You are on page 1of 93

uRANS-ALM Modelling of Vertical Axis Turbines

Zulkeefal Dar

1st September 2017

Supervised by:

Mr Georgios Deskos, Department of Earth Science & Engineering, Imperial College London

Professor Matthew Piggott, Department of Earth Science & Engineering, Imperial College London

A thesis presented to Imperial College London in partial fulfilment of the requirements for the

degree of Master of Science in Sustainable Energy Futures and for the Diploma of Imperial

College

Energy Futures Lab

Imperial College London

SW7 2AZ
Thesis
MSc Sustainable Energy Futures Imperial College London

Abstract
This thesis aims to assess unsteady Reynolds-averaged Navier Stokes equations- Actuator line
model’s (uRANS-ALM) validity as a computationally inexpensive high-fidelity tool to analyse
vertical axis turbines (VATs). A high solidity VAT at turbine rotor diameter Reynolds number
ReD=106 is used as a validation case. The ALM combines a modified conventional Blade Element
Theory (BET) model with Navier-Stokes based fluid flow models. Several “correction models”, like
dynamic stall, flow curvature, added mass and blade end effects, were incorporated in the
actuator line concept to capture the unsteady behaviour of the VAT more realistically. Validation
parameters included coefficient of power and drag of the turbine and mean velocity and turbulent
kinetic energy (TKE) in the near wake. 3D k- ε and k- ω SST uRANS simulations were performed
in two open source computational fluid dynamics (CFD) codes, OpenFOAM and Fluidity. The aim
of using multiple options was to investigate the most suitable turbulence closure model and CFD
code for uRANS-ALM implementation alongside establishing its validity. OpenFOAM is a widely
used, highly optimised code which has been utilized to analyse VATs using uRANS-ALM in the
past. So, an existing VAT ALM code was used for OpenFOAM, whereas, a horizontal axis turbine
(HAT) ALM code was modified to include VAT for Fluidity. Fluidity supports adaptive unstructured
meshing which automatically adjusts mesh during the simulation to concentrate it in the regions
where more resolution is needed producing high-resolution results efficiently. Simulation results
showed OpenFOAM to be better able to capture the contribution each correction model made
towards bringing the results closer to the experimental values. Dynamic stall was found to have
a very significant role in the correct prediction of the wake characteristics. However, the remaining
models were not found to affect the results significantly. Overall, OpenFOAM gave reasonable
performance and wake predictions. Quantitatively speaking, Fluidity produced poorer predictions,
except for turbulent kinetic energy (TKE) using k- ε turbulence model, owing mainly to
inappropriate implementation of dynamic stall and non-inclusion of struts in the model. However,
Fluidity gave much better qualitative results than OpenFOAM. Fluidity was able to capture
important physical phenomena like blade trailing edge vortex shedding and produced high-
resolution wake because of its adaptive meshing feature. In turbulence models, k- ω SST
performed better than k- ε in providing turbine and wake characteristics and capturing the physical
description of the flow using both CFD codes. However, TKE produced was an order of magnitude
lower than k- ε due to inappropriate initial value of ω chosen. Also, the ALM implementation
resulted in three to four orders of magnitude reduction in computational expense as compared to
blade resolved simulations. In general, the two turbulence models and CFD codes produced

i
Thesis
MSc Sustainable Energy Futures Imperial College London

satisfactory results in specific areas only, and no conclusions can be drawn regarding their overall
validity. Thus, this thesis is to be considered a first step in the comprehensive validation study of
uRANS-ALM for VATs, and future work needs to be done to provide a conclusive result.

ii
Thesis
MSc Sustainable Energy Futures Imperial College London

Acknowledgements
“I have no special talents. I am only passionately curious.”
Albert Einstein

I would first like to thank my supervisor, Georgios Deskos, for his patience in guiding me in the
use of simulation software; for insightful comments and valuable ideas; and for instant help that
he always provided when I was in need.
Secondly, I would like to express my sincere gratitude to my co-supervisor, Professor Matthew
Piggott, for his continuous support and guidance on my thesis; for his motivation and inspiration
towards the sustainable world; and for his invaluable experience and knowledge that he has
passed on to myself and others.
Personally, I would like to specially thank my family and friends, for providing me unfailing
support and encouragement throughout my study life in the UK. Finally, my gratitude goes to
Commonwealth Scholarship Commission in the UK who gave me an opportunity to study here.

iii
Thesis
MSc Sustainable Energy Futures Imperial College London

Table of Contents

Chapter 1. Introduction and Research Motivation .............................................................. 1


1.1 Wind and tidal energy as an instrument against climate change ....................................... 1
1.2 Revival of VATs .......................................................................................................................... 2
1.3 Motivation for research .............................................................................................................. 4
1.4 Objectives and chapter layout .................................................................................................. 5
Chapter 2. Theory.................................................................................................................. 8
2.1 VAT operation ............................................................................................................................. 8
2.2 Modern modelling approaches ................................................................................................. 9
2.3 Turbine parameterisation ........................................................................................................ 11
2.3.1 Actuator Disk (AD)............................................................................................................ 11
2.3.2 Actuator Surface (AS) ...................................................................................................... 12
2.3.3 Actuator Line (AL) ............................................................................................................ 12
2.3.3.1 Methodology .............................................................................................................. 13
2.3.3.2 Dynamic stall ............................................................................................................. 14
2.3.3.3 Added mass .............................................................................................................. 17
2.3.3.4 Flow curvature .......................................................................................................... 17
2.3.3.5 Blade end effects ...................................................................................................... 18
2.3.3.6 Projection on flow field ............................................................................................. 19
2.4 Turbulence modelling .............................................................................................................. 19
2.4.1 Direct Numerical Simulation (DNS) ............................................................................... 20
2.4.2 Large Eddy Simulation (LES) ......................................................................................... 20
2.4.3 (Unsteady) Reynolds Averaged Navier-Stokes (RANS/uRANS) Modelling ............ 21
2.4.3.1 Prandtl’s Mixing Length Theory (zero-equation).................................................. 23
2.4.3.2 k-model (one-equation) ........................................................................................... 23
2.4.3.3 k- ε model (two-equation) ........................................................................................ 26
2.4.3.4 k- ω model (two-equation) ...................................................................................... 26
2.4.3.5 k- ω SST model (two-equation).............................................................................. 27
2.5 Domain discretisation (Meshing)............................................................................................ 28
2.5.1 Finite Difference Method (FDM) ..................................................................................... 29
2.5.2 Finite Volume Method (FVM).......................................................................................... 29
2.5.3 Finite Element Method (FEM) ........................................................................................ 30

iv
Thesis
MSc Sustainable Energy Futures Imperial College London

2.5.4 Fluidity and adaptive meshing ........................................................................................ 31


Chapter 3. Modelling and Validation ...................................................................................35
3.1 Previous work ........................................................................................................................... 35
3.2 Experiment ................................................................................................................................ 36
3.2.1 Setup .................................................................................................................................. 36
3.2.2 Results and analysis ........................................................................................................ 39
3.3 OpenFOAM ............................................................................................................................... 40
3.3.1 Modelling ........................................................................................................................... 41
3.3.2 Meshing and verification.................................................................................................. 41
3.3.3 Results and analysis ........................................................................................................ 43
3.3.3.1 k- ε uRANS- ALM ..................................................................................................... 43
3.3.3.2 k- ω SST uRANS- ALM ........................................................................................... 46
3.4 Fluidity ........................................................................................................................................ 48
3.4.1 Modelling ........................................................................................................................... 49
3.4.2 Meshing and verification.................................................................................................. 49
3.4.3 Results ............................................................................................................................... 52
3.4.3.1 k- ε uRANS- ALM ..................................................................................................... 52
3.4.3.2 k- ω SST uRANS- ALM ........................................................................................... 57
3.5 Validation ................................................................................................................................... 61
3.5.1 Quantitative validation ..................................................................................................... 61
3.5.2 Qualitative validation ........................................................................................................ 64
Chapter 4. Conclusion .........................................................................................................69
4.1 Summary ................................................................................................................................... 69
4.2 Future Work............................................................................................................................... 70
References …………………………………………………………………………………………… ..72

v
Thesis
MSc Sustainable Energy Futures Imperial College London

List of Figures

Figure 1: Installed global capacity of renewable power ....................................................................... 2


Figure 2: Lift generation by a foil ............................................................................................................. 8
Figure 3: Vector diagram of VAT ............................................................................................................ 9
Figure 4: Lift (left) and drag (right) coefficients of foil over one turbine revolution for different λ . 9
Figure 5: Schematic diagram of numerical modelling ........................................................................ 10
Figure 6: Break-down of different numerical simulation techniques with the preferred methods
for this thesis highlighted in orange ....................................................................................................... 11
Figure 7: ALM model of 3-bladed turbine ............................................................................................ 13
Figure 8: Stages of dynamic stall ......................................................................................................... 15
Figure 9: Variation of Cl, Cd and Cm with AoA during dynamic stall ................................................. 15
Figure 10: CN for a NACA 0012 oscillatory test (α=15 deg+10 deg sinωt) with Orig. LB DS model
and New Sheng et al. model ................................................................................................................. 16
Figure 11: General features of commonly used turbulence models ............................................... 19
Figure 12: Velocity resolution based on different turbulence models, adapted from .................... 20
Figure 13: Representation of flow velocity as a sum of mean and fluctuating component ........... 21
Figure 14: A plot showing function being approximated by linear combination of shape functions
.................................................................................................................................................................... 30
Figure 15: Some common 2D and 3D linear and quadratic elements along with their nodes ..... 31
Figure 16: Mesh before and after h adaptivity ..................................................................................... 32
Figure 17: Mesh before and after r adaptivity ...................................................................................... 33
Figure 18: Mesh before and after p adaptivity ..................................................................................... 33
Figure 19: Turbine model dimensions and tow tank experimental setup ....................................... 37
Figure 20: Wake measurement coordinate system and measurement locations in meters ......... 38
Figure 21: Mean power (left) and drag (right) coefficients plotted versus tip speed ratio based on
experimental values ................................................................................................................................ 39
Figure 22: Mean velocity profile (left) and TKE (right) in the near-wake at x/D=1 and z/H=0
based on experimental data, adapted from ........................................................................................ 39
Figure 23: Mean velocity (top) and TKE (bottom) contours at x/D=1 based on experimental
results. Vectors in the top figure are cross-stream and vertical velocities; contours are stream-
wise velocity. The view is looking upstream, with the turbine frontal area indicated by the solid

vi
Thesis
MSc Sustainable Energy Futures Imperial College London

lines. The upper half of the turbine is shown with horizontal plane through z/H=0 considered as
the plane of symmetry (Bachant & Wosnik, 2014) .............................................................................. 40
Figure 24: Cutaway of 3D mesh used in OpenFOAM ........................................................................ 42
Figure 25: Comparison of mean velocity (left) and TKE (right) profiles at x/D=1 and z/H=0 based
on experimental and OpenFOAM k- ε uRANS- ALM using all correction models ......................... 43
Figure 26: Mean velocity (top) and TKE (bottom) contours at x/D=1 using OpenFOAM k- ε
uRANS- ALM............................................................................................................................................. 44
Figure 27: Contribution of the correction models towards mean velocity profile using
OpenFOAM k- ε uRANS- ALM, DS=Dynamic stall, FC=Flow curvature, EE=End effects ........... 45
Figure 28: Contribution of the correction models towards TKE profile using OpenFOAM k- ε
uRANS- ALM, DS=Dynamic stall, FC=Flow curvature, EE=End effects ......................................... 45
Figure 29: Comparison of mean velocity (left) and TKE (right) profiles at x/D = 1 and z/H = 0
based on experimental and OpenFOAM k- ω uRANS- ALM using all correction models ............ 46
Figure 30: Mean velocity (top) and TKE (bottom) contours at x/D=1 using OpenFOAM k- ω
uRANS- ALM............................................................................................................................................. 47
Figure 31: Contribution of the correction models towards mean velocity profile using
OpenFOAM k- ω uRANS- ALM, DS=Dynamic stall, FC=Flow curvature, EE=End effects .......... 48
Figure 32: Contribution of the correction models towards TKE profile using OpenFOAM k- ω
uRANS- ALM, DS=Dynamic stall, FC=Flow curvature, EE=End effects ......................................... 48
Figure 33: Cutaway of the initial (left) and final adapted (right) 3D mesh used in Fluidity............ 50
Figure 34: Cp and CD spatial resolution sensitivity results for Fluidity k- ε uRANS- ALM............. 50
Figure 35: Mean velocity (left) and TKE (right) spatial resolution sensitivity results for Fluidity k-
ε uRANS- ALM.......................................................................................................................................... 50
Figure 36: Cp and CD spatial resolution sensitivity results for Fluidity k- ω uRANS- ALM ........... 51
Figure 37: Mean velocity (left) and TKE (right) spatial resolution sensitivity results for Fluidity k-
ω uRANS- ALM ........................................................................................................................................ 51
Figure 38: Comparison of mean velocity (left) and TKE (right) profiles at x/D = 1 and z/H = 0
based on experimental and Fluidity k- ε uRANS- ALM using all correction models ...................... 52
Figure 39: Mean velocity (top) and TKE (bottom) contours at z/H=0, viewing from top, using
Fluidity k- ε uRANS- ALM ....................................................................................................................... 53
Figure 40: Mean velocity (top) and TKE (bottom) contours at x/D=1 using Fluidity k- ε uRANS-
ALM ............................................................................................................................................................ 54

vii
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 41: Contribution of the correction models towards Cp and CD using Fluidity k- ε uRANS-
ALM ............................................................................................................................................................ 55
Figure 42: Variation of AoA and the corresponding Cl vs turbine revolutions ................................ 55
Figure 43: Contribution of the correction models towards mean velocity profile using Fluidity k- ε
uRANS- ALM, DS=Dynamic stall, FC=Flow curvature, AM=Added mass ...................................... 56
Figure 44: Contribution of the correction models towards TKE profile using Fluidity k- ε uRANS-
ALM, DS=Dynamic stall, FC=Flow curvature, AM=Added mass ...................................................... 56
Figure 45: Comparison of mean velocity (left) and TKE (right) profiles at x=D = 1 and z=H = 0
based on experimental and Fluidity k- ω uRANS- ALM using all correction models ..................... 57
Figure 46: Mean velocity (top) and TKE (bottom) contours at z/H=0 using Fluidity k- ω uRANS-
ALM ............................................................................................................................................................ 58
Figure 47: Mean velocity (top) and TKE (bottom) contours at x/D=1 using Fluidity k- ω uRANS-
ALM ............................................................................................................................................................ 59
Figure 48: Contribution of the correction models towards Cp and CD using Fluidity k- ω uRANS-
ALM ............................................................................................................................................................ 60
Figure 49: Contribution of the correction models towards mean velocity profile using Fluidity k-
ω uRANS- ALM, DS=Dynamic stall, FC=Flow curvature, AM=Added mass .................................. 60
Figure 50: Contribution of the correction models towards TKE profile using Fluidity k- ω uRANS-
ALM, DS=Dynamic stall, FC=Flow curvature, AM=Added mass ...................................................... 60
Figure 51: Comparison of mean velocity contours at x/D=1 using OpenFOAM k- ε uRANS-ALM
(top-left), Fluidity k- ε uRANS-ALM (top- right), OpenFOAM k- ω uRANS-ALM (centre-left),
Fluidity k- ω uRANS-ALM (centre-right) and experimental (bottom) results ................................... 65
Figure 52: Comparison of TKE contours at x/D=1 using OpenFOAM k- ε uRANS-ALM (top-left),
Fluidity k- ε uRANS-ALM (top- right), OpenFOAM k- ω uRANS-ALM (centre-left), Fluidity k- ω
uRANS-ALM (centre-right) and experimental (bottom) results ......................................................... 66
Figure 53: Comparison of mean velocity contours at z/H=0 using OpenFOAM k- ε uRANS-ALM
(top-left), Fluidity k- ε uRANS-ALM (top- right), OpenFOAM k- ω uRANS-ALM (bottom-left) and
Fluidity k- ω uRANS-ALM (bottom-right) results ................................................................................. 67
Figure 54: Comparison of TKE contours at z/H=0 using OpenFOAM k- ε uRANS-ALM (top-left),
Fluidity k- ε uRANS-ALM (top- right), OpenFOAM k- ω uRANS-ALM (bottom-left) and Fluidity k-
ω uRANS-ALM (bottom-right) results.................................................................................................... 68

viii
Thesis
MSc Sustainable Energy Futures Imperial College London

List of Tables
Table 1: Values of coefficients used in Sheng et al. dynamic stall model ..................................41
Table 2: Mean of absolute error in simulation results with respect to experimental results ........63
Table 3: Standard deviation of absolute error in simulation results with respect to experimental
results .......................................................................................................................................63
Table 4: Mean of percentage relative error in simulation results with respect to experimental
results .......................................................................................................................................63
Table 5: Standard deviation of percentage relative error in simulation results with respect to
experimental results ..................................................................................................................63

ix
Thesis
MSc Sustainable Energy Futures Imperial College London

Nomenclature

Symbols

Cl coefficient of lift
CD coefficient of turbine’s drag/thrust
Cp coefficient of power
Urel velocity of the fluid relative to the moving blade
FT tangential force on blades
FN tangential force on blades
α angle of attack
𝜃 azimuthal angle
Cd blade drag
λ tip speed ratio
U, u fluid velocity
𝑝 pressure
𝜌 density
𝑣 kinematic viscosity
Cm coefficient of pitching moment
Cn coefficient of normal force
Ct tangential force
𝑈𝑛 normal component of velocity
𝑈𝑡 tangential component of velocity
Aelem planform area of a blade element
𝑐 blade chord
𝑈∞ free-stream velocity
𝐶𝑛𝐴𝑀 coefficient of normal force incorporating added mass
𝐶𝑡𝐴𝑀 coefficient of tangential force incorporating added mass
𝐶𝑚𝐴𝑀 coefficient of pitching moment incorporating added mass
𝛽𝑝 blade pitch angle
𝑉𝑎𝑏𝑠 magnitude of local inflow velocity at the blade
𝜃𝑏 blade azimuthal position
𝛽 direction of the inflow velocity

x
Thesis
MSc Sustainable Energy Futures Imperial College London

𝛺 turbine’s angular velocity


𝑟 the blade element radius
x the fractional distance of the mounting point of a blade from the AL node
𝑉𝑟𝑒𝑓 reference flow velocity for calculating angle of attack
𝑆 blade span length
𝜂 spherical Gaussian function
𝜖 Gaussian width parameter
𝑣𝑇 eddy or turbulent viscosity
𝑘 turbulent kinetic energy
𝛿𝑖𝑗 Kronecker delta
̅
𝑆𝑖𝑗 mean rate of strain tensor
𝑙𝑚 mixing length
P rate of production of turbulence
ε rate of its dissipation
𝜎𝑘 turbulent Prandtl number
ω specific dissipation rate
A area of the control surface
h characteristic length of the elements
Np nodes of previous mesh
Nn nodes of the new mesh
D turbine diameter
Nx number of cells in stream-wise direction
∆t time step
R turbine radius
αds0 dynamic stall initiation angle
The symbols have been arranged in the order of appearance

Super and sub scripts

𝑥̇ rate of change with time


𝑥 mean/ average component
𝑥́ fluctuating component
𝑥⃗ vector quantity
𝑥𝑖,𝑗 ith, jth component

xi
Thesis
MSc Sustainable Energy Futures Imperial College London

𝑥𝑆𝑖𝑚 simulation value


𝑥𝐸𝑥𝑝 experimental value
Where 𝒙 denotes any variable

Acronyms

ABL Atmospheric Boundary Layer


ADM Actuator Disk Theory
AD-R Actuator Disc with Rotation
ADV Acoustic Doppler Velocimeter
ALM Actuator Line Model
ALM Actuator Line Model
AoA Angle of Attack
AS Actuator Surface
ASSM Actuator Swept Surface Model
BEMT Blade Element Momentum Theory
BET Blade Element Theory
CFD Computational Fluid Dynamics
CFL Courant Friedrichs Lewy
CFT Cross Flow Turbine
DDES Delayed Detached Eddy Simulation
DMST Double Multiple Stream Tube
DNS Direct Numerical Simulation
EU European Union
FC Flow curvature
FDM Finite Difference Method
FEM Finite Element Method
FVM Finite Volume Method
GHG Green House Gases
HAT Horizontal Axis Turbine
HATT Horizontal Axis Tidal Turbine
HAWT Horizontal Axis Wind Turbine
LB Leishman Beddoes
LCOE Levelised Cost of Electricity

xii
Thesis
MSc Sustainable Energy Futures Imperial College London

LES Large Eddy Simulation


M Mean
MPI Message Passing Interface
NACA National Advisory Committee for Aeronautics
NRC National Research Council
NS Navier Stokes
Re REynolds number
RSMs Reynold Stress Models
SA Spalart Allmaras
SD Standard Deviation
SGS Sub Grid Scale
SNL Sandia National Labs
SST Shear Stress Transport
TKE Turbulent Kinetic Energy
TSR Tip Speed Ratio
uRANS unsteady Reynold- Averaged Navier-Stokes
VAT Vertical Axis Turbine
VATT Vertical Axis Tidal Turbine
VAWT Vertical Axis Wind Turbine

xiii
Thesis
MSc Sustainable Energy Futures Imperial College London

Chapter 1. Introduction and Research Motivation

1.1 Wind and tidal energy as an instrument against climate change


The Kyoto Protocol (1997) was a milestone in human history- it held anthropogenic CO2 emissions
responsible for climate change and marked the beginning of humanity’s efforts to mitigate climate
change (UNFCCC, 1998). Climate change has resulted in severe and unpredictable weather
patterns, like droughts and floods, resulting in billions of dollars in economic losses to several
countries every year (Kreft, Eckstein & Melchios Inga, 2017). It also threatens food security by
significantly affecting crop yields (Gregory, Ingram & Brklacich, 2005). Realizing this, in Kyoto
Protocol and later on in Paris Agreement, more than 170 nations around the globe set GHG
emission reduction targets for themselves, e.g. the UK aims to decrease its GHG emissions by
80%, as of 1990 level, by 2050 (Committee on Climate Change, 2008).

One way of mitigating climate change is shifting electricity generation from fossil fuels to
renewable sources. Figure 1 shows the breakdown of global renewable power installed capacity.
As it can be seen, wind is the second biggest contributor to renewable electricity generation and
has been growing exponentially. Tidal energy, though in its development phase, has a promising
future because of its predictable nature and vast resource present (Lewis, Neill, Robins, et al.,
2015). Development of tidal stream1 technology has also resolved the environmental concerns
which have haunted tidal range2 technology (Thomas, Khan & Gillman, 2016). According to an
estimate, harnessing just 20% of global wind and tidal potential can provide seven times the global
electricity demand (Jacobson & Delucchi, 2011; Archer & Jacobson, 2005).

Though wind and tidal energies are fundamentally different in the sense of their source, they both
use the same technology to extract power. Indeed, wind and tidal turbine technologies are similar,
with the later been slightly modified to deal with the higher rotor loads (thrust) and the corrosive
nature of the marine environment. So, deeper understanding of such devices can help in
extracting both, wind and tidal, energies more efficiently and can help many countries in tackling
climate change.

_____________________________________________________________________________________
1 Tidal stream involves extracting kinetic energy from the horizontal flow of water currents resulting from
the tidal pull of the moon. It does not involve construction of any reservoir.
2 Tidal range involves constructing a barrage or lagoon impoundment to allow development of height

difference in water on the either side during tidal cycle. The water at the higher side is then allowed to
flow to the lower side through turbines, generating power.

1
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 1: Installed global capacity of renewable power (IRENA, 2015)

1.2 Revival of VATs


VATs saw extensive research in the 1970s to 1990s mainly by the National Research Council
(NRC) of Canada and the Sandia National Labs (SNL) of the USA, before being abandoned in
favour of HATs. This is because vertical axis turbines (VATs) cost twice as much as horizontal
axis turbines (HATs) (Sutherland, Berg & Ashwill, 2012), as well as, suffer from more fatigue,
reducing their life to 3.5 years only as compared to 20+ years of HATs life (Veers, 1983).

Nevertheless, VATs have some inherent advantages which make them a better candidate than
HATs for the following applications:

Compact onshore wind farms

Vertical axis wind turbines (VAWTs) have a rectangular cross-section and better wake recovery
properties- a wake can recover to 95% of upstream velocity at four diameters distance as opposed
to 16 diameter distance required for horizontal axis wind turbines (HAWTs) (Kinzel, Mulligan &
Dabiri, 2012), which can result in more compact placement of onshore wind farms. Also,
employing pairs of counter-rotating VAWTs in a wind farm can lead to an order of magnitude
higher power extraction as compared to a HAWT farm of the same size (Dabiri, 2011). Due to
their noise and visual impact, rejection rates for the approval of onshore wind farms increased to

2
Thesis
MSc Sustainable Energy Futures Imperial College London

57% in 2015 from 21% in 2008 in UK (Tait, 2015). VAWT farms can be smaller than HAWT ones
and hence can potentially facilitate easier planning consent.

Floating offshore wind farms

Saturation of shallow waters with wind farms have necessitated the development of floating wind
turbine technology to harness wind energy present above deep waters. VAWTs also have the
edge over HAWTs in this case as they can have the generator and the gear box at the base
instead of the hub. This lowers the centre of gravity resulting in more stable floating turbines
(Shires, 2013).

250 m diameter turbines and beyond

Wind turbines are aimed to reach a size of 250m rotor diameter in the future to improve their
efficiency and levelised cost of electricity (LCOE). The EU supported UpWind research project
concluded that the weight of such large blades of HAWT would undergo significant bending under
gravity (Sieros, Chaviaropoulos, Sørensen, et al., 2012). VAWTs can be used for such large
scales as their blades do not undergo gravitational bending.

Urban wind turbines

A key advantage of VAWTs is that they can accept wind from any direction and hence are an
ideal candidate for the urban environment wind projects like rooftop wind turbines. They can also
be used in other places which do not have any prevailing wind direction (Balduzzi, Bianchini,
Carnevale, et al., 2012).

Tidal Fencing

Tidal fencing is achieved by closely spacing the tidal turbines so that water does not easily find a
path to escape around a turbine and must pass through the closely spaced turbines. Due to their
rectangular cross section, vertical axis tidal turbines (VATTs) can be packed more tightly in a tidal
channel and hence offer more blockage3 and thus a higher power extraction than horizontal axis
tidal turbines (HATTs), which have circular cross-section.

_____________________________________________________________________________________
3Blockage is defined as the ratio of rotor swept area of all tidal turbines in a row to channel cross-section
area

3
Thesis
MSc Sustainable Energy Futures Imperial College London

1.3 Motivation for research


In the light of a potential preference for VATs over HATs for the application areas mentioned
above and continued growth in renewable energy requirement, an increase in VATs demand is
guaranteed in the future. Structurally stronger and more cost-effective VATs are required to
realise this potential. This requires replicating the real life fluid-turbine interactions more
accurately during their design phase to have a more realistic estimate of loads that will be
experienced by them during operation. The design of VATs can be broken down into two
categories- experimental setups and computation modelling, each having its own merits and
shortcomings. A well-designed experiment can replicate the actual environment exactly and
hence can be very accurate. However, conducting experimental testing on real turbine blades,
which can be as long as 80 m, can be very expensive and time-consuming. One way this issue
is resolved is by scaling down the actual turbine and working on the model in the laboratory, but
this has implications for the accuracy of the laboratory results when applying to the actual turbine.

On the other hand, design approaches based on computational modelling has the potential to be
accurate, as well as, economical and quick at the same time. However, this aim is yet to be
achieved for VATs. For example, 3D blade-resolved Direct Numerical Simulation (DNS) is
extremely reliable but a 108 fold increase in the computation power of the fastest super computer
is required for DNS of the high Reynolds number (Re) flows experienced by VATs (Kurien &
Taylor, 2005). On the other hand, classic Blade Element Momentum Theory (BEMT) can perform
turbine array simulations on a personal computer but produces unreliable results (Badreddinne,
Ali & David, 2005).

Designing VATs based on the computational modelling approach involves integration of two sub-
models- a model for the turbine and a model to represent the fluid behaviour. An effective way of
reducing the computational expense is to parameterize the turbine i.e. represent turbine loads
using simpler models instead of via actual interaction of a 3 D turbine with the fluid. The
foundation for turbine/rotor parameterization was laid by Froude (1911) in which he represented
the rotor by a uniformly loaded permeable Actuator Disk (AD) placed normal to the flow direction.
Thrust on the rotor was represented with a pressure drop across the disc, calculated using energy
and momentum balance between the disk and fluid. The elementary AD theory was substantiated
by vortex theories of Joukowsky (1912) and Betz (1921) and resulted in the formulation of Betz

4
Thesis
MSc Sustainable Energy Futures Imperial College London

limit4-the upper limit of turbine efficiency. AD theory results applied to turbines with an infinite
number of blades only and did not take aerofoil shape into consideration, therefore could not be
used to design actual turbines. Glauert (1935) combined the works of Drzewiecki SK (1910) and
Betz (1919) to produce BEMT accounting for the effects of air foil shape and a finite number of
blades. Paraschivoiu(1988) later on developed Double Multiple Stream Tube (DMST), which was
a variation of BEMT suitable for VATs. BEMT relies on fluid inflow conditions to calculate the
loads experienced by a turbine and over the years many variations of BEMT have been developed
to take more and more fluid flow phenomenon into account to improve load estimates. The
actuator line (AL) concept, developed by Sorensen & Shen (2002) as an extension of classic
BEMT, represents turbine blades as rotating lines to compute the time varying loads on a turbine.
Recent research on Actuator Line Models (ALMs), involving coupling of AL concept with advanced
Computational Fluid Dynamics (CFD) techniques like unsteady Reynolds Averaged Navier-
Stokes (uRANS) and Large Eddy Simulation (LES) to predict HATs and VATs performance, has
produced results which are in good agreement with experiments (Martínez-Tossas et al. 2015;
Draper et al. 2016; Mendoza et al. 2016; Bachant et al. 2016; Baba-Ahmadi & Dong 2017).
Furthermore, Bachant et al. (2016) showed that ALM provides two orders of magnitude of
computational savings over blade-resolved simulation and so does uRANS over LES. Advanced
meshing techniques like unstructured adaptive meshing can modify the mesh during a simulation
to capture flow behaviours with higher resolution or provide similar results in less time by allowing
bigger cells and time steps to be used in simulations (Abolghasemi, Piggott, Spinneken, et al.,
2016). Hence, combining all these features into a mesh-adaptive uRANS-AL Model for VATs can
potentially provide an accurate and fast analysis tool for stand alone, as well as, arrays of VATs.

1.4 Objectives and chapter layout


Based on the above reasoning this thesis is aimed at formulating and validating mesh-adaptive
uRANS-ALM as a computationally inexpensive modelling technique which can significantly
reduce the uncertainty during the design phase of VATs. The steps required to fulfil the above
mentioned objectives are as follows:

1. Perform a literature review of the numerical modelling techniques developed so far for
VATs to ensure uRANS-ALM is the best potential candidate to be the inexpensive and
accurate technique.

_____________________________________________________________________________________
4No turbine can harness more than 59.8% of the kinetic energy present in the wind (or any other fluid in
general)

5
Thesis
MSc Sustainable Energy Futures Imperial College London

2. Find experimental data for a VAT which would be compared with the modelling results to
draw conclusions regarding the validity of uRANS-ALM. It is to be ensured that the data
is comprehensive and refined enough to make an in depth comparison.
3. Select a suitable CFD software having following features:
 Widely validated and optimised to have fast solution times
 Incorporates mesh additivity
 Flexibility to incorporate and alter code easily
 Troubleshooting assistance
4. Model and analyse the VAT selected in step 2. This step involves implementation and
refinement of the uRANS-ALM. Literature is also sought at this stage to facilitate
implementation of the model in the CFD software. The model is verified by ensuring that
as temporal and spatial step size is refined, solution converges to a specific value. The
important turbine performance parameters to be analysed include the turbine’s coefficient
of power and thrust/drag, Cp and CD. Wake characteristics like mean stream velocity and
turbulent kinetic energy will also be examined to analyse wake behaviour and recovery.
5. Compare the gathered results with experimental results. The turbine performance and
wake characteristics will be compared with the data, both qualitatively and quantitatively.
Possible explanations for the deviation of results from the experimental values will be
sought. Based on the analysis and reasons of inaccuracy, validity of uRANS-ALM as a
highly accurate predictive tool will be determined.

The work done to accomplish the objectives mentioned above is presented in this thesis as:

Chapter 2 Theory: VAT operation principle is introduced at the beginning. This is followed by the
description of modern modelling techniques used for VAT analysis with a focus on uRANS-ALM.
Meshing techniques commonly employed by CFD solver are then described. Finally, the Fluidity
code is introduced along with its adaptive meshing feature.

Chapter 3 Model Validation: First of all, recent work on VAT modelling is presented.
Subsequently, a description of the experimental setup of a VAT along with the results is given.
The results of 3D k- ε and k- ω SST uRANS simulations in OpenFOAM and Fluidity, the two CFD
codes used, are then analysed. This is followed by quantitative and qualitative validation of ALM.
Possible explanations for the discrepancies between the simulation and experimental results are
also presented.

6
Thesis
MSc Sustainable Energy Futures Imperial College London

Chapter 4 Conclusion: Key findings and summary of the overall research in presented in this
chapter. Potential future work to be built upon this study is also investigated.

It is important to highlight certain limitations which might have affected the scope of this research.
Being a Master’s degree thesis, the main limitation confronted was the lack of time. Also, as a
result of having just elementary knowledge of super-computers, use of super-computer was
limited to a single node to prevent issues arising from poor communication between different
nodes. Hence the resource usage could not be optimised leading to higher simulation times.
Finally, this thesis is to be considered as a part of an ongoing research project, and the tasks
which could not be completed due to these limitations are listed as potential future work at the
end of the thesis.

7
Thesis
MSc Sustainable Energy Futures Imperial College London

Chapter 2. Theory

This chapter presents an overview of the modern modelling techniques for VATs and their
implementation in CFD solvers. The principle of VAT operation is explained first. Subsequently,
a description of the popular turbine parameterisation methods and turbulence models with the
focus on methodology and advantages of ALM and uRANS- the two sub-models of the chosen
modelling technique, is presented. This is followed by an account of the meshing techniques
commonly used by numerical solves. Finally, the Fluidity CFD code is introduced along with its
feature of adaptive meshing.

2.1 VAT operation


VATs are classified into two main types- lift and drag type. Lift type VATs are preferred over drag
type because of their higher efficiency (Castelli & Benini, 2011). The lift generation mechanism of
an air/hydrofoil has been a subject of controversy for decades. Some of the incorrect theories
used to explain lift are the longer path, skipping stone, venture theories (NASA, 2015). In reality,
fluid is deflected downward by a foil, and upward lift is generated as per Newton’s third law. As
fluid molecules are free to move, they tend to follow the foil profile as shown in Figure 2. Thus, a
foil is either given a particular shape or kept at a certain angle to produce net downward deflection
of fluid resulting in lift generation.

Figure 2: Lift generation by a foil (NASA, 2015)

Figure 3 shows the velocity and force vectors acting on a VAT blade during its rotation. Drag acts
in the direction of the relative fluid velocity Urel, the velocity of the upstream fluid relative to the
moving blade, whereas, lift acts perpendicular to it. The sum of the components of lift and drag,
acting tangential FT and normal FN to the direction of the blade motion, are responsible for torque
generation and stresses in the blade, shaft bearings, etc. The main difference between a HAT
and a VAT’s operation is that in a VAT, the angle of attack (AoA) α, the angle between relative
fluid velocity and the chord of the aerofoil, and the torque produced by the turbine varies
continuously with azimuthal angle 𝜃 during the rotation. Thus, the motion of VATs can be said to

8
Thesis
MSc Sustainable Energy Futures Imperial College London

be more “dynamic” than the HATs when the upstream velocity is uniform and no yaw misalignment
exists. Figure 4 shows how coefficient of lift and drag, Cl and Cd, vary during a VAT’s revolution
for different tip speed ratios (λ).

Figure 3: Vector diagram of VAT (Ferreira, van Bussel, van Kuik, et al., 2011)

Figure 4: Lift (left) and drag (right) coefficients of foil over one turbine revolution for different λ (Alaimo, Esposito,
Messineo, et al., 2015)

2.2 Modern modelling approaches


Since the availability of high computation power, computational fluid dynamics (CFD) has become
the preferable option for turbine design. These numerical techniques involve combining a turbine
and a fluid flow model in an iterative solver to compute a turbine’s performance and wake

9
Thesis
MSc Sustainable Energy Futures Imperial College London

characteristics. Since flows around a turbine have Re>106, the flow is highly turbulent5 which
necessitates the use of turbulence models, which are based on mass continuity and Navier-
Stokes (NS) equations, to be used as the fluid flow model. The equations are as follows:

𝜕𝑢𝑗
∑ = 0, 𝑓𝑜𝑟 𝑗 = 1,2,3 (𝑚𝑎𝑠𝑠 𝑐𝑜𝑛𝑡𝑖𝑛𝑢𝑖𝑡𝑦) (1)
𝜕𝑥𝑗
𝑗

𝜕𝑢𝑖 𝜕𝑢𝑖 1 𝜕𝑝 𝜕 2 𝑢𝑖 1
+ ∑ 𝑢𝑗 + − 𝑣 ∑ 2 − 𝐹𝑇𝑥𝑖 = 0, 𝑓𝑜𝑟 𝑖, 𝑗 = 1,2,3 (𝑁𝑆) (2)
𝜕𝑡 𝜕𝑥𝑗 𝜌 𝜕𝑥𝑖 𝜕𝑥 𝑗 𝜌
𝑗 𝑗

Where u1, u2, u3 are x, y, z (represented by x1, x2, x3) components of fluid velocity vector, 𝑝 is the
fluid pressure, 𝜌 is density of the fluid, 𝑣 is kinematic viscosity and 𝐹𝑇 momentum source term.

By converting the second term into its conservative form6, Eq(2) can be rewritten as:

𝜕𝑢𝑖 𝜕𝑢𝑗 𝑢𝑖 1 𝜕𝑝 𝜕 2 𝑢𝑖 1
+∑ + − 𝑣 ∑ 2 − 𝐹𝑇𝑥𝑖 = 0 (3)
𝜕𝑡 𝜕𝑥𝑗 𝜌 𝜕𝑥𝑖 𝜕𝑥 𝑗 𝜌
𝑗 𝑗

NS equation, in effect, is Newton’s second law applied to fluids where first two terms represent
the inertial forces and third, fourth and fifth term represents pressure, viscous and external body
force (momentum source), respectively, in Eq (3). In the case of turbine simulation, this external
body force is the force exerted by the turbine on the fluid. Because of the high computational
expense associated with blade-resolved simulations, turbine parameterisation is used to calculate
the turbine loads; the reaction of which is exerted on the fluid as the momentum source term in
NS equation. As NS equation cannot be solved analytically for such complex cases, the fluid
domain is discretised, and numerical solvers are used to solve the flow field iteratively. The
interaction of two sub-models is shown in Figure 5.

Figure 5: Schematic diagram of numerical modelling

_____________________________________________________________________________________
5 “Turbulence or turbulent flow is a flow regime in fluid dynamics characterized by chaotic changes in
pressure and flow velocity” (Batchelor, 1967).
6 A term is called conservative if it is represented as the divergence of a flux

10
Thesis
MSc Sustainable Energy Futures Imperial College London

Turbulence and turbine parameterisation models available for VATs are shown in Figure 6.

VAT Modelling

Turbulence Turbine
Modelling Parameterisation

DNS AD/AD-R

LES AS

RANS/uRANS AL

Figure 6: Break-down of different numerical simulation techniques with the preferred methods for this thesis
highlighted in orange

These models are now discussed with emphasis on uRANS and AL methodology and why they
are potentially better options than the other models.

2.3 Turbine parameterisation


2.3.1 Actuator Disk (AD)
As discussed in the first chapter, the AD concept represents a turbine as an infinitely thin and
porous actuator disk, having diameter equal to that of the turbine, which exerts a constant force
over the entire rotor swept area. Original AD represented the effect of turbine on fluid using thrust
force only. However, turbines also impart angular momentum to the fluid. So, a rotating actuator
disc (AD-R) concept was developed to take both the axial and angular momentum imparted by
the turbine into account.

Since the turbine is represented by a solid disk and does not take rotor parameters into account,
the AD concept is not used to compute turbine performance characteristics. Bachant & Wosnik
(2016) also showed that ADM does not predict the near-wake of VATs, characterised by the
turbine geometry, with precision.

11
Thesis
MSc Sustainable Energy Futures Imperial College London

2.3.2 Actuator Surface (AS)


Dobrev, Massouh & Rapin (2007) & Shen, Sørensen & Zhang (2007) developed the AS concept
which is similar to AL but instead of representing turbine using one-dimensional lines, AS uses
two-dimensional surfaces to represent turbine loads. The turbine blade is divided into elements
along its span and chord, each of which is represented by a force calculated using air foil-
coefficients. Using this surface representation, flow over the chord can be simulated more
accurately as compared to AL. However, use of AS technique necessitates the use of very fine
mesh around the blades to capture the flow correctly thus resulting in higher computational
expense. Moreover, AL can also be adjusted to distribute the turbine loads over a width
corresponding to its blade chord and with this done, AS was shown to offer only a slight
improvement over AL (Shen, Sørensen, Zhang, et al., 2007).

2.3.3 Actuator Line (AL)


Sorensen & Shen (2002) developed the AL concept as a modification of Blade Element Theory
(BET) so that it could be coupled with turbulence models to capture the time dependent behaviour
of turbines. Traditionally, BET, a turbine parameterisation technique, has been used with
momentum theory, forming BEMT, to find forcing on a turbine. This integrated model has been
quite successful for HATs and is among the most common engineering tools for computation of
loads acting on wind/tidal turbines (Madsen, Mikkelsen, Øye, et al., 2007). Though HATs do
experience some level of unsteady loading because of pitching, yawing and atmospheric
boundary layer (ABL) effects, etc., but in general, either the magnitude or the number of cycles of
these variations is small enough to classify them as quasi-static machines. Hence, BEMT, which
provides a good estimate for steady state conditions, can be employed to compute their
performance reliably. It is to be noted here that BEMT only provides good estimate for HAT
performance and not that for the associated wake. Accurate prediction of wake characteristics
requires modelling of fluid interaction within the wake, as well as, with the surrounding fluid. BEMT
is completely devoid of such a model and hence does not predict wake properties accurately even
for HATs. In the case of VATs, which experience unsteadiness consistently, BEMT is unable to
provide even performance characteristics reliably. As VAT blade rotates, it experiences a
constantly changing angle of attack. Moreover, each blade has to pass through the wake of the
other blades. These unsteady effects and some other phenomena unique to VATs called for the
up-grading of classic BET to AL technique. The AL methodology is now explained followed by the
description of some of the effects which make classic BEMT unsuitable for VATs, and the models
used to incorporate them within the AL concept.

12
Thesis
MSc Sustainable Energy Futures Imperial College London

2.3.3.1 Methodology
The basic formulation of AL is similar to classical BET. The turbine blades are divided into a
number of elements, each represented by a point at their quarter-chord position. These points are
assigned lift, drag and pitching moment coefficients, Cl, Cd and Cm, corresponding to the lift, drag
and pitching moment experienced by the associated elements. This yields actuator lines, through
the combination of the quarter-chord dots, representing the turbine blades loading shown in
Figure 7.

Figure 7: ALM model of 3-bladed turbine (Troldborg & Sørensen, 2014)

Cl, Cd and Cm of an aerofoil are pre-computed experimental or DNS band values and are a function
of inflow conditions, i.e. AoA and chord Reynolds number (Rec). The velocity triangle shown in
Figure 3 is used to compute the local AoA and Urel for each element.

𝑈𝑛
𝛼 = tan−1 (4)
𝑈𝑡

𝑈𝑟𝑒𝑙 = √𝑈𝑛 2 + 𝑈𝑡 2
(5)

where 𝑈𝑛 and 𝑈𝑡 are normal and tangential components of relative velocity, respectively.

Rec can then be computed as:

𝑈𝑟𝑒𝑙 𝑐
𝑅𝑒 =
𝜈 (6)

Where 𝑐 is the blade chord and 𝜈 is kinematic viscosity of fluid.

Once AoA and Rec are known, aerofoil coefficients can be found using 2-D profile look-up tables.
Sheldahl & Klimas (1981) is a popular database for these coefficients used for VATs. The blade
element’s normal force F𝑁 , tangential force F𝑇 and pitching moment M can then be calculated as:

13
Thesis
MSc Sustainable Energy Futures Imperial College London

1
F𝑁 = 𝜌𝐴𝑒𝑙𝑒𝑚 C𝑛 𝑈𝑟𝑒𝑙 (7)
2
1
F𝑇 = 𝜌𝐴𝑒𝑙𝑒𝑚 C𝑡 𝑈𝑟𝑒𝑙 (8)
2
1
M = 𝜌𝐴𝑒𝑙𝑒𝑚 𝑐C𝑚 𝑈𝑟𝑒𝑙
2 (9)

Where,

C𝑛 = C𝑙 cos 𝛼 + C𝑑 sin 𝛼 (10)

C𝑡 = −C𝑙 sin 𝛼 + C𝑑 cos 𝛼


(11)

Where C𝑛 and C𝑡 are coefficients of normal and tangential force, respectively, and 𝐴𝑒𝑙𝑒𝑚 is the
planform area of a blade element.

After this, the VAT specific effects not captured by BET are incorporated in ALM.

2.3.3.2 Dynamic stall


Description

This is a transient stall effect that can result in unsteady aero/hydrodynamic forces being
produced that are considerably in excess of what would be expected or predicted under steady
(static) conditions (Leishman, 2002). VATs, by virtue of their operation, undergoes large cyclic
variation in the angle of attack continually resulting in dynamic stall. Dynamic stall (DS)
phenomenon is comprised of the following stages:

Step 1: Static angle of attack is exceeded and flow reversal starts occurring
Step 2: Flow separation, accompanied by the formation of vortex, start taking place at the leading
edge
Step 2-3: The vortex starts convecting towards the trailing edge. Extra lift is produced during
convection
Step 3-4: Stalling takes place- flow is entirely separated from the upper aerofoil region
Step 5: The flow re-attaches, starting from the leading edge, once the AoA has become small
enough
These steps are represented diagrammatically in Figure 8.

14
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 8: Stages of dynamic stall (Leishman, 2002)

It can be seen in Figure 9 that dynamic stall causes hysteresis, i.e. the value of Cl, Cd and Cm is
not unique for a given AoA and depends on whether the AoA was approached from higher or a
lower value. Also, the loads during dynamic stall are much higher than that for static stall. So, for
an accurate prediction of VAT performance, dynamic stall needs to be taken into account.

Figure 9: Variation of Cl, Cd and Cm with AoA during dynamic stall (Leishman, 2002)

15
Thesis
MSc Sustainable Energy Futures Imperial College London

ALM correction

In ALM, dynamic stall is incorporated using semi-empirical models to offer more realistic aero foil
coefficients as compared to the static ones. These models use experimentally measured
constants, which might depend on aerofoil or fluid flow conditions, to predict dynamic stall onset,
loads during the stall and flow reattachment after stall. Beddoes (1982) produced the first such
model which contained constants general enough to be used in many cases. Leishman &
Beddoes (1989) and other works further refined the model resulting in the final general form of
LB DS model called “the third generation DS model” (Beddoes, 1993). Sheng, Galbraith & Coton
(2008) further modified LB DS model to make it suitable for low Mach number applications like
wind/tidal turbines. Based on the experiments at the University of Glasgow, they concluded that
there were three major problems with the LB DS methods when applied to low Mach number
applications (like VATs):

1. It predicts smaller AoA for DS initiation


2. It does not account for stalled flow convection over the upper surface of aerofoil and into
the free stream during recovery from stall
3. No negative chord wise force is produced by LB DS model

Based on the better physical understanding and more reliable data from the Glasgow
experiments, an improved DS model was devised. Figure 10 shows that the new DS model is
exceptionally close to experimental values and offers significant improvement over original LB DS
model.

Figure 10: CN for a NACA 0012 oscillatory test (α=15 deg+10 deg sinωt) with Orig. LB DS model and New Sheng et
al. model (Sheng, Galbraith & Coton, 2008)

16
Thesis
MSc Sustainable Energy Futures Imperial College London

2.3.3.3 Added mass


Description

As explained before, HATs can be assumed to operate in a quasi-static state, i.e. the velocity of
their blades and the upwind and downwind flow field does not change much with time. In their
𝜕𝑢
case, it is assumed that the fluid possesses only “spatial acceleration”, 𝑈∞ , where x is the flow
𝜕𝑥
𝜕𝑢
direction of one-dimensional flow, but no temporal acceleration . However, as VATs rotate, the
𝜕𝑡

AoA on their blades changes continuously resulting in their temporal acceleration. As these
blades accelerate, they also cause the fluid surrounding them to accelerate which is experienced
as an added inertial force by the blades. So, in order to correctly calculate the forces on a VAT,
the effect of this added mass (AM) of fluid needs to be considered.

ALM correction

Strickland, Smith & Sun (1981) derived a model for added mass using a pitching flat plate in a
flow field. Normal C𝑛𝐴𝑀 , tangential C𝑡𝐴𝑀 and moment C𝑚𝐴𝑀 correction coefficients were calculated
as:

𝜋𝑐𝑈̇𝑛
C𝑛𝐴𝑀 = − (12)
8 𝑈𝑟𝑒𝑙
𝜋𝑐𝛼̇ 𝑈𝑛
C𝑡𝐴𝑀 = − (13)
8 𝑈𝑟𝑒𝑙
C𝑛𝐴𝑀 𝑈𝑛 𝑈𝑡
C𝑚𝐴𝑀 = − − (14)
4 8 𝑈𝑟𝑒𝑙
Where 𝑈̇𝑛 and 𝛼̇ denotes time derivative. These coefficients are then added to those calculated
by dynamic stall modelling.

2.3.3.4 Flow curvature


Description

As VAT blade moves in a curvilinear path, the AoA is not the same along its chord-wise direction
at a given azimuthal angle. The thicker the chord, the larger is the variation in AoA. So, as
explained before, AoA of the blade not only varies as it rotates but also varies along its chord at
a given instant of time. This varying point of incidence along chord makes it difficult to specify an
AoA to define foil coefficients. So, an “effective AoA” needs to be used to take into account the
flow curvature (FC) effects.

17
Thesis
MSc Sustainable Energy Futures Imperial College London

ALM correction

A correction for this purpose was proposed by Goude (2012) on the basis of his experiments
involving a flat plate moving in a circular path. It defines the effective angle of attack as:

𝑉𝑎𝑏𝑠 cos(𝜃𝑏 − 𝛽) 𝛺x𝑐 𝛺𝑐


𝛼 = 𝛽𝑝 + tan−1 − − (15)
𝑉𝑎𝑏𝑠 sin(𝜃𝑏 − 𝛽) + 𝛺𝑟 𝑉𝑟𝑒𝑓 4𝑉𝑟𝑒𝑓

Where 𝛽𝑝 is the blade pitch angle, 𝑉𝑎𝑏𝑠 is the magnitude of local inflow velocity at the blade, 𝜃𝑏 is
the blade azimuthal position, 𝛽 is the direction of the inflow velocity, 𝛺 is the turbine’s angular
velocity, 𝑟 is the blade element radius, x is the fractional distance of the mounting point from the
AL node and 𝑉𝑟𝑒𝑓 is the reference flow velocity for calculating angle of attack.

2.3.3.5 Blade end effects


Description

Prandtl & Betz (1927) related the local lift produced by a blade element with circulation of a bound
vortex. According to Helmholtz 2nd vortex law, vortex lines should either form a closed loop or end
at a boundary. As a consequence, the bound vortex and the associated lift should become zero
at the blade ends. The famous Glauert (1935) correction for this effect for HATs relies on rotor
parameters like tip flow angle, no. of blades, etc., which does not necessarily depict VAT flow
conditions. So, a more fundamental definition of these effects is needed to account for VAT blade
end effects.

ALM correction

A generalised correction for blade end effects can be derived using Prandtl’s lifting line theory.
Based on this, the circulation distribution Γ(θ) is given by

Γ(θ) = 2𝑆𝑈∞ ∑ 𝐴𝑛 sin 𝑛𝜃 (16)


1

Where 𝑆 is the span length and 𝑈∞ is free stream velocity.

Using Kutta-Joukowski theorem, the lift coefficient is given by

−Γ(θ)
C𝑙 (𝜃) = (17)
1
2 𝑐𝑈∞
This can be converted into a correction factor varying from 0 to 1 by dividing with C𝑙 (𝜃)𝑚𝑎𝑥 .

18
Thesis
MSc Sustainable Energy Futures Imperial College London

All these corrections are incorporated into the basic AL concept in order to get a much more
realistic estimate of turbine loadings than the classical BET.

2.3.3.6 Projection on flow field


Once the corrected blade loadings have been calculated, they need to be projected on the fluid
domain to compute the momentum source term 𝐹𝑇𝑥𝑖 in Eq(3). As ALM consists of node points and
introduction of point sources into the fluid model can lead to instability, the loads are smoothed
using spherical Gaussian function 𝜂. The function is given by:

1 |𝑟⃗|2
𝜂= exp (− ) (18)
𝜖 3 𝜋 3/2 𝜖2
Where |𝑟⃗| is the magnitude of distance of the mesh point from the AL node and 𝜖 is Gaussian
width.

2.4 Turbulence modelling


Turbulence is described as being made up of eddies of different sizes. Turbulence models shown
in Figure 6 mainly differ with respect to the required spatial mesh resolution with which they are
able to explicitly capture the turbulence in the flow field and what effects remain to be ‘modelled’.
Figure 11 shows different turbulence regimes and which of these regimes are captured by each
turbulence model.

Increasing physical description and computational cost


Increasing use of Empirical results

Figure 11: General features of commonly used turbulence models (Altair, 2014)

19
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 12 shows a comparison of the resolution of results, say velocity at a point in flow field,
achieved by these models.

Figure 12: Velocity resolution based on different turbulence models, adapted from (Staffelbach, Senoner, Gicquel, et
al., 2008)

Now, DNS and LES are briefly described followed by in-depth explanation of uRANS.

2.4.1 Direct Numerical Simulation (DNS)


DNS aims to solve the NS equations for the turbulence length scales down to the dissipation
scale, called the Kolmogorov scale (Kolmogorov, 1941). However, resolving the flow up to the
Kolmogorov scale is very computationally expensive with computational cost ∝ Re9/4 (Pope, 2000)
in terms of spatial discretisation requirements. As a result, DNS has been limited to low Re and
simple geometry based flows to understand the fundamentals of turbulence better.

2.4.2 Large Eddy Simulation (LES)


LES involves representing turbulence below a certain spatial resolution threshold using
turbulence models known as sub-grid scale (SGS) models. A consequence of Kolmogorov (1941)
is theory of self-similarity, which states that the turbine parameters dictates only the large scale
eddies while the smaller scale eddies show universal behaviour. So, only the larger and relatively
important turbulent scales are resolved, substantially reducing computational cost as compared
to DNS (Bachant & Wosnik, 2016). LES can be applied to High Re flows but is troublesome near
walls. Also, Courant–Friedrichs–Lewy (CFL) condition limits the maximum time interval which can
be used with a given length interval to achieve convergence in explicit LES. As a result, LES is
mainly used in research and is slowly progressing into industry.

20
Thesis
MSc Sustainable Energy Futures Imperial College London

2.4.3 (Unsteady) Reynolds Averaged Navier-Stokes (RANS/uRANS) Modelling


RANS modelling involves decomposing the governing equations, continuity and NS equation, into
mean and fluctuating components and applying Reynolds Averaging function. This yields
equations describing the mean flow of the fluid because the Reynold average of fluctuating
components is zero. The mean flow equations contain some products of varying components
which are approximated using models. This approach results in even faster computation as
compared to LES as far coarser mesh resolution can be used, but can compromise accuracy to
some extent (Bachant & Wosnik, 2016). Because of its faster solution times, it is a standard
approach in industry.

uRANS is a variation of RANS used to analyse problems where flow field is inherently unsteady
because of the nature of forcing applied like in the case of VATs, where the interaction of blades
and fluid induces periodic disturbances in the flow field (Iaccarino, Ooi, Durbin, et al., 2003).
Impact of turbulence on mean velocity is still calculated using models but the effects of periodic
disturbance are captured by selecting a suitable time scale to compute temporally varying flow
properties.

As shown in Figure 13, first of all, the variables in the governing equations are broken into their
mean and fluctuating components. Eq (19) shows this approach being applied to fluid velocity.

𝑢𝑖 (𝑥⃗, 𝑡) = 𝑢̅𝑖 (𝑥⃗, 𝑡) + 𝑢́ 𝑖 (𝑥⃗, 𝑡) (19)

Figure 13: Representation of flow velocity as a sum of mean and fluctuating component (ANSYS FLUENT, 2006)

21
Thesis
MSc Sustainable Energy Futures Imperial College London

Where (−) denotes the averaged and (‘) denotes the fluctuating components. The mean quantities
are defined using Reynolds averaging. There are two kinds of Reynolds averaging commonly
used- Time averaging and ensemble averaging represented in Eq (20) and Eq (21).

𝑇
1
𝑢̅𝑖 (𝑥⃗) = lim ∫ 𝑢𝑖 (𝑥⃗, 𝑡)𝑑𝑡 (time averaging) (20)
𝑡→∞ 𝑇
0
𝑛
1
𝑢̅𝑖 (𝑥⃗, 𝑡) = lim ∑ 𝑢𝑖 (𝑖) (𝑥⃗, 𝑡) (ensemble averaging) (21)
𝑛→∞ 𝑛
𝑖=1

Ensemble averaging is a statistical approach in which results from a large number of independent
experiments are averaged to find a mean value. Velocity and pressure in a turbulent flow can be
regarded as random variables and can be ensemble averaged to find the mean values.

Using averaging and simplifying, Eq(1) and Eq(3) without the momentum source term which can
be added later, can be written as:

𝜕𝑢̅𝑗
∑ =0 (22)
𝜕𝑥𝑗
𝑗

𝜕𝑢̅𝑖 𝜕𝑢̅𝑗 𝑢̅𝑖 𝜕𝑢́̅̅̅̅̅


𝑗 𝑢́ 𝑖 1 𝜕𝑝̅ 𝜕 2 𝑢̅𝑖
+∑ +∑ + −𝑣∑ 2 = 0 (23)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜌 𝜕𝑥𝑖 𝜕𝑥 𝑗
𝑗 𝑗 𝑗

̅̅̅̅̅
𝑢́ 𝑗 𝑢́ 𝑖 are called Reynolds stresses and arise due to Reynolds averaging of viscous terms.

Appearance of these terms increases the unknowns from four (𝑢1 , 𝑢2 , 𝑢3 , 𝑝) to ten
(𝑢̅1 , 𝑢̅2 , 𝑢̅3 , 𝑝̅ , ̅̅̅̅̅̅
𝑢́ 1 𝑢́ 1 , ̅̅̅̅̅̅
𝑢́ 1 𝑢́ 2 , ̅̅̅̅̅̅
𝑢́ 1 𝑢́ 3 , ̅̅̅̅̅̅
𝑢́ 2 𝑢́ 2 , ̅̅̅̅̅̅
𝑢́ 2 𝑢́ 3 , ̅̅̅̅̅̅
𝑢́ 3 𝑢́ 3 ). Since there are four governing equations and ten
unknowns, more equations are required to calculate the six unknown Reynolds stresses. This is
known as the closure problem. The two main models used to close uRANS equations are:

Reynold Stress Models (RSMs)

RSMs use an additional transport equation for each of the unknown Reynolds stresses. These
models are the most physically accurate of all the uRANS model. However, these models are
troublesome with respect to convergence and solving six new transport equations makes them
consume considerably more time and memory.

Eddy Viscosity Models (using Boussinesq hypothesis)

In an approach analogous to molecular viscosity, Eddy viscosity models use isotropic “eddy or
turbulent viscosity” (𝑣𝑇 ) to relate Reynold stresses with mean velocity gradients.

22
Thesis
MSc Sustainable Energy Futures Imperial College London

2 𝜕𝑢̅𝑖 𝜕𝑢̅𝑗
̅̅̅̅̅
𝑢́ 𝑗 𝑢́ 𝑖 − 𝑘𝛿𝑖𝑗 = −𝑣𝑇 ( + ̅
) = −2𝑣𝑇 𝑆𝑖𝑗 (24)
3 𝜕𝑥𝑗 𝜕𝑥𝑖

Where 𝑘 is turbulent kinetic energy, which is the mean kinetic energy per unit mass associated
̅ is the mean rate of strain tensor.
with eddies in turbulent flow, 𝛿𝑖𝑗 is the Kronecker delta and 𝑆𝑖𝑗
Eddy viscosity models have been verified widely over the years and are preferred over RSMs, in
general, because of their faster (and easier) convergence and less computational expense. The
eddy viscosity used in Eq (24) can be calculated using zero-equation, one-equation or two-
equation models, listed in the increasing order of complexity and accuracy. Some of the
commonly used models are now described.

2.4.3.1 Prandtl’s Mixing Length Theory (zero-equation)


Using dimensional analysis, the eddy viscosity can be defined as:

𝑣𝑇 = 𝑈𝑙𝑚
(25)

Where 𝑈 is a typical velocity defined by

𝜕𝑢̅
𝑈 = 𝑙𝑚 | |
𝜕𝑦 (26)

And 𝑙𝑚 is a mixing length. Putting value of 𝑈 in Eq (25) we get,

𝜕𝑢̅
𝑣𝑇 = 𝑙𝑚 2 | |
𝜕𝑦 (27)

This model applies to simple shear flows only and is incomplete, i.e. it requires 𝑙𝑚 to be defined
as an input to the model. Moreover, it solely relates 𝑣𝑇 with velocity gradient, so 𝑣𝑇 is zero in the
absence of gradients which is not always the case like in the centerline of a jet and in separating
flows. Also, flow history effects are not taken into account.

2.4.3.2 k-model (one-equation)


This approach uses turbulent kinetic energy (k) to compute typical velocity scale 𝑈 as

𝑈 = 𝑐𝑘 1/2
(28)

where c is a constant of proportionality. Combining Eq (28) and (25), we get

23
Thesis
MSc Sustainable Energy Futures Imperial College London

𝑣𝑇 = c𝑘1/2 𝑙𝑚 (29)

This decoupling of 𝑈 with velocity gradients solves the problem of the eddy viscosity becoming
zero in gradient less flows described above.

A transport equation is used for 𝑘 as:

𝜕𝑘 𝜕𝑘 𝜕 𝟏 𝜕𝑘 𝜕𝑢̅𝑖
= − ∑ 𝑢̅𝑗 −∑ 𝒖́𝒋 𝒑́ + ̅̅̅̅̅̅
( ̅̅̅̅̅̅ 𝒖́𝒋 𝒌́ − 𝑣 ) − ∑ ∑ ̅̅̅̅̅
𝑢́ 𝑗 𝑢́ 𝑖
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝝆 𝜕𝑥𝑗 𝜕𝑥𝑗
𝑗 𝑗 𝑖 𝑗
(30)
̅̅̅̅̅̅̅̅̅̅
𝝏𝒖́𝒊 𝝏𝒖́𝒊
−∑ ∑ 𝒗
𝝏𝒙𝒋 𝝏𝒙𝒋
𝒊 𝒋

Where the first term on the right represents convection, second and third represent turbulent
diffusion, fourth represents viscous diffusion, fifth represents production and sixth represents
dissipation. The production term is computed using the Boussinesq approximation. This leaves
Eq (30) with three unclosed terms, i.e. the second, third and sixth term, shown in bold, which
needs to be closed (modelled).

Closure of dissipation term:

Production of turbulence is governed by scales corresponding to the turbulence generating


object’s geometry. Using dimensional analysis

𝑈3
P=
𝐿 (31)

Where U and L are typical velocity and length scales of the mean flow. In a turbulent flow, the
rate of production of turbulence (P) should be equal to rate of its dissipation (ε) to prevent the
turbulence from increasing or diminishing without bound (Versteeg & Malalasekera, 2007). So,

𝑈3
ε=𝑃=
𝐿 (32)

Combining Eq (28) and Eq (32) gives

𝑘 3/2
ε = 𝐶𝐴
𝑙𝑚 (33)

24
Thesis
MSc Sustainable Energy Futures Imperial College London

where CA is a model constant

Closure of turbulent transport terms:

The gradient diffusion hypothesis is used to model turbulent transport as:

𝑣𝑇 𝜕𝑘
𝑢́ 𝑖 𝑝́ + 𝑢́̅̅̅̅̅
̅̅̅̅̅ ́
𝑗𝑘 =
𝜎𝑘 𝜕𝑥𝑗 (34)

Where 𝜎𝑘 is turbulent Prandtl number and is generally taken to be equal to 1. Eq (34) shows that
the turbulent transport due to correlation in fluctuating pressure and velocity is along the gradient
of 𝑘.

Once all the terms have been suitably modelled, the k one equation model can be written as:

𝜕𝑘 𝜕𝑘 𝜕 𝑣𝑇 𝜕𝑘
= − ∑ 𝑢̅𝑗 +∑ ((𝑣 + ) )+𝑃−𝜀 (35)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜎𝑘 𝜕𝑥𝑗
𝑗 𝑗

Where:

𝑘 3/2
ε = 𝐶𝐴 (36)
𝑙𝑚

𝜕𝑢̅𝑖
𝑃 = − ∑ ∑ ̅̅̅̅̅
𝑢́ 𝑗 𝑢́ 𝑖 (37)
𝜕𝑥𝑗
𝑖 𝑗

𝑣𝑇 = c𝑘1/2 𝑙𝑚 (38)

2
̅̅̅̅̅ ̅
𝑢́ 𝑗 𝑢́ 𝑖 − 𝑘𝛿𝑖𝑗 = −2𝑣𝑇 𝑆𝑖𝑗
3 (39)

With 𝑐 = 0.55 and 𝐶𝐴 = 𝑐 3

25
Thesis
MSc Sustainable Energy Futures Imperial College London

The main advantage it offers over Prandtl’s mixing length theory is that it takes flow history effects
into account. However, 𝑙𝑚 still needs to be given as an input to the model, so the one equation
model is still incomplete.

2.4.3.3 k- ε model (two-equation)


k- ε is the most widely validated and used turbulence model for engineering applications. This
model introduces an equation for the dissipation rate 𝜀 to close eddy viscosity without the need
for defining the flow dependent parameter 𝑙𝑚 . Using dimensional analysis 𝑙𝑚 is defined as:

𝑘 3/2
𝑙𝑚 = (40)
𝜀

So, 𝑣𝑇 is now given by

𝑘 3/2 𝑘2
𝑣𝑇 = UL = C𝜇 𝑘 1/2 = C𝜇
𝜀 𝜀 (41)

It uses the same equation for k as used in one equation k model. A further transport equation for
𝜀 is developed analogous to the k transport equation. The 𝜀 transport equation is written as:

𝜕𝜀 𝜕𝜀 𝜕 𝑣𝑇 𝜕𝜀 𝑃 𝜀
= − ∑ 𝑢̅𝑗 +∑ (𝑣 + ) + 𝐶𝜀1 − 𝐶𝜀2
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜎𝜀 𝜕𝑥𝑗 𝑇 𝑇 (42)
𝑗 𝑗

Where

𝑘
𝑇=
𝜀 (43)

With C𝜇 , 𝐶𝜀1, 𝐶𝜀2, 𝜎𝑘 and 𝜎𝜀 as constants.

The major issue with k- ε model is that it is too diffusive and does not give accurate results near
no-slip walls. The value of ε is undefined at the wall and necessitates the use of a wall function.

2.4.3.4 k- ω model (two-equation)


Wilcox (1994) used specific dissipation rate (ω) instead of 𝜀 to solve the indeterminacy issue k- ε
model faces near walls. Specific dissipation rate is defined as:

26
Thesis
MSc Sustainable Energy Futures Imperial College London

ε 1
ω≈ ∝
𝑘 𝑇 (44)

So, 𝑣𝑇 is now given by

𝑘
𝑣𝑇 = 𝛼 ∗
ω (45)

And the modified k equation and an equation for ω is written as:

𝜕𝑘 𝜕𝑘 𝜕 𝑣𝑇 𝜕𝑘
= − ∑ 𝑢̅𝑗 +∑ (𝑣 + ) + 𝑃 − 𝛽 ∗ 𝑓𝛽 𝑘ω (46)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜎𝑘 𝜕𝑥𝑗
𝑗 𝑗

𝜕ω 𝜕ω 𝜕 𝑣𝑇 𝜕ω ω
= − ∑ 𝑢̅𝑗 +∑ (𝑣 + ) + 𝛼 𝑃 − 𝛽 𝑓𝛽 ω2
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜎ω 𝜕𝑥𝑗 𝑘 (47)
𝑗 𝑗

Where 𝛼 ∗ , 𝛼 , 𝛽, 𝛽 ∗, 𝑓𝛽 , 𝜎𝑘 , 𝜎ω are constants.

Wilcox’s k- ω model is accurate and can be applied to a variety of flow regimes. However, it has
one drawback- it is very sensitive to the inlet free stream value of ω.

2.4.3.5 k- ω SST model (two-equation)


Menter (1994) suggested a shear stress transport (SST) k- ω model combining the good attributes
of k- ε and Wilcox k- ω model. It uses blending function to switch between k- ε and k- ω models.
It uses k- ε in the outer part of the boundary layer, away from the wall, to get rid of the high
sensitivity of the k- ω model to ω boundary conditions. Whereas a modified k- ω approach is used
near the walls as k- ε performs poorly in this region. The modified SST k- ω has following set of
equations:

𝜕𝑘 𝜕𝑘 𝜕 𝑣𝑇 𝜕𝑘
= − ∑ 𝑢̅𝑗 +∑ (𝑣 + ) + 𝑃 − 𝛽 ∗ 𝑘ω (48)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜎𝑘 𝜕𝑥𝑗
𝑗 𝑗

27
Thesis
MSc Sustainable Energy Futures Imperial College London

𝜕ω 𝜕ω 𝜕 𝑣𝑇 𝜕ω γω
= − ∑ 𝑢̅𝑗 +∑ (𝑣 + ) + 𝑃 − 𝛽 ω2 + ∑ 2(1
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜎ω 𝜕𝑥𝑗 𝑘
𝑗 𝑗 𝑗

1 𝜕𝑘 𝜕ω (49)
− 𝐹1 )𝜎ω2
ω 𝜕𝑥𝑗 𝜕𝑥𝑗

Where

4
√𝑘 500𝑣 4𝜎ω2 𝑘
𝐹1 = tanh {{min [max ( ∗ , ), ]} } (50)
𝛽 ωy ωy 2 𝐶𝐷𝑘ωy 2

1 𝜕𝑘 𝜕ω
𝐶𝐷𝑘ω = max(2𝜌𝜎ω2 , 10−10 ) (51)
ω 𝜕𝑥𝑗 𝜕𝑥𝑗
And 𝜎ω2 is a constant.

2.5 Domain discretisation (Meshing)


Real life fluids have continuous velocity and pressure fields comprising of infinite points. However,
in order to make use of numerical methods on finite components, the flow field needs to be broken
down into a finite number of cells/elements/nodes on which the solution of the Partial Differential
Equations (PDEs) is approximated numerically. But before proceeding to the time discretisation
methods, it is important to establish concept of weak and strong representation forms of the PDEs.

Strong form

The strong form of an equation describes some behaviour pointwise and applies to each point in
the domain. For example, Eq (3) is the NS equation in its strong form as it will hold individually for
every point in the fluid flow field. Since the solution to the strong form needs to hold for each point
in the domain, it needs to be continuous and differentiable. However, in reality, there might not be
a smooth solution to the problem especially in case of complex domains and/or different material
surfaces. Furthermore, inclusion of boundary conditions to the solution of strong form can be very
challenging. This restricts the use of strong form.

Weak form

The weak solutions of an equation describe the behaviour over an interval (areas, volumes, etc.)
and hence relaxes the requirement of continuity as it is satisfied in an average sense, i.e. over

28
Thesis
MSc Sustainable Energy Futures Imperial College London

volumes instead for each point. Conversion of strong form to weak form requires multiplying the
equation with a test function and integrating it over some interval or subdomain Z as:

𝜕𝑢𝑖 𝜕𝑢𝑗 𝑢𝑖 1 𝜕𝑝 𝜕 2 𝑢𝑖 1
∫ ψ𝑖 ( +∑ + − 𝑣 ∑ 2 − 𝐹𝑇𝑥𝑖 ) = 0 (52)
𝜕𝑡 𝜕𝑥𝑗 𝜌 𝜕𝑥𝑖 𝜕𝑥 𝑗 𝜌
Z 𝑗 𝑗

The equation of continuity can be represented in its weak form similarly. Now, if a function u is a
solution of Eq (3), it is also a solution to the Eq (52). Conversely, if a trial function u satisfies
Eq(52) for a sufficient number of different test functions ψ𝑖 , belonging to some test space, the trial
function should be the solution to Eq(3) as well (Burgerscentrum, 2015).

Now the standard discretization methods are as follows:

2.5.1 Finite Difference Method (FDM)


The FDM involves solving the NS equation in strong form. The domain is divided into a finite set
of grid points, also called nodes, and exact partial derivatives are approximated using forward,
backward or central difference methods with central difference equation shown below.

𝜕𝑢 Δ𝑢 𝑢(𝑥𝑖+1 )−𝑢(𝑥𝑖−1 )
≅ = (53)
𝜕𝑥 Δ𝑥 𝑥𝑖+1 − 𝑥𝑖−1
As a result, a system of algebraic equations is formed which can be solved to find the values at
nodes. FDM is applicable to structured grids only (Li, 2014). Structured grids are generally
comprised of quadrilaterals in 2D and hexahedra in 3D having regular connectivity over the
domain. This makes grid formation (meshing) simpler but complex geometries cannot be captured
using it.

2.5.2 Finite Volume Method (FVM)


The FVM involves discretising the domain into a finite set of sub-volumes. Volumes usually
comprise of simple polyhedral shapes like cuboids, wedges, tetrahedra, etc., with the unknown
variable located at the centroid of the control volume. It involves representing the governing
equations in integral form and applying to each control volume. The problem generally reduces
to flux calculation across surfaces of the control volume.

𝜕
∫ 𝑸𝑑𝑉 = ∫ 𝑭𝑑𝑨 (54)
𝜕𝑡
Z δZ

29
Thesis
MSc Sustainable Energy Futures Imperial College London

Where Q and F are vectors of conserved quantities and fluxes, respectively, and A is the vector
area of the control surface. FVM can be applied to structured, unstructured or hybrid grids.
Unstructured grids are generally comprised of triangles in 2D and tetrahedra in 3D having irregular
connectivity over the domain. Meshing is complicated but complex geometries can be captured
far more easily than FDM. Hybrid grids combine the benefits of structured and unstructured grids.

2.5.3 Finite Element Method (FEM)


The FEM relies on discretising the domain into finite elements and uses the weak form of the NS
equation. The solution can be represented by a linear combination of N functions, the trial or
shape functions ψ𝑖 , as follows:

𝑢(𝑥) ≅ ∑ 𝑢𝑖 ψ𝑖 ((𝑥)) (55)


𝑖=1

The shape functions are non-zero over a small region of the domain or elements only. Each
element has a number of nodes which depends on the order of the shape function. The purpose
of the FEM is to transform the solution from every point of domain, an infinite or huge space, to
finite set of vectors belonging to the trial space, which can be solved in matrix form.

Figure 14 shows u(x) being approximated by linear shape functions with N=7. The functions have
value 1 at their respective nodes and 0 at other nodes. The elements are chosen to be linear but
can be a function of any order.

Figure 14: A plot showing function being approximated by linear combination of shape functions (COMSOL, 2017)

In the Galerkin method, the same function space is chosen for the test function in Eq (52) and
shape function in Eq (55). A common approach is to use polynomials of degree n ≤ N, referred to
as PN discretisation, for this purpose. Continuous Galerkin method restricts the choice to only the

30
Thesis
MSc Sustainable Energy Futures Imperial College London

functions which are continuous between different elements. However, the discontinuous Galerkin
method allows for the use of any function but extra care should be taken while applying Green’s
theorem to the weak form.

Some of the commonly used elements are shown in Figure 15.

After discretisation, the weak form is solved by solving a matrix system for each node which
contains contributions from all the elements towards that node. In general, the denser the mesh
the closer the approximation gets to the actual solution.

Figure 15: Some common 2D and 3D linear and quadratic elements along with their nodes (COMSOL, 2017)

2.5.4 Fluidity and adaptive meshing


Fluidity is an open source, general purpose, computational fluid dynamics code based on the
finite element formulation of NS equation. It can be used in parallel on thousands of processors
using Message Passing Interface (MPI). An advantageous feature of Fluidity over other CFD
codes is that it offers unstructured meshing with adaptive meshing techniques which can reduce
the simulation time without compromising accuracy (Hiester, Piggott, Farrell, et al., 2014; Davies,
Wilson & Kramer, 2011). Deskos, Abolghasemi & Piggott (2017); Deskos & Piggott (2017)
analysed HAT arrays using ALM and mesh adaptivity in Fluidity. A traditionally unstructured mesh
is used to capture complex geometries and a locally refined mesh is applied in specific areas
where more “physics” is taking place. But the mesh is developed at the start and does not change
during the simulation. In unsteady problems like VAT analysis, there are some areas in which
mesh resolution requirement changes with time. For example, the region behind the VAT has no
vortices or turbulence initially and hence a coarse mesh can give good results inexpensively. But

31
Thesis
MSc Sustainable Energy Futures Imperial College London

as the wake starts to develop and progress downstream, the mesh refinement requirement of
these areas start to increase. Adaptive meshing techniques use error estimates to decide the
regions which need mesh modification and correspondingly modify/refine the mesh. This ensures
that the mesh is optimised, i.e. it is neither under nor over-resolved to capture the flow behaviour
correctly in the least time possible. There are three primary ways of adapting the finite element
mesh.

1. h adaptivity

This approach fulfils the mesh requirement by increasing or decreasing the number of elements,
keeping their type the same. This is done by changing the characteristic length ‘h’ of the elements
used. It increases the no. of elements in the region having error estimates above a user-defined
acceptable limit. Similarly, number of elements is decreased if the error estimate is below the
allowed limit. h adaptivity is the simplest of all adaptivity techniques and is suitable mainly for
steady state problems (Flaherty, 2010). Figure 16 show how h adaptivity refines the mesh. h
adaptivity is offered by Fluidity.

Figure 16: Mesh before and after h adaptivity

2. r adaptivity

The r adaptivity (relocation adaptivity) technique keeps the no. of elements the same and adapts
the mesh by repositioning nodes so that the elements are more suitably shaped/placed which is
particularly useful for transient problems (Flaherty, 2010). Fluidity also implements this type of
adaptivity as shown in Figure 17.

32
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 17: Mesh before and after r adaptivity

3. p adaptivity

In p adaptivity, the polynomial order of elements is increased or decreased to offer mesh


refinement or coarsening as per requirement. The number of elements is kept constant. This is
represented in Figure 18.

Figure 18: Mesh before and after p adaptivity

p adaptivity offers some problems while coupling with h adaptivity and has not been incorporated
in Fluidity yet (AMCG, 2014).

While using adaptivity there are a few points to be kept in mind:

 The need to specify maximum and minimum allowed edge length based on geometry and
physics involved, respectively
 Mesh should be reconstructed approximately every 10-20 time steps as the reconstruction
of mesh for every time step can be very time consuming and may introduce distortions in
the solution
 Edge length scales should be varied smoothly across adjacent elements

33
Thesis
MSc Sustainable Energy Futures Imperial College London

Once an optimised mesh has been developed based on the above guidelines, the results for flow
field needs to be transferred from previous mesh to this one. This process is called mesh to mesh
interpolation. A simple and commonly used technique for this purpose is Consistent Interpolation.
If Np are the nodes of previous mesh and Nn are the nodes of the new mesh, then a containing
element Kp is identified for each node from the previous mesh and the solution is evaluated at the
physical location of Nn. R-tree spatial indexing algorithm can be used to efficiently find Kp
(Guttman, Antonin, Guttman, et al., 1984). Consistent interpolation has several short-comings like
it is non-conservative or not suitable for discontinuous fields, etc. As discussed before,
discontinuous Galerkin method produces discontinuous fields which are not pointwise well
defined. For such fields, Galerkin projection method can be used for mesh to mesh interpolation
as it is free from the issues faced by Consistent Interpolation method (Farrell & Maddison, 2011).

34
Thesis
MSc Sustainable Energy Futures Imperial College London

Chapter 3. Modelling and Validation


The chapter starts with the presentation of a literature review of the attempts conducted so far to
model VATs. This is followed by a brief description of the experiment selected as a validation
case and its results. Simulation setup of uRANS-ALM in OpenFOAM and Fluidity using 3D k- ε
and k- ω SST closure models are then described along with the results and analysis. The
contribution of each correction model towards the correct estimation of results is also established.
A qualitative and quantitative validation is then carried out at the end.

3.1 Previous work


The history of tidal/wind turbine modelling was briefly described in the introduction chapter. Here,
significant recent efforts, first in VAT experiments- to provide reliable data to be used for model
validation studies, and then in VAT modelling are presented. Vermeulen, Builtjes, Dekker, et al.
(1979) experimentally analysed the wake behind a full-scale VAWT. The data acquired was not
very reliable because the inflow conditions were not controlled stringently. Sheldahl (1981) found
an excellent agreement between wind tunnel and on-site tests for 2 m Darrieus NACA 0012
blades for the same Re and rotational speeds. Brochier, Fraunie, Beguier, et al. (1986) used
laser-doppler velocimeter for velocity measurement to study dynamic stall on a two bladed H-
rotor VAWT in a water channel. The measurements were carried out at Re=104, which was not
high enough to provide Re independent data set. Tescione, Ragni, He, et al. (2014) used
stereoscopic particle image velocimetry to study the blade tip vortices and development of
asymmetry in the wake of an H-rotor VAT. In the domain of modelling, Balduzzi, Bianchini, Maleci,
et al. (2016) summarised the recent 2D modelling efforts of VATs and proposed an optimum
methodology regarding mesh size, domain, etc., to get the best results. Howell, Qin, Edwards, et
al. (2010) analysed surface finish effects on two and three bladed NACA 0022 VAWTs using 2D
and 3D k- ε closure methods concluding that 2D analysis overpredicts coefficient of power (Cp)
significantly due to the absence of vertical flow. The same was observed by Alaimo, Esposito,
Messineo, et al. (2015) using ANSYS FLUENT to study different VAWT geometries. Hence, a 3D
modelling approach was selected for this thesis as well. Boudreau & Dumas (2015) used 3D
Delayed Detached Eddy Simulation (DDES) to investigate the wake recovery mechanisms of
different Marine Hydrokinetic turbines (MHKTs) including HAT, cross flow turbine (CFT) and
oscillating CFT. They found that turbulent transport dominated wake recovery in HATs. However,
wake recovery of both types of CFTs was dominated by mean spanwise velocity. Lam & Peng
(2016) found good agreement between k- ω SST simulation of a low solidity VAWT’s wake and

35
Thesis
MSc Sustainable Energy Futures Imperial College London

stereoscopic PIV test data. All of these studies were blade fitted and agreed well, especially the
3D studies, with experiments. Ouro & Stoesser (2017) used immersed boundary technique to
represent turbines which is quite useful to represent moving bodies without the need for the
continued regeneration of the mesh. They performed an LES of VATT blade to study vortex
interaction for both, laminar and turbulent, flows and found good agreement with the experiments.
Shamsoddin & Porté-Agel (2014) performed the first ever LES-ALM study of VAT and found ALM
to capture the unsteady wake much better than the Actuator Swept Surface Model (ASSM).
Bachant & Wosnik (2014) laid the foundation for their comprehensive work on VATs/CFTs with
an experimental study of the near wake along with uRANS-ADM simulation. uRANS-ADM was
found to be unable to predict the near wake accurately as the near wake is governed by the
turbine’s geometry. This study was followed by a blade resolved Spalart–Allmaras (SA) and k- ω
SST uRANS study to analyse the near wake of the same VATT. k- ω SST matched experimental
results quite accurately while SA failed to capture turbulent kinetic energy (TKE) in the wake
satisfactorily (Bachant & Wosnik, 2016). Bachant, Goude & Wosnik (2016) replicated the blade
resolved study using uRANS-ALM and LES-ALM in OpenFOAM. LES-ALM was found to captured
the mean velocity and asymmetry of the wake very accurately. However, it produced much less
TKE in the wake. uRANS-ALM gave satisfactoy results for both mean velocity and TKE. The
parameterised models were found to significantly reduce simulation time over blade resolved
studies, still producing results which were in reasonable agreement with experiments.

Based on the comprehensiveness of the studies carried out by Bachant and Wosnik, their
experiment is used as the validation case. Also, their ALM in OpenFOAM is used to analyse the
effect of correction models like dynamic stall, flow curvature, etc., on the accuracy of uRANS-
ALM results. Fluidity was also used to evaluate the usefulness of this modelling option as
OpenFOAM does not offer mesh adaptivity.

3.2 Experiment
3.2.1 Setup
A 3-bladed high solidity H-rotor VAT turbine was analysed in a tow-tank. The turbine was kept
simple with rotor height (H) and diameter (D) equal to 1 m. The blades, with zero preset pitch,
were attached at half-chord and half-span to 0.095 m diameter shaft via three support struts, one
strut for each blade. All blades and struts had NACA 0020 section with a chord of 0.14 m.
Experiments were conducted in a 36 m long, 3.66 m wide and 2.44m deep towing tank in
University of New Hampshire. The turbine was mounted in a vertical position on an aluminium

36
Thesis
MSc Sustainable Energy Futures Imperial College London

frame made up of NACA 0020 foils to minimise flow disturbance in the region adjacent to the
turbine and guy wires were used to prevent the frame from swinging. The drag/thrust force was
calculated by two S-beam load cells, whereas, the power produced was captured by an internal
regeneration resistor of a servo motor attached to the turbine’s shaft. The torque generated by
the turbine was measured by an inline rotary torque transducer. The turbine and the experimental
setup are shown in Figure 19.

Figure 19: Turbine model dimensions and tow tank experimental setup (Bachant & Wosnik, 2014)

37
Thesis
MSc Sustainable Energy Futures Imperial College London

An Acoustic Doppler Velocimeter (ADV) was used to capture the water velocity in the wake. All
the three components of velocity were captured at one diameter downstream distance (x/D=1m)
with a total of 270 probes whose positions are depicted in Figure 20.

Figure 20: Wake measurement coordinate system and measurement locations in meters (Bachant & Wosnik, 2014)

The fluid inlet velocity to the turbine was 1 m/s and the turbine rotated with tip speed ratios (TSR,
𝜆) varying from 0 to 3.2. The turbine diameter and blade chord Reynolds numbers, ReD and Rec,
are shown in Eq (56) and Eq (57), respectively.
𝑈∞ 𝐷
𝑅𝑒𝐷 = = 106 (56)
𝜈

𝜆𝑈∞ 𝑐
𝑅𝑒𝑐 = = 2.7 ∗ 105 (57)
𝜈

This high value of Re was necessary to ensure Re-independent results as the whole basis of the
experimental study was to provide data set to be used for model validations.
To make sure that the data set was free from errors due to drag produced by the frame (tear drag)
and friction in shaft bearings (tear torque), tear drag and torque were measured, and the results
were adjusted accordingly. Tear drag was calculated by removing the turbine and measuring the
drag due to the frame only. This was then subtracted from the total drag. Tear torque was

38
Thesis
MSc Sustainable Energy Futures Imperial College London

computed by rotating turbine support shaft and bearings in the air. The values were then
interpolated for different TSRs and added to the measured torque to get the actual torque
produced by the blades.
3.2.2 Results and analysis
The turbine was analysed over different TSRs and the resultant Cp and CD are shown in Figure
21. Turbines optimal TSR was found to be 𝜆 = 1.9.

Figure 21: Mean power (left) and drag (right) coefficients plotted versus tip speed ratio based on experimental values
(Bachant & Wosnik, 2014)

For the validation study, turbine and wake characteristics corresponding to optimum TSR and at
t=6 seconds were selected. Turbine performance parameters, Cp and CD, were found to be 0.26
and 0.96, respectively. For wake, mean velocity and TKE, nondimensionalized by their free
stream values as 𝑈/𝑈∞ and 𝑘/𝑈∞ 2 , respectively, were computed at x/D=1. Figure 22 shows mean
velocity and TKE profiles at x/D=1m and z/H=0. Mean velocity and TKE contours on a y-z planar
slice at x/D=1 are shown in Figure 23.

Figure 22: Mean velocity profile (left) and TKE (right) in the near-wake at x/D=1 and z/H=0 based on experimental
data, adapted from (Mendoza, Bachant, Wosnik, et al., 2016)

39
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 23: Mean velocity (top) and TKE (bottom) contours at x/D=1 based on experimental results. Vectors in the top
figure are cross-stream and vertical velocities; contours are stream-wise velocity. The view is looking upstream, with
the turbine frontal area indicated by the solid lines. The upper half of the turbine is shown with horizontal plane
through z/H=0 considered as the plane of symmetry (Bachant & Wosnik, 2014)

3.3 OpenFOAM
OpenFOAM is a widely used, highly optimised open source CFD software used for solving
turbulence and heat transfer problems in industry and academia. Bachant, Goude & Wosnik
(2016) performed 3D k- ε uRANS-ALM and Smagorinsky LES-ALM in OpenFOAM producing
reasonably accurate results. So, their AL method, called turbineFoam (Bachant, Goude & Wosnik,
2016b), is used with 3D k- ε and k- ω SST uRANS to examine turbine and wake performance
characteristics. The contribution of each correction model in predicting these characteristics
accurately is also determined.

40
Thesis
MSc Sustainable Energy Futures Imperial College London

3.3.1 Modelling
The whole turbine, including blades, shaft and struts, was parameterised using the AL concept.
The number of elements used for blade, strut and shaft was equal to 14, 6 and 20, respectively.
The actuator lines were rotated with TSR= 1.9, corresponding to the optimal operation point.
Sheng et al. (2008) DS model, as described in the Section 2.3.3.2, was used with coefficients
given in Table 1, which are also the default values. For flow curvature, added mass and blade
end effects, Goude’s, Stickland’s and Prandtl’s lifting line corrections described in Chapter 2 were
used, respectively. NACA 0021 foil data was used because of the scarcity of the data
corresponding to Rec>105 for NACA 0020 foils with the assumption that the small difference in foil
thickness is negligible. Data sets corresponding to Rec=3.6x105 and Rec=6.8x105 were used, and
values for other Reynolds number were interpolated. The rotor shaft and struts were assigned the
drag coefficients of 1.1 and 0.05, respectively.
Coefficient Value
Tp 1.7
Tf 3
Tα 6.25
αds0-αss 3.8
r0 0.01
Tv 11
Tvl 8.7
B1 0.5
B2 0.2
η 0.98
E0 0.16

Table 1: Values of coefficients used in Sheng et al. dynamic stall model

3.3.2 Meshing and verification


The domain used was 3.66 m long, 3.66 m wide and 2.44 m deep. The domain length was
reduced from the experimental setup length of 36 m to reduce computational expense with
boundary conditions ensuring similar conditions at inlet and outlet as that of the experiment. The
turbine shaft’s actuator line was placed at 1.52 m from the inlet and at the centre of the width and
depth of the domain. A structured hexahedral finite volume mesh, shown in Figure 24, was used

41
Thesis
MSc Sustainable Energy Futures Imperial College London

which was refined by a factor of 2 in the vicinity of the actuator lines to capture fluid-blade
interaction and near wake more accurately.
To mimic the tow tank conditions, following initial and boundary conditions were used:
 Velocity at the inlet was kept at 1 m/s with a zero gradient boundary condition.
 Velocity at the bottom and side walls was fixed to 1 m/s with no slip condition.
 Top free surface was applied using rigid slip condition.
 Pressure was held constant at the outlet.
 For k- ε, inlet turbulent kinetic energy and dissipation were set to 2x10-4 and 3x10-5,
respectively (Bachant & Wosnik, 2014).
 For k- ω, inlet turbulent kinetic energy and specific dissipation rate were set to 2x10-4 and
0.525, respectively. The same value of k was used as in k- ε case, whereas, the value of
ω was calculated using CFD-online turbulence properties calculator (CFD online, 2017).
The simulation was run for 6 seconds to allow the wake to develop and come out of the initial
transient phase.

Figure 24: Cutaway of 3D mesh used in OpenFOAM

In their study, Bachant et al. verified their mesh by varying spatial and temporal resolutions, one
by one, while keeping the other resolution constant. The whole spatial grid resolution was a
function of the number of cells in the stream-wise direction, Nx. They chose Nx= 48, resulting in

42
Thesis
MSc Sustainable Energy Futures Imperial College London

24576 cells, and time step ∆t= 0.01 seconds. The same spatial and temporal resolution was used
for this study as well.
3.3.3 Results and analysis
3.3.3.1 k- ε uRANS- ALM
With all the correction models included, Cp and CD values at TSR=1.9 came out to be 0.30 and
0.99, respectively, against the experimental values of 0.26 and 0.96. Both Cp and CD are slightly
over predicted. The simulation values were obtained by averaging over the last rotation cycle only
to get rid of the transient effects. Figure 25 shows variation of mean velocity and TKE at x/D=1
and z/H=0. The mean velocity can be seen to be skewed towards y+ side of the turbine showing
that the asymmetry in the wake, which would be absent in the case of ADM, was captured well
by ALM. This asymmetry is attributed to the interaction of the fluid with the shaft which acts as a
rotating cylinder. Reversing the rotation direction would result in the wake being skewed towards
y- side providing a mirror image of the current profile. The asymmetry observed in the simulation,
however, is not as pronounced as shown by the experiment. Asymmetry in TKE is also captured
well by ALM, but TKE is focused in the region around the turbine centre. This is in contrast to the
experimental results which show TKE almost absent at the centre with large spikes around the
blades as a result of strong blade tip vortex shedding. Magnitude wise, mean velocity relates well
with the experiment, whereas, TKE does not give good agreement in the y- region.

Figure 25: Comparison of mean velocity (left) and TKE (right) profiles at x/D=1 and z/H=0 based on experimental and
OpenFOAM k- ε uRANS- ALM using all correction models

Mean velocity and TKE contours at x/D=1 are shown in Figure 26. This further strengthens the
above-mentioned analysis that though some extent of asymmetry can be observed, the high level
of variation in velocity and TKE present in the experimental results is not captured. Note that for

43
Thesis
MSc Sustainable Energy Futures Imperial College London

contours, y-direction coordinates are used on x-axis instead of its non-dimensional version y/R.
The same approach would be applied to the other contour plots as well.

Figure 26: Mean velocity (top) and TKE (bottom) contours at x/D=1 using OpenFOAM k- ε uRANS- ALM

Now, the effect on mean velocity and TKE is analysed for each model. AL code used incorporated
added mass with no option provided to turn it off. So, added mass effects were not analysed
separately as it required modification of the code. Also, turbine parameters showed negative
values when the models were removed, so the effect of models on wake was analysed only.
Figure 27 shows how mean velocity changes when one by one each model is removed. It can be
seen that dynamic stall plays an essential role in bringing the results close to the experimental
values. Removing flow curvature effects slightly increases the mean velocity of the wake.

44
Thesis
MSc Sustainable Energy Futures Imperial College London

Whereas, removal of blade end effects reduces the wake velocity. This is because blade end
effects reduce the lift generated near the blade ends. Removal of blade end effects results in
more energy capture which appears as decreased velocity of the wake. A similar trend is observed
for TKE as shown in Figure 28. The correction effect of dynamic stall is even more prominent for
TKE as dynamic stall causes vortex shedding leading to higher TKE. Other models do not have
a reasonable effect on TKE similar to what was observed in the case of mean velocity.

Figure 27: Contribution of the correction models towards mean velocity profile using OpenFOAM k- ε uRANS- ALM,
DS=Dynamic stall, FC=Flow curvature, EE=End effects

Figure 28: Contribution of the correction models towards TKE profile using OpenFOAM k- ε uRANS- ALM,
DS=Dynamic stall, FC=Flow curvature, EE=End effects

45
Thesis
MSc Sustainable Energy Futures Imperial College London

3.3.3.2 k- ω SST uRANS- ALM


Cp and CD values for k- ω SST were 0.28 and 0.94, respectively, which are even closer to
experimental values showing k- ω SST gives better performance characteristics as compared to
k- ε. Figure 29 shows variation of mean velocity and TKE at x/D=1 and z/H=0. For mean velocity,
the asymmetry can be seen to be even more pronounced than k- ε. Moreover, the values are
almost an exact match to the experimental values. However, there are two slight discrepancies.
First, the wake is not that much spread as seen in the experiment. Secondly, the flow acceleration
on the outer side of the blades is less strong. But other than these slight differences, k- ω SST
produced excellent results for turbine characteristics and mean velocity. TKE, however, was found
to be more than an order of magnitude lower than the experimental results. This can be attributed
to the inappropriate value of ω used as the initial value. Though less sensitive than Wilcox’s k- ω,
k- ω SST is still sensitive to initial free stream ω values as a percentage of its results are based
on k- ω results (the percentage increases as wall Is approached). It can be seen in Figure 30 that
using k- ω SST results in the more realistic spread of wake velocity and TKE than k- ε. Note that
the rendering scale used for TKE is an order of magnitude smaller than the one used for k- ε for
clarity and same approach will be applied for other k- ω SST TKE contour plots as well. The shape
of the wake has shifted from a perfect circle, observed in the case of k- ε, to a circle-rectangle
mixed shape which is closer to the experimental results. Thus, as we move above or below the
centre line in the vertical direction, k- ω performs much better in capturing the qualitative aspects
of the wake as compared to k- ε. Also, TKE can be seen to be focused near the wake edge instead
of its centre.

Figure 29: Comparison of mean velocity (left) and TKE (right) profiles at x/D = 1 and z/H = 0 based on experimental
and OpenFOAM k- ω uRANS- ALM using all correction models

46
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 30: Mean velocity (top) and TKE (bottom) contours at x/D=1 using OpenFOAM k- ω uRANS- ALM

For correction models, the same trend is observed as k- ε with dynamic stall offering the biggest
contribution and other models not altering the results much as shown in Figure 31 and Figure 32.

47
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 31: Contribution of the correction models towards mean velocity profile using OpenFOAM k- ω uRANS- ALM,
DS=Dynamic stall, FC=Flow curvature, EE=End effects

Figure 32: Contribution of the correction models towards TKE profile using OpenFOAM k- ω uRANS- ALM,
DS=Dynamic stall, FC=Flow curvature, EE=End effects

Concerning simulation time, uRANS-ALM in OpenFOAM converged to solutions four orders of


magnitude faster than the uRANS blade resolved simulations (Bachant & Wosnik, 2016). This
shows the potential of ALM to be an inexpensive computational tool.

3.4 Fluidity
Fluidity CFD code was introduced in Chapter 2. Fluidity supports unstructured adaptive meshing
which refines the mesh in the region where more physics is occurring. Thus, fluidity is expected
to give better-resolved details for the wake. The AL code written for OpenFOAM could not be
used in Fluidity. So, an AL code, previously written for HATs, was modified to model VATs as
well. Similar turbulence models, i.e. 3D k- ε and k- ω SST uRANS were used in Fluidity as well.

48
Thesis
MSc Sustainable Energy Futures Imperial College London

3.4.1 Modelling
The new AL code, written in Fortran 90 for Fluidity (Deskos, Abolghasemi & Piggott, 2017), was
similar in general to the code used in OpenFOAM. However, there were some minor modifications
incorporated. Turbine blades and tower were divided into 20 and 21 elements instead of 14 and
20, respectively, to achieve better resolution. For simplicity, the struts were not modelled. For DS,
instead of using values against two Re only, a bigger set of foil parameters vs Re was used to get
more accurate parameters. Also, the Sheng et al. DS model was tested with different coefficients
in literature, but the default values were utilized in the end as they were found to give the best
results. OpenFOAM showed that the blade end effects did not contribute much to the results. So,
the blade end effects were omitted. Added mass correction was added to study its effect.

3.4.2 Meshing and verification


A domain and initial and boundary conditions similar to OpenFOAM were used. The domain was
discretised using unstructured finite elements of 2nd order polynomial by discontinuous Galerkin
method. Adaptivity was turned on, and minimum and maximum element size were specified. The
mesh at the start and end of a simulation is shown in Figure 33. Velocity contours are shown in
the figure to show how the mesh refines and adapts its shape around the wake. The mesh was
verified for grid independence before the simulations were performed. For this purpose, maximum
edge length (hmax) was kept constant at 0.5 m, and minimum edge length (hmin) was varied to
analyse the effect of mesh refinement on turbine and wake characteristics. Time step ∆t was kept
constant at 0.01 seconds similar to OpenFOAM (Bachant, Goude & Wosnik, 2016a).

OpenFOAM results were found to converge as the cell length was reduced to approximately 0.1
m, which was chosen as the cell length for the simulations. So, hmin=0.1 was chosen as the starting
point, and it was further refined to ensure the independence of the grid. Figure 34 and Figure 35
show results of grid refinement for k- ε. The values did not vary much for Cp, CD and mean
velocity, expect for hmin=0.00625, which indicates that the grid independent region is reached.
TKE, however, varied substantially with a local maxima at hmin=0.025. Local minima/maxima were
also observed at or around the same hmin for other quantities as well. Note that the time step ∆t
was kept constant at 0.01 seconds and the spatial grid was refined. So, the refinement of grid
beyond 0.025 led to CFL→1 resulting in divergence of results, with hmin=0.00625 giving CFL>1
and hence, arbitrary results. Based on this, hmin=0.025 was selected to be used for k- ε
simulations.

49
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 33: Cutaway of the initial (left) and final adapted (right) 3D mesh used in Fluidity

Figure 34: Cp and CD spatial resolution sensitivity results for Fluidity k- ε uRANS- ALM

Figure 35: Mean velocity (left) and TKE (right) spatial resolution sensitivity results for Fluidity k- ε uRANS- ALM

50
Thesis
MSc Sustainable Energy Futures Imperial College London

A similar set of values was used for hmin to check the grid independence for k- ω SST with results
shown in Figure 36 and Figure 37. In this case, the solver did not converge for edge lengths
smaller than 0.05 with ∆t fixed at 0.01 s. Again for Cp, CD and mean velocity, the results did not
vary much with hmin=0.05 giving values closest to the experimental data. TKE was observed to
increase drastically as the mesh was refined but the values were over an order of magnitude
smaller than the experimental results, consistent with OpenFOAM, and the spread of TKE in the
y-direction also did not change with grid refinement. Thus, TKE convergence was not considered
a deciding factor while choosing grid resolution and hmin=0.05 was used for k- ω SST simulations.

Figure 36: Cp and CD spatial resolution sensitivity results for Fluidity k- ω uRANS- ALM

Figure 37: Mean velocity (left) and TKE (right) spatial resolution sensitivity results for Fluidity k- ω uRANS- ALM

51
Thesis
MSc Sustainable Energy Futures Imperial College London

3.4.3 Results
3.4.3.1 k- ε uRANS- ALM
With all correction models included, Cp and CD values of 0.15 and 0.93 were obtained against
experimental values of 0.26 and 0.96, respectively. The values were again obtained by averaging
over the last cycle only to get rid of arbitrary values during the transient phase. C D is in good
agreement with the experiments, but the value of Cp is significantly lower. As will be discussed
later, dynamic stall could not be incorporated in Fluidity appropriately. Thus, the extra lift during
dynamic stall was not produced resulting in lower Cp. Figure 38 shows variation of mean velocity
and TKE at x/D=1 and z/H=0. The velocity deficit predicted is much lower than the experiment.
Also at x/D=1, no clear asymmetry can be observed in the mean velocity in the y-direction. The
spread is also quite small as compared to experimental results. It can be seen in Figure 39 that
the wake is contracting, instead of expanding, as it progresses downstream resulting in smaller
spread. As shown in Figure 38, TKE levels comparable to experimental values were achieved.
Asymmetry was also captured, but the curve was skewed in y+-direction instead of y--direction
seen in the experiment. Turbulence was also found to be less spread and focussed around the
turbine centre, whereas, the experiment showed the concentration of TKE around the blades.
Higher TKE concentration at the centre can also explain less spread of mean velocity in the wake.
TKE helps promote turbulent mixing leading to expansion of wake. Since TKE is focused at the
centre of the wake and not around the edges; more mixing is happening inside the wake than at
the edges. Thus the fluid in the wake does not interact much with the free-stream fluid, and no
expansion in the wake is observed. Contours for mean velocity and TKE at x/D=1 are shown in
Figure 40. The wake can be seen to have expanded in the z-direction to match the spread
observed in the experiment. This is different from the behaviour of wake in the y-direction which,
as discussed above, shows wake contraction. No TKE is found representing blade tip vortices.

Figure 38: Comparison of mean velocity (left) and TKE (right) profiles at x/D = 1 and z/H = 0 based on experimental
and Fluidity k- ε uRANS- ALM using all correction models

52
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 39: Mean velocity (top) and TKE (bottom) contours at z/H=0, viewing from top, using Fluidity k- ε uRANS- ALM

53
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 40: Mean velocity (top) and TKE (bottom) contours at x/D=1 using Fluidity k- ε uRANS- ALM

The impact of different correction models on results is now analysed. Figure 41 shows how these
models affect the values of Cp and CD. Flow curvature and added mass are found to have a
negligible effect on Cp. DS does not alter the results at all, even in the case of other results,
suggesting that it could not be incorporated in the model successfully. Figure 42 shows that AoA
surpasses dynamic stall initiation angle αds0 =17.91 several times showing that the blades undergo
significant dynamic stall. However, when the corresponding Cl were analysed, they corresponded
to the values for static foil even for α> αds0 showing that DS was not incorporated properly.

54
Thesis
MSc Sustainable Energy Futures Imperial College London

Removal of FC and AM led to an underestimation of CD values highlighting the importance of the


models for correct CD prediction.

Figure 41: Contribution of the correction models towards Cp and CD using Fluidity k- ε uRANS- ALM

Figure 42: Variation of AoA and the corresponding Cl vs turbine revolutions

Figure 43 shows how mean velocity changes with subsequent removal of correction models. As
expected, no difference is found with and without DS. So, a single curve is drawn for both the
cases. Though added mass affected CD significantly, it had no effect on mean velocity. So, a
separate curve could not be drawn for it. Removal of flow curvature gives somewhat higher values
for velocity deficit with a little bit of asymmetry induced as well. A similar trend for added mass

55
Thesis
MSc Sustainable Energy Futures Imperial College London

and dynamic stall is observed for TKE as shown in Figure 44. Exclusion of flow curvature reduces
TKE by a reasonable amount. In Fluidity, the effects of flow curvature are found to be more
apparent than OpenFOAM showing FC is probably better captured by Fluidity.

Figure 43: Contribution of the correction models towards mean velocity profile using Fluidity k- ε uRANS- ALM,
DS=Dynamic stall, FC=Flow curvature, AM=Added mass

Figure 44: Contribution of the correction models towards TKE profile using Fluidity k- ε uRANS- ALM, DS=Dynamic
stall, FC=Flow curvature, AM=Added mass

56
Thesis
MSc Sustainable Energy Futures Imperial College London

3.4.3.2 k- ω SST uRANS- ALM


Cp and CD values for k- ω SST were 0.13 and 0.93, respectively, which are almost the same as
that of k- ε with a minute difference in the value of Cp. CD value is quite close to the experimental
result, but Cp was reduced to almost half the experimental value. Figure 45 shows variation of
mean velocity and TKE at x/D=1 and z/H=0. The velocity deficit is closer to experimental values
as compared to k- ε model. Though the wake is not skewed towards any direction, a little
asymmetry can be observed in the velocity deficit, as well as, flow acceleration near the blades,
i.e. near y/R=±1. The spread of wake is also significantly less than that observed in experiments.
TKE produced is considerably lower than the experimental values. The same discrepancy was
observed for the OpenFOAM k- ω SST as well. Contours of velocity and TKE in x-y plane are
also shown in Figure 46. It can be seen clearly that due to the absence of dynamic stall, no
reasonable TKE is produced by the blades in the y- sides. TKE can be seen to be originating from
the fluid-blade interaction on y+ side and around y=0 only. The same behaviour could be noticed
in the case of k- ε but it was not as obvious. This lack of generation of TKE on one side results in
wake contraction and skewness towards that side, y- side, due to turbulent mixing.

Figure 45: Comparison of mean velocity (left) and TKE (right) profiles at x=D = 1 and z=H = 0 based on experimental
and Fluidity k- ω uRANS- ALM using all correction models

Contours for mean velocity and TKE at x/D=1 are shown in Figure 47. It can be seen that the
velocity contours are not a regular shape and qualitatively match the experimental results.
Significant mean vertical velocity is also observed around the blade ends on the y+ side but it is
directed out of the wake instead of towards the centre of the turbine, which is the momentum
deficit region. An expansion in the z-direction is also evident. TKE contours, though having very
low values, can be seen to be resolved with high detail and concentrated around the wake edge
instead of at the centre.

57
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 46: Mean velocity (top) and TKE (bottom) contours at z/H=0 using Fluidity k- ω uRANS- ALM

58
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 47: Mean velocity (top) and TKE (bottom) contours at x/D=1 using Fluidity k- ω uRANS- ALM

Different correction models show a behaviour almost similar to that observed for k- ε. Figure 48
shows the results for Cp and CD similar to k- ε. Figure 49 and Figure 50 shows different models
contribution to mean velocity and TKE, respectively. For mean velocity, removal of flow curvature
reduces the velocity deficits, improves the spread and induces some skewness towards y+ side.
For TKE, removal of flow curvature makes wake more asymmetric as well. The effect is however
not as pronounced as in for the case of k- ε.

59
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 48: Contribution of the correction models towards Cp and CD using Fluidity k- ω uRANS- ALM

Figure 49: Contribution of the correction models towards mean velocity profile using Fluidity k- ω uRANS- ALM,
DS=Dynamic stall, FC=Flow curvature, AM=Added mass

Figure 50: Contribution of the correction models towards TKE profile using Fluidity k- ω uRANS- ALM, DS=Dynamic
stall, FC=Flow curvature, AM=Added mass

60
Thesis
MSc Sustainable Energy Futures Imperial College London

With respect to simulation time, uRANS-ALM in Fluidity converged to solutions two to three
times as fast as the blade resolved simulations (Bachant & Wosnik, 2016). Slow convergence of
Fluidity as compared to OpenFOAM is due to the implementation of finite element discretisation
using discontinuous unstructured mesh instead of structured finite volume mesh. Also, Fluidity
has not been as optimised as OpenFOAM has been over years. Without adaptivity, Fluidity
would have taken even longer.

3.5 Validation
The results for uRANS-ALM simulations in OpenFOAM and Fluidity using k- ε and k- ω SST
turbulence models were presented and analysed above. Now a quantitative and qualitative
comparison is made between turbine and wake characteristics results from the experiment, and
OpenFOAM and Fluidity to assess the validity of the simulation results using all correction models.

3.5.1 Quantitative validation

For quantitative comparison, absolute and percentage relative errors were calculated. For Cp and
CD, the approach was straight forward and the equations used were:

𝐶𝑥𝑎𝑏𝑠 𝑒𝑟𝑟𝑜𝑟 = |𝐶𝑥,𝐸𝑥𝑝 − 𝐶𝑥,𝑆𝑖𝑚 | (58)

𝐶𝑥𝑎𝑏𝑠 𝑒𝑟𝑟𝑜𝑟
𝐶𝑥𝑟𝑒𝑙 𝑒𝑟𝑟𝑜𝑟 = 𝑥 100% (59)
𝐶𝑥,𝐸𝑥𝑝

Where 𝑥 = 𝑝 𝑜𝑟 𝐷 for coefficient of power and drag, respectively. | | represents omission of the
sign of the answer. 𝐸𝑥𝑝 and 𝑆𝑖𝑚 represent experimental and simulated values, respectively.

For wake characteristics, the difference between the experiment and simulation curves was found
at points with y/R coordinates corresponding to that of the experimental data set. This gave the
absolute error at these coordinates. Mean (M) of the absolute values of this absolute error is then
calculated to get a single number representing the absolute error. Absolute values are used so
that negative and positive errors do not cancel out each other. Standard deviation (SD) was also
calculated to get the spread of absolute error. So, for Umean and TKE, following equations were
used to calculate absolute error:

𝑁
1
|𝑀𝑥𝑎𝑏𝑠 𝑒𝑟𝑟𝑜𝑟 | = ∑|𝑋𝐸𝑥𝑝,𝑖 − 𝑋𝑆𝑖𝑚,𝑖 | (60)
𝑁
𝑖=1

61
Thesis
MSc Sustainable Energy Futures Imperial College London

Where 𝑥, 𝑋 = 𝑈𝑚𝑒𝑎𝑛 or 𝑇𝐾𝐸, 𝑁 is equal to total points in the data set.

𝑁
1 2
𝑆𝐷𝑥𝑎𝑏𝑠 𝑒𝑟𝑟𝑜𝑟 = √ ∑((𝑋𝐸𝑥𝑝,𝑖 − 𝑋𝑆𝑖𝑚,𝑖 ) − 𝑀𝑥𝑎𝑏𝑠 𝑒𝑟𝑟𝑜𝑟 ) (61)
𝑁
𝑖=1

For wake characteristics, the percentage relative error for Umean and TKE was found by:

|𝑀𝑥𝑎𝑏𝑠 𝑒𝑟𝑟𝑜𝑟 |
|𝑀𝑥𝑟𝑒𝑙 𝑒𝑟𝑟𝑜𝑟 | = 𝑥100% (62)
(U∞ − UExp,min ) 𝑜𝑟 TKEExp,max

|𝑆𝐷𝑥𝑎𝑏𝑠 𝑒𝑟𝑟𝑜𝑟 |
| 𝑆𝐷𝑥𝑟𝑒𝑙 𝑒𝑟𝑟𝑜𝑟 | = 𝑥100% (63)
(U∞ − UExp,min ) 𝑜𝑟 TKEExp,max

For mean velocity, the minimum velocity in the wake is subtracted from the free-stream velocity
(U∞) because it is the amount of velocity deficit that needs to be compared, and greater the
difference between free-stream and wake velocity, the greater is the velocity deficit. So Eq (62)
is based on the comparison between the error in the wake velocity estimation as compared to the
maximum wake deficit seen in experiments. Another important point to note is that instead of
calculating relative error at each point and then taking the mean and SD, mean and SD of absolute
error was converted to the relative form. This is because the relative error weightage should not
be the same for every point along the y-direction. For example, in the region of y/R= [-1,1], the
velocity deficit has higher values, greater than 0.5, and a relative error with respect to
experimental values in this region will communicate the correct information regarding the level of
inaccuracy. However, in the area outside y/R= [-1,1], the velocity deficit or flow acceleration does
not have substantial values and are less than 0.1. A slight difference in simulated results in this
region can translate into a huge relative error. A mean or SD of such results would be representing
the accuracy of results outside y/R= [-1,1] region, whereas, our priority is the region right behind
the turbine where actual flow physics of the near wake occurs. UExp,min value corresponding to
maximum velocity deficit, was chosen as a representative value of y/R= [-1,1] region for simplicity
of calculations. The same argument applies to TKEExp,max. Absolute and percentage relative
errors are shown in Table 2-Table 5.

62
Thesis
MSc Sustainable Energy Futures Imperial College London

k- ε k- ω SST

Cp CD Umean TKE Cp CD Umean TKE

Fluidity 0.11 0.03 0.24 0.020 0.13 0.03 0.22 0.011

OpenFOAM 0.04 0.03 0.14 0.013 0.03 0.02 0.08 0.013

Table 2: Mean of absolute error in simulation results with respect to experimental results

k- ε k- ω SST

Cp CD Umean TKE Cp CD Umean TKE

Fluidity - - 0.26 0.020 - - 0.26 0.015

OpenFOAM - - 0.15 0.017 - - 0.08 0.015

Table 3: Standard deviation of absolute error in simulation results with respect to experimental results

k- ε k- ω SST

Cp CD Umean TKE Cp CD Umean TKE

Fluidity 42.3 3.1 31.0 28.9 50.0 3.1 28.0 15.8

OpenFOAM 15.4 3.1 17.6 19.1 11.5 2.0 10.4 19.0

Table 4: Mean of percentage relative error in simulation results with respect to experimental results

k- ε k- ω SST

Cp CD Umean TKE Cp CD Umean TKE

Fluidity - - 32.7 35.9 - - 33.6 21.9

OpenFOAM - - 18.9 24.9 - - 10.4 21.4

Table 5: Standard deviation of percentage relative error in simulation results with respect to experimental results

63
Thesis
MSc Sustainable Energy Futures Imperial College London

From Table 2 and Table 4 it can be concluded that quantitatively speaking:

 All models capture CD very accurately


 k- ω SST performs better than k- ε in general
 Fluidity performs worse than OpenFOAM in general with abnormally inaccurate results
for Cp
 k- ω SST using OpenFOAM gives the best overall results

Table 3 and Table 5 show SD values close to mean values suggesting a large spread of error
values. This is particularly true for Fluidity k- ε TKE results, in which sufficient TKE was generated
but in the “wrong places” as compared to the experiment and thus still resulting in a significant
amount of error. Most of the shortcomings in the quantitative accuracy of Fluidity results can be
attributed to the incorrect incorporation of the DS model. This is further evident by the fact that
without DS the peak of mean velocity deficit is the same for the two codes though the spread is
still different. Also, struts and strut connections were not modelled for simplicity in Fluidity. Their
exlusion can also explain the lower values shown by Fluidity (Rawlings, 2008).

3.5.2 Qualitative validation


As discussed, even though Fluidity k- ε TKE results visually seemed to be the closest to the
experimental results, they gave the largest value of error. This is due to the lack of overlap
between the experimental and Fluidity results. Moreover, it was seen previously that Fluidity was
able to capture the flow physics more accurately than OpenFOAM. So, it is important to compare
the results qualitatively as well before firm conclusions are drawn regarding the validity of the
simulation results. For this purpose, some of the figures already shown in the section “results
and analysis” would be reproduced for the ease of understanding. Figure 51 shows mean velocity
contours in the top half of the plane at x/D=1 for all simulations and the experiment using the
same scale for rendering. OpenFOAM is found to be better able to capture the spread of wake.
The wake results using Fluidity are less spread in the y-direction and more spread in the z-
direction as compared to experimental results. However, the “shape” of the wake is better
captured by Fluidity, giving a rectangular wake instead of a circular one. Specifically, note the dip
in the velocity near the top of the turbine for Fluidity k- ω SST results, which is in accordance with
the experimental results.

64
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 51: Comparison of mean velocity contours at x/D=1 using OpenFOAM k- ε uRANS-ALM (top-left), Fluidity k- ε
uRANS-ALM (top- right), OpenFOAM k- ω uRANS-ALM (centre-left), Fluidity k- ω uRANS-ALM (centre-right) and
experimental (bottom) results

Figure 52 shows TKE contours in the top half of the plane at x/D=1 for all simulations and
experiments. Note that the scale of rendering for TKE based on k- ω SST has been reduced by
an order of magnitude for visual clarity. Also, simulation results show TKE contour, whereas,
experimental results show k/0.5U∞ contours. It is stated here again that this is a qualitative
analysis only and no quantitative results should be deduced from the figures in this section.
Regarding spread and shape of the wake, the same trend is observed here as was seen for the
mean velocity. Another important observation is that the concentration of TKE at the wake edge,
instead of the centre, is captured much better by k- ω SST for both CFD codes, with the Fluidity
results being more prominent. All the simulations predict higher concentration of TKE on the y+
side, whereas, experimental results show concentration on the y- side. TKE due to blade tip and
dynamic stall vortex shedding is missing from all simulation results.

65
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 52: Comparison of TKE contours at x/D=1 using OpenFOAM k- ε uRANS-ALM (top-left), Fluidity k- ε uRANS-
ALM (top- right), OpenFOAM k- ω uRANS-ALM (centre-left), Fluidity k- ω uRANS-ALM (centre-right) and
experimental (bottom) results

Results from experiments are not available for wake characteristics in the x-y plane. So, a
comparison is drawn between the simulations only to see the detail of flow captured by each code.
Figure 53 shows velocity contours on the x-y plane through z/H=0. Distinct trailing edge vortices,
a prominent feature of VAT wakes, can be seen in Fluidity results, whereas, no such vortices can
be seen in OpenFOAM results showing that Fluidity can better capture the actual physics
happening. One reason for this could be that Fluidity uses a more refined mesh, but it is the very
feature of mesh adaptivity in Fluidity which refines the mesh enough in this region to capture this
detail. Fluidity results can also be seen to be skewed towards y- sides which could not be seen in
the y-z contour plot as it was taken at x/D=1 only. The wake also seems to be contracting which
is not what would be observed in reality (Mendoza & Goude, 2017). As discussed before, these
behaviours can be a result of inappropriate incorporation of dynamic stall. Using a longer mean
might also result in better establishment of wake width.

66
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 53: Comparison of mean velocity contours at z/H=0 using OpenFOAM k- ε uRANS-ALM (top-left), Fluidity k- ε
uRANS-ALM (top- right), OpenFOAM k- ω uRANS-ALM (bottom-left) and Fluidity k- ω uRANS-ALM (bottom-right)
results

Figure 54 shows TKE contours on the x-y plane through z/H=0. Rendering scale for k- ω SST has
again been reduced by an order of magnitude to enhance clarity. Though k- ω SST produces
much less TKE, it resolves TKE much better than k- ε. Also, more concentrated TKE can be seen
to appear as wake progresses downstream in the case of k- ω SST. Fluidity results show the
stronger generation of TKE by blades, whereas, OpenFOAM shows very diffused results. Distinct
vortices being generated by blades are again clearly visible in Fluidity results.

67
Thesis
MSc Sustainable Energy Futures Imperial College London

Figure 54: Comparison of TKE contours at z/H=0 using OpenFOAM k- ε uRANS-ALM (top-left), Fluidity k- ε uRANS-
ALM (top- right), OpenFOAM k- ω uRANS-ALM (bottom-left) and Fluidity k- ω uRANS-ALM (bottom-right) results

So, even though OpenFOAM gives better quantitative results, Fluidity can be seen to resolve the
flow details and give qualitative results much closer to the experimental results.

68
Thesis
MSc Sustainable Energy Futures Imperial College London

Chapter 4. Conclusion

4.1 Summary
A huge increase in wind and tidal energy demand is expected in the near future owing to the
decarbonising targets set by countries around the globe. VATs can help in fulfilling this
requirement by offering application areas not feasible for HATs. This necessitates the
establishment of high fidelity and low computationally expensive analysis tools which can reduce
the uncertainty during the designing of VATs. The actuator line model, a potential such tool, was
analysed in two open-source CFD codes, OpenFOAM and Fluidity. OpenFOAM is a highly
optimised widely used code, whereas, Fluidity offers adaptive mesh features. An existing AL code
was used in OpenFOAM, whereas, a HATs ALM code was modified to include VATs for
implementation in Fluidity. The final ALM incorporated different correction models like dynamic
stall, flow curvature, added mass and blade end effects on top of conventional blade element
representation to apply more realistic loadings on the fluid. uRANS turbulence models, closed
using k- ε and k- ω SST, were used for simulations and experimental data was used to validate
the results. Choice of turbulence models was based on their proven track record for a wide variety
of cases and reasonable solution times with little compromise on the accuracy of results. The
coefficients of power and drag were used to evaluate the turbine’s performance. Both,
OpenFOAM and Fluidity, predicted coefficient of drag with good accuracy. OpenFOAM gave
slightly higher values for coefficient of power, whereas, Fluidity greatly underestimated power
extraction- giving almost half the values for coefficient of power. The wake, being a very complex
flow structure, demands several parameters to be analysed at different downstream distances
from the turbine for a thorough analysis. However, for the purpose of validation, mean velocity
and TKE was investigated in the near wake only. k- ω SST performed better in capturing mean
velocity profile for both codes with k- ε giving better results for TKE values. OpenFOAM captured
mean velocity values better, whereas, Fluidity gave higher and more realistic TKE values. Overall,
OpenFOAM gave a better spread of wake resulting in less mean and standard deviation of error.
Though Fluidity produced better “visual” results in some cases, like in the case of TKE, it focused
the results at the centre of the wake resulting in large errors. To inspect the codes’ ability to
capture wake correctly, the quantitative analysis was supplemented with a qualitative analysis.
Qualitative analysis showed that though OpenFOAM gave more quantitatively correct results,
Fluidity was better able to capture the details of the wake structure and underlying physical
phenomena. The spread of the wake was better captured by OpenFOAM in both, y and z,

69
Thesis
MSc Sustainable Energy Futures Imperial College London

directions but Fluidity gave a rectangular wake instead of OpenFOAM’s circular one. Also,
turbulence generation and vortex shedding was clearly seen in Fluidity’s results, whereas,
OpenFOAM was unable to resolve these details.

Effect of different correction models was also analysed to see their relative importance for
accurate prediction of VAT characteristics. OpenFOAM simulation revealed dynamic stall to be of
vital importance as it was found to alter results by up to 20% in case of mean velocity and up
to100% in the case of TKE. Dynamic stall model could not be incorporated successfully in the
Fluidity for the reasons not known. The difference in velocity profiles generated by OpenFOAM
and Fluidity can be attributed to this, as with dynamic stall inactivated, OpenFOAM’s mean
velocity profiles were just a wider spread version of Fluidity’s results. Added mass improved
turbine performance characteristic slightly. Flow curvature proved to be significant only in the case
of Fluidity k- ε simulation. Blade end effects were not found to make a tangible difference in any
case. Overall, dynamic stall was found to be the only model which can affect numerical values
significantly.

The uRANS-ALM implementation in Fluidity and OpenFOAM resulted in three and four orders of
magnitude reduction in computational expense as compared to blade resolved simulations.
uRANS-ALM implementation in OpenFOAM can be used to analyse VATs on personal
computers, whereas, Fluidity will require low-end supercomputers. Regarding accuracy, both
OpenFOAM and Fluidity gave mixed results when compared with experimental results. Moreover,
OpenFOAM performed better quantitatively, whereas, Fluidity performed better qualitatively
showing that uRANS-ALM can capture both of these aspects. With DS implemented correctly in
Fluidity, it is expected to even match the quantitative results produced by OpenFOAM. However,
this reasoning cannot be used to declare the uRANS-ALM a valid modelling technique for VATs
until a single ALM captures both of these aspects. The study conducted was not comprehensive
enough, due to the limited time of three months, and more research is required, especially in the
areas where the simulation results are not in agreement with the experimental results, to decide
the validity of uRANS-ALM as a high-fidelity tool. The further research to be pursued is highlighted
in the next section.

4.2 Future Work


The study was able to provide key insights regarding the implementation of uRANS-ALM in
OpenFOAM and Fluidity. As discussed, a continuation of this study is required to establish a
conclusive result. The tasks needed to make this study complete and conclusive are listed below:

70
Thesis
MSc Sustainable Energy Futures Imperial College London

 OpenFOAM lacked in capturing the high resolution of fluid flow through and behind the
turbine. So, OpenFOAM studies need to be conducted with higher mesh resolution to
establish if the mesh resolution can resolve this issue or if there is some fundamental
difference in ALM implementation in the two software. As OpenFOAM does not support
adaptivity, the high resolution will greatly affect simulation times as the refined mesh would
be used throughout the simulation. So, this increase in simulation time needs to be
weighed against better flow description as well.
 Dynamic stall incorporation in Fluidity needs to be correctly implemented. Struts need to
be represented in the ALM of Fluidity as well. With these modifications, Fluidity is expected
to produce quantitative results very close to OpenFOAM.
 Trailing edge vortices were captured by ALM. However, blade tip and dynamic stall
vortices were not captured. Incorporation of turbulence injection models used by James,
Seetho, Jones, et al. (2010) in ALM can result in the capture of these vortices.
 A sensitivity study with regards to the effect of initial ω values on TKE in the near wake
also needs to be undertaken. TKE of the order of experimental values was produced at a
much larger downstream distance using k- ω SST. Selection of a more appropriate ω can
resolve this.
 Though a few different combinations of coefficients for Sheng et al. DS model were used,
a comprehensive study analysing the difference in DS model results involving a larger set
of coefficients should be carried out to make sure that the correct set of coefficients is
being used.

All these studies are expected to result in better conformity between experimental and simulation
results. Once the ALM has been validated conclusively, following future works can be undertaken
with regards to application of uRANS-ALM:

 Analysis of a VAT in an actual operation environment. The environment can be a rooftop


or a water channel. This will help in analysing the effects of turbulence in the wind and
roughness of river/sea bed on VAT performance and wake recovery. ALM used in
OpenFOAM might need to be adjusted to use unstructured mesh to capture the flow over
rough bed more accurately.
 The analysis of a stand-alone turbine can then be extended to array analysis. Here
Fluidity’s adaptive mesh is expected to provide a considerable advantage. As the array
modelling requires a vast domain to be used, adaptive mesh refinement only in the regions
where it is needed can offer a significant reduction in computational expense.

71
Thesis
MSc Sustainable Energy Futures Imperial College London

References
Abolghasemi, M.A., Piggott, M.D., Spinneken, J., Vire, A., et al. (2016) Simulating tidal turbines
with multi-scale mesh optimisation techniques. Journal of Fluids and Structures. 6696–90.

Alaimo, A., Esposito, A., Messineo, A., Orlando, C., et al. (2015) 3D CFD Analysis of a Vertical
Axis Wind Turbine. Energies. [Online] 8 (4), 3013–3033. Available from:
doi:10.3390/en8043013.

Altair (2014) European Altair Technology Conference. In: Multiphysics & CFD. 2014 Munich. p.

AMCG (2014) Fluidity manual v4.1.10. [Online] (April), 1–321. Available from:
doi:10.6084/m9.figshare.1089457.

ANSYS FLUENT (2006) Introductory FLUENT Notes FLUENT v6.3.

Archer, C.L. & Jacobson, M.Z. (2005) Evaluation of global wind power. J . Geophys . Res.
[Online] 110. Available from: doi:10.1029/2004JD005462.

Baba-Ahmadi, M.H. & Dong, P. (2017) Numerical simulations of wake characteristics of a


horizontal axis tidal stream turbine using actuator line model. Renewable Energy. [Online]
113669–678. Available from: doi:10.1016/j.renene.2017.06.035.

Bachant, P., Goude, A. & Wosnik, M. (2016a) Actuator line modeling of vertical-axis turbines.

Bachant, P., Goude, A. & Wosnik, M. (2016b) turbinesFoam: v0.0.7. Zenodo. 2016.

Bachant, P. & Wosnik, M. (2014) Characterizing the near-wake of a cross-flow turbine. Journal
of Turbulence. 5248 (0), 11.

Bachant, P. & Wosnik, M. (2016) Modeling the near-wake of a vertical-axis cross-flow turbine
with 2-D and 3-D RANS. Journal of Renewable and Sustainable Energy. [Online] 8 (5).
Available from: doi:10.1063/1.4966161.

Badreddinne, K., Ali, H. & David, A. (2005) Optimum project for horizontal axis wind turbines
OPHWT. Renewable Energy. [Online] 30 (13), 2019–2043. Available from:
doi:10.1016/j.renene.2004.12.004.

Balduzzi, F., Bianchini, A., Carnevale, E.A., Ferrari, L., et al. (2012) Feasibility analysis of a
Darrieus vertical-axis wind turbine installation in the rooftop of a building. Applied Energy.
[Online] 97921–929. Available from: doi:10.1016/j.apenergy.2011.12.008.

72
Thesis
MSc Sustainable Energy Futures Imperial College London

Balduzzi, F., Bianchini, A., Maleci, R., Ferrara, G., et al. (2016) Critical issues in the CFD
simulation of Darrieus wind turbines. Renewable Energy. [Online] 85419–435. Available
from: doi:10.1016/j.renene.2015.06.048.

Batchelor, G.K. (1967) An Introduction to Fluid Dynamics. [Online]. Cambridge University Press.
Available from: doi:10.1063/1.3060769.

Beddoes, T.S. (1993) A Third Generation Model for Unsteady Aerodynamics and Dynamic Stall.
Westland Helicopter Limited.

Beddoes, T.S. (1982) Practical computation of unsteady lift. ASSOCIATION AERONAUTIQUE


ET ASTRONAUTIQUE DE FRANCE.

Betz, A. (1919) Schraubenpropeller mit geringstem Energieverlust. Mit einem Zusatz von l.
Prandtl. Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen,
Mathematisch-Physikalische Klasse. 193–217.

Betz, A. (1921) The theory of the screw propeller. NACA Tech Notes. 831–19.

Boudreau, M. & Dumas, G. (2015) Wake Analysis of Various Hydrokinetic Turbine Technologies
through Numerical Simulations. In: 62nd CASI Aeronautics Conference and AGM 3rd
GARDN Conference. 2015 p.

Brochier, G., Fraunie, P., Beguier, C. & Paraschivoiu, I. (1986) Water channel experiments of
dynamic stall on Darrieus wind turbine blades. Journal of Propulsion and Power. [Online] 2
(5), 445–449. Available from: doi:10.2514/3.22927.

Burgerscentrum, J.M. (2015) Finite element methods for the incompressible Navier-Stokes
equations.

Castelli, M. & Benini, E. (2011) Comparison between Lift and Drag-Driven VAWT Concepts on
Low-Wind Site AEO.

CFD online (2017) CFD Online - Turbulence Properties, Conversions & Boundary Estimations.
[Online]. 2017. Available from: https://www.cfd-online.com/Tools/turbulence.php
[Accessed: 27 June 2017].

Committee on Climate Change (2008) Carbon budgets: how we monitor emissions targets |
Committee on Climate Change. [Online]. 2008. Available from:
https://www.theccc.org.uk/tackling-climate-change/reducing-carbon-emissions/carbon-

73
Thesis
MSc Sustainable Energy Futures Imperial College London

budgets-and-targets/ [Accessed: 28 March 2017].

COMSOL (2017) Detailed Explanation of the Finite Element Method (FEM). [Online]. 2017.
Available from: https://www.comsol.com/multiphysics/finite-element-method [Accessed: 29
March 2017].

Dabiri, J.O. (2011) Potential order-of-magnitude enhancement of wind farm power density via
counter-rotating vertical-axis wind turbine arrays. Journal of Renewable and Sustainable
Energy. [Online] 3 (4), 43104. Available from: doi:10.1063/1.3608170.

Davies, D.R., Wilson, C.R. & Kramer, S.C. (2011) Fluidity: A fully unstructured anisotropic
adaptive mesh computational modeling framework for geodynamics. Geochemistry,
Geophysics, Geosystems. [Online] 12 (6). Available from: doi:10.1029/2011GC003551.

Deskos, G., Abolghasemi, M.A. & Piggott, M.D. (2017) Wake predictions from two turbine
parameterisation models using mesh-optimisation techniques. In: EWTEC. 2017 p.

Deskos, G. & Piggott, M.D. (2017) Mesh-adaptive simulations of horizontal-axis turbine arrays
using the actuator line method. under preparation for submission to Wind Energy journal.

Dobrev, I., Massouh, F. & Rapin, M. (2007) Actuator surface hybrid model. Journal of Physics:
Conference Series. [Online] 75 (1), 12019. Available from: doi:10.1088/1742-
6596/75/1/012019.

Draper, M., Guggeri, A. & Usera, G. (2016) Validation of the Actuator Line Model with coarse
resolution in atmospheric sheared and turbulent inflow. Journal of Physics: Conference
Series. [Online] 753 (8), 82007. Available from: doi:10.1088/1742-6596/753/8/082007.

Drzewiecki SK (1910) Theory of air propellers and the method of their calculation. foreward by
N.B. Delone. Publ R.K. Lubkovsky.

Farrell, P.E. & Maddison, J.R. (2011) Conservative interpolation between volume meshes by
local Galerkin projection. Computer Methods in Applied Mechanics and Engineering.
[Online] 200 (1–4), 89–100. Available from: doi:10.1016/j.cma.2010.07.015.

Ferreira, C.J.S., van Bussel, G.J.W., van Kuik, G.A.M. & Scarano, F. (2011) On the Use of
Velocity Data for Load Estimation of a VAWT in Dynamic Stall. Journal of Solar Energy
Engineering. [Online] 133 (1), 11006. Available from: doi:10.1115/1.4003182.

Flaherty, J.E. (2010) Adaptive Finite Element Techniques.

74
Thesis
MSc Sustainable Energy Futures Imperial College London

Froude, R.E. (1911) The acceleration in front of a propeller. Trans Inst Naval Archit . 53139–
182.

Glauert, H. (1935) Airplane Propellers : Aerodynamic Theory. Springer. [Online] Available from:
doi:10.1007/978-3-642-91487-4_3.

Goude, A. (2012) Fluid Mechanics of Vertical Axis Turbines - Simulations and Model
Development.

Gregory, P.., Ingram, J.S.. & Brklacich, M. (2005) Climate change and food security.
Philosophical Transactions of the Royal Society of London B: Biological Sciences. 360
(1463).

Guttman, A., Antonin, Guttman & Antonin (1984) R-trees. In: Proceedings of the 1984 ACM
SIGMOD international conference on Management of data - SIGMOD ’84. [Online]. 1984
New York, New York, USA, ACM Press. p. 47. Available from:
doi:10.1145/602259.602266.

Hiester, H.R., Piggott, M.D., Farrell, P.E. & Allison, P.A. (2014) Assessment of spurious mixing
in adaptive mesh simulations of the two-dimensional lock-exchange. Ocean Modelling.
[Online] 7330–44. Available from: doi:10.1016/j.ocemod.2013.10.003.

Howell, R., Qin, N., Edwards, J. & Durrani, N. (2010) Wind tunnel and numerical study of a
small vertical axis wind turbine. Renewable Energy. [Online] 35 (2), 412–422. Available
from: doi:10.1016/j.renene.2009.07.025.

Iaccarino, G., Ooi, A., Durbin, P.A. & Behnia, M. (2003) Reynolds averaged simulation of
unsteady separated flow. International Journal of Heat and Fluid Flow. [Online] 24 (2),
147–156. Available from: doi:10.1016/S0142-727X(02)00210-2.

IRENA (2015) Renewable Energy Capacity Statistics 2015.

Jacobson, M.Z. & Delucchi, M.A. (2011) Providing all global energy with wind, water, and solar
power, Part I: Technologies, energy resources, quantities and areas of infrastructure, and
materials. Energy Policy. [Online] 39 (3), 1154–1169. Available from:
doi:10.1016/j.enpol.2010.11.040.

James, S.C., Seetho, E., Jones, C. & Roberts, J. (2010) Simulating environmental changes due
to marine hydrokinetic energy installations. In: OCEANS 2010 MTS/IEEE SEATTLE.

75
Thesis
MSc Sustainable Energy Futures Imperial College London

[Online]. September 2010 IEEE. pp. 1–10. Available from:


doi:10.1109/OCEANS.2010.5663854.

Joukowsky, N.E. (1912) Vortex Theory of Screw Propeller. i Trudy Otdeleniya Fizicheskikh,
Nauk Obshchestva Lubitelei Estestvoznaniya. 16 (1), 1–31.

Kinzel, M., Mulligan, Q. & Dabiri, J.O. (2012) Energy exchange in an array of vertical-axis wind
turbines. Journal of Turbulence. [Online] 13N38. Available from:
doi:10.1080/14685248.2012.712698.

Kolmogorov, A.N. (1941) Dissipation of Energy in the Locally Isotropic Turbulence (1941).
Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering
Sciences. 434 (1890).

Kreft, S., Eckstein, D. & Melchios Inga (2017) Global Climate Risk Index 2017.

Kurien, S. & Taylor, M.A. (2005) Direct Numerical Simulations of Turbulence Data Generation
and Statistical Analysis. Los Alamos Science Number.29 (29).

Lam, H.F. & Peng, H.Y. (2016) Study of wake characteristics of a vertical axis wind turbine by
two- and three-dimensional computational fluid dynamics simulations. Renewable Energy.
[Online] 90386–398. Available from: doi:10.1016/j.renene.2016.01.011.

Leishman, J.G. (2002) Challenges in Modeling the Unsteady Aerodynamics of Wind Turbines.
American Institute of Aeronauticy and Astronautics. [Online] 37 (February), 1–28. Available
from: doi:10.1115/wind2002-37.

Leishman, J.G. & Beddoes, T.S. (1989) A Semi-Empirical Model for Dynamic Stall. Journal of
the American Helicopter Society. [Online] 34 (3), 3–17. Available from:
doi:10.4050/JAHS.34.3.

Lewis, M., Neill, S.P., Robins, P.E. & Hashemi, M.R. (2015) Resource assessment for future
generations of tidal-stream energy arrays. Energy. [Online] 83403–415. Available from:
doi:10.1016/j.energy.2015.02.038.

Li, W. (2014) Efficient Implementation of High-Order Accurate Numerical Methods on


Unstructured Grids. [Online] M123–125. Available from: doi:10.1007/978-3-662-43432-1.

Madsen, H.A., Mikkelsen, R., Øye, S., Bak, C., et al. (2007) A Detailed investigation of the
Blade Element Momentum (BEM) model based on analytical and numerical results and

76
Thesis
MSc Sustainable Energy Futures Imperial College London

proposal for modifications of the BEM model. Journal of Physics: Conference Series.
[Online] 7512016. Available from: doi:10.1088/1742-6596/75/1/012016.

Martínez-Tossas, L.A., Churchfield, M.J. & Leonardi, S. (2015) Large eddy simulations of the
flow past wind turbines: actuator line and disk modeling. Wind Energy. [Online] 18 (6),
1047–1060. Available from: doi:10.1002/we.1747.

Mendoza, V., Bachant, P., Wosnik, M., Goude, A., et al. (2016) Validation of an Actuator Line
Model Coupled to a Dynamic Stall Model for Pitching Motions Characteristic to Vertical
Axis Turbines. Journal of Physics: Conference Series. [Online] 753 (2), 22–43. Available
from: doi:10.1088/1742-6596/753/2/022043.

Mendoza, V. & Goude, A. (2017) Wake Flow Simulation of a Vertical Axis Wind Turbine Under
the Influence of Wind Shear. Journal of Physics: Conf. Series. [Online] 854. Available from:
doi:10.1088/1742-6596/854/1/012031.

Menter, F.R. (1994) Two-equation eddy-viscosity turbulence models for engineering


applications. AIAA Journal. [Online] 32 (8), 1598–1605. Available from:
doi:10.2514/3.12149.

NASA (2015) Incorrect Lift Theory. [Online]. 2015. Available from:


https://www.grc.nasa.gov/www/k-12/airplane/wrong1.html [Accessed: 28 March 2017].

Ouro, P. & Stoesser, T. (2017) An immersed boundary-based large-eddy simulation approach to


predict the performance of vertical axis tidal turbines. Computers & Fluids. [Online] 15274–
87. Available from: doi:10.1016/j.compfluid.2017.04.003.

Paraschivoiu, I. (1988) Double-multiple streamtube model for studying vertical-axis wind


turbines. Journal of Propulsion and Power. [Online] 4 (4), 370–377. Available from:
doi:10.2514/3.23076.

Pope, S.B. (2000) Turbulent flows. Cambridge University Press.

Prandtl, L. & Betz, A. (1927) Vier Abhandlungen zur Hydrodynamik und Aerodynamik.

Rawlings, G.W. (2008) Parametric characterization of an experimental vertical axis hydro


turbine. [Online] Available from: doi:10.14288/1.0068009.

Shamsoddin, S. & Porté-Agel, F. (2014) Large Eddy Simulation of Vertical Axis Wind Turbine
Wakes. Energies. [Online] 7 (2), 890–912. Available from: doi:10.3390/en7020890.

77
Thesis
MSc Sustainable Energy Futures Imperial College London

Sheldahl, R.E. (1981) Comparison of Field and Wind Tunnel Darrieus Wind Turbine Data.

Sheldahl, R.E. & Klimas, P.C. (1981) Aerodynamic characteristics of seven symmetrical airfoil
sections through 180-degree angle of attack for use in aerodynamic analysis of vertical
axis wind turbines. Chemistry &. [Online] SAND80-211118. Available from:
doi:10.2172/6548367.

Shen, W.Z., Sørensen, J.N., Zhang, J.H. & W. Z. Shen J. H. Zhang, J.N.S. (2007) Actuator
surface model for wind turbine flow computations. In: European Wind Energy Conference
and Exhibition 2007, EWEC 2007. 2007 pp. 1198–1205.

Sheng, W., Galbraith, R.A.M. & Coton, F.N. (2008) A modified dynamic stall model for low mach
numbers. Journal of Solar Energy Engineering. [Online] 130 (3), 31013. Available from:
doi:10.1115/1.2931509.

Shires, A. (2013) Design optimisation of an offshore vertical axis wind turbine. Energy. [Online]
166 (EN1), 7–18. Available from: doi:10.1680/ener.12.00007.

Sieros, G., Chaviaropoulos, P., Sørensen, J.D., Bulder, B.H., et al. (2012) Upscaling wind
turbines: theoretical and practical aspects and their impact on the cost of energy. Wind
Energy. [Online] 15 (1), 3–17. Available from: doi:10.1002/we.527.

Sorensen, J.N. & Shen, W.Z. (2002) Numerical Modeling of Wind Turbine Wakes. Journal of
Fluid Engineering. [Online] 124 (June 2002), 393–399. Available from:
doi:10.1115/1.1471361.

Staffelbach, G., Senoner, J.M., Gicquel, L. & Poinsot, T. (2008) Large Eddy Simulation of
combustion on massively parallel machines.

Strickland, J.H., Smith, T. & Sun, K. (1981) A vortex model of the darrieus turbine: An analytical
and experimental study.

Sutherland, H.J., Berg, D.E. & Ashwill, T.D. (2012) A Retrospective of VAWT Technology.
Security. [Online] (January), 1–64. Available from: doi:10.2172/1035336.

Tait, C. (2015) Transition by consent: Meeting Britain’s energy needs together.

Tescione, G., Ragni, D., He, C., Simão Ferreira, C.J., et al. (2014) Near wake flow analysis of a
vertical axis wind turbine by stereoscopic particle image velocimetry. Renewable Energy.
[Online] 7047–61. Available from: doi:10.1016/j.renene.2014.02.042.

78
Thesis
MSc Sustainable Energy Futures Imperial College London

Thomas, A.R.B., Khan, P.S.Z. & Gillman, S.B.J. (2016) Current tidal power technologies and
their suitability for applications in coastal and marine areas. Journal of Ocean Engineering
and Marine Energy. [Online] 2 (2), 227–245. Available from: doi:10.1007/s40722-016-0044-
8.

Troldborg, N. & Sørensen, J. (2014) A simple atmospheric boundary layer model applied to
large eddy simulations of wind turbine wakes. Wind Energy. [Online] 17 (April 2013), 657–
669. Available from: doi:10.1002/we.

UNFCCC (1998) Kyoto Protocol to the United Nations Framework Convention on Climate
Change.

Veers, P.S. (1983) A General Method for Fatigue Analysis of Vertical Axis Wind Turbine Blades.
Contract. SAND82-254 (October), 1–4.

Vermeulen, P.E.J., Builtjes, P., Dekker, J. & Lammerts van Bueren, G. (1979) An experimental
study of the wake behind a full scale vertical-axis wind turbine.

Versteeg, H.K. (Henk K. & Malalasekera, W. (Weeratunge) (2007) An introduction to


computational fluid dynamics : the finite volume method. Pearson Education Ltd.

Wilcox, D.C. (1994) Turbulence Modeling for CFD. DCW Industries Inc.

79

You might also like