The effect of wind farming on mesocale flow

Similar documents
Wind Flow Modeling The Basis for Resource Assessment and Wind Power Forecasting

Wake meandering under non-neutral atmospheric stability conditions theory and facts. G.C. Larsen, E. Machefaux and A. Chougule

Wake modeling with the Actuator Disc concept

Mesoscale meteorological models. Claire L. Vincent, Caroline Draxl and Joakim R. Nielsen

GENERALISATION OF THE TWO-SCALE MOMENTUM THEORY FOR COUPLED WIND TURBINE/FARM OPTIMISATION

Linking mesocale modelling to site conditions

Large-eddy simulations of the internal boundary layer and wake flow within large wind farms

Wake decay constant for the infinite wind turbine array Application of asymptotic speed deficit concept to existing engineering wake model

Simulating the Vertical Structure of the Wind with the WRF Model

Effect of Wind Turbine Wakes on the Performance of a Real Case WRF-LES Simulation

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

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

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

TNO BOUNDARY LAYER TUNNEL

Offshore wind farm. Copenhagen, 10 March Charlotte Hasager. Senior scientist DTU Wind Energy. Application of mesoscale models with wind farm

Atmospheric stability-dependent infinite wind-farm models and the wake-decay coefficient

The Forcing of Wind Turbine Rotors by True Weather Events as a Function of Atmospheric Stability State*

Three dimensional actuator disc modelling of wind farm wake interaction

ADAPTATION OF THE REYNOLDS STRESS TURBULENCE MODEL FOR ATMOSPHERIC SIMULATIONS

A comparative and quantitative assessment of South Africa's wind resource the WASA project

Two-scale momentum theory for very large wind farms

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

Detailed validation of dynamic loading simulation of offshore wind turbines operating in wake

IMPLEMENTATION OF PRESSURE BASED SOLVER FOR SU2. 3rd SU2 Developers Meet Akshay.K.R, Huseyin Ozdemir, Edwin van der Weide

Turbulence - Theory and Modelling GROUP-STUDIES:

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

Kommerciel arbejde med WAsP

The Long distance wake behind Horns Rev I studied using large eddy simulations and a wind turbine parameterization in WRF

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

Comparison of turbulence models for the computational fluid dynamics simulation of wind turbine wakes in the atmospheric boundary layer

Influence of sea structures on wind measurements: CFD analysis

WASA WP1:Mesoscale modeling UCT (CSAG) & DTU Wind Energy Oct March 2014

Generating Virtual Wind Climatologies through the Direct Downscaling of MERRA Reanalysis Data using WindSim

Extreme Winds in the Western North Pacific. Søren Ott

WASA Project Team. 13 March 2012, Cape Town, South Africa

Simulating wind energy resources with mesoscale models: Intercomparison of stateof-the-art

Reducing Uncertainty of Near-shore wind resource Estimates (RUNE) using wind lidars and mesoscale models

Comparison of a Mesoscale Model with FINO Measurements in the German Bight and the Baltic Sea

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

Iterative Learning Control for Smart Rotors in Wind turbines First Results

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

Application of mesoscale models with wind farm parametrisations in EERA-DTOC

Study of the atmospheric wake turbulence of a CFD actuator disc model

Beyond the Point Particle: LES-Style Filtering of Finite-Sized Particles

Canopy structure effects on the wind at a complex forested site

Wind conditions based on coupling between a mesoscale and microscale model

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

What is a wind atlas and where does it fit into the wind energy sector?

Local blockage effect for wind turbines

Wake modeling and simulation

STATOR/ROTOR INTERACTION

Measurements and Simulations of Wakes in Onshore Wind Farms Julie K. Lundquist 1,2

2.1 OBSERVATIONS AND THE PARAMETERISATION OF AIR-SEA FLUXES DURING DIAMET

A turbine interaction model for choosing operating points in wind farms

Carbon Trust Offshore Wind Accelerator (OWA) Designing an Offshore Wind Measurement Campaign to Promote the Development of Wake Effects Models

This wind energy forecasting capability relies on an automated, desktop PC-based system which uses the Eta forecast model as the primary input.

PRELIMINARY STUDY OF COMPUTATIONAL SETUP FOR URBAN STREET CANYONS. by MUHAMMAD NOOR AFIQ WITRI, M.Eng

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

UNCERTAINTY QUANTIFICATION WITH DEPENDENT INPUTS: WIND AND WAVES

Mechanical Engineering for Renewable Energy Systems. Wind Turbines

Mellor-Yamada Level 2.5 Turbulence Closure in RAMS. Nick Parazoo AT 730 April 26, 2006

WIND CLIMATE ESTIMATION USING WRF MODEL OUTPUT: MODEL SENSITIVITIES

SAWEP Workshop. and. Hans E. Jørgensen Head of programme Meteorology Wind Energy Division Risø DTU. Cape Town, 4 th March 2010

Aerodynamics of Centrifugal Turbine Cascades

Validation of Boundary Layer Winds from WRF Mesoscale Forecasts over Denmark

Extreme wind atlases of South Africa from global reanalysis data

WIND FARM PERFORMANCE MONITORING WITH ADVANCED WAKE MODELS

Direct and Large Eddy Simulation of stably stratified turbulent Ekman layers

Publication 97/2. An Introduction to Turbulence Models. Lars Davidson, lada

Self-similarity and turbulence characteristics of wind turbine wakes via large-eddy simulation

LARGE-EDDY SIMULATIONS OF A WIND TURBINE WAKE ABOVE A FOREST

EROSION. D1.1 Disdrometers in EROSION. Sagsnr B. Deliverable: D1.1 (Public) Authors: Charlotte Hasager (DTU) and Flemming Vejen (DMI)

LARGE-EDDY SIMULATIONS OF A WIND TURBINE WAKE ABOVE A FOREST

A Detached-Eddy-Simulation study

W I N D R E S O U R C E A S S E S S M E N T

Comparisons of winds from satellite SAR, WRF and SCADA to characterize coastal gradients and wind farm wake effects at Anholt wind farm

Influence of Atmospheric Stability on Wind-Turbine Wakes: A Large-Eddy Simulation Study

Assessing WRF PBL Schemes for Wind Energy Applications

A Reexamination of the Emergy Input to a System from the Wind

Mesoscale Modelling Benchmarking Exercise: Initial Results

Sergej S. Zilitinkevich 1,2,3. Division of Atmospheric Sciences, University of Helsinki, Finland

CFD calculations of the test 2-4 experiments. Author: G. de With

A primal-dual mixed finite element method. for accurate and efficient atmospheric. modelling on massively parallel computers

Speedwell High Resolution WRF Forecasts. Application

Offshore wind power prediction in critical weather conditions

Aerodynamic Rotor Model for Unsteady Flow and Wake Impact

2.3 The Turbulent Flat Plate Boundary Layer

Validation of Offshore load simulations using measurement data from the DOWNVInD project

WINDS FROM METEOSAT AND GMS WV-IMAGERY USING VARIATIONAL TECHNIQUES. M. Rohn. Institut für Meteorologie, Freie Universität Berlin Berlin, F.R.G.

Sea Ice Model for Marginal Ice Zone

Modeling of Coastal Ocean Flow Fields

Descripiton of method used for wind estimates in StormGeo

350 Int. J. Environment and Pollution Vol. 5, Nos. 3 6, 1995

Numerical Modelling for Optimization of Wind Farm Turbine Performance

Wind velocity measurements using a pulsed LIDAR system: first results

Validation and comparison of numerical wind atlas methods: the South African example

Vindkraftmeteorologi - metoder, målinger og data, kort, WAsP, WaSPEngineering,

ESCI 485 Air/sea Interaction Lesson 3 The Surface Layer

EXCITATION OF GÖRTLER-INSTABILITY MODES IN CONCAVE-WALL BOUNDARY LAYER BY LONGITUDINAL FREESTREAM VORTICES

Wind Power Meteorology

Transcription:

The effect of wind farming on mesocale flow Validation and prediction A.J. Brand This report has been presented at: - the European Wind Energy Conference 2009, Marseille, France, 16-19 March 2009; - IEA Wind Task 23 Back-to-back Workshop Offshore wind farms - Wake effects and power fluctuations, Risø-DTU, Risø campus, Denmark, 25-26 February 2009; - We-at-sea/ATO/NWEA Offshore Wind Energy Conference Essential innovations, Den Helder, Netherlands, 12-13 February 2009 ECN-M--09-051

The effect of wind farming on mesoscale flow - Validation and prediction Arno J Brand ECN Wind Energy, P.O. Box 1, NL 1755 ZG Petten, Netherlands E: brand@ecn.nl, T: +31 224 56 4775, F: +31 224 56 8214 Abstract This paper addresses validation of the planetary boundary layer method MFwWF and predictions as obtained with MFwWF. The validation considers measured wind farm wakes. The predictions include the modification of the wind profile due to wind farming plus the impact of wind farm design parameters and meteorological parameters on the wind profile. A validation of MFwWF is presented on basis of velocities measured downstream of the Horns Rev and the Nysted wind farms. The calculated decrease of the relative velocity deficit with increasing distance downstream a wind farm is supported by measured data from both wind farms. As to the calculated decrease of this deficit with increasing upstream velocity the measured data is inconclusive. Quantitative conclusions have not been drawn because the calculations and the measurements are evaluated in different spanwise positions. In a qualitative sense the calculated relative velocity deficits are in agreement with the measured ones. Resolved profiles show how most of the velocity change occurs in the lower part of the boundary layer whereas most of the wind direction change occurs in the upper part, and that the thinner the boundary layer or the larger the surface roughness, the larger the wind direction change. Near a 5 MW wind turbine with a rotor diameter of 100 m operating at full load the velocity deficit is of the order of 5%, the wind direction change is increased with 1 2 deg, and the velocity recovery distance is 20 rotor diameters. For a wind farm with 22 of these turbines separated at 10 rotor distances these numbers are 15%, 2 3 deg, and at least 2 wind farm length scales. Initial velocity deficits and velocity recovery distances show the impact of nominal power density and geostrophic velocity for a wind farm which consists of 22 wind turbines with a nominal power of 5 MW. The initial velocity deficit relative to the upstream velocity decreases with increasing geostrophic velocity in general, and ranges from 6% (at a turbine separation of 14 rotor diameters) to 32% (at a separation of 5 rotor diameters) if the velocity at hub height is halfway cut-in and nominal. At this hub-height velocity the absolute initial velocity deficit reaches a maximum (of 1.2 m/s in the case of a nominal power density of 5 MW/km 2 ) and the velocity recovery distance relative to the wind farm length scale is of the order of 20. The relative velocity recovery distance for other geostrophic velocities varies between 0 (at low geostrophic velocities) and a limit value of the order of 40 (at high velocities). Key words Wind farm wake, Wind resource assessment 1. Introduction Offshore wind farms tend to be placed closer together over the years, as already illustrated by Offshore Windfarm Egmond aan Zee OWEZ and Princess Amalia Wind Farm (separated 15 km) in the Netherlands or Horns Rev I and II (separated 23 km) in Denmark. Since these separation distances are between 5 and 10 times the wind farm's horizontal length scale, the velocity deficit due to an upstream wind farm may be considerable [1]. If so, energy production loss and mechanical load increase are expected to be significant. For this reason the dedicated planetary boundary layer method Mesoscale Flow with Wind Farming (MFwWF) has been developed, which method computes the interaction between a wind farm and the prevailing wind. The methodology behind MFwWF and predictions obtained with MFwWF have been reported separately [2][3][4]. In this paper we address validation of MFwWF and predictions obtained with MFwWF. We start with brief descriptions of prior work on modelling wind farm wakes (section 2) and the new dedicated planetary boundary layer method MFwWF (section 3). Next validation on basis of measured wind farm wakes is addressed (section 4). We then present predictions in the form of the modification of the wind profile due to wind farming plus the impact of wind farm design parameters and meteorological parameters on the wind profile (section 5). Finally, we summarize the findings (section 6). 2. Prior work The mean (in the sense of Reynolds averaged) flow in the neutral planetary boundary layer usually is described by the momentum equations in combination with the continuity equation and a set of boundary conditions [5]. These momentum equations represent equilibrium between convective forces, pressure gradient forces, Coriolis

forces, forces due to turbulent flux gradients, and external forces. A wind farm wake study requires simulation of mesoscale atmospheric flow together with energy extraction/redistribution due to wind turbines. The studies that have been published so far can be subdivided into two categories: self-similar approaches and mesoscale approaches. In a selfsimilar approach [6] [7] the convective force and the spanwise turbulent momentum flux gradients are assumed to dominate the flow, allowing for standard wake-like solutions. In a mesoscale approach, on the other hand, the flow is assumed to be dominated by the Coriolis force and the vertical turbulent momentum flux gradients, opening the door to either extra surface drag approaches [8] or more generic mesoscale approaches [9][10][11]. As is shown in other publications [2][3], neither the self-similar wake approach nor the extra surface drag approach is valid on the mesoscale because over the separation distance between wind farms the convective and the Coriolis forces are of equal order of magnitude so that neither can be neglected. Although this was already implicitly recognized in the more generic approaches, these studies lack realistic formulations for the turbulence and the wind turbines. 3. Flow model 3.1. Principles and outline of method This subsection presents a brief description of the planetary boundary layer method MFwWF; details have been reported elsewhere [2][3]. MFwWF is a CFD method that is based on three principles: (1) Neutral planetary boundary layer flow with wind farming essentially is steady and twodimensional. In addition the convective forces, the Coriolis forces, the vertical and spanwise gradients of the turbulent momentum fluxes, and the external forces that represent wind turbines all have the same order of magnitude. (2) A numerical representation of the momentum equations in the form of backward differences allows for (a) an implicit solution of the two horizontal velocity components in vertical direction, iterating on the turbulent viscosity while employing a mass-energy conserving scheme, and (b) a marching solution in the horizontal directions. (3) The continuity equation is satisfied by employing the Lagrange multiplier method to the velocity components that satisfy the continuity equation. Because of its mixed implicit/explicit character, the planetary boundary layer method MFwWF is computationally fast and cheap. L y w i n d f a r m L x d y grid point (centre of cell) grid point where calculated velocity is evalu- point where measured velocity is evaluated Figure 1: Horizontal lay-out of the grid cells in relation to the horizontal area covered by the wind farm, and definition sketch of downstream distance d x, spanwise distance d y, streamwise wind farm length scale L x and spanwise wind farm length scale L y. Also indicated are the points where measured velocity is evaluated in the validation study Apart from the hub height and power curve of the turbines in the wind farm and information on the location and the lay-out of the wind farm, the input to the planetary boundary layer method MFwWF includes the geostrophic velocity, the geostrophic height (i.e. the height where the geostrophic velocity is reached) and the surface roughness length (i.e. the height where the velocity is zero). The output of MFwWF consists of the grid-cellaveraged values of the two horizontal velocity components and the turbulent viscosity. The turbulent viscosity is modeled with a Baldwin-Lomax model [12] which is modified such that realistic values of the velocity and the turbulence intensity at heights between 10 m and 100 m are obtained. Figure 1 shows the horizontal lay-out of the grid cells in relation to the horizontal area covered by the wind farm. Also indicated in figure 1 are the locations where measured velocity is evaluated in the validation study (section 4); note these are at a different spanwise location. 3.2. Numerical stability In this subsection we briefly address the numerical stability of the solution procedure in MWwWF. To this end consider the matrix-vector representation of the discretized momentum equations. Now in general an error in a matrix element will propagate into an error in the resolved vector. In order to prevent instability we require that the absolute value of the growth rate remains smaller than 1. At the bottom of the numerical domain, where errors grow from initial velocities u 0 and v 0, we have shown that this requirement is met if d x

( Δx, Δy) f max( u ) φ g, v g min Δz 4 min c b1 k 2 m ( z ) where Δx and Δy indicate the horizontal grid spacings, Δz min indicates the spacing between the lower two points in vertical direction, c b1 indicates a constant of proportionality, f φ is the Coriolis parameter, u g and v g are the streamwise and the spanwise component of the geostrophic velocity, and k m (z 0 ) indicates the turbulent viscosity at the bottom of the numerical domain. This means that for the solution procedure to remain stable the horizontal grid sizes must be larger than a critical value which is proportional to the vertical grid size and the value of the Coriolis parameter, and which is inversely proportional to the turbulent viscosity. At the top of the numerical domain, on the other hand, errors grow from initial velocities u u g and v v g, and we have shown that the requirement on preventing error propagation is met if min( 2Δx Δy,2Δy Δx) max( u ) g, v g c 2 t1 Δz k ( h ) max m geo where Δz max indicates the spacing between the upper two points in vertical direction, c t1 indicates a constant of proportionality and k m (h geo ) is the turbulent viscosity at the top of the numerical domain. As an illustration of the stability requirements figure 2 shows the empirically determined minimal horizontal grid size for different combination of the geostrophic height and the surface roughness length in the case of a geostrophic wind speed of 14.1 m/s. Figure 2: Smallest horizontal grid spacing if the solution procedure is to remain stable as a function of geostrophic height h geo and surface roughness length z 0 4. Validation In this section we address validation of the planetary boundary layer method MFwWF. Validation data consists of 10-minute averaged velocities measured at or near hub height at met masts located upstream and downstream of the Horns Rev wind farm and the Nysted wind farm. Note that in both cases the velocity is measured along 0 the central wind turbine row, which implies that measured and calculated velocities are evaluated at different spanwise positions in this study (see figure 1). The Horns Rev wind farm consists of 80 wind turbines with a hub height of 70 m, a rotor diameter of 80 m and a nominal power of 2 MW. It covers an area of approximately 20 km 2, and has wind turbine rows directed 90-270 deg (from West to East). The Nysted wind farm, on the other hand, consists of 72 wind turbines with a hub height of 69 m and a rotor diameter of 84 m. These turbines can operate at two modes, reaching the full nominal power of 2.3 MW at one mode but a lower nominal power of 0.4 MW at the other. Nysted covers an area of approximately 23 km 2, and has rows directed 98-278 deg. Figure 3 shows the velocity W relative to the upstream velocity W ups as a function of the downstream distance d x from the Horns Rev wind farm for upstream velocities of 6 m/s, 8 m/s and 10 m/s in combination with wind directions between 255 deg and 285 deg. In the calculations the surface roughness length is 0.1 mm, the geostrophic height is 500 m, and the geostrophic velocity is set such that the correct velocity at hub height is obtained. The number of grid points in vertical direction is 100. The calculated velocities are average values in grid cells with a horizontal area of 3.92 x 3.92 km 2 and the centre at a spanwise distance of 1.96 km at a height of 66 m. The measured velocities, on the other hand, are point values at a spanwise distance zero and at downstream distances of 2 km and 6 km at a height of 70 m. Upstream turbulent intensities have the same order of magnitude in the calculations and in the measurements (10% resp. between 7% and 8%). The data in figure 3 clearly show that the calculated relative velocity deficit (W ups -W)/W ups is smaller than the measured one. This is explained by the different spanwise positions of the points where calculated and measured velocity are evaluated in combination with the different character of these velocities (cell averaged versus point value). There however is qualitative agreement in the sense that the relative velocity deficit decreases with downstream distance. Figure 3 also shows that at a spanwise distance d y of 0 km relative velocity deficits of 15 17% have been measured at a downstream distance of 2 km, and 7 10% at 6 km. This means that for a given downstream distance and upstream velocity between 6 m/s and 10 m/s the measured relative velocity deficit is proportional to the upstream velocity. In contrast, at a spanwise distance d y of 1.96 km (which corresponds to d y /L y = 0.50 where L y is the spanwise length scale of the wind farm), the calculated relative velocity deficits are found to decrease with increasing upstream velocity, with values in the range 14 13% at a d x of 2 km

and 5 4% at 6 km (first value corresponds to the lower upstream velocity). Calculations with higher values of the surface roughness length and/or the geostrophic height did not give significantly different results. W / W ups [-] 1.05 1.00 0.95 0.90 0.85 calc 285 calc 255 calc 270 meas min 255-285 meas ave 255-285 meas max 255-285 W / W ups [-] 1.05 calc 293 calc 263 1.00 calc 278 meas min 263-293 0.95 meas ave 263-293 meas max 263-293 0.90 0.85 0.80 0.75-5 0 5 10 d x [km] W / W ups [-] W / W ups [-] 0.80 0.75-5 0 5 10 1.05 1.00 0.95 0.90 0.85 0.80 calc 285 calc 255 calc 270 meas min 255-285 meas ave 255-285 meas max 255-285 d x [km] 0.75-5 0 5 10 1.05 1.00 0.95 0.90 0.85 0.80 calc 285 calc 255 calc 270 meas min 255-285 meas ave 255-285 meas max 255-285 d x [km] 0.75-5 0 5 10 d x [km] Figure 3: Velocity W relative to the upstream velocity W ups as a function of distance d x downstream the Horns Rev wind farm; upstream velocity 6 m/s (top), 8 m/s (centre) and 10 m/s (bottom). Calculated grid cell averaged velocities are presented for three different wind directions: aligned with turbine rows (wind direction 270 deg) and misaligned over ±15 deg (285 resp. 255 deg). Horizontal error bars indicate the size of the grid cell. Bold vertical error bars indicate the minimum and the maximum of the point measurements. As a reference thin vertical error bars show a variation of 7.5% about the average of the point measurements W / Wups [-] 1.05 1.00 0.95 0.90 0.85 0.80 calc 293 calc 263 calc 278 meas min 263-293 meas ave 263-293 meas max 263-293 0.75-5 0 5 10 d x [km] Figure 4: Velocity relative to the upstream velocity as a function of distance downstream the Nysted wind farm; upstream velocity 8 m/s (top) and 10 m/s (bottom). Calculated velocities are presented for three different wind directions: aligned with turbine rows (wind direction 278 deg) and misaligned over ±15 deg (293 resp. 263 deg). See caption of figure 3 for interpretation of error bars Figure 4 shows the same information for the Nysted wind farm for upstream velocities of 8 m/s and 10 m/s in combination with wind directions between 263 deg and 293 deg. Apart form the wind turbine and the wind farm data the input to the calculations is the same as for the Horns Rev case, which however implies that the calculations are evaluated at d y /L y = 0.49 as the Nysted wind farm has a smaller spanwise length scale. The velocities are measured at downstream distances of 2 km and 5 km at a height of 69 m. Measured upstream turbulent intensity is between 7% and 9%. The data in figure 4 reveal much the same information as in figure 3 but with one exception: the measured relative velocity deficits are found to decrease with increasing upstream velocity which is in agreement with the trend in the calculated velocity deficits. In conclusion, the calculated relative velocity deficits have been found to decrease with downstream distance from a wind farm, which is in agreement with the measurements in the wake of the Horns Rev and the Nysted wind farms. Also these calculated deficits have been found to decrease with upstream velocity, which is in contrast to measured deficits from Horns Rev but in agreement with those from Nysted. Quantitative conclusions can however not be drawn because of the difference between the spanwise positions

where the calculations and the measurements are evaluated. Nevertheless, in a qualitative sense the predicted relative velocity deficits have been found to be in agreement with the measured ones because the calculated deficits (evaluated at a spanwise position of 1.95 km) are smaller than the measured deficits (evaluated at a spanwise position zero). MW/km 2. Figure 9 shows that if the wind farm operates at full load the initial velocity deficit is of the order of 15% and that the velocity twist is increased with 2 3 deg. The velocity recovery distance is at least 2 streamwise wind farm length scales L x, where the streamwise length scale is equal to the length of the wind farm. 5. Predictions 5.1. Resolved velocity profiles 5.1.1. Empty set In this section the resolved velocity profiles for the empty set, that is a domain without wind farming, are presented. Figure 5 shows four vertical profiles, each in a corner of a 200 x 200 km 2 domain, as valid for a geostrophic height of 500 m and a surface roughness length of 0.1 mm. The number of grid points in vertical direction is 25. The figures display the streamwise velocity versus height, the spanwise velocity, the angle between the streamwise and the spanwise velocity, and a hodograph of the two velocity components. The data in the figure is in qualitative agreement with the observed height dependence of undisturbed wind, where most of the velocity change occurs in the lower part and most of the direction change occurs in the upper part of the boundary layer, but it is too early to decide on the quantitative agreement. The figures 6 and 7 display a much thicker boundary layer (1500 m) with the same surface roughness, and the same boundary layer thickness in combination with a much rougher surface (1 cm). The figures show that the thinner the boundary layer or the larger the surface roughness, the larger the twist in the velocity profile. Again this qualitative agreement with observations is to be collaborated with quantitative data. 5.1.2. Wind turbine and wind farm The modification of the wind profile due to an hypothetical wind turbine is studied for a turbine with a nominal power of 5 MW operating at full load, and having a rotor diameter of 100 m and a hub height of 70 m. Figure 8 shows that the relative initial velocity deficit is of the order of 5% and that the velocity twist is increased with 1 2 deg. (The relative initial velocity deficit is the ratio of the velocity deficit at d x = 0 and the upstream velocity.) The velocity recovery distance d rec is 20 rotor diameters. This distance is the downstream distance where the velocity is 99.9% of the upstream velocity, and is determined by fitting a decay law to the velocities and the corresponding downstream distances. The hypothetical wind farm consists of 22 turbines with a rotor diameter of 100 m, a hub height of 70 m and a nominal power of 5 MW. The turbine separation distance is 1 km (10 rotor diameters) so that the nominal power density is 5

z / hgeo [-] 1.0E-06 Point SW Point SE Point NW Point NE z / hgeo [-] 1.0E-06 1.0E-07 Point SW Point SE Point NW Point NE 1.0E-07 0.0 0.2 0.4 0.6 0.8 1.0 1.0E-08 0.0 0.2 0.4 0.6 0.8 1.0 u / G [-] u / G [-] z / hgeo [-] 1.0E-06 z / hgeo [-] 1.0E-06 1.0E-07 1.0E-07-0.4-0.3-0.2-0.1 0.0 0.1 0.2 1.0E-08-0.4-0.3-0.2-0.1 0.0 0.1 0.2 v / G [-] v / G [-] z / hgeo [-] 1.0E-06 z / hgeo [-] 1.0E-06 1.0E-07 1.0E-07-15.0-10.0-5.0 0.0 5.0 1.0E-08-15.0-10.0-5.0 0.0 5.0 atan( v / u ) [deg] atan( v / u ) [deg] 0.1 0.1 0.0 0.0 v / G [-] -0.1 v / G [-] -0.1-0.2-0.2-0.3 0.0 0.2 0.4 0.6 0.8 1.0-0.3 0.0 0.2 0.4 0.6 0.8 1.0 u / G [-] u / G [-] Figure 5: Vertical profiles of streamwise velocity u and spanwise velocity v in four corners of the numerical domain as non-dimensionized with the geostrophic velocity G and the geostrophic height h geo ; valid for a geostrophic height of 500 m and a surface roughness length of 0.1 mm Figure 6: Vertical profiles of streamwise velocity u and spanwise velocity v in four corners of the numerical domain as non-dimensionized with the geostrophic velocity G and the geostrophic height h geo ; valid for a geostrophic height of 1500 m and a surface roughness length of 0.1 mm

z / hgeo [-] Point SW Point SE Point NW Point NE z / hgeo [-] 1.0E-06 dx = -2 km dx = 0 km dx = 2 km dx = 4 km 0.0 0.2 0.4 0.6 0.8 1.0 1.0E-07 0.0 0.2 0.4 0.6 0.8 1.0 u / G [-] u / G [-] z / hgeo [-] z / hgeo [-] 1.0E-06-0.4-0.3-0.2-0.1 0.0 0.1 0.2 1.0E-07-0.4-0.3-0.2-0.1 0.0 0.1 0.2 v / G [-] v / G [-] z / hgeo [-] z / hgeo [-] 1.0E-06-15.0-10.0-5.0 0.0 5.0 1.0E-07-15.0-10.0-5.0 0.0 5.0 atan( v / u ) [deg] atan( v / u ) [deg] 0.1 0.1 0.0 0.0 v / G [-] -0.1 v / G [-] -0.1-0.2-0.2-0.3 0.0 0.2 0.4 0.6 0.8 1.0 u / G [-] Figure 7: Vertical profiles of streamwise velocity u and spanwise velocity v in four corners of the numerical domain as non-dimensionized with the geostrophic velocity G and the geostrophic height h geo ; valid for a geostrophic height of 500 m and a surface roughness length of 1 cm -0.3 0.0 0.2 0.4 0.6 0.8 1.0 u / G [-] Figure 8: Vertical profiles of streamwise velocity u and spanwise velocity v for different distances d x behind a wind turbine as non-dimensionized with the geostrophic velocity G and the geostrophic height h geo ; valid for a geostrophic height of 500 m and a surface roughness length of 0.1 mm

z / hgeo [-] 1.0E-06 1.0E-07 dx = -2 km dx = 0 km dx = 2 km dx = 4 km 0.0 0.2 0.4 0.6 0.8 1.0 Wini / Wups [-] 1.0 0.9 0.8 0.7 0.6 0 5 10 15 20 25 P nom / A farm [MW / km 2 ] u / G [-] 50 z / hgeo [-] z / hgeo [-] 1.0E-06 1.0E-07-0.4-0.3-0.2-0.1 0.0 0.1 0.2 v / G [-] 1.0E-06 1.0E-07-15.0-10.0-5.0 0.0 5.0 atan( v / u ) [deg] 0.1 d rec / L x [-] 40 30 20 10 0 0 5 10 15 20 25 P nom / A farm [MW / km 2 ] Figure 10: Initial downstream velocity W ini relative to the upstream velocity W ups and velocity recovery distance d rec relative to the streamwise wind farm length scale L x as a function of nominal power density P nom /A farm for a wind farm in a geostrophic height of 500 m, a geostrophic velocity of 14 m/s, and a surface roughness length of 0.1 mm Wini / Wups [-] 1.00 0.95 0.90 0.85 0.80 0 10 20 30 40 50 G [m / s] v / G [-] 0.0-0.1-0.2 d rec / L x [-] 50 40 30 20-0.3 10 0.0 0.2 0.4 0.6 0.8 1.0 u / G [-] 0 0 10 20 30 40 50 Figure 9: Vertical profiles of streamwise velocity u and spanwise velocity v for different distances d x behind a wind farm as non-dimensionized with the geostrophic velocity G and the geostrophic height h geo ; valid for a geostrophic height of 500 m and a surface roughness length of 0.1 mm G [m / s] Figure 11: Initial downstream velocity W ini relative to the upstream velocity W ups and velocity recovery distance d rec relative to the streamwise wind farm length scale L x as a function of geostrophic velocity G for a wind farm with a nominal power density of 5 MW/km 2

5.2. Impact of wind farm design parameters and meteorological parameters Wind farm design parameters include separation distance from and layout (spacing, nominal power density) of the wind farm, and hub height and rotor diameter of the wind turbine. The impact of nominal power density is studied by changing the turbine separation distance in the hypothetical wind farm between 5 and 14 times the rotor diameter and keeping the geostrophic velocity at a constant value such that the hub-height velocity is halfway cut-in and nominal. Figure 10 shows that the relative initial velocity deficit (W ups -W ini )/W ups increases with the nominal power density from 6% (turbine separation 14 rotor diameters) to 32% (5 rotor diameters), and that the relative velocity recovery distance d rec /L x is of the order of 20. Meteorological parameters include geostrophic velocity, geostrophic height and surface roughness length. Figure 11 shows the impact of the geostrophic velocity for the hypothetical wind farm for hub height velocities near cut-in, halfway cut-in and nominal, near nominal, halfway between nominal and cut-out, and beyond cut-out. The relative initial velocity deficit (W ups -W ini )/W ups is found to decrease with increasing geostrophic velocity, and the largest absolute initial velocity deficits W ups -W ini (of in this case 1.2 m/s) occur when the hub-height velocity is halfway cut-in and nominal. Also the relative velocity recovery distance is found to increase with the geostrophic velocity, from 0 at low geostrophic velocities to a limit value near 40 at high geostrophic velocities. 6. Summary A validation of the dedicated planetary boundary layer method MFwWF has been presented on basis of wakes measured downstream of the Horns Rev and the Nysted wind farms. The calculated decrease of the relative velocity deficit with increasing distance behind a wind farm is shown to be supported by data measured in both wind farms. On the other hand, the measured data is shown to be inconclusive as to the calculated decrease of this deficit with increasing upstream velocity. In addition it has been shown that quantitative conclusions can not be drawn because the spanwise positions where the calculations and the measurements are evaluated are different. May this be as it is, it has been shown that in a qualitative sense the calculated relative velocity deficits are in agreement with the measured ones. Resolved profiles have been presented that show how most of the velocity change occurs in the lower part of the atmospheric boundary layer whereas most of the wind direction change occurs in the upper part. The profiles also show that the thinner the boundary layer or the larger the surface roughness, the larger the wind direction change. Furthermore it has been shown that near a wind turbine with a rotor diameter of 100 m operating at a full load of 5 MW the velocity deficit is of the order of 5%, the wind direction change is increased with 1 2 deg, and the velocity recovery distance is 20 rotor diameters. And it has been shown that for a wind farm with 22 of these turbines these numbers are 15%, 2 3 deg, and at least 2 wind farm length scales, respectively. Initial velocity deficits and velocity recovery distances have been presented that show the impact of nominal power density and geostrophic velocity for a wind farm which consists of 22 wind turbines with a nominal power of 5 MW. It has been shown that the initial velocity deficit relative to the upstream velocity decreases with increasing geostrophic velocity in general, and ranges from 6% (at a turbine separation of 14 rotor diameters) to 32% (at a separation of 5 rotor diameters) if the velocity at hub height is halfway cut-in and nominal. Also it has been shown that at this hub-height velocity the absolute initial velocity deficit reaches a maximum (of 1.2 m/s in the case of a nominal power density of 5 MW/km 2 ), and the velocity recovery distance relative to the wind farm length scale is of the order of 20. Finally it has been shown that the relative velocity recovery distance for other geostrophic velocities ranges between 0 (at low geostrophic velocities) and a limit value of the order of 40 (at high velocities). Acknowledgements Work funded by the Dutch Ministry of Economic Affairs in BSIK programme We@Sea, project "Windenergiecentrale Noordzee - Parkinteractie" (We@Sea/BSIK 2005/002). Measured data by courtesy of L.E. Jensen of Dong Energy A/S as prepared in the framework of the European UPWIND research project under contract with the European Commission (CE Contract Number 019945 (SES6)). References [1] M.B. Christiansen and C.B. Hasager, 2005, "Wake studies around a large offshore wind farm using satellite and airborne SAR", 31st Int. Symp on Remote Sensing of Environment, St Petersburg, Russian Federation [2] A.J. Brand, 2007, "Modelling the effect of wind farming on mesoscale flow - Part 1: Flow model", J. of Physics: Conf. Series 75 (2007) 012043 [3] A.J. Brand, 2008, "Modelling the effect of wind farming on mesoscale flow - Part 2: Modification of the velocity profile", European Wind Energy Conference 2008, Brussels, Belgium [4] A.J. Brand, 2008, "The effect of wind farming on mesoscale flow", Global Wind Energy Conference 2008, Beijing, China [5] J.R. Holton, 1992, An Introduction to

dynamic meteorology (3rd ed.), Academic Press [6] S. Frandsen et al., 2004, "The necessary distance between large wind farms offshore - Study", Report Risø-R- 1518(EN) [7] T. Hegberg, 2004, "Turbine interaction in large offshore wind farms - Atmospheric boundary layer above a wind farm", Report ECN-C--04-033 [8] T. Hegberg, 2002, "The effect of large wind farms on the atmospheric boundary layer", Global Wind Power Conference 2002, Paris, France [9] M-K. Liu et al., 1983, "Mathematical model for the analysis of wind-turbine wakes", J. Energy, Vol. 7, No. 1, pp. 73-78 [10] S. Baidya Roy et al., 2004, "Can large wind farms affect local meteorology?", J. Geoph. Research, Vol. 109, D19101 [11] P. Rooijmans, 2004, "Impact of a largescale offshore wind farm on meteorology - Numerical simulations with a mesoscale circulation model", Universiteit Utrecht, Masters thesis [12] D.C. Willcox, 1998, Turbulence modelling for CFD (2nd ed.), DCW Industries Inc.

Wind Farm Design When other wind farms are close Arno J. Brand Motivation Source: Rem Koolhaas' Office for Metropolitan Architecture (OMA) 2

Motivation Source: Noordzeeloket 3 Motivation Day-ahead Source: Wind Service Holland 4

Motivation Source: DONG Energy 5 Motivation Large offshore wind farms may affect each other s wind resource velocity deficit WF 1 disturbed wind rose 6

Challenge Can we design a wind farm when other sites are close? 7 Challenge Can we design a wind farm when other sites are close? Yes How? 8

Challenge Can we design a wind farm when other sites are close? Yes How? By determining the local wind climate How? 9 Challenge Can we design a wind farm when other sites are close? Yes How? By determining the local wind climate How? By modelling planetary boundary layer flow with wind farming 10

Conclusion minimum safe distance WF velocity deficit disturbed wind rose velocity recovery distance 11 Outline Concepts Approach Validation Predictions 12

Concepts Design parameters Wind farm design parameters Farm separation distance Turbine spacing Hub height Rotor diameter Nominal power Meteorological design parameters Geostrophic velocity Geostrophic height Surface roughness length 13 Concepts Design parameters Streamwise turbine spacing WF 1 WF 2 Streamwise farm separation distance 14

Concepts Design parameters Height Geostrophic height Geostrophic velocity Hub height Surface roughness length Wind farm Downstream distance 15 Concepts Impact parameters Velocity deficit Velocity recovery distance Minimum safe distance Disturbed sectors 16

Concepts Velocity deficits and recovery distances Velocity Initial velocity deficit Wind farm Downstream distance Streamwise velocity recovery distance 17 Concepts Spanwise recovery distance Velocity Wind farm Spanwise distance Spanwise velocity recovery distance 18

Concepts Velocity deficit velocity deficit WF 19 Concepts Minimum safe distance minimum safe distance velocity deficit WF 20

Concepts Disturbed sectors minimum safe distance velocity deficit WF 1 WF 2 21 Concepts Disturbed sectors minimum safe distance velocity deficit WF 1 disturbed sector 22

Concepts Disturbed wind rose velocity deficit WF 1 disturbed wind rose 23 Approach Flow solver Neutral planetary boundary layer flow with wind farming Steady and two-dimensional Equilibrium between convective and Coriolis forces and vertical and spanwise turb. mom. flux gradients and forces due to the wind turbines Numerical representation Implicit solution in the vertical Marching solution in horizontal directions Implicit Lagrange multiplier velocity correction 24

Approach Flow problem spanwise distance downstream distance spanwise wind farm length scale upstream wind speed streamwise wind farm length scale 25 Approach Velocity decomposition x-axis y-axis u(h geo ) u(z) U(z) G u(z ns ) v(z) v(h geo ) 26

Vertical and horizontal length scales Approach z [m] 10 3 top layer h geo 10 2 turbine layer h hub mesoscale motion farm wake 10 1 farm turbine far wake 10 0 turbine near wake 10-1 surface layer z 0 turbine x,y [m] 10 0 10 1 10 2 10 3 10 4 10 5 10 6 27 Approach Length and velocity scales U y U z U x L z 1 km L y 100 km L x 100 km 28

Approach Governing equations ( ) u' v' u' w' v v a g x u u u + v = + f + φ x y y z ( ) v' v' v' w' u u a g y v v u + v = f + φ x y y z u v + x y = 0 29 Approach Layout of grid cells d y L y w i n d f a r m d x L x grid point (centre of cell) grid point where calculated velocity is evaluated point where measured velocity is evaluated 30

Approach Decay of velocity deficit ΔW ΔW farm ( x) farm,ini = W(x) x 2D rot m ΔW farm (x) ΔW farm,ini Wind farm x Inspiration: R.J. Barthelmie et al., 2008, EWEC 2008, Brussels, Belgium 31 Decay of velocity deficit Validation ΔW ΔW farm ( x) farm,ini = W(x) x 2D rot m ΔW farm (x) ΔW farm,ini ΔW farm,ini and m from calculations and measurements Wind farm x Inspiration: R.J. Barthelmie et al., 2008, EWEC 2008, Brussels, Belgium 32

Validation Grid-cell versus one-point velocities Translation ( x, x ) = W f ( m, ΔW ; N, D, x, x, Δy) Wcell 1 2 0 ini t 1 2 33 Validation Horns Rev wind farm Grid-cell values y [m] 6154000 N t = 80, D= 80 m, H= 70 m, P nom = 2 MW Wind Turbine Met Mast One-point values 6152000 6150000 6148000 1.0 6146000 422000 424000 426000 428000 430000 432000 434000 436000 438000 x [m] 0.9 Source: DONG Energy W / W ups [-] calc 285 0.8 calc 255 calc 270 meas min 255-285 meas ave 255-285 0.7 meas max 255-285 -5 0 5 10 d x [km] 34

Validation Nysted wind farm Grid-cell values y [m] 6052000 6050000 N t = 72, D= 84 m, H= 69 m, P nom = 2.3 MW Wind Turbine Met Mast One-point values 6048000 6046000 1.0 6044000 670000 672000 674000 676000 678000 680000 682000 684000 686000 x [m] 0.9 Source: DONG Energy W / W ups [-] calc 293 0.8 calc 263 calc 278 meas min 263-293 meas ave 263-293 0.7 meas max 263-293 -5 0 5 10 d x [km] 35 Velocity profile without a wind farm Predictions geo region y [km] 200 180 160 140 120 100 80 60 40 20 0 0 20 40 60 80 100 120 140 160 180 200 Point SW Point SE Point NW Point NE logarithmic layer sub layer z / hgeo [-] 1.0E-06 1.0E-07 Point SW Point SE Point NW Point NE 0.0 0.2 0.4 0.6 0.8 1.0 ( u**2 + v**2 ) / G [-] x [km] 36

Velocity profile without a wind farm Predictions geo region y [km] 200 180 160 140 120 100 80 60 40 Point SW Point SE Point NW Point NE logarithmic layer sub layer z / hgeo [-] 1.0E-06 1.0E-07 Point SW Point SE Point NW Point NE -15.0-10.0-5.0 0.0 5.0 20 0 0 20 40 60 80 100 120 140 160 180 200 atan( v / u ) [deg] x [km] 37 Predictions Velocity profile near a hypothetical wind farm y [m] 4000 N t = 22, D= 100 m, H= 70 m, P nom = 5 MW Wind Turbine velocity deficit 2000 0-2000 -4000-8000 -6000-4000 -2000 0 2000 4000 6000 8000 y [km] 200 dx/lx = -3.06 dx/lx = -1.35 180 dx/lx = 0.35 dx/lx = 2.06 160 140 120 100 80 60 40 20 0 0 20 40 60 80 100 120 140 160 180 200 x [km] x [m] z / hgeo [-] 1.0E-06 1.0E-07 dx/lx = -3.06 dx/lx = -1.35 dx/lx = 0.35 dx/lx = 2.06 0.0 0.2 0.4 0.6 0.8 1.0 ( u**2 + v**2 ) / G [-] 38

Predictions Conclusion minimum safe distance WF velocity deficit disturbed wind rose velocity recovery distance 39 Extra 40

Approach Momentum equations Discretisation Representation Solution procedure Formal order of method Numerical stability Conservation of mass and energy 41 Approach Other aspects Lagrange-multiplier approach Turbulence parameterization Wind turbine parameterization Initial and boundary conditions Discretization error estimation 42

Predictions Impact of nominal power density on initial velocity deficit 43 Predictions Impact of nominal power density on velocity recovery distance 44

Summary Summary 45