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

Similar documents
Two-scale momentum theory for very large wind farms

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

Local blockage effect for wind turbines

Manhar Dhanak Florida Atlantic University Graduate Student: Zaqie Reza

3D hot-wire measurements of a wind turbine wake

Turbulence - Theory and Modelling GROUP-STUDIES:

Turbulence Modeling I!

Interaction(s) fluide-structure & modélisation de la turbulence

Simulating Drag Crisis for a Sphere Using Skin Friction Boundary Conditions

A Detached-Eddy-Simulation study

Wind turbine wake interactions at field scale: An LES study of the SWiFT facility

Fluid Dynamic Simulations of Wind Turbines. John Abraham, Brian Plourde, Greg Mowry University of St. Thomas

Tutorial School on Fluid Dynamics: Aspects of Turbulence Session I: Refresher Material Instructor: James Wallace

Numerical Simulation of a Blunt Airfoil Wake

Assessment of Various Diffuser Structures to Improve the Power Production of a Wind Turbine Rotor

Before we consider two canonical turbulent flows we need a general description of turbulence.

arxiv: v1 [physics.flu-dyn] 1 Feb 2017

Numerical Methods in Aerodynamics. Turbulence Modeling. Lecture 5: Turbulence modeling

NONLINEAR FEATURES IN EXPLICIT ALGEBRAIC MODELS FOR TURBULENT FLOWS WITH ACTIVE SCALARS

ρ t + (ρu j ) = 0 (2.1) x j +U j = 0 (2.3) ρ +ρ U j ρ

Numerical simulations of heat transfer in plane channel flow

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

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

Numerical Investigation of the Transonic Base Flow of A Generic Rocket Configuration

A Computational Investigation of a Turbulent Flow Over a Backward Facing Step with OpenFOAM

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

Turbulence modelling. Sørensen, Niels N. Publication date: Link back to DTU Orbit

Fluid Dynamics Exercises and questions for the course

The mean shear stress has both viscous and turbulent parts. In simple shear (i.e. U / y the only non-zero mean gradient):

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

Flow characteristics of curved ducts

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

RECONSTRUCTION OF TURBULENT FLUCTUATIONS FOR HYBRID RANS/LES SIMULATIONS USING A SYNTHETIC-EDDY METHOD

Numerical Simulation of Flow Around An Elliptical Cylinder at High Reynolds Numbers

Investigation of the Numerical Methodology of a Model Wind Turbine Simulation

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

A dynamic global-coefficient subgrid-scale eddy-viscosity model for large-eddy simulation in complex geometries

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

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

Mechanical Engineering for Renewable Energy Systems. Wind Turbines

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

Modelling of turbulent flows: RANS and LES

Numerical Investigation of Thermal Performance in Cross Flow Around Square Array of Circular Cylinders

Instability of wind turbine wakes immersed in the atmospheric boundary layer

Turbulent Boundary Layers & Turbulence Models. Lecture 09

Turbulence Solutions

Turbulent Rankine Vortices

Turbulence Modeling. Cuong Nguyen November 05, The incompressible Navier-Stokes equations in conservation form are u i x i

Experimental Study on Influence of Pitch Motion on the Wake of a Floating Wind Turbine Model

EVALUATION OF FOUR TURBULENCE MODELS IN THE INTERACTION OF MULTI BURNERS SWIRLING FLOWS

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

Turbulence and its modelling. Outline. Department of Fluid Mechanics, Budapest University of Technology and Economics.

Numerical investigation of swirl flow inside a supersonic nozzle

2.2 The Turbulent Round Jet

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

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

Zonal hybrid RANS-LES modeling using a Low-Reynolds-Number k ω approach

Turbulence: Basic Physics and Engineering Modeling

Comparison of Turbulence Models in the Flow over a Backward-Facing Step Priscila Pires Araujo 1, André Luiz Tenório Rezende 2

Introduction to Turbulence and Turbulence Modeling

CHARACTERISTIC OF VORTEX IN A MIXING LAYER FORMED AT NOZZLE PITZDAILY USING OPENFOAM

Aeroacoustic calculations of a full scale Nordtank 500kW wind turbine

Aerodynamic Rotor Model for Unsteady Flow and Wake Impact

Probability density function (PDF) methods 1,2 belong to the broader family of statistical approaches

Experimental and Numerical Investigation of Flow over a Cylinder at Reynolds Number 10 5

Effects of Free-Stream Vorticity on the Blasius Boundary Layer

LARGE EDDY SIMULATION OF FLOW OVER NOZZLE GUIDE VANE OF A TRANSONIC HIGH PRESSURE TURBINE

Numerical Heat and Mass Transfer

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

Vortex shedding from slender surface mounted pyramids

DEVELOPMENT OF CFD MODEL FOR A SWIRL STABILIZED SPRAY COMBUSTOR

Comparison of two LES codes for wind turbine wake studies

Boundary-Layer Theory

COMPUTATIONAL SIMULATION OF THE FLOW PAST AN AIRFOIL FOR AN UNMANNED AERIAL VEHICLE

LES ANALYSIS ON CYLINDER CASCADE FLOW BASED ON ENERGY RATIO COEFFICIENT

CFD Time Evolution of Heat Transfer Around A Bundle of Tubes In Staggered Configuration. G.S.T.A. Bangga 1*, W.A. Widodo 2

1. Introduction, tensors, kinematics

Computational Fluid Dynamics 2

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

Prediction of noise from a wing-in-junction flow using computational fluid dynamics

Model Studies on Slag-Metal Entrainment in Gas Stirred Ladles

Numerical simulation of dispersion around an isolated cubic building: Model evaluation of RANS and LES. Yoshihide Tominaga a and Ted Stathopoulos b

The Simulation of Wraparound Fins Aerodynamic Characteristics

FLOW-NORDITA Spring School on Turbulent Boundary Layers1

O. A Survey of Critical Experiments

A Full Scale Elliptic CFD Analysis of the Anisotropic Flow in the Wake of a Wind Turbine

Hydrodynamic Characteristics of Gradually Expanded Channel Flow

SHEAR-LAYER MANIPULATION OF BACKWARD-FACING STEP FLOW WITH FORCING: A NUMERICAL STUDY

WALL ROUGHNESS EFFECTS ON SHOCK BOUNDARY LAYER INTERACTION FLOWS

Chapter 7 The Time-Dependent Navier-Stokes Equations Turbulent Flows

CFD modelling of lab-scale anaerobic digesters to determine experimental sampling locations

Active Control of Separated Cascade Flow

Iterative Learning Control for Smart Rotors in Wind turbines First Results

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

NUMERICAL STUDY ON WIND TURBINE WAKES UNDER THERMALLY-STRATIFIED CONDITIONS

Fluid Mechanics. Chapter 9 Surface Resistance. Dr. Amer Khalil Ababneh

Actuator Surface Model for Wind Turbine Flow Computations

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

Wall treatments and wall functions

Direct Numerical Simulation of fractal-generated turbulence

Transcription:

Low-Order Modelling of Blade-Induced Turbulence for RANS Actuator Disk Computations of Wind and Tidal Turbines Takafumi Nishino and Richard H. J. Willden ariv:20.373v [physics.flu-dyn] Oct 202 Abstract Modelling of turbine blade-induced turbulence (BIT) is discussed within the framework of three-dimensional Reynolds-averaged Navier-Stokes (RANS) actuator disk computations. We first propose a generic (baseline) BIT model, which is applied only to the actuator disk surface, does not include any model coefficients (other than those used in the original RANS turbulence model) and is expected to be valid in the limiting case where BIT is fully isotropic and in energy equilibrium. The baseline model is then combined with correction functions applied to the region behind the disk to account for the effect of rotor tip vortices causing a mismatch of Reynolds shear stress between short- and long-time averaged flow fields. Results are compared with wake measurements of a two-bladed wind turbine model of Medici and Alfredsson [Wind Energy, Vol. 9, 2006, pp. 29-236] to demonstrate the capability of the new model. Introduction Modelling of turbulent mixing behind turbines is an important area of research for the wind and tidal power industries. Whilst recent eddy-resolving computations (such as large-eddy simulations and detached-eddy simulations) using a so-called actuator line method have given valuable insight into time-dependent features of wind/tidal turbine wakes (e.g., []), it is still of great importance to develop a loworder model that satisfactorily predicts the characteristics of time-averaged flow around turbines at much lower computational cost. Such a low-order model would be useful for the design of wind and tidal power farms for the future. Takafumi Nishino Engineering Science, University of Oxford, e-mail: takafumi.nishino@eng.ox.ac.uk Richard H. J. Willden Engineering Science, University of Oxford e-mail: richard.willden@eng.ox.ac.uk

2 Nishino and Willden The objective of the present study is to develop a reasonably generic low-order model that describes time-averaged flow around a horizontal-axis wind/tidal turbine of various designs and operating conditions. Our current model is based on the threedimensional Reynolds-averaged Navier Stokes (RANS) equations coupled with an actuator disk model. To account for the effect of turbulence (or velocity fluctuations) induced by turbine blades we refer to this as blade-induced turbulence (BIT) in this paper we first introduce a simple baseline model, which requires the specification of BIT energy and scale but does not require any additional model coefficients (other than those used in the original RANS turbulence model). The baseline model is then combined with correction functions to account for the effect of rotor tip vortices. Details of the model are described below, followed by the results of a model validation study. 2 Baseline Model The baseline model used in this study is the one recently proposed by the authors [5]. The governing equations of the flow are the three-dimensional incompressible RANS equations, where the Reynolds stress tensor, u i u j, is modelled using the standard k-ε model of Launder and Spalding [2]. A simple actuator disk concept is used, where a turbine rotor is modelled as a stationary permeable disk of zero thickness placed perpendicular to the incoming flow. The effect of the rotor on the mean flow is considered as a loss of streamwise (or axial) momentum at the disk plane. The change in momentum flux (per unit disk-area and per unit fluid-density) is locally calculated as S U = 2 KU d 2, where U d is the local streamwise velocity at the disk plane and K is a momentum loss factor (constant over the disk) to determine the thrust acting on the disk. It is known that this actuator disk model (without taking account of BIT) yields much weaker turbulent mixing behind the disk compared to that measured behind a turbine rotor. In the baseline BIT model we assume that, at the disk plane, virtual turbine blades generate turbulence that is characterised by its turbulent kinetic energy k b and dissipation rate ε b. Two physical parameters are introduced to determine the values of k b and ε b : (i) the ratio of the energy converted to BIT to that removed from the mean flow at the disk plane, β, and (ii) a representative length scale for BIT, l b. Note that β and l b are model variables rather than model coefficients, as they depend on actual turbine design and operating conditions. k b is then calculated (locally over the disk plane) as k b = βs U = 2 βku d 2, whereas ε b is estimated (based on the high Reynolds number equilibrium hypothesis) as ε b = C 3 4 µ k 3 2 b /l b, where C µ = 0.09 (following the standard k-ε model). To account for the combined effect of BIT and the turbulence coming from upstream of the turbine, we further assume that: (i) the disk plane is a special internal boundary that does not allow the transport of k or ε via their diffusion (so that the transport of k and ε through the disk plane is only via the convection from upstream to downstream) and (ii) at the disk plane, the BIT of k b and ε b is mixed (in the time-

Low-Order Modelling of Blade-Induced Turbulence 3 averaged sense) with the ambient (or upstream) turbulence of k a and ε a, resulting in the mixed turbulence of k m and ε m just downstream of the disk. Here k m is calculated as k m = k a + k b, whereas ε m is estimated by assuming the conservation of the time-integral of linearly decaying turbulent kinetic energy, i.e., k m τ m = k a τ a + k b τ b, where τ a = k a /ε a, τ b = k b /ε b and τ m = k m /ε m represent the initial eddy turnover time or lifetime for the ambient, blade-induced and mixed turbulence, respectively [5]. Eventually, the changes in k and ε to be added to their transport equations at the disk plane (per unit disk-area) are calculated as S k = U d (k m k a ) = U d k b, () [ (k a + k b ) 2 ] S ε = U d (ε m ε a ) = U d (ka/ε 2 a ) + (kb 2/ε b) ε a. (2) Generally, the two model variables, β and l b, should depend on the actual turbine design and may be given as functions of the distance, r, from the disk axis. In the present study, however, uniform values of β and l b are given either to the entire disk surface or to the disk edge region defined by (0.5d w edge ) r 0.5d, where d is the disk diameter and w edge (= 0.d in this study) is the width of the disk edge region. For the latter case, β = 0 is given to the rest of the disk surface. 3 Tip Vortex Correction As will be shown later, the baseline model tends to yield too strong/fast turbulent mixing in the core region (r/d < 0.4) in the near wake, even when BIT is given only around the disk edge. This suggests that further modifications are required on the modelling of turbulent mixing in the near wake region. An important issue to be considered here is how large the Reynolds shear stress is (relative to the turbulent kinetic energy) behind a turbine, especially in the region where strong rotor tip vortices exist. It is known that for many two-dimensional turbulent shear flows where the production and dissipation of k are close to equilibrium, the value of u xu y/k is around 0.3 (when the mean shear du/dy < 0) or 0.3 (when du/dy > 0) and this is the basis of the turbulent viscosity constant C µ = (0.3) 2 = 0.09 used in the standard k-ε model (see, e.g., Pope [6]). In the mean shear layer behind a turbine rotor, however, experimental results have shown that the magnitude of u xu y/k can be significantly smaller than 0.3 [4]. This is most likely due to the effect of tip vortices causing a mismatch between short-time (or phase) averaged and long-time averaged wake shear profiles. Specifically, tip vortices periodically create adverse velocity gradient regions, u / r < 0 (where φ denotes a short-time average of φ over a time scale relevant to the passing-through time of each tip vortex), and hence adverse Reynolds shear stress regions, u xu r > 0, whilst the long-time averaged wake shear direction is U/ r > 0 and therefore u xu r < 0. This explains why the long-time-averaged wake shear behind a rotor does not diffuse quickly despite its high level of turbulent kinetic energy.

4 Nishino and Willden To account for the above effect of tip vortices within the framework of RANS actuator disk computations, we consider applying empirical correction functions to the so-called tip vortex region. The streamwise extent of the region will be given as a model variable, l tip, which may be further modelled as a function of turbine design and operating conditions in future studies. For the present study, we introduce three correction functions, f µ, f k and f ε, into the standard k-ε model: Dε Dt = Dk Dt = u i u j = ν t ( νt ( ) νt k u U i i σ k x u j f k ε, (3) j σ ε ε ) ( Ui + U j x i u U i ε i u j f ε C ε k C ε2 ) 23 kδ i j, ν t = f µ C µ k 2 ε 2 k, (4) ε, (5) where C µ = 0.09, C ε =.44, C ε2 =.92, σ k =.0 and σ ε =.3 (following the standard k-ε model). To reduce the eddy viscosity and thus mitigate turbulent mixing in the tip vortex region behind the disk (located at x = 0), we model f µ as follows: f µ = for x 0, l tip x, (6) f µ = f r ( f µmin ) for 0 x ltip, (7) ( f µ = f r ( f µmin ) + cos π(x l tip ) ) 2 l tip ltip for ltip x l tip, (8) where f µmin = 0. is given in this study, following experimental observations [4] (note, however, that this is also a model variable and should generally depend on the rotor tip-speed ratio and the number of blades). ltip is the streamwise extent of the region where f µ does not change (corresponding to the region where tip vortices are stable); we assume ltip = 2 l tip in the present model. f r is a damping (sinusoidal) function in the radial (r) direction, normalised such that f r = at r = 0.5d (where the rotor edge is located) and monotonically decreases to f r = 0 at r = 0 and d. A major difficulty in this correction is that the change in ν t due to the introduction of f µ affects not only the strength of turbulent diffusion but also the production of k and ε. This necessitates the introduction of f k and f ε in Eqs (3) and (4), respectively; however, we assume f ε = in this study for simplicity. Preliminary Here the introduction of f k is justified since, in the tip vortex region, the production of k should be linked to the mean of the magnitude of short-time averaged velocity gradient and the corresponding Reynolds stress rather than those for the long-time averaged flow field, i.e., f k accounts for the difference between u i u U i j (known in the model) and u i u j u i (unknown in the model). Meanwhile, physical interpretation of f ε seems less straightforward since the production term of ε includes ε and k, which should also be divided into two components corresponding to the short-time and long-time averaged flow fields, respectively. This seems far beyond the capability of single-scale eddy viscosity models (perhaps we require a proper multi-scale model for such discussion; see, e.g., Sagaut et al. [7]) and hence we assume f ε = in this study.

Low-Order Modelling of Blade-Induced Turbulence 5 computations with f k = yielded too small k in the near wake, whereas those with f k = / f µ (i.e., fully cancelling out the effect of f µ on the production of k) resulted in too large k in the near wake. To maintain the level of turbulent kinetic energy in the near wake comparable to that measured in the experiments [4] (and also to minimise its sensitivity to the value of l tip ), we model f k as f k = + γ k ( f µ )/ f µ with the model coefficient γ k = 0.85 in this study (note that γ k = 0 and correspond to the two extreme cases, f k = and / f µ, respectively). 4 Model Validation A model validation study was performed for wind tunnel tests of a two-bladed wind turbine model of Medici and Alfredsson [3, 4]. The tunnel test section is.2m wide and 0.8m high. The tunnel ceiling is adjustable, so that the freestream velocity, U in, is constant throughout the test section (when the turbine is not installed). The rotor diameter d is 0.8m and its hub height from the floor is 0.24m. The hub diameter is 20mm. Table summarises experimental conditions; for this particular set of experiments, a turbulence-generating grid was installed, providing freestream turbulence (FST) intensities of 4.5% at x/d = 0 (turbine location) and 2.5% at x/d = 9. Here the thrust coefficient C T and the tip-speed ratio (TSR) are based on the freestream velocity and therefore slightly different from those reported earlier [3]. For the computations, the turbine rotor was modelled as an actuator disk of zero thickness with a small non-permeable disk embedded at the centre to account for the effect of the rotor hub. The supporting tower of the turbine was not modelled. Table Experimental conditions U in TSR C T Yaw angle FST intensity 8.4 [m/s] 3.87 0.899 0 [deg.] 4.5% (x/d = 0), 2.5% (x/d = 9) Fig. Contours of turbulent kinetic energy across a horizontal plane at hub height: (a) without BIT, (b) baseline model (for the entire actuator disk), (c) baseline model (only for 0.4 < r/d < 0.5), and (df) baseline model (only for 0.4 < r/d < 0.5) with the tip vortex correction (l tip /d =, 3, 5 for d, e, f, respectively).

6 Nishino and Willden 2 x/d = 3 2 x/d = 3 0.5 0 0 0 0.5 2 0 2 4 6 8 0 2 (a) (b) (c) (d) Exp. (a) (b) (c) (d) Exp. 2 0 0.2 0.4 0.6 0.8.2 2 0 0.05 0. Fig. 2 Centreline streamwise velocity (left), streamwise velocity profiles at x/d = 3 (centre) and turbulent kinetic energy profiles at x/d = 3 (right): (a) without BIT, (b) baseline model (for the entire actuator disk), (c) baseline model (only for 0.4 < r/d < 0.5), (d) baseline model (only for 0.4 < r/d < 0.5) with the tip vortex correction (l tip /d = 3), and experimental data [3]. Computations were performed using FLUENT 2 (together with its User-Defined Function module). The solver is based on a finite volume method and is nominally 2nd-order accurate in space. All computations were performed as steady state. Figures and 2 present results for C T = 0.899, β = 0.30 and l b = 0.d (except for the case without BIT). The case without BIT yields too small turbulent kinetic energy and thus too slow wake recovery. The baseline model applied to the entire disk yields too large turbulent kinetic energy in the core region (r/d < 0.4) and hence too early wake recovery behind the disk. The baseline model applied only to the edge region still yields too strong turbulent mixing in the core region in the near wake, and the tip vortex correction provides better results in both near and far wake. Further validation studies are needed to examine the robustness of the model. Acknowledgements The authors gratefully acknowledge the support of the Oxford Martin School, University of Oxford, who have funded this research. The authors would also like to thank Dr Davide Medici and Prof. Henrik Alfredsson for providing us with their experimental data. References. Ivanell, S., Sørensen, J.N., Mikkelsen, R., Henningson, D.: Analysis of numerically generated wake structures. Wind Energy 2, 63 80 (2009). 2. Launder B.E., Spalding, D.B.: The numerical computation of turbulent flows. Comput. Method. Appl. M. 3, 269 289 (974). 3. Medici, D., Alfredsson, P.H.: Measurements on a wind turbine wake: 3D effects and bluff body vortex shedding. Wind Energy 9, 29 236 (2006). 4. Medici, D., Alfredsson, P.H.: Personal communication (20). 5. Nishino, T., Willden, R.H.J.: Effects of 3-D channel blockage and turbulent wake mixing on the limit of power extraction by tidal turbines. Int. J. Heat Fluid Flow (202, in press). 6. Pope, S.B.: Turbulent Flows. Cambridge University Press, Cambridge, UK (2000). 7. Sagaut, P., Deck, S., Terracol, M.: Multiscale and Multiresolution Approaches in Turbulence. Imperial College Press, London, UK (2006).