You are on page 1of 9

47th Symposium of Applied Aerodynamics Paris, FRANCE, 26-28 MARCH

EXPERIMENTAL AND NUMERICAL STUDY OF FLOW AROUND A WIND TURBINE ROTOR 47 APPLIED AERODYNAMICS SYMPOSIUM Paris, FRANCE, 26-28 MARCH 2011 Ivan Dobrev , Fawaz Massouh , Asif Memon
(1) (2) (1) (2) (3) th

Arts et Mtiers-Paristech, 151, bd LHpital, Paris 75013, France, Email: ivan.dobrev@ensam.eu Arts et Mtiers-Paristech, 151, bd LHpital, Paris 75013, France, Email:fawaz.massouh@ensam.eu (3) Arts et Mtiers-Paristech, 151, bd LHpital, Paris 75013, France Email:asif.memon@paris.ensam.fr
An improved hybrid model of actuator surface is proposed to represent the flow around a wind turbine. The model, coupled with a Navier-Stokes solver and a blade element method, permits to calculate the power and the wake development. The blades are replaced with thin surfaces; and a boundary condition of pressure discontinuity is applied, which is determined using rotor inflow and blade-section characteristics. The proposed improvement consists to apply, in addition to pressure discontinuity which represents normal blade-section forces, body forces which represent tangential forces acting along the chord. The proposed model is validated for the flow around a horizontal axis wind turbine. The results obtained using the proposed model are compared with the experimental results obtained using PIV technique in the wind tunnel. The comparison shows applicability of the proposed method to reproduce the wake behind rotor. Its rapidity, compared to full geometry modeling, seems to be very promising for wind farm simulation.

INTRODUCTION The optimum placement of wind turbines to design a wind farm has not been an easy task, and several factors can influence the choice of optimised positioning. Evidently, only specialised software tools can help design engineers to improve energy production in view of restrictions. Presently, there exist advanced CFD software tools for wind farm design which can precisely take into account the wind profile and complexity of terrain. Mostly, in these tools, a wind turbine rotor is represented by an actuator disk model; this model gives good results if the flow is close to axial flow conditions, but the results are incorrect when the inflow is perturbed [1, 2, 3]. The aim of this study is to improve the wind turbine simulation using actuator surface (AS) model [4, 5, 6, 7, 8]. This model represents better the rotor blades, when compared to other hybrid models as the actuator disk (AD) [9] and the actuator line (AL) [10]. The AS model replaces the blades with surface of pressure discontinuity, and reproduces well normal blade section forces. However, the tangential forces are not generally taken into account, which can affect the correct flow representation for some angles of attack. In this study, an improvement of the existing AS model is proposed, and tangential blade

section forces are represented as source terms imposed in the actuator surface vicinity. To validate the proposed method, an experimental study is carried out in the wind tunnel of Arts et Mtiers ParisTech; where flow field behind the rotor is explored using particle image velocimetry (PIV) technique. The obtained velocity and vorticity fields serve as reference to compare the results obtained using improved AS model. 1. EXERIMENTAL STUDY The investigation of flow through the wind turbine is important for several reasons. Due to kinetic energy extraction, the wind speed behind the rotor decreases. An optimal wind turbine placement in a wind farm needs to avoid a rotor to operate in the wake of another machine. Thus, it is necessary to have knowledge about the development of wake behind the rotor. The flow in the wake is dominated by the vortices, which trail from the blade tips. These tip vortices play a significant role for wake meandering and diffusion. However, it is not easy to obtain useful information by conducting experiments carried out in situ. These kinds of measurements are interesting, but rarely precise and complete. The most important difficulties come from wind irregularities. For example, after several attempts, NREL abandoned in situ

measurements and carried out a successful study in the world largest wind tunnel at NASA Ames Research Center, Hand et al [11]. These results are largely used as reference for comparison with numerical simulations. Lately [12] presents project MEXICO concerning the flow downstream a three bladed horizontal axis wind turbine with rotor diameter of 4.5m. Thus, besides in situ measurements, which are important for industrial applications, it remains necessary to carry out experiments under controlled conditions. In this case, it is possible to obtain quantitative data with sufficient quality, which are needed for validating numerical simulation. 1.1. Test Bench The experiments are conducted in the wind tunnel of Arts et Mtiers - ParisTech, Fig.1. This wind tunnel is a closed circuit type and has a three-blade axial fan with a rotor diameter of 3m. The fan is driven by frequency controlled asynchronous motor with a power of 120 kW. Behind the fan, flow is decelerated in a settling chamber, which is equipped with honeycomb straighteners and wire mesh in order to smooth the flow. The tunnel nozzle accelerates the wind, from settling chamber to test section, up to 40 m/s. The nozzle has contraction ratio of 12.5, which ensures a uniform velocity profile with a turbulence ratio less than 0.25%. The semi-guided test section has a cross-section of 1.35 m x1.65 m and a length of 2 m. The static pressure in the test section is equal to atmospheric pressure. Hence, the upstream velocity depends on the stagnation pressure in the settling chamber which is measured by pressure transducer (Furness Control FC20).

The wind tunnel is equipped with sensitive 6 components balance. The forces are measured by means of strain gauges connected to data acquisition system (HBM MGCPlus). The wind turbine mast is installed on the balance, which permits to measure the rotor forces. The tested horizontal axis wind turbine has a three-blade rotor with a diameter D=540 mm. The tapered blade has a root chord of 40 mm and a tip chord of 30mm. The blade is twisted, and the pitch angle varies from 14 at the root to 3 at the tip. The blade section airfoil is NACA 4418. The nacelle diameter is only 0.08D and its length is 1.2D, Fig. 1. Thus, the vortex wake behind the rotor is non-perturbed by the mast. The rotor is mounted on a shaft coupled with a DC generator. The rotor load is controlled with a rheostat connected to the generator. The coupling between the rotor shaft and the generator is made via a contactless torque transducer (HBM T20WN). This transducer also emits 360 pulses per revolution. Because the measured power is quite low, the seals of the rotor ball bearings are removed, and the grease is replaced by thin silicon oil. A fiber optic sensor (Keyence FS20V), which detects a reflective target on the shaft, permits to locate the passage of the blade considered to be the reference. Thus, by counting the number of square signals delivered by the torque-meter, after the passage of the reference signal, it is possible to obtain the rotor's angular position with an accuracy of 1 . The acquisition of data from the sensors is carried out by data acquisition card, which emits a TTL signal for triggering the PIV measurements at a desired rotor angular position. During experiments, the upstream flow velocity V is maintained constant at 9.3 m/s, while the wind turbine rotational speed varies from 1600 rpm to 2300 rpm. Thus, Reynolds number calculated with blade tip cord and tip peripheral velocity U, varies approximately between 90,000 and 125,000. Data acquisition is carried out by a computer equipped with acquisition card. At each operating point, torque T, thrust F, rotational velocity n and upstream wind velocity V are measured continuously with a rate of 1 sample per second. To minimize the torque fluctuations, the results are averaged over a period of 5 minutes before calculating the rotor power. The power coefficient Cp, which

Figure 1. Test bench

represents the ratio of obtained power P to available power, is calculated as follows:

Cp =

P 3 V A 2 F V2 A 2

(1)

Where, is air density, and A is the rotor area. Then, the thrust coefficient is obtained:

CT =

(1)

The experiment permits to obtain the variation of the power coefficient Cp, Fig. 6 and the thrust coefficient CT, Fig. 7; as the tip speed ratio (TSR) varies. The TSR is calculated as the ratio between peripheral velocity and upstream wind velocity:

Before the tests, the calibration images are taken in order to calculate the scale, and to match spatially the camera with the blades. The plane of PIV exploration is shown on Fig. 2. In this study yaw angle is zero. Taking into account the frequency of the camera which is 7 Hz, the images are taken once every five or six revolutions. For each pair of images, the time delay between the first and second image is set to 40 s. This value has been experimentally established to ensure the best cross-correlation. The crosscorrelation with adaptation is applied to interrogation windows of 32x32 pixels and a 50% overlap, which provides a spatial resolution of 3 mm.

U TSR = V .
1.2. PIV investigation

(2)

The PIV system is controlled by Dantec software DynamicsStudio 2.30. The images are taken by using a green Nd-Yag laser with 532 nm light (Litron Nano-L 200-15), with an impulse power of 2x200 mJ, a camera of 2048x2048 px (Dantec FlowSense 4M), equipped with the lens (Micro-Nikkor AF-S 105 mm f/2.8G IF-ED), a frame grabber card and a synchronization system for synchronizing image sensor and laser flashes with the blade angular position. The flow is seeded with micro-droplets of olive oil generated by a mist generator (10F03 Dantec). The droplets diameter is approximately 2-5 m. The study is carried out for four different TSR of 4.86, 5.47, 6.08 and 7 corresponding respectively to 1600 rpm, 1800 rpm, 2000 rpm and 2300 rpm. For each point of exploration, a series of 200 pairs of images is taken. The image capture is synchronized with the rotation of the rotor, and the PIV system is triggered when the blade is positioned at the desired angular position, in this case when the reference blade is vertical. The laser sheet passes through the rotor axis and from central position of reference blade tip. To reduce the reflections of laser radiation, the blade and the rotor casing are painted with a fluorescent paint. When the blade and rotor casing are illuminated, they reflect orange light. The lens equipped with narrow band filter is transparent only for laser wavelength, and thus the saturation is overcome.

Despite image saturation near the blades due to reflections, the velocity field is exploitable and can be used for comparison with CFD simulation. Finally, the treatment of all PIV images has permitted to establish a database containing instantaneous and average velocity fields for each of the four TSR. Averaged flow and vorticity fields for azimuth position of 0 are presented on Fig. 8 and Fig. 9

Figure 2. PIV plane of measurement

2. NUMERICAL MODEL 2.1. Hybrid modelling The flow around a wind turbine is governed by the 3D incompressible Navier-Stokes equations:

r r r V r 1 r = f grad p + V (V )V + t

Here, the velocity field is obtained as a solution of the boundary value problem with non-slip boundary condition on the wind turbine surfaces V=0. However, in order to obtain a solution with accuracy, high mesh density is needed, in the vicinity of the rotor blades in order to resolve the blade section boundary layer. Usually, the number of nodes needed to model the rotor is equivalent to those used for the wake modelling. To facilitate the wake modelling and to reduce the computational cost, so-called hybrid modelling has been developed [1]. In this modelling, the real rotor

angles of attack, it is important to represents these forces. In this work, in order to improve the forces representation, tangential forces are imposed as body forces in a thin layer near the surface of pressure discontinuity. The actuator hybrid modelling is carried out by using two modules. The first module is a CFD solver, which calculates the flow in simulation domain with appropriate boundary condition. The second module extracts the inflow, and use the blade element theory to determine the pressure discontinuity. The solution is carried out iteratively, exchanging data between the blade element and CFD modules. Starting from initial conditions of the upstream flow, and using the blade geometry and the airfoil data, the blade element module calculates the pressure jump distribution on the surface, which replaces the blade. The CFD module computes the flow of velocity field, using a boundary condition of the pressure distribution previously obtained from the blade element module. The calculation stops, after several iterations, as the flow in the wake reaches convergence. The calculation of pressure discontinuity is based on the blade element approach. At the blade radius r, the elementary forces acting in the normal and tangential directions on a blade element with span dr and chord c are:

geometry is replaced by source terms f which create the equivalent forces. Depending on

how the source terms f are distributed, there are three types of hybrid models: actuator disk, actuator line and actuator surface. In the case of wind turbine, the boundary value problem is solved for Navier-Stokes equations using appropriate non-slip boundary condition on the surfaces, which corresponds to walls. As a result, the obtained velocity field depends directly on the blade sections geometry. Contrarily, in the case of hybrid modelling, there are no walls within the fluid domain; and the blades are replaced by the volume or surface forces. In this case, the boundary conditions are applied only on external surfaces of the domain. A priori, the blade forces are unknown, and their values are obtained iteratively during the calculation using the blade element method. If the flow along the blade sections is close to the two-dimensional flow, the blade forces depend on inflow and blade sections aerodynamics properties. This assumption is justified for most blade sections except at the root and at the tip, where the flow is 3D. 2.2. Actuator surface model In the actuator surface model, the rotor geometry is simplified, and the blades are replaced by surfaces with a boundary condition of the type pressure discontinuity. The surface forces, which replace the rigid blade wall, are calculated from inflow and airfoil performance. In this case, the grid density depends only on pressure distribution gradient. Hence, the number of computation nodes is significantly reduced; as there is no need to model the boundary layer. It must be noted that the actuator surface model can represent only the blade section normal forces. Usually, the blade section tangential forces are lower than normal forces and can be neglected. However, for some

d Fn =
and

1 W 2 c Cn ( )d r 2 1 W 2 c Ct ( )d r 2

(2)

d Ft =

(3)

In the above formulas, the force coefficients Cn and Ct are determined using the aerodynamic blade sections performances CL=CL () and CD=CD (). The angle of attack is:

(r ) = (r ) (r ),

(4)

where is the blade section pitch angle, and is the flow angle between the plane of rotation and the reference relative velocity W. It is very important to choose the right point where the reference velocity W and flow angle are to be determined. Unfortunately, these flow parameters cannot be extracted directly from the inflow. The theory of Froude-Rankine, applied for the wind turbines, states that due to energy extraction, the flow through the rotor slows. Betz shows that the maximum extracted energy corresponds to a velocity in the rotor plane equals to 2/3 V. Therefore, the reference point must be chosen sufficiently

close to the rotor. However, the flow field in the rotor plane is perturbed by the blade sections and a method is needed to obtain the reference velocity. The simplest method of obtaining this velocity is to calculate the mean velocity in the plane of rotation, similarly as of Glauert, but in the case of yaw, the results become incorrect. Another method to choose this reference point is proposed in [7]. The method is based on assumption that the inflow is a result of a local induced flow, created by the presence of the blade section airfoils and the actuator disk, which extracts kinetic energy and decelerates the flow, passing through the rotor. As a result, the reference point can be located sufficiently close to the rotor, if the blade perturbations are excluded by subtracting the velocity field induced by the blade from those obtained by CFD solver. 3. NUMERICAL MODEL The studied wind turbine model is similar to that tested in the wind tunnel. The choice of studying this wind turbine model permits to compare the results of wake simulation with experimental data of high quality. If the large wind turbine is studied, like in MEXICO project [12], only a limited comparison of flow field will be available. In fact, the PIV measurements are constrained by the laser power and camera resolution to limited areas.

5D, and here the grid approaching the rotor becomes dense. The second part, containing the pressure discontinuity surface, is meshed finely. This high grid density is needed in order to represent velocity profile behind the rotor, which is important for initial development of the wake. In order to describe the form of pressure discontinuity, the surface which replaces the blade is divided into 80 intervals in chordwise direction and into 100 intervals along the blade span. The mesh is finest near leading edge and, blade tip and root because the pressure gradient in these areas is more important. The initial cell layer height is equal to 0.02 chord length, and cell growth ratio is 1.05, which is needed to distribute precisely the body forces acting in chordwise direction. The third part of simulation domain has a length of 1D, and finely meshed, similarly as that of the second part. This part contains most of the length of near wake, and the high grid density is needed in order to represent the vortices, trailing from the tips of the blades. In this part, the cell size represents 0.005D. The fourth and last part of the volume is similar to the first, and the mesh density decreases away from the rotor. This part represents the development of the far wake. In radial direction, each of four parts is divided into three annuli: internal, intermediate and external. The internal annulus has diameter of 1D and contains the most part of the wake. The intermediate annulus has a diameter of 1.4D and contains the tip vortices and wake shear layer. Here, the cells are cube-shaped in order to avoid grid oriented numerical diffusion of the vortices. The internal and the intermediate annuli have a denser grid as compared to the external annulus, because the wake is developed here. The external volume is needed to represent free stream velocity, not perturbed by rotor. The total number of cells of this multi-block structured grid is 13 million. The grid density is limited in order to perform the simulations on a single multiprocessor machine. Thus, in case of wind farm, where the simulation is performed on cluster of PCs, one computer will calculate the flow around one wind turbine. All simulations are carried out using CFD solver ANSYS Fluent 12.1. The simulations are performed with upstream velocity of 9.3 m/s and turbulence of 0.3%, which correspond to wind tunnel conditions. To calculate the

Figure 3. Simulation domain.

To facilitate the calculation, the simulation domain represents cylindrical sector having a blade period, which represents one third of the entire domain as the rotor has three blades. The domain, having a diameter of 5D and a length of 15D, is divided into four parts, Fig. 3. The first part, rotor upstream, has a length of

performance of studied wind turbine, only the angular velocity is varied, keeping the upstream velocity and turbulence constant. A velocity inlet boundary condition is applied on the inlet and on the peripheral surfaces, and the rotational periodicity boundary condition is applied on lateral surfaces of the simulation domain. On the surface, which represents the blade, boundary condition of a fan is applied. The pressure distribution is calculated from blade section inflow, blade geometry and airfoil performance of NACA 4418, see Fig.4, [13]. Because the experimental pressure distribution is unknown, a simplified pressure distribution is applied [4, 5]. In this study, the reference velocity is calculated at a distance of one chord upstream the blade section. At this distance, the perturbation due to the blade is relatively important. To take into account this perturbation, the velocity induced by the blade sections is subtracted from velocity calculated by the CFD solver. This induced velocity is assumed to be equal to that induced by an airfoil in potential flow which has a vorticity distribution equivalent to blade pressure distribution, Fig. 5.

for momentum equation is a second order upwind.

Figure 5. Hybrid model calculation

Calculation is carried out iteratively and after numerous iterations, according to the number of the nodes and the value of the residuals required, the convergence is achieved. The computed rotor power reaches a constant value quickly, for example, after 500 iterations; but additional iterations, more than 3000, are needed to obtain the wake development. 4. NUMERICAL AND EXPERIMENTAL RESULTS 4.1. Power and Thrust The comparison between wind turbine performance obtained experimentally and numerically using actuator surface is presented in Fig. 6 for coefficient of power and Fig. 7 for the coefficient of thrust. Curves are denoted with calc for simulation results and exp for experiment results.

Figure 4. NACA 4418 aerodynamic performance

To obtain this pressure distribution, for each node of pressure discontinuity surface, it is needed to determine the velocity, the local angle of attack , the distance s/c to the leading edge and the relative radius r/R. To take into account the angular rotation of the rotor, a multi-reference frame (MRF) is used. The turbulence model is detached eddy simulation (DES), and discretization scheme

Figure 6.Variation of coefficient of power with tip speed ratio

The simulation results for the power are satisfactory for TSR equal or greater than optimum. The reason for the discrepancy can be found in high angles of attack which occurs for low TSR, and flow becomes threedimensional and blade element theory is irrelevant for 3D flow. Additionally, as experimental results show [12], at internal blade sections, close to the root, the bladesection performance becomes considerably different from those obtained for 2D airfoils.

the vortices which turn in clockwise direction, as is intended for a wind turbine. The tip vortices induce velocities which are added to the free stream velocity. As a result, the flow in the external zone is accelerated, because the induced velocities have same direction as free stream velocity, Fig.8. Conversely, in the internal zone, the vortices induce velocity in opposite direction of the free stream, and the flow is decelerated. This deceleration is more important close to the tip vortices. Downstream of the rotor, the helical vortices decelerate the internal flow, and thus the diameter of wake increases gradually.

Figure 7. Variation of coefficient of thrust with tip speed ratio

4.2. Wake analysis The analysis of flow downstream the rotor permits to understand better the wake development. The Fig. 8 shows the dimensionless average velocity field, and the Fig. 9 shows average vorticity field, depending on TSR. In these figures, the plane of exploration passes through axis of rotation, the reference blade is vertical, and the flow is developed from left to right. The tip vortices emanating from the blades move with fluid, and as a consequence, their form is helical. The plane of exploration cuts the helices, and on raw PIV images, the tip vortices appear as dark circles. In fact, the centrifugal forces, due to rotation of the vortex core, separate the olive droplets which reflect the laser lights. It is easy to observe tip vortices, trailing from the blades, which govern the flow in the wake. In fact, the flow around the rotor is divided into two zones: an internal zone, when the wake restrained by the tip vortices, and an external zone, less perturbed free flow. On the Fig. 9, the tip vortices with high negative values are represented as blue circles; these negative values correspond to
Figure 8. Averaged velocity field downstream of wind turbine rotor

On the velocity field shown on Fig. 8, the positions of the tip vortices can be found, where the velocity varies rapidly in radial direction and its gradient becomes stronger. Experimental results show that at vortex core center, velocity varies radially from values close to zero to values two times greater than upstream velocity. Generally, the wake simulations are satisfactory for axial and radial positions of the tip vortices, except for the case of high TSR, where the wake expanding is less than experimental measurements, Fig. 10.

The result for the rotor thrust coefficient, Fig. 7 shows, how the thrust increases with increasing TSR. This result is confirmed by the velocity field measurements; and the wake diameter increases with increasing TSR. Thus, high thrust leads to more deceleration. The velocity gradient obtained from simulations is less noticeable: tip vortex has large diameter and its intensity is much lower. The difference between experiment and simulations is due to relatively important grid coarseness. In fact, the cell size is comparable with vortex core radius. As a result, the vortex core is not well reproduced, and numerical diffusivity increases. However, as is noted earlier, due to computer memory limitation, the cell size cannot be reduced. The first vortex, represented on the left side of the images, is emitted from the blade which passes just before the reference blade has . Here, vortex age is vortex age of VA=120 defined as the azimuthal travel of the blade since vortex creation. Experiment shows that vortex sheet trailing from the blade is completely rolled up and almost all of the vorticity concentrates at tip vortex. Contrarily, in simulations, the blade wake is well apparent for all TSR, Fig. 9.

Figure 10. Vortex center positions

5. CONCLUSION An improved hybrid model based on the actuator surface is proposed in order to facilitate the simulation of wind farms. The objective of the paper has been to demonstrate capabilities of the proposed model to represent the wake behind the rotor. The simulation domain is simplified and blades are replaced with pressure discontinuity surfaces. In the case of complete geometry simulation, the boundary condition applied on the blades is

Figure 9. Averaged vorticity field downstream of wind turbine rotor

non-slip, which needs a high quality mesh to represent boundary layer. However, the appropriate modeling of the boundary layer is crucial for simulation accuracy. Conversely, in the case of hybrid modeling, there are no walls and the blades are represented by the volume and the surface forces. These forces are determined using blade sections inflow and their aerodynamics properties, thus the grid density can significantly be reduced. The actuator surface has the same pitch angle and chord as the original blade. On these surfaces, a pressure discontinuity is applied. The airfoil aerodynamics characteristics are obtained from wind tunnel experiments, but the pressure distribution is unknown. In this work, a pressure discontinuity is applied, which is similar to a thin flat plate pressure distribution, but without singularity at the leading edge point. The proposed model is tested in the case of flow around a horizontal axis wind turbine model for four tip speed ratios. The results of simulations are compared with the results of experiments carried out in the wind tunnel. The comparison shows efficiency of the proposed actuator surface model to reproduce the rotor mechanical power and forces, but for low tip speed ratios some discrepancy with experimental results is revealed. In case of low tip speed ratios, the angles of attack increase along the blade span, and the flow becomes detached and highly three-dimensional, where blade element method, used to determine the aerodynamic forces, is not adequate. Generally, the wake simulations are satisfactory for axial and radial positions of the tip vortices, except in the case of high TSR, where the wake expanding is less than experimental. Actually, it is difficult to find the reason of this discrepancy because the rotor thrust is very well calculated. Probably the tip vortices instability, which occurs for high TSR, plays an important role. It must also be noted that the simulations cannot reproduce the recirculation zone behind the rotor.

3. Sanderse, B., van der Pijl, SP. & Koren, B. (2011). Review of computational fluid dynamics for wind turbine wake aerodynamics. Wind Energy, Vol.14, 7, 799-819. 4. Dobrev, I. & Massouh, F. (2005). Etude d'un modle hybride pour reprsenter l'coulement travers un rotor olien. 17 CFM, Troyes, France. 5. Dobrev, I., Massouh, F. & Rapin, M. (2007). Actuator surface hybrid model. J. of Physics: Conf. Ser. 75,1-7. 6. Sibuet Watters, C., Breton, S.P. & Masson, C. (2007). Recent advances in modeling of wind turbine wake vortical structure using a differential actuator disk theory. J. of Physics: Conf. Ser. 75. 7. Shen, W., Zhang, J., & Srensen, J. (2009). The Actuator Surface Model: A New NavierStokes Based Model for Rotor Computations. Journal of Solar Energy Engineering, 131. 8. Sibuet-Watters, C., Breton, S.P. & Masson C. (2010). Application of the actuator surface concept to wind turbine rotor aerodynamics, Wind Energy, vol.13 (5). 9. Mikkelsen, R. (2003). Actuator Disc Methods Applied to Wind Turbines. Ph.D. Thesis, Technical University of Denmark. 10. Srensen, J., & Shen, W. (2002). Numerical modelling of wind turbine wakes. J Fluids Eng, 124 (2), pp. 393399. 11. Hand M., Simms D., Fingersh L.J., et al, (2001). Unsteady Aerodynamics Experiment Phase VI: Wind Tunnel Test Configurations and Available Data Campaigns. TR NREL/TP-500-29955. 12. Schepers, JG. & Snel, H. (2007). Model Experiments in Controlled Conditions, ECN Wind Energy, ECN-E--07-042. 13. Ostowari C. & Naik D. (1985). Post Stall Studies of Untwisted Varying Aspect Ratio Blades with NACA 44XX Series Airfoil Sections - Part II. Wind Engineering Vol. 9, No.3, 149-164. 14. Wang S. Rusak Z (1997). The dynamics of a swirling flow in a pipe and transition to axisymmetric vortex breakdown. J. Fluid Mech. vol. 340, pp. 177-223. 15. Brcker, C. (2003). Some Observations of Vortex Breakdown in a Confined Flow with Solid Body Rotation.Flow, Turbulence and Combustion 69: 6378.

6. REFERENCES 1. Vermeer, LJ., Sorensen, JN. & Crespo, A. (2003). Wind turbine wake aerodynamics. Progr. Aerosp. Sci. 39 46751. 2. Chen , Z., Marathe, N., Dharmarathne, S., Hu, Y. & Parameswaran, S. (2011). Review of wind turbine wake studies in the past 5 years. ICWE 13, Netherlands.

You might also like