You are on page 1of 11

Energy 90 (2015) 680e690

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Comparative analysis of turbulence models for the aerodynamic


simulation of H-Darrieus rotors
szlo
La  Daro bor Janiga a, *, Klaus Petrasch b, Michael Webner c,
 czy a, Ga
Dominique The venin a
a €tsplatz 2, Magdeburg D-39106, Germany
Lab. of Fluid Dynamics and Technical Flows, University of Magdeburg “Otto von Guericke”, Universita
b
Windpower4u, Tro €bigauer Str. 1a, Schmo
€lln-Putzkau D-01877, Germany
c
VENTEGO AG, Chemnitzer Straße 75, Limbach-Oberfrohna D-09212, Germany

a r t i c l e i n f o a b s t r a c t

Article history: The importance of wind energy has increased at a rapid pace in the last years. As a result, increasing
Received 24 March 2015 efforts are taken to improve the efficiency and extend the applicability of wind turbines to all suitable
Received in revised form locations. Although HAWTs (horizontal axis wind turbines) are clearly the most well-spread, VAWTs
5 July 2015
(vertical axis wind turbines) show several advantages. In this category, the H-Darrieus configuration is
Accepted 23 July 2015
Available online 24 August 2015
particularly popular.
For evaluating the performance of an H-Darrieus wind turbine, two-dimensional simulations relying
on CFD (computational fluid dynamics) have been widely employed, since detailed full three-
Keywords:
Wind energy
dimensional simulations for an optimization remain beyond reach with current computer power. Un-
Darrieus fortunately, there is up to now no agreement in the scientific literature concerning the best possible
CFD model and model parameters for such conditions. At the same time, different models, different solvers
Turbulence and different meshing strategies may yield very different results.
Validation The aim of the current study is to perform a systematic numerical analysis in order 1) to identify the
H-rotor necessary mesh resolution for the different turbulence models, 2) to choose a model suitable for opti-
mization and 3) to compare the characteristic curves obtained with the different models for four
different configurations.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction increasingly considered. For instance, Fischer et al. [3] performed


multi-objective optimization of horizontal axis wind turbines with
The energy consumption of the world increases rapidly. Due to Blade Element Momentum (BE-M) code.
the increasing awareness of the population, to alarming reports Most installed turbines are Horizontal Axis Wind Turbines, but
concerning global warming, air pollution and nuclear accidents, VAWTs (vertical axis wind turbines), and especially the H-Darrieus
these increasing needs have to be covered by clean and reliable wind turbines are very interesting as well, especially for small-scale
energy sources. Wind energy can be a suitable answer to these applications, due to their low cost and easy installation at almost
needs. As a consequence, the gross production of wind energy in any location, and due to more favorable interaction effects at low
the EU27 (European Union with 27 members) increased from 80 PJ inter-turbine distances. As a yaw system is absent and since all
(2000) to 537 PJ (2010) just over a period of ten years [1], and an electrical and mechanical components are close to the ground, it
increase to 2300 PJ is prognosed for 2030 [2]. has a relatively simple design [4], which decreases installation and
In order to improve further wind energy outcome, efficiency has operation costs dramatically. Darrieus turbines show relatively low
become a crucial factor and optimization of wind turbines is being sound emissions and have increased performance in skew [5] as
shown, e.g., by Bianchini et al. [6]. Furthermore, they are omni-
directional and are thus better suited for areas with strongly
* Corresponding author. inhomogeneous topology, like in urban and peri-urban environ-
E-mail addresses: laszlo.daroczy@ovgu.de (L. Daro czy), janiga@ovgu.de ments [7]. Last but not least, Darrieus turbines can withstand larger
(G. Janiga), klaus.petrasch@gmx.de (K. Petrasch), michael.webner@ventego-ag. windspeeds due to their aerodynamic behavior.
com (M. Webner), thevenin@ovgu.de (D. Thevenin).

http://dx.doi.org/10.1016/j.energy.2015.07.102
0360-5442/© 2015 Elsevier Ltd. All rights reserved.
czy et al. / Energy 90 (2015) 680e690
L. Daro 681

Nomenclature T torque, time period


TVR turbulent viscosity ratio
A projected area u wind speed
c camber/chord length
CD/L/T/N drag/lift/thrust/normal coefficient Greek letters
CFL CouranteFriedrichseLewy number D difference
Cp power coefficient ε turbulent dissipation rate
CQ torque coefficient l,TSR tip-speed ratio
Cx,y global force coefficients m dynamic viscosity
D rotor diameter r density
F force s solidity
H height 4 azimuth angle
I turbulence intensity J uniformity factor
k turbulent kinetic energy u angular velocity, specific dissipation
L integral length scale
N number of blades Sub/superscripts
P performance act actual value
R rotor radius ref reference value
Re Reynolds number th. theoretical
t time x,y,z Cartesian coordinate system directions

The original patent of the Darrieus rotor from Georges Jean [18], annual energy yield [19], aero-elastic behavior and loading
Marie Darrieus dates back to 1927, but its very complex aero- [20,21], optimal layout [22] or robustness [23] could also be
dynamics are still not completely understood. Recently, a few ar- considered later in a multi-objective optimization. Very few such
ticles dealt with the analysis and optimization of Darrieus turbines. combined studies are already available, like that of Go ۍmen and
Using CFD techniques, Mohamed [4] investigated different airfoils €
Ozerdem [24], using optimization for noise reduction.
for performance improvement of an H-Darrieus rotor. Bedon et al. The aerodynamic efficiency of commercially available rotors is
[8] used a BE-M model for the optimization of a Darrieus rotor, already very high (large-size HAWTs can reach up to 50%). A further
while Castelli et al. [9] proposed a new performance prediction improvement is only possible using advanced optimization
model for Darrieus turbines. Gosselin et al. analyzed the flow methods. Even with such techniques, the resulting increase in
around an H-Darrieus rotor while varying solidity, pitch angle or performance cannot exceed more than a few percents. Therefore, it
number of blades [10]. is essential to reach sufficient precision in corresponding CFD
The Darrieus rotor is a vertical-axis lift-type rotor, i.e, the lift forces simulations, in order to be sure that the resulting improvement is
on the airfoil are used to generate the performance (in contrast to not simply due to model uncertainties. Moreover, as computing
Savonius rotors, mostly drag driven [11]). The flow physics is inher- time is an essential factor, one has to identify the really necessary
ently unsteady due to the varying angle of attack, which leads to a mesh resolution and modeling approach, delivering a high accuracy
higher complexity compared to standard HAWTs. In particular, dy- at acceptable computational cost.
namic stall is of major importance but is still not completely under- The final goal is to perform an optimization based on EA
stood [12]. As a result, the turbulent flow field is extremely complex (evolutionary algorithm) for the performance coefficient using the
and there are still contradictory statements in the literature con- selected model and parameter setting. For this purpose, based on
cerning the most appropriate turbulence model. For instance, our past experience with similar challenges, the model should be
Mukinovi c et al. [13] have used SpalarteAllmaras and k  u SST fast enough to perform at least 500 simulations with no more than
model, while Lanzafame et al. used k  u SST and Transitional SST 30 computers in no more than 3 months runtime. Besides, the
model in 2D CFD simulations to evaluate the performance of H-Dar- model should provide mesh and temporal independency with no
rieus rotors [14] and concluded that Transitional model is the best. In more than 2% error.
contrast, Gosselin et al. have tested SpalarteAllmaras (with modified
strain-based formulation), k  u SST with low Reynolds corrections
and Transition SST and concluded, that the k  u SST model was the 1.2. Performance evaluation of Darrieus wind turbines
most appropriate [10]. The k  u SST model was also applied by Jer-
icho et al. for a novel wind turbine concept [15] or by Campobasso Early models for the evaluation of wind turbines were based on
et al. for HAWTs with harmonic balance method [16]. Castelli et al. the Glauert actuator disk theory [25]. Starting from the Single
found that k  u SST model is the most appropriate in 3D, while k  ε Streamtube Model [26], the Multiple Streamtube Model and the
Realizable model is more accurate in 2D [17]. Double Multiple Streamtube Model [27] were progressively
developed. Following the development of vortex-based models, the
1.1. Aim of the current study next natural step was to rely on CFD simulations. There are three
main families of turbulent CFD computations, namely the DNS
The ultimate goal of the present project is to increase the per- (direct numerical simulations), the LES (large eddy simulations),
formance of H-Darrieus rotors by optimizing the airfoil shape based and the simplified approach relying on Reynolds-Averaged Navier
on 2D CFD simulations. For this purpose, the first questions that Stokes equations (RANS, or URANS for Unsteady RANS). DNS for
must be answered concern mesh independency, turbulence Darrieus rotors is still infeasible, since resolving all scales at this
modeling and boundary conditions, in order to ensure sufficient Reynolds number is too expensive. LES has been employed only in a
precision. Although only performance is considered here, noise few studies (e.g., [28]), but due to the very high computational costs
682 czy et al. / Energy 90 (2015) 680e690
L. Daro

compared to RANS, LES is clearly incompatible with any optimi- For higher tip speed ratios the loss of efficiency is the result of
zation process and can only be used to investigate specific aero- secondary losses: strut losses, wingtip losses, etc. These losses
dynamic aspects. As a result, the present manuscript considers only could only be properly computed with detailed 3D computations.
URANS simulations.
2. Limitations in the comparison of numerical and
1.3. Physics of the H-Darrieus wind turbine experimental investigations

The operation of wind turbines is usually characterized by the Although in some articles a nearly perfect agreement between
characteristic curve, or TSR-Cp curve. The tip-speed-ratio (TSR, or l) CFD and experiment is discussed, such an observation should not
is the ratio of the highest velocity of the rotating blade compared to indeed be expected. Even if CFD can be used to resolve the flow field
the wind speed (u): around the rotating turbine, important details will always be
missing in the CFD. As a whole, the resulting differences can be
vtip Ru quite large. In many studies the predicted peak efficiency exceeds
TSR ¼ l ¼ ¼ ; (1)
u u the measured one by more than 75% [12].
Minor issues encompass the uncertainty of the measurement
where R is the radius of the rotor and u is the angular velocity of the devices in the experiment, geometrical tolerances during
rotor. In case of the H-type Darrieus rotor the cross-section does not manufacturing, insufficient temporal or spatial resolution in the
vary with height, i.e., the rotor has the same speed for each cross- simulation, inadequate level of convergence, rounding errors,
section (not like Troposkein rotors). In the current article 4 ¼ 0 missing fluctuations in the rotation speed (present in the reality,
means, when the blade starts moving against the wind and 0e180 but not in the simulation), unknown turbulence conditions (they
denotes upstream position, see Fig. 1. cannot be exactly measured). For 2D simulations, the lack of end-
The power coefficient (Cp) is the indicator of efficiency. It is the plate friction and the missing exponential wind profile addition-
ratio of the power generated by the wind turbine compared to the ally lead to mistakes.
available energy in the wind: More significant terms are the generator losses (if only the po-
wer output is measured) and bearing losses. However, these terms
Pmech Tu Tu
Cp ¼ ¼1 ¼1 ; (2) can be in principle measured and taken into account.
Ptotal 2 ru3A
2 ru 3 2RH
The major source of discrepancy is connected to the strut losses
and strutejunction wake interaction. Maitre et al. came to the
where A is the projected area of the rotor, r is the density, H is the conclusion, that the main losses do not come from the arms friction,
height of the rotor and T is the (average) torque generated by the but from the tip and arm-blade junction vortices [35]. In contrast
rotor (average, because a torque ripple is present [29]). This value with this observation, the 3D CFD study of Castelli et al. indicated
shows a theoretical upper limit, which is known as Betz limit that, for small aspect ratio rotors the arm friction losses can result
(59.3%). In reality, even modern HAWTs usually cannot reach more in more than 20% loss of performance [17]. Fortunately, the strut
than Cp ¼ 45e50% [30,31]. losses can be analytically estimated with a good accuracy [36,37].
Although the characteristic curve (l  Cp) is often regarded as a The largest and most difficult terms are the wing-tip vortex losses.
non-dimensional indicator of the performance, the same tip-speed In wind turbines with small aspect ratios these losses can be large
ratio can correspond to different local chord Reynolds numbers (up to 25% [17]), while at high aspect ratio these effects are small. As
(Relocal ¼ cRu/n), where c is the chord length and n is the kinematic shown in the study of Gosselin et al., the last 20e30% of the blade
viscosity of air. The dependence of the power coefficient on the height are less effective for energy generation [10]. Amet also came
Reynolds number can be experimentally measured [32]. to a similar conclusion, with 22% loss of performance due to the tip
The optimal tip-speed ratio depends on the airfoil shape and on vortices and blade/arm junctions [38], while Qin et al. reported 40%
the solidity, which is defined as s ¼ Nc/D [33], where N is the losses compared to 2D simulations [39]. Finally, the uncertainty of
number of blades. Lower solidity rotors operate at higher tip-speed the turbulence models can also result in large errors and will be
ratios, while higher solidity rotors typically have low optimal tip- considered more specifically in what follows. Moreover, in the re-
speed ratios, as discussed also in Ref. [34]. The loss of perfor- ality many revolutions can be measured to average out the effect of
mance at lower tip-speed ratios originates from the stall mecha- vortex shedding and wind gusts; in a simulation computing such
nism (when the angle-of-attack, AOA of the blade increases above large number of revolutions is very time consuming and thus not
12e13 [35]). feasible.

3. Computational method

This section describes the complete workflow of the simulation.


For the current study a geometry was chosen, which is currently
under development for an industrial project. The rotor has three
blades with R ¼ 1.5 m radius, each blade is symmetric. The blades
have a chord of c ¼ 160 mm, thus solidity is s ¼ 0.32 and the
optimal tip-speed ratio is 3 < l < 4. The exact airfoil profile cannot
be disclosed yet due to confidentiality requirements from the in-
dustrial partners of this project.

3.1. Automatic workflow

Systematic studies are always subject to human errors. Thus, in


Fig. 1. Schematical representation of the conventions used for the H-Darrieus rotor in order to eliminate this problem, the complete workflow has been
the current study. fully automated. The analysis and the evaluation of the different
czy et al. / Energy 90 (2015) 680e690
L. Daro 683

configurations have been carried out using the OPtimization Al-


gorithm Libraryþþ (called simply OPALþþ). OPALþþ is an object-
oriented multi-objective optimization and parameterization
framework developed at the University of Magdeburg “Otto von
Guericke” [40]. The software has already been successfully applied
to many different problems [40,41] and is focused especially on
CFD-based problems.
For the present study a module has been created for OPALþþ. In
this module the complete CFD workflow was parameterized, from
the geometry creation (number of blades, geometry of the blade,
radius of the rotor, etc.), through the meshing (mesh size, size of
domain, etc.) to the CFD setup (choice of CFD software, choice of
solver, turbulence model, boundary conditions, etc.).

3.1.1. Domain decomposition


As the flow around a Darrieus rotor is highly transient, it has to
be resolved not only in space, but in time as well: a moving mesh
has to be created with a sliding mesh technique. This means, that
the domain of the computation has to be decomposed into a sta-
tionary domain, which will be used to simulate the free stream flow
around the rotor (Fig. 2), and a moving domain, which will rotate
with the blades (Fig. 3). Between the two domains an interface is
used in order to interpolate the flow quantities. In this study the
same circular decomposition was applied as e.g., by Castelli et al. Fig. 3. Rotating domain of the mesh.
[17]. Another possible decomposition could have been a ring
structure [7,14]. Preliminary tests have shown no significant dif-
ferences between both. 3.1.2. Mesh size and mesh independency
The mesh was created with ANSYS Gambit 2.4.6. Unstructured In published studies very different mesh resolutions and tem-
parts are plotted with a brick pattern in Figs. 2 and 3, the density of poral resolutions have been applied. For a corresponding review
the pattern indicating the mesh resolution. If a shaft is not present please refer to Trivellato and Castelli [42].
in the computation, the inner structured domain can be simply The mesh size is very important for the turbulence as well. If the
removed and replaced with an unstructured mesh. For the struc- mesh spacing is too coarse, the decay of turbulence will be grossly
tured parts the Map algorithm is applied, while for the unstruc- overestimated [43]. The reference mesh sizes are defined at seven
tured parts the quadrilateral Pave method is used. In the current different locations (S1,…,S7): on the sides of the middle rectangle in
study only quadrilateral cells are used (following the recommen- the stationary domain (90 mm), inside the middle rectangle in the
dation of [12]), as triangular cells are well-known to be dissipative stationary domain (60 mm), on the interface (20 mm), on the wall
in nature. However, tests have indicated in the present study no of the shaft (2.5 mm), on the blade (0.375 mm) and in the middle of
large differences between triangular or quadrilateral cells for the the rotating domain (13.5 mm) and around the blade in the addi-
mesh-independent resolution. tional control area (4.8 mm). The Cþþ module uses additional size
The employed hybrid structuredeunstructured grid (as shown functions to ensure a smooth transition between the different sizes.
schematically in Figs. 2e3) provides the advantage, that it increases The nodes on the surface of the blade were uniformly distributed,
the precision near the interface (where interpolation errors can be except at the trailing edge, where a refinement was used with
large) and in the wake of the blades, while being still very flexible, exponential growth factor. Throughout the studies, a calibration
thus enabling automatic mesh generation of different blades. A coefficient (ccalib) was used to compute the actual mesh sizes from
completely structured approach was tested as well, but tests did the reference size (Sact,i ¼ ccalibSref,i). In case of the boundary layer
pffiffiffiffiffiffiffiffiffiffi
not indicate significant differences. the refinement is modified to yact ¼ yref ccalib (with tested values
of 1.5, 1.25, 1.0, 0.75, 0.5). The quality of the mesh was checked and
recorded in all cases by the software and reported in the output log
file. Orthogonality (as defined by Fluent) was always above 0.7, and
maximum skewness angle (as reported by StarCCMþ) was always
below 40 .
In order to minimize the interpolation error between the two
domains, the mesh resolution in the validations is chosen so, that
throughout the rotation the nodes will overlap with each other.
For a precise computation the boundary layers have to be
resolved up to the laminar sublayer. This is ensured by computing
automatically the thickness of the first layer using the Schlichting
correlation with the camber of the blade as reference length and
U ¼ (l þ 1)uwind as the relative speed. Then, yþ þ
avg z0:2, ymax z1:0
was maintained for all computations (except for the test with the
k  ε Realizable model with Wall function). This strong criterion
was applied following the recommendations of Maitre et al. [35].
The final mesh around the trailing edge of a sample geometry can
Fig. 2. Stationary domain of the mesh. be seen in Fig. 4.
684 czy et al. / Energy 90 (2015) 680e690
L. Daro

0.1
CQ
0.05

-0.05

-0.1 C15 P15


C24 P24
-0.15 C50 P50
C100 P100
φ
Fig. 4. Example of the obtained mesh resolution near a random blade geometry.

-0.2
3.2. CFD setup 0 60 120 180 240 300 360

3.2.1. Material properties & discretization Fig. 5. Torque coefficient with different solvers and number of inner iterations in
As small Darrieus rotors do not have high rotational speed, the ANSYS Fluent (C denotes Coupled Solver, P denotes PISO solver, the number denotes
the number of inner iterations).
relative speed of the blades always stays well below M ¼ 0.3
(z100 m/s). As a result, air can be modeled as an incompressible
medium, as widely done in the literature (e.g., [17]). least 75 inner iterations were necessary for convergence with the
Reference density r ¼ 1.225 kg/m3 was chosen according to DIN SIMPLE method.
EN 61400 [44], and the dynamic viscosity was chosen corre- This proved to be a very strict convergence criterion. The
spondingly as m ¼ 1.7894$105 Pa s. different residuals, as reported by Fluent, generally dropped to
In the present work a second-order implicit temporal dis- 105e1010 depending on turbulence model and angular position.
cretization is always applied, together with second-order (central All residuals always were reduced at least to 103.
or upwind) derivatives, if not stated otherwise. If for a specific case
or model (e.g., with k  ε RNG always) divergence is detected, the 3.2.3. Boundary conditions
simulation is restarted using first-order discretization and Segre- The four sides of the computational domain are defined as ve-
gated solver. locity inlet, pressure outlet and symmetry boundaries. Proper inlet
conditions are clearly the most important ones. Unfortunately, the
3.2.2. Segregated vs coupled solution of the fluid dynamic equations specification of inlet turbulence properties is a very difficult ques-
When solving the fluid dynamics equations, there are two tion, particularly so for wind turbines. Different sources propose
possibilities for the pressureevelocity coupling: the segregated incompatible values and procedures. Although one could use
approach (SIMPLE, SIMPLEC or PISO in ANSYS Fluent and Segre- values derived from corresponding experimental measurements,
gated in CD-Adapco StarCCMþ), where a predictor-corrector turbulent variations are usually post-processed from measured
approach is applied, and the Coupled solver, where the mo- wind fluctuations during 10e60 min [44,50]. As noted by Spalart
mentum and continuity equations are directly coupled. In transient and Rumsey [43], in the atmospheric boundary layer the typical
cases the coupled algorithm is necessary for poor mesh quality or length scales are around 100 m and the eddy viscosity can reach
large time steps [45]; this solver can dramatically increase the 50 m2/s on windy days (corresponding to a turbulent viscosity ratio
performance even for incompressible flows [46]. According to the TVR ¼ 3.3$106!). However, the boundary layer will couple only with
CD-Adapco StarCCMþ User Guide [47], the Coupled solver is always length scales in the order of 1 cm [43] and, due to the large rotation
recommended when sufficient resources are available due to its speed, only a part of the turbulent spectra will really interact with
increased robustness. Another advantage of the Coupled solver is the blades in the CFD [51].
that the CPU time scales linearly with the number of cells, i.e., the In our first numerical studies different values were tested. After
convergence rate will not deteriorate when refining the mesh confirming that small values are almost equivalent to each other,
resolution. final values of I ¼ 0.1% and TVR ¼ 10 were retained for all further
Again, different authors do not agree concerning this topic. The configurations, in agreement with recommendations from the
SIMPLE method was applied, e.g., in the study of Lain and Osorio literature [14,52]. Such low intensities are not rare in wind tunnels.
[48], while the PISO method was finally retained by Lanzafame et al. E.g., the wind tunnel of TU Delft has I ¼ 0.015% at 10 m/s [5].
[14]. Instead, Qin et al. [39] or Balduzzi et al. [49] have applied the
Coupled solver due to its increased robustness. 3.2.4. Temporal discretization e size of time step
For the geometry discussed later for numerical analysis (Oper- To determine the time step size, the CFL (CouranteFrie-
ating condition 2, k  u SST model), the PISO and Coupled solvers drichseLewy) number can be used. In CFD, CFL ¼ 1 is usually rec-
were compared using ANSYS Fluent 14.0. As a convergence crite- ommended for stability. However, with an implicit time integration
rion, the number of inner iterations per time step was fixed to 15, scheme, much larger values are allowed. In the current study the
24, 50 and 100 for both solvers. Fig. 5 shows the torque coefficient time-step was thus defined as
for blade 1, for the 15th revolution. One can see, that the results
with the coupled solver stay identical when enforcing more than 24 2p
Dt ¼ ; (3)
iterations, while for PISO the agreement is only achieved for more uD4
than 100 iterations. Thus, the Coupled solver was always employed
with 24 inner-iterations in what follows. This is the faster solution. where D4 ¼ 2k(k ¼ 0,1,2) is the angle of rotation for each tested
These results support the conclusions of Maitre et al. [35], where at time step. This criterion is strongly correlated with the CFL number.
czy et al. / Energy 90 (2015) 680e690
L. Daro 685

Then, the post-processing becomes very simple, as a single revo- Table 1


lution will be 360$2k time steps. Of course, the independence from Comparison of performance coefficients for four different domain sizes (l ¼ 2.28,
k  u SST model).
the temporal resolution still has to be analyzed.
This strategy offers another advantage: if the nodes on the D1 D2 D3 D4
interface are equidistant, and equal in number to n$360$2k, they Linlet[m] 15 20 30 45
will overlap at each time step. Instead of interpolation, the values Loutlet[m] 30 40 60 90
can be copied directly between the moving and stationary domain Lsym[m] 18.75 25 37.5 56.25
Cp[%] 23.58 23.46 23.46 23.75
(conformal match). This is automatically detected by some solvers
(e.g., CD-Adapco StarCCMþ). In the experimental validations 720
points were always used. The study of Gosselin et al. proposed boundary of the acceptable domain. Hence, D2 was retained for all
similar values, i.e., 1000 time-steps [10], while Castelli et al. [17] further simulations.
used 360 time-steps per revolutions.
4.2. Impact of boundary layer resolution

3.2.5. Temporal discretization e number of revolutions In order to analyze the effect of the boundary layer resolution,
Reaching a quasi periodic solution can require a large number of the thickness of the first layer and the growth rate were varied. A
revolutions, leading to very long computations. Moreover, when growth rate of 1.1, 1.175 and 1.225 was tested with different thick-
starting the computation, the flow field is not realistic around the
nesses resulting in yþ þ
max 2½0:23  3:65 and yave 2½0:04  0:63 for
blade. To solve this problem, a two-level approach was applied. In
the last revolution. With l ¼ 2.28, I ¼ 0.1%, TVR ¼ 10 and k  u SST
the first part the flow field around the blade does not necessarily
model the resulting coefficients were all within the range of Cp2
have to be computed precisely, thus D4 ¼ 4...6 is used to speed up
[23.41e23.70], and the hysteresis curves indicated only very small
the initialization of the flow field. Later, in the second part D4 ¼ 1,
differences as well. This confirms that when using the default
D4 ¼ 1/2 , D4 ¼ 1/4 or an even smaller value is employed.
values (growth rate z1.2, yþ < 1), the models will deliver suitable
In the current study at least 10 (coarse)þ5 (detailed) revolutions
results in all cases.
were always used (for slower wind speed or faster rotation, the
number of revolutions were increased). Although this seems to be
4.3. Operating condition 1 - low turbulence intensity l ¼ 3.8
large at first look, similar recommendations are found in the liter-
ature (e.g., Gosselin et al. used 20 revolutions [10], Maitre et al.
In the first numerical study the turbulence models were
considered 11 revolutions [35]). To evaluate the average perfor-
analyzed for l ¼ 3.8 and low turbulence, I ¼ 0.25%, integral length
mance, only the last revolution is post-processed.
scale L ¼ 0.15 m, In this case the angle of attack was rather low
resulting in only weak stall effects and in smooth torque curves.
4. Systematic numerical analysis Analysis of the vortex structure revealed a wide filament, as also
found by Maitre et al. [35]. For all cases five different mesh con-
In order to identify the optimal setup and mesh for the CFD figurations were analyzed, with ccalib ¼ 1.5, 1.25, 1.0, 0.75, 0.5. For
computations, an extensive numerical analysis was performed, the time-steps D4 ¼ 1 was applied. When changing the time-step
involving many different tests. In particular, the domain size, to D4 ¼ 0.5 , the total change in the performance coefficient with
boundary layer resolution, mesh resolution, temporal resolution, the k  ε Realizable model was only DCp ¼ 0.02%. Investigating the
different turbulence specifications and different tip-speed-ratios performance and torque not only in an integral manner, but also in
were analyzed. Unless otherwise specified, all comparisons have a time-dependent sense, revealed no significant changes either.
been done with ANSYS Fluent 14.0. This is in agreement with the study of Maitre et al. [35], where
In order to enable a short and quick description of the turbu- D4 ¼ 1 was found to be sufficient for appropriate accuracy.
lence models, the following notations are used in all Tables: As strong dynamic effects are missing, in Fluent solver CFL ¼ 200
T0 ¼ k  ε Realizable with Enhanced Wall Treatment; T0* ¼ k  ε (default value) was used. In StarCCMþ lower values (between 25
Realizable with Standard Wall Function (yþ > 30); T2 ¼ k  ε RNG, and 35) were applied to ensure stability. The results are presented
T3 ¼ SpalarteAllmaras, T4 ¼ SAS, T5 ¼ k  u SST, T6 ¼ k  u; in Table 2, together with the number of cells and with the number
T7 ¼ Transitional k  kl  u; T8 ¼ Transitional SST in the standard of nodes on the interface (int.).
formulation implemented in ANSYS Fluent [45]. For selected cases
Table 2
the simulations were performed using StarCCMþ as well to allow
Comparison of performance coefficients for the last revolution for operating con-
for comparison: all models starting with S denote the equivalent dition 1 with 8 different turbulence models Ti, 5 mesh resolutions Mi and 2 CFD
model in StarCCMþ with All-yþ Wall Treatment. As different tur- codes.
bulence intensities and tip speed ratios result in different stall
M5 M4 M3 M2 M1
mechanisms and angle of attacks, four different cases were tested.
cell 201k 296k 368k 588k 1089k
int. 360 720 720 1080 1440
4.1. Impact of domain size CpT0 40.04 40.14 40.19 40.22 40.23
CpT0 36.86 37.84 38.93 39.72 40.50
In order to analyze the effect of the domain size on the perfor- CpT2 35.66 35.24 35.18 34.89 34.38
mance coefficient, four different sizes were tested (D1-D4). Table 1 CpT3 34.49 34.45 34.50 34.47 34.43
presents the performance coefficients for the corresponding CpT4 42.44 42.53 42.66 42.83 43.00
domain size (Linlet, Loutlet and Lsym represent the distance from the CpT5 42.26 42.36 42.48 42.52 42.54
central rectangular area to the inlet, outlet and symmetry bound- CpT7 49.40 49.50 49.66 49.78 49.87
ary, respectively; Lrect ¼ 8 was kept constant in all cases). The CpT8 45.47 45.70 46.01 46.51 46.42
computations were done with l ¼ 2.28, I ¼ 0.1%, TVR ¼ 10 and k  u 40.89 40.99 41.09 41.10 41.44
CpST0
SST model. All coefficients agreed and the hysteresis curves did not 35.17 35.27 35.34 35.36 35.41
CpST3
reveal significant differences either. Choosing the first variant 42.50 42.56 42.82 42.95 43.38
CpST5
might be inappropriate for some cases, as it could be on the
686 czy et al. / Energy 90 (2015) 680e690
L. Daro

The standard k  u model was tested as well, but resulted in Table 3


unphysical results. This is no surprise, since the k  u standard Comparison of performance coefficients for the last revolution for operating con-
dition 2 with 6 different turbulence models Ti, 5 mesh resolutions Mi and 2 CFD
model is intended only for fully turbulent flows. In StarCCMþ the codes.
AKN k  ε Realizable model resulted also in unacceptable oscilla-
tions. Model T0* (k  ε Realizable with Wall Function) did not M5 M4 M3 M2 M1

provide appropriate results either (in this case the boundary layer cell 226k 266k 370k 547k 1091k
was calibrated for yþ > 30). The reason for this failure was that int. 540 540 720 900 1440
CpT0 35.49 36.30 37.66 38.0 38.65
maintaining yþ > 30 and a cell aspect ratio of 1 was impossible for
CpT2 33.54 33.89 34.77 35.18 35.83
the finer mesh resolutions.
  CpT3 25.27 25.54 26.21 26.27 26.59
 
All models managed to reach DCp ¼ CpM1  CpM3  < 1%, and most 21.01 20.80 22.60 22.05 19.90
CpT4
of them provided even stricter results. Thus, the grid corresponding CpT5 21.62 21.77 23.63 23.95 24.05
to M3 provides an appropriate resolution. The agreement between CpT8 16.53 17.69 18.44 20.29 21.4
StarCCMþ and Fluent is satisfactory as well, although CpST5 19.83 19.68 23.35 22.93 24.65
StarCCMþ always overestimated the power coefficient by approx-
imately 1% compared to Fluent. However, the hysteresis curves did
correspond to the appearance of dynamic stall and wake interac-
not reveal significant differences between StarCCMþ and Fluent.
tion. The k  u based models result in higher fluctuations.
The 4  CQ ð4Þ curves are presented for blade ‘1’ in Fig. 6 for all
turbulence models computed in Fluent.
4.4.1. Impact of turbulence intensity
When comparing different, but low turbulence specifications,
4.4. Operating condition 2 e low turbulence intensity l ¼ 2.28 the results were found to be very similar. The comparison was done
for the k  u SST model, which provided the most fluctuating but
For operating condition 2 the turbulence was still kept low still converged solution for the previous, stall conditions. When
(I ¼ 0.25%, L ¼ 0.15 m), but the tip-speed-ratio was decreased to varying I, no significant changes were detected in the time-
increase the angle of attack, and thus stall effects. Due to their dependent curves. The performance coefficients were very similar
inappropriate performance in the first test, following approaches as well, for instance 23.50% for I ¼ 0.1%, TVR ¼ 10, compared to
have not been considered further: k  ε Realizable (T0*) with Wall 23.63% for I ¼ 0.25%, L ¼ 0.15 m. As the former set of values has been
Function; k  u (T6), and Transitional k  kl  u (due to instability widely used in publications [14,52], it was retained for the exper-
problems with the tested Fluent version when running it in parallel imental validations.
on Linux). The results are shown in Table 3.
For the temporal resolution D4 ¼ 0.5 was used. Coarser reso- 4.5. Operating conditions 3 and 4 e high turbulence intensity
lutions did not result in temporal independency (the hysteresis l ¼ 2.28
curves indicated noticeable differences). Using D4 ¼ 0.125 resulted
in only 1.5% difference for k  ε Realizable model, 0.8% for Spa- In the 3rd test moderately high turbulence conditions were
larteAllmaras and 2.2% for k  u SST in the performance coefficient. tested (I ¼ 7%, L ¼ 0.3 m; based on the NRC wind tunnel [53]). SAS
As the maximal allowable error was chosen as z2%, D4 ¼ 0.5 was and Transitional models failed to converge, SpalarteAllmaras and
deemed acceptable. One has to keep in mind, that this excellent k  ε RNG models have experienced significantly decreased per-
temporal independency has been achieved due to the robustness of formances, while other models delivered similar results. The
the Coupled solver, even at large CFL numbers. models were again able to reach for all cases
   
   
All models were able to satisfy DCp ¼ CpM1  CpM3  < 2%, except DCp ¼ CpM1  CpM3  < 2%.
SAS and Transitional SST, where M2 resolution is necessary (in case In the 4th test case (I > 20%, according to DIN [44]) most tur-
of T4 (SAS) model the SIMPLE solver was used due to convergence bulence models start to have problems. The k  ε Realizable, RNG
problems). and SpalarteAllmaras lead to significantly reduced performance
The phase-angle dependent torque values are presented in (z27%, 22% and 8%, respectively), while the curves become very
Fig. 7. As one can see, the predictions are very different. The largest smooth and the strong oscillations and peaks in the hysteresis
differences are seen at 4 ¼ 60e160 and 4 ¼ 200e300 , which
0.1
0.05 CQ CQ
0.05
0
0

-0.05 -0.05
k-ε Real. SAS
k-ε Real.* k-ω SST -0.1 k-ε Real. SAS
-0.1 k-ε RNG k-ω SST
k-ε RNG k-kl-ω
Spal.-All. Tr. SST -0.15 Spal.-All. Trans. SST
-0.15 ϕ ϕ
-0.2
0 60 120 180 240 300 360 0 60 120 180 240 300 360
Fig. 6. Comparison of torque coefficients for the last revolution for operating condition Fig. 7. Comparison of torque coefficients for the last revolution for operating condition
1 with 8 different turbulence models. 2 for 6 different turbulence models.
czy et al. / Energy 90 (2015) 680e690
L. Daro 687

curves completely disappear. SAS and Transitional SST models fail total time domain of the measurement is denoted with T, the
to converge and k  u SST model provides completely unrealistic average turbulence intensity can be defined as:
values. In fact, as already stated earlier, the DIN standard results in
unrealistically high turbulence intensities (TVR > 106), which is the IðNDtÞ ¼ Iðt0 ; NÞ ct0 2T: (4)
explanation for these problems. As a result, only small turbulence
The I(NDt) curve was computed and plotted in Fig. 8 for all the
intensities could be used in a meaningful manner for URANS sim-
measurements (denoted as all), for a single measurement (stat.;
ulations of wind turbines.
length of the measurement: 2821 s), where the mobile platform
was standing and for a single measurement, where the mobile
4.6. Final recommendations
platform was moving (mov.; length of the measurement: 1169 s). As
one can see, the standing and moving spectra are significantly
Considering all the previous results, M3 resolution was finally
different; in the moving case the vibrations of the platform during
chosen, with 367k cells. This corresponds to at least 720 nodes
movement increase significantly the measured intensities.
along the interface and at least 720 nodes along the blades. A
However, although the turbulence is high, the turbines are
coarser resolution is not sufficient.
operating at smaller time scales than the atmospheric turbulence.
Concerning turbulence intensity, I ¼ 0.1% and turbulent viscosity
Thus, only a part of the spectra will interact with the blades, where
ratio TVR ¼ 10 is recommended. At least 10 þ 5 revolutions with
the frequency of the wind speed oscillations is similar to the fre-
D4 ¼ 0.5 are needed for proper time-resolution. Resolving 720
quency of the rotation [51]. This means, that the rotor will feel
time steps per revolutions together with at least 720 points on the
turbulence and gusts rather as a quasi-steady flow. In the NDt < 4s
interface ensures CFL  1 for stability.
range (inlaid plot) the curves collapse with very low turbulence
intensities. This explains and justifies the common practice of using
5. Experimental validation process
low-turbulence intensities in URANS-based simulations of wind
turbines, even under real wind conditions.
Although most standard turbulence models managed to reach
As measurements themselves are highly unsteady concerning
mesh independency at M3-level, they delivered significantly
wind speed, rotation speed and torque, a filtering method was
different results concerning performance. In what follows, four
applied. Only measurement intervals (NDt) were accepted, where
experimental measurement campaigns are compared with the re-
the root-mean-square of the rotation speed and wind speed stayed
sults of 2D CFD computations in order to help identifying the most
below a specified limit compared to the average (Ju ¼ uRMS < Jmax
u ).
suitable model. u
For all comparisons the previous recommendations have been Afterwards, the corrected efficiency can be determined using
taken into account. Although the integral Cp values are not able to the energy balance:
capture all aspects of the flow (e.g., the time-dependent distribu-
Ee  Es þ Eoutput þ Ebearing þ Estrut
tion of forces and torque in a single rotation), PIV or LDV mea- h2D z ; (5)
surements are rarely available in the literature and difficult to Ewind
obtain. Using such data would make it impossible to compare a
wide range of rotors and Reynolds numbers. Due to the lack of data, where Es ¼ 0:5Iz uðt0 Þ2 , Ee ¼ 0:5Iz uðt0 þ NDtÞ2
in the followings only performance coefficients will be compared.
t0Z
þNDt
The main purpose of the validation process was to see, if the
differences between experimental measurements and 2D simula- Ewind ¼ 0:5rA uðtÞ3 dt; (6)
tions show a systematic tendency, then opening the door for t¼t0
quantitative predictions based on 2D simulations. At the end this
was indeed observed, but only for the k  ε Realizable model and t0Z
þNDt
the k  u SST model. The differences compared to experiments Ebearing zMbearing uðtÞdt; (7)
were found to be constant, or cubic with the tip-speed-ratio,
t¼t0
respectively. A constant offset was already observed for the k  ε
Realizable model in other works as well [17]. The cubic offset found Finally, all measurement points were formed with N ¼ 8 length
for the k  u SST model is probably associated to the strut losses, (see Eq. (6)), and these measurement points were filtered to include
though further studies are needed to prove this point.
In all cases the original experimental measurements and the
50 10
characteristic curves of the 2D simulations are first shown unal- I[%] I[%]
tered, along with the modified curves taking into account the offset 8
(constant for k  ε Realizable, cubic for k  u SST) obtained from a 40 6
best fit.
4
5.1. Validation 1 30 2
N t
00
The first validation is based on the investigated geometry in our 2 4 6 8 10
current industrial project. In the first measurement the rotor was 20
placed on a platform moving at constant speed and an anemometer
was fixed at midheight in front of the rotor. The wind speed was
10 mov.
recorded every second (with 10 Hz averaging). The angular velocity
all
was measured with 1 Hz frequency; the torque was measured
stat N t
directly with the same frequency.
In a first step the turbulence spectrum was analyzed by defining 0
0 40 80 120 160 200
the turbulence intensity as a function of the time window (NDt, i.e.,
the duration used to compute the turbulence intensities). If the Fig. 8. Measured turbulence spectra for different time windows.
688 czy et al. / Energy 90 (2015) 680e690
L. Daro

only points with 6 < u < 11 m/s and with uniformity factor 200 mm (c ¼ 400 mm) resulted in a performance loss of almost
Jmax
u ¼ 0:16, Jmax
u ¼ 0:06. Cp ¼ 0.086.
From these points, tip-speed-ratio intervals were formed The performance curve in the wind tunnel and on the rooftop
(lmin;i  l  lmax;i ), and for each interval a single point was (thus in turbulent urban environment) were very close to each
computed by averaging the tip-speed-ratio and the performance other [51]. Even with 25% fluctuations the performance dropped
for the bin, forming a characteristic curve. The filtered measure- only by 10%, showing that the urban environment has very little
ments points, together with the obtained characteristic curve are impact on the performance.
presented in Fig. 9. The characteristic curve has been already In the simulation the rotor was placed in an open field. The
compensated for strut losses (based on an analytical model [36]) curves were computed using ANSYS Fluent. About 1700 points were
and bearing losses. used to resolve the blade with u ¼ const. ¼ 10 m/s. The resulting
For this experimental comparison the curves were computed number of cells was around 530,000.
using CD-Adapco StarCCMþ, with k  ε Realizable, k  u SST and Fig. 10 shows the characteristic curves with “Exp.” as obtained in
SpalarteAllmaras model. As one can see, the overall prediction of Ref. [54]. The shaft losses were added to this curve as measured in
the location of the maximal performance is excellent, but the tur- Ref. [33] (denoted with Exp. (þshaft)). As the struts were not
bulence models exhibit a very interesting tendency. With the k  ε aerodynamic, the assumption must hold that large cubic losses are
Realizable model, the numerically evaluated curve shows a con- present. The results with the different turbulence models are
stant offset of DCp ¼ 0.1 compared to the experiment. Alternatively, shown as well. In case of the Transitional model the trend is not
the k  u SST model exhibits roughly a cubic offset with correct as the optimal tip-speed ratio is shifted to larger values.
DCp ¼ 0.002l3. As the curves were compensated for bearing and Interestingly, for the present geometry all other models agree
strut losses, the differences can only originate from junction wake approximately with each other. All models predict the optimal tip-
interaction, wing-tip losses and turbulence effects. Currently speed ratio almost correctly. When using a cubic correction (“Exp.
running 3D simulations will be used to evaluate these losses. Also, (þshaft þ corr.)”, DCp ¼ 0.02l3, computed for the best fit), the
stationary measurements will be soon available in order to rule out agreement is even better. Unfortunately, without knowing the
the possible effect of platform vibrations. Before these results are exact geometry it is not possible to determine more accurately the
available, it is not possible yet to decide if k  ε Realizable or k  u corresponding cubic losses.
SST delivers the best agreement.
5.3. Validation 3
5.2. Validation 2
The third validation is based on the experimental work of Kjellin
The second validation from the literature is based on the et al. [55]. In this field test a three-bladed Darrieus rotor with 12 kW
experimental work of Bravo et al. [54], where a small (H ¼ 3 m, rated power output (u ¼ 12 m/s), H ¼ 5 m height and D ¼ 6 m
D ¼ 2.5 m), high-solidity (N ¼ 3, c ¼ 0.4 m) rotor was tested in the diameter was tested and measured for around 350 h. The blades
NRC 9 m  9 m Low Speed Wind Tunnel in Ottawa (I < 2% [51]). The were NACA0021 airfoils, with c ¼ 0.25 m chord length and tapered
rotor used NACA0015 profiles. The mount position was at 0.5 c [33] end to reduce the wingtip losses (the mounting position was not
and the trailing edge rounding was 4 mm [51]. indicated in the original article, but the difference in the pitch angle
Experiments have confirmed, that the characteristic curves of between 0.25 c and 0.5 c corresponds to only Db ¼ 0.6 ). The rotor
the Cleanfield rotor were almost invariant for 8 m/s  u  16 m/s used passive stall regulation with a direct drive. The test was con-
(u ¼ const.). The experiments were performed for the 0.8 < l < 2.2 ducted at 48e57 rpm. The blades were mounted with two struts at
domain, the maximum power coefficient achieved was Cp ¼ 0.3 for 17.6 , the struts having NACA0025 profile with 280e320 mm chord
l ¼ 1.6. It was further revealed in the work of Fiedler and Tullis [33], length. With the tested constant rotational speeds, the optimal
that the frictional and strut losses were quite large, 50e500 W performance of Cp ¼ 0.29 was found at l ¼ 3.30. Comparisons are
between 40 and 200 rpm. The study has also revealed, that the presented in Fig. 11. The computations were carried out using
mounting position (i.e., location of the chord-strut connection) has ANSYS Fluent. About 1200 points were used to resolve the blade. A
quite a large effect on the rotor. Mounting at 145 mm instead of

0.5 Exp. k-ε RNG


k-ω SST Exp. (+shaft) Spal.-All.
Cp k-ω SST-0.002λ3
k-ε Real. Exp. (+shaft+corr.) k-ω SST
0.4 k-ε Real.-0.1 k-ε Real. Trans. SST
Exp. 0.5
Spal.-All. Cp [%]
0.4
0.3
0.3
0.2 0.2

0.1
0.1
0 TSR [-]
TSR
0 0 0.5 1 1.5 2 2.5
1 2 3 4 5 6
Fig. 9. Experimental validation (1) based on own measurements. Fig. 10. Experimental validation (2) based on [54].
czy et al. / Energy 90 (2015) 680e690
L. Daro 689

Experiment (estimated max. error) CFD (52.5 rpm (Spal.-All.)) Castelli (exp.) k-ω SST (T5)
Experiment (48 &57 rpm) CFD (52.5 rpm (k-ω SST)) Castelli (CFD [5]) k-ε Real. (T0)
CFD (52.5 rpm (k-ε Real.)) CFD (52.5 rpm (k-ω SST corr.)) Spalart-Allmaras (T3) k-ε Real. (T0)*
CFD (52.5 rpm (k-ε Real. corr.))
CFD (52.5 rpm (k-ε RNG))
CFD (52.5 rpm (Trans. SST))
0.6

Power coefficient (Cp)


0.45
Power coefficient (Cp)

0.4 0.5
0.35
0.4
0.3
0.25 0.3
0.2
0.15
0.2
0.1 0.1
0.05 TSR(λ)
0 0
1.5 2 2.5 3 3.5 4 4.5 5 1.5 2 2.5 3 3.5
TSR(λ)
Fig. 12. Experimental validation (4) based on [17].
Fig. 11. Experimental validation (3) based on [55].
acceptable agreement. In spite of an overestimation at lower tip
shaft with 200 mm was added to the simulation as well. Total mesh speed ratios, the optimal tip speed ratio is correctly predicted. The
size was around 815,000 elements. overproduction at low l might be a consequence of the mounting
In Fig. 11, the typical behavior of the turbulence models can be position, since T0* shows the characteristic curve obtained for
seen again. The Transitional SST model overestimates the perfor- 0.25 c mounting (as computed with StarCCMþ). In case of the k  u
mance coefficient and the optimal tip-speed-ratio as well, due to SST model the performance curve is shifted slightly to higher tip-
the lack of calibration of the empirical correlation function. The speed-ratios. The underestimation at low TSR values for this
k  ε Realizable model is able to predict the shape of the charac- model was observed as well in many different studies [10,35], but
teristic curve and the location of the maximal performance, but not remains unexplained yet. The agreement between the present
the exact values. Instead, it shows a constant offset (DCpcorr ¼ 0:14). computations and the computations of Castelli et al. is very good
concerning the location of the maximal value, but the absolute
The modified curve (see k  ε Real. corr.) shows a very good
values do not agree, though all models are identical. The origin of
agreement. A similar behavior was experienced by Castelli et al. for
this discrepancy could not be explained yet.
a different wind turbine [17]. For k  ε RNG neither the values nor
the optimal tip-speed-ratio are predicted correctly.
In case of the SpalarteAllmaras and k  u SST models the ten- 6. Conclusion
dency is completely different. At low tip-speed-ratios the predic-
tion is rather accurate; but, at higher l-values the differences In the present study URANS-based 2D computations of H-Dar-
increase. It is interesting to note, that the difference is again cubic in rieus wind turbines were analyzed and compared. Deriving a
nature (fl3). The same tendency can be found, e.g., in Ref. [12]. proper URANS tool is of considerable importance, since such an
When using a correction factor of DCpcorr ¼ 0:0021l3 , the resulting approach would open the door to performance optimization, as
corresponding simulations are relatively fast. For this purpose, the
curve (k  u SST corr.) shows a very good agreement for the whole
needed requirements in terms of turbulence modeling, resolutions
domain. However, estimating the strut losses analytically
in space and time, etc., must first be clarified.
(CD0 ¼ 0.01288), they can account only for about 30% of these cubic
For this purpose mesh independency, temporal discretization,
losses. It remains unclear yet where the further difference finds its
the effect of turbulence intensity and the impact of different tur-
explanation.
bulence models have been checked. After identifying an optimal
value for these parameters, an extensive validation campaign was
5.4. Validation 4 performed by comparison with independent experimental mea-
surements. Experimentally measured characteristic curves were
The 4th and last validation was based on the work of Castelli compared with the results of different turbulence models using
et al. [17]. The authors have numerically and experimentally CFD computations with StarCCMþ and Fluent. The results with the
analyzed a three-bladed Darrieus-rotor with D ¼ 1030 mm diam- different software showed a very good agreement. Beside an own
eter and c ¼ 85.8 mm camber length using a NACA0021 airfoil. The experimental campaign, comparisons with three additional rotors
rotor had a small aspect ratio, with only H ¼ 1456.4 mm height. For from the literature have been considered.
the location of the spoke-blade connection 0.5 c was used [9]. Finally, two turbulence models were identified that lead to the
The blunt edge (0.38 mm) of the airfoil was disregarded. In the best systematic agreement and could therefore be used to obtain
experiment, 9 m/s wind speed was tested in a wind tunnel with with sufficient accuracy the characteristic curves of the turbine:
4000  3840 mm cross-section. The blockage was not considered in
the simulation. About 840 points were used to resolve the blade, 1. The k  ε Realizable model was able to predict correctly the
the shaft was disregarded. Total mesh size was around 330,000 location of the optimal tip-speed-ratio in all four cases, with an
elements, ANSYS Fluent was used for the simulations. The results almost constant offset of Cp compared to the experimental
can be seen in Fig. 12. measurements.
In their study, Castelli et al. have found, that k  u SST model is 2. The k  u SST model with a cubic correction delivers an excel-
the most appropriate model in 3D, while k  ε Realizable model is lent prediction for the first three cases. It remains an open
the most accurate in 2D. One can see, that the present study further question, why the last configuration behaves differently and if
confirms this observation. Indeed, k  ε Realizable shows an this observation can be generalized.
690 czy et al. / Energy 90 (2015) 680e690
L. Daro

Further 3D simulations with much higher computational efforts [25] Glauert H. Airplane propellers. In: Aerodynamic theory. Berlin Heidelberg:
Springer; 1935. p. 169e360.
will be required to evaluate the nature of wing-tip and
[26] Templin RJ. Aerodynamic performance theory for the NRC vertical-axis wind
strutejunction losses, and decide finally which model should be turbine. LTR-LA-160, Tech. rep. National Research Council of Canada; June
used. 1974.
[27] Paraschivoiu I, Delclaux F. Double multiple streamtube model with recent
improvements (for predicting aerodynamic loads and performance of Dar-
Acknowledgement rieus vertical axis wind turbines). J Energy 1983;7(3):250e5.
[28] Ferreira CJS, Bijl H, van Bussel G, van Kuik G. Simulating dynamic stall in a 2D
The authors gratefully acknowledge the financial support by the VAWT: modeling strategy, verification and validation with particle image
velocimetry data. J Phys Conf Ser 2007;75(1):012023.
BMBF (German Federal Ministry of Education and Research) [29] Winchester J, Quayle S. Torque ripple and variable blade force: a comparison
through the AiF-ZIM project KF2473103WZ3. of Darrieus and Gorlov-type turbines for tidal stream energy conversion. In:
Proceedings of the 8th European wave and tidal energy conference,
EWTEC2009, Uppsala, Sweden; 2009. p. 668e76.
References [30] ENERCON GmbH. ENERCON wind energy converters e product overview.
2010.
[1] European Commission ENERGY e Country factsheets V.1.3. [31] Bianchi FD, de Battista H, Mantz RJ. Wind turbine control systems. Springer;
[2] European Commission e Directorate-General for Energy, EU energy trends to 2007.
2030. [32] Sheldahl RE, Klimas P, Feltz L. Aerodynamic performance of a 5-metre-
[3] Fischer G, Kipouros T, Savill A. Multi-objective optimisation of horizontal axis diameter Darrieus turbine with extruded aluminium NACA-0015 blades,
wind turbine structure and energy production using aerofoil and blade SAND80-0179. Tech. rep. Sandia National Laboratories; March 1980.
properties as design variables. Renew Energy 2014;62:506e15. [33] Fiedler AJ, Tullis S. Blade offset and pitch effects on a high solidity vertical axis
[4] Mohamed M. Performance investigation of H-rotor Darrieus turbine with new wind turbine. Wind Eng 2009;33(3):237e46.
airfoil shapes. Energy 2012;47(1):522e30. [34] Mohamed M. Impacts of solidity and hybrid system in small wind turbines
[5] Ferreira S, van Bussel G, Scarano F, van Kuik G. 2D PIV visualization of dynamic performance. Energy 2013;57:495e504.
stall on a vertical axis wind turbine. In: Collection of technical papers e 45th [35] Maître T, Amet E, Pellone C. Modeling of the flow in a Darrieus water turbine:
AIAA aerospace sciences meeting, vol. 23; 2007. wall grid refinement analysis and comparison with experiments. Renew En-
[6] Bianchini A, Ferrara G, Ferrari L, Magnani S. An improved model for the per- ergy 2013;51:497e512.
formance estimation of an H-Darrieus wind turbine in skewed flow. Wind Eng [36] Paraschivoiu I. Wind turbine design: with emphasis on Darrieus concept.
2012;36(6):667e86. 
Ecole Polytechnique de Montre al; 2002.
[7] Nobile R, Barlow MVJ, Mewburn-Crook A. Dynamic stall for a vertical axis [37] Goude A, Lundin S, Leijon M. A parameter study of the influence of struts on
wind turbine in a two-dimensional study. In: World Renewable energy the performance of a vertical-axis marine current turbine. In: Proceedings of
congress, Linko €ping, Sweden; 2011. the 8th European wave and tidal energy conference, EWTEC2009, Uppsala,
[8] Bedon G, Castelli MR, Benini E. Optimization of a Darrieus vertical-axis wind Sweden; 2009. p. 477e83.
turbine using blade element e momentum theory and evolutionary algo- [38] Amet E. Simulation nume rique d'une hydrolienne a  axe vertical de type
rithm. Renew Energy 2013;59:184e92. Darrieus [Ph.D. thesis]. France, Grenoble: Polytechnical National Institute;
[9] Castelli MR, Englaro A, Benini E. The Darrieus wind turbine: proposal for a 2009.
new performance prediction model based on CFD. Energy 2011;36(8): [39] Qin N, Howell R, Durrani N, Hamada K, Smith T. Unsteady flow simulation and
4919e34. dynamic stall behaviour of vertical axis wind turbine blades. Wind Eng
[10] Gosselin R, Dumas G, Boudreau M. Parametric study of H-Darrieus vertical- 2011;35(4):511e28.
axis turbines using uRANS simulations (Paper CFDSC-2013 178). In: 21st [40] Daro czy L, Janiga G, The venin D. Systematic analysis of the heat exchanger
annual conference of the CFD Society of Canada, Sherbrooke, Canada; 2013. arrangement problem using multi-objective genetic optimization. Energy
[11] D'Alessandro V, Montelpare S, Ricci R, Secchiaroli A. Unsteady aerodynamics 2014;65:364e73.
of a Savonius wind rotor: a new computational approach for the simulation of [41] Daro czy L, Mohamed M, Janiga G, The venin D. Analysis of the effect of a
energy performance. Energy 2010;35(8):3349e63. slotted flap mechanism on the performance of an H-Darrieus turbine
[12] Almohammadi K, Ingham D, Ma L, Pourkashan M. Computational fluid dy- using CFD (GT2014-25250). In: ASME turbo expo conference, Düsseldorf;
namics (CFD) mesh independency techniques for a straight blade vertical axis 2014.
wind turbine. Energy 2013;58:483e93. [42] Trivellato F, Castelli MR. On the CouranteFriedrichseLewy criterion of
[13] Mukinovi c M, Brenner G, Rahimi A. Analysis of vertical axis wind turbines. In: rotating grids in 2D vertical-axis wind turbine analysis. Renew Energy
New results in numerical and experimental fluid mechanics VII e contribu- 2014;62:53e62.
tions to the 16th STAB/DGLR symposium Aachen, Germany; 2010. p. 587e94. [43] Spalart PR, Rumsey CL. Effective inflow conditions for turbulence models in
[14] Lanzafame R, Mauro S, Messina M. 2D CFD modeling of H-Darrieus wind aerodynamic calculations. AIAA J 2007;45(10):2544e53.
turbines using a transition turbulence model. Energy Procedia 2014;45: [44] DIN Deutsches Institut für Normung e.V. und VDE Verband der Elektrotechnik
131e40. Elektronik Informationstechnik e. V.. Design requirements for small wind
[15] Jericha H, Go €ttlich E, Selic T, Sanz W. Novel vertical-axis wind turbine with turbines, DIN EN 61400-2 (VDE 0127-2):2007-2. Tech. rep.. 2007.
articulated blading (GT2012-68969). In: ASME turbo expo conference, [45] ANSYS Inc. Ansys fluent 14.0: users guide. 2011. Canonsburg, PA.
Copenhagen; 2012. [46] Kelecy FJ. Coupling momentum and continuity increases CFD robustness.
[16] Campobasso MS, Gigante F, Drofelnik J. Turbulent unsteady flow analysis of ANSYS Advant II 2008;(2):49e51.
horizontal axis wind turbine airfoil aerodynamics based on the harmonic [47] CD-adapco. USER guide Star-CCMþ version 9.06. CD-adapco; 2014.
balance Reynolds-averaged NaviereStokes equations (GT2014-25559). In: [48] Lain S, Osorio C. Simulation and evaluation of a straight-bladed Darrieus-type
ASME turbo expo conference, Düsseldorf; 2014. cross flow marine turbine. J Sci Ind Res 2010;69(12):906e12.
[17] Castelli MR, Ardizzon G, Battisti L, Benini E, Pavesi G. Modeling strategy and [49] Balduzzi F, Bianchini A, Maleci R, Ferrara G, Ferrari L. Blade design criteria to
numerical validation for a Darrieus vertical axis micro-wind turbine compensate the flow curvature effects in H-darrieus wind turbines (GT2014-
(IMECE2010-39548). In: ASME 2010 international mechanical engineering 26389). In: Proceedings of the ASME turbo expo; 2014.
congress and exposition, vol. 7; 2010. p. 409e18. [50] Şahin AD. Progress and recent trends in wind energy. Prog Energy Combust
[18] Mohamed M. Aero-acoustics noise evaluation of H-rotor Darrieus wind tur- Sci 2004;30(5):501e43.
bines. Energy 2014;65:596e604. [51] Kooiman S, Tullis S. Response of a vertical Axis wind turbine to time. Varying
[19] Bianchini A, Ferrara G, Ferrari L. Design guidelines for H-Darrieus wind tur- wind conditions found within the urban environment. Wind Eng 2010;34(4):
bines: optimization of the annual energy yield. Energy Convers Manag 389e401.
2015;89:690e707. [52] Langtry RB. A correlation-based transition model using local variables for
[20] Capuzzi M, Pirrera A, Weaver P. A novel adaptive blade concept for large-scale unstructured parallelized CFD codes [Ph.D. thesis]. Universitaet Stuttgart;
wind turbines. Part I: aeroelastic behaviour. Energy 2014;73:15e24. 2006.
[21] Capuzzi M, Pirrera A, Weaver P. A novel adaptive blade concept for large-scale [53] Wall A, Penna P. The role of NRC wind tunnels in the study of the aerodynamic
wind turbines. Part II: structural design and power performance. Energy performance of small wind turbines. In: European wind energy conference &
2014;73:25e32. exhibition; 2010.
[22] Gao X, Yang H, Lu L. Investigation into the optimal wind turbine layout pat- [54] Bravo R, Tullis S, Ziada S. Performance testing of a small vertical-axis wind
terns for a Hong Kong offshore wind farm. Energy 2014;73:430e42. turbine. In: Proceedings of the 21st Canadian Congress of applied mechanics
[23] Liu Z, Wang X, Kang S. Stochastic performance evaluation of horizontal axis (CANCAM07), Toronto, Canada; 2007. p. 470e1.
wind turbine blades using non-deterministic CFD simulations. Energy [55] Kjellin J, Bülow F, Eriksson S, Deglaire P, Leijon M, Bernhoff H. Power coeffi-
2014;73:126e36. cient measurement on a 12 kW straight bladed vertical axis wind turbine.
[24] Go €
ۍmen T, Ozerdem B. Airfoil optimization for noise emission problem and Renew Energy 2011;36(11):3050e3.
aerodynamic performance criterion on small scale wind turbines. Energy
2012;46(1):62e71.

You might also like