Vous êtes sur la page 1sur 14

Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Contents lists available at ScienceDirect

Journal of Wind Engineering & Industrial Aerodynamics


journal homepage: www.elsevier.com/locate/jweia

Effects of blade design on ice accretion for horizontal axis wind turbines
G.M. Ibrahim, K. Pope *, Y.S. Muzychka
Department of Mechanical Engineering, Faculty of Engineering and Applied Science, Memorial University of Newfoundland, St. John's, NL, A1B 3X5, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: This paper presents a numerical study to predict ice loads on a wind turbine blade section at 80% of blade span.
Wind power Air flow, droplet impingement and ice accretion simulations are conducted in FENSAP ICE. The effect of low and
Atmospheric icing high liquid water content (LWC) conditions on blade thickness is presented. All NREL airfoil families used for
Numerical predictions HAWTs are examined to study blade design's effect on ice accretion. Ice shapes are numerically predicted at
Aerodynamics different atmospheric temperatures and LWC conditions. The degradation in aerodynamic characteristics due to
ice formation is predicted. The numerical predictions suggest a 10%–65% reduction in lift coefficient due to ice
accretion, which is highly dependent on the accreted ice shape.

1. Introduction the atmospheric temperature range. Additionally, numerous different ice


shapes can be formed, however, shape prediction is difficult due to the
The installed capacity of wind turbines has expanded rapidly in recent rapid changes in the atmospheric conditions during icing events (Fikke
years. However, current major issues for further development of HAWTs, et al., 2006). Glaze ice usually forms transparent, horn shapes of ice and
such as fatigue problems associated with upscaling of turbines (Rezaeiha is typically caused by freezing rain/drizzle or wet icing conditions with
et al., 2017), floating offshore installations (Borg et al., 2014; Paulsen high LWC concentrations Reid et al., 2013; Virk et al., 2012; Battisti,
et al., 2014) and operation in harsh northern locations should be 2015). It is formed at temperature ranges between 0  C and 5  C and
addressed to facilitate continued rapid expansion of installed capacity. droplet diameter ranges from 0 μm to 500 μm. Its density is higher than
Cold regions typically have abundant wind power resources making them rime ice, usually around 900 kg/m3 (Reid et al., 2013; Virk et al., 2012;
attractive locations for wind farm installations (Reid et al., 2013). Cattin, 2013).
However, wind turbines installed in cold climates can experience blade Rime ice is caused by frozen fog with smaller range of droplet sizes
icing causing significant power losses (Hochart et al., 2008), and atmo- compared to glaze ice, ranging from 0 μm to 10 μm (Reid et al., 2013;
spheric icing is a common cause of ice accretion on a wind turbine (Reid Battisti, 2015). Rime ice usually occurs at lower temperatures below
et al., 2013). Ice accretion on wind turbines can have significant impacts 6  C with lower concentrations of LWC, as the temperature is well
on its operation (i.e. aerodynamic performance and controlling systems), below the freezing point water droplets freezes instantaneously upon
safety (i.e. site accessibility due to ice throwing) and economics (i.e., collision with object surface. Rime ice usually forms grainy and opaque
annual energy production and turbine lifespan) (Virk et al., 2012). At- ice shapes (Reid et al., 2013). Depending on droplet size, soft or hard
mospheric icing on wind turbines occurs when super-cooled water rime can be formed. Typically, rime ice density ranges from 100 kg/m3 to
droplets collide with the surface of the blades and freeze upon impact 600 kg/m3 (Reid et al., 2013; Virk et al., 2012; Cattin, 2013). Atmo-
(Reid et al., 2013). Formed ice type is based on the atmospheric and spheric measurements indicated that LWC usually varies from 0 g/m3 to
operational conditions, including temperature, pressure, wind velocity, 5 g/m3. Low LWC concentrations below 1 kg/m3 can be existed during
liquid water content (LWC), median volume diameter of droplets (MVD), stratiform clouds environment. In contrast to cumuliform clouds envi-
rotor blade curvature, roughness, heat flux, icing event time, and ronment which are characterized by higher LWC concentrations usually
collection efficiency of droplets (Battisti, 2015; Kraj and Bibeau, 2010). higher than 1 kg/m3 (Cober et al., 2001; Masters, 1985). LWC value is
Typically, ice accumulates on the turbine blades at temperatures be- highly influenced by elevation and atmospheric temperature. The rela-
tween 0  C and 15  C (Virk et al., 2012; Battisti, 2015). Ice accretion tionship between LWC, MVD and temperature was experimentally eval-
mechanisms are usually classified into two different types: (i) glaze ice uated showing that LWC decreases as the droplet size increases (Jeck,
and (ii) rime ice (Makkonen, 2000), which are typically formed based on 2002). LWC value decreases with altitude and temperature.

* Corresponding author.
E-mail addresses: gmi588@mun.ca (G.M. Ibrahim), kpope@mun.ca (K. Pope), yurim@mun.ca (Y.S. Muzychka).

https://doi.org/10.1016/j.jweia.2017.11.024
Received 28 June 2017; Received in revised form 23 November 2017; Accepted 25 November 2017
Available online 11 December 2017
0167-6105/© 2017 Elsevier Ltd. All rights reserved.
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Nomenclature n Surface normal vector


XOffset Offset distance (m)
ρ Density (kg/m3) TMax Maximum thickness (m)
V Velocity (m/s) XMean Average distance (m)
T Temperature (K)
m Mass flow (kg) Subscripts
d Droplet diameter (μm) a Air
A Surface area (m2) d Droplet
p Static pressure (Pa) f Fluid
μ Dynamic viscosity (kg/m⋅s) evap Evaporation
E Internal energy (J) s Solid
H Enthalpy (J/kg) ∞ Reference
α Local water volume fraction (kg/m3) Rec Recovery
β Collection efficiency h Water film height
C Chord length (m) l Lift
C Coefficient d Drag
F Force (N) U Upper
K Droplet inertia parameter L Lower
τij Shear stress tensor Abbreviations
g Gravity vector (m/s2) LWC Liquid water content
Fr Froude number MVD (Droplet) Median volumetric diameter
Re Reynolds number CFD Computational fluid dynamics
L Latent heat (J/kg) HAWTs Horizontal axis wind turbines
Q Heat flux (W/m2) NREL National Renewable Energy Laboratory
σ Stefan-Boltzmann constant (σ ¼ 5.67  108 W/m2⋅K4) AoA Angle of attack
ε Solid emissivity

The most common solvers applied to investigate icing on wind tur- accumulation at blade sections near the tip and at the leading edge of a
bines are LEWICE (Ruff and Berkowitz, 1990), TURBICE (Marjaniemi blade section (Bose, 1992a; Fu and Farzaneh, 2010).
et al., 2000; Makkonen et al., 2001) and FENSAP ICE (He et al., 2003; Ice accumulation along the blade is highly affected by variation of
Habashi et al., 2001). LEWICE is 2-D an ice accretion code developed by centripetal and gravity forces along blade span sections which can cause
NASA Glenn Research Center. The code applies a time-stepping proced- highly irregular ice shapes (Fu and Farzaneh, 2010; Bose, 1992b). The
ure code to predict ice accretion shape. The flow field calculations are rate of ice accretion increases with water droplet size (Homola et al.,
computed using Douglas Hess-Smith 2-D panel code. Then the obtained 2010) and atmospheric temperature significantly influences ice shape.
solution is used to calculate the trajectories of particles and the Streamlined ice shapes are formed at lower temperatures and horn ice is
impingement points on the body. The icing model is then used to predict formed at higher temperatures (Homola et al., 2010). Glaze ice condi-
the ice growth rate and shape (Ruff and Berkowitz, 1990). Turbine Blade tions can produce higher power loss compared to rime ice conditions, due
Icing Model (TURBICE) is a two-dimensional ice accretion simulation to the greater ice thickness and shape of glaze ice (Reid et al., 2013). An
program. It was developed to predict icing on wind turbine blades. investigation of ice accretion on four different wind turbines sizes indi-
TURBICE uses panel methods to calculate potential flow field around the cated that rime ice conditions are less severe for large wind turbines
blade. The program uses Lagrangian technique for droplet trajectories compared in terms of ice mass and thickness (Virk et al., 2010b; Homola
and impingements calculations. The program is capable of calculating et al., 2010). The most severe icing problems are typically from glaze ice
collision efficiencies and locating the droplet impingement locations conditions during freezing rain and drizzle (Bose, 1992a), which can
along the blade surface. The program can estimate the amount of energy produce higher losses in lift coefficient than rime ice conditions (Hudecz
required for heating to prevent ice accretion on blade surface (Ruff and et al., 2013).
Berkowitz, 1990; Makkonen et al., 2001). FENSAP ICE is a 3-D premier Limited data is available on airfoil profile's effect on ice accretion, in
in-flight icing simulation solver. The solver simulates flow field, droplet terms of ice quantity and shape, as well as the effect on aerodynamic
impingements, ice shapes and predicting anti/de-icing heat loads using a performance. To further examine ice accretion impacts on wind turbines,
built in CFD modules. It uses 3D Navier-Stokes and energy equations for numerical studies are performed to determine the connection between
flow field calculations and 3-D Eulerian model for droplet calculations icing parameters and blade designing parameters. In this paper, ice ac-
(He et al., 2003; Habashi et al., 2001). cretion on wind turbines is investigated in terms of the following:
Ice accretion on wind turbines can greatly degrade the aerodynamic
performance and reduce the power production by up to 60% at lower  Airfoil type of the NREL airfoil families of S801 to S835 (inclusive)
wind speeds (Reid et al., 2013). In a study on icing of large wind turbines,  Blade thickness of 50%–150% of airfoil blade for NREL S809
the results indicated that ice growth can be greatly influenced by com-  The effect of ice accretion on aerodynamic performance (lift and
bined changes in blade design (i.e., size and profile) and relative velocity drag) for different icing conditions
at each blade section (Virk et al., 2012).
Furthermore, ice accretion was observed to be more at blade sections Studying blade design parameters such as blade thickness is useful to
near the tip compared to root section, where both blade thickness and investigate ice accretion physics at different locations along the blade
chord length are minimized (Virk et al., 2012). A study by Homola (Virk span, as typical turbine designs often have varying airfoil curvature/
et al., 2010a) indicated that ice growth in terms of ice rate and accretion thickness from the blade's root to tip. Lastly, the effect of different ice
shape is highly influenced by angle of attack of fluid flow. A study of icing shapes, caused by differing icing conditions, on aerodynamic perfor-
on small wind turbines identified that blade rotation can lead to large ice mance provides important insights for wind turbine design and

40
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Fig. 1. Solution scheme for (a) single-shot approach and (b) multi-shot approach.

 
placement strategies. The results of this paper provide valuable new in- 2
σ ij ¼ δij pa þ μa δjk ∇k vi þ δik ∇k vj  δij ∇k vk ¼ δij pa þ τij (2)
sights and new numerical predictions to design wind turbines that can 3
operate in harsh northern locations.
where pa and μa represent air static pressure and dynamic viscosity of air,
2. Formulation of multiphase flow and phase change

Typically, the numerical simulation of ice accretion mainly consists of


the following modules: air flow solution, droplets collection efficiency
calculations, evaluation of boundary layer characteristics, and ice mass
evaluation by thermodynamic model computation (Reid et al., 2013).

2.1. Flow field calculations

The CFD simulations are steady incompressible flow. The air flow
solution is solved by FENSAP solver with the 3D Navier-Stokes and en-
ergy equations (ANSYS, 2015). The Navier-Stokes equation is
expressed as

∂ρa Va ⇀
 ⇀ ⇀  ⇀
þ ∇ : ρa Va Va ¼ ∇ :σ ij þ ρa :g (1)
∂t

where ρa is the air density and Va is the velocity vector. The stress tensor,
σ ij , is represented by

Fig. 2. One of NREL airfoil families, S809 profile.

41
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Fig. 3. Computational domain discretizations for NREL S809, (a) O grid domain elevation view, (b) cell distribution around the blade profile, and (c) 3-D span wise view.

respectively. The conservation of energy equation is adds a single additional variable for an undamped kinematic eddy vis-
    cosity. It is designed for a low Reynolds number and can compute the
∂ρa Ea ⇀  ⇀  ⇀ ⇀ ⇀ ⇀ entire domain flow field to the airfoil surface. The model was originally
þ ∇ : ρa Va Ha ¼ ∇: Ka ∇ Ta þ vi τij þ ρa g: Va (3)
∂t created for aerodynamic applications and is beneficial for its moderate
resolution requirements, stability and convergence ability. The SIMPLE
where E and H are the total internal energy and enthalpy, respectively. algorithm is used for the pressure velocity coupling. The heat fluxes at the
The airflow around a blade is typically turbulent at high Reynolds blade wall are computed with second order accuracy by solving the en-
numbers, thus one-equation Spalart Allmaras turbulence model (Hudecz ergy equation. Lift and drag coefficients can be calculated from
et al., 2013), with low free stream turbulence (using a very low Eddy/-
laminar viscosity ratio of 1  105) is selected (ANSYS, 2015; Spalart and 2Fl
Cl ¼ (4)
Allmaras, 1992). The Spalart-Allmaras model was selected as a ρa;∞ Va;∞
2 A

compromise between acceptable computational cost and the required
accuracy in turbulent flow simulations (Virk et al., 2010a). The model

42
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

2Fd fluid. The variables T∞ ; V∞ ; LWC and β represent the airflow and droplets
Cd ¼ (5)
ρa;∞ Va;∞
2 A
∞ reference parameters. The collection efficiency is calculated by
⇀ ⇀
where Fl and Fd are lift and drag forces, respectively, and A∞ is the αVd :n
β¼ (11)
reference surface area (based on C, chord length and blade span-wise ðLWCÞV∞
length). The torque coefficient, Cy, is determined by

where n is surface normal vector. The flow and droplet solutions are
Cy ¼ Cl sin ϕ  Cd cos ϕ (6) input to the ICE 3D solver to predict ice accretion (ANSYS, 2015).
In this paper, single- and multi-shot solutions methodologies are
where ϕ is the angle between the relative wind direction and plane of employed. Single shot is selected for the study of ice and the airfoil type
blade rotation. because of its lower computational expense compared to multi-shot
computations. A single-shot methodology is useful for capturing the
2.2. Droplet calculations differences of ice mass with airfoil type, but less accurate at predicting
the localized ice accretion quantity for specific icing conditions. A multi-
Water concentration, droplet velocity vectors, collection efficiency, shot approach provides more accurate predictions for the localized ice
impingement, shadow zone properties and impingement limits over the accumulation as it accounts for the changes in droplet collection distri-
entire domain are provided by the DROP 3D solver (ANSYS, 2015). The bution that is caused by accumulated ice changing thing the aero-
general Eulerian two-fluid model is used, which consists of Navier-Stokes dynamics of the airfoil. In this paper, a multi-shot solution methodology
equations augmented by the droplets’ continuity and mo- is used for investigating the aerodynamic effects of ice accumulation.
mentum equations, For single shot calculations, flow field and droplet solutions are
∂α ⇀  ⇀ 
computed once and icing solution is obtained for total icing time as one
þ ∇ : αVd ¼ 0 (7) interval (Fig. 1a). Flow field solutions are used as an input to the droplet
∂t
solver to compute droplet collision efficiencies and impingement pro-
 ⇀  files. Then droplet solutions are transferred manually to ICE 3D solver to
∂ αVd  ⇀  C R ⇀    predict ice mass/shape for one interval of an icing time. For an accurate
⇀ ⇀ D ed ⇀ ρ 1
þ ∇ αVd  Vd ¼ α Va  Vd þ α 1  a prediction of ice shape, multi-shot approach is recommended. As for
∂t 24K ρd Fr2
multi shot calculations (Fig. 1b), the transient changes in geometry,
(8)
surface heat fluxes and the droplet impingement profile due to ice for-
mation are taken into account. The total icing time is divided into short
where α and Vd are water volume fraction and droplet velocity, respec-
time intervals and an icing solution is obtained for each interval (i.e. for
tively. The first term in the momentum equation represents the drag on a
total icing time of 1800 s the time interval for 3 shots equals 600 s)
droplet with MVD. The second term represents the buoyancy and gravity
(ANSYS, 2015). Number of shots means the number of divisions of a total
forces. Collection efficiency is a dimensionless parameter that describes
icing time and each time interval is an iteration. Flow field, droplet and
the effective collision forces of droplets on a blade surface. Its peak value
is typically near the flow stagnation point. The droplet distribution is set
as mono-disperse (uniform droplet diameters), which has provided ac- Table 1
curate results in previous papers (Virk et al., 2010a). Additionally, a Test conditions for validation of numerical predictions.
comparison of mono-dispersed and Langmuir-D distribution (which Parameter Quantity
contains 7 different droplet sizes) (ANSYS, 2015), indicated that the Chord length 0.53 m
composite solution for the distribution is fairly close to solution of the Air speed 67.1 m/s
mono disperse model. To represent the icing event, the median volu- LWC 1.0 g/m3
metric diameter (MVD) of the droplets is chosen as a single quantity. MVD 20 μm
AoA 4
Temperature 7.78  C
2.3. Icing calculations Icing time 6 min

The ICE 3D solver uses mass conservation and energy conservation


equations to solve complex thermodynamics of ice formation (ANSYS,
2015). The equation expressing mass conservation is
 
∂hf ⇀ 
ρf þ ∇ : V f hf ¼ V∞ LWCβ  m_ evap  m_ ice (9)
∂t

where f and β represent water film and collection efficiency, respectively.


The first term represents the mass of water impingement, the second term
is mass of evaporated water and the third term is ice mass. The conser-
vation of energy equation is
2 3 2

2 3

e
6∂hf cf Tf ⇀  7 

Vd

ρf 4 þ ∇ : V f hf cf T~ f 5 ¼ 4cf T~ ∞  T~ f þ 5V∞ LWCβ


∂t 2
   
Levap m_ evap þ Lfusion  cs T~ m_ ice þ σε T∞ 4  Tf 4  ch T~ f  T~ ice;rec
þ Qantiicing
(10)
Fig. 4. Ice accretion on NACA 0 012 obtained numerically and experimentally
(ANSYS, 2015).
where ρf ; cf ; cs ; ε; σ; Levap and Lfusion represent the physical properties of

43
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Fig. 5. Cross-section of NREL airfoil profiles investigated in this paper.

Table 2 Table 3
Test conditions to examine curvature impact on ice mass. Tested ranges of airfoil curvature parameters.

Parameter Quantity Curvature parameter Examined range (m)

NREL airfoil families From S801 to S835 (S824 data is not available) TMax 0.11496–0.26868
Wind speed (m/s) 75 XOffset 0.07983–0.2454
AoA ( ) 0 XMean 0.251755–0.5107445
Icing temperature ( C) 4 XL 0.168058–0.546335
MVD value (μm) 30 XU 0.24461–0.508787
LWC value (g/m3) 0.5
Blade thickness 100% (normal value corresponds to 1 m chord length)
Icing time (minutes) 30

Fig. 6. Airfoil curvature characterizations. Fig. 7. The effect of airfoil geometric parameters on airfoil icing for a NREL airfoil.

ice calculations are computed for the first interval of time to predict first
ice layer formation on blade. A new computational grid is generated for each shot (ANSYS, 2015). In this paper, other solution methodologies are
the formed geometry due to ice accretion. The generated grid is then used maintained for single- and multi-shot simulations (e.g., turbulence
as an input to predict the second ice layer and the process is continued for model, conservation equations, among others). For both single and
multi-shot approaches, a convergence criteria of 1010 for fluid flow

44
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

calculations and 1010 for the energy equations is maintained. For all is discretized into one cell of 0.25 m and along the chord there are 360
multi-shot simulations, full convergence is achieved for each shot. discretizations, in a mapped geometric progression scheme. Ice accretion
is mostly formed at the leading edge due to droplet collection efficiency.
2.4. Mesh discretizations Smaller elements are created at the leading edge with equal spacing of
0.001 m to accurately evaluate both ice mass and shape. The cell size
The NREL airfoil families are selected to investigate icing as they are gradually changes along the airfoil surface, as illustrated in Fig. 3b and c.
widely used for HAWTs blade industry (Tangier and Somers, 1995). For The grid independence study investigated different mesh resolutions,
all airfoils, a chord length of 1 m is maintained along a blade span of approximately 25,000, 35,000 and 45,000 elements, at a specific icing
0.25 m to represent a small portion of a rotor blade near the tip. A span of condition resulting in minimal changes in ice mass (approximately ± 4 g).
0.25 was shown in a previous study to provide useful results (Virk et al., Lift and drag values are obtained after convergence with minimal dif-
2010a). A fine 3D structured grid is created for each airfoil with identical ferences of ± 0.0001 in their quantities. Convergence is achieved for all
discretization schemes. As for a NREL S809 profile (Fig. 2), a fine 3D variables (i.e., momentum, mass, energy and turbulence model equations
structured grid with 35,280 cells is created with a minimum orthogonal and droplet solution residuals).
quality of 0.835 (where values close to 0 correspond to low quality and 1
represents high quality) and maximum orthogonal skew of 0.194 (where 3. Results validation
values close to 1 correspond to low quality and 0 represents high quality).
The yþ value is 0.2. The grid's first boundary layer thickness over the A study (Shin and Bond, 1992) on icing of NACA 0 012 airfoils using
blade is 1  106 m with a growth rate of 1.1 for 98 layers to accurately LEWICE code and wind tunnel tests is used to validate this work in terms
capture ice shape and mass on blade surface. The main advantages of of ice shape. A computational grid is created around NACA 0 012 with
using a structured grid are higher resolution and better accuracy. The the same process and configurations used for NREL S809. As presented in
structured domain uses hexahedra cells which are characterized by reg- Table 1, one condition from the experiment (Shin and Bond, 1992) is
ular connectivity and efficient spacing. As shown in Fig. 3a, a circular investigated using FENSAP ICE solver, to obtain an ice shape comparison.
domain (O grid type) with a diameter of 20 m is created. The blade span As illustrated in Fig. 4, the predicted ice shapes agree well with

Fig. 10. Total droplet collection with NREL blades of various thicknesses.
Fig. 8. Ice mass variations with TMax/XMean.

Fig. 9. Droplets impingement profile (droplet collection) for 3 different blade sizes. Fig. 11. Numerically predicted effect of blade thickness and droplet collection on accreted
ice mass on a HAWT blade.

45
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

previous numerical and experimental data. The total icing time is divided results in this paper agree well with the experimental results by Broeren
in 5 intervals (72 s) to evaluate an accurate ice shape for comparison. For and Bragg Broeren et al. (2004); Horak et al.(2008), with a reduction in
each interval of time, one ice shot/layer is added. Then the grid lift ranging from 10% to 65%.
displacement is obtained and transferred to start the calculations for the
second shot taking into account transient changes happening in flow 4. Results and discussion
field, droplet impingement profile. For 5 ice shots, the ice horn thickness
(approximately x/c ¼ 0.025) at the leading edge is nearly equal to the In this section, numerical predictions of ice accretion on HAWTs with
one predicted by experimental investigation (Shin and Bond, 1992). NREL airfoils and the effect on aerodynamic performance are presented.
Minimal changes (especially after the third shot) in ice shapes are pre- The analysis is in terms of (i) the quantity of ice accretion for 35 different
dicted. The more shots the accurate ice shapes and high computational NREL airfoils (S801 to S835, inclusive), (ii) the quantity of ice accretion
time to obtain simulated ice shapes. Therefore, based on nearly equal ice for a NREL airfoil S809 for different blade thicknesses, and (iii) the effect
shapes when comparing FENSAP ICE data with experimental data, ice of ice accretion on a NREL airfoil S809 at different angles of attack.
mass is nearly identical and 3 shots is suitable for this paper. Significant parameters that are important for this study include wind
For aerodynamic efficiency prediction due to ice accumulation, the speed/relative wind velocity, air temperature, LWC, MVD, blade curva-
experimental results analysed by Broeren and Bragg (Spalart and All- ture parameters, angle of attack (AOA), blade surface roughness and
maras, 1992; Tangier and Somers, 1995) at a Reynolds number of icing time, among others. In the numerical predictions, the following
6.5  106 m showed that the reduction in the aerodynamic efficiency operating conditions are investigated.
can range from 15% to 80% due to ice accretion. The current predicted
 Relative wind velocity - a relative wind speed at a span wise location
80% from the rotor blade root is represented, as 80% from the rotor
blade root is typical location of high ice accretion (Virk et al., 2010a).
Therefore, assuming a TSR of 6 and a wind speed of 15.3 m/s, a
relative wind speed of 75 m/s is calculated and maintained for all
simulations in this paper. A relative wind speed of 75 m/s can also be
calculated from a TSR of 7 and a wind speed of 13.2 m/s.
 Air temperature and blade surface temperature are assumed uniform
(i.e., air temperature ¼ blade surface temperature). Two tempera-
tures are selected, 4  C and 8  C, to represent temperatures that
can cause glaze and rime icing conditions, respectively.
 Current wind turbine towers can reach up to 135 m, and blade tip
height can reach 200 m above ground. Atmospheric icing conditions
mainly depend on the surrounding environment with lowest cloud
extent at 480 m above sea level for cumuliform clouds and more than
1 km for stratiform clouds. As LWC changes with altitude, the
assumption of no significant change in values of LWC and MVD is
maintained for atmospheric icing of wind turbines (Fikke et al.,
2006). Since Meteorological data at lower altitudes (wind towers) are
not available, different icing conditions with higher and lower LWC
Fig. 12. Ice loads on NREL airfoil families for HAWTs.

Table 4
Test conditions to examine blade thickness effect at various icing conditions.

46
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Fig. 13. Relationship between LWC and MVD for the examined conditions of atmospheric
icing on wind turbines.

concentrations can be numerically predicted using cloud


characteristics.
 The LWC and MVD values investigated in blade section effect are
selected from measured values of LWC variation with MVD at
different temperatures, which is in the DROP 3D solver. The selection
is based on different cloud conditions (stratiform cloud for low con-
centrations and cumuliform for higher ones) (Jeck, 2002), to provide
conditions of LWC that exist at specific MVD. For lower and higher
LWC concentrations, stratiform and cumuliform clouds conditions at
lowest altitude are applied to investigate different icing events. This
method captures the impact of different icing conditions, in terms of
high and low LWC, at specific MVD ranges, for a specified
temperature.
 Icing time - an icing time of 30 min is simulated, which is a critical
time for determining whether to stop turbine operation. Fig. 14. Predicted ice accretion on different NREL S809 airfoil thicknesses for (a) low
 Blade roughness - the sand-grain roughness model is used and LWC conditions and (b) high LWC conditions.
maintained at 0.0005 m, which has been shown to be an accurate
representation for typical icing predictions based on experimental
validation (ANSYS, 2015). Surface roughness is a significant factor in Table 5
predicting/calculating heat fluxes and shear stresses for icing simu- Test conditions to examine blade aerodynamics during icing conditions.

lations, although not the focus of this paper. Parameter Quantity


 Blade curvature - parameters are investigated in terms of the airfoil Airfoil designated NREL S809
maximum thickness and its location. name
 Airfoil thickness (Section II of the Results and Discussion) - the NREL AOA ( ) 0
S809 airfoil is selected to study the effect of blade thickness variation Wind speed (m/s) 75
Icing temperature 4, 8
on ice mass during different icing events characterized by high and ( C)
low LWC conditions. MVD value (μm) 30
 To examine the airfoil design or curvature impact on ice mass all LWC value (g/m3) 0.5, 1
NREL families are used. As ice accumulates on blade surface airfoil Blade thickness 0.21 m (Normal thickness corresponds to a chord length of
1 m)
curvature changes noticeably causing a significant impact of aero-
Icing time (minutes) 30
dynamics, therefore various ice shapes at different icing conditions
are used to study this impact.
YL, XOffset, TMax, and XMean. The variable XOffset is the offset distance be-
5. The effect of airfoil shape on ice accumulation quantity for a tween the points (XU,YU) and (XL,YL), TMax is blade maximum thickness
HAWT with an NREL airfoil and XMean is the average offset distance from the origin (TMax location on
X axis). Blade curvature/leading edge radius is highly affected by the
In this section, ice accumulation is investigated for the NREL airfoil location of these two points (XU, YU) and (XL, YL) for both upper and
families. As shown in Fig. 5, all NREL airfoil families are analysed to lower sides, respectively. Table 3, shows the examined ranges of curva-
study the relationship between their curvature's significant parameters ture parameters for NREL airfoil families analyzed in this paper.
and ice accretion. As presented in Table 2, one icing condition is main- The combined effects of the airfoil's geometry on the flow stream and
tained to investigate the effect of blade design on ice accumulation. ice accretion are difficult to distinguish. The geometric variables that can
As illustrated in Fig. 6, airfoil curvature can be defined by XU, YU, XL, be used to identify an airfoil's profile were systematically analyzed based

47
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Fig. 15. Numerically predicted ice shapes for different icing conditions at (a) LWC ¼ 0.5 g/m3, T ¼ 4  C, (b) LWC ¼ 1 g/m3, T ¼ 4  C, (c) LWC ¼ 0.5 g/m3, T ¼ 8  C and (d)
LWC ¼ 1 g/m3, T ¼ 8  C.

on the numerical predictions of ice accumulation. As illustrated in Fig. 7, curve profile is not identical for the pressure side and suction side of a
the geometric variable TMax and YMean have the most significant effect on blade section because of asymmetry of the airfoils. Higher collection
ice accumulation. The vertical axis is the ice accumulation and the hor- efficiency peak magnitude can occur for thinner blade sections, thus ice
izontal axis represents a normalized geometric variable, whereby the accretion is more affected by the total collection efficiency rather than its
values are divided by their highest quantity. The trend line's slope for peak value.
TMax, YMean, and YOffset are 0.56, 0.44, and 4.3, respectively, with a co- As illustrated in Fig. 10, total droplet collection has a linear trend
efficient of determination (r2) of 0.79, 0.82, and 0.74 (representing (R2 ¼ 0.7405) with NREL blade thickness. The total droplet collection is
medium to high correlation), respectively. The geometric variables XMean defined as the integral of collection efficiency (i.e., total area under the
and XOffset show only a minor correlation with ice accumulation with collection efficiency curve, Fig. 9), which represents the quantity of
trend line slope's of 0.36 and 0.07, respectively and coefficient of de- droplets impacting the wind turbine blade. The integral values are ob-
terminations of 0.17 and 0.03 respectively (representing low tained for all NREL airfoils and suggest that blade thickness has a sig-
correlation). nificant effect on the quantity of droplets colliding with the blade surface.
As illustrated in Fig. 8, the location of maximum thickness has a Furthermore, as illustrated in Fig. 11, total droplet collection is nearly
significant impact on ice load and can be represented by plotting ice mass perfectly correlated (R2 ¼ 0.9997) with accreted ice mass, thus accreted
with TMax/XMean. Higher values of TMax/XMean ratio suggests that the ice mass on a wind turbine blade can be predicted based on the airfoil's
location of maximum thickness is near the leading edge causing greater thickness, with moderate agreement (R2 ¼ 0.74).
ice accumulation. Locating XOffset (distance between the two points As illustrated in Fig. 12, ice load significantly increases with airfoil
(XU,YU) and (XL,YL)) near the blade's leading edge can lead to a large thickness as thicker airfoils (e.g., S804, S808, S815, S821) tend to have
stagnation area and increased ice accumulation. However, locating XOffset greater ice accumulation compared to thinner airfoils (e.g., S802, S806,
near the blade's trailing edge can reduce the stagnation area and decrease S817, S832). The differences in ice quantities between thicker
the quantity of accumulated ice. (S821 ¼ 0.8) and thinner (S806 ¼ 0.43) airfoils is significant. Thick
In Fig. 9, droplets collection efficiencies are illustrated for three blades tend to have more droplet collection area at the leading edge, thus
different airfoils to show how the impingement area is affected due to the droplet impingement area is large and can cause greater ice accu-
blade thickness. For thinner airfoils (e.g., S806), the collection efficiency mulation. Differences between the upper and lower points on the X-axis
profile shows the lowest y/c (0.02 to 0.04), while thicker airfoils (e.g., (XOffset) or maximum thickness location, can lead to various stagnation
S821) have the largest y/c (0.1 to 0.06), which illustrates a contributing area configurations, which is the likely cause of airfoils with lower
factor to increased ice accumulation for thicker airfoils. The peak value of thickness having slightly equal or large ice quantity (i.e. comparing S802
collection efficiency can be found near the point of flow stagnation. The with S803, S808 with S811 and S821 with S823). For example, S823

48
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Fig. 16. Computational grids around airofils with accreted ice for different icing conditions: (a) LWC ¼ 0.5 g/m3, T ¼ 4  C, (b) LWC ¼ 1 g/m3, T ¼ 4  C, (c) LWC ¼ 0.5 g/m3, T ¼ 8  C
and (d) LWC ¼ 1 g/m3, T ¼ 8  C.

(0.786) is having nearly equal ice mass with S821 (0.785), however S821 6. The effect of airfoil thickness on ice accumulation quantity for
have comparatively higher thickness, this can highlight the significance a HAWT with an NREL S809 airfoil
of other curvature important parameters like TMax/XMean ratio or
maximum thickness location. This section presents predicted ice accumulation on five different
thicknesses of the same airfoil type, NREL S809, to investigate the

49
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

significance of blade thickness on ice load. For lower MVD, below 20 μm,
differences in ice mass do not change to the same extent as larger ones,
highlighting the significance of MVD during ice accumulation. For larger
droplet sizes, ice accretion variation with blade thickness is significant.
Icing environments with small droplets at high and low LWC concen-
trations has a minor impact on ice amount for different blade thicknesses
compared to large droplets. LWC content have a significant impact on ice
amount as it can be shown when comparing Fig. 14a with 14c, that ice
mass values is increased at fixed droplet size comparison.

7. The effect of ice accretion on aerodynamic performance for a


HAWT with an NREL S809 airfoil

Aerodynamic characteristics significantly depend on blade curvature.


During an ice accretion process, the blade profile changes with time,
changing both lift and drag characteristic. As shown in Table 5, different
icing conditions are examined to numerically obtain various ice shapes.
This comparison between clean and iced profiles is used to predict the
changes in aerodynamic performance.
As ice accretes on the blade surface, aerodynamics change and
considerably affect flow field, heat fluxes and droplet impingement dy-
namics. Multi-shot approach in FENSAP ICE solver accounts for the
transient changes occurring during an ice formation process. Thus, multi-
shot approach is used to accurately predict ice shapes for this analysis.
As illustrated in Fig. 15, ice shapes at 4  C and 8  C temperatures
with two LWC conditions are numerically predicted using a multi-shot
approach. Generally, icing shape and location are controlled by flow
speed and angle of attack. Ice thickness at the stagnation point of the flow
is highly influenced by the atmospheric temperature, as ice thickness at
this point is increased when comparing ice shapes at 4  C and 8  C,
under fixed LWC and MVD conditions.
At 4  C, LWC significantly affects the quantity of ice accumulation,
as well as its propagation along the blade section. At 8  C, a horn ice
shape is formed and its height is affected by LWC, as a significantly more
pronounced ice horn accumulates at a LWC of 1 g/m3 compared to 0.5 g/
m3. As illustrated in Fig. 16, a computational grid around an iced NREL
S809 blade is created by the grid displacement tool existed in the solver
for each ice shape to numerically predict the lift and drag coefficients.
The effect of the four ice shapes (Fig. 15) on aerodynamic perfor-
mance, is investigated by analysing the lift and drag at several angles of
Fig. 17. Aerodynamic performance for predicted ice shapes, in terms of (a) lift coefficient
and (b) drag coefficient.
attack (i.e., the ice shapes are formed at an AoA of 0 , and then the AoA is
changed to study the effect of ice on lift and drag). As illustrated in
Fig. 17, ice accretion can produce degradation of the lift coefficient, as
relationship between blade thickness and icing conditions in terms of
well as increase drag. Every ice distribution can cause different flow
LWC and MVD at specific temperature. As presented in Table 4, blade
characterises around the blade, including location of flow stagnation
thickness is varied from 50% to 150% (where 100% value represents the
point (affected by angle of attack) and flow separation behaviour
original equivalent blade thickness for a chord length of 1 m), thus 50%
(affected by height/thickness of the ice horn). A significant reduction in
thickness represents a thin blade and 150% thickness represents a
lift, ranging from 10% to 65%, is predicted. Furthermore, the increase in
thick blade.
drag and decrease in lift is substantial for large horn ice. As wind turbine
Flow field solutions are computed by FENSAP solver, and then droplet
blade lift decreases the overall torque will decrease leading to a reduction
calculations for each icing conditions (LWC and MVD) are obtained using
in output power.
DROP 3D solver. Finally, icing calculations are predicted for each con-
In Fig. 18, flow fields (in terms of velocity magnitude) are presented
dition using a single shot approach. Icing conditions with low and high
for the clean blade and four iced blades, at an AoA of 8 . The larger ice
LWC conditions are investigated. Typically, smaller droplet sizes exist at
shape formation causes more flow separation (e.g., comparing ice shape
higher LWC conditions while large droplets are at lower LWC conditions
(d) with a clean blade). As flow behaviour around the blade changes due
(Jeck, 2002).
to the formed ice, the overall lift force from the blade section is signifi-
As illustrated in Fig. 13, eight different LWC/MVD ratios are selected
cantly reduced. The flow fields show the drastic changes the accumulated
to represent various atmospheric icing conditions. The eight different
ice has on the airfoil aerodynamics, which can significantly impair
LWC quantities, at a temperature of 4  C, from 0.17 g/m3 to 3.7 g/m3
HAWT performance.
(selected as discussed above from appendix C) are investigated, which
To determine the effect of reduced lift and increased drag due to ice
represent icing events with low and high LWC conditions. Eight different
accumulation on the performance of a HAWT, the torque coefficient is
MVD values, from 15 μm to 50 μm, are investigated at a temperature of
determined for various AoA for the clean and the four iced blades. To
4  C, which is useful to study the impact of droplet size during ice
determine the angle of relative wind, a 40 m blade radius is assumed,
formation on blade profile.
therefore, at 80% blade radius (32 m), the torque coefficient is deter-
As illustrated in Fig. 14, for larger blade thicknesses higher quantities
mined. A TSR of 6 and a wind speed of 15.3 m/s is also assumed. As
of ice are formed for differing icing conditions, which highlights the
illustrated in Fig. 19, the different ice shapes have severe impacts on

50
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

Fig. 18. Predicted velocity magnitude contours for clean and iced NREL S809 airfoils on a HAWT.

HAWT performance. The optimal relative angle of attack for best aero- 8. Conclusions
dynamic performance is at a different location, depending on the
ice shape. In this paper, ice accretion on a wind turbine blade was investigated
using FENSAP ICE solver. The predicted results identified that droplet
size has a significant effect on ice accumulation for fixed temperature and
LWC conditions. The results of this paper provide the following
key insights.

 From predictions on ice loads on NREL airfoil families, it was shown


that blade thickness and its location from the leading edge has a
greater impact on ice quantity than any other blade curvature
parameter.
 Thicker blades have a collection efficiency characterized by large
coverage area (y/c) at the leading edge when compared to thinner
ones, thus ice loads increase significantly with blade thickness.
 Accreted quantity of ice is dependent on the blade thickness due to
the increased droplet collection of thicker blades
 Ice mainly formed/found at the blade's leading edge and its location is
controlled by angle of attack.
 Atmospheric icing temperature affects ice thickness at the location of
flow stagnation, as ice thickness at this location is increased signifi-
cantly when comparing ice shapes at 4  C and 8  C, under same
icing conditions.
 Propagation of the accreted ice along blade section is expected for an
icing temperature of 4  C, with higher LWC conditions.
Fig. 19. The effect of ice accumulation on aerodynamic performance of a HAWT.

51
G.M. Ibrahim et al. Journal of Wind Engineering & Industrial Aerodynamics 173 (2018) 39–52

 At a temperature of 8  C or below, horn ice shapes are expected, its Hochart, C., Fortin, G., Perron, J., Ilinca, A., 2008. Wind turbine performance under icing
conditions. Wind Energy 11 (4), 319–333.
thickness and height are controlled by LWC concentration. Aero-
Homola, M.C., Virk, M.S., Wallenius, T., Nicklasson, P.J., Sundsbø, P.A., 2010. Effect of
dynamic behaviour prediction mainly depends on accreted ice shape atmospheric temperature and droplet size variation on ice accretion of wind turbine
on the blade's surface and ice accretion significantly increased drag blades. J. Wind Eng. Industrial Aerodynamics 98 (12), 724–729.
coefficient. Horak, V., Rozehnal, D., Chara, Z., Hyll, A., 2008. CFD and Experimental Study of
Aerodynamic Degradation of Iced Airfoils.
 The loss in lift coefficient due to ice accretion was between 10% and Hudecz, A., Koss, H., Hansen, M.O.L., 2013. Ice accretion on wind turbine blades. In: 15th
65%, depending on accreted ice shape. International Workshop on atmospheric icing of structures (IWAIS XV).
Jeck, R.K., 2002. Icing Design Envelopes (14 CFR Parts 25 and 29, Appendix C) Converted
to a Distance Based Format (No. DOT/FAA/AR-00/30). Federal Aviation
Acknowledgments Administration Technical Center, Atlantic City, NJ.
Kraj, A.G., Bibeau, E.L., 2010. Phases of icing on wind turbine blades characterized by ice
The authors of this paper gratefully acknowledge the financial sup- accumulation. Renew. Energy 35 (5), 966–972.
Makkonen, L., 2000. Models for the growth of rime, glaze, icicles and wet snow on
port of the Research and Development Corporation of Newfoundland structures, Philosophical Transactions of the Royal Society of London A:
and Labrador. Mathematical. Phys. Eng. Sci. 358 (1776), 2913–2939.
Makkonen, L., Laakso, T., Marjaniemi, M., Finstad, K., 2001. Modelling and prevention of
ice accretion on wind turbines. Wind Eng. 25 (1), 3–21.
References Marjaniemi, M., Makkonen, L., Laakso, T., 2000. TURBICE-the wind turbine blade icing
model. In: In Proceedings of the BOREAS V conference.
ANSYS, 2015. FENSAP-ICE User Manual R1.0, Canada. Masters, C.O., 1985. A New Characterization of Super Cooled Clouds below 10,000 Feet
Battisti, L., 2015. Wind Turbines in Cold Climates: Icing Impacts and Mitigation Systems. AGL.
Springer. Paulsen, U.S., Madsen, H.A., Kragh, K.A., Nielsen, P.H., Baran, I., Hattel, J.,
Borg, M., Shires, A., Collu, M., 2014. Offshore floating vertical axis wind turbines, Berthelsen, P.A., 2014. DeepWind-from idea to 5 MW concept. Energy Procedia 53,
dynamics modelling state of the art. Part I: aerodynamics. Renew. Sustain. Energy 23–33.
Rev. 39, 1214–1225. Reid, T., Baruzzi, G., Ozcer, I., Switchenko, D., Habashi, W., January, 2013. FENSAP-ICE
Bose, N., 1992. Icing on a small horizontal-axis wind turbine - Part 1: glaze ice profiles. simulation of icing on wind turbine blades, part 1: performance degradation. In: 51st
J. Wind Eng. Industrial Aerodynamics 45 (1), 75–85. AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace
Bose, N., 1992. Icing on a small horizontal-axis wind turbine - Part 2: three dimensional Exposition, Grapevine, Texas.
ice and wet snow formations. J. Wind Eng. Industrial Aerodynamics 45 (1), 87–96. Rezaeiha, A., Pereira, R., Kotsonis, M., 2017. Fluctuations of angle of attack and lift
Broeren, A.P., Bragg, M.B., Addy, H.E., 2004. Effect of intercycle ice accretions on airfoil coefficient and the resultant fatigue loads for a large horizontal axis wind turbine.
performance. J. Aircr. 41 (1), 165–174. Renew. Energy 114 (B), 904–916, 2017.
Cattin, R., 2013. Icing of Wind Turbines, Vindforsk Projects, a Survey of the Development Ruff, G.A., Berkowitz, B.M., 1990. User's Manual for the NASA Lewis Ice Accretion
and Research Needs, Elforsk Report 13. Prediction Code (LEWICE).
Cober, S.G., Isaac, G.A., Strapp, J.W., 2001. Characterizations of aircraft icing Shin, J., Bond, T.H., 1992. Experimental and Computational Ice Shapes and Resulting
environments that include supercooled large drops. J. Appl. Meteorology 40 (11), Drag Increase for a NACA 0012 Airfoil.
1984–2002. Spalart, P., Allmaras, S., January, 1992. A One-equation Turbulence Model for
Fikke, S.M., Ronsten, G., Heimo, A., Kunz, S., Ostrozlik, M., Persson, P.E., Laakso, T., Aerodynamic Flows. 30th Aerospace Sciences Meeting and Exhibit.
2006. Atmospheric Icing on Structures: Measurements and Data Collection on Icing: Tangier, J.L., Somers, D.M., 1995. NREL Airfoil Families for HAWTs National Renewable
State of the Art, Meteo Schweiz. Energy Laboratory, pp. 117–123.
Fu, P., Farzaneh, M., 2010. A CFD approach for modeling the rime ice accretion process Virk, M.S., Homola, M.C., Nicklasson, P.J., 2010. Relation between angle of attack and
on a horizontal axis wind turbine. J. Wind Eng. Industrial Aerodynamics 98 (4), atmospheric ice accretion on large wind turbine's blade. Wind Eng. 34 (6), 607–613.
181–188. Virk, M.S., Homola, M.C., Nicklasson, P.J., 2010. Effect of rime ice accretion on
Habashi, W.G., Morency, F., Beaugendre, H., December 2001. FENSAP-ICE: a aerodynamic characteristics of wind turbine blade profiles. Wind Eng. 34 (2),
comprehensive 3D simulation tool for in-flight icing. In: 7th International Congress of 207–218.
Fluid Dynamics and Propulsion, Sharm-El-Sheikh, Egypt. Virk, M.S., Homola, M.C., Nicklasson, P.J., 2012. Atmospheric icing on large wind turbine
He, lo-iacute, Beaugendre, S., Fran-atilde, Morency, O., Habashi, W.G., 2003. FENSAP- blades. Int. J. Energy Environ. 3 (1), 1–8.
ICE's three-dimensional in-flight ice accretion module. ICE3D. J. Aircr. 40 (2),
239–247.

52

Vous aimerez peut-être aussi