You are on page 1of 20

ARTICLE IN PRESS

Journal of Wind Engineering and Industrial Aerodynamics 93 (2005) 951970 www.elsevier.com/locate/jweia

Experimental and numerical investigations of ow elds behind a small wind turbine with a anged diffuser
K. Abea,, M. Nishidab, A. Sakuraia, Y. Ohyac, H. Kiharaa, E. Wadad, K. Satod
a

Department of Aeronautics and Astronautics, Kyushu University, Hakozaki, Higashi-ku, Fukuoka 812-8581, Japan b Department of Aerospace Systems Engineering, Sojo University, Ikeda, Kumamoto 860-0082, Japan c Research Institute for Applied Mechanics, Kyushu University, Kasuga-Kouen, Kasuga 816-8580, Japan d Graduate student, Department of Aeronautics and Astronautics, Kyushu University, Hakozaki, Higashi-ku, Fukuoka 812-8581, Japan Received 5 October 2004; received in revised form 8 September 2005; accepted 19 September 2005 Available online 27 October 2005

Abstract Experimental and numerical investigations were carried out for ow elds of a small wind turbine with a anged diffuser. The present wind-turbine system gave a power coefcient higher than the Betz limit ( 16=27) owing to the effect of the anged diffuser. To elucidate the ow mechanism, mean velocity proles behind a wind turbine were measured using a hot-wire technique. By processing the obtained data, characteristic values of the ow elds were estimated and compared with those for a bare wind turbine. In addition, computations corresponding to the experimental conditions were made to assess the predictive performance of the simulation model presently used and also to investigate the ow eld in more detail. The present experimental and numerical results gave useful information about the ow mechanism behind a wind turbine with a anged diffuser. In particular, a considerable difference was seen in the destruction process of the tip vortex between the bare wind turbine and the wind turbine with a anged diffuser. r 2005 Elsevier Ltd. All rights reserved.
Keywords: Wind turbine; Flanged diffuser; Tip vortex; Separation; CFD; Turbulence; Non-linear Eddy-viscosity model; Disk-loading method

Corresponding author. Tel.: +81 926 423723; fax: +81 926 423752.

E-mail address: abe@aero.kyushu-u.ac.jp (K. Abe). 0167-6105/$ - see front matter r 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.jweia.2005.09.003

ARTICLE IN PRESS
952 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

1. Introduction The power in wind is well known to be proportional to the cubic power of the wind velocity approaching the wind turbine. This means that even a small amount of acceleration gives a large increase in the energy output. Therefore, many research groups have tried to nd a way to accelerate the approaching wind effectively [17]. Recently, Ohya et al. [8,9] have developed an effective windacceleration system. Although it adopts a diffuser-shaped structure surrounding a wind turbine like the others previously proposed, the feature that distinguishes it from the others is a large ange attached at the exit of diffuser shroud. Fig. 1 illustrates an overview of the present windacceleration system. A ange generates a large separation behind it, where a very low-pressure region appears to draw more wind compared to a diffuser with no ange. Owing to this effect, the ow coming into the diffuser can be effectively concentrated and accelerated. In this system, the maximum velocity is obtained near the inlet of diffuser and thus a wind turbine is located there as shown in Fig. 1. Although this windturbine system

(a)

(b)
Fig. 1. Schematic view of a diffuser-shrouded wind turbine: (a) overview of system; (b) ow mechanism around a anged diffuser.

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 953

has already shown some promising results [8,9], more detailed investigations of ow elds inside the diffuser are still needed to achieve the best performance. Having considered the above background, in this study experimental and numerical investigations are carried out for ow elds of a small wind turbine with a anged diffuser (i.e. a diffuser-shrouded wind turbine). In the experiments, mean velocity proles behind the wind turbine are measured using a hot-wire technique. Furthermore, computations corresponding to the experimental conditions are made to assess the predictive performance of the simulation model presently used and also to investigate the ow eld in more detail. By processing the obtained data, characteristic values of the ow elds are estimated and compared with those for a bare wind turbine. 2. Experimental apparatus All measurements were performed in a large wind tunnel at the Department of Aeronautics and Astronautics, Kyushu University. It has a measurement section of 2.5 m (width) 1:5 m (height), with a maximum wind velocity of 20 m/s. In this study, a wind turbine is located 500 mm downstream of the wind-tunnel exit. The wind turbine was a three-bladed rotor with a diameter of 388 mm. The blade was designed for the diffusershrouded wind turbine currently tested, with the aid of the design theory developed by Inoue et al. [10]. Fig. 2 gives detailed information on the diffuser-shrouded wind turbine. The diffuser consists of a main diffuser, a ange attached at the rear of the diffuser and an inlet shroud attached at the front. As shown in Fig. 2, the diameter of the diffuser throat was 400 mm, where the maximum velocity was obtained and thus where the wind turbine was located. The length of the main diffuser was 500 mm and the ange height was 200 mm. In this

Fig. 2. Detailed information of the present diffuser-shrouded wind turbine.

ARTICLE IN PRESS
954 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

experiment, the diameter of the center hub was 80 mm and hence the hub/throat ratio was 0.2. An AC servomotor was installed in the hub to change the rotational velocity of the wind turbine arbitrarily. Also, a torque meter was installed to measure the axial torque generated by the wind turbine. Fig. 3 shows the traverse system used in the present experiment. In this study, a cylindrical coordinate system was adopted, with x, r and y being the streamwise, radial and rotational coordinates, respectively. The origin was set at the location of the blades as shown in Fig. 3(b). The velocity elds behind the wind turbine were measured using a hotwire technique. In this study, a hot-wire X-probe (KANOMAX) was used, each of which consists of a tungsten wire 4 mm in diameter and 5 mm in length. Since an X-probe can provide only two components of a velocity, we performed measurements for both xy and xr components independently to obtain three-dimensional mean velocity elds. As seen in Fig. 3(a), in order not to disturb the ow eld, the probe was inserted with a supporting bar 500 mm in length. Figs. 3(b) and 4 illustrate schematic images of the measurement

Fig. 3. Traverse system: (a) overview of system; (b) detailed information.

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 955

Fig. 4. Schematic image of measurement: (a) bare wind turbine; (b) diffuser-shrouded wind turbine.

positions. In the experiments, to measure velocity elds just behind the wind turbine, we rst set the hot wire at x 48 mm (just 10 mm behind the root of the blade) and r 55 mm (15 mm away from the hub surface). From this position, the probe was traversed in the radial direction and velocity data were obtained in the ry plane (see Fig. 3(b)). Moreover, to obtain velocity distributions for the whole ow region, the probe was traversed in the streamwise direction (see Fig. 4). The free stream velocity (U 0 ) was specied as 11 and 6.8 m/s for the bare wind turbine and the diffuser-shrouded wind turbine respectively. The latter condition (6.8 m/s for the diffuser-shrouded wind turbine) was chosen so that the approaching wind velocity could be similar to that in the case of the bare wind turbine. The Reynolds number based on the diameter of the wind turbine and the free stream velocity was about 2:9 105 for the bare wind turbine and 1:8 105 for the diffuser-shrouded wind turbine. The measurement at a given position was started by a trigger signal detected when a selected blade passed a laser beam. The data-sampling rate was set to between 2 and 10 kHz so as to be enough to

ARTICLE IN PRESS
956 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

resolve in detail the rotational direction. The resolution was less than 1 (i.e. more than 360 samples/rotation) in each case. The same procedure was repeated 150300 times and the obtained data were ensemble-averaged (i.e. phase-averaged) to calculate mean values at each point. 3. Numerical method and computational conditions The present ow eld is generally expressed by the continuity and the incompressible Reynolds-averaged NavierStokes equations as follows: qU i 0, qxi qU i qP q rU j qxi qxj qxj     qU i qU j rn ru i u j F i , qxj qx i (1)

(2)

where denotes a Reynolds-averaged value. In the equations, r, P, U i , ui and n respectively denote the density, mean static pressure, mean velocity, turbulent uctuation and kinematic viscosity. In Eq. (2), F i is the body-force term imposed for the representation of a load. The present computational procedure is almost the same as that used in the previous report by Abe and Ohya [11], with some modications in specifying the newly incorporated load. The grid system and computational conditions are shown in Fig. 5. In this study, the ow was assumed to be an axisymmetric steady ow, with x and r being the streamwise and radial coordinates respectively. In Fig. 5, L, D, f and h respectively denote the diffuser length, diffuser diameter at the throat, diffuser opening angle and height of ange used in the present experiment, i.e., L=D 1:25, h=D 0:5 and f 12 . In Fig. 5(a), the subscripts 0, 1, 2 and b denote values at the inlet (free stream), in front of the load (approaching), behind the load and at the exit of the diffuser, respectively. Note that the computational domain was determined so as not to create any serious problems for the obtained results [11]. In this study, the load inside the diffuser was represented by the following general expression [11]: Ct 1 rU x jU x j; F r 0, (3) D2 where U x is the streamwise velocity. In Eq. (3), C t and D are the loading coefcient and its streamwise width imposed, respectively. In this study, C t was determined by use of a diskloading method (see for example, [12,13]), as illustrated in Fig. 6. It was found that C t can be evaluated in the two-dimensional cylindrical (xr) coordinate as follows: Fx ncC L cos b C D sin bf1 ro=U x 2 g , (4) 2 pr where n is the number of blades, o is the angular velocity of the wind turbine, c is the chord length and b tan1 U x =ro. In Eq. (4), C D and C L are, respectively, the drag and lift coefcients for the relative angle of attack, a b g, where g is the angle of blade setting (see Fig. 6). In this study, the proles of C D and C L were taken from the database [10] for two-dimensional blade sections adopted for the wind turbine. They are shown in Fig. 7. Ct

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 957

(a)

(b)
Fig. 5. Grid system and computational conditions: (a) grid system; (b) computational conditions.

On the other hand, from the relation shown in Fig. 6, the torque generated by the blades can be estimated as follows: Z r0 Z r0 1 2 rfU 2 T dT x ro gC L sin b C D cos bncr dr rh rh 2 Z r0 1 2 rU 2 x C L sin b C D cos bf1 ro=U x gncr dr rh 2 Z r0 1 2 r U2 5 x C Z 2pr dr, 2 rh where r0 and rh denote the radius of blade and hub, respectively. Similarly to the discussion on C t in Eq. (4), C Z is dened as CZ ncC L sin b C D cos bf1 ro=U x 2 g . 2pr (6)

Therefore, the total torque coefcient generated by the blades can be expressed as Z r0 Z r0 T 1 1 C trq 1 d T C Z K 2 2pr2 dr, 2 2 1 AD r U AD r U AD r r h h 0 0 2 2

(7)

ARTICLE IN PRESS
958 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

Fig. 6. Overview of disk-loading method.

1.5 1 CL 0.5 CD CL

0.06

0.04 CD 0.02 0 (Symbols: original data) 0.5 5 0 0 5 Angle of attack (deg) 10

Fig. 7. Drag and lift coefcients for relative angle of attack (a).

where A pD=22 and K U 2 =U 0 is the acceleration factor. The power coefcient is estimated as follows: Cw 1 To o 1 3 3 2 rU 0 A 2 rU 0 A Z
r0

dT
rh

1 A

r0 

rh

 ro C Z K 3 2pr dr. Ux

(8)

Such being the case, it is of interest that the performance of a wind turbine can be virtually evaluated with Eqs. (7) and (8), though no swirl effect is incorporated into the calculation.

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 959

It will be later shown that this evaluation procedure provides a powerful tool to obtain useful information of ow elds at low computational cost. As shown in Fig. 5(b), concerning the inlet boundary condition, a uniform ow (U 0 ) was specied. At the outlet boundary, zero-streamwise gradients were prescribed. Concerning the top boundary, slip conditions were adopted. No-slip conditions were specied at the walls, where the nearest node was properly placed inside the viscous sublayer. The number of grid points in this study was 156 86. The Reynolds number (Re U 0 D=n) was set to 20000. It is noted that the grid dependency and the Reynolds-number dependency of the computational results were carefully examined in the previous report [11], where it was shown that the grid resolution used was enough to obtain almost grid-independent solutions and also that the relatively lower Reynolds-number condition did not cause any serious problems in investigating fundamental ow characteristics, at least, for this kind of ow cases. This leads to a practical way to reduce the computational cost (see for detail [11]). Calculations were performed with the nite-volume procedure STREAM of Lien and Leschziner [14], followed by several improvements and substantially upgraded by Apsley and Leschziner [15]. This method uses collocated storage on a non-orthogonal grid and all variables are approximated on cell faces by the UMIST scheme [16], a TVD implementation of the QUICK scheme. The solution algorithm is SIMPLE, with a RhieChow interpolation for pressure. In this study, to predict complex turbulent ow elds with massive ow separation, the non-linear Eddy-viscosity model proposed by Abe et al. [11] was incorporated into the code. This model is based on former models by Abe et al. [1719], with some modications for the length-scale equation. Detailed descriptions of the model are given in the reference paper [11]. 4. Results and discussion Fig. 8 compares the distributions of torque coefcient (C trq ) and power coefcient (C w ) between the bare and diffuser-shrouded wind turbines. In the gure, l is a tip speed ratio dened as r0 o l . (9) U0 It is found that the shapes of both C w and C trq are similar between the bare and diffusershrouded wind turbines. However, the diffuser-shrouded wind turbine returns much higher output power compared to the bare wind turbine. The power coefcient obtained from the diffuser-shrouded wind turbine is about four times as high as that from the bare wind turbine. The basic performance of such a wind turbine has been discussed by many research groups (see for example, Rauh and Seelert [3]; Hansen et al. [5]; Inoue et al. [10]), from which it has been pointed out that a diffuser-shrouded wind turbine has the possibility of giving C w exceeding the Betz limit 16=27 owing to the effect of the diffuser placed around the wind turbine. On the other hand, Fig. 9 compares the distributions of torque coefcient (C trq2 ) and power coefcient (C w2 ) normalized by the local mean velocity just behind the turbine blades (U 2 , see Fig. 5(a)). They are dened as C trq2 1 T C trq 2 ; 2 K 2 rU 2 AD C w2 1 To Cw 3. 3 K 2 rU 2 A (10)

ARTICLE IN PRESS
960 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

(a)

(b)
Fig. 8. Basic performance of wind turbines: (a) bare wind turbine; (b) diffuser-shrouded wind turbine.

Fig. 9. Performance normalized by local mean velocity just behind turbine blades.

In the gure, l2 is alternative denition of a tip speed ratio dened as l2 r0 o l . U2 K (11)

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 961

These values (C trq2 , C w2 and l2 ) are used to investigate the performance of the wind turbine under the condition using the local mean velocity just behind the blades. It is of interest that both the bare and diffuser-shrouded wind turbines return almost the same peak performance as seen in Fig. 9, though some discrepancy is seen in the high l2 region. This may indicate that the present wind turbine operates similarly against the approaching wind even with a anged diffuser. Having considered these facts, it is thought that the augmentation of the power in the diffuser-shrouded wind turbine is mainly caused by the acceleration of the approaching wind by the anged diffuser. This is supported by the previous study of Hansen et al. [5], which evaluated the effect of a diffuser on the power augmentation obtained from a wind turbine. Abe and Ohya [11] also investigated the acceleration mechanism of the approaching wind when the anged diffuser is placed around a wind turbine. Fig. 10 gives an overall view of the ow eld around the present anged diffuser at l 4:5, in the form of the calculated streamfunction plots. As seen in the gure, a ange generates a large separation behind it. This is a notable feature of this kind of ow eld, which is different from that around a general airfoil-shaped diffuser with no ange. In fact, this separation generates a low-pressure region, owing to which more wind can be drawn into the diffuser [11]. Fig. 11 compares the computational results of C trq and C w with the experimental data. It is found that the present calculation gives satisfactory predictions for both the bare and diffuser-shrouded wind turbines. Although some overprediction is seen in higher l region of the diffuser-shrouded wind turbine, the peak performance is well predicted and the difference between the bare and diffuser-shrouded wind turbines is sufciently reproduced. Fig. 12 shows distributions of the angles in Fig. 6 at l 2:8 for the bare wind turbine. For this condition, the wind turbine gives high performance and thus no separation is expected to appear at the blade surface. As seen in the gure, the calculated relative angle of attack (a) is less than 10 degrees at every position, being sufciently involved in the range shown in Fig. 7. On the other hand, in the low l region, it is easily expected that a may become very large beyond the range in Fig. 7. To investigate ow phenomena in detail, the angular momentum was estimated by the streamwise and rotational velocities. The proles for two conditions, i.e., l 2:8 and 1:7 for the bare wind turbine, are

Fig. 10. Streamlines of the ow eld (l 4:5).

ARTICLE IN PRESS
962 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

(a)

(b)
Fig. 11. Comparison of basic performance: (a) torque coefcient; (b) power coefcient.

80 Inflow angle 60 Angle (deg) Angle of blade setting Relative angle of attack 40

20

0 0 0.2 r/D
Fig. 12. Distributions of inow angle (b), angle of blade setting (g) and relative angle of attack (a b g) at l 2:8 for the bare wind turbine.

0.4

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 963

Fig. 13. Radial distributions of angular momentum for the bare wind turbine.

compared in Fig. 13. It is readily seen that the computational results at l 2:8 correspond well to the measured data. This means that the present calculation can provide a reasonable distribution in the radial direction as well as a bulk performance for conditions with no separation at the blade surface. On the contrary, the measured data in Fig. 11 show that the performance decreases at l 1:7. As will be shown later, under this condition, separations and their related large vortex-like structures appear on the outer (tip) side. They occupy the region of 1=3 of the blade radius and are highly three-dimensional. Fig. 13 denitely explains that these large ow structures do not work effectively to generate the angular momentum (i.e. torque). Generally, such complex ow phenomena with large ow structures can never be modeled just by 2-D database with some simple corrections of 3-D effect, even if 2-D database of an airfoil are available for high angles of attack including two-dimensional massive separation. Unfortunately, at least now, there is no data (or model) available to represent such complex ow phenomena. Such being the case, calculations in the low l region were not carried out in this study. Although the present results are encouraging from the engineering viewpoint, further investigations will be needed to resolve this issue. Computational results of the acceleration factor K are compared with those of the experiment in Fig. 14, where U 2 is calculated by averaging the streamwise velocity in the section just behind the wind turbine (i.e. the load imposed by Eq. (4)). It is seen from the gure that the present computation returns generally reasonable trends for the variation of K against l, though the computation gives some underpredictions in the higher l region especially for the diffuser-shrouded wind turbine. Fig. 15 compares the calculated streamwise-velocity proles immediately behind the wind turbine with those of the experiment. The conditions were l 2:8 (l2 3:1) for the bare wind turbine and l 4:5 (l2 3:2) for the diffuser-shrouded wind turbine respectively. In the gures, the plots of the experimental data have been made by averaging the streamwise velocity in the rotational (y) direction. As described earlier, the similar condition of l2 allows us to compare the results under the condition of similar local mean velocity (i.e. local tip speed ratio) just behind the blades. As seen in the gure, a similar trend is seen in the velocity proles. The velocity slightly decreases in the region around r=D 0:4, whereas it increases

ARTICLE IN PRESS
964 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

Fig. 14. Comparison of acceleration factor.

2.5
Experiment

2 1.5 1 0.5 0 0 (a) 2.5

Calculation

Ux/U0

0.2 r/D

0.4

0.6

Experiment

2 1.5 1 0.5 0 0 (b)

Calculation

Ux/U0

0.2 r/D

0.4

0.6

Fig. 15. Comparison of streamwise velocity (x 48 mm): (a) bare wind turbine; (b) diffuser-shrouded wind turbine.

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 965

again at the end of the blade tip (around r=D 0:5). In particular, in the region close to the diffuser wall (Fig. 15(b)), the ow is greatly accelerated due to the tip clearance. The present computation successfully predicts such a trend of the velocity distribution in the r direction, though some underprediction is seen especially in Fig. 15(b). Velocity vectors in the xr plane are compared in Fig. 16, where vector plots of the experiment have been made by averaging the measured velocities in the y direction. It is found that the computed velocity vectors generally correspond well to those measured in the experiment, though some discrepancy is seen particularly in the downstream region of the bare wind turbine. Fig. 16(b) indicates that the present computation can provide useful information on the deceleration mechanism inside the diffuser. As seen in the gure, there is no separation on

(a)

(b)
Fig. 16. Comparison of velocity vectors in xr direction: (a) bare wind turbine; (b) diffuser-shrouded wind turbine.

ARTICLE IN PRESS
966 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

the diffuser wall even with a large diffuser opening angle of f 12 . This notable fact is reasonably predicted by the present computation. From these results, it can be said that the present computation sufciently reproduces fundamental features of the ow elds of this kind. Such being the case, the present calculation generally returns reasonable results. It is, however, seen that the approaching wind speed commonly tends to be a little underpredicted compared to the measured data. Since this trend is seen even in the cases of the bare wind turbine, it is thought that the main reason may lie in some insufciency in the model of the load and/or some inuence of ignoring the swirl effect in the calculation. In this respect, further efforts are needed to construct an advanced prediction model, including a new accurate disk-loading model applicable to the low l range. Although there still remains some margin for improvement, it is said that the present computational procedure is very useful for the rst assessment of performance in developing new diffusershrouded wind turbines. Fig. 17 shows a comparison of the measured velocity vectors just behind the blades for three representative rotating frequencies, i.e., 2100, 1500 and 900 rpm. It should be noted again that the present experiment adopts such conditions that local tip speed ratios (l2 ) can be similar between the bare and diffuser-shrouded wind turbines at the same rotational frequency. As seen in Fig. 9, the conditions of 2100 rpm (l2 $4) and 1500 rpm (l2 $3) are in the region of a higher tip speed ratio, where the ow is thought to be smooth with no massive separation on the blades. On the other hand, 900 rpm (l2 $2) is in the region of a lower tip speed ratio and then the results give us information of the ow structure in the stall condition. In the cases of 2100 and 1500 rpm, the ow successfully rotates in the y direction, which means that the blades effectively generate torque. In the region close to the blade tip, some vortex structures are seen, which are thought to be the tip vortices. It is of interest that similar patterns can be seen for both the bare and diffuser-shrouded wind turbines, though the tip vortex of the diffuser-shrouded wind turbine looks stronger than that of the bare wind turbine. On the other hand, in the case of 900 rpm, very large ow structures are clearly seen. These structures occupy the region of 1=41=3 of the blade radius from the blade tip. This kind of large structure is usually thought to be caused by massive separations occurring at the blade surfaces. They prevent the blades from generating torque as described in Fig. 13 and thus the performance of the wind turbine decreases as shown in Figs. 8 and 9. From these gures, it is generally said that the ow structures of the bare and diffuser-shrouded wind turbines are basically similar to each other, at least, at the location just behind the turbine blades. This corresponds well to the results of the performance normalized by the local mean velocity (U 2 ) shown in Fig. 9. In the region close to the blade tip, however, some discrepancies are seen mainly due to the effect of the diffuser wall (i.e. the tip clearance). To compare the ow structures of the bare and diffuser-shrouded wind turbines in the region downstream of the blades, Fig. 18 shows the experimental data of velocity vectors and streamwise vorticity distributions at three locations in the streamwise direction, i.e., x 48 mm (x=D 0:12), 148 mm (0.37) and 248 mm (0.52). As mentioned above, it is found again that the ow structures of the diffuser-shrouded wind turbine just behind the turbine blades are generally similar to those of the bare wind turbine except for the region close to the blade tip. On the other hand, it is of great interest that completely different trends are seen in the downstream region. In particular, rapid destruction of the vortex structure is seen in the diffuser-shrouded wind turbine, while similar vortex structures

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 967

0.5

0.5

-0.5

5 [m/s]

-0.5

5 [m/s]

-0.5

0.5

-0.5

0.5

2100rpm ( = 3.6, 2 = 3.8)

2100rpm ( = 6.3, 2 = 3.9)

0.5

0.5

-0.5

5 [m/s]

-0.5

5 [m/s]

-0.5

0.5

-0.5

0.5

1500rpm ( = 2.8, 2 = 3.1)

1500rpm ( = 4.5, 2 = 3.3)

0.5

0.5

-0.5

5 [m/s]

-0.5

5 [m/s]

-0.5

0.5

-0.5

0.5

(a)

900rpm ( = 1.7, 2 = 1.8)

(b)

900rpm ( = 2.7, 2 = 2.1)

Fig. 17. Comparison of velocity vectors behind turbine blades (x 48 mm): (a) bare wind turbine; (b) diffusershrouded wind turbine.

ARTICLE IN PRESS
968 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970
0.5 0.5

vorticity
20 16 12 8 4 0 -4 -8 -12 -16 -20 [1/s]

vorticity
20 16 12 8 4 0 -4 -8 -12 -16 -20 [1/s]

-0.5
5 [m/s]

-0.5

5 [m/s]

-0.5

0 x=48mm (x / D = 0.12)

0.5

-0.5

0 x=48mm (x / D = 0.12)

0.5

0.5

0.5

vorticity
20 16 12 8 4 0 -4 -8 -12 -16 -20 [1/s]

vorticity
20 16 12 8 4 0 -4 -8 -12 -16 -20 [1/s]

-0.5

-0.5

5 [m/s]

5 [m/s]

-0.5

0 0.5 x=148mm (x / D = 0.37)

-0.5

0 0.5 x=148mm (x / D = 0.37)

0.5

0.5

vorticity
20 16 12 8 4 0 -4 -8 -12 -16 -20 [1/s]

vorticity
20 16 12 8 4 0 -4 -8 -12 -16 -20 [1/s]

-0.5

5 [m/s]

-0.5

5 [m/s]

-0.5

(a)

0 0.5 x=248mm (x / D = 0.52)

-0.5

(b)

0 0.5 x=248mm (x / D = 0.52)

Fig. 18. Comparison of velocity vectors and vorticity distributions in ry planes at 1500 rpm: (a) bare wind turbine (l 2:8; l2 3:1); (b) diffuser-shrouded wind turbine (l 4:5; l2 3:3).

ARTICLE IN PRESS
K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970 969

clearly remain even in the far downstream region of the bare wind turbine (see Fig. 18, x=D 0:52). This is also supported by the variation of the streamwise vorticity at the three locations. 5. Concluding remarks Experimental and numerical investigations were carried out for ow elds of a small wind turbine with a anged diffuser (i.e. a diffuser-shrouded wind turbine). By processing the data obtained, characteristic values of the ow elds were estimated and compared with those for a bare wind turbine. The main conclusions derived from the study are as follows:

The present diffuser-shrouded wind turbine provided much a higher power output compared to the bare wind turbine. The power coefcient of the diffuser-shrouded wind turbine was about four times as high as that of the bare wind turbine. However, when the performance was normalized by the local mean velocity just behind the turbine blades, both the bare and diffuser-shrouded wind turbines returned almost the same peak performance. This may indicate that the wind turbine operates similarly against the approaching wind even with a anged diffuser. It is also thought that the power augmentation of this kind of wind turbine is mainly caused by the acceleration of the approaching wind by a anged diffuser. The present computational results generally showed reasonable agreement with the corresponding experimental data in the range from moderate to high l, where no separation at the blade surface was expected. On the other hand, in the low l range, the present experiment elucidated that very large vortex-like structures were generated. These structures occupied the region of 1=41=3 of the blade radius from the blade tip. This may indicate that further efforts are needed to construct a new disk-loading model representing such complex ow phenomena in the low l range. Although there still remains some margin for improvement, it is said that the present computational procedure is very useful for the rst assessment of performance in developing new diffuser-shrouded wind turbines. Flow structures of the diffuser-shrouded wind turbine just behind the blades were generally similar to those of the bare wind turbine, though some discrepancy was seen in the region close to the blade tip. On the other hand, completely different trends were seen in the region downstream of the wind turbine. In particular, it is worth noting that rapid destruction of the vortex structure was seen in the diffuser-shrouded wind turbine, while similar vortex structures were clearly seen even in the far downstream region of the bare wind turbine. This is thought to be a notable feature of this kind of diffusershrouded wind turbine.

Acknowledgements This research was partially supported by the Ministry of Education, Culture, Sports, Science and Technology, Japan (Grant-in-Aids for Scientic Research, No. 14205139, No. 15360450), the Ministry of Economy, Trade and Industry (Matching Fund), Sumitomo Fund (Environment Protection Research), Harada Memorial Fund (Fluid Machinery Research) and Kyushu University (Program and Project for Education and Research). The

ARTICLE IN PRESS
970 K. Abe et al. / J. Wind Eng. Ind. Aerodyn. 93 (2005) 951970

authors wish to express their appreciation to Mr. M. Matsubara of Kyushu University, Fukuoka, Japan for his technical help in the experiments. KA also wishes to express his appreciation to Professor M.A. Leschziner of Imperial College of Science, Technology and Medicine, London, UK for the support in using the STREAM code. References
[1] O. Igra, Research and development for shrouded wind turbines, Energy Conversion Manage. 21 (1981) 1348. [2] B.L. Gilbert, K.M. Foreman, Experiments with a diffuser-augmented model wind turbine, Trans. ASME, J. Energy Res. Technol. 105 (1983) 4653. [3] A. Rauh, W. Seelert, The Betz optimum efciency for windmills, Appl. Energy 17 (1984) 1523. [4] D.G. Phillips, P.J. Richards, R.G.J. Flay, CFD modelling and the development of the diffuser augmented wind turbine, Proceedings of the Comp. Wind Eng. 2000, Birmingham, 2000, pp. 189192. [5] M.O.L. Hansen, N.N. Sorensen, R.G.J. Flay, Effect of placing a diffuser around a wind turbine, Wind Energy 3 (2000) 207213. [6] M. Nagai, K. Irabu, Momentum theory for diffuser augmented wind turbine, Trans. JSME 53-489 (1987) 15431547 (in Japanese). [7] I. Ushiyama, Introduction of Wind Turbine, Sanseido Press, Tokyo, 1997, pp. 7784 (in Japanese). [8] Y. Ohya, T. Karasudani, A. Sakurai, Development of high-performance wind turbine with brimmed diffuser, J. Japan Soc. Aeronaut. Space Sci. 50 (2002) 477482 (in Japanese). [9] Y. Ohya, T. Karasudani, A. Sakurai, M. Inoue, Development of high-performance wind turbine with a brimmed-diffuser: Part 2, J. Japan Soc. Aeronaut. Space Sci. 52 (2004) 210213 (in Japanese). [10] M. Inoue, A. Sakurai, Y. Ohya, A simple theory of wind turbine with a brimmed diffuser, Turbomachinery 30 (2002) 497502 (in Japanese). [11] K. Abe, Y. Ohya, An investigation of ow elds around anged diffusers using CFD, J. Wind Eng. Indust. Aerodyn. 92 (2004) 315330. [12] J.N. Sorensen, C.W. Kock, A model for unsteady rotor aerodynamics, J. Wind Eng. Indust. Aerodyn. 58 (1995) 259275. [13] H. Kume, Y. Ohya, T. Karasudani, K. Watanabe, Design of a shrouded wind turbine with brimmed diffuser using CFD, Proceedings of the Annual Conference of JSAS, The West Side Division, 2003, pp. 5154 (in Japanese). [14] F.S. Lien, M.A. Leschziner, A general non-orthogonal collocated nite volume algorithm for turbulent ow at all speeds incorporating second-moment turbulence-transport closure, Part 1: Computational implementation, Comput. Methods Appl. Mech. Eng. 114 (1994) 123148. [15] D.D. Apsley, M.A. Leschziner, Advanced turbulence modelling of separated ow in a diffuser, Flow, Turbul. Combust. 63 (1999) 81112. [16] F.S. Lien, M.A. Leschziner, Upstream monotonic interpolation for scalar transport with application to complex turbulent ows, Int. J. Numer. Methods Fluids 19 (1994) 527548. [17] K. Abe, Y.J. Jang, M.A. Leschziner, An investigation of wall-anisotropy expressions and length-scale equations for non-linear Eddy-viscosity models, Int. J. Heat Fluid Flow 24 (2003) 181198. [18] K. Abe, T. Kondoh, Y. Nagano, On Reynolds stress expressions and near-wall scaling parameters for predicting wall and homogeneous turbulent shear ows, Int. J. Heat Fluid Flow 18 (1997) 266282. [19] K. Abe, T. Kondoh, Y. Nagano, A new turbulence model for predicting uid ow and heat transfer in separating and reattaching owsI. Flow eld calculations, Int. J. Heat Mass Transfer 37 (1994) 139151.

You might also like