You are on page 1of 19

WIND ENERGY Wind Energ. 2002; 5:151169 (DOI: 10.1002/we.

64)

Research Article

NavierStokes Predictions of the NREL Phase VI Rotor in the NASA Ames 80 ft 120 ft Wind Tunnel
N. N. Srensen, Wind Energy Department, Ris National Laboratory, DK-4000 Roskilde, Denmark J. A. Michelsen, Department of Mechanical Engineering, Technical University of Denmark, DK2800 Lyngby, Denmark S. Schreck, National Renewable Energy Laboratory, 1617 Cole Boulevard, Golden, CO 80401, US

Key words: NavierStokes; computational uid dynamics; wind turbine aerodynamics; NREL Phase VI

This article describes the application of an incompressible Reynolds-averaged NavierStokes solver to several upwind cases from the NREL/NASA Ames wind tunnel tests. In connection with the NREL blind code comparison the present results showed the overall best agreement with experimental measurements. Based on this, it is of great interest to demonstrate the quality that can be obtained in 3D CFD rotor computations. All six cases we present have 0 yaw angle and 3 tip pitch angle. All computations are performed as rotor-only computations, excluding the tower and nacelle. In this article we compare computed results and measurements in the form of shaft torque, ap and edge moments, aerodynamic coefcients and pressure distributions as a function of wind speed. The spanwise force distributions are compared with measurements for all wind speeds, along with the pressure distributions at ve spanwise positions. Finally, we show how 3D CFD computations can be used to extract information about three-dimensional aerodynamic effects. Copyright 2002 John Wiley & Sons, Ltd.

Introduction
In the design process of new rotor blades the need for accurate aerodynamic predictions is very important. During the last couple of years a large effort has gone into developing CFD tools for prediction of wind turbine ows. Previously, full NavierStokes simulations of wind turbine rotors have been performed by Duque,1 Xu and Sankar,2,3 Srensen and Hansen4 and Srensen and Michelsen.5 Additionally, a large number of rotor simulations have been performed in the helicopter community. In the model development phase, validation against measurements is very important. Typically, only integrated loads are available from fullscale wind turbine measurements under atmospheric conditions. Additionally, full-scale measurements are often contaminated by varying wind speed, changes in wind direction, etc. Detailed sectional information in the form of pressure distributions has only been measured to a very limited degree.6 The NREL/NASA Ames wind tunnel test7 and the following NREL/NWTC aerodynamics blind comparison8 therefore offered a unique opportunity to assess and demonstrate the capability of modern CFD tools to predict rotor ows in comparison with other approaches. Herein all computations were performed without prior knowledge of the measured results. The following is a presentation of the results computed using the EllipSys3D CFD code for
Correspondence to: N. N. Srensen, Ris National Laboratory, Wind Energy Department, PO Box 49, DK-4000 Roskilde, Denmark. E-mail: niels.sorensen@risoe.dk Contract/grant sponsor: Danish Energy Agency; Contract/grant number: ENS 1363/00-0007.

Copyright 2002 John Wiley & Sons, Ltd.

Received 13 November 2001 Revised 5 March 2002 Accepted 20 March 2002

152

N. N. Srensen, J. A. Michelsen and S. Schreck

the blind comparison. Additionally, computations of a conguration consisting of a simplied representation of the wind tunnel are included to investigate the inuence of tunnel blockage effects on the results.

Method
NavierStokes Solver
In the present work an incompressible NavierStokes solver is applied to predict the aerodynamics of the unsteady aerodynamics experiment Phase VI rotor from the National Renewable Energy Laboratory. The two-bladed 10058 m diameter Phase VI rotor geometry is based on the S809 aerofoil, and details about the blades can be found in Reference 9. For the cases considered in the present study, the rotor cone angle was set at 0 . The blade pitch angle was set at 3 , which rotated the blade tip chordline 3 towards feather, relative to the rotor plane, pointing the leading edge into the oncoming wind. In this investigation, only the upwind conguration will be examined, and the operational conditions for the cases computed can be found in Table I. The Reynolds number varies between 07 and 14 106 at the root and between 10 and 11 106 at the blade tip. The inuence of the tower and nacelle on the rotor aerodynamics will be neglected, which is a fair approximation for an upwind turbine. Besides removing the one-per-revolution periodic interference between the tower and the individual rotor blades, this also greatly simplies the geometrical complexity of the ow problem. Additionally, assuming zero vertical shear in the incoming ow and zero yaw misalignment, the blades see the same inow conditions irrespective of the azimuthal position of the rotor. These simplications result in a problem where only the rotor needs to be modelled. As uctuations from local ow separations on the blades may still lead to ow unsteadiness, both steady and unsteady computations are carried out. For further details of the turbine properties, see Reference 10. The computations submitted for the NREL aerodynamics blind comparison were performed assuming that the tunnel walls could be ignored, and only computations of a free rotor were carried out. Having compared the results with measurements, the question of whether inclusion of the tunnel could bring the predictions even closer to measurements arose. A new series of computations was performed where a cylindrical approximation of the actual tunnel was included to determine the effect of tunnel blockage on the computed results. In the present work, only one of the blades is explicitly modelled in the computations. The remaining blade is accounted for using periodic boundary conditions, exploiting the 180 symmetry of the two-bladed rotor. The information needed for this type of CFD computation is the geometry of the blade and the operational parameters of the rotor (RPM, density, etc.). No empirical tuning is needed for a given rotor. The in-house ow solver EllipSys3D is used in all computations presented in the following. The code was developed in co-operation between the Department of Mechanical Engineering at DTU and the Department of Wind Energy at Ris National Laboratory.11 13 The EllipSys3D code is a multiblock nite volume discretization of the incompressible Reynolds-averaged NavierStokes (RANS) equations in general curvilinear co-ordinates. The code uses a collocated variable arrangement, and Rhie/Chow interpolation14 is used to avoid odd/even pressure decoupling. As the code solves the incompressible ow equations, no equation of state exists for the pressure, and the SIMPLE algorithm of Patankar and Spalding15 is used to enforce the
Table I. Operational conditions for the runs computed in the present work Run RPM Wind speed (m s 1 ) 70 100 130 151 201 251 Density (kg m 3 ) 1246 1246 1227 1224 1221 1220 Viscosity 105 (kg m 1 s 1 ) 1769 1769 1781 1784 1786 1785

S070000 S100000 S130000 S150000 S200000 S250000

719 721 721 721 720 721

Copyright 2002 John Wiley & Sons, Ltd.

Wind Energ. 2002; 5:151169

NavierStokes Predictions

153

pressure/velocity coupling. The EllipSys3D code is parallelized with MPI for execution on distributed memory machines, using a non-overlapping domain decomposition technique. For rotor computations a non-inertial reference frame attached to the rotor blades is used, and terms accounting for the Coriolis and centripetal forces are added to the equations. Polar velocities (vr , v , vz ) are used to allow simple treatment of periodic boundary conditions in the azimuthal direction.5,16 The solution is advanced in time using a second-order iterative time-stepping (or dual time-stepping) method. In each global time step the equations are solved in an iterative manner, using underrelaxation. First, the momentum equations are used as a predictor to advance the solution in time. At this point in the computation the ow eld will not full the continuity equation. The rewritten continuity equation (the so-called pressure correction equation) is used as a corrector to make the predicted ow eld satisfy the continuity constraint. This two-step procedure corresponds to a single subiteration, and the process is repeated until a convergent solution is obtained for the time step. When a convergent solution is obtained, the variables are updated and the computation continues with the next time step. For steady state computations the global time step is set to innity and dual time stepping is not used. This corresponds to the use of local time stepping. To accelerate the overall algorithm, a three-level grid sequence is used in the steady state computations. The convective terms are discretized using a second-order upwind scheme, implemented using the deferred correction approach rst suggested by Khosla and Rubin.17 Central differences are used for the viscous terms. In each subiteration, only the normal terms are treated fully implicitly, while the terms from non-orthogonality and the variable viscosity terms are treated explicitly. Thus, when the subiteration process is nished, all terms are evaluated at the new time level. The three momentum equations are solved decoupled using a red/black GaussSeidel point solver. The solution of the Poisson system arising from the pressure correction equation is accelerated using a multigrid method. In the present work the turbulence in the boundary layer is modelled by the k SST model of Menter.18 The details of the model will not be given here. The model is chosen because of the very promising results for 2D separated ows.19,20 The equations for the turbulence model are solved after the momentum and pressure correction equations in every subiteration/pseudo time step. In the present work, all computations are performed assuming fully turbulent ow, excluding any laminar and transitional effects at the leading edge region of the rotor.

Computational Mesh
As previously discussed, two different congurations were computed in the present work. One is called the free conguration, where the rotor is modelled without the presence of any external disturbances using a very large domain. The other is called the tunnel conguration, where a cylindrical approximation of the actual tunnel is included in the computations. In the tunnel conguration a cylindrical cross-section with an area equal to the actual tunnel cross-section (244 m 366 m) is used, corresponding to a radius of 1686 m. The cylindrical approximation of the tunnel is used to avoid the need for sliding meshes, where one part of the computational mesh is moving with respect to another, which would be necessary if the actual tunnel were to be modelled. The mesh consists of three main components, shown in Figures 1 and 2. First, an inner ve-block OO topology is used locally around the blade. This mesh component is identical for the two congurations. Second, an outer section is wrapped around the inner OO section, containing three blocks for the free conguration and ve blocks for the tunnel conguration. The inner OO section and the outer section both cover only 90 in the azimuthal direction. Finally, an additional mesh component is used to cover the remaining 90 of the computational domain. This section is simply generated by rotating the periodic plane of the 90 section. In total, the mesh consists of 12 blocks of 643 cells or a total of 31 106 cells for the free conguration and 16 blocks of 643 cells or a total of 42 106 cells for the tunnel conguration. The size of the inner OO section is difcult to see in Figures 1 and 2, but the upstream and downstream faces are placed approximately 1 m away from the rotor plane. It has 64 cells in the direction normal to the surface, 256 cells around the aerofoils and 64 cells in the spanwise direction. To facilitate the resolution of the tip of the blade, a 64 64 block is placed at the tip. The total number of cells for the ve-block inner OO
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

154

N. N. Srensen, J. A. Michelsen and S. Schreck

Inlet

Symmetry Planes

Figure 1. View through the outlet part of the mesh for the free conguration showing half of the domain (180 azimuth), with the inner OO section and the blade in the central part of the picture, the two 180 symmetry planes and the inlet/far-eld boundary

section is 13 106 . To resolve the boundary layer, the y C values at the wall are kept below 2 everywhere on the blade surface. Given the long computing times required for these 3D computations, the effect of grid size on the solution accuracy cannot be readily evaluated at present. Instead, experience from previous 2D aerofoil computations and 3D rotor computations regarding the necessary number of mesh points was used when designing the 3D mesh.

Boundary Conditions
In the following the boundary conditions are described for the two congurations, starting with the free conguration and nishing with the tunnel conguration. Free Conguration For the free conguration, zero axial gradient is enforced at the outlet at the downstream end of the spherical domain where the ow leaves the domain. At the inner cylindrical face near the rotational axis, Euler/slip conditions are applied, while no-slip conditions are applied at the surface of the rotor blade. Fully implicit periodic conditions are applied at the 180 cyclic boundaries. At the upstream part of the spherical domain the undisturbed wind speed is specied. In the present computations, no attempt was made to include the tunnel walls and the corresponding blockage effects. Instead, a very large computational domain was used and the rotor was computed in an undisturbed environment. With the present computational domain, which
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

NavierStokes Predictions
Tunnel Wall Symmetry Plane

155

Outlet Inlet

Symmetry Plane

Figure 2. Overview of the tunnel mesh showing half of the domain (180 azimuth), with the inner OO section and the blade in the central part of the picture, the two 180 symmetric planes, the inlet and outlet planes and part of the outer tunnel wall. The gure shows every second mesh point only

placed the far-eld boundary approximately 6 rotor diameters away from the rotor blades, no correction for the presence of the rotor was necessary at the inlet boundary. Tunnel Conguration For the tunnel conguration a zero-axial-gradient condition is applied at the downstream boundary of the cylindrical domain where the ow leaves the domain. At the inner face near the rotational axis, Euler/slip conditions are applied, while no-slip conditions are applied at the surface of the rotor blade. Fully implicit periodic conditions are applied at the 180 cyclic boundaries. At the upstream boundary of the cylindrical domain the undisturbed wind speed is specied, while slip conditions are applied on the outer cylindrical part of the domain.

Computing Time
For all computations shown in the present article, the Ris IBM SP parallel computer was used. It is equipped with eleven 375 MHz Power-3 SC Thin Nodes, each with two CPUs, and a 150 Mb s 1 switch connection between the nodes. For the free conguration, both steady and time-accurate computations were performed. The steady state computations were continued for approximately 4500 iterations and took approximately 30 h per case on four CPUs. The unsteady computation used the steady state ow eld as a starting condition and carried out 830 time steps per revolution for 34 revolutions, which took approximately 408 h using four CPUs. For the tunnel conguration, only steady state computations were performed, as the previous free conguration had shown very limited difference between the steady and time-accurate computations. The
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

156

N. N. Srensen, J. A. Michelsen and S. Schreck

steady state computations took approximately 40 h for 4500 iterations using four CPUs. Even though the lack of time-dependent behaviour may be caused by the relatively large time step used in the computations, limitation due to computing time prevents further time step renement. Based on these facts, the application of time-accurate computations was not considered worthwhile for the tunnel conguration.

Results and Discussion


In the following, the results of the NavierStokes simulations are compared with the results of the NREL/NASA Ames wind tunnel test. As stated earlier, the computations for the free conguration were all completed before the release of the results from the measuring campaign. In contrast to this, the tunnel conguration computations were completed after the release of the measurements. In the following, integrated quantities (low-speed shaft torque, root ap and edge moments) will be presented rst. Then the spanwise distribution of normal and tangential forces will be compared to measurements, followed by a comparison of computed and measured pressure distributions. All measurements except the pressure distributions are averaged over several full revolutions, and one standard deviations are shown where available to indicate the variation over one revolution. The standard deviation of the measurements comprises several sources, including the disturbance from tower passage, dynamic loads, etc. The measured pressure distributions represent averages of data acquired over the upper half of the rotor disc, where the inuence of the tower is negligible.

Integrated Loads
Looking rst at the integrated loads, these are obtained from the computations by integrating the pressure and skin friction over the blade surface. In Figure 3 the low-speed shaft torque computed by the EllipSys3D CFD code is compared to measurements. Even though quantitative differences exist, the overall shape of the computed torque curve agrees well with the measured curve for both the free and the tunnel conguration. The shaft torque at stall onset (10 m s 1 ) is overpredicted by 20%. This amount of overprediction is in the range found in previous rotor computations.21 At higher wind speeds under heavily stalled conditions the simulations
1700 Measured 1600 Comp. Free Comp. Tunnel

Low Speed Shaft Torque [Nm]

1500 1400 1300 1200 1100 1000 900 800 700 6 8 10 12 14 16

18

20

22

24

26

Wind Speed [m/s]


Figure 3. Comparison of measured and computed shaft torques for the NREL Phase VI rotor. For all measurements, one standard deviations are shown to indicated the variation over one revolution
Copyright 2002 John Wiley & Sons, Ltd.

Wind Energ. 2002; 5:151169

NavierStokes Predictions
5000 Measured 4500 4000 3500 3000 2500 2000 1500 1000 Comp. Free Comp. Tunnel

157

Root Flap Moment [Nm]

10

12

14

16

18

20

22

24

26

Wind Speed [m/s]


Figure 4. Comparison of measured and computed root ap moments. For the three highest wind speeds the deviations from measurements are less than the standard deviation
1600 1400 1200 Measured Comp. Free Comp. Tunnel

Root Edge Moment [Nm]

1000 800 600 400 200 0 200 400 600 6 8 10 12 14 16 18 20 22 24 26

Wind Speed [m/s]


Figure 5. Comparison of measured and computed root edge moments. The large standard deviation for the measurements is caused by the inuence of gravity during one rotation

underpredict the shaft torque by approximately the same amount. The underprediction at high wind speeds does not agree with previous ndings for the LM17.0 and LM19.1 blades, where NavierStokes solvers are know to overpredict during stalled conditions.21 One reason for this difference may be that the present rotor is equipped with completely different aerofoil sections compared to the other rotors. A recent study has shown that the aerofoil shape, and thereby the ow mechanisms controlling the stalling behaviour, may have a large impact on the quality of the CFD solution.22,23 When comparing computed and measured root ap moments,
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

158
(a)

N. N. Srensen, J. A. Michelsen and S. Schreck

1 0.9 0.8 S0700000

CN

0.7 0.6 0.5 0.4 0.2 1.6 1.4 S1000000

(b)

0.3

0.4

0.5

0.6

0.7

0.8

0.9

CN

1.2 1 0.8 0.6 0.2 2.4 2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 S1300000 1

(c)

CN

1.6 1.2 0.8 0.4 0.2 2.8 2.4 2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 S1500000 1

(d)

CN

1.6 1.2 0.8 0.4 0.2 2.8 2.4 2 S2000000

(e)

0.3

0.4

0.5

0.6

0.7

0.8

0.9

CN

1.6 1.2 0.8 0.4 0.2 2.8 2.4 2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 S2500000 1

(f)

CN

1.6 1.2 0.8 0.4 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

r/R

Figure 6. Spanwise distribution of normal force coefcients for the six cases; for the measurements, one standard deviations are shown

Copyright 2002 John Wiley & Sons, Ltd.

Wind Energ. 2002; 5:151169

NavierStokes Predictions
(a)

159

0.14 0.12 0.1 S0700000

CT

0.08 0.06 0.04 0.02 0.2 0.25 0.2 0.15 S1000000

(b)

0.3

0.4

0.5

0.6

0.7

0.8

0.9

CT

0.1 0.05 0

-0.05 0.2 (c) 0.3 0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

S1300000

CT

0.1 0

-0.1 0.2 0.2 0.16 0.12 0.08 0.04 0 -0.04 -0.08 0.2 (e) 0.04 (d)

0.3

0.4

0.5

0.6

0.7

0.8

0.9

S1500000

CT

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0.02 0

S2000000

CT

-0.02 -0.04 -0.06 -0.08 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(f)

0.04 0.02 0 -0.02 -0.04 -0.06 -0.08 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 S2500000

CT

r/R

Figure 7. Spanwise distribution of tangential force coefcients for the six cases; for the measurements, one standard deviations are shown

Copyright 2002 John Wiley & Sons, Ltd.

Wind Energ. 2002; 5:151169

160

N. N. Srensen, J. A. Michelsen and S. Schreck

good qualitative agreement is again found, as shown in Figure 4. For the three highest wind speeds (151, 201 and 251 m s 1 ) the deviations from measured data are less than the standard deviations. For the two lowest wind speeds the computed values are approximately 16% overpredicted. Comparing the measured and computed root edge moments, good qualitative agreement is again found, as shown in Figure 5. Here the computed values are as much as 50% low compared to measurements. The large standard deviation for the measurements is mainly caused by the inuence of gravity during one rotation. The differences between the results from the free and tunnel congurations are very small for all three integrated quantities and most pronounced in all cases for the 10 m s 1 wind speed, where separated ow starts to cover the blade.

Blade Forces
Turning to the spanwise distribution of the normal and tangential force coefcients, good agreement is generally found, as shown in Figures 6 and 7. Generally, the computed sectional force coefcients fall within one standard deviation of the measurements. The main exception is the S100000 run (10 m s 1 ), where large deviations are seen on the inboard part of the rotor, especially in Figure 7. Looking at the development of the tangential force components for increasing wind speed, the following explanation is suggested. When the wind speed reaches approximately 10 m s 1 , the separation point is suddenly moved to the leading edge near the r/R D 047 section. As the wind speed increases further, the stalled area spreads, as seen by the widening of the dip of the tangential force coefcient in Figure 7. In connection with the computations there can be several reasons for not capturing this behaviour exactly. First, there is the generally accepted problem of accuracy of most turbulence models in highly separated ows. Second, the trends observed in the measurements may be connected to the presence of a laminar separation at the leading edge, a phenomenon that will require the computations to include laminar-to-turbulent transition. Previously, we have suggested that the tunnel blockage effect may also inuence this,24 and this is indeed seen in Figure 7. Even though a positive effect is seen on the tangential force component in Figure 7, a degradation is seen for the normal force in Figure 6. Close examination of the 10 m s 1 case has led us to
(a)

(b)

(c)

Figure 8. Limiting streamlines on the suction side of the blade. The cases shown are, from the top, 7, 10 and 20 m s 1 . The ve vertical lines indicates the location of the 30%, 47%, 63%, 80% and 95% sections
Copyright 2002 John Wiley & Sons, Ltd.

Wind Energ. 2002; 5:151169

NavierStokes Predictions
(a) 3 2.5 2 1.5

161
r/R = 0.30
(b) 3.5 3 2.5 2 1.5 1 0.5 0 -0.5 -1 -1.5

r/R = 0.47

-Cp

0.5 0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

-Cp

0.2

0.4

0.6

0.8

X/Chord
(c)
3 2.5 2 1.5 1 1 0.5 0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1 0.5 0 -0.5 -1 -1.5 0 0.2

X/Chord
(d)
2 1.5

r/R = 0.63

r/R = 0.80

-Cp

-Cp

0.4

0.6

0.8

X/Chord
(e)
1.5 1 0.5

X/Chord r/R = 0.95

-Cp

0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

X/Chord

Figure 9. Pressure distributions for the 7 m s

case: circles, measurements; full curves, comp. free; dotted curves, comp. tunnel

believe that the tunnel blockage effect is not the main parameter. In fact, the tunnel blockage effect is very small. Instead, the fact that the ow is extremely unstable for this specic wind speed right at the onset of massive stall is the key issue. This fact makes the ow sensitive to the smallest disturbance, such as the small change in wind speed from the inclusion of the tunnel walls in the computation.

Pressure Distributions
In the following, comparisons of measured and computed pressure distributions will be shown. In the experiment, pressure distributions for ve spanwise sections are measured at r/R D 030, 047, 063, 080 and
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

162
(a) 9 8 7 6 5 4 3 2 1 0 -1 -2

N. N. Srensen, J. A. Michelsen and S. Schreck

r/R = 0.30

(b) 8 7 6 5 4 3 2 1 0 -1

r/R = 0.47

-Cp

-Cp

0.2

0.4

0.6

0.8

0.2

0.4

0.6

0.8

X/Chord
(c) 7 6 5 4 3 2 1 0 -1 -2

X/Chord
(d) 6 5 4 3

r/R = 0.63

r/R = 0.80

-Cp

-Cp

2 1 0 -1 -2

0.2

0.4

0.6

0.8

0.2

0.4

0.6

0.8

X/Chord
(e) 3 2.5 2 1.5 1 0.5 0 -0.5 -1 -1.5

X/Chord r/R = 0.95

-Cp

0.2

0.4

0.6

0.8

X/Chord
Figure 10. Pressure distributions for the 10 m s
1

case: circles, measurements; full curves, comp. free; dotted curves, comp. tunnel

095. In connection with the evaluation of the blind comparison a distinct deviation at the 95% section was observed for the highest wind speeds. In the blind comparison the EllipSys3D computations were performed using an approximate tip shape, and this was suggested as a possible explanation of the deviations near the tip. For the computations of the tunnel conguration the actual tip geometry was used, as kindly provided by E. P. N Duque from Northern Arizona University. The denition used for the pressure coefcient in the following is P1 P0 Cp D 1 C W2 C r 2 1 2
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

NavierStokes Predictions

163

For the lowest computed wind speed of 70 m s 1 , good agreement is found for all ve spanwise sections, as shown in Figure 9. Referring to Figure 8, showing the limiting streamlines on the suction side of the blade, only the innermost station closest to the root experiences stalled ow conditions at this wind speed. From the previous discussion of the integrated quantities and the spanwise force distributions it was observed that the 10 m s 1 case is poorly predicted. Looking at Figure 10, it is obvious that the measurements show a leading edge separation at the r/R D 047 section, while the computations preserve a sharp suction peak. A closer look at Figure 10 shows that the computation of the tunnel conguration has a slightly less pronounced
(a)

3.5 3 2.5 2 1.5 1 0.5 0 -0.5 -1 -1.5

r/R = 0.30

(b) 2 1.5 1

r/R = 0.47

-Cp

-Cp
0 0.2 0.4 0.6 0.8 1

0.5 0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

X/Chord
(c) 2 1.5 1

X/Chord
(d)

r/R = 0.63

r/R = 0.80
2.5 2 1.5 1

-Cp

0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

-Cp

0.5 0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

X/Chord
(e) 6 5 4

X/Chord r/R = 0.95

-Cp

3 2 1 0 -1 0 0.2 0.4 0.6 0.8 1

X/Chord
Figure 11. Pressure distributions for the 15 m s
1

case: circles, measurements; full curves, comp. free; dotted curves, comp. tunnel

Copyright 2002 John Wiley & Sons, Ltd.

Wind Energ. 2002; 5:151169

164

N. N. Srensen, J. A. Michelsen and S. Schreck

(a)

3 2.5 2 1.5

r/R = 0.30

(b)

2 1.5 1

r/R = 0.47

-Cp

0.5 0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

-Cp

0.5 0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

X/Chord
(c) 1.5 1 0.5

X/Chord
(d) 1.5 1 0.5

r/R = 0.63

r/R = 0.80

-Cp

-Cp

0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

X/Chord
(e) 0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 -1 0 0.2 0.4 0.6 0.8 1

X/Chord r/R = 0.95

-Cp

X/Chord
Figure 12. Pressure distributions for the 20 m s
1

case: circles, measurements; full curves, comp. free; dotted curves, comp. tunnel

suction peak and therefore is slightly closer to the measured conditions. This is related to the previously mentioned unstable ow conditions present near the r/R D 047 section. Here small spanwise shifts in the position of the separation near the leading edge may cause signicant changes to the pressure distribution, as shown in Figure 8. For the higher wind speeds (15, 20 and 25 m s 1 ) the computed results show surprisingly good agreement with the measured results, as shown in Figures 1113. The stall behaviour of the S809 aerofoil is believed to be one explanation for the surprisingly good agreement for the high-wind-speed cases.
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

NavierStokes Predictions
r/R = 0.30 r/R = 0.47

165

(a)

2.5 2 1.5 1

(b)

2 1.5 1

-Cp

0.5 0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

-Cp

0.5 0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

X/Chord
(c)

X/Chord
(d)

1.5
1

r/R = 0.63

1.5 1 0.5

r/R = 0.80

0.5

-Cp

-Cp

0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

0 -0.5 -1 -1.5 0 0.2 0.4 0.6 0.8 1

X/Chord
(e)

X/Chord r/R = 0.95

0.6 0.4 0.2 0

-Cp

-0.2 -0.4 -0.6 -0.8 -1 0 0.2 0.4 0.6 0.8 1

X/Chord

Figure 13. Pressure distributions for the 25 m s

case: circles, measurements; full curves, comp. free; dotted curves, comp. tunnel

Recent investigation of the capability of the 2D NavierStokes solver EllipSys2D to model several aerofoils has shown that the S809 belongs to a group of aerofoils where acceptable agreement between experiment and computation is found even at high angles of attack.22 Additionally, computations using the DES technique have indicated that resolving the vortex interaction in the wake at high angles of attack for the S809 aerofoil has a limited effect on the computed loads compared to other investigated aerofoils.25 It is believed that the Reynolds-averaged turbulence models perform well in 2D, and the limited dependence of the aerofoil loads on the 3D wake structure is mainly responsible for the quality of the rotor computations during heavily stalled conditions.
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

166
(a) 2 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 10 20 30 40 50 60

N. N. Srensen, J. A. Michelsen and S. Schreck


r/R = 0.30
2D 3D (b) 1.2 1.1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0 10 20 30 40 50 60

r/R = 0.47
2D 3D

Cl

AOA [deg]
(c) 1.2 1.1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0 10 20 30 40 50 60

Cl

AOA [deg]
(d)

r/R = 0.63
2D 3D

1.2 1.1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0 10 20

r/R = 0.80
2D 3D

Cl

Cl

30

40

50

60

AOA [deg]
(e) 1.2 1.1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0 10 20 30 40 50

AOA [deg] r/R = 0.95


2D 3D

Cl

60

AOA [deg]

Figure 14. Comparison of the computed lift for the ve spanwise sections with 2D computed values

The deviation at the r/R D 095 section for the 20 m s 1 case is believed to be caused by an inaccurate representation of the extremely complicated separated ow at the tip. The computations of the tunnel conguration have the correct tip geometry and show slightly better agreement at the tip. However, as seen from the surface streamlines in Figure 8, the ow in the tip region is extremely complicated and dominated by three-dimensional effects from the tip vicinity.

3D Effects
To investigate the changes to the 2D aerofoil characteristics when the aerofoils are used on rotating blades, lift and drag curves were extracted and compared to computed 2D curves. The method used to extract the
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

NavierStokes Predictions
(a)

167
r/R = 0.30
2D 3D

1.6 1.4 1.2 1

(b) 1.2 1 0.8

r/R = 0.47
2D 3D

Cd

Cd
0 10 20 30 40 50 60

0.8 0.6 0.4 0.2 0

0.6 0.4 0.2 0 0 10 20 30 40 50 60

AOA [deg]
(c) 1.2 1 0.8

AOA [deg]
(d) 1.2 1 0.8

r/R = 0.63
2D 3D

r/R = 0.80
2D 3D

Cd

0.6 0.4 0.2 0 0 10 20 30 40 50 60

Cd

0.6 0.4 0.2 0 0 10 20 30 40 50 60

AOA [deg]
(e) 1.2 1 0.8

AOA [deg] r/R = 0.95


2D 3D

Cd

0.6 0.4 0.2 0 -0.2 0 10 20 30 40 50 60

AOA [deg]

Figure 15. Comparison of the computed drag for the ve spanwise sections with 2D computed values

lift and drag is based on matching the stagnation point location between 2D and 3D CFD computations, both having been computed using the EllipSys code, as described in Reference 26. In Figures 14 and 15 the computed 3D lift and drag curves are compared with the 2D curves at the ve spanwise sections. From the lift curves it is evident that there is large increase in lift for the inboard stations for high angles of attack. This is in good qualitative agreement with theory, where the increased lift is explained by radial pumping of the ow in the separated area. When approaching the tip (r/R D 080 and 095), the ow is inuenced by the proximity of the blade tip, having the opposite effect on the lift. This is especially observed for the 95% section, where the lift is markedly reduced compared to 2D values. Next, looking at the drag, it is observed that the increase in lift for the inboard stations is accompanied by increased drag also, as shown in Figure 15.
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

168

N. N. Srensen, J. A. Michelsen and S. Schreck

Conclusions
A series of computations of the NREL Phase VI rotor as tested in the NASA Ames wind tunnel during upwind operation at 0 yaw and 3 tip pitch have been made. Generally, the computed results show good qualitative and quantitative agreement with the measurements, the only exception being the 10 m s 1 wind speed case. The large deviation seen in the 10 s 1 case is believed to be connected to the very unstable ow behaviour in this wind speed range. As the onset of massive ow separation on the blade is taking place for wind speeds in the region of 10 m s 1 , the prediction is very sensitive to very small variations. Changes in the wind speed, inclusion of laminar/turbulent transition, and the well-known problem of all turbulence models to capture the exact onset of stall will be extremely important in this connection. It is also shown that the 3D CFD computations can be used to determine the 3D aerodynamic effects present on wind turbine rotors. For the lowest wind speed, below onset of stall, the present CFD predictions accurately predict the low-speed shaft torque, while an overprediction of nearly 25% is observed at the onset of stall. For deep stall conditions an underprediction of 25% is observed. Additionally, the NREL/NWTC aerodynamics blind comparison has shown that CFD computations are competitive with BEM-type computations for new and unknown rotors. Here the fact that CFD or NavierStokes computations do not rely on empirical input has proven to have distinct advantages. Additionally, the NREL/NASA Ames wind tunnel experiment has proven to be extremely valuable in connection with code development and in the process of understanding the ow physics of modern wind turbine rotors.

Acknowledgements
This work was carried out under a contract with the Danish Energy Agency, ENS 1363/00-0007, Program for research in aeroelasticity 20002001. Computations were made possible by the use of the IBM RS/6000 SP at the Ris central computing facility.

References
1. Duque EPN, van Dam CP, Hughes SC. NavierStokes simulations of the NREL combined experiment phase II rotor. AIAA Paper 99-0037, 1999. 2. Xu G, Sankar LN. Computational study of horizontal axis wind turbines. AIAA Paper 99-0042, 1999. 3. Xu G, Sankar LN. Effects of transition, turbulence and yaw on the performance of horizontal axis wind turbines. AIAA Paper 2000-0048, 2000. 4. Srensen NN, Hansen MOL. Rotor performance predictions using a NavierStokes method. AIAA Paper 98-0025, 1998. 5. Srensen NN, Michelsen JA. Aerodynamic predictions for the unsteady aerodynamics experiment Phase-II rotor at the National Renewable Energy Laboratory. AIAA Paper 2000-0037, 2000. 6. Schepers JG, Brand AJ, Bruining A, Graham JMR, Hand MM, Ineld DG, Madsen HA, Paynter RJH, Simms DA. Final report of IEA Annex XIV: eld rotor aerodynamics. Technical Report ECN-C-97-027, Netherlands Energy Research Foundation, 1997. 7. Fingersh L, Simms D, Hand M, Jager D, Cortrell J, Robinson M, Schreck S, Larwood S. Wind tunnel testing of NRELs unsteady aerodynamics experiment. AIAA 2001-0035, 2001. 8. Simms D, Schreck S, Hand M, Fingersh LJ. NREL unsteady aerodynamics experiment in the NASA-Ames wind tunnel: a comparison of predictions to measurements. NREL/TP-500-29494, NREL, Golden, CO, 2001. 9. Giguere P, Selig MS. Design of a tapered and twisted blade for the NREL combined experiment rotor. NREL/SR500-26173, NREL, Golden, CO, 1999. 10. Hand M, Simms D, Fingersh LJ, Jager D, Larwood S, Cotrell J, Schreck S. Unsteady aerodynamics experiment Phase VI: wind tunnel test congurations and available data campaigns. NREL/TP-500-29955, NREL, Golden, CO, 2001. 11. Michelsen JA. Basis3Da platform for development of multiblock PDE solvers. Technical Report AFM 92-05, Technical University of Denmark, Lyngby, 1992. 12. Michelsen JA. Block structured multigrid solution of 2D and 3D elliptic PDEs. Technical Report AFM 94-06, Technical University of Denmark, Lyngby, 1994. 13. Srensen NN. General purpose ow solver applied to ow over hills. Ris-R-827(EN), Ris National Laboratory, Roskilde, 1995.
Copyright 2002 John Wiley & Sons, Ltd. Wind Energ. 2002; 5:151169

NavierStokes Predictions

169

14. Rhie CM. A numerical study of the ow past an isolated airfoil with separation. PhD Thesis, University of Illinois, Urbana-Champaign, IL, 1981. 15. Patankar SV, Spalding DB. A calculation procedure for heat, mass and momentum transfer in three-dimensional parabolic ows. International Journal of Heat and Mass Transfer 1972; 15: 17871806. 16. Michelsen JA. General curvilinear transformation of the NavierStokes equations in a 3D polar rotating frame. Technical Report ET-AFM 98-01, Technical University of Denmark, Lyngby, 1998. 17. Khosla PK, Rubin SG. A diagonally dominant second-order accurate implicit scheme. Computers and Fluids 1974; 2: 207209. 18. Menter FR. Zonal two equation k turbulence models for aerodynamic ows. AIAA Paper 93-2906, 1993. 19. Wilcox DC. A half century historical review of the k model. AIAA Paper 91-0615, 1991. 20. Menter FR. Performance of popular turbulence models for attached and separated adverse pressure gradient ows. AIAA Journal 1992; 30: 20662072. 21. Chaviaropoulos PK, Nikolaou IG, Aggelis K, Srensen NN, Montgomerie B, von Geyer G, Hansen MOL, Kang S, Voutsinas S, Dyrmose SZ. Viscous and aeroelastic effects on wind turbine blades: the VISCEL project. European Wind Energy Conference and Exhibition, Copenhagen, July 2001; 347350. 22. Bertagnolio F, Srensen NN, Johansen J, Fuglsang P. Wind turbine airfoil catalogue. Ris-R-1280(EN), Ris National Laboratory, Roskilde, 2001. 23. Bertagnolio F, Srensen NN, Johansen J, Fuglsang P. Conclusion from the comparison of numerous 2D airfoil computations with experiments. AIAA Paper 2002-0034, 2002. 24. Srensen NN, Michelsen JA, Schreck S. Detailed aerodynamic prediction of the NREL/NASA Ames wind tunnel tests using CFD. European Wind Energy Conference and Exhibition, Copenhagen, July 2001; 4853. 25. Johansen J, Srensen NN, Michelsen JA, Schreck S. Detached-eddy simulation of ow around the S809 airfoil. European Wind Energy Conference and Exhibition, Copenhagen, July 2001; 414417. 26. Srensen NN. Evaluation of 3D effects from 3D CFD computations. IEA Joint Action. Aerodynamics of Wind Turbines. 14th Symposium, Boulder, CO, December 2000; 1122.

Copyright 2002 John Wiley & Sons, Ltd.

Wind Energ. 2002; 5:151169

You might also like