 320, 468 cells. It was found
that there was no significant increase in static
pressure coefficient beyond 320,468 cells (< 1%).
For VARC configuration, mesh was tested for
280,410 cells, 325,780 cells and 339,678 cells.
Further increase in number of cells did not improve
the static pressure coefficient beyond 1%. Similarly
for VARC rotor with guide vanes, the generated
mesh was tested with 340, 456 cells, 360, 789 cells
and 378, 980 cells. Further increase in cells did not
alter the static pressure coefficient value. The final
mesh distribution is shown in Table 2.
Fig 1(b) shows the final mesh on VARC rotor with
guide vanes and Fig. 1(c) shows mesh on plane of

The computational domain for VARC rotor with


guide vanes is shown in Fig. 1(a). Apart from the
blade surfaces, the other surfaces enclosing the
domain are the inlet, outlet, periodic pairs, hub and
the casing surface. Inlet and outlet surfaces were
developed normal to the axis of turbine. These
surfaces were kept at 3 times chord length on either
side of the blade. The periodic curves on one side of
the blades were joined to get one periodic surface.
This periodic surface was then rotated through 600
in case of CONC rotor and 510 in case of VARC
rotor. This was necessary as the pair of periodic
surfaces should be symmetric about the axis of
rotation and the mesh developed on one of the pair
should show one-to-one map with the other pair of
surfaces. The seven surfaces, inlet, outlet, periodic
pairs, hub surface, casing surface and blade surface
constitute the boundary surfaces of the domain.

Fig. 1(a). Computational domain for VARC rotor


with guide vanes
For the geometry of turbine with guide vanes, the
whole computational domain was divided into three
volumes, one volume that includes rotor and one
each for guide vanes on either side of rotor, (Fig.
1(a)). For guide vanes blade sections were created
using circular arc profiles. The untwisted blade
sections were projected in radial directions in order
to get guide vanes volume. Hub and casing
surfaces were created following the same procedure
as used for rotor. As the total number of guide
vanes used were 7, the two periodic surfaces were
51 degree apart from each other.

59

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

constant radius. The effect of size functions is


clearly seen on the blade surfaces, edges and on the
hub wall.

modeled using RNG k- model. This model is


chosen for present investigations in view of its
success in turbulent swirling flow predictions.

Table 2. Grid Details for all domains


Type of Geometry
Total number of
cells
Constant
Chord
Rotor 320,468
(CONC)
Variable
Chord
Rotor 339,678
(VARC)
Variable Chord Rotor with 378,980
Guide Vanes

3. RESULTS AND DISCUSSION

Exit Axial Velocity Coefficient

3.1 Validation
The aim of validation is to show that the present
code used for the analysis is reliable and could be
used with confidence. The experimental results of
Swaminathan (1990) are compared with the present
investigations. A typical graph comparing axial
velocity at the exit of the VARC rotor is shown in
Fig. 2. The results agree very well.
0.16

Experimental (Swaminathan)
Computational

0.12
0.08
0.04
0.00
0.60

0.65

0.70

0.75

0.80

0.85

0.90

0.95

Radius Ratio

Fig. 2. Comparison of exit axial velocity with


experimental results
3.2 Performance characteristics
Fig. 3 shows the variation of pressure coefficient
with flow coefficient for CONC rotor and its
comparison with experimental results. For a given
rotor speed, the non-dimensional value of pressure
coefficient p* is proportional to pressure drop and
flow coefficient is proportional to the volume flow
rate. Predicted values of pressure drop with flow
coefficient show almost a linear variation. The
agreement between predicted and experimental
values is quite good. At higher flow coefficient,
predicted value is more than the experimental value.
The discrepancy may be explained by the fact that
the hub and casing walls were treated as inviscid
surfaces, hence in these regions turbine will be
operating at higher angle of incidence than the
corresponding experimental turbine. In the present
study only one blade was modeled and therefore
interference effects due to neighboring blades are
not taken into account.
The variation of power coefficient for both rotors is
shown in Fig. 4. The power coefficient for the
CONC rotor increases with flow coefficient up to
= 0.2, and thereafter power characteristic drops due
to stalling. On the other hand, power coefficient for
the VARC rotor increases with flow coefficient

Fig. 1(b). Mesh on VARC rotor with guide vanes

Fig. 1(c). Mesh on a plane of constant radius for


VARC rotor with guide vanes
At inlet, velocity inlet was specified whereas at
outlet, static pressure was specified in accordance
with the experimental values. The hub and casing
surfaces were treated as inviscid surfaces and only
blade was treated as viscid surface. For both rotors,
single rotating frame of reference was used. As
guide vanes are stationary, multiple frames of
reference were used for this model. Turbulence is

60

1.00

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

without any stalling characteristics. In addition,


VARC rotor produces almost double the peak
power output compared to the CONC rotor. The
operating range of the VARC rotor is = 0.08 0.29 which is about 30 % more than the CONC
rotor. The only disadvantage with the VARC rotor
is its poor starting characteristics.

80
70

Efficiency

60
50
40

CONC - with out Guide Vanes


VARC - with out Guide Vanes
VARC - with Guide Vanes

30
20
10

0.16
0.14

Experimental (Swaminathan)
Computational

Pressure Coefficient, p

0
0.00

0.12

0.06

0.12

0.15

0.18

0.21

0.24

0.27

In CONC rotor the blade chord (66.7 mm) was kept


constant from hub to tip. Therefore, the chord to
pitch ratio at the hub (0.781) is greater than the
chord-to pitch ratio at the tip (0.484). This results
in non-uniform flow area in the rotor. The opening
available for the flow in the rotor increases from
hub to tip. In addition to the flow area, incidence
angle also varies from hub towards tip. This is due
to the fact that the inlet axial velocity distributions
are more or less uniform in the annulus but the
peripheral velocity increases from hub to tip.
VARC rotor is designed to have constant pitch to
chord ratio of 1.667 from hub to tip. As pitch chord
ratio is constant, chord length varies from hub to tip.
The performance of the VARC rotor indicate that
use of varying chord blades in place of constant
chord blades in a Wells turbine rotor would
improve stalling limit and peak power output. For
given size of the turbine, use of variable chord
blades would double the possible peak power
output with increased operating range.
The
experimental results obtained by Swaminathan
(1990), Dhanasekaran and Govardhan (2001)
suggested the same trend. The reason for late
occurrence of stalling and subsequent improvement
in the range of operation with better efficiency is
mainly due to the presence of better inlet conditions
to the blades arising out of more favourable
incidence angle along the entire blade height
compared to CONC rotor.
If there are no inlet guide vanes, the tangential
velocity at the rotor inlet would be zero. The
momentum transfer in the rotor results in creation
of tangential velocity at the exit of the rotor. In the
absence of any kinetic energy system like
downstream guide vanes (DGV), the energy in this
tangential velocity would be wasted. This is one of

0.04

0.03

0.06

0.09

0.12

0.15

0.18

0.21

0.24

Flow Coefficient,

Fig. 3. Variation of pressure coefficient for CONC


rotor
0.006

CONC Rotor
VARC Rotor

0.005

Power Coefficient, W

0.09

Fig. 5. Effect of guide vanes on efficiency of


VARC rotor

0.08

0.004
0.003
0.002
0.001
0.000
0.00

0.06

Flow Coefficient,

0.10

0.02
0.00

0.03

0.03

0.06

0.09

0.12

0.15

0.18

0.21

0.24

0.27

0.30

Flow Coefficient,

Fig. 4. Variation of power coefficient with flow


coefficient for CONC and VARC rotors
The efficiency of CONC and VARC rotors is
shown in Fig.5. In general, the efficiency of rotors
without guide vanes is low compared to VARC
rotor with guide vanes. The efficiency of CONC
rotor decreases with flow coefficient. The peak
efficiency is 62% at = 0.08 whereas VARC rotor
attains a peak efficiency of 71% at = 0.177.
Efficiency of VARC rotor has improved with guide
vanes, this improvement is more at lower flow
coefficients.
Peak efficiency is about 77%
occurring at = 0.136. By providing guide vanes, a
maximum improvement of six points is obtained
when compared to the same turbine without guide
vanes. The efficiency curve for the VARC rotors
with guide vanes is flat.

61

0.30

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

values at = 0.232 (Fig. 6(d)). With a higher


positive pressure on the pressure surface and higher
negative pressure on the suction surface, the blade
loading at = 0.232 is high compared to the case
with = 0.082. This is expected as turbine
produces more power at higher flow coefficients.
At = 0.232, a region of low pressure starts
developing near the hub side of the blade. Effect
of tip leakage flow on suction surface static
pressure can be seen from the above figure. Tip
clearance effect seems negligible at low flow
coefficient (Fig. 6(c)), but it is rather significant at
higher flow coefficient (Fig. 6(d)) especially in the
region from 30% to 80 % of blade chord length
from the leading edge. Air passes through the tip
gap due to pressure difference across the blade. As
this leakage flow emerges on the suction side, it
rolls up to form tip leakage vortex that changes the
static pressure distribution over the blade surface
near the tip region. As this change in pressure
distribution depends on pressure difference across
the blade, it is found to be more at higher flow
coefficient due to increased pressure difference.
Fig. 7 shows the static pressure distribution on
pressure side of the VARC rotor blade. Similar to
CONC rotor, high pressure region increases as the
flow coefficient is increased. Static pressure
remains almost same along the blade height in midchord region. At = 0.288, (Fig. 7(b)), pressure
distribution is more uniform over the blade pressure
surface as compared to CONC blade. When guide
vanes are incorporated, the magnitudes of pressure
on the pressure surface are less compared to the
same rotor without guide vanes (Figs. 7(c) and (d)).
Fig. 8 shows static pressure distribution on the
suction side of blade surface. Negative pressure
region near the leading edge of the blade increases
as the flow coefficient is increased. The static
pressure distribution with guide vanes at = 0.102
(Fig. 8(c)) is uniform over the entire blade surface
from the leading edge to the trailing edge and from
hub surface to tip surface. But at = 0.243, (Fig.
8(d)), the static pressure seems to be disturbed over
the blade surface.
Positive pressure region
increases from the leading edge to about 55% of the
blade chord.
The static pressure distribution
suggests the advancement of stall due to the
presence of guide vanes.

the reasons for poor performance of the Wells


turbine compared to conventional turbine. The
tangential velocity would be in the opposite
direction to the peripheral velocity. If there were no
inlet guide vanes (IGV), the tangential velocity at
the rotor outlet would be higher by the inlet
tangential velocity, which means even without the
presence of DGV, the exit kinetic energy losses
would be less with the presence of IGV alone.
The flow leaving the rotor is three-dimensional and
the incidence angle to the downstream guide vanes
will also be non-uniform across the annulus
resulting in higher mixing and consequent total
pressure losses. These losses would be higher at
higher flow coefficients. Due to these reasons, the
VARC rotor with guide vanes exhibits considerable
efficiency improvements at lower flow coefficients
compared to higher flow coefficients. In addition to
recovery of tangential velocity, DGV will create
sub-atmospheric static pressure at the exit of the
rotor as the exit of the DGV is exposed to
atmosphere. This results in higher static pressure
drop across the rotor compared to the same rotor
without guide vanes.
3.3 Rotor aerodynamics
Static pressure on blade pressure surface, suction
surface, hub and tip surfaces is presented in Figs. 612 to analyse the blade loading, tip clearance effect,
stagnation point on the blade profile. This analysis
throws light on internal aerodynamics of the
turbines which otherwise is difficult to obtain from
experimental investigations.
Static pressure distribution on pressure and suction
sides of the turbine blade is shown in Fig.6 (a) &
(b). The static pressure in the leading region
increases as the flow coefficient is increased due to
increase in incidence angle. High-pressure region
increases and occupies more area along the chord as
the flow coefficient is increased. At = 0.232, the
high-pressure region exceeds the position of
maximum thickness of airfoil (30.3 % of chord
from the leading edge) leading to stalling of the
turbine. At = 0.082, (Fig. 6(a)) static pressure
varies from hub to tip in mid chord region, whereas
at = 0.232, static pressure almost remains constant
from hub to tip (Fig. 6(b)). A region of high
pressure begins to develop near the hub surface of
the blade at this flow coefficient. Fig. 6(c-d) shows
static pressure distribution on suction surface of
blade. Static pressure in general has more negative

62

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

No

1
2
3
4
5
6
7
8

p*
0.0387
0.0379
0.0348
0.0426
0.0487
0.0543
0.0582
0.0621

9
10
11
12
13
14
15

p*
0.0660
0.0699
0.0738
0.0972
0.1127
0.1361
0.1400

No
1
2
3
4
5
6
7

N/m2
0.0995
0.1096
0.1197
0.1298
0.1399
0.1500
0.1602

No
9
10
11
12
13
14
15

N/m2
0.1804
0.1905
0.2008
0.2107
0.2208
0.2309
0.2714

0.1702

No

No

1
2
3
4
5
6
7

p*
-0.1501
-0.1310
-0.1096
-0.1013
-0.0839
-0.0583
-0.5400

p*
-0.0499
-0.0411
-0.0194
-0.0026
0.0144
0.0273
0.0401

-0.0497

No
1
2
3
4
5
6
7

p*
-0.3119
-0.2745
-0.2183
-0.1663
-0.1188
-0.1062
-0.0815

-0.0690

No

(a) = 0.082

(b) = 0.232

Pressure surface
9
10
11
12
13
14
15

(c) = 0.082
No
9
10
11
12
13
14
15

(d) = 0.232

Suction surface
Fig. 6. Static pressure contours of CONC rotor

63

p*
-0.0625
-0.0542
-0.0254
-0.0126
0.0057
0.0429
0.0554

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

No
1
2
3
4
5
6
7
8

p*
0.0348
0.0387
0.0426
0.0465
0.0504
0.0543
0.0582
0.0621

No
9
10
11
12
13
14
15

p*
0.0660
0.0700
0.0737
0.0972
0.1127
0.1360
0.1400

(a) = 0.082
1
2
3
4
5
6
7

p*
0.0692
0.0825
0.1096
0.1197
0.1298
0.1399
0.1472

0.1602

No
1
2
3
4
5
6
7

p*
0.0889
0.0112
0.0164
0.0191
0.0250
0.0285
0.0388

0.0423

No
1
2
3
4
5
6
7

p*
0.0497
0.0693
0.0823
0.1007
0.1149
0.1214
0.1280

0.1345

No

No
9
10
11
12
13
14
15

p*
0.1804
0.1905
0.2008
0.2107
0.1208
0.2309
0.2714

(b) = 0.288

Without guide vanes


No
9
10
11
12
13
14
15

p*
0.0454
0.0493
0.0527
0.0631
0.0908
0.105
0.108

(c) = 0.102
No
9
10
11
12
13
14
15

(d) = 0.243

With guide vanes


Fig. 7. Static pressure contours on pressure surface of VARC rotor
64

p*
0.1410
0.1474
0.1541
0.1606
0.1670
0.1801
0.2388

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

No
1
2
3
4
5
6
7
8

p*
-0.0975
-0.0741
-0.0664
-0.0508
-0.0430
-0.0352
-0.0274
-0.0195

No
9
10
11
12
13
14
15

p*
-0.0118
-0.0040
0.0037
0.0115
0.0193
0.0271
0.0348

(a) = 0.082
1
2
3
4
5
6
7

p*
-0.4230
-0.3340
-0.2748
-0.2455
-0.1864
-0.1716
-0.1568

-0.1417

No
1
2
3
4
5
6
7

p*
-0.0990
-0.0857
-0.0579
-0.0442
-0.0338
-0.0234
-0.0165

-0.0130

No
1
2
3
4
5
6
7

p*
-0.0480
-0.0610
-0.0602
-0.0588
-0.0806
-0.0545
-0.0349

-0.0219

No

No
9
10
11
12
13
14
15

p*
-0.1272
-0.1124
-0.0977
-0.0827
-0.0385
-0.0237
0.0058

(b) = 0.288

Without guide vanes


No
9
10
11
12
13
14
15

p*
0.0096
0.0061
0.0008
0.0042
0.0075
0.0215
0.0285

(c) = 0.102
No
9
10
11
12
13
14
15

(d) = = 0.243

With guide vanes


Fig. 8. Static pressure contours on suction surface of VARC rotor

65

p*
-0.0089
0.0058
0.0043
0.0172
0.0436
0.0497
0.0628

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

No
1
2
3
4
5
6
7
8

p*
-0.0011
-0.0057
-0.0102
-0.0137
-0.0515
-0.0745
0.0677

9
10
11
12
13
14
15

p*
0.0539
0.0493
0.0401
0.0447
0.0264
0.0172
0.0035

No
9
10
11
12
13
14
15

N/m2
0.1079
0.0652
0.0408
0.0285
0.0102
0.0041
0.0019

No

0.0585

(a) = 0.082
No
1
2
3
4
5
6
7
8

N/m2
-0.0080
-0.0158
-0.0265
-0.0385
-0.0568
-0.0995
0.1139
0.1262

(b) = 0.232

Fig. 9. Contours of static pressure on hub surface of CONC rotor

No
1
2
3
4
5
6
7

N/m2
0.1452
-0.1197
-0.1071
-0.0830
-0.0774
-0.0650
-0.0563

-0.350

No
9
10
11
12
13
14
15

N/m2
-0.2230
0.0139
-0.0011
0.0031
0.0072
0.0188
0.0369

No
1
2
3
4
5
6
7
8

(a) = 0.082

p*
-0.1527
-0.1461
-0.1153
-0.0906
-0.0721
-0.0413
-0.0352
-0.0290

No
9
10
11
12
13
14
15

(b) = 0.232

Fig. 10. Contours of static pressure on tip surface of CONC rotor

66

p*
-0.0166
-0.0104
0.0080
0.0141
0.0511
0.0634
0.0667

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

No
1
2
3
4
5
6
7
8

p*
-0.0011
-0.0057
-0.0102
-0.0137
-0.0515
-0.0745
0.0677
0.0585

No
9
10
11
12
13
14
15

p*
0.0539
0.0493
0.0401
0.0447
0.0264
0.0172
0.0035

No
1
2
3
4
5
6
7
8

p*
-0.0219
-0.0487
-0.0624
-0.0762
-0.1440
-0.1712
-0.2526
0.2087

No
9
10
11
12
13
14
15

p*
0.1952
0.1816
0.1539
0.1409
0.1273
0.0323
0.0187

(b) = 288

(a) = 0.082

Without guide vanes

1
2
3
4
5
6
7

p*
0.0527
0.0562
0.0596
0.0423
-0.0373
-0.0130
-0.0231

-0.0026

No

No
9
10
11
12
13
14
15

p*
0.0008
0.0077
0.0181
0.0285
0.0458
0.0423
0.0388

1
2
3
4
5
6
7

p*
0.1475
0.1410
0.1410
0.1280
-0.1196
-0.0675
-0.0480

-0.0154

No

No
9
10
11
12
13
14
15

p*
-0.0936
0.0041
0.0237
0.0367
0.0627
0.1345
0.0648

(d ) = 0.232

(c) = 0.102

With guide vanes


Fig. 11. Contours of static pressure on hub surface of VARC rotor

67

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

(a) = 0.082
N0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15

p*
-0.1483
-0.0739
-0.0705
-0.0739
-0.0671
-0.0570
-0.0457
-0.4008
-0.0333
-0.0232
-0.0164
-0.0029
0.0005
0.0375
0.0410

(b) = 0.288
N0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15

(c) = 0.102

p*
-0.2572
-0.2215
-0.0566
-0.0366
-0.0266
0.0343
0.0134
0.0334
0.0445
0.534
0.0734
0.0875
0.1018
0.1601
0.1312

N0
1
2
3
4
5
6
7
8
9
10
11
12

Without IGV

p*
-0.0476
-0.0268
-0.0130
-0.0061
0.0008
0.0042
0.0110
0.0215
0.0354
0.0388
0.0354
0.0388

(d) = 0.243
N0
1
2
3
4
5
6
7
8
9
10
11
12

p*
0.0432
0.0528
0.0566
0.0628
0.0693
0.0758
0.0886
0.0953
0.0840
0.0693
0.0562
0.0366

With IGV

Fig. 12. Contours of static pressure on tip surface of VARC rotor


60 % of the chord negative pressure exists over the
blade tip and negative pressure magnitude decreases
from the leading edge towards the trailing edge. On
the other hand, positive pressure region near the
trailing edge decreases as the flow coefficient is
increased. With the result, tip leakage flow will be
considerably higher in the leading edge portion at
lower flow coefficient. But, as the flow coefficient
increases, leakage flow region advances towards
trailing edge causing large mass flow of air to leak
through the gap.
The entrainment of the fluid in the trailing edge
region towards suction surface is less in VARC rotor
without guide vanes (Figs. 11(a) and (b)) compared

In order to find out the stagnation point on the blade


profile and effective leakage area through the gap,
static pressure distribution on hub and blade tip are
plotted and shown in Figs.9-13. Stagnation point for
CONC rotor is very near to the leading edge of the
turbine blade at = 0.082, (Fig. 9(a)). As the flow
coefficient increases stagnation point moves away
(towards mid portion of the blade) from the leading
edge (Fig. 9(b)). Blade passage flow enters through
the trailing edge to mid chord portion of the suction
surface. This entrainment is considerably higher at
higher flow coefficient (near stall region). From Fig.
10 where static pressure on tip surface is presented, it
can be observed that from the leading edge to about
68

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

to CONC rotor. With the result wakes shed by the


rotor will be less giving rise to lower total pressure
loss and higher efficiency. When VARC rotor is
fitted with guide vanes, (Fig. 11(d)), the pressure
distribution is non-uniform on the suction surface.
The flow coming out of IGV will not be uniform and
consequently the turbine rotor receives flow with
varying incidence angles. The incidence angle and
its variation will be more at higher flow coefficient.
Guide vanes themselves generate total pressure loss
in addition to offering resistance to the flow. With
the result, the highest flow coefficient operable will
be less ( = 0.243 with guide vanes and = 0.288
without guide vanes). Hence turbine with guide
vanes stalls early. Non-uniform pressure observed on
the suction surface and positive pressures observed
on the blade tip at = 0.243 (Fig. 12(d)) suggests the
beginning of stall. At other flow coefficients (not
presented here), the entrainment and pressure nonuniformity were found to be less compared to CONC
rotor. The efficiency curves in Fig.5 indicate the
same trend. The turbine without guide vanes does not
exhibit the above trend (Figs. 12(a) and (b)). The
magnitude of negative pressure near the leading edge
and positive region near the trailing edge increase as
the flow coefficient is increased.

NOMENCLATURE

4Q

cm

average axial velocity

d
l
P0
PI
ps
p0

diameter [m]
blade chord [m]
power output to the turbine [W]
power input to the turbine [W]
static pressure [N/m2]
total pressure [N/m2]

p*

pressure coefficient

flow rate [m3/s]

W*

power coefficient

angular velocity [rad/s]

efficiency 0
PI

d2 - d2
t
h

po
2
2
dt

P0

d 5t

flow coefficient m
ut
Subscripts
h hub
o outlet
Superscript
* non-dimensionalised.

4. CONCLUSIONS
The ability of CFD to predict the performance of
Wells turbine has been tested. The predicted turbine
aerodynamics are consistent with the experimental
data. Wells turbine with VARC has wider operating
range than CONC rotor. The power coefficient of this
rotor is almost twice that of CONC rotor. In VARC
rotor, efficiency of 60 % and above is possible for
flow coefficient ranging from 0.1 to 0.28, which is a
considerable improvement over rotors having
constant chord blades.
Due to generation of
tangential velocity at rotor inlet by inlet guide vanes
and due to the recovery of exit kinetic energy by the
downstream guide vanes, the pressure drop across the
turbine is increased resulting in higher efficiency.
The peak efficiency of turbine with VARC rotor with
guide vanes is about 77 %. Static pressure in the
leading edge region increases with the flow
coefficient for all rotors. The high pressure region
increases and occupies more area along the chord as
flow coefficient is increased leading to stalling of the
turbine. CONC rotor stalls early compared to VARC
rotor.

REFERENCES
1. Govardhan M, Dhanasekaran TH (2001). Effect
of guide vanes on the performance of a self
rectifying air-turbine with constant chord and
variable chord rotors. Renewable Energy 26:
201-219.
2. Kim TH, Raghunanthan S (2002). Numerical
investigation on the effect of blade sweep on the
performance of Wells turbine. Renewable Energy
25(2): 235-248.
3. Raghunathan S, Tan CP, Wells NAJ,
Mc.Illhagger DS (1981a). Efficiency, starting
torque and prevention of run-away with Wells
self-rectifying turbines. Proceedings of Second
International Symposium on Wake and Tidal
Energy, BHRA. 23-25 September 1981,
Cambridge, U.K., 207-217.
4. Raghunathan S, Tan CP, Wells NAJ (1981b).
Wind tunnel tests on airfoils in tandem cascade.
AIAA, 19:1490-1492.

69

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 1 (2007)

5. Raghunathan S, Tan CP (1983). Aerodynamic


performance of Wells turbine. Energy 7(3):226230.
6. Raghunathan S, Tan CP (1985). Effect of blade
profile on the performance of the Wells self
rectifying air turbine. Heat and Fluid Flow 6: 1722.
7. Raghunathan S, Setoguchi T, Kaneko K (1990).
Prediction of aerodynamic performance of Wells
turbines from aerofoil data. Trans. ASME,
Turbomachinery 112:792-795.
8. Swaminathan G (1990). Performances and Flow
Investigations on Wells Turbine. Ph. D. Thesis,
I.I.T. Madras, India.
9. Takao M, Ajit T, Rahil A, Setoguchi T (2006).
Effect of blade profile on the performance of a
large scale Wells turbine for wave energy
conversion. Sustainable Energy 25(1):53-61.
10. Takao M, Setoguchi T, Kim TH, Kaneko K,
Inoue M (2001). The performance of a Wells
turbine with 3D guide vanes. Off Shore and Polar
Engineering 11(1):64-71.
11. Thakker A, Hourigan F (2004). Modeling and
scaling of the impulse turbine for wave power
applications. Renewable Energy 29(3):305-317.
12. Watterson JK, Raghunathan S (1998). Computed
effect of solidity on Wells turbine performance.
JSME Series B 41(1):199-205.

70

S-ar putea să vă placă și