Sunteți pe pagina 1din 10

A Novel High-Efficiency Impulse Turbine for Use in

Oscillating Water Column Devices


Shahab Natanzi
1
, Joao Amaral Teixeira
2
, George Laird
1
1
Dresser-Rand Company Limited
85, Papyrus Road, Peterborough, United Kingdom
SNatanzi@dresser-rand.com
GLaird@dresser-rand.com
2
Cranfield University, Department of Power and Propulsion
Cranfield, Bedfordshire, United Kingdom
j.a.amaral.teixeira@cranfield.ac.uk
Abstract This paper presents a novel impulse turbine
configuration for use in OWC systems which was developed as
part of a UK Department of Trade and Industry part-funded
research programme carried out by Dresser-Rand Company Ltd
and Cranfield University. The turbine demonstrates increased
levels of efficiency in comparison with existing impulse designs
across a wide range of incident flows. This is achieved by a
design which does not contain any moving parts besides the
rotor, meaning that good reliability and economic cost of
production should be achievable. The results presented are based
on a validated and calibrated numerical approach and describes
that a total-static efficiency of about 75% for a full scale turbine
has been achieved. The numerical results have been validated
against the data obtained at unique test facility located at
Cranfield University which is capable of simulating the
aerodynamic conditions that turbines are exposed to in OWC
power plants. Moreover, a unique numerical study of the effects
of inlet flow distortion on the performance of the impulse turbine
for use in OWC power plants is also presented.
Keywords Wave Energy, Impulse Turbine, HydroAir,
Efficiency, Output Power
I. INTRODUCTION
Oscillating Water Column (OWC) devices have been the
subject of research for over thirty years and a small number of
these devices have been operated, mostly in prototype or pre-
commercial forms.
An OWC device comprises a structure forming an air
chamber on the sea surface which encloses a volume of air.
The incident waves make the free surface of the water inside
the chamber rise and fall. The rising water level forces the air
through a turbine at the top of the structure. As the water level
falls, the air from outside the chamber is sucked back into the
chamber through the turbine. It should be noted that in most
OWC configurations the turbine rotates only in one direction
irrespective of the direction of the air flow.
OWC systems comprise of a chain of energy conversions,
wave to pneumatic (buoy), pneumatic to mechanical (air
turbine) and mechanical to electrical (generator). Of these the
first two are the most problematic when the overall efficiency
of the conversion is taken into account. Accepting that the
wave to pneumatic conversion efficiency can be reasonably
high for a judiciously sized and appropriately designed
collector buoy, knowing that the mechanical to electrical
conversion efficiency for electrical systems can be over 80%,
it leaves the turbine as the single most critical element in the
efficiency chain of an OWC system. Figure 1 shows a
schematic view of an OWC power plant. In practice, the
collector chamber can be either a land based structure or a
tethered buoy.
Figure 1 - Oscillating Water Column (OWC) wave energy converter
Over the last thirty years, there have been a large number
of research projects addressing turbine design for use in OWC
systems. However, mainly due to the low efficiency of the
OWC turbines, the reported overall wave-to-wire efficiencies
are often low, and hence such plants are still not economically
feasible. This is because, the design of a turbine for use in the
reciprocating flow conditions, experienced in these
applications, is complicated due to the alternating direction of
the flow. Therefore, a self-rectifying turbine that can provide a
unidirectional rotation, regardless of direction of flow, is a
solution for these systems. Additional difficulties arise from
Wave Propagation
Water
Air
Turbine
Generator
Airflow
OWC
the fact that the turbine has to operate over a wide range of
flow conditions associated with a variety of sea states; to say
nothing of the fact that waves are composed by trains of
largely irregular water fluctuations.
Over the last 30 years, many different turbine designs have
been examined some of which are listed in Table 1.
TABLE 1
Turbines for Wave Power Generation
Turbine Type Sub-type
Wells Turbine Without guide vanes
With guide vanes
Self pitch controlled blades
Leading edge slat
Biplane without guide vanes
Biplane with outer guide vanes
Biplane with outer and inner
guide vanes
Contra-rotating
Impulse turbines Impulse turbine with self pitch
controlled guide vanes
Impulse turbine with fixed guide
vanes
McCormick counter-rotating
turbine
Dennis Auld turbine
Radial turbines
Cross flow turbine
Savonius Turbine
In 2006 Setoguchi and Takao [3] reviewed the status of the
turbines listed in Table 1 on the basis of:
- The starting characteristic
- The running characteristic i.e. the efficiency versus flow-rate
characteristic
- Peak efficiency, and stall margin
The authors presented the turbine peak efficiency values
under steady flow conditions (Table 2).
TABLE 2
PEAK EFFICIENCY OF VARIOUS AIR TURBINES FOR WAVE ENERGY [3]
Turbine
Type
Wells
turbine
with
guide
vanes
Wells
turbine
with self-
pitch-
controlled
blades
Biplane
Wells
turbine
with
guide
vanes
Impulse
turbine
with self-
pitch-
controlled
guide
vanes
Impulse
turbine
with
fixed
guide
vanes
Peak
Turbine
Efficiency
0.492 0.496 0.534 0.564 0.390
The same authors also evaluated the conversion efficiency
of wave energy systems for various turbines under irregular
flow conditions. This was carried out because the authors
believed that the characteristics of the turbines under steady
flow conditions do not provide sufficient information about
the suitable turbine for wave power conversion. This lack of
information exists because the performance of the wave power
converter also depends on the OWCs energy absorption
efficiency, which is closely related to the pressure difference
across the turbine. Therefore, they presented the energy
conversion efficiency of the system which is obtained by
multiplying the efficiency of the air chamber and the turbine
efficiency. This is presented in Table 3.
TABLE 3
PEAK CONVERSION EFFICIENCY OF WAVE ENERGY SYSTEM UNDER
IRREGULAR FLOW CONDITIONS [3]
Turbine
Type
Wells
turbine
with
guide
vanes
Wells
turbine
with self-
pitch-
controlled
blades
Biplane
Wells
turbine
with
guide
vanes
Impulse
turbine
with self-
pitch-
controlled
guide
vanes
Impulse
turbine
with
fixed
guide
vanes
Peak
Conversion
Efficiency
0.30 0.44 0.34 0.47 0.37
They concluded that the impulse type turbines have the
potential to be superior to the Wells type turbines in overall
performance under irregular flow conditions; to say nothing of
the fact that in comparison, the starting characteristic of the
impulse turbine with self pitching guide vanes was superior;
that the lower rotational speed of an impulse turbine was
desirable to reduce noise and ease mechanical design; and that
the impulse turbine maintain its efficiency over a wider range
of flow rates.
Most of the OWC plants operated to date have utilised
Wells turbines, except for the Port Kembla OWC which is
using a variable pitch Denniss-Auld turbine. The outputs of
the Wells turbine OWCs have been around 400-500 kW. In
terms of overall wave-to-wire efficiency under real sea
conditions, their performance have been found to be relatively
low [4].
Figure 2 - Schematic of fixed guide vane impulse turbine showing [19]
The main problematic design element of an impulse
turbine utilised in bidirectional flows, is the requirement to
have guide vanes both upstream and downstream of the rotor
blade as shown in Figure 2. Since the function of the vanes is
to direct the inlet flow onto the rotor with the correct
orientation, and given that the outflow of impulse turbine rotor
is broadly axial, it is inevitable that the angle of incidence
onto the downstream guide vanes is rather poor. This results
in large areas of separated flow around the downstream guide
vanes and correspondingly large losses. This is demonstrated
graphically in Figure 3 which shows a velocity vector plot
around the impulse turbine blades from a CFD calculation.
Although changing the shape of the guide vanes are proved to
be effective, but similar separated flow has been observed for
various studies with different guide vane profiles [5].
Figure 3 - Velocity vector plot around impulse turbine blades
The observed separation increase the losses through the
turbine and therefore decreases the turbine efficiency
significantly. One solution to this problem is to have variable
geometry guide vanes that are able to provide the desired flow
turning when they are upstream of the rotor, and adopt a low
loss configuration when they are downstream. The experience
with this approach in the field, however, has been that there
are significant reliability and maintainability issues associated
with this choice [3].
II. VARYING RADIUS TURBINE (VRT)
The Dresser-Rand/Cranfield University HydroAir project,
which the UK DTi co-sponsored, commenced by an extensive
literature review undertaken by Herring [7], and numerical
studies of the baseline Wells and impulse turbines performed
by Banks [19]. After that a decision was taken to focus on
impulse turbines for further development. The rationale
behind this decision was that it seemed more feasible to
improve the peak efficiency of the impulse turbine and
maintain the wide operating range (at high efficiencies) than
to broaden the efficient operating range of the Wells turbine.
Although various methods exist to delay the onset of stall for
aerofoils used in aircraft, none of these seemed to offer the
promise of sufficiently wide operating range to make the
Wells turbine an ideal solution for use in real OWC flows.
Initially the key approach that seemed likely to give high
efficiencies over a wide range of incident flows was to
implement the variable pitch mechanism for the Wells turbine
blades [6]. However, this would require a complex mechanism
to actuate the blades as part of a rotating assembly and it was
felt that it would be significantly difficult to make such a
system reliable in a real marine environment.
At this stage, extensive Computational Fluid Dynamic
(CFD) analysis was performed on Wells turbine, mainly by
Banks [19]. Comparing the results with the available
experimental data in the literature, some confidence had been
gained in the ability of CFD to be used as a tool in the design
of impulse turbines. Therefore, it was decided to attempt to
produce a baseline impulse turbine geometry with equal or
better performance than those reported in the literature over
which improvements could then be sought.
In the course of developing the baseline impulse turbine, it
became clear that it was likely to prove difficult to achieve a
substantial increase in efficiency if the flow separation on the
downstream guide vane could not be removed or at least
substantially decreased. Since there was no obvious way to
remove the separation around the downstream guide vane
without recourse to variable geometry or boundary layer
blowing (neither of which is desirable in a marine
environment due to the associated reliability concerns), it was
realised that decreasing the flow velocity in this region would
substantially reduce the pressure drop, even if the separation
was still present. This can be demonstrated by adopting the
approach of assigning loss coefficients to the guide vanes.
According to the classical turbine theory, the stagnation
pressure loss coefficient equation in incompressible flow can
be calculated for the guide vanes using the equations below in
which P is the total pressure, V is the inlet flow velocity to the
guide vanes, and Y represents the stagnation pressure loss
coefficient.
2
02 01
5 . 0 V
P P
Y
GV

=
This equation describes that for a particular guide vane
profile, which has a constant pressure loss coefficient (
GV
Y
),
there is a correlation between the pressure drop across the
component and the inlet velocity (
2
V P A ). This shows that
the total pressure drop across the guide vanes is related to the
second power of inlet velocity. In other words, if the velocity
of the flow through the guide vane row is reduced, the result
would be a greater reduction in pressure drop which results in
having a more efficient guide vane row. An obvious way to
decrease the velocity is to increase the cross sectional area of
the flow path for a given volumetric flow rate. This could be
achieved with a moderate flaring of the annulus such that the
span of the guide vanes would be larger than that of the rotor
blades. Simply increasing the height of the annulus at the
guide vanes however, would lead to problems with the flow
angle at the inlet to the rotor, because the axial velocity would
increase, in the convergent section of the annulus,
disproportionately to the tangential velocity.
The solution to achieving both a velocity reduction while
keeping control of the variation of the velocity triangles
involved moving the guide vanes outwards towards a larger
radius by reference to their original position. This
arrangement was patented and designated the varying radius
turbine (VRT) and a schematic view of it is shown in Figure
4.
Figure 4 - Schematic of the varying radius turbine layout [19]
There were two approaches here that could be taken; firstly
keeping the height of the blades (guide vanes and rotors)
constant, or calculating the effect that changing height of the
blades can have on the flow angles [20]. In this section, to
simplify the explanation, the height of the blades is kept
constant, although the results presented in the next section are
for uneven blade heights.
By maintaining the height of the guide vanes, while
moving them outward from the original annulus to a greater
diameter, so that the guide vanes are placed at a greater radius
than the rotor, the axial and tangential components would both
be accelerated as the fluid proceeded towards the rotor,
because angular momentum must also be conserved. This
would maintain a good angle of incidence when acting as an
IGV. After exiting the rotor, the fluid would then be
decelerated both axially and tangentially meaning that
although the angle of incidence onto the downstream guide
vane would remain poor, the velocities and hence pressure
drop would be much smaller.
Due to the novelty of the arrangement, there was no guidance
in the literature as to what an appropriate offset and duct
geometry might be. Initially, it was calculated that to give a
predicted increase in efficiency to 60 % from the 40 % of the
baseline design, the required duct offset was 3.6 blade heights.
This analysis was based on the fact that in the absence of any
other effects (e.g. increased wetted area losses), the pressure
drop (and hence the efficiency) should obey the simple
relation [7]:
P
r
r
P A |
.
|

\
|
'
= ' A
2
where P A is the pressure drop with a constant mean radius of
r and P' A is the pressure drop that would be expected from a
VRT arrangement with the guide vanes located at a mean
radius of r' .
A large number of CFD simulations carried out indicated
that the concept was effective in increasing performance with
the initial case giving a predicted efficiency of 58.9 % at the
design point. This suggested that the simplistic analysis of
Herring lead to reasonably accurate predictions of the benefit
that might be expected for modest guide vane offsets.
III. EXPERIMENTAL PROCEDURE
As part of the Dresser-Rand and DTi sponsored HydroAir
project, a reciprocating flow test facility was designed and
installed at the Gas Turbine Laboratories at Cranfield
University for testing bidirectional turbines intended for use in
OWCs. A full description of the design, construction and
operation of this test rig can be found in [7] and [8]; however,
an outline is presented here.
The rig consists of a circular chamber containing a flap that
can be controlled by means of a computer to deliver an
oscillating airflow to a duct in which test turbines can be
located. A schematic of this arrangement is shown in Figure 5.
The fixed baffle means that by rotating the flap backwards and
forwards, the air on the turbine side of the baffle is alternately
compressed and rarified and an oscillating flow is set up
through the turbine located in the test section. Both the
frequency and the amplitude of the flap movements can be
accurately controlled to allow the simulation of airflows
caused by waves of different heights and periods. Although
the results that will be presented in this work are for constant
frequency and amplitude sinusoidal flows, the rig is also
capable of producing time varying irregular pressure
fluctuations that are more representative of the flow
conditions seen in a real OWC. A CAD model of the test rig is
shown in Figure 6.
Figure 5 - Schematic of the oscillating flow bidirectional turbine
Figure 6 - CAD model of the test rig
The rig and the VRT turbine were designed in tandem such
that the turbine could be tested in a range of flows and
maintain adequate Reynolds numbers for scaling
purposes(
5
10 2 > ). Due to the low velocities of the
flow
1
50

s ms , there were not expected to be any scaling issues
associated with Mach number effects and so these were not
considered [7]. In order to capture any possible frequency
effects, Herring concluded that it is necessary to attain
Strouhal number similarity in order to be able to scale the
results to full size. Strouhal number is defined as:
a
v
fD
St =
where f is the frequency of the flow oscillations, D is a
characteristic length, in this case taken to be the turbine
diameter, and
a
v
is the mean axial flow velocity. The turbine
was designed to be a one quarter scale of a turbine that would
be suitable for an OWC similar to LIMPET. Full-scale waves
in locations suitable for the installation of OWC devices
typically have a period of around 10s. This, in tandem with a
design axial flow velocity of 25
1
ms for the turbine, meant
that to maintain Strouhal similarity, the test rig would have to
be capable of producing oscillations of period T equal to
Scale Full
T
_
314 . 0
. The thesis of Herring shows that this is
possible with the rig that was built, but only for a relatively
small range of flows.
TABLE 4
FLOW CAPABILITIES OF THE OSCILLATING TEST RIG [7]
Parameter Value Units
Total swept volume 16.43 3
m
Max flow rate 5.88 1 3
s m
Max RMS flow rate 4.15 1 3
s m
Max frequency 0.14
Hz
Min period for full oscillation 7 s
Max frequency for small oscillations 1
Hz
The maximum flows that test rig was capable of producing
are summarised in Table 4. The flow rate could be calculated
by measuring the flap movements by means of a position
encoder located on the motor shaft, or by static and total
pressure sensors positioned on the guide vane leading edges.
Both static and total pressures could be measured at these
positions and thermocouple measurements were also made for
temperature correction of the pressures.
The turbine speed and mechanical torque were measured
using a torque transducer located on the shaft between the
turbine and electrical generator and the maximum turbine
speed was limited by the rating of the electrical generator
which was 1500 rpm. With measurements of flow rate,
pressure drop, turbine rotational speed and mechanical torque,
the efficiency of a test turbine could be determined both
instantaneously (subject to the data acquisition limits) and on
average.
IV. NUMERICAL PROCEDURE
The ANSYS TurboGrid package was used to generate the
computational meshes for the guide vanes and rotor blade. All
meshes were body fitted hexahedral meshes with O-grids
around the blades and H/J/L-type topology. The tip region of
the rotor blade was meshed with a non-matching H-grid. First
cell heights from the blade surfaces were set to give an
estimated y+ < 2 at maximum Reynolds number. In all cases,
the maximum mesh angles were > 25. Three different mesh
densities were prepared for these cases; coarse (<500,000
nodes), medium (<1,000,000 nodes) and fine (>1,000,000
nodes).
The flow in the stator domains were calculated in the
stationary frame while the rotor blade flow was solved in the
rotating frame. Stage type interfaces were used to join the
domains. These interfaces are an option within CFX package
which allows for both pitch and frame change with all flow
quantities being circumferentially averaged and appropriately
scaled. Such interfaces incur a one time mixing loss (due to
the averaging) but are fully conservative. Rotationally
periodic boundaries were set along the sides of all domains
and all hub, shroud, and blade surfaces were set as non-slip
walls, with the shroud casing of the rotor domain being set as
a counter rotating wall. A generalised grid interface (GGI)
was used to join the two sides of the non-matching H-grid
used in the rotor tip gap. A total pressure was specified at the
inlet to the upstream guide vane and an average static pressure
(atmospheric) imposed at the outlet of the downstream guide
vane. The values of mass flow were therefore calculated based
on the pressure variation across the passage.
Figure 7 shows a typical comparison between the
experimental and the numerical data, obtained for one of the
earlier versions of the VRT, and describes that a sufficient
compatibility can be expected between numerical and
experimental results [19].
Figure 7 - Comparison of experimental and numerical results [19]
V. RESULTS AND DISCUSSION
After verifying the effectiveness of the VRT design,
several optimisation processes were carried out on this
concept which led to introduction of other generations of VRT
over the last three years. The results presented here
correspond to a full scale design denoted as VRT-2 and were
obtained through a calibrated and validated CFD approach
established at Dresser-Rand Company. It ought to be
mentioned that a prototype scale of this turbine was
manufactured at Dresser-Rand's Peterborough facility and was
installed on an off-shore OWC platform in Australia during
2010. A schematic view of the full scale VRT-2 is shown in
Figure 8 while a manufactured prototype is shown in Figure 9.
Figure 8 - Full scale VRT-2 configuration
Figure 9 - Manufactured VRT-2 prototype
Although the VRT-2 make use of a new guide vane profile,
new 3D rotor profile, as well a new duct concept, the whole
system is still based on the same approach outlined in the
previous section. Figure 10 shows the predicted efficiency of
the VRT-2 with a 2 meter diameter rotor as obtained from
numerical analysis. The present status of the design represents
the outcome of a series of optimisations studies on various
design parameters such as hub-tip ratio, guide vanes turning
angles, guide vane twist, tip clearance effects, and axial and
radial offset of the guide vanes in comparison to the rotor
position.
As can be seen from Figure 10, the present VRT-2 design
can achieve a peak efficiency of over 75 percent. This can be
considered as a step change in efficiency of the impulse
turbine with fixed guide vanes for use in OWC power plants.
0
0.25
0.5
0.75
1
0 0.5 1 1.5 2 2.5 3 3.5
Flow Coefficient
E
f
f
i
c
i
e
n
c
y
Figure 10 - Full scale VRT-2 Characteristic (numerical results)
Figure 11 - Velocity vectors on downstream guide vane (a) Flow-Coefficient
= 0.97 (b) Flow-Coefficient = 1.57
The new design approach was also beneficial to reduce
(and in some cases to completely remove) the separation on
the downstream guide vanes. This is thought to be the main
factor which allows the VRT-2 to achieve these peak
efficiency values. Figure 11 shows the velocity vectors on the
downstream guide vanes for two flow coefficients. Figure 11
(b) shows the velocity vectors on a downstream guide vane of
the VRT-2 operating at a flow coefficient of 1.57. As can be
seen, the majority of the separation on the downstream guide
vane is recovered compared to what was observed in Figure
11 (a) which is the same turbine operating at a flow
coefficient of 0.97. It should be highlighted that both cases
(Figure 11 a and b) achieved similar efficiency values. This
can be explained by that the operating under higher flow
coefficients means the flow velocity is increased inside the
turbine and therefore greater losses are associated with the
upstream turbine elements when compared to the turbine
operating under a lower flow coefficient. This approach can
be used to reduce the noise levels from the turbine since
boundary layer separation can be attributed to the generated
noise of the turbo-machines.
VI. ENVIRONMENTAL EFFECTS - INLET FLOWCONDITIONS
The performance of turbo-machines is dependent on
various environmental factors, amongst which inlet flow
conditions play a significant role. The importance of studying
the inlet flow conditions is magnified in the OWC turbine
investigation process due to the nature of the environment in
which OWC plants operate. There are a number of published
papers which have investigated the effects of inlet flow
distortion on various types of turbo-machines such as gas
turbines [9], [10], [11] and compressors [12], [13]. However,
due to the flow conditions in which these turbo-machines
usually operate (i.e. high flow velocities), these results are not
directly relevant to the OWC turbine studies.
In 1987, Raghunathan et al. [14] experimentally
investigated the effects of inlet flow distortion and high levels
of turbulence on the performance of a Wells turbine with
guide vanes, using an average Reynolds number of around
2.6
5
10 . The authors concluded that a large change in the
turbulence level (from 0.4% to 6.7%) increases the turbine
total-to-total efficiency by 2%. However, the stall point was
observed to be postponed from a flow coefficient of 0.24 to
0.3. They also concluded that inlet flow distortion leads to a
very small performance loss in the Wells turbine.
There is no literature available on the effect of inlet flow
distortion on the performance of the impulse turbine for use in
OWC power plants. Therefore, this section focuses on
evaluating the performance of the designed impulse turbine
(VRT-2), in presence of inlet flow distortion, in order to fill
the present gap in this knowledge. An investigation of the
predicted turbine performance in the operational environment
has been performed. The inlet flow distortion in the presented
model was caused by introducing a moderate wind flow (10
m/s wind velocity) to the system.
A numerical model was set up to include a complete OWC
chamber, a full scale VRT-2, and a large domain set to
represent the atmosphere. The atmospheric domain surrounds
the whole system. Figure 12 displays a schematic view of the
numerical model.
Figure 12 - A whole OWC power plant model
(a)
(b)
VRT-2
Atmospheric Domain
OWC
Chamber
Rotor sub-
domain
Oscillating
sea surface
Turbine
inner duct
Turbine
outer duct
Since the inclusion of all the turbine blading in the model
involved some very large grid counts, ~100million nodes, and
correspondingly large computational resources and running
times, a decision was taken to dispense with the detailed
representation of the blades. Instead it was the effects of the
blades on the flow that were modeled. This involved the
numerical representation, through source terms, of the turning
exerted by the blades together with the momentum losses. The
appropriate quantities were obtained from previous VRT-2
stage simulations. Figure 13 shows the added turning of the
flow passing through the guide vane sub-domain. This guide
vane turning shown in Figure 13 was achieved by imposing
directional loss on the desired flow direction using an extra
polar coordinate system. The pressure drop imposed on the
rotor domain was related to the mass flow rate by using a
quadratic equation.
Figure 13 - Imposed numerical turning on guide vane sub-domain and and
pressure drop on the rotor sub-domain
The simulations were performed using 32 processors out of
which 16 processors had 16GB memory and the rest had 8GB
memory. The solutions time-step was set to be 0.01 second
and the simulation was run for 1500, 2000, and 3000 time-
steps each with a maximum of 3 and 5 coefficient loops and a
residual target of 0.0001. It was shown that the results were
closely comparable for all the cases and every cycle. Hence,
for the purposes of clarity, the results presented below are
only for one full cycle. Figure 14 illustrates the velocity
contours on a X-Z plane. This is a meridional plane aligned
lengthways. Four cases are shown corresponding to maximum
exhalation and inhalation flows, with and without wind.
Figure 14 - Velocity contours during maximum exhalation and inhalation
Figure 15 shows the P and mass flow rate across the
turbine for one full cycle and for both cases with and without
wind. The P shown in this figure is the difference in pressure
measured between the inside of the chamber and the outside
of the turbine outlet which is atmospheric pressure. Therefore,
a positive P corresponds to the exhalation period when the
pressure inside the chamber is higher than atmospheric
pressure. A negative P is indicative of the inhalation cycle,
when the pressure inside the chamber is lower than the
atmospheric pressure and therefore the flow is being sucked
into the device. Although as expected the maximum and
minimum flow rates correspond to maximum and minimum
P respectively there is interestingly a small lag between the
two peaks.
Moreover, comparing Figures 15 (a) and (b), the results
show that in the presence of wind, the mass flow rate
increases during inhalation and decreases during exhalation
when compared to the case without wind. This can be
explained by the fact that during the inhalation, wind will
freely flow into the turbine which results in having around 7%
more flow during inhalation; while during exhalation the wind
blocks the turbines outlet resulting in a reduction of some
12% less mass flow. This is expected to correspond to a slight
increase in efficiency during inhalation and about a 3%
decrease in efficiency during exhalation. The results also
showed that the inlet flow distortion causes a certain amount
of non-uniformity of the flow entering the rotor blades during
inhalation period. This can on its own cause further disruption
to the performance of the turbine.
-6000
-4000
-2000
0
2000
4000
6000
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14

P

(
P
a
)
-300
-200
-100
0
100
200
300
M
a
s
s
f
l
o
w
r
a
t
e
(
k
g
/
s
)
-6000
-4000
-2000
0
2000
4000
6000
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Time (s)

P

(
P
a
)
-300
-200
-100
0
100
200
300
M
a
s
s
f
l
o
w
r
a
t
e
(
k
g
/
s
)
P (Pa) Mass flow rate (kg/s)
Figure 15 - P and mass flow rate across the turbine in a full cycle (a) with
wind (b) without wind
There are various other data that can be extracted from this
kind of simulations and hysteresis data can be one of them. As
regards to wave energy devices, it is usually presumed that the
10m/s wind
10m/s wind
No wind
No wind
With 10 m/s wind
Without wind
hysteresis effect is minor due to a very low air-flow
frequency. The hysteresis effect may have various causes such
as asymmetry in the boundary layer on the blade surfaces.
There are a number of studies on this subject mainly
performed by Raghunathan and Setoguchi et al [15], [16],
[17], [18]. However, all of them have studied the hysteresis
effects in a small scale model and mostly on the turbine
solely, not as a whole system as performed here.
-6000
-5000
-4000
-3000
-2000
-1000
0
1000
2000
3000
4000
5000
6000
-250 -225 -200 -175 -150 -125 -100 -75 -50 -25 0 25 50 75 100 125 150 175 200 225 250
Mass flow rate (kg/s)

P

(
P
a
)
Without wind With wind
Figure 16 - Flow rate against pressure drop across the turbine
Figure 16 shows the hysteresis loop on the OWC system in
the two cases, with and without wind effects. In other words,
it shows the differential pressure across the turbine against the
mass flow rate passing through the turbine. This indicates that
the relationship between the mass flow rate and the pressure
differential is not unique, but depends on whether the air is
forced out of the turbine (exhalation) or sucked in (inhalation).
It also explains that the magnitudes of hysteresis for both
cases (with and without wind) are very similar.
VII. CONCLUSIONS
The results presented in this paper are a selection from the
data acquired over a 5 year research program. From these a
number of conclusions based on this research program can be
drawn:
A new turbine design (VRT-2) was generated which
introduces a considerable step up in efficiency of the
impulse turbine with fixed guide vanes for use in
oscillating water column power plants. Various turbine
design elements were investigated and optimised. This
includes hub-tip ratio, guide vanes twist, tip clearance,
blade profiles, and axial and radial offset of the duct. As
a result, a peak efficiency of over 75% was observed for
this design.
The negative effect of the flow separation on the outlet
guide vane was minimised and in some cases was
completely removed, using the VRT approach as well as
optimised elements such as new guide vane profiles and
duct shape.
A novel transient OWC power plant model was created.
Based on this model, it was observed that the turbine
mass flow and efficiency is likely to be reduced when it
is exposed to windy conditions. Also, the transient
simulations helped to increase the understanding of the
hysteresis effects on the OWC systems.
ACKNOWLEDGEMENT
The authors would like to acknowledge the financial
assistance provided by Dresser-Rand Company Ltd and the
UK Department of Trade and Industry (now BERR). The
works of Professor Christopher Freeman on the design of the
turbine, Dr. Kevin Banks in performing various optimisation
analyses, and Dr. Steven Herring in carrying out the physical
experiments reported in this paper are also gratefully
acknowledged.
REFERENCES
[1] R. Shaw. Wave Energy a Design Challenge. Ellis Horwood Ltd, 1982.
[2] A. Sarmento, F. Neumann, and V. La Regina. The need and available
means for international cooperation in marine renewable technology. In
World Renewable Energy Congress, 2005.
[3] T. Setoguchi, and M. Takao,. Current status of Self rectifying air turbines
for wave energy conversion. Energy Conversion and Management, vol. 47,
no. 15-16, pp. 2382-2396.
[4] I. Webb, C. Seaman, and G. Jackson. Oscillating water column wave
energy converter evaluation report. Technical report, Carbon Trust, 2005.
[5] T. Setoguchi, S. Santhakumar, H. Maeda, M. Takao, and K. Kaneko. A
review of impulse turbines for wave energy conversion. Renewable Energy,
23:261292, 2001.
[6] S. Herring. OWC Performance Matrix and Design Review (Unpublished
report). 2005.
[7] S. Herring. Design and evaluation of turbines for use in OWC power
plants. PhD thesis, School of Engineering, Cranfield University, 2007.
[8] S. Herring and G. Laird. A new test facility for evaluating turbines for use
in OWC power plants. In Proceedings of the 7th European Wave and Tidal
Energy Conference, Porto, Portugal, 2007.
[9] M. Davis, and A. Hale, (2007), "A parametric study on the effects of inlet
swirl on compression system performance and operability using numerical
simulations", Vol. 1, pp. 1.
[10] M. Davis, A. Hale, and D. Beale, (2002), "An argument for enhancement
of the current inlet distortion ground test practice for aircraft gas turbine
engines", Journal of Turbomachinery, vol. 124, no. 2, pp. 235-241.
[11] Anon (1987), "Engine response to distorted inflow conditions (Paper
presented at the Propulsion and Energetics 68TH Annual Specialists
Meeting)".
[12] V. G. Filipenco, S. Deniz, J. M. Johnston, E. M. Greitzer, and N. A.
Cumpsty, (2000), "Effects of inlet flow field conditions on the performance of
centrifugal compressor diffusers: Part 1-discrete-passage diffuser", Journal of
Turbomachinery, vol. 122, no. 1, pp. 1-10.
[13] P. Bry, P. Laval, and G. Billet, (1985), "Distorted Flow Field in
Compressor Inlet Channels.", Journal of Engineering for Gas Turbines and
Power, vol. 107, no. 3, pp. 782-791.
[14] S. Raghunathan, T. Setoguchi, and K. Kaneko, (1987), "Wells air turbine
subjected to inlet flow distortion and high levels of turbulence.", International
Journal of Heat and Fluid Flow, vol. 8, no. 2, pp. 165-167.
Exhalation
Inhalation
[15] T. Setoguchi, Y. Kinoue, T. H. Kim, K. Kaneko, and M. Inoue, (2003),
"Hysteretic characteristics of Wells turbine for wave power conversion",
Renewable Energy, vol. 28, no. 13, pp. 2113-2127.
[16] S. Raghunathan, (1995), "The wells air turbine for wave energy
conversion", Progress in Aerospace Sciences, vol. 31, no. 4, pp. 335-386.
[17] M. Mamun, Y. Kinoue, T. Setoguchi, T. H. Kim, K. Kaneko, and M.
Inoue, (2004), "Hysteretic flow characteristics of biplane Wells turbine",
Ocean Engineering, vol. 31, no. 11-12, pp. 1423-1435.
[18] Y. Kinoue, T. H. Kim, T. Setoguchi, M. Mohammad, K. Kaneko, and M.
Inoue, (2004), "Hysteretic characteristics of monoplane and biplane Wells
turbine for wave power conversion", Energy Conversion and Management,
vol. 45, no. 9-10, pp. 1617-1629.
[19] K. Banks. Optimisation of bidirectional impulse turbines for wave power
generation. PhD thesis, School of Engineering, Cranfield University, 2009.
[20] S. Natanzi. Design and investigation of turbines for use in oscillating
water column power plants. PhD thesis, School of Engineering, Cranfield
University, 2010.

S-ar putea să vă placă și