arxiv: v1 [physics.flu-dyn] 13 Sep 2017

Similar documents
VORTEX METHOD APPLICATION FOR AERODYNAMIC LOADS ON ROTOR BLADES

Validation of the actuator line and disc techniques using the New MEXICO measurements

Actuator Surface Model for Wind Turbine Flow Computations

Validation of Chaviaro Poulos and Hansen Stall Delay Model in the Case of Horizontal Axis Wind Turbine Operating in Yaw Conditions

Experimental and numerical investigation of 3D aerofoil characteristics on a MW wind turbine

CFD RANS analysis of the rotational effects on the boundary layer of wind turbine blades

Aerodynamic Rotor Model for Unsteady Flow and Wake Impact

Insight into Rotational Effects on a Wind Turbine Blade Using Navier Stokes Computations

Calculation of Wind Turbine Geometrical Angles Using Unsteady Blade Element Momentum (BEM)

Analysis of Counter-Rotating Wind Turbines

Analysis of the high Reynolds number 2D tests on a wind turbine airfoil performed at two different wind tunnels

CHAPTER 4 OPTIMIZATION OF COEFFICIENT OF LIFT, DRAG AND POWER - AN ITERATIVE APPROACH

1 Introduction. EU projects in German Dutch Wind Tunnel, DNW. 1.1 DATA project. 1.2 MEXICO project. J.G. Schepers

Dynamic Stall Modeling for Wind Turbines. M. A. Khan

ANALYSIS OF HORIZONTAL AXIS WIND TURBINES WITH LIFTING LINE THEORY

Final Results from Mexnext-I: Analysis of detailed aerodynamic measurements on a 4.5 m diameter rotor placed in the large German Dutch Wind Tunnel DNW

Actuator disk modeling of the Mexico rotor with OpenFOAM

Numerical Study on Performance of Curved Wind Turbine Blade for Loads Reduction

aerodynamic models for wind turbine design codes can be improved, developed and validated. Although the emphasis of the IEA Annex XIV/XVIII activities

Adjustment of k ω SST turbulence model for an improved prediction of stalls on wind turbine blades

Wind turbine power curve prediction with consideration of rotational augmentation effects

Some effects of large blade deflections on aeroelastic stability

Numerical Study on Performance of Innovative Wind Turbine Blade for Load Reduction

Results of the AVATAR project for the validation of 2D aerodynamic models with experimental data of the DU95W180 airfoil with unsteady flap

Aerodynamic Performance 1. Figure 1: Flowfield of a Wind Turbine and Actuator disc. Table 1: Properties of the actuator disk.

Analysis of Wind Turbine Pressure Distribution and 3D Flows Visualization on Rotating Condition

Assessment of low-order theories for analysis and design of shrouded wind turbines using CFD

Aeroacoustic calculations of a full scale Nordtank 500kW wind turbine

Numerical Simulations of Wakes of Wind Turbines Operating in Sheared and Turbulent Inflow

Mechanical Engineering for Renewable Energy Systems. Dr. Digby Symons. Wind Turbine Blade Design

Effect of Geometric Uncertainties on the Aerodynamic Characteristic of Offshore Wind Turbine Blades

Research in Aeroelasticity EFP-2005

Prediction of airfoil performance at high Reynolds numbers.

Large-eddy simulations for wind turbine blade: rotational augmentation and dynamic stall

Numerical Investigation of Flow Control Feasibility with a Trailing Edge Flap

Research on Dynamic Stall and Aerodynamic Characteristics of Wind Turbine 3D Rotational Blade

MODELLING OF ROTATIONAL AUGMENTATION BASED ON ENGINEERING CONSIDERATIONS AND MEASUREMENTS

A Short History of (Wind Turbine) Aerodynamics

arxiv: v1 [physics.flu-dyn] 11 Oct 2012

A Hybrid CFD/BEM Analysis of Flow Field around Wind Turbines

Validation of a vortex ring wake model suited for aeroelastic simulations of floating wind turbines

Iterative Learning Control for Smart Rotors in Wind turbines First Results

The Pennsylvania State University. The Graduate School. College of Engineering. Inviscid Wind-Turbine Analysis Using Distributed Vorticity Elements

Indicial lift response function: an empirical relation for finitethickness airfoils, and effects on aeroelastic simulations

Comparison of Blade Element Method and CFD Simulations of a 10 MW Wind Turbine

Mechanical Engineering for Renewable Energy Systems. Wind Turbines

Lecture 4: Wind energy

An Insight into the Rotational Augmentation on HAWTs by means of CFD Simulations Part I: State of the Art and Numerical Results

Comparison of aerodynamic models for Vertical Axis Wind Turbines

Modelling of Vortex-Induced Loading on a Single- Blade Installation Setup

Analysis of the effect of curtailment on power and fatigue loads of two aligned wind turbines using an actuator disc approach

Assessment of aerodynamics models for wind turbines aeroelasticity

Manhar Dhanak Florida Atlantic University Graduate Student: Zaqie Reza

Unsteady Analysis of a Horizontal Axis Marine Current Turbine in Yawed Inflow Conditions With a Panel Method

Steady State Comparisons HAWC2 v12.2 vs HAWCStab2 v2.12

Sensitivity of Key Parameters in Aerodynamic Wind Turbine Rotor Design on Power and Energy Performance

A Computational Aerodynamics Simulation of the NREL Phase II Rotor

Airfoil data sensitivity analysis for actuator disc simulations used in wind turbine applications

Determining Diffuser Augmented Wind Turbine performance using a combined CFD/BEM method

Aerodynamic force analysis in high Reynolds number flows by Lamb vector integration

Advanced Load Alleviation for Wind Turbines using Adaptive Trailing Edge Flaps: Sensoring and Control. Risø-PhD-Report

INCLUSION OF A SIMPLE DYNAMIC INFLOW MODEL IN THE BLADE ELEMENT MOMENTUM THEORY FOR WIND TURBINE APPLICATION

Aeroelastic Stability of Suspension Bridges using CFD

The CFD Investigation of Two Non-Aligned Turbines Using Actuator Disk Model and Overset Grids

Unsteady Viscous-Inviscid Interaction Technique for Wind Turbine Airfoils. PhD Thesis

A simplified model for a small propeller with different airfoils along the blade

Blade Element Momentum Theory

Effects of ambient turbulence on the near wake of a wind turbine

AERODYNAMIC ANALYSIS OF THE HELICOPTER ROTOR USING THE TIME-DOMAIN PANEL METHOD

APPENDIX A. CONVENTIONS, REFERENCE SYSTEMS AND NOTATIONS

Keywords: Contoured side-walls, design, experimental, laminar boundary layer, numerical, receptivity, stability, swept wing, wind tunnel.

Research on Propeller Characteristics of Tip Induced Loss

3D hot-wire measurements of a wind turbine wake

A New Implementation of Vortex Lattice Method Applied to the Hydrodynamic Performance of the Propeller-Rudder

STRUCTURAL PITCH FOR A PITCH-TO-VANE CONTROLLED WIND TURBINE ROTOR

UNSTEADY AERODYNAMIC ANALYSIS OF HELICOPTER ROTOR BY USING THE TIME-DOMAIN PANEL METHOD

Aerodynamic Performance Assessment of Wind Turbine Composite Blades Using Corrected Blade Element Momentum Method

Near-Wake Flow Simulations for a Mid-Sized Rim Driven Wind Turbine

Aeroelastic modelling of vertical axis wind turbines

Large eddy simulations of the flow past wind turbines: actuator line and disk modeling

Lecture 7 Boundary Layer

Performance and Equivalent Loads of Wind Turbines in Large Wind Farms.

Aerodynamics. High-Lift Devices

INSTITUTE OF AERONAUTICAL ENGINEERING (Autonomous) Dundigal, Hyderabad

A Numerical Blade Element Approach to Estimating Propeller Flowfields

Wind Turbine Blade Analysis using the Blade Element Momentum Method. Version 1.0

Performance and wake development behind two in-line and offset model wind turbines "Blind test" experiments and calculations

High Speed Aerodynamics. Copyright 2009 Narayanan Komerath

PEMP ACD2505. M.S. Ramaiah School of Advanced Studies, Bengaluru

1. Fluid Dynamics Around Airfoils

Numerical Investigation of Aerodynamic Performance and Loads of a Novel Dual Rotor Wind Turbine

θ α W Description of aero.m

Aeroelasticity in Dynamically Pitching Wind Turbine Airfoils

Fan Stage Broadband Noise Benchmarking Programme

Effects of the Leakage Flow Tangential Velocity in Shrouded Axial Compressor Cascades *

Validation and comparison of aerodynamic modelling approaches for wind turbines

Three dimensional actuator disc modelling of wind farm wake interaction

Kinetic energy entrainment in wind turbine and actuator disc wakes: an experimental analysis

Fluid Dynamics: Theory, Computation, and Numerical Simulation Second Edition

DETERMINATION OF HORIZONTAL AXIS WIND TURBINE PERFORMANCE IN YAW BY USE OF SIMPLIFIED VORTEX THEORY

Transcription:

arxiv:1709.04298v1 [physics.flu-dyn] 13 Sep 2017 Evaluation of different methods for determining the angle of attack on wind turbine blades with CFD results under axial inflow conditions H. Rahimi 1,2, J.G Schepers 3, W.Z Shen 4, N. Ramos García 4, M.S. Schneider 5, D. Micallef 6, C.J. Simao Ferreira 7, E. Jost 8, L. Klein 8, I. Herráez 9 1 ForWind, Institute of Physics, University of Oldenburg, 26129 Oldenburg, Germany 2 Fraunhofer IWES, Küpkersweg.70, 26129 Oldenburg, Germany 3 ECN Wind Energy Technology, Petten, Netherlands 4 Department of Wind Energy, Technical University of Denmark, Lyngby, Denmark 5 German Aerospace Center (DLR), Institute of Aeroelasticity, Göttingen, Germany 6 Department of Environmental Design, University of Malta, Msida, Malta 7 Department of Wind Energy, Delft University of Technology, Delft, Netherlands 8 Institute of Aerodynamics and Gasdynamics, University of Stuttgart, Stuttgart, Germany 9 Faculty of Technology, University of Applied Sciences Emden/Leer, Emden, Germany E-mail: hamid.rahimi@uni-oldenburg.de, schepers@ecn.nl Abstract. This work presents an investigation on different methods for the calculation of the angle of attack and the underlying induced velocity on wind turbine blades using data obtained from three dimensional computational fluid dynamics. Several methods were examined and their advantages as well as shortcomings were presented. The investigations were performed for flows past two 10MW reference wind turbines under axial inflow conditions, namely the 10MW turbines designed in the EU AVATAR and INNWIND.EU projects. The results shown that the evaluated methods are in good agreement with each other at the mid-span, though some deviations were observed at the root and tip regions of the blades. Using any of the proposed methods, it is possible to obtain airfoil characteristics for lift and drag coefficients as a function of the angle of attack. Furthermore, by using the extracted data new correction models can be derived for blade element momentum calculations. Nomenclature AoA Angle of Attack [ ] θ Twist angle [ ] φ Inflow angle [ ] a Axial induction factor at each section [-] r Radial coordinate [m] R Rotor radius [m] Ω Rotational speed [rad/s] U Free stream wind speed at hub height [m s 1 ]

V rel Relative velocity at each section [m s 1 ] U ind Induced velocity at each section [m s 1 ] ρ Air density [kg m 3 ] C l Lift coefficient [-] C d Drag coefficient [-] Γ circulation [m 2 /s] CF D Computational Fluid Dynamics BEM Blade Element Momentum ABL Atmospheric boundary layer T SR Tip Speed Ratio MW Mega Watt 1. Introduction The traditional approach in wind turbine design codes used to simulate the aerodynamic behavior of wind turbines is the Blade Element Momentum (BEM) theory. The low computational cost of BEM simulations, which is the result of many simplifications (e.g. steady and two dimensional flow), makes this an affordable approach even for the more than 7 million time steps which wind turbine manufacturers needs to perform in the calculation of the design load spectrum for certification [1]. In the BEM theory, the aerodynamic forces are interpolated from sectional airfoil characteristics such as C l and C d as a function of the Angle of Attack (AoA). Hence, the outcome of a BEM code depends heavily on these airfoil characteristics and AoA which are derived from two Dimensional (2D) steady wind tunnel experiments, viscous-inviscid panel codes or Computational Fluid Dynamics (CFD) simulations [2]. In BEM, the assumption of 2D steady flow might be valid for the mid-span region and steady uniform inflow. Nonetheless, on a rotating blade, in a fluctuating 3D wind flow, which leads to a misalignment of the velocity vector with respect to the rotor plane and near the thick root section and the blade tip where the flow is highly complex and 3D, this assumption is not valid anymore. Therefore the complex 3D flows cannot be captured accurately by the 2D airfoil characteristics and basic BEM theory [3]. Hence empirical correction models are generally added to a BEM model to mitigate its lace of accuracy when specific phenomena are observed, as: 3D effects [4, 5, 6, 7], dynamic stall [8, 9] or skewed wake [1, 10, 11]. In the research community, there is a continuous effort for improving these engineering addons. An example is the EU project AVATAR [12], which aims at improving and validating aerodynamic models, and to ensure their applicability for designing 10MW+ turbines [13]. One could think of improving the accuracy of these correction models for the BEM codes from measurements using hot wires or Particle Image Velocimetry (PIV) [14, 15, 16, 17]. However, these measurements are very expensive in terms of both time and cost where they often limited (spatial and temporal) resolution. An alternative method would be the usage of high fidelity numerical tools like 3D CFD [18]. One of the major limitations for the improvement of these correction models using measurements or 3D CFD lies in the many uncertainties involved on the definition of AoA and underlying axial and tangential induced velocity for 3D flows. The AoA is a 2D steady concept defined as the angle between the inflow velocity at farfield and the airfoil chord, as shown in Figure 1. The AoA is not the angle measured with a probe/piv or calculated with CFD ahead of the airfoil at a monitor point. In reality, trailing vortices due to variations in blade circulation alter the AoA. The local inflow velocity is influenced by the axial induction resulting from the momentum equilibrium. As such, if the influence of the presence of the airfoil,

i.e. the influence of bound circulation is removed from any point of the flow field, the AoA is obtained, and thus the induced velocity. Figure 1: Schematic representation of the relative velocity V rel, axial velocity U (1 a), tangential velocity Ωr(1 + a ), AoA (α), inflow angle (φ) and twist angle (θ). Several methods have been introduced so far for calculating the AoA at the rotor in axial inflow conditions using 3D CFD simulations, such as inverse BEM [19], Average Azimuthal Technique (AAT) [3, 20] and the Shen models from [21, 22]. However, there are very few studies such as Guntur et al [23] examining all of these methods for different conditions. The goal of present work is therefore to examine all the available methods for the calculation of the AoA for axial inflow conditions without wind shear from 3D CFD results. Fir this purpose, a numerical simulation of the 10MW AVATAR [12] and the INNWIND.EU [24] turbines, by means of fully resolved CFD under uniform and axial inflow conditions are used. The freestream velocities considered for this study are 6 and 9 m/s with the rotor speeds of 6 and 7.3 rpm, respectively, and also no wind shear was considered. The operational conditions set for the turbines are the operating controlled conditions described in their reference manual. In Table 1 the main parameters of the considered turbine are listed. The CFD results used in this work and the pressure distributions around the blades have been calculated using the EllipSys3D CFD solver [25, 26, 27]. It should be noted that the present work serves also as a verification of several methods since main participants used the same methods so that the correct implementation could be checked. Moreover, this contribution could be regarded as a survey on the assessment of each method under uniform and axi-symmetric inflow conditions. Turbine parameter INNWIND.EU AVATAR Rated power (MW) 10 10 Rotor diameter (m) 178.3 205.8 Axial induction 0.3 0.24 Rotor speed (rpm) 9.6 9.6 Tip speed (m/s) 90 103.4 Hub height (m) 119 132.7 Table 1: Basic reference turbine characteristics for the INNWIND.EU and AVATAR rotors.

2. Methods Several techniques have been implemented for the calculation of the induction factor and AoA so far. In the following, these techniques are presented: Inverse BEM method: The computed or measured load distribution is used as an input to estimate the induction and AoA using the general BEM theory formulation [28, 29, 19]. This technique may be expected to give reasonable results in axial flow conditions. However due to the fact that BEM theory is one-dimensional, when it comes to flow separation or yawed conditions the accuracy of this model is questionable [1, 10, 11]. This also holds for the root and tip where the flow shows its 3D nature. In addition, just like in BEM, the simple momentum theory does not hold anymore when the axial induction factor becomes larger than approximately 0.4. Therefore, several empirical relations between the thrust coefficient and the induction factor have been proposed based on measurements in order to correct for this effect [30, 19] and it is essential to use such a correction in the inverse BEM, just like the BEM itself [31]. Measurements or CFD results are supposed to be used as validation material for BEM codes and to improve the uncertainties in BEM. Therefore it would be better if the inverse BEM method would not be a part of the validation materials. However, it is also interesting to look at the validity of the general BEM theory formulation at different conditions. Average Azimuthal method (AAT): This technique is based on the annular average values of the axial velocity by using the data at several upstream and downstream locations [3, 20]. The velocity field for an annulus upstream and downstream of the rotor plane at a given radial location are then averaged. Afterwards, the value of the velocity in the rotor plane is estimated by interpolating the upstream and downstream positions velocities. In Figure 2, a schematic representation of this method is presented. AAT provides an annular averaged induced velocity which is known to differ from the local induced velocity near the tip. The shortcoming of this method lies in the fact that this model is only valid for axial conditions (or for the mean value of axial induction factor over one rotation) and it can not capture the dynamics of the induced velocity of yawed flows. Furthermore, since the method is based on averaged data, many input points have to be sampled, and, therefore the computational cost might be high. It is also noted that the results might depend on the positions of the monitor point and the interpolation algorithm. Figure 2: Average Azimuthal Technique (AAT) for the calculation of axial induction factor. 3-Point method: This method first introduced in Ref. [10] uses only three points along the chord length on each side. This great simplicity is the main advantage of this method, which makes the calculation of the AoA very straightforward. Moreover by using three

points unllike the AAT method, this method is able to reproduce the dynamic behavior of the induction for each azimuthal positions, e.g at yawed flow. In addition, the local induced velocity near the tip and root of the blade can be reproduced and finally, by choosing three points at each section, the effect of upwash and downwash on the averaged velocity can be reduced. In Figure 3, a schematic representation of this method is presented. For this method, 3 points on each side of a particular section which is modeled as an airfoil will be assigned. These points are located at 25, 50 and 75 % of chord length along the airfoil. Then, each pair of points (P1,P4), (P2,P5) and (P3, P6) are averaged independently to get the three velocity at the airfoil section. Finally, the resulting velocity points are averaged to get the estimated velocity which is induced at the blade section (red point). Figure 3: Schematic representation of the 3-Point method for the calculation of the axial induction factor. Shen 1 method: The technique presented in [21] is a simple way to determine the AoA on a rotor blade. As input to this technique, the force distribution along the blade (typically projected along the chordwise and normal to the chordwise directions) and the velocity at a set of monitor points in the vicinity of the blade is assumed to be known from CFD computations or experiments. The AoA is determined by (1) estimating the lift force by projecting the force along the incoming and normal to the incoming directions; (2) calculating the bound vortex using the Kutta-Joukowski law; (3) calculating the induced velocity by the bound vortex using Biot-Savart s law; (4) computing the relative velocity at the monitor points by subtracting the induced velocity from the bound vortex; (5) computing the AoA from the relative velocity. This procedure continues until the convergence is reached. The positions of the monitoring points are important for this method as the bound vortex is considered as a line vortex along the blade. The points cannot be chosen too close to the bound vortex since this is a singularity where the induced velocity approaches infinite. Shen 2 method: In order to overcome the difficulty of singularity in the previous method, an alternative technique was presented in [22], where a distributed bound circulation along the airfoil/blade surface is used instead of the concentrated bound vortex at the force center. In this case, the monitor points can be chosen closer to the blade. Another advantage is that this method takes the chordwise variation of aerodynamic forces into account which is neglected when the vorticity is concentrated in a bound vortex. Additionally, this method is not iterative. However, the difficulty of using this method is to find the separation point (SP) where the local circulation changes sign [22]. Ferreira-Micallef method: Assuming that the flow around a blade section is 2D, incompressible and irrotational, then the velocity at any point near the blade section can be decomposed in the linear addition

Vorticity points calculated at the center of each grid point Control point locus Figure 4: Schematic of free vortex points distributed at each grid point. Figure 5: Schematic of the point vortices bound to the blade s surface and locus of the control points for velocity evaluation. of three elements; (i) induced velocity by the free vorticity in the flow, (ii) induced velocity by the vorticity bound to the surface of the airfoil, (iii) local uniform wind velocity (the result of U and the induction of all other elements in the 3D domain). Of these three elements, only the distribution of free vortices can be calculated directly from the velocity/vorticity field. To determine the local wind velocity, a system of equations based on potential flow vortex theory is solved (see [32]) simultaneously, thus determining the local wind velocity and the strength of the vortices bound to the blade section surface. In this approach, the free vorticity and the vorticity bound to the surface of the airfoil are approximated by point vortices with a Rankine vortex distribution (see [32]), having a strength Γ ω = ω z x y. Figure 4 shows a schematic of the free vortex points distributed at each grid point. Bound vortex points of strength Γ bi are placed on the blade surface, where i is the index of the point vortex. The control points for the evaluation of the system of equations are collocated in the vicinity of the blade section, as represented in Figure 5; however, it is not necessary for the locus of the control points to be a closed contour around the blade section, as represented in Figure 5. The unknowns of the system of equations are the vortex strengths of each element and the velocities (U l, V l ) which are the components of the local wind velocity at the blade s quarter chord. This results in an over-constrained system which is solved using a least squares approach. Finally (U l, V l ) can then be used to determine the angle of attack. The relative inflow angle is thus given by: ( ) φ = tan 1 Vl (1) U l Line Average method: The Line average (LineAve) method determines the AoA by averaging the flow velocities along a symmetric, closed line around the rotor blade [33]. In the present study, a circle was chosen for this purpose as illustrated in Figure 6. The circle center is placed at the quarter chord position, where, like in a lifting line representation the bound vortex is located. The idea is that the induced velocity at opposed points on the circle extinguish each other. By averaging the flow velocities along the circle, the influence of bound circulation is eliminated and the local inflow velocity and AoA can be determined. Since it needs to be solely ensured that the shapes are symmetric to the quarter chord point, in theory a variety of closed shapes are possible. However, in earlier works [33] the

Z Y X vind,bound Γbound vind,bound vind,bound vind,bound Figure 6: Line average method - 2D and 3D [33]. circle shape provided convincing results. The local circle radius is varied along the blade span and chosen dependent on the local chord length c. In this study, rcirc = 4c is used. Herr aez Method: This method obtains the undisturbed flow rotor velocities by extracting them directly from a position in the rotor plane where the influence of the blade bound circulation from each blade is canceled out by the other blades [34]. In the case of axisymmetric, homogeneous inflow, this position corresponds to the bisectrix of the angle between two arbitrary blades. For a wind turbine with 3-blades, the undisturbed velocities can be obtained along the radial traverses located 60 ahead and behind an arbitrary blade, as shown in Figure 7. In the case of a 2-bladed rotor, the velocities should be probed at the radial traverses located 90 ahead and behind an arbitrary blade. The difference between the free stream velocity and the axial velocity component obtained in this way for each radial position corresponds to the local axial wake induction. The local tangential velocity extracted from the radial traverse corresponds to the local tangential wake induction after changing its sign. The main advantage of this method is its simplicity, which makes the calculation of the AoA very straightforward. Furthermore, in opposition to other methods (see section 3.1), it is not sensitive to input parameters like engineering correction models, monitoring point location, etc. The main limitation of the method is that its use for nonaxi-symmetric or inhomogeneous inflow becomes much more complicated because of the dependence of the blade bound circulation on the azimuthal blade position. 3. Results In this section, the results from the above mentioned methods for the calculation of the AoA and underlying induced velocity are presented for the INNWIND.EU and AVATAR turbines at two wind velocities of 6m/s and 9m/s. The selected cases endeavour to be representative for the results of both the turbines which have been investigated. 3.1. Sensitivity study on input parameters: Uind, AoA As explained in Section 2, different methods could have dependency on various input parameters e.g: the monitoring point location, turbulent wake correction or interpolation method etc. In this section, a sensitivity study on the input parameters for the AAT, Shen1 and Shen2 methods is presented. 3P and Line Average method also show similar behavior as the AAT method, therefore, they are not presented here. For the Ferreira-Micallef and Herr aez methods, no dependency from input-parameters was reported.

Figure 7: Three bladed rotor with axial and homogeneous inflow. The bound circulation distribution is the same for all blades. The red line represents a position where the influence of the bound circulation from all blades is cancelled out and the undisturbed velocity field can be probed. 3.1.1. AAT method: The AAT method is based on probing the velocity at certain locations, hence, the numerical results can be affected by the location of the monitoring points. In Figure 8, the dependency of U ind and the AoA on the monitor point location is presented. The monitoring points location is changed from 0.5 chords to 4 chords along the span. Although from Figure 8 it can be seen that at the very tip (r/r >0.85 ) and root (r/r <0.35 ) U ind is slightly dependent on the position of the monitoring point, the extracted AoA generally remains in a very close agreement. It should also be noted that in this study, the location of monitoring points depends on the chord length where the original AAT method assumes the position to be fixed along the entire span. Since the chord is strongly variable along the span, the chord dependent monitoring point location is believed to be a more meaningful choice. Moreover, the dependency of U ind and the AoA on the azimuthal resolution of the AAT method for the INNWIND.EU turbine at 9m/s is presented in Figure 9. It can be seen that when more than 80 points are used, U ind and AoA remain in a close agreement. Hence, it can be concluded that the azimuthal resolution has little effect on the results. The inner plot in Figure 8b represents AoA at the outer span location in a smaller range. It should be mentioned that the points are distributed in equidistant spacing in annulus upstream and downstream of the rotor plane. However concentrating the points near the blade did not present any differences.

U ind [m/s] 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 AOA [ ] 0 10 20 30 40 50 (a) U ind (b) AoA 6 8 10 0.5*C 2*C 3*C 4*C Figure 8: Dependency of U ind and AoA on the monitoring point location in the AAT method for the INNWIND.EU turbine at 9m/s. U ind [m/s] 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 AOA [ ] 0 10 20 30 40 50 (a) U ind (b) AoA 6 8 10 60 Points 80 Points 100 Points 120 Points Figure 9: Dependency of U ind and AoA on the azimuthal resolution in the AAT method for the INNWIND.EU turbine at 9m/s.

3.1.2. Shen methods: The influence of changing the position of the monitoring points for Shen 1 and Shen 2 is studied in this section and presented in Figures 10 and 11, respectively. The monitoring point distances to the blade leading-edge is changed from 1 to 4 chord lengths. The results indicate that the dependency on the monitoring point locations has limited effect from r/r =0.35 of rotor span onwards. However, substantial dependency is observed in both U ind and AoA at the most inner-part of the blade as the monitoring points near the root are too close to the own blade or other blades. In addition, the dependency on the monitoring points near the root are too location is higher for Shen 1 in comparison with the Shen 2 method. This is attributed to the fact that the method Shen1 monitors points located close to the blade that are influenced by the singular bound vortex. U ind [m/s] 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 AOA [ ] 0 10 20 30 40 50 (a) U ind (b) AoA 6 8 10 1*C 2*C 3*C 4*C Figure 10: Dependency of U ind and AoA on the monitoring point location in the Shen 1 method for the INNWIND.EU turbine at 9m/s.

U ind [m/s] 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 AOA [ ] 0 10 20 30 40 50 (a) U ind (b) AoA 6 8 10 1*C 2*C 3*C 4*C Figure 11: Dependency of U ind and AoA on the monitoring point location in the Shen 2 method for the INNWIND.EU turbine at 9m/s.

3.1.3. Inverse BEM methods: In the inverse BEM method, the blade loads and the momentum equations of the BEM method were used to derive the AoA. There are several implementations of this method available such as [28, 29, 19], which for axial inflow conditions mainly differ on the correction method of high induction and the tip and root corrections. In the momentum equations, when the axial induction factor becomes larger than approximately 0.4 (denoted as a critical ), the simple momentum theory breaks down. Hence, several empirical relations between the thrust coefficient and a have been made using measurements to correct this effect [30, 19]. In Figure 12, the result of two implementations for a critical is presented. As it can be seen in Figure 12, although a critical has a certain impact on U ind, the AoA remains in a close agreement due to the fact that the rotation velocity plays a major role for the derivation of the AoA. The dependency of the results on different tip and root correction models is not presented here but can be found in Ref [21]. U ind [m/s] 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 (a) U ind AOA [ ] 0 10 20 30 40 50 a critical = 0.23 a critical = 0.20 6 8 10 (b) AoA Figure 12: Dependency of U ind and AoA on different corrections for high values of induction factor for the INNWIND.EU turbine at 9m/s.

3.2. Comparison of all the methods: U ind, AoA, C l and C d In this section, the methods which were explained in section 2 are used to extract the AoA and also the normalized aerodynamic force coefficients (C l and C d ). The results presented in this section are for the INNWIND.EU and AVATAR wind turbines at the inflow velocity of 9m/s. From every method one result is presented. Thereto, the results from the sensitivity study in Section 3.1 have been analyzed and it is concluded that representative results are obtained when the monitor point is located at 4 times the chord length in front of the blade for the AAT, 3P and the Shen methods. For the AAT method 90 points in the annular ring are selected and for the inverse BEM the a critical is 0.23. The numerical results are also compared with BEM calculations using FAST V8 [35], developed by the National Renewable Energy Laboratory (NREL). For the BEM calculation, polars from CFD computations are used. A uniform axial inflow without shear is used and no 3D correction model is applied in order to isolate the effects of the different airfoil polars. Figures 13a and 14a, the induced velocity extracted from the CFD calculations using different methods is presented and also compared with the induced velocity extracted from BEM. From the induced velocity the AoA is extracted and presented in Figures 13b and 14b. Finally in Figures 13c, 14d, 13d and 14d the aerodynamic lift and drag coefficients are presented. From Figures 13 and 14 the following observation can be reported: (r/r <0.30): In terms of induced velocity (Figures 13a and 14a), deviations are observed at the root of the blades. Herráez and Shen methods are forming one group of results (since they share similar monitoring point location at this part) and all others (AAT, Line Average, 3P, InvBEM and Ferreira-Micallef) are forming the second group. One reason for this deviations could be the 3D nature of the flow at the root. Hence since different methods are using different way of monitoring velocity, this could lead to some deviation in this area. The AoA is presented in Figures 13b and 14b. At the most inner part of the blade, the AoA is different from the 2D case due to the Coriolis and centrifugal forces as well as to the large area with flow separation. The deviation which was presented in the induced velocity is translated to the AoA and hence it leads to up to 5 degrees difference between the two groups of results. Finally in Figures 13c, 14d, 13d and 14d the aerodynamic lift and drag coefficients are presented. The thicker airfoil results in lower lift and higher drag with respect to the other parts of the blade. The rapid changes in C l and C d can be due to the root vortex or spanwise variations in the laminar to turbulent transition location. Also due to the Coriolis and centrifugal forces the extracted C l and C d are much higher than those extracted from 2D calculations without stall delay models. C l is not affected by changes in the AoA, however C d between the two groups of results differs up to 20%. (0.30 <r/r <0.75): In general, a good agreement is observed between all the methods at the mid-span in terms of induced velocity (Figures 13a and 14a). Also in terms of the AoA (Figures 13b and 14b) and C l and C d (Figures 13c, 14d, 13d and 14d) the level of agreement between the methods remains close. (r/r >0.75): At the tip of the blade deviations are observed in terms of induced velocity (Figures 13a and 14a) between the Line Average and Ferreira-Micallef and the remaining methods. The flow at the blade tip is 3D and due to the presence of the tip vortex it is very difficult to capture the flow physics correctly. Hence by having different way of monitoring velocity the results could differ. Nevertheless, the level of the agreement between all the methods beside the Line Average and FM model at the tip remain close. Near the tip it is known that there

is a strong dependency of the induced velocity on the azimuthal position in the rotor plane. As regards to the FM and Line Average approach the main reasons for the discrepancies are associated with the fact that these methods calculates the bound circulation on a 2D airfoil section from a highly 3D flow field as found towards the tip. These methods would therefore seem to necessitate some form of tip correction which is left for future work. In terms of AoA, it is important to note that the AoA (α) is the difference between inflow angle (φ) and twist angle (θ). The inflow angle as indicated in Figure 1 is the arctangent of the rotational speed and the axial velocity. Due to the high value of the rotational velocity in particular near the tip, differences in induced velocity are hidden and so the AoA from the different methods will remain in a very close agreement all over the blade span. However at the very tip of the blade where the Line Average and Ferreira-Micallef method have up to 2 degree deviation from other methods. The flow is also 3D which can cause a higher C d comparing to that at the mid span. Nevertheless, a very high level of agreement for C l remains. For C d the Line Average and Ferreira-Micallef model show small deviations from the average behaviour of all other methods. As before, the deviation between the Line Average and Ferreira-Micallef methods with others could come from the fact that, these methods calculates the bound circulation on a 2D airfoil section from a highly 3D flow field as found towards the tip U ind [m/s] 0 1 2 3 4 AOA [ ] 0 10 20 30 40 50 (a) U ind (b) AoA C L [ ] 0.5 1.0 1.5 2.0 C D [ ] 0.0 0.1 0.2 0.3 0.4 0.5 0.6 5 6 7 8 9 10 0.01 0.03 0.05 AAT Line Average UpWash2 UpWash1 3P InvBEM 0 Gamma BEM (c) C l AAT Line Average Shen2 Shen1 3P InvBEM Herraez BEM (d) C d Figure 13: Comparison of all the considered methods in terms of U ind, AoA, C l and C d for the INNWIND.EU turbine at 9m/s. For a better visibility the inner part of sub figures b,c and d represents the AoA, C l and C d at the outer span location in a smaller range.

U ind [m/s] 0.0 0.5 1.0 1.5 2.0 2.5 3.0 AOA [ ] 0 10 20 30 40 50 (a) U ind (b) AoA C L [ ] 0.4 0.6 0.8 1.0 1.2 C D [ ] 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0 1 2 3 4 5 0.01 0.01 0.03 (c) C l AAT Line Average Shen2 Shen1 3P InvBEM Ferreira Micallef Herraez BEM (d) C d Figure 14: Comparison of all the considered methods in terms of U ind, AoA, C l and C d for the AVATAR turbine at 9m/s. For a better visibility the inner part of sub figures b,c and d represents the AoA, C l and C d at the outer span location in a smaller y-range.

4. Conclusion and future work In this work, several methods for extracting the induced velocity and AoA from CFD computations in axial inflow conditions for flows past two 10MW class wind turbines, namely the EU AVATAR and INNWIND.EU turbines, are presented. From the results presented in the sensitivity study section, it is concluded that the AAT, Line Average and 3P methods show converged behavior when the monitor point is located at 4 times the chord length in front of the blade. Moreover for AAT 90 points in the annular ring seems to be sufficient. For the inverse BEM the a critical has an impact on the U ind. However, this deviation is almost negligible for AoA. For the Shen1 and Shen2 methods, results indicate that the dependency on the monitoring point locations has limited effect from r/r =0.35 of rotor span onwards. However, dependency is still observed in both U ind and AoA at the most inner part of the blade. For all the other methods, no dependency was reported. Based on the results presented, it can be concluded that all the methods are in good agreements at the mid-span. However, at the very root and very tip of the blade deviations are observed. At the root (r/r =0.0 to 0.3) The Herráez method, Shen1 and Shen2 methods are in a very close agreement with respect to each other. This could be due to the fact that the position of monitoring point for these methods are very similar. They are forming the first group of results. The AAT, 3P, Ferreira-Micallef and Line Average methods are also forming the second group of results which are in a closer agreement with each other. At the tip (r/r =0.75 to 1.0) Ferreira-Micallef and Line Average are forming one group of the results. The similarity in the approach and the position of monitoring point at the tip between these two methods could be the reason for this close agreement. This group (Ferreira-Micallef and Line Average) deviate in that region with respect to other methods. These deviations have less impact on the resulting AoA and also C l and C d due to the high value of the rotational velocity in particular near the tip. New correction models can be derived using the extracted airfoil characteristics for blade element momentum calculations. In future work, these methods will be tested in asymmetric inflow conditions. Acknowledgments The authors would like to thank Niels N. Sørensen from DTU for providing the CFD simulations. This project has received partly funding from the European Union 7th framework program for research, technological development and demonstration under the grant agreement No FP7- ENERGY-2013-1/no 608396 (AVATAR project). This work has also partly supported by the Energy Technology Development and Demonstration Program (EUDP-2014, J. nr. 64014-0543) under the Danish Energy Agency. References [1] Schepers J.G. Engineering models in wind energy aerodynamics. Ph.D. Thesis Delft University of Technology 2012. [2] Sant T, Kuik G, Bussel G.J.W. Estimating the angle of attack from blade pressure measurements on the NREL Phase VI rotor using a free wake vortex model: axial conditions. Wind Energy. 2006;9(6):549 577. [3] Johansen J, Sørensen N.N. Airfoil characteristics from 3D CFD rotor computations. Wind Energy. 2004;7:283 294. [4] Chaviaropoulos P.K, Hansen M. OL. Investigating three-dimensional and rotational effects on wind turbine blades by means of a quasi-3d Navier-Stokes solver. Journal of Fluids Engineering. 2000;122(2):330 336. [5] Shen W.Z, Sørensen J.N. Quasi-3D Navier Stokes model for a rotating airfoil. Journal of Computational Physics. 1999;150(2):518 548. [6] Du Z, Selig M.S. A 3-D stall-delay model for horizontal axis wind turbine performance prediction. In: AIAA; 1997. [7] Snel H, Houwink R, Bosschers J, Piers W.J, Van Bussel G.J.W, Bruining A. Sectional prediction of 3D effects for stalled flow on rotating blades and comparison with measurements. In: Proc. European Community Wind Energy Conference, HS Stevens and Associates, LÃ1; 1993.

[8] Leishman J.G. Challenges in Modelling the Unsteady Aerodynamics of Wind Turbines. Wind Energy. 2002;5(2-3):85-132. [9] Theodorsen T. General theory of aerodynamic instability and the mechanism of flutter. : NASA. Ames Res. Center; 1979. [10] Rahimi H, Hartvelt M, Peinke J, Schepers JG. Investigation of the current yaw engineering models for simulation of wind turbines in BEM and comparison with CFD and experiment. 2016;753(2):022016. [11] Rahimi H, Dose B, Stoevesandt B, Peinke J. Investigation of the validity of BEM for simulation of wind turbines in complex load cases and comparison with experiment and CFD. 2016;749(1):012015. [12] (EERA) European Energy Research Alliance. AVATAR: Advanced Aerodynamic Tools for Large Rotors Accessed May 29, 2016;. [13] Schepers G.J, Ceyhan O, Boorsma K, al.. Latest results from the EU project AVATAR: Aerodynamic modelling of 10 MW wind turbines. In: Journal of Physics: Conference Series. [14] Micallef D, Akay B, Sant T, Ferreira CJ, Bussel G. Experimental and numerical study of radial flow and its contribution to wake development of a HAWT. Proceedings of the European wind energy association. 2011;. [15] Micallef D. 3D flows near a HAWT rotor: A dissection of blade and wake contributions. Ph.D. Thesis Delft University of Technology 2012. [16] Campo V, Ragni D, Micallef D, Akay B, Diez F.J, Ferreira C.J. 3D load estimation on a horizontal axis wind turbine using SPIV. Wind Energy. 2014;17(11):1645 1657. [17] Campo V, Ragni D, Micallef D, Diez F.J, Ferreira C.J. Estimation of loads on a horizontal axis wind turbine operating in yawed flow conditions. Wind Energy. 2015;18(11):1875 1891. [18] Rahimi H, Garcia A.M, Stoevesandt B, Peinke J, Schepers JG. An engineering model for wind turbines under yawed conditions derived from high fidelity models. Wind Energy. 2017;. under review. [19] Lindenburg C. Investigation into rotor blade aerodynamics. techreport ECN-C 03-025: Energy Research Centre of the Netherlands, ECN, Petten, The Netherlands; 2003. [20] Hansen M.O.L, Sørensen N.N, Sørensen J.N, Michelsen J.A. Extraction of Lift, Drag and Angle of Attack from Computed 3-D Viscous Flow around a rotating Blade. In: Irish Wind Energy Association; 1998. [21] Shen W.Z, Hansen M.O.L, Sørensen J.N. Determination of angle of attack (AOA) for rotating blades. In: Berlin Heidelberg: Springer-Verlag; 2006. [22] Shen W.Z, Hansen M.O.L, Sørensen J.N. Determination of the angle of attack on rotor blades. Wind Energy. 2009;12(1):91 98. [23] Guntur S, Sørensen N.N. An evaluation of several methods of determining the local angle of attack on wind turbine blades. In: Journal of Physics: Conference Series; 2014. [24] INNWIND.EU: Design of state of the art 10-20MW offshore wind turbines Accessed May 29, 2016;. [25] Michelsen J.A. Basis3D - a Platform for Development of Multiblock PDE Solvers. techreport AFM 92-05: Department of Fluid Mechanics, Technical University of Denmark; 1994. [26] Michelsen J.A. Block structured Multigrid solution of 2D and 3D Elliptic PDEs. techreport AFM 94-06: Department of Fluid Mechanics, Technical University of Denmark; 1994. [27] Sørensen N.N. General Purpose Flow Solver Applied to Flow Over Hills. techreport Risø-R-827: Risø National Laboratory Denmark; 1995. [28] Snel H, Houwink R, Bosschers J, others. Sectional prediction of lift coefficients on rotating wind turbine blades in stall. Netherlands Energy Research Foundation Petten, Netherlands; 1994. [29] Laino D.J, Hansen A, Minnema J.E. Validation of the aerodyn subroutines using NREL unsteady aerodynamics experiment data. In: American Society of Mechanical Engineers; 2002. [30] Hansen M. Aerodynamics of Wind Turbines: second edition. Earthscan Publications Ltd; 2 ed.2015. [31] Schneider M.S, Nitzsche J, Hennings H. Accurate load prediction by BEM with airfoil data from 3D RANS simulations. Journal of Physics: Conference Series. 2016;753(8):082016. [32] Katz J., Plotkin A.. Low-Speed Aerodynamics. Cambridge Aerospace SeriesCambridge University Press; 2001. [33] Jost E, Klein L, Leipprand L, Lutz T, Krämer E. Extracting the angle of attack on rotor blades from CFD simulations. Wind Energy. 2017;. under review. [34] Herraez I, Daniele E, Schepers J.G. Brief communication: Extraction of the wake induction and the angle of attack on rotating wind turbine blades from PIV and CFD results. Wind Energy Science. 2017;. under review. [35] Jonkman J.M, Buhl Jr M.L. FAST user s guide. NREL/EL-500-38230: National Renewable Energy Laboratory; 2005.