Sunteți pe pagina 1din 11

Applied Ocean Research 33 (2011) 321–331

Contents lists available at ScienceDirect

Applied Ocean Research


journal homepage: www.elsevier.com/locate/apor

Numerical investigation on the performance of Wells turbine with non-uniform


tip clearance for wave energy conversion
Zahari Taha a , Sugiyono b,c,∗ , T.M.Y.S. Tuan Ya b , Tatsuo Sawada d
a
Department of Manufacturing Engineering, University of Malaysia Pahang, Lebuhraya Tun Razak, 26300 Gambang Kuantan, Pahang, Malaysia
b
Centre for Product Design and Manufacturing, University of Malaya, Lembah Pantai, 50603 Kuala Lumpur, Malaysia
c
Department of Mechanical and Industrial Engineering, Gadjah Mada University, Jl. Grafika No. 2, Yogyakarta 55281, Indonesia
d
Department of Mechanical Engineering, Keio University, 3-14-1 Hiyoshi, Kohoku-ku, Yokohama 223-8522, Japan

a r t i c l e i n f o a b s t r a c t

Article history: The performance of a Wells turbine with various non-uniform tip clearances was investigated using com-
Received 27 December 2010 putational fluid dynamics (CFD). The investigation was performed on numerical models of a NACA0020
Received in revised form 24 May 2011 blade profile under steady flow conditions. The performance of turbines with uniform and non-uniform
Accepted 2 July 2011
tip clearances was compared. The results were also compared with experimental results in literature. It
Available online 30 July 2011
was shown that the performance of turbine with non-uniform tip clearance is similar with that of turbine
with uniform one in terms of torque coefficient, input power coefficient, and efficiency. However, the
Keywords:
turbine with non-uniform tip clearance seems to have a preferable overall performance. An investigation
Wells turbine
NACA0020 blade profile
on the flow-field around the turbine blade was performed in order to explain the phenomena.
Tip clearance © 2011 Elsevier Ltd. All rights reserved.
CFD

1. Introduction converts the wave energy into pneumatic energy in the form of a
bi-directional airflow. A Wells turbine which constitutes one of self-
In nature, ocean has provided many potential resources of rectifying turbines is commonly used in order to extract mechanical
renewable energy which can be exploited to reduce the world’s energy from the bi-directional airflow, which is then converted into
dependence upon conventional fuels such as coal, oil, and natural electricity by means of the generator.
gas, or exhaustible sources of energy. One of the potential resources As described above, there are three main stages of energy con-
is wave energy. In its development, this energy has received much version process in the OWC system. Nevertheless, the performance
attention due to its potency which is relatively abundant, sustain- of turbine plays the most important role, which will give a crucial
able, and pollutant free. By converting into more usable forms of impact on the overall efficiency. Hence, various parameters which
energy, the ocean wave energy can give a significant contribution induce the performance of turbine have to be considered properly.
to cover the energy requirements, particularly for coastal nations One of the parameters is tip clearance. As revealed in Raghunathan
with island communities and correspondingly high energy costs [5], the Wells turbine is very sensitive to tip clearance when com-
[1]. pared to a conventional turbine.
Among the wide variety of possible technologies for the purpose Fig. 2 shows the schematic view of tip clearance of a Wells tur-
of wave energy conversion, the oscillating water column (OWC) bine, which in general has a uniform shape. This shape has the size
system is relatively mature and promising. The system is said to be of the gap between the tip of turbine blade and the turbine casing
one of the most successful devices in harnessing wave energy [2,3]. which is constant from the leading to trailing edges of the turbine
The schematic view of an OWC system is depicted in Fig. 1, which blade. Takao et al. [6] has modified the shape of the tip clearance to
substantially consists of a capture pneumatic chamber that opens be non-uniform and has investigated it experimentally. He found
at the bottom front to the incident wave, an air turbine, and an that the non-uniform tip clearance is preferable to the uniform one.
electrical generator [4]. In its energy conversion chain, the system However, it is necessary to clarify this further.
Along with the developments in computer hardware and soft-
ware, computational fluid dynamics (CFD) has become an efficient
∗ Corresponding author at: Centre for Product Design and Manufacturing, Univer-
means of investigating the performance of a Wells turbine. This is
because of its modeling capabilities on a wide range of fluid flow
sity of Malaya, Lembah Pantai, 50603 Kuala Lumpur, Malaysia. Tel.: +60 379675200;
fax: +60 379675330. problems and highly accurate predictions. It also provides detailed
E-mail address: sugiyono ugm@yahoo.com ( Sugiyono). descriptions of flow which is impossible to be obtained through

0141-1187/$ – see front matter © 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.apor.2011.07.002
322 Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331

Fig. 1. Oscillating water column system.

experimental work. It is also helpful in interpreting experimental This paper describes the use of CFD to investigate the perfor-
results and significantly reducing the amount of experimentation. mance of a Wells turbine with various non-uniform tip clearances.
There are several reports of investigations on Wells turbines The investigation was performed on numerical models of a
using CFD. Watterson and Raghunathan [7] studied the effect of NACA0020 blade profile under steady flow conditions. The com-
solidity on the performance of a Wells turbine with NACA0015 putational results of this study were compared with the values
blade profile. The effect of blade geometry with several hub-to- obtained experimentally by Takao et al. [6]. Moreover, they were
tip ratios and aspect ratios on the performance of a Wells turbine also compared with the computational results of the turbine with
with NACA0020 blade profile was studied by Kim et al. [8]. Kim uniform tip clearance from the previous study of Zahari Taha et al.
et al. [9] investigated the effect of blade sweep on the performance [15]. Furthermore, a flow-field investigation around the turbine
of a Wells turbine, and made a comparison of the performance blade was performed in order to explain the turbine performance.
between NACA0020 and CA9 blade profiles. Hysteretic phenomena
of a Wells turbine has been investigated under unsteady flow condi- 2. Numerical method
tions [10–12]. Dhanasekaran and Govardhan [13] investigated the
performance and aerodynamics of a Wells turbine with NACA0021 The present study was carried out using a CFD code called
constant chord blade, while performance improvement using a FLUENT. The method solves the three-dimensional, steady,
variable chord blade was studied by Govardhan and Chauhan [14]. incompressible, Reynolds-averaged Navier–Stokes equations. The
Takao et al. [3] used CFD to clarify the performance improvement Reynolds stress in the mentioned equations is related to the mean
of a Wells turbine due to the effect of end plates. Zahari Taha et al. velocity gradients by employing the Boussinesq approach [16]
[15] studied the effect of uniform tip clearance on the performance which has an advantage associated with the relatively low cost
of a Wells turbine with NACA0020 blade profile, and compared the in computation of the turbulent viscosity. Meanwhile, the turbu-
computational results with the experimental ones of Takao et al. lence model adopts the Realizable k–ε model [17] which is likely to
[6]. provide superior performance for flows involving rotation, bound-
ary layer under strong adverse pressure gradients, separation, and
recirculation. The non-equilibrium wall functions [18] are applied
for the near-wall region modeling to produce accurate predictions
due to the wall-bounded turbulent flows. Furthermore, the govern-
ing equations are solved in the absolute frame and discretized by
the finite volume technique while imposing several discretization
schemes. The standard scheme is used for pressure by means of
momentum equation coefficients [19] as it is acceptable for most
cases. Because of steady state calculations conditions, the SIM-
PLE algorithm [20] is used for pressure–velocity coupling. In order
to improve the accuracy of the solution, the second-order accu-
rate upwind scheme is adopted for momentum, turbulence kinetic
energy (k), and turbulence dissipation rate (ε) by applying a multi-
dimensional linear reconstruction approach [21].
The specification of the turbine model for this study is summa-
rized in Table 1. Note that the adopted turbine rotor is the most
promising one in previous studies [22,23]. The configuration of
the non-uniform tip clearance of the turbine model is described in
later section. By considering the symmetry of the turbine geometry,
and in order to make the computation practical, the computational
domain of the turbine model is restricted to one blade-to-blade
Fig. 2. Schematic view of tip clearance. passage with periodic boundaries. The computational domain is
Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331 323

Table 1 the sub-mapping grids are only used in this region. The total num-
Specification of the turbine model.
ber of grids is approximately 295,500.
Blade profile NACA0020 The computations on each case of non-uniform tip clearance
Number of planes 1 are carried out under steady flow conditions for various values of
Number of blades 6 the flow coefficient (), which is defined as the ratio of the axial
Blade chord length 90 mm velocity of inlet airflow (v) to the circumferential velocity of the
Solidity at mean radius 0.67 blade at mean radius (U),
Casing diameter 300 mm
Hub diameter 210 mm v
= (1)
Mean radius 127.5 mm U
Blade sweep ratio 0.35
Width of flow passage 45 mm Meanwhile, this study was conducted regardless of the sig-
nificant Reynolds number effects. The various values of the flow
coefficient were achieved by varying the rotational speed of the
rotor at a constant value of the axial velocity, which are similar
with those performed by Kim et al. [8,9], Takao et al. [3] and Zahari
et al. [15]. Reynolds number is based on the blade chord length
and relative velocity at mean radius, which is about 0.68 × 105 to
4.41 × 105 .
Further, for the solution initialization, the computations are per-
formed under the absolute reference frame and the axial velocity
is set to zero. During the solution process, the default under-
relaxation factors are selected to control the update of computed
variables, which are found to be near optimal for a large number
of cases. The convergence of solution is monitored by checking the
residuals of the numerically solved governing equations. Moreover,
in order to judge the convergence, the behaviour of other quanti-
ties, such as the total pressure at the inflow and outflow boundaries,
and torque coefficient generated by the rotor, are also monitored.
Fig. 3. Perspective view of computational domain (l: blade chord length).
Here, the default convergence criterion of each residual is reduced
to reasonable values in order to allow the monitored quantities to
stagnate at consistent values. Finally, the convergence of the solu-
also limited to four and eight blade chord lengths upstream and
tion is checked for mass balance. In this study, the computations
downstream from the blade row, respectively. Fig. 3 shows the per-
on each case of non-uniform tip clearance result in the net mass
spective view of the computational domain for the turbine model.
imbalance of less than 0.006%.
For the boundary conditions, the inlet velocity and static
pressure are imposed on the inflow and outflow boundaries,
respectively, which are normal to the axis of turbine. Rotational 3. Results and discussion
periodic boundaries are applied on the pairs of surfaces which con-
stitute the circumferential sides of the domain. A moving reference The performance of turbine is expressed in the form of dimen-
frame is employed on the fluid zone which has the rotational speed sionless parameters which are torque coefficient (CT ), input power
equivalent to that of the rotor. No-slip conditions are used for the coefficient (CA ), and efficiency (). These parameters are defined as
blade and the hub surfaces. follows [6]:
Fig. 4 depicts the mesh employed on the computational domain. T
CT = (2)
The mesh consists of structured, hexahedral grids. This has been {[v2 + U 2 ]hlzr/2}
achieved by performing decomposition of the model geometry into
meshable pieces for mapping and sub-mapping grid algorithms. As pQ
CA = (3)
shown in the figure, O-type mapping grids are applied surround- {[v2 + U 2 ]hlz v/2}
ing the turbine blade. Further, there are five grids in the span-wise Tω CT
direction which are imposed in the tip clearance region, and then = = (4)
pQ CA 
where T, Q, p, ω, , l, h, r, and z denote the turbine output torque,
flow rate, total pressure drop across the turbine, turbine angular
velocity, density of air, chord length, width of flow passage, mean
radius, and number of rotor blades, respectively. In addition, as
stated above, the computations are carried out for various values of
the flow coefficient. However, this study is confined to flow coef-
ficients up to when stall occurs. As presented in Kim et al. [9], the
stall point is decided by a decrease in the turbine torque coefficient.

3.1. Preliminary investigation

The effect of non-uniform tip clearance configuration on the


performance of Wells turbine is not known exactly yet. Therefore,
computations are firstly performed on the turbine model with two
types of the non-uniform tip clearance configuration. Each type
has the smallest and largest sizes of the gap which are 0.5 mm and
Fig. 4. Computational grid. 1 mm, respectively. The two types of configuration are as follows:
324 Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331

Fig. 5. Computational results of turbine efficiency for the cases of TC 0.5G1 mm and
TC 0.5UMG1 mm. Fig. 6. Computational results of turbine efficiency for the cases of TC 0.5G0.75 mm
and TC 0.5UMG1 mm.

Fig. 7. Comparison of turbine performance between computational and experimental results for TC 0.5G1 mm or TC** = 0.0083.
Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331 325

in the first type (TC 0.5G1 mm), the size of the gap increases
gradually from 0.5 mm at the leading edge to 1 mm at the
trailing edge; and in the second type (TC 0.5UMG1 mm), the
size of the gap is kept constant at 0.5 mm from the leading
edge to the mid chord and then increases gradually to 1 mm
at the trailing edge. The aim of these preliminary computa-
tions is solely to interpolate a near optimal configuration of the
non-uniform tip clearance regardless of the average size of the
gaps of the two. The computational results of the performance
of both turbines are then presented in terms of the turbine
efficiency.
Fig. 5 illustrates the computational results of the turbine effi-
ciency for the cases of TC 0.5G1 mm and TC 0.5UMG1 mm. As shown
in the figure, the efficiency of TC 0.5UMG1 mm is slightly higher
than that of TC 0.5G1 mm at low values of the flow coefficient, but
lower at values of the flow coefficient greater than 0.29. How-
ever, the peak efficiency of TC 0.5UMG1 mm is higher than that of
TC 0.5G1 mm, and occurs at a lower value of the flow coefficient.
TC 0.5UMG1 mm stalls earlier as well. These tendencies might be
reasonable when associated with those pointed out by Zahari Taha Fig. 8. Computational results of turbine efficiency for the cases of TC 0.5G1.25 mm
et al. [15] in which the peak efficiency of turbine is higher and occurs and TC 0.75G1 mm.
at a lower value of the flow coefficient, and the turbine stalls earlier,
as the uniform tip clearance becomes smaller. In this case, it should

Fig. 9. Computational results of turbine performance with non-uniform tip clearance for various values of the tip clearance to the chord length ratio.
326 Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331

be noted that the average size of the gap for TC 0.5UMG1 mm (i.e.
0.63 mm) is smaller than that for TC 0.5G1 mm (i.e. 0.75 mm).
Furthermore, in terms of operational range and overall effi-
ciency, it is shown in Fig. 5 that the performance of TC 0.5G1 mm
is better than that of TC 0.5UMG1 mm. Immediately, it can be con-
cluded that a configuration of non-uniform tip clearance which is
similar with that of TC 0.5G1 mm is preferable and more suitable
to be considered further in the study. However, this needs to be
clarified because of the difference of the average gaps between
TC 0.5G1 mm and TC 0.5UMG1 mm. Therefore, TC 0.5UMG1 mm is
then compared with TC 0.5G0.75 mm which has a configuration
similar to that of TC 0.5G1 mm (i.e. the size of the gap increases
gradually from 0.5 mm at the leading edge to 0.75 mm at the trail-
ing edge), but has the same average gap with TC 0.5UMG1 mm (i.e.
0.63 mm). A comparison of the turbine efficiency between both tur-
bines is depicted in Fig. 6. As shown in the figure, the performance
of TC 0.5G0.75 mm is better than that of TC 0.5UMG1 mm. Thus, the
figure confirms the aforementioned conclusion.
In order to show that the numerical method used for the
study is reliable, the computational results are compared with the
experimental values of Takao et al. [6], particularly for the case
of TC 0.5G1 mm, or in terms of the average tip clearance to the
chord length ratio, TC** = 0.0083. Fig. 7 depicts the mentioned com-
parison of the turbine performance between the computational
and experimental results. In general, it is shown in the figure
that within the range of the flow coefficients given, good agree-
ment exists between the computational and experimental results.
Meanwhile, it can be observed that the torque coefficients of both
computational and experimental results is almost the same, except
for the flow coefficient values near the stall point in which the
torque coefficient is predicted slightly lower. On the other hand,
the input power coefficients are predicted slightly lower within
almost the whole range of the flow coefficients given. As a con-
sequence, the predicted turbine efficiency is slightly higher than
the experimental one, except for the flow coefficient values near Fig. 10. Variation of mean turbine efficiency and stall incidence angle with ratio of
the stall point in which the turbine efficiency of both computa- the tip clearance to the blade span: (a) computational results of the Wells turbine
tional and experimental results is almost the same. Further, the with non-uniform tip clearance; (b) experimental results quoted in Raghunathan
value of peak efficiency is predicted higher by approximately 1%, [5].

and occurs at the value of the flow coefficient which is very close
to the experimental one as well as that where the stall point is
predicted. input power coefficient, and efficiency have the same tendencies
with those explained in Zahari Taha et al. [15] with regard to the
3.2. Performance of Wells turbine with non-uniform tip clearance turbine performance for various cases of uniform tip clearance.
The most conspicuous feature is that a turbine with a larger tip
As a follow up to the preliminary investigation, a study is clearance would have a wider operational range of flow without
performed on the turbine model with various configurations of stalling. The peak efficiency of the turbine decreases and shifts
non-uniform tip clearance which are similar with the configuration towards a higher value of the flow coefficient as the tip clearance
of TC 0.5G1 mm as shown in Table 2. However, there are two turbine increases.
models with the same average size gap, which are TC 0.5G1.25 mm Furthermore, based on Fig. 9(c), the mean turbine efficiency () ¯
and TC 0.75G1 mm. Therefore, it is necessary to compare the per- in the wide range of flow coefficients for each case of TC** can
formances of both turbines in order to determine which turbine be estimated, which is the average value of an integration value
model would be better and more proper to be considered further. of turbine efficiency with the variation in the flow coefficient [9].
Fig. 8 presents the computational results of the turbine per- Fig. 10(a) illustrates the variation of the mean turbine efficiency
formance for TC 0.5G1.25 mm and TC 0.75G1 mm in terms of the and stall incidence angle (˛s ) with the ratio of the tip clearance
turbine efficiency. From the figure, it can be observed that for the to the blade span ( c ). Here, the incidence angle is defined as the
range of the flow coefficients without stalling, the performances of arctangent of t in degrees, while the term of t is the ratio of the
both turbines are relatively almost the same. However, the stall axial velocity of inlet airflow to the circumferential velocity of the
margin of TC 0.75G1 mm is wider than that of TC 0.5G1.25 mm. blade at tip. Further, it can be seen in Fig. 10(a) that an increase in
Therefore, it can be then concluded that the performance of  c improves the stall incidence angle, or delays the stall occurrence.
TC 0.75G1 mm is better than that of TC 0.5G1.25 mm. However, this reduces the mean turbine efficiency. Meanwhile,
The computational results of the turbine performance for Fig. 10(b) illustrates a set of experimental results due to the vari-
the cases of TC 0.5G0.75 mm, TC 0.5G1 mm, TC 0.75G1 mm, ation of the mean turbine efficiency and stall incidence angle of
TC 0.75G1.25 mm, and TC 1G1.25 mm, or in term of the aver- the Wells turbine with uniform tip clearance for blade profiles of
age tip clearance to the chord length ratio, TC** = 0.0070, 0.0083, NACA18 and NACA0021 [24–26] which is quoted in Raghunathan
0.0098, 0.0111, and 0.0126, are presented in Fig. 9. In general, [5]. It can be observed that the same tendency of the respective
it can be seen that the three parameters of torque coefficient, parameters is shown between Fig. 10(a) and (b).
Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331 327

Fig. 11. Flow patterns for TC** = 0.0070: (a) contours of circumferential velocity on a plane of constant radius through 95% h; (b) contours of circumferential velocity on a
plane perpendicular to blade chord line through 95% l; (c) relative velocity vector on suction surface.

Fig. 12. Contours of circumferential velocity on a plane of constant radius through 95% h, at  = 0.39: (a) TC** = 0.0070; (b) TC** = 0.0083; (c) TC** = 0.0111; (d) TC** = 0.0139;
(e) TC** = 0.0126.
328 Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331

Fig. 13. Relative velocity vector on suction surface, at  = 0.39: (a) TC** = 0.0070; (b) TC** = 0.0083; (c) TC** = 0.0111; (d) TC** = 0.0139; (e) TC** = 0.0126.

A study is then performed on the flow-field around the turbine and 0.32. On the other hand at  = 0.39, the contour lines disappears,
blade to explain the performance of turbine with non-uniform tip and the turbine stalls. Further, as described in Zahari Taha et al. [15],
clearance. Fig. 11(a)–(c) depict the flow patterns for TC** = 0.0070, there is a strong relationship between the velocity contours such
which consist of the circumferential velocity contours on the plane as those shown in Fig. 11(a) and the effect of tip leakage flow. This
of constant radius through 95% width of flow passage (h), the cir- can be observed clearly in Fig. 11(b) of which the tip leakage flow
cumferential velocity contours on the plane perpendicular to the strongly coerces the boundary layer near the tip to separate from
blade chord line through 95% blade chord length (l), and the rela- the suction surface, in particular around the trailing edge region.
tive velocity vector on the suction surface, respectively, at  = 0.09, Moreover, the tip leakage flow also has an effect in the span-wise
0.20, 0.32, and 0.39. It is shown obviously in Fig. 11(a) that a bound- direction from the tip. These occurrences take place increasingly as
ary layer separation occurs considerable on the plane of constant the flow coefficient increases. It can then be seen in Fig. 11(c) that
radius through 95% h which is just beneath the tip. This separation the effect of tip leakage flow seems physically powerful to force
increases on the mentioned plane as the flow coefficient increases. the turbine to be in a stall condition at  = 0.39, which is indicated
Besides, it also can be seen that dense contour lines emerge near by a reversed flow region with swirl occupying a large portion of
the suction surface, which then dissipate gradually as the flow the suction surface. On the other hand at  = 0.09, 0.20, and 0.32,
coefficient increases. The emergence of the dense contour lines is most of the suction surface is relatively occupied by the main flow
attributed to counter-acting of vortex [15], which might have an velocity, which is associated with the counter-acting effect of the
effect on the turbine operating without stalling at  = 0.09, 0.20, vortex, therefore the turbine holds out to operate without stalling.

Table 2
Configurations of non-uniform tip clearance.

Turbine model Configuration of tip clearance Average size of the gap (mm)

TC 0.5G0.75 mm The size of the gap increases gradually from 0.5 mm at the leading edge to 0.75 mm at the trailing edge 0.63
TC 0.5G1 mm The size of the gap increases gradually from 0.5 mm at the leading edge to 1 mm at the trailing edge 0.75
TC 0.5G1.25 mm The size of the gap increases gradually from 0.5 mm at the leading edge to 1.25 mm at the trailing edge 0.88
TC 0.75G1 mm The size of the gap increases gradually from 0.75 mm at the leading edge to 1 mm at the trailing edge 0.88
TC 0.75G1.25 mm The size of the gap increases gradually from 0.75 mm at the leading edge to 1.25 mm at the trailing edge 1
TC 1G1.25 mm The size of the gap increases gradually from 1 mm at the leading edge to 1.25 mm at the trailing edge 1.13
Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331 329

Fig. 12(a)–(e) present the circumferential velocity contours on


the plane of constant radius through 95% h for TC** = 0.0070, 0.0083,
0.0098, 0.0111, and 0.0126, respectively, at  = 0.39. From the fig-
ures, it can be seen that dense contour lines near the suction surface
are still encountered for the cases of TC** = 0.0083, 0.0098, 0.0111,
and 0.0126, whereas not for the case of TC** = 0.0070 (stall con-
dition). Furthermore, it is interesting to observe TC** = 0.0083, in
which the dense contour lines are shown vaguely around 35% chord
length. A possible reason may be that the turbine of TC** = 0.0083 is
imminent to be in a stall condition (predicted at  = 0.41). Besides,
it also can be seen that the velocity contours for the cases of
TC** = 0.0098, 0.0111, and 0.0126 are almost the same when the
performances of the three cases are very close (see Fig. 9). How-
ever, in general it can be revealed that at  = 0.39, the dense contour
lines near the suction surface tend to emerge more explicit as TC**
increases. Hence, it is understood that the stall for TC** = 0.0126 is
more delayed than that for the others (see Fig. 9).
Fig. 13(a)–(e) depict the relative velocity vector on the suc-
tion surface of the turbine blade for TC** = 0.0070, 0.0083, 0.0098,
0.0111, and 0.0126, respectively, at  = 0.39. It can be observed
that the portion of the suction surface which is occupied by the
main flow velocity tends to become larger as TC** increases. Nev- Fig. 14. Comparison of efficiency between the turbines with uniform (TC*) and non-
uniform (TC**) tip clearances for the tip clearance to the chord length ratio of 0.0083.
ertheless, the flow patterns on the suction surface for the cases of
TC** = 0.0098, 0.0111, and 0.0126 seem almost the same when the
performances of the three cases are very close. Furthermore, it is
also shown that the turbine of TC** = 0.0083 has a strong tendency
to create a reversed flow near the tip when the mentioned turbine
is imminent to be in a stall condition as described above.

3.3. Comparison of performance between the turbines with


uniform and non-uniform tip clearances

In order to make a performance comparison between the tur-


bines with uniform and non-uniform tip clearances, computational
results of Zahari Taha et al. [15] are adopted. The comparison is
firstly presented in terms of the turbine efficiency which is based
on the same value of the tip clearance to the chord length ratio.
In this sense, there are two cases of ratio which could be consid-
ered, i.e., 0.0083 and 0.0111. Additionally comparison is presented
in terms of the mean turbine efficiency and stall incidence angle.
Figs. 14 and 15 show the efficiencies of turbines with uniform
(TC*) and non-uniform (TC**) tip clearances, with the tip clearance
to the chord length ratios of 0.0083 and 0.0111, respectively. As Fig. 15. Comparison of efficiency between the turbines with uniform (TC*) and non-
shown in the figures, the efficiency of the turbine with non-uniform uniform (TC**) tip clearances for the tip clearance to the chord length ratio of 0.0111.
tip clearance is higher than that of the turbine with uniform one
for both ratios. The peak efficiency value differs by approximately
0.76% and 0.83% for the ratios of 0.0083 and 0.0111, respectively. It
is interesting that the peak efficiencies of each case of ratio occur at
the same value of the flow coefficient of both tip clearance shapes,
which are  = 0.24 and 0.25 for the ratios of 0.0083 and 0.0111,
respectively. However, the stall margin of the turbine with uni-
form tip clearance is predicted narrower for both ratios. Further,
for the range of the flow coefficients without stalling, the turbines
with non-uniform tip clearances have the values of mean efficiency
which are higher by around 0.95% and 1.07% for the ratios of 0.0083
and 0.0111, respectively. Finally, it can be concluded that the over-
all performance of the turbine with non-uniform tip clearance is
superior to that of the turbine with uniform one.
The superiority in performance of the turbine with non-uniform
tip clearance also can be observed in Fig. 16. The figure presents a
comparison of the mean turbine efficiency and stall incidence angle
between the turbines with uniform and non-uniform tip clearances.
It can be seen that the variation of the mean turbine efficiency with
the ratio of the tip clearance to the blade span for the case of the Fig. 16. Comparison of mean turbine efficiency and stall incidence angle between
the turbines with uniform and non-uniform tip clearances.
turbine with non-uniform tip clearance is higher than that for the
330 Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331

Fig. 17. Comparison of flow patterns between the turbines with uniform (TC*) and non-uniform (TC**) tip clearances for the tip clearance to the chord length ratio of 0.0083:
(a) contours of circumferential velocity on a plane of constant radius through 95% h; (b) contours of circumferential velocity on a plane perpendicular to blade chord line
through 95% l.

Fig. 18. Comparison of flow patterns between the turbines with uniform (TC*) and non-uniform (TC**) tip clearances for the tip clearance to the chord length ratio of 0.0111:
(a) contours of circumferential velocity on a plane of constant radius through 95% h; (b) contours of circumferential velocity on a plane perpendicular to blade chord line
through 95% l.

case of the turbine with uniform one as well as the variation of the tip clearance occurs more considerable than that of the turbine with
stall incidence angle. non-uniform one. It is then clarified in Figs. 17(b) and 18(b) that the
In order to make clear due to the performance of Wells tur- mentioned occurrence is caused by the adverse effect of tip leak-
bine with uniform and non-uniform tip clearances, a comparison age flow of the turbine with uniform tip clearance which seems
of flow patterns was made. The flow patterns which are compared stronger than that of the turbine with non-uniform one. These phe-
encompass the circumferential velocity contours on the plane of nomena might give sufficient reasons why the performance of the
constant radius through 95% h and circumferential velocity con- turbine with non-uniform tip clearance is superior to that of the
tours on the plane perpendicular to the blade chord line through turbine with uniform one.
95% l. The comparison is performed at  = 0.32 and 0.39.
Figs. 17 and 18 show comparisons of the flow patterns between 4. Conclusions
the turbines with uniform (TC*) and non-uniform (TC**) tip clear-
ances, which are for the ratios of 0.0083 and 0.0111, respectively. The performance of a Wells turbine with various non-uniform
From Figs. 17(a) and 18(a) it can be observed that the boundary tip clearances has been investigated under steady flow conditions
layer separation on the plane at 95% h of the turbine with uniform using computational fluid dynamics (CFD). Good agreement has
Z. Taha et al. / Applied Ocean Research 33 (2011) 321–331 331

been achieved when the computational results of this study were [7] Watterson JK, Raghunathan S. Computed effects of solidity on Wells turbine
compared with experimental values. Further, it was shown that performance. JSME International Journal, Series B 1998;41(1):199–205.
[8] Kim TH, Setoguchi T, Kinoue Y, Kaneko K. Effects of blade geometry on perfor-
the performance of the turbine with non-uniform tip clearance mance of Wells turbine for wave power conversion. Journal of Thermal Science
has the same tendency with that of the turbine with uniform one. 2001;10(4):293–300.
The most conspicuous feature is that a turbine with a larger tip [9] Kim TH, Setoguchi T, Kaneko K, Raghunathan S. Numerical investigation on the
effect of blade sweep on the performance of Wells turbine. Renewable Energy
clearance would have a wider operational range of flow without 2002;25:235–48.
stalling. Besides, the peak efficiency of the turbine decreases and [10] Setoguchi T, Kinoue Y, Kim TH, Kaneko K, Inoue M. Hysteretic charac-
shifts towards a higher value of the flow coefficient as the tip clear- teristics of Wells turbine for wave power conversion. Renewable Energy
2003;28:2113–27.
ance increases. Nevertheless, it was found that the turbine with
[11] Mamun M, Kinoue Y, Setoguchi T, Kim TH, Kaneko K, Inoue M. Hysteretic flow
non-uniform tip clearance seems to have the overall performance characteristics of biplane Wells turbine. Ocean Engineering 2004;31:1423–35.
which is preferable. Note that this is determined by the configu- [12] Kinoue Y, Mamun M, Setoguchi T, Kaneko K. Hysteretic characteristics of Wells
turbine for wave power conversion (effects of solidity and setting angle). Inter-
ration type of non-uniform tip clearance. Furthermore, from the
national Journal of Sustainable Energy 2007;26:51–60.
flow-field investigation it was shown that the tip clearance has a [13] Dhanasekaran TS, Govardhan M. Computational analysis of performance and
strong impact on the flow patterns around the turbine blade which flow investigation on Wells turbine for wave energy conversion. Renewable
then determine the turbine performance. In this sense, the tip leak- Energy 2005;30:2129–47.
[14] Govardhan M, Chauhan VS. Numerical studies on performance improvement
age flow constitutes an important factor which plays considerable of self-rectifying air turbine for wave energy conversion. Engineering Applica-
role in developing the mentioned flow patterns. tions of Computational Fluid Mechanics 2007;1(1):57–70.
[15] Zahari Taha, Sugiyono, Sawada T. A comparison of computational and exper-
imental results of Wells turbine performance for wave energy conversion.
Acknowledgements Applied Ocean Research 2010;32:83–90.
[16] Hinze JO. Turbulence. New York: McGraw-Hill Publishing Co.; 1975.
The authors would like to thank AUN/SEED-Net (JICA) for the [17] Shih TH, Liou WW, Shabbir A, Yang Z, Zhu J. A new k–ε Eddy-viscosity model
for high Reynolds number turbulent flows—model development and validation.
providing financial support. The authors would also like to thank Computers Fluids 1995;24(3):227–38.
the Department of Engineering Design and Manufacture at Uni- [18] Kim SE, Choudhury D. A near-wall treatment using wall functions sensitized to
versity of Malaya, the Department of Mechanical and Industrial pressure gradient. In: ASME FED vol. 217, Separated and Complex Flows. ASME.
1995.
Engineering at Gadjah Mada University, and the Department of [19] Rhie CM, Chow WL. Numerical study of the turbulent flow past an airfoil with
Mechanical Engineering at Keio University. trailing edge separation. AIAA Journal 1983;21(11):1525–32.
[20] Patankar SV. Numerical heat transfer and fluid flow. Washington, DC: Hemi-
sphere; 1980.
References [21] Barth TJ, Jespersen D. The design and application of upwind schemes on
unstructured meshes. Technical Report AIAA-89-0366, AIAA 27th Aerospace
[1] Curran R, Whittaker TJT, Stewart TP. Aerodynamic conversion of ocean power Sciences Meeting; 1989.
from wave to wire. Energy Conversion and Management 1998;39:1919–29. [22] Kaneko K, Setoguchi T, Inoue M. Performance of Wells turbine in oscillating
[2] Kaneko K, Setoguchi T, Raghunathan S. Self-rectifying turbine for wave energy flow. In: Proceedings of current practices and new technology in ocean engi-
conversion. In: Proceedings of the 1st offshore and polar engineering confer- neering. 1986. p. 447–52.
ence, ISOPE. 1. 1991. p. 385–92. [23] Setoguchi T, Kaneko K, Inoue M. Determine of optimum geometry of Wells
[3] Takao M, Setoguchi T, Kinoue Y, Kaneko K. Wells turbine with end plates for turbine rotor for wave power generator. In: Proceedings of 3rd symposium on
wave energy conversion. Ocean Engineering 2007;34:1790–5. ocean wave utilization, JAMSTEC. 1986. p. 141–9.
[4] Brito-Melo A, Gato LMC, Sarmento AJNA. Analysis of Wells turbine design [24] Inoue M, Kaneko K, Setoguchi T, Raghunathan S. The fundamental characteris-
parameters by numerical simulation of the OWC performance. Ocean Engi- tics and future of Wells turbine for wave power generator. Science of Machines
neering 2002;29:1463–77. 1987;39:275–80.
[5] Raghunathan S. The Wells air turbine for wave energy conversion. Progress in [25] Raghunathan S, Setoguchi T, Kaneko K. Aerodynamics of monoplane Wells
Aerospace Sciences 1995;31:335–86. turbine—a review. In: Proc. offshore mechanics and polar engineering conf.
[6] Takao M, Nagata S, Setoguchi T, Toyota K. A study on the effects of blade profile 1991.
and non-uniform tip clearance of the Wells turbine. In: Proceedings of the ASME [26] Tagori R, Arakawa C, Suzuki M. Estimation of prototype performance and opti-
27th international conference on offshore mechanics and arctic engineering, mum design of Wells turbine. Research in Natural Energy SPEY20 1987:127–32.
OMAE2008. 2008. p. 1–8.

S-ar putea să vă placă și