Sie sind auf Seite 1von 6

L. M. C.

Gato
Department of IVIectianical Engineering,
Instituto Superior T^cnico,
Universidade T6cnica de Lisboa,
1096 Lisbon, Portugal

V. Warfield
A. Thakker
Department of l\/leciianicai and Aeronautical
Engineering,
University of Limerick,
Limerick, Ireland

Performance of a High-Solidity
Wells Turbine for an OWC Wave
Power Plant
The paper describes an experimental investigation, and presents the results of the
aerodynamic performance of a high-solidity Wells turbine for a wave power plant.
A monoplane turbine of 0.6 m rotor diameter with guide vanes was built and tested.
The tests were conducted in unidirectional steady airflow. Measurements taken include flow rate, pressure drop, torque, and rotational speed, as well as velocity and
pressure distributions. Experimental results show that the presence of guide vanes
can provide a remarkable increase in turbine efficiency.

Introduction
Potentially, one of the most successful devices used in the
harnessing of wave energy has been the oscillating water column (OWC) wave energy converter. The OWC is a chamber,
either floating or bottom standing, with the immersed end
opened to the action of the sea (see Gato et al., 1991a). A
reciprocating airflow is created by the action of the free surface
of the water within the chamber. The conversion of this airflow
into mechanical energy may be achieved by means of a number
of devices.
A reciprocating airflow may be rectified in order to produce
unidirectional flow, by a series of nonreturn valves. This unidirectional flow may be used to drive a conventional turbine such
as a Francis turbine. Although the Francis turbine may reach a
high peak efficiency, its average cyclic efficiency in oscillating
airflow is likely to be considerably small, since the turbine will
be working far from the best efficiency point during a large
part of the cycle (Raghunathan et al., 1985).
No rectifying valve system is required if a self-rectifying air
turbine is used. Versions of such turbines have been proposed
and built; namely, McCormick's counter-rotating turbine (Richards and Weiskopf, 1986; McCormick and Surko, 1989), the
impulse turbine with self-pitch-controlled guide vanes (Setoguchi et al., 1993), and the Wells turbine (Raghunathan et al.,
1985, 1989; Gato and Falcao, 1984, 1988). The rotor of the
latter consists basically of a set of symmetrical airfoil blades,
fixed to a hub, with their plane of symmetry normal to the axis
of rotation. Versions of the Wells turbine have been studied
with and without stationary vanes (Gato and Falcao, 1990). In
biplane or multiplane Wells turbines, the rotor blades are placed
in more than one plane, in order to allow for greater total bladed
area (Raghunathan and Tan, 1983; Kaneko et al., 1991; Gato
et a l , 1991b). A modified Wells turbine with variable pitch
rotor blades was found to provide a way of controlling the
airflow rate and pressure difference independently from each
other, and therefore it can be used to enable an OWC wave
energy device to be phase-controlled (Sarmento et al., 1990).
In addition to this, the capability of controlling the blades setting
angle can constitute an effective way to extend the flow range
within which the turbine can respond with fairly high efficiency
(Gato and Falcao, 1989; Gato et al., 1991a).
The paper describes an experimental study concerning the
design of a Wells turbine to equip the European wave energy
pilot plant to be built on the island of Pico, Azores (Falcao et
Contributed by the Advanced Energy Systems Division for publication in the
JOURNAL OF ENERGY RESOURCES TECHNOLOGY . Manuscript received by the AES

Division, October 14,1993; revised manuscript received June 12,1996. Associate


Technical Editor: R. J. Krane.

al., 1991). The input data for the turbine design were obtained
from experimental tests conducted on a 1/35-scale model of
the wave power plant in an irregular wave basin. The design
method for the air turbine is based upon the quasi-three-dimensional analysis of the flow and uses the minimization of the
outer rotor diameter as an optimization criteria, thus yielding
the turbine dimensions and rotational speed. A comparison of
the turbine performance and size using several geometries (Gato
et al., 1991) showed that the most suitable geometry to accommodate the given design conditions was the monoplane rotor
with guide vanes.
A turbine of 0.6-m diameter was jointly manufactured both
at the University of Limerick and at Instituto Superior Tecnico
(1ST), Lisbon. Two sets of rotor blades were manufactured,
each set giving the same solidity. The first set of blades was of
constant thickness with a NACA 0015 profile, along the blade
length, while the second set was of linearly variable thickness
with NACA 0021 at the hub and NACA 0009 at the tip. The
turbine was fitted with two sets of guide vanes, which could be
removed for comparative testing. Experiments were conducted
with unidirectional steady airflow, in the test rig at 1ST.
Details of the guide vane design are given in Section 2. The
experimental procedure as well as the test rig used are described
in Section 3. In Section 4, experimental results are presented and
compared for three different turbine configurations: constant
thickness rotor blade and variable thickness rotor blade, without
guide vanes; constant thickness rotor blade with guide vanes.
2

Guide Vane Design


The quasi-three-dimensional analysis described in detail in
Gato and Falcao (1990) was used to determine the flow angle
distribution at the exit of the inlet guide vanes, which yields
the required design flow conditions at the entrance to the outlet
guide vanes, i.e., a2(r) = TT - a\{r) (see Gato and Falcao,
1990), Fig. 1. Once a2{r) = TT a i ( r ) is known, the shape
of the guide vanes can be calculated from two-dimensional
theory. Due to manufacturing constraints, a circular arc profile
of constant radius was adopted for the guide vanes. We assume
axial inlet flow and specify the exit flow angle a^ (or, equivalently, fix the inlet flow angle a2 = TT - ai for the second row
of guide vanes and assume axial exit flow). Furthermore, we
impose shock-free flow at the leading edge and Kutta condition
at trailing edge. A panel method code (Ferro, 1990) is then
used to find iteratively the chord-to-pitch ratio and the cascade
stagger angle, for a given profile radius of curvature and thickness. The same cascade geometry is obtained regardless of
whether inlet or outlet guide vanes are considered. The profile
radius of curvature is chosen to obtain a chord-to-pitch ratio
distribution that closely satisfies the Zweifel criterion, applied
DECEMBER 1996, Vol. 1 1 8 / 2 6 3

Journal of Energy Resources Technology

Copyright 1996 by ASME


Downloaded From: http://energyresources.asmedigitalcollection.asme.org/ on 04/09/2015 Terms of Use: http://asme.org/terms

Fig. 1 Two-dimensional cascade geometry and velocity diagrams for


Wells turbine witli guide vanes

to the decelerating flow in the outlet guide vane row (Horlock,


1966; Wilson, 1984).
3

Test Rig and Measuring Technique

3.1 Test Rig. A schematic layout of the test rig is shown


in Fig. 2. The turbine was mounted in a cylindrical annular
duct, with a blade tip clearance of 1 mm. The turbine was
coupled in line through a torquemeter to a motor/generator.
Airflow was provided by a set of three centrifugal fans. The
main fan provided a continuously variable flow rate, by means
of a variable speed controller connected to the fan motor. The
two auxiliary fans were connected in series to boost the maximum flow rate. The air was drawn into the open end of the
annular duct through the turbine, then passed through the plenum chamber. The plenum chamber settled the flow prior to
passing through a calibrated nozzle and then exhausting at the
fan outlets.
The plenum chamber and the nozzle were designed in accordance with the Air Moving and Conditioning Association
(AMCA), standard 210/67. The working section of the nozzle
had a diameter of 0.4 m, the trumpet having a diameter of 0.8
m. The overall length of the nozzle was 0.67 m. The plenum
chamber was a large rectangular box measuring 2.0 m in length
by 1.72 m high, and 1.78 m wide. The air entered through a
0.63-m-dia opening in the front of the chamber, and passed
through the chamber directly into the calibrated nozzle. A honeycomb-type flow straightener, located at the midplane of the

chamber, was used to remove the swirl from the flow leaving
the turbine.
The motor/generator was mounted in line with the turbine.
A stepless variable speed control maintained the turbine at a
pre-set speed, regardless of the flow rate. When the torque on
the turbine was low, or negative, the motor/generator performed
as an electric motor, drawing power from the mains in order to
maintain the set speed. When the torque was positive, the machine behaved as a generator, supplying power to the grid, and
preventing the turbine from overspeeding.
The blades were fixed to the hub by means of a tapered
dovetail. This provides the accurate location of the blades on
the hub, withstands the centrifugal forces exerted when running
at maximum speed, and allows a quick exchange of rotor blades.
Geometric data for the rotor are presented in Table 1.
The guide vane system consisted of two sets of guide vanes,
one set being the mirror image of the other. Handmade from a
brass sheet 1 mm thick, each set of guide vanes consisted of
60 blades, with a circular arc profile. The blades were equally
spaced around the circumference of an inner and outer ring.
These rings were designed to fit neatly within the cylindrical
annular duct that formed the turbine casing. The guide vanes
were designed as described in Section 2 for a dimensionless
flow rate U* = 0.13. The data for the inlet guide vanes are
given in Table 2.
Although the airflow obtained from the wave energy converter is intrinsically transient, the blade passing frequency of
the turbine rotor, under normal operating conditions, is so much
higher than the typical period of a wave that the flow can be
considered quasi-steady (Raghunathan et al, 1985). The turbine tests were then performed with unidirectional steady airflow. Prior to each experimental test, recordings of atmospheric
pressure, and both wet and dry temperatures were taken. These
data were used in the calculation of air density.
3.2 Instrumentation. Torque and rotation speed were
measured by means of an inductive torque transducer and a
photo-electric speed pick-up. The calibration curves for the
torquemeter and the speed pick-up were found to deviate less
than 0.02 Nm and 0.5 rpm within the range of the measured
values. The volume flow rate was calculated from the measured
pressure drop across the nozzle located at the exit of the plenum
chamber. Calibration curves were previously established for a
wide range of flow rates, by traversing the duct with Pitot tubes
along four equally spaced radii at a cross section. The average
discharge coefficient was found to be 0.9526 (Pereira and Pereira, 1991), for Reynolds number above 2.5 X 10''.
All pressure measurements were made using four micromanometers amounting to a total range from 0 to 20 kPa. By
selecting the niost suitable micromanometer for a given pressure
reading, an accuracy of 1 percent of reading or 0.25 Pa, which-

Nomenclature
c = blade chord
E = pressure plus kinetic energy
flux
L = energy loss per unit of time
Ma = Mach no.
p = pressure
Apo = energy available to turbine per
unit volume fluid
ApS = Apo/(pu>^R^)
P = turbine power output
Q = volume flow rate
r, 6, x = cylindrical coordinate system
r* = rlR
R, Ri = outer, inner radius

Re = Reynolds no.
t = cascade pitch
U = inlet flow (average) velocity
U* = U/(uiR), flow rate coefficient
V, W = absolute, relative velocity
(x, y) = Cartesian coordinate system
a I, a'o = inlet, outlet blade angle
a, P = angle of absolute, relative velocity
$ = ratio of minimum and maximum flow rate at which efficiency is ?7 = 0.6r7,x
rj = turbine efficiency

p = density
Lo = rotor angular speed
Subscripts
X, 9 = axial, tangential velocity
component
0, I, 2, B, C = far upstream (in the
atmosphere), upstream
of rotor, downstream of
rotor, downstream of
outlet guide vanes, inside
the plenum chamber
Superscript
* = dimensionless value

264 / Vol. 118, DECEMBER 1996

Downloaded From: http://energyresources.asmedigitalcollection.asme.org/ on 04/09/2015 Terms of Use: http://asme.org/terms

Transactions of the ASME

Schematic Layout of Test Rig

Auxiliary fans

Turbine shaft and support bearings


Exit gui(
Torque transducer

Main fan

IH 4 ^ 4 : # ^ 3 ^

Ahn-'

20 m

1.3 m

Pressure tappings
2.0 n
0.6 m
Fig. 2 Schematic layout of the test rig

ever is the greater, could be obtained for values of pressure


differences above 2 Pa.
A cylindrical three-hole directional probe of 4-mm diameter
was used to take flow velocity and flow direction measurements.
The probe had holes radially drilled at 50 deg, in a cross section
6 mm forward from the rounded end of the tube. A Vernier
protractor and spirit level incorporated into the probe provided
an accurate method of measuring flow direction and a reliable
system of ascertaining a horizontal datum. The instrument was
mounted on a dial height gage which enabled it to be accurately
traversed across the flow field. The three-hold directional probe
was used as a null-type probe to measure yaw angle, as well
as the dynamic and total pressure. The flow direction was determined by rotating the probe about its vertical axis: when the
recorded pressures at both side holes are equal, the horizontal
axis of the probe indicates the flow direction. The total pressure
and the dynamic pressure were obtained, respectively, from the
central hole reading and from the pressure difference recorded
between the central hole and any of the side orifices by means
of a previous calibration in a wind tunnel against a standard
Pitot tube.

rates, at three rotational speed settings, both with and without


guide vanes (Warfield, 1994). The tests involved the collection
of vast quantities of data. To facilitate the reliable and speedy
recording of these data, a computerized data acquisition system
was developed.
With the exception of torque, rotational speed, and flow
angle, all the other values recorded were pressure differences.
All pressure measurements within the same range were recorded
using the same micromanometer. This procedure was adopted
to minimize errors.
All pressure tappings on the rig, as well as the micromanometers used for recording pressure drops, were connected to a
bank of solenoid valves which were controlled by a computer.
When recording a pressure drop, signals from the computer
operated the relevant valves on the lines to link the pressure
drop being measured to the selected recording instrument. The
outputs from all measuring instruments were connected to a
data acquisition board. This board converted the analogue input
from the measuring instruments and generated the digital output
to control the solenoid valves.

3.3 Data Acquisition. The experimental programme required very detailed tests to be conducted on the performance
of the turbine. These tests were carried out over a range of flow

Table 1 Rotor specifications


Outer casing diameter
Hub diameter
Inner/outer diameter ratio
Constant thickness blade
Variable thickness blade
Number of blades
Rotor sohdity at hub
Blade chord

590 mm
400 mm
0.680
NACA 0015
NACA 0021 at hub, NACA 0009 at tip
8
0.79
125 mm

Journal of Energy Resources Technology

Experimental Results

Figures 3 and 4 represent the results from a test on a single


set of stator guide vanes. The guide vanes were tested in situ
without the rotor blades to enable a true assessment of guide
vane performance. Two flow rates were used for this test, whose
average inlet flow velocities, U, were 8.2 m/s and 13.5 m/s.

Table 2 Geometric data for the inlet guide vanes


rlR
al (deg)
oio (deg)
c/t
c (mm)

0.687
88.1
152.3
4.77
101

0.769
87.6
145.3
3.87
92

0.837
87.3
140.9
3.33
86

0.898
86.9
137.5
2.94
81

0.956
86.6
134.9
2.64
78

1.000
86.4
132.9
2.44
75

DECEMBER 1996, Vol. 1 1 8 / 2 6 5

Downloaded From: http://energyresources.asmedigitalcollection.asme.org/ on 04/09/2015 Terms of Use: http://asme.org/terms

1.25
1

~---3:>^

'
U

0.75 1 r

^
- ^ U = 13.Sm/s
- o - U = 8.2ni/s

0.25
0.68

0.76

0.84
*

^ 11

0.92

Fig. 3 Velocity profiles upstream of the stator biades, for two inlet flow
average velocities

Upstream of the guide vanes one velocity profile per flow rate
was measured, whereas downstream seven profiles per flow rate
were measured. The downstream profiles were measured along
radii at 1-deg spacing, over one blade pitch. Figure 3 shows the
dimensionless velocity profile of the incoming flow measured
lAx = - 2 8 0 mm. The profile is quite flat outside the boundary
layers, and was found to be practically unchanged in the presence of other blade rows in subsequent turbine tests. Figure 4
shows the circumferentially averaged values of the flow angle
measured at the exit of the inlet guide vanes {x = - 7 0 mm),
plotted as a function of the radial coordinate. The curve in the
same figure represents theoretically predicted values for the
same conditions. They have been calculated by the quasi-threedimensional method referred in Section 2. It can be seen that
the theory overestimates the flow deflection in the outer region
by about 4 deg.
Results from the overall turbine performance are presented
in Fig. 5. These were recorded at a rotor speed of 1500 rpm.
Based on relative velocity this gives Re 3.9 X 10^ and Ma
i=a 0.14 at the blade tip. The figure shows a dimensionless plot
of stagnation pressure drop across the turbine, and efficiency
against the flow rate. Values of the stagnation pressure drop
across the turbine, Apo, were obtained by measuring the pressure difference between the atmosphere and the plenum chamber. The efficiency is defined as /? = P{Ql^po)~^, where Q is
the flow rate and P is the power output. The power output
includes mechanical losses which were determined by measuring the torque required to drive a bladeless rotor. All values of

Fig. 5 Measured values of efficiency (open symbols) and dimensionless


stagnation pressure drop across the turbine (solid symbols) against flow
rate coefficient for the variable thickness bladed rotor (NACA 0021 at
the hub and NACA 0009 at the tip) without guide vanes and the NACA
0015 bladed rotor with and without guide vanes

torque measured during the tests were then adjusted to compensate for the power consumed in overcoming mechanical losses.
This procedure yields the true torque value developed by the
rotor. At 1500 rpm the torque value measured with a bladeless
rotor was 0.31 Nm. This represents about 5.5 percent of the
maximum measured turbine torque at the same speed.
Figure 5 presents the overall performance of two different
high solidity monoplane configurations, both with and without
guide vanes: constant thickness rotor blade (NACA 0015) and
variable thickness rotor blade (NACA 0021 at the hub and
NACA 0009 at the tip). Both have the same solidity, blade
chord length, and number of blades. It can be seen in Fig. 5
that, for stall-free conditions, both turbines without guide vanes
have very similar experimental curves for A / J * and rj. The main
difference between the aerodynamic performance of these two
rotors is the maximum value of the flow rate prior to a sharp
drop in turbine torque, due to blade stalling. For the Reynolds
number prevailing in the experiments, the results reveal that
stall occurs at f/* 0.23 for the constant thickness bladed
rotor, whereas for the variable thickness bladed rotor this will
occur for a smaller value of [/* ! 0.19. Defining $ =
Uf/Uf, where Uf and U^ are the minimum and maximum
flow rate coefficients, respectively, at which the efficiency is
nominally 77 = 0.6rjrax< we obtain, for these turbines, $ =
0.189 and $ = 0.223, respectively, which clearly shows that a
judicious choice of the rotor blade profile is required to widen
the range of flow rates for which the turbine performs efficiently.
Figure 5 also shows the overall performance of the constant
thickness rotor blade turbine equipped with a double row of
guide vanes. If we compare the monoplane rotor (NACA 0015)
with and without guide vanes, experimental results show that
the presence of guide vanes has the effect of decreasing the
slope of the (approximately straight) curve of stagnation pressure drop versus flow rate. An improvement in the maximum
efficiency of 17.1 percentfrom 0.603 to 0.706was obtained

Table 3 Aerodynamic losses for the monoplane turbine


without guide vanes

0.68

0.76

0.84

0.92

Fig. 4 Measured (symbols) and predicted (curve) radial distribution of


flow angle downstream of the Isolated Inlet guide vanes, for two inlet
flow average velocities

U*

r*

0.095
0.134
0.183
0.219
0.243

0.002
0.003
0.004
0.004
0.005

I*
0.328
0.302
0.271
0.396
0.550

0.072
0.132
0.178
0.166
0.139

v
0.598
0.563
0.547
0.434
0.306

266 / Vol. 118, DECEMBER 1996

Downloaded From: http://energyresources.asmedigitalcollection.asme.org/ on 04/09/2015 Terms of Use: http://asme.org/terms

0.043
0.036
0.044
0.050
0.051

L%

L% (1 - nr

0.034
0.072
0.098
0.090
0.083

0.085
0.165
0.216
0.159
0.120

Transactions of the ASME

Table 4 Aerodynamic losses for the monoplane turbine


with guide vanes

u*
0.098
0.183
0.251

1 *
'-'OX

J*

1*

0.009
0.012
0.016

0.272
0.410
0.763

0.026
0.048
0.183

1*

0.025
0.036
0.045

0.693
0.530
0.038

0.001
0.001
0.002

by the use of guide vanes. However, stall onset was observed


at U* ^ 0.19, which left us with a smaller range of useful flow
rates, $ = 0.226.
The error analysis for the experimental data shows that the
accuracy of the calculated values of the turbine efficiency depends largely onU*. Using the data obtained for the NACA
0015 bladed rotor with guide vanes and considering the uncertainty of the instrumentation presented in Section 3.2, an accuracy of 5.5 and 2.4 percent is predicted for the experimental
values of the turbine efficiency obtained, respectively, at U* =
0.1 and [/* = 0.175 (Warfield, 1994).
Assuming axisymmetric flow as an approximation (or taking
circumferentially averaged values), the kinetic plus pressure
flux across a section normal to the turbine axis is given by
27r

/:['

-{yl + Vj)

- ^ U* = 0.05
- B - U * = 0.1

V ^

^ t - U * = 0.15
- e - U * = 0.2
- ^ U * - 0 25
S,^

--U*=0.28

V,rdr

/ ^

/W

where p, V^, and Ve are taken as functions of r. Numerical


values of E could be obtained from turbine tests at 2000 rpm
1

0.68

'

0.76

0.84

- ^ u = o.os

0.92

(b)

-B-ll=0.1
-<-U=0.15

Fig. 7 Radial distribution of the axial (a) and tangential (b) velocity
components measured at the rotor exit for the NACA 0015 bladed rotor
turbine with guide vanes, for several flow rate coefficients

= 0.25
= 0.28

u
1

3E^
1

0.68

0.76

0.84

0.92

*
r
(a)

- - U * = 0.05
- B - U * = 0.1
- * - U * = 0.15

t.^.

- - U = 0.2
--C*=0.25

\
0.68

0.76

0.84

1
0.92

1
1

*
r
(b)
Fig. 6 Radial distribution of the axial (a) and tangential (b) velocity
components measured at the rotor exit for the NACA 0015 bladed rotor
turbine without guide vanes, for several flow rate coefficients

Journal of Energy Resources Technology

by integration from values of p , V,, and Vg that were measured


at X = 70 mm (section 1) and x = -t-70 mm (section 2) along
the radius, for several values of the flow rate coefficient [/*. In
the absence of guide vanes, the flow was assumed axisymmetric
at sections 1 and 2. For the turbine with guide vanes, the flow
was assumed axisymmetric at section 2 (between rotor and
outlet guide vanes), whereas at Section 1 (between inlet guide
vanes and rotor), a double numerical integration was made
from results of p , V^, and Vg measured along seven radial lines
circumferentially spaced by 1 deg (i.e., over an annular segment
corresponding to one stator blade pitch).
Mechanical energy balances were performed between a section 0, far upstream of the rotor (in the atmosphere), sections
1, 2 and a section C, far downstream of the rotor in the plenum
chamber, for the calculation of both the rotor and the energy
losses upstream and downstream of the turbine rotor. Assuming
gage pressures, it seems adequate to define the following coefficients: Loi = -Ei(ApoQ)~^ for the losses upstream of the
rotor (including guide vanes );Lfc=i+
2(ApoG)"' for the
losses downstream of the rotor; Lji = (^i E2 P ) ( Apofi) '
for the rotor losses.
Results from the foregoing calculations are presented in Tables 3 and 4 for the NACA 0015 bladed rotor with and without
guide vanes. Also shown in the same tables are the axial and
tangential kinetic energy coefficients
np
ApoQ

V L rdr and

VlgV2,rdr
Apo2 JR,

calculated from values of V^ and Vg measured at sections 2 and


DECEMBER 1996, Vol. 1 1 8 / 2 6 7

Downloaded From: http://energyresources.asmedigitalcollection.asme.org/ on 04/09/2015 Terms of Use: http://asme.org/terms

B (x = +280 mm) for the turbine with and without guide vanes,
respectively.
Results from Table 3 show that large exit losses occur for
the turbine without guide vanes. For U* = 0.183, Ljg represents
21.6 percent of total dimensionless turbine losses (1 77).
Comparing the values of Table 3 for axial and tangential kinetic
energy coefficient with the exit loss coefficient for the same
flow rate [/*, it can be found that non-negligible aerodynamic
losses occur in the rods supporting the bearings, downstream
of the rotor, due to large flow incidence angles.
It can be seen from Table 4 that the outlet guide vanes perform with reasonable efficiency within the range of flow rates
for which the stall-free conditions exist at turbine rotor.
Figures 6 and 7 present the radial distribution of the axial
and tangential velocity components measured at the rotor exit,
at X = 70 mm, for the NACA 0015 bladed rotor with and
without guide vanes. The trends of the radial variation of the
experimentally obtained values are generally in qualitative
agreement with those previously obtained at 1ST for low solidity
rotors (Gato and Falcao, 1988; 1990). From the velocity profiles shown in Fig. 7(b), for values of U* above 0.2, it can be
clearly seen that the tested rotor can produce positive torque
with partial stall.
Figure 8 shows the radial distribution of the absolute flow
angle calculated from the experimental values shown in Fig. 7
for the turbine with guide vanes. As predicted by the theory
(Gato and Falcao, 1990), absolute flow angles are not substantially affected by the flow rate variation in the range of flow
for which high values of rotor efficiency can be observed. Annular-wall boundary layer and tip clearance effects produce a high
variation of the flow angle near the inner and outer walls, which
affects the performance of the outlet guide vanes in an unpredicted manner.
5

Conclusions
In the present experiments, a remarkable difference was
found in the flow rate coefficient prior to stalling for two similar
rotors with slightly different blade profiles. This indicates that
a judicious choice of the blade profile, using an accurate modeling of the viscous flow around the turbine blades (90 deg stagger
angle cascades), is required to widen the range of flow rates
for which the turbine can perform with good efficiency.
Kinetic energy losses measured at the turbine exit are substantially smaller than viscous losses observed at the rotor.
However, the use of guide vanes to remove flow swirl at the
exit can result in a considerable improvement in efficiency.

160

u* = o.os
-B-

120

U* = 0.1

ii^
\
\

--U*=0.2
-ii-U*=0.25
\ -e-U* = 0.28

40

0.68

0.76

Acknowledgments
The authors wish to thank J. Ponciano for his assistance in
the preparation of the experiments. The work was financially
supported by CTAMFUTL, and contract JOUR-CT91-0133
from EU. Mobility grants were obtained under ERASMUS programme.
References
Falcao, A. F. de O., Sarmento, A. J. N. A., Gato, L. M. C , and Pontes, M. T.,
1991,' 'Preliminary Design of a Demonstration OWC Wave Plant for tiie Azores,''
Third International Symposium on Ocean Wave Energy Utilization, eds., T. Miyazaki and H. Hotta, JAMSTEC, Yokosuka, pp. 395-403.
Ferro, L. M. C , 1990, "Design of a Kaplan Bulb Turbine Rotor," Internal
Report (in Portuguese), Dept. Mechanical Engineering, 1ST, Lisbon, Portugal.
Gato, L, M. C , and Falcao, A. F. de O., 1984, "On the Theory of the Wells
Turbine," ASME Journal of Engineering for Gas Turbines and Power, Vol. 106,
pp. 628-633.
Gato, L, M. C , and Falcao, A. F, de O., 1988, "Aerodynamics of the Wells
Turbine," International Journal of Mechanical Sciences, Vol. 30, No. 6, pp. 3 8 3 395.
Gato, L. M. C , and FalcSo, A. F. de O., 1989, "Aerodynamics of the Wells
Turbine: Control by Swinging Rotor Blades," International Journal of Mechanical Sciences, Vol. 31, No. 6, pp. 425-434.
Gato, L. M. C and Falcao, A. F. de O., 1990, "Performance of Wells Turbine
with Double Row of Guide Vanes," JSME International Journal, Series II, Vol.
33, No. 2, pp. 262-271.
Gato, L. M. C , Efa, L. R. C , and Falcao, A. F. de 0., 1991a, "Performance
of the Wells Turbine with Variable Pitch Rotor Blades," ASME JOURNAL OP
ENERGY RESOURCES TECHNOLOGY, Vol. 113, pp. 141-146.

Gato, L. M. C , Falcao, A. F. de O., Pereira N. H. C , and Pereira,


R. M. R. J., 1991b, "Design of Wells Turbine for OWC Wave Power Plant,"
Proceedings, First International Offshore and Polar Engineering Conference,
eds., J. Chung, B. Natvig, K. Kaneko, and A. Ferrante, International Society of
Offshore and Polar Engineers (ISOPE), Vol. 1, pp. 380-384, Golden, CO.
Horlock, J. H., 1966, Axial Flow Turbines, Butterworths, London, U.K.
Kaneko, K., Setoguchi, T., Harakawa, H., and Inoue, M., 1991, "Biplane Axial
Turbine for Wave Power Generator," International Journal of Offshore and Polar
Engineering, Vol. 1, No. 2, pp. 122-128.
McCormick, M. E., and Surko, S. W., 1989, "An Experimental Study of the
Performance of the Counter-rotating Wave Energy Conversion Turbine," ASME
JOURNAL OF ENERGY RESOURCES TECHNOLOGY, Vol. I l l , pp. 167-173.

Pereira, N. H. C , and Pereira, R. M. R. J., 1991, "Calibration of a Flow


Meter," Internal Report (in Portuguese), Dept. Mechanical Engineering, 1ST,
Lisbon, Portugal.
Raghunathan, S., and Tan, C. P., 1983, "The Performance of Biplane Wells
Turbine," Journal of Energy, Vol. 7, No. 6, pp. 721-742.
Raghunathan, S., Tan, C. P., and Ombaka, O. O., 1985, "Performance of the
Wells Self-Rectifying Air Turbine," Aeronautical Journal, Vol. 89, pp. 3 6 9 379.
Raghunathan, S., and Ombaka, O. O., 1985, "Effect of Frequency of Air Flow
on the Performance of the Wells Turbine,'.' International Journal of Heat and
Fluid Flow, Vol. 6, No. 2, pp. 127-132.
Raghunathan, S., Setoguchi, T., and Kaneko, K., 1989, "The Effect of Inlet
Conditions on the Performance of Wells Turbine," ASME JOURNAL OF ENERGY
RESOURCES TECHNOLOGY, Vol. I l l , pp. 37-42.

80

'\

which, for the constant thickness NACA 0015 rotor considered


in this paper, was found to attain 17.1 percent.
The results presented here indicate that the use of stator
blades produce a non-negligible reduction of the value in the
ratio <E>, which means a narrower range of flow rates for which
the turbine performs efficiently.

0.84

0,92

r
Fig. 8 Radial distribution of average absolute flow angle measured at
the rotor exit for the NACA 0015 bladed rotor turbine with guide vanes,
for several flow rate coefficients

Richards, D and Weiskopf Jr., F. B., 1986, "Studies With, and Testing of the
McCormick Pneumatic Wave Energy Turbine with Some Comments on PWECS
Systems," Proceedings of the International Symposium Utilization of Ocean
Waves-Waves to Energy Conversion, eds., M. E. McCormick and Y. C. Kim,
ASCE, New York, NY, pp. 80-102.
Sarmento, A. J. N. A., Gato, L. M. C , and Falcao, A. F. de O., 1990, "TurbineControlled Wave Energy Absorption by Oscillating-water-column Devices,"
Ocean Engineering, Vol. 17, No. 5, pp. 481-497.
Setoguchi, T Kaneko, K., Maeda, H., Kim, T. W., and Inoue, M., 1993,
"Impulse Turbine with Self-Pitch-Controlled Guide Vanes for Wave Power Conversion: Performance of Mono-Vane Type," International Journal of Offshore
and Polar Engineering, Vol. 3, No. 1, pp. 73-78.
Warfield, V. P., 1994, "An Investigation into the Performance of a High
Solidity Wells Turbine for an Oscillating Water Column Wave Power Plant," M.
Tech. thesis, Department of Mechanical and Production Engineering, University
of Limerick, Ireland.
Wilson, D. G., 1984, The Design of High-Efficiency Turbomachinery and Gas
Turbines, The MIT Press, Cambridge, MA.

268 / Vol. 118, DECEMBER 1996

Downloaded From: http://energyresources.asmedigitalcollection.asme.org/ on 04/09/2015 Terms of Use: http://asme.org/terms

Transactions of the ASME

Das könnte Ihnen auch gefallen