Gravo-Aeroelastic Scaling for Extreme-Scale Wind Turbines

Similar documents
Some effects of large blade deflections on aeroelastic stability

HARP_Opt: An Optimization Code for System Design of Axial Flow Turbines

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

Benefits of Preview Wind Information for Region 2 Wind Turbine Control

Adaptive Control of Variable-Speed Variable-Pitch Wind Turbines Using RBF Neural Network

Unsteady structural behaviour of small wind turbine blades Samuel Evans Supervisor: A/Prof Philip Clausen

Measured wind turbine loads and their dependence on inflow parameters

Structural Analysis of Wind Turbine Blades

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

Using Pretwist to Reduce Power Loss of Bend-Twist Coupled Blades

Available online at ScienceDirect. IFAC PapersOnLine 50-1 (2017)

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

Geometric scaling effects of bend-twist coupling in rotor blades

Aeroelastic effects of large blade deflections for wind turbines

0000. Finite element modeling of a wind turbine blade

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

Parameter Sensitivities Affecting the Flutter Speed of a MW-Sized Blade

Engineering Tripos Part IB. Part IB Paper 8: - ELECTIVE (2)

Effect of linear and non-linear blade modelling techniques on simulated fatigue and extreme loads using Bladed

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

Identification of structural non-linearities due to large deflections on a 5MW wind turbine blade

A comparison study of the two-bladed partial pitch turbine during normal operation and an extreme gust conditions

Generally, there exists an optimum tip-speed-ratio, λ that maximized C p. The exact λ depends on the individual wind turbine design

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

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

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

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

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

Dynamic Characteristics of Wind Turbine Blade

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

Statistical Analysis of Wind Turbine Inflow and Structural Response Data From the LIST Program

Implementation of an advanced beam model in BHawC

UNIT V : DIMENSIONAL ANALYSIS AND MODEL STUDIES

Simulation of Aeroelastic System with Aerodynamic Nonlinearity

Anomaly detection of a horizontal wind turbine using an Extended Kalman Filter

Buffeting Response of Ultimate Loaded NREL 5MW Wind Turbine Blade using 3-dimensional CFD

Multidisciplinary Design Optimization Of A Helicopter Rotor Blade

θ α W Description of aero.m

Limit Cycle Oscillations of a Typical Airfoil in Transonic Flow

Aeroelasticity in Dynamically Pitching Wind Turbine Airfoils

Design and modelling of an airship station holding controller for low cost satellite operations

Aero-Propulsive-Elastic Modeling Using OpenVSP

ANALYSIS OF HORIZONTAL AXIS WIND TURBINES WITH LIFTING LINE THEORY

Rotor reference axis

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

Aeroelastic Gust Response

ROTATING RING. Volume of small element = Rdθbt if weight density of ring = ρ weight of small element = ρrbtdθ. Figure 1 Rotating ring

AE 714 Aeroelastic Effects in Structures Term Project (Revised Version 20/05/2009) Flutter Analysis of a Tapered Wing Using Assumed Modes Method

Aero-Structural Optimization of a 5 MW Wind Turbine Rotor THESIS

THE ANALYSIS OF LAMINATE LAY-UP EFFECT ON THE FLUTTER SPEED OF COMPOSITE STABILIZERS

Proof of Concept Test for Dual-Axis Resonant Phase-Locked Excitation (PhLEX) Fatigue Testing Method for Wind Turbine Blades

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

Design of an autonomous flap for load alleviation

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

Dynamic Response of an Aircraft to Atmospheric Turbulence Cissy Thomas Civil Engineering Dept, M.G university

ANALYSIS AND OPTIMIZATION OF A VERTICAL AXIS WIND TURBINE SAVONIUS-TYPE PANEL USING CFD TECHNIQUES

Theoretical Modeling, Experimental Observation, and Reliability Analysis of Flow-induced Oscillations in Offshore Wind Turbine Blades

CHAPTER 3 ANALYSIS OF NACA 4 SERIES AIRFOILS

ENGR 4011 Resistance & Propulsion of Ships Assignment 4: 2017

A new configuration of vertical axis wind turbine: an overview on efficiency and dynamic behaviour 垂直轴风力涡轮机的一种新配置 : 对其效率与动态行为之概览

Experimental Studies on Complex Swept Rotor Blades

Efficient Modelling of a Nonlinear Gust Loads Process for Uncertainty Quantification of Highly Flexible Aircraft

GyroRotor program : user manual

Effect of Blade Number on a Straight-Bladed Vertical-Axis Darreius Wind Turbine

Effect of Blade Number on a Straight-Bladed Vertical-Axis Darreius Wind Turbine

APPENDIX A. CONVENTIONS, REFERENCE SYSTEMS AND NOTATIONS

Analysis Of Naca 2412 For Automobile Rear Spoiler Using Composite Material *

Analysis of Counter-Rotating Wind Turbines

EFFECT OF TAPER AND TWISTED BLADE IN STEAM TURBINES

Aeroelastic Wind Tunnel Testing of Very Flexible High-Aspect-Ratio Wings

Individual Pitch Control for Load Mitigation

A comparison study of the two-bladed partial pitch turbine during normal operation and an extreme gust conditions

Low-Dimensional Representations of Inflow Turbulence and Wind Turbine Response Using Proper Orthogonal Decomposition

Damage Detection in Wind Turbine Towers using a Finite Element Model and Discrete Wavelet Transform of Strain Signals

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

German Aerospace Center (DLR)

Individual Pitch Control of A Clipper Wind Turbine for Blade In-plane Load Reduction

Detailed Load Analysis of the baseline 5MW DeepWind Concept

Dynamic Responses of Jacket Type Offshore Wind Turbines using Decoupled and Coupled Models

Strain-Based Analysis for Geometrically Nonlinear Beams: A Modal Approach

An Experimental Validation of Numerical Post-Stall Aerodynamic Characteristics of a Wing

Control Volume Analysis For Wind Turbines

Given the water behaves as shown above, which direction will the cylinder rotate?

ACTIVE SEPARATION CONTROL ON A SLATLESS 2D HIGH-LIFT WING SECTION

Optimization Framework for Design of Morphing Wings

Effect of functionally graded materials on resonances of rotating beams

AN ENHANCED CORRECTION FACTOR TECHNIQUE FOR AERODYNAMIC INFLUENCE COEFFICIENT METHODS

NUMERICAL INVESTIGATION OF VERTICAL AXIS WIND TURBINE WITH TWIST ANGLE IN BLADES

AEROSPACE ENGINEERING

Active Control of Separated Cascade Flow

Reduction of unwanted swings and motions in floating wind turbines

364 VIBROENGINEERING. JOURNAL OF VIBROENGINEERING. MARCH VOLUME 15, ISSUE 1. ISSN

ANALYSIS OF AEROELASTIC EFFECTS ON THE 3-DIMENSIONAL INTERFERENCE OF WIND-TURBINE ROTORS

AIRFRAME NOISE MODELING APPROPRIATE FOR MULTIDISCIPLINARY DESIGN AND OPTIMIZATION

Shape Optimization of Revolute Single Link Flexible Robotic Manipulator for Vibration Suppression

Definitions. Temperature: Property of the atmosphere (τ). Function of altitude. Pressure: Property of the atmosphere (p). Function of altitude.

FREQUENCY DOMAIN FLUTTER ANALYSIS OF AIRCRAFT WING IN SUBSONIC FLOW

EVALUATE THE EFFECT OF TURBINE PERIOD OF VIBRATION REQUIREMENTS ON STRUCTURAL DESIGN PARAMETERS: TECHNICAL REPORT OF FINDINGS

8 Lidars and wind turbine control

A Framework for the Reliability Analysis of Wind Turbines against Windstorms and Non-Standard Inflow Definitions

Transcription:

AIAA AVIATION Forum 5-9 June 2017, Denver, Colorado 35th AIAA Applied Aerodynamics Conference 10.2514/6.2017-4215 Gravo-Aeroelastic Scaling for Extreme-Scale Wind Turbines a c const. C C! c! E g G I Eric Loth 1, Meghan Kaminski 2, and Chao Qin 3 University of Virginia, Charlottesville, VA, 22904 Lee Jay Fingersh 4 National Renewable Energy Laboratory, Golden, Colorado, 80401 and D. Todd Griffith 5 Sandia National Laboratories, Albuquerque, New Mexico, 87185 A scaling methodology is described in the present paper for extreme-scale wind turbines (rated at 10 MW or more) that allow their sub-scale turbines to capture their key blade dynamics and aeroelastic deflections. For extreme-scale turbines, such deflections and dynamics can be substantial and are primarily driven by centrifugal, thrust and gravity forces as well as the net torque. Each of these are in turn a function of various wind conditions, including turbulence levels that cause shear, veer, and gust loads. The 13.2 MW rated SNL100-03 rotor design, having a blade length of 100-meters, is herein scaled to the CART3 wind turbine at NREL using 25% geometric scaling and blade mass and wind speed scaled by gravo-aeroelastic constraints. In order to mimic the ultralight structure on the advanced concept extreme-scale design the scaling results indicate that the gravoaeroelastically scaled blades for the CART3 are be three times lighter and 25% longer than the current CART3 blades. A benefit of this scaling approach is that the scaled wind speeds needed for testing are reduced (in this case by a factor of two), allowing testing under extreme gust conditions to be much more easily achieved. Most importantly, this scaling approach can investigate extreme-scale concepts including dynamic behaviors and aeroelastic deflections (including flutter) at an extremely small fraction of the full-scale cost. = speed of sound in air = airfoil chord length = constant = centrifugal force = lift coefficient = thrust coefficient = modulus of elasticity of blade material = gravitational acceleration = gravity force = area moment of inertia of blade Nomenclature 1 Professor, Department of Mechanical and Aerospace Engineering, 122 Engineer s Way, AIAA Fellow 2 Graduate Researcher, Department of Mechanical and Aerospace Engineering, 122 Engineer s Way 3 Research Scientist, Department of Mechanical and Aerospace Engineering, 122 Engineer s Way 4 Senior Research Engineer, National Wind Technology Center, 15013 Denver West Parkway 5 Principal Member of the Techincal Staff, Wind Energy Technologies Department, P.O. Box 5800 MS 1124 1 Copyright 2017 by Eric Loth, Meghan Kaminski, Chao Qin, Lee Jay Fingersh and D. Todd Griffith. Published by the, Inc., with permission.

m = mass m = mass per unit length M = Mach number r = radius R = blade length Re = Reynolds Number T = thrust force in the downwind direction TSR = tip-speed ratio V = windspeeds δ = aeroelastic blade deflection angle φ = azimuthal angle of rotor η = geometric scaling of sub-sclae lengths to full-scale lengths (R s /R f ) ρ = density ω = flap-wise frequency ϖ = non-dimensional frequency Ω = rotor angular velocity Subscripts ( ) b = blade value ( ) f = full-scale ( ) m = mass load distribution ( ) s = sub-scale ( ) rat = rated condition ( ) r/r = blade spanwise location ( ) tip = blade tip ( ) w = wind value ( ) ω = harmonic excitation mode A I. Introduction verage wind turbine rated power has increased twenty-fold since 1985 and has been associated with a commensurate increase in rotor swept area. Continued increases in turbine sizes are needed to address the growth in wind energy sought by the DOE in the Wind Vision report. The next generations of turbines from the wind industry are expected to be extreme-scale, i.e. 10 MW are more rated power. For example, Sandia National Laboratories recently designed a series of 100-meter blades for a 13.2 MW system 1 4. The most advanced version of these is the SNL100-03, which was computationally designed to primarily include fiberglass layups, but with carbon fiber added for reinforced shells and girders to carry the gravity, inertial, and aerodynamic loads. The rotors must consider aeroelastic behavior over a wide range of conditions that may occur over the turbine lifetime. In particular, the SNL100-03 design was highly constrained by gravity loads, aeroelastic flutter, as well as buckling and strength requirements under hurricane-level winds. To optimize this blade further, validation with experimental blade deflections and dynamics would be of very high value. However, it would be very expensive to fabricate several blades and test them. As such, there is need for an aeroelastic scaling method to allow high-fidelity testing of extreme-scale blade concepts at cost-effective sub-scale sizes. Such a scaling method would not only reduce testing cost, but would enable testing of a larger number of design options for new concepts. Ideally, one would develop, design, fabricate, and test a fully integrated turbine so as to experimentally demonstrate novel technology such as ultra-light and/or segmented blades 5 over the range of load conditions that represent the full-scale aeroelastic loading and dynamics with high fidelity. A cost-effective development solution is to use scaling in terms of aerodynamic/aeroelastic performance, segmentation, control performance, etc. to conduct testing on a sub-scale demonstrator. To prove the viability of wind turbine design at extreme-scales (10+MW), the sub-scale demonstrator should ideally have all the same critical characteristics of the full-scale turbine, i.e. same geometry, same load-path distributions, same aeroelastic deflection angles and same dynamical behavior (on a per revolution basis). This requires careful consideration of the forces (gravity, centrifugal, and thrust) and system frequencies (e.g. rotational, inertia, and stiffness-based structural frequencies) as discussed by Bottasso et al. 6. This is particularly challenging at extreme-scales because of the large geometric scaling factor 2

required and the increased significance of gravity loads and blade flexibility. The present scaling technique includes the effect of gravity, which is shown to be critical to the aero-elastic deflections and dynamics for extreme-scale turbines. To the authors knowledge, this study describes the first gravo-aeroelastic scaling method (GAS) for extreme-scale wind turbines and is the first to provide scaling characteristics for a sub-scale demonstrator to mimic an extreme-scale turbine. In the present study, the sub-scale demonstrator is based on the CART3 platform located at NREL (Fig. 1a) while the extreme-scale turbine is based on the SNL100-03 rotor designed by Sandia. a) b) Figure 1. Wind turbines viewed from the side assuming wind approaching from the left: a) the 3-bladed CART3 experimental wind turbine at NWTC, and b) side-view forces on a rotor blade including centrifugal (C), gravity (G) and thrust (T) forces (torque force not shown) II. Scaling method A. Geometric and Flow Angle Scaling In order to ensure the non-dimensional aerodynamics are similar between the full-scale and sub-scale systems, geometric scaling is generally employed for all the external dimensions. For geometric scaling, one may define a length-scaling factor (η) using the blade length (R) as η!!!! applies to all external dimensions for fixed geometric scaling (1) In this expression, the subscript s refers to the sub-scale demonstrator and subscript f refers to the full-scale system. For the example case of a 100-meter long blade of a 13.2MW rotor scaled to a 25-meter blade, η=0.25. This same geometric scaling of Eq. (1) is then applied to all other external geometric parameters at each radial location (r). For example, blade chord at each non-dimensional blade radius station (r/r) is preserved via [c s ] r/r = η [c f ] r/r to maintain external geometric scaling (2) A similar scaling for blade thickness is ensured if the geometric outer mold line shape is preserved so only the size is reduced by Eq. (1). Geometrically, it is also important to ensure the relative flow angles relative to the rotor reference frame are kept fixed. At rated conditions, these flow angles near the blade tip are preserved if the rated Tip Speed Ratio (TSR) is the same for the full-scale and sub-scale turbines. This is the ratio between the rated wind speed (V w,rat ) and the rated blade tip speed (V tip =Ω rat R), where Ω rat is the rated angular rotational rate. If geometric scaling is used, flow 3

angle symmetry is retained at all non-dimensional radial locations (r/r) and non-dimensional flow speeds (V/V rat ) so long as the rated Tip Speed Ratio is kept constant, i.e. TSR = V tip /V w,rat = Ω rat R/V w,rat (3a) TSR s = TSR f to maintain flow angles (3b) The relative shape of the blade with respect to the flow is also ideally preserved by enforcing the key geometric blade and rotor angles are fixed (including blade twist, blade coning angle, rotor tilt, etc.). Some adjustments may be needed for pitch angle to account for aerodynamic scaling as discussed below. B. Importance of Load Ratios To allow sub-scale turbines to capture the key blade dynamics and aeroelastic deflections that can lead to fatigue and failure, it is important to ensure that the ratio of the centrifugal, thrust and gravity forces of Fig. 1b (as well as ratio of the torque loads) are preserved in the scaling. This will ensure that non-dimensional flapwise deflections and loads are appropriately reproduced in the scaled turbine. To further ensure non-dimensional edgewise deflections and loads are reproduced, it is important to consider the ratio of the above forces and the ratio of torques. In particular, to ensure that load path angles associated with gravitational, centrifugal, and thrust loads are the same on the full-scale system as they are on the sub-scale system, one must maintain the correct ratios of thrust to centrifugal forces (T/C) and gravity to centrifugal forces (G/C), as a function of azimuthal angle (φ) for a given tip speed ratio. To show why this is important, it is helpful to consider the magnitude of these force ratios for extremescale turbines in terms of changes in blade mass. If blade technology did not change in terms of the materials and structural design (same structural shape and non-dimensional thicknesses of structural elements using same materials), the blade mass would scale cubically with m b,s = η 3 m b,f for constant blade technology (4) However, in practice the increase of rotor mass as technology evolves as shown in Fig. 2 is more mass efficient, due to improvements in the aero-structural-material design. The trend in this figure is consistent with previous authors 7,8 and can be expressed as m b,s η 2.1 m b,f for evolving blade technology as rotor size increases (5) This technology scaling has occurred over time whereby innovations have made blade mass to increase at a subcubic rate, e.g., η 2.1. 4

40 60 80 100 Figure 2. Blade mass dependency on blade length associated with blade volution, where innovation changes result in blade mass scaling with R 2.1 and load angles changing as well (in comparison, fixing load angles requires the blade mass to instead be proportional to R 3 ). Let us now examine the effect that technology scaling has on the size of the force ratios. The ratio of gravitational to centrifugal forces (G/C) is the ratio of gravitational to centrifugal accelerations for a given blade mass. As such, this ratio can be expressed as the gravitational acceleration (g) to centrifugal acceleration (Ω! R!" ). If one considers rated conditions and assumes the center of gravity is about one-third of the blade radius (r CG /R 1/3), then this ratio can be expressed for a given rotor mass as G/C 3Rg/(V tip ) 2 (6) This ratio is thus proportional to a rotational Froude number. If one assumes that rated tip speed (V tip ) is approximately fixed as blade size increases, it can be seen that this ratio of forces increases linearly with blade length, i.e. G/C scales with R. This linearly proportional scaling is represented by the straight line in Fig. 3 assuming a rated tip-speed of 95 m/s (a typical limit due to acoustic and other considerations) and is reasonably consistent with the conditions for the CART-3 and SNL-100-03 conditions as shown by the data symbols. It is shown that gravitational forces can no longer be reasonably neglected for extreme-scale turbines such as the SNL- 100-03. A scaled version of this extreme-scale turbine should ideally fix the ratio of G/C and thus should be represented by the dashed horizontal line on Fig. 3. 5

Figure 3. Ratio of gravity force to centrifugal force as a function of blade length, comparing changes for technology evolution to keeping this ratio fixed for gravo-aeroelastic scaling The result of Eq. (6) shows that gravity force effects should be included in a scaling method to properly mimic the loads and deflections of extreme-scale turbines. However, including gravity loads in wind turbine scaling is uncommon and often not needed. For example, Bottaso et al. 6 noted that neglecting gravitational scaling is acceptable unless very large machines are considered. This neglect of gravity scaling is quite reasonable for wind turbines in the 1-5 MW regime. However, the present study specifically seeks to consider extreme-scale (10+MW) wind turbines where blade masses become very large, such that gravity effects become a significant fraction of overall loading 9. Since the present study is specifically exploring extreme-scale turbines, gravitational scaling (where by G/C is held constant) can no longer be neglected. The importance of the thrust force on deflections and dynamics can be similarly shown to be critical at extremescales by noting that power scales with thrust force (T), swept area (proportional to R 2 ) and wind energy (proportional to V 3 w ). If one assumes that the rated wind speed is fixed, then thrust varies with R 2 while centrifugal forces vary with R 1.1 based on Eq. (5) and Eq. (6). As a result, T/C scales with R 0.9 indicating the thrust forces alos becomes more significant for extreme-esclae systems. And since thrust forces are on the order of the gravity forces 10, it is important to account for thrust and gravity forces via fixing G/C and T/C for aeroelastic scaling. C. Gravo-Aeroelastic Scaling with Load Ratios The ratio of centrifugal to gravitational forces (G/C) can be fixed for gravo-aeroelastic scaling if the ratio of the centrifugal acceleration (Ω 2 R) to gravitational acceleration (g) is fixed for rated conditions. For a fixed tip speed ratio, the wind velocity and rotation speed of the sub-scale wind turbine scale with that of the full-scale turbine as V!,! = V!,! η for fixed ratio of centrifugal to gravity forces (7a) Ω s = Ω f / η for fixed ratio of centrifugal to gravity forces (7b) This scaling is needed in order to maintain the same load path angles for a given extreme-scale technology and indicates that a sub-scale demonstrator must operate at lower speeds in order to properly emulate the effect of gravitational forces experienced on the extreme-scale turbine. To fix the ratio of centrifugal to thrust forces when scaling, note that the centrifugal force is proportional to the product of blade mass (m b ) and centrifugal acceleration (Ω 2 R) while the thrust force (T) is proportional to the 6

product of rotor area (πr 2 ), dynamic pressure (½ρ w V 2 ) and aerodynamic thrust coefficient (c T ). Since TSR is fixed by Eq. (3b) and assuming c T is constant, fixing C/T for rated conditions requires fixing m b /(ρ w R 3 ), which yields the following blade mass and thrust force scaling m b,s = η 3 (ρ w,s /ρ w,f ) m b,f for fixed ratio of centrifugal to thrust forces (8a) T s = η 3 (ρ w,s /ρ w,f ) T f for fixed ratio of centrifugal to thrust forces (8b) This cubic dependence is necessary to ensure that the sub-scale system is exposed to the same net load-path angle as the full-scale system, so that it can properly mimic any extreme-scale conditions and technology innovations. While this scaling for the overall mass and thrust is important for net force ratio, the thrust and mass distribution along the radius should also be kept constant to ensure that the net moments are preserved. This can be ensured by scaling the blade mass per unit length (m ) such that the full-scale and sub-scale turbines have the same non-dimensional distributions as a function of non-dimensional radius (r/r), i.e. m s r/r = η 2 (ρ w,s /ρ w,f ) m f r/r for fixed C/T distribution (9) It is important to note that the above load path scaling is much different from the technology evolution scaling of Fig. 2 which has occurred over time to make rotor blades more structurally efficient as they become larger, and is a result of changes in aero-structural-material design. As such, blade mass scaling will be cubic if force angles are fixed to represent a specific technology or concept (Eq. (8a)) but will be sub-cubic if technology innovation is included so that small-scale system does not represent force angles of large-scale system (Eq. (5)). As a result, the blade mass of a small rotor designed to mimic the dynamics of a large system will be lighter than the blade mass of a conventional rotor that has the same dimensions of the small rotor. D. Blade Dynamics, Aeroelastic Deflection and Buckling For fixed aeroelastic and dynamic scaling, the important scaling parameters to keep fixed are the nondimensional flapping frequency (ϖ =ω/ω) and the aeroelastic deflection angle (δ). If the non-dimensional flapping frequency for the full-scale and sub-scale cases are equal (ϖ s =ϖ f ), and the rotation rate scales per Eq. (7b), then this yields ω s = ω f η -1/2 for fixed non-dimensional flapping frequency (10) This dynamic criterion can be shown to be the same as fixing the steady-aeroelastic deflection by considering elastic beam theory. The flapping frequency (ω) in this case is based on the blade stiffness and is given as ω 2 = (const ω ) 4 EI/(m b R 3 ) (11) In this expression, E is the reference modulus of elasticity of the blade, I is the area moment of inertia of the blade (inertia per unit mass), and const ω is a constant based on the harmonic excitation mode. If the rotor radius scales per Eq. (1) and the blade mass scales per Eq. (8a), then this yields (EI) s = (EI) f (ρ w,s /ρ w,f ) η 5 for fixed non-dimensional flapping frequency (12) If one employs Euler beam theory with a distributed gravity force (G=m b g) over a beam of length R, the nondimensional blade cantilever deflection angle can be expressed as δ=const. m R 2 m b g/(ei) (13) In this expression, const m is a constant based on the particular mass load distribution for the beam (which is fixed based on Eq. (9)). If the gravity driven deflection angles for the full-scale and sub-scale cases are equal (δ s =δ f ) and Eq. (1) and Eq. (8a) are employed so that the ratio of gravity to thrust and centrifugal forces are kept fixed (by Eq. (7) and Eq. (8)), then this yields 7

(EI) s = (EI) f (ρ w,s /ρ w,f ) η 5 for fixed aeroelastic deflection angles (14) This is the same as Eq. (12), indicating that dynamic scaling ensures aeroelastic scaling and vice-versa. Similarly, the ratio of gravity force to buckling force for a beam is proportional to R 2 m b g/(ei) so that if this ratio is fixed, then in combination with Eq. (1) and Eq. (8a) yields the same scaling as Eq. (12). Thus, aeroelastic scaling, dynamic scaling, and buckling scaling can match either the non-dimensional flapping frequency or the non-dimensional deflection angles under gravitational loadings. In concert, this gravo-aeroelastic scaling allows the load-angles, aeroelastic deflection, and blade dynamics of an extreme-scale system to be faithfully represented in a small-scale demonstrator E. Aerodynamic Scaling Aerodynamic scaling can be ensured if the non-dimensional lift, drag and thrust forces along the blade length are retained at the same relative span wise location (r/r). This is ensured if geometric scaling is use, Eq. (1), and flow angles are preserved, Eq. (3), so long as lift, drag and thrust coefficients are independent of scaling. For airfoils with a fixed shape and fixed relative flow angles, this is ensured so long as the effects of Mach number (M) and rated Reynolds number (Re) effects are weak. To determine the potential effects of these parameters, it is helpful to considerer reference values for a given rotor size and speed. If one defines the reference chord (c ref ) as that at r/r=2/3, a as the air speed of sound, ρ w as the wind air density, and µ as the air viscosity, this yields reference values as follows M = (Ω rat R)/a Re = ρ w (Ω rat R)c ref /µ Unfortunately, the combination of geometric and load-path angle scaling will generally not allow matching of the Mach and Reynolds numbers if the flow density and viscosity can not be adjusted. Thus, there will be some effects of compressibility (from Mach number) and of boundary layer transition and growth (from Reynolds number) that may alter the lift, drag and thrust coefficients for a given angle of attack. However, these two effects are generally weak if the Mach number is small enough (e.g. M<0.25) and if the chord Reynolds number is high enough to ensure turbulent boundary layers with weak sensitivity of Reynolds number (e.g. Re>2x10 7 ). For most full- and sub-scaled wind turbines, the Mach number criterion can be reasonably met. However, the lower Reynolds numbers for a subscale turbine may make Reynolds number effects on aerodynamic coefficients to be significant. These may be accommodated in two ways. Firstly, transition strips may be used on the sub-scale airfoils so that the flows are ensured to be fully turbulent for most of the chord length. This is especially important for the suction side of the blade as flow separation and hysteresis effects can be very different between laminar and turbulent flow conditions. Secondly, the Reynolds number effect on the lift-curve slope can be accommodated by ensuring that the lift coefficients as a function of the non-dimensional blade radius (r/r) are preserved. (15a) (15b) C L,s r/r = C L,f r/r for fixed aerodynamic load distribution (16) A small adjustment of the pitch angle or twist angle can be used to accommodate this effect. Noting that drag to lift ratios are very small for a well-designed wind turbine blade and the objective of the present gravo-aeroelastic scaling approach is to preserve blade dynamics and deflections (and not aerodynamic efficiency), the effects of drag can generally be ignored. III. Results While the above scaling can generally be applied to scale any extreme-scale turbine to a sub-scale demonstrator, an example for a three-bladed upwind turbine was chosen to evaluate the practicality of such scaling. In the present case study, aeroelastic scaling is applied to a fully-designed extreme-scale turbine based on the 100-meter long blade for a 13.2 MW turbine designed by Sandia National Labs, denoted SNL100-03 1. This design represents a publically-available internationally-accepted extreme-scale standard which meets IEC criteria but which has not been experimentally tested. Indeed, building and testing an off-shore prototype could cost hundreds of millions of dollars. 8

In this study, the chosen sub-scale demonstrator platform is the CART3 Wind Turbine at NREL (Fig. 1a), which has a 20 m radius and a rated power of 550 kw 11. It has been extensively retrofitted for performing controls research and includes a custom, real-time control system, fully grid-decoupled torque control (type 4), and independent pitch control with removable blades. The turbine provides substantial sensing capabilities, including strain gauges and accelerometers located inside the turbine. The CART3 also has a dedicated meteorological tower, with wind sensing measurements at multiple locations including hub height, to measure inflow wind speed, direction, shear and turbulence. As such, it is a good platform for demonstrating and documenting aeroelastic behavior of a turbine. Table 1 shows the results for applying the above scaling to the chosen extreme-scale blade (SNL-100-03) and sub-scale demonstrator blade. The first data column in Table 1 shows the characteristics of the SNL100-03 full-scale design, based on IEC standards with Class IB winds. The second data column in Table 1 shows the characteristics of the SNL100-03 scaled down by 25%, such that G/C is held constant per Figure 3. Table 1: Scaling SNL-100-03 wind turbine to the CART-3 Platform Wind Turbine and Associated Properties Full-Scale SNL-100-03 Rotor Gravo-Aeroelastic Scaled to CART3: 25% GAS SNL 9 Conventional CART3 Rotor Design Innovation Extreme-Scale Extreme-Scale Conventional Rotor Radius (R) 100 m 25 m 20 m Length Scaling Factor (η) 1 0.25 Not Applicable Number of Blades 3 3 3 Blade mass (m b ) 49,519 kg 620.3 kg 1807 kg Rated Wind Speed (V) 11.3 m/s 5.65 m/s 13.5 m/s Cut-out Wind Speed (V) 25 m/s 12.5 m/s ~24 m/s Stowing Wind Speed (V) N/A ~18 m/s Not Applicable 50-year Wind Speed (V) 70 m/s 35 m/s 70 m/s Rated Power 13.2 MW 82.7 kw 550 kw Wind Air Density (ρ w ) 1.21 kg/m 3 0.97 kg/m 3 0.97 kg/m 3 Tip-Speed Ratio (λ) 9.5 9.5 7.0 Rated Rotor speed (Ω) 10.25 RPM 20.5 RPM 37.1 RPM Reference chord at r/r=2/3 (c) 3.47 m 0.87 m 1.4 m Ref. Reynolds Number (Re)* 1.54x10 7 1.54x10 6 3.67x10 6 Ref. Mach Number (M)* 0.224 0.112 0.166 A key result of the scaling is that the scaled CART3 blade weight for the 25% GAS SNL turbine must be drastically reduced relative the conventional CART3 blade (nearly 3 times lighter!). This highlights the difficulty of designing a blade that can handle the load ratio of an extreme-scale turbine (i.e. the same ultra-light structural design challenges that need to be faced for SNL-100-03 must also be faced for the sub-scale demonstrator). It should be noted that this ultra-light demonstrator blade will be subjected to aerodynamic forces about 5.8 times smaller than that of the conventional CART3 blade since the wind speed scaling of Eq. (2) reduces the rated and cut-out velocities by a factor of about 2.4. The Reynolds and Mach numbers are based on the reference chord length (c) and the rated rotational speed at r/r=2/3. The reduction in Reynolds number of the scaled blade suggests use of an aerodynamic trip strip to retain the relative chord-wise transition of turbulence. Since the reference Mach number at rated conditions is less than 0.25, compressibility effects are fortunately weak for these conditions and so the Mach number need not be scaled. However, the reference Reynolds number of 1.6x10 6 compared to the sub-scale value of 1.5x10 5 indicates that there can be significant differences in the boundary layer transition, and thus on the aerodynamic properties. These may be addressed with slight modifications of the sub-scale turbine blade s aerodynamic shape, as well as forced tripping of the boundary layer on the sub-scale rotor blades at the non-dimensional transition location expected for the fullscale blades. It should be noted that geometric changes in the sub-scale aerodynamic shape (to improve

aerodynamic scaling to the full-scale behavior) should be modest so that the structural response is not significantly affected. In general, the above results indicate that it is feasible to test operational capabilities of a scaled rotor on the CART3 platform while ensuring appropriate scaling for geometric, load, dynamic and aeroelastic characteristics so that the demonstration can replicate all these key features of a 13.2 MW turbine. It should be noted that this ultra-light demonstrator blade will be subjected to rated aerodynamic forces which are four times smaller than that of the conventional CART3 blade, since the wind speed scaling of Eq. (4a) reduces the rated and cut-out velocities by a factor of 2 for η =0.25. An advantage of testing the system on the CART3 platform at NWTC (National Wind Technology Center) is that the effective 50-year gust wind speed of 70 m/s (representing hurricane-like conditions for parked operation) scales down to 35 m/s, a wind speed that is generally achievable within any single year campaign at the NWTC as shown in Fig 4. Testing may be ideal for the months between July and September as this would allow access to a large range of wind speeds within the scaled 50-year gust limit, and would most likely avoid any wind speeds above this limit. 25% scaled 50-year gust Figure 4. Peak wind speeds at the National Wind Technology Center from 2010 to 2015 with 25% scaled 50-year gust wind speeds. The Reynolds and Mach numbers are based on the reference chord length (c) and the rated rotational speed at r/r=2/3. Since the reference Mach number at rated conditions is less than 0.25, compressibility effects are fortunately weak for these conditions and so the Mach number need not be scaled. However, there is nearly a tenfold reduction in Reynolds number of the scaled blade suggests that an aerodynamic trip strip should be used to retain the relative chord-wise location of the transition to a turbulent boundary layer. In addition, small adjustments in pitch control can be used to ensure the criterion of Eq. (14) is reasonably maintained. IV. Conclusions A scaling method taking into account gravitational scaling was developed in order to capture the full-scale dynamic behaviors and aeroelastic deflections of extreme-scale wind turbine blades. The method takes into account the gravitational and thrust loadings, which become a significant factor in extreme-scale turbines as compared to conventional 1-5 MW wind turbines. The gravo-aeroealstic scaling (GAS) can be used to test extreme-scale wind turbine concepts at sub-scale sizes at a far more reasonable cost than full-scale testing. The gravo-aeroealstic scaling method was herein applied to the SNL100-03 rotor to propose a 25% gravitationally-aeroelastically scaled rotor to be placed on the CART3 tower at the NWTC. The development of this rotor shows a scaled blade using GAS to be 25% longer and 1/3 rd the mass of the CART3 original blade. 10

1 2 3 4 5 6 7 8 9 10 11 References Griffith, D. T., and Richards, P. W., The SNL100-03 Blade : Design Studies with Flatback Airfoils for the Sandia 100- meter Blade, 2014. Griffith, D. T., The SNL100-01 Blade: Carbon Design Studies for the Sandia 100-meter Blade, Sand2013-1178, 2013, pp. 1 26. Griffith, D. T., and Ashwill, T. D., The Sandia 100-meter All-glass Baseline Wind Turbine Blade : SNL100-00, Baseline, 2011, pp. 1 67. Griffith, D. T., The SNL100-02 Blade : Advanced Core Material Design Studies for the Sandia 100- meter Blade, Sand2013-10162, 2013. Ichter, B, Steele, A, Loth, E., Moriarty, P. and Sleig, M. A morphing downwind-aligned rotor concept based on a 13- MW wind turbine, Wind Energy, 2016, Vol. 19, pp. 625-637. Bottasso, C. L., Campagnolo, F., Croce, A., and Maffenini, L., Development of a Wind Tunnel Model for Supporting Research on Aero-servo-elasticity and Control of Wind Turbines, 13th International Conference on Wind Engineering, Amsterdam, ICWE13, July 10-15. Ashwill, T. D., Developments in Large Blades for Lower Cost Wind Turbines, SAMPE Journal-Society for the Advancement of Material and Process Engineering, vol. 40, 2004, pp. 65 73. Griffin, D. a, Blade System Design Studies Volume II : Preliminary Blade Designs and Recommended Test Matrix, Design, vol. II, 2004, pp. 3 75. Hillmer, B., Borstelmann, T., Schaffarczyk, P. a, and Dannenberg, L., Aerodynamic and Structural Design of MultiMW Wind Turbine Blades beyond 5MW, Journal of Physics: Conference Series, vol. 75, 2007, p. 12002. Noyes, C., Qin, C., and Loth, E., Ultralight, Morphing Rotor for Extreme-Scale Wind Turbines, AIAA SciTech Forum, Grapevine: 2017, pp. 1 6. Scholbrock, A. A. A., Fleming, P. A., Fingersh, L. J., Wright, A. D., Schlipf, D., Haizmann, F., and Belen, F., Field Testing LIDAR-Based Feed-Forward Controls on the NREL Controls Advanced Research Turbine, 51st AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition, 2013, pp. 1 8. 11