Вы находитесь на странице: 1из 13

Comparison of different numerical approaches to the study of the H-Darrieus

turbines start-up
A. Rossetti
*
, G. Pavesi
University of Padua, Department of Mechanical Engineering, Via Venezia 1, 35131 Padova (PD), Italy
a r t i c l e i n f o
Article history:
Received 2 December 2011
Accepted 7 June 2012
Available online 16 July 2012
Keywords:
VAWT
Darrieus
Self-start
3D effects
CFD
BEM
a b s t r a c t
Self-start capability is an important feature of wind turbines. It allows to obtain simpler and cheaper
turbines not actively controlled. Different approaches to describe the self-start of an H-blade Darrieus
rotor are presented and compared in the present work. The Blade Element Momentum (BEM) approach
was compared with two and three-dimensional CFD simulations. The tipespeed ratio versus power
coefcient curves and the evolution of the trust forces over a blade revolution highlighted the limits and
the strengths of each approach.
The BEM model showed remarkable limits to describe to describe the self-start behaviour of the tested
geometry. The principal limits of the BEM approach can be ascribed to the absence of well documented
aerofoil databases for low Reynolds number and the inadequate modelling of dynamics effects. The 2D
simulation allowed to highlight the unsteady features of the ow elds, and the presence of a complex
vortices pattern which interact with the blade. Furthermore the comparison between 2D and 3D data
demonstrated the importance of 3D effects such as secondary ows and tip effects. These effects were
proved to have a positive effect on start-up, increasing the torque characteristic for tipespeed ratio of 1.
The start-up capability of H-Darrieus appears to be inuenced by many different factors, which include
secondary ows, three-dimensional aerodynamic effects and the nite aspect-ratio of the blades.
2012 Elsevier Ltd. All rights reserved.
1. Introduction
The increasing energy cost and the presence of support policy
towards renewable energy are feeding the interest to develop small
power wind turbines suitable for urban and suburban applications.
The soft sound emissions, due to the lower optimum tipespeed
ratios compared with the three blades HAWT, the insensitivity to
the wind direction, and the lack of need of passive or active yawl
control are attractive aspects. Moreover, the easy accessibility of
the electrical and mechanical part of the VAWT decreases the
maintenance costs of the turbine. Notwithstanding, the higher
performances of the HAWT and the more complex structural
design of the VAWT, the Darrieus turbines presents many
advantages when small applications in urban agglomerations are
considered.
The ability of self-start is an important quality/property for
small turbines, to obtain simpler and cheaper actively controlled
turbines. The need of the turbine assisted launch causes energy
expense, could impact the turbine overall production and increases
the control system complexity [1]. The self-start capability of xed
pitch Darrieus turbine is not completely documented yet and
conicting data and conclusions can be found. Baker [2] identied
the presence of a negative torque band, on a l range from 0.5 to 2,
depending on the blade geometry consistently with the experi-
mental data reported in literature [3]. Baker suggested the use of
inclined or wrapped blades and a series of desirable aerofoil section
feature to aid self-starting. Other authors [4,5] have pointed out the
effect of the blade Reynolds number on the self-start behaviour,
demonstrating the positive effects of high Re numbers. This has
encouraged the construction of turbines with long chord blades
that is with high solidity. The capability of high solidity Darrieus to
self-start is consistent with some producers claims [6], but the
excessive increase in solidity leads to the decrease of the maximum
power coefcient. For the Troposkien blade turbine, the optimum
solidity was calculated by Strickland to be about 0.3, on the base of
Blade Element Momentum analyses [7] and conrmed by wind
tunnel experimental tests conducted by Sandia National Labora-
tories [8]. Recent papers demonstrated the capability of self-
starting of an H-blade, low solidity Darrieus in open section wind
tunnel at low Re [9,10]. The papers highlighted the difculty of
predicting self-start using analytical models, due to the uncertainly
of aerodynamic database at low Re and to the presence of complex
* Corresponding author. Tel.: 39 049 827 7474; fax: 39 049 827 6785.
E-mail address: antonio.rossetti@unipd.it (A. Rossetti).
Contents lists available at SciVerse ScienceDirect
Renewable Energy
j ournal homepage: www. el sevi er. com/ l ocat e/ renene
0960-1481/$ e see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.renene.2012.06.025
Renewable Energy 50 (2013) 7e19
and unsteady phenomena at low tipespeed ratios [1,11]. In the
same way, the studies on the reduced aspect-ratio effects carried
out contradictory conclusions. Musgrove and Mays [5] suggested
a positive effect whereas Kirke [12] presented opposite argumen-
tation based on theoretical considerations. Despite the apparent
discrepancy between the cited papers, all concur to suggest that the
xed blade Darrieus could self-start but only under some circum-
stances and with suitable geometry.
In this paper, different approaches to describe the self-start of
an H-blade Darrieus rotor are presented and compared. The
start-up characteristics of the turbine are computed using both
a Blade Element Momentum (BEM) model and 2D and 3D CFD
analyses. The 3D analyses are compared to 2D approach to
identify the effects of nite aspect-ratio on start-up process, to
highlight the strong 3D effects on start-up mechanics and the
inadequacy of analytical models to predict self-start worsened
by the absence of accurate aerodynamic database for low Rey-
nolds numbers.
2. Self-starting
Two different approaches can be distinguished in the self-
starting denition. The rst one denes the self-starting deni-
tion on the basis of the output power delivered by the turbine.
Ebert and Wood considered the starting process completed when
usable power extraction starts up [1]. In a similar way, Kirke
dened the self-starting feature as the capability to accelerate from
not running to the point at which a useful output is produced [12].
The second approach denes the starting as a function of the
tipespeed ratio (TSR) l. Lunt identied the self-starting feature as
the turbine capability to reach a TSR greater than 1 [13]. This
denition is based on the assumption that to reach TSR values
greater than 1 the turbine has to overcome the drag-driven motion
and enter into the lift-driven motion. These denitions are limited.
The rst approach suffers from the absence of a precise threshold
for the output power. Moreover, the output power depends on the
wind velocity. On the basis of the given denition, every machine
that allows a drag-driven motion can be dened as self-starting
under high wind. As opposed to the second denition, which
assures the presence of a lift-driven behaviour, but does not assure
any energy production.
For the present work a restrictive formulation was considered.
The self-start capability of the turbine was dened as the ability to
reach the escape velocity, starting from rest, under constant wind
speed when no power extraction occurs in the start-up process.
This denition is compatible with the previous denitions. The
request to accelerate the turbine up to the escape velocity assures
the presence of a net power production because the turbine over-
steps the TSR with the maximum Cp and consequently the turbine
gets to l greater than 1. Moreover the dened start-up capability is
explicitly related to the wind velocity and to the inuence of the
Reynolds number on the start-up [4,5,14] and consistently allows to
dene the minimum wind velocity at which the start-up takes
place.
3. Reference turbine
The reference turbine is an H-Darrieus 3-bladed rotor [10] with
the main geometrical characteristics listed in Table 1. According to
Hill, this Darrieus geometry was proved to be able to self-start with
a constant wind of 6 [m/s] when no electrical load is applied to the
rotor. The turbine tipespeed ratio verses time from rest to the
equilibrium velocity is reported in Fig. 1 [10].
The velocity trend shows a sequential behaviour. After quite
a linear acceleration, due to the drag-driven motion of the turbine,
the rotor speed shows a relatively long plateau period before it is
accelerate by the lift-driven motion and reaches the equilibrium
condition. The small inertia of the rotor and the persistence in the
plateau zone allows to exclude the inertial effects to help overcome
the dead zone. The experimental data are compatible both with
a slightly positive torque for l in the range between 1.2 and 1.3 and
with the presence of a negative torque band in the same range.
In eld test, the rise to overcome the dead zone could be
reasonably explained by wind instability according to Baker [2], and
a temporary drop in wind intensity results in an increase of l to
a value outside the negative torque region. Despite signicant
uctuation observed in eld test, the wind tunnel test conditions
are carried out with marginal wind variation. According to this
observation, only a small dead band in the range of l 1.2e1.3 could
match with the experimental data.
4. BEM-software
The Blade Element Momentum (BEM) model, originally pre-
sented by Templin [15] and developed by Strickland [7], adapted
the actuator disk approach to predict the performance of a Darrieus
turbine. The model was improved over the years by many authors
[16e19], to take into account dynamic stall and secondary effects.
The BEM-software used for this paper was implemented in Matlab,
according to the Double-Disk Multiple stream tubes presented by
Paraschivoiu. Moreover, the basic model was improved using
appropriate sub models to copy dynamic behaviour and secondary
effects:
- Momentum equation: The original momentum equation
resulting from the axial equilibrium of the turbine is inappro-
priate for turbulent wake. The empirical momentum equation
proposed by Glauert [17,20] was implemented to overcome the
original limits.
Table 1
Reference geometry.
Aerofoil NACA 0018
Chord 0.083 [m]
Span 0.600 [m]
Diameter 0.750 [m]
Moment of inertia 0.018 [kg m
2
]
Turbine Ar. 0.80 []
Blade Ar. 7.22 []
Fig. 1. Start-up characteristic of the reference rotor under constant wind of 6 [m/s].
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 8
- Dynamic stall model: The dynamic stall was modelled according
to Mass and Berg [21].
- Stream tube expansion model was implemented to improve the
software accuracy [18,19].
The aerodynamic proles were characterized for different Rey-
nolds numbers in the range from 40,000 to 1e7 and for an angle of
attach from 0

to 180

using the data published by Sheldahl and


Klimas [22].
The BEMmodel was validated comparing the model output with
the experimental data published by the Sandia National Laborato-
ries for a 5 m rotor [23]. The geometrical features of the test
geometry are reported in Table 2, while the results of the
comparison are reported in Fig. 2.
5. CFD set-up
CFD commercial code Ansys CFX 12 [24] was used to analyse
torque and power variations versus time at low tipespeed ratio.
Both 2D and 3D unsteady analyses were carried out on the refer-
ence geometry. Two different domains were dened for the 2D and
the 3D models, corresponding to the turbine rotor and the outer
stationary domain. For the 3D domain only half of the turbine span
was considered taking advantage of the symmetry of the system.
The 3D domains and the main dimensions of the volume consid-
ered are presented in Fig. 3. The 2D model was obtained extracting
the symmetry surface from the 3D domain of Fig. 3.
Unstructured mesh was employed for both the stationary and
rotating domains. Grid renements were provided near the
domains interface as well as in the wake region downstream the
turbine and near the rotor blades. A structured O-grid of hexahe-
dral elements was adopted to improve the near wall accuracy. The
3D mesh was obtained extruding the 2D mesh along the direction
orthogonal to the symmetry plane and renements were adopted
near the blade tip along the span direction.
The nal grid size was dened after grid sensitivity analyses on
the 2D case. The meshes were obtained scaling the average
dimension of elements except for the maximum height on blade,
which was kept constant. The height of the rst near wall element
was selected to obtain y lower 5 for all the working conditions.
The sensitivity study was also extended to the temporal dis-
cretization, comparing results obtained using different time steps.
The results of the sensitivity are reported for l 1 in Fig. 4. Only
small changes in the power coefcient were observed when more
than 100.000 nodes per layer and a time step smaller than 2

/u
were employed. Results of the sensitivity analyses encouraged the
use of the nest mesh for 2D simulation, while 100.000 nodes per
layer were assumed to be a good compromise between accuracy
and computational time for the 3D case.
The mesh structure and renements are presented in Fig. 5. The
grid size and elements dimensions are reported in Tables 3 and 4
for the different domains.
The inlet condition was set on the upwind face of the stator
domain, imposing a constant wind prole equal to 6 [m/s] with
stochastic turbulence of 5% of the free stream velocity. The static
atmospheric pressure was imposed on the remaining surfaces of
the model. Symmetry was employed on both the lower and the
upper faces for the 2D simulation, while for the 3D analyses it was
substituted on the upper part with a static pressure condition.
Table 2
Benchmark geometry for the BEM-software.
Aerofoil NACA 0015
Chord 0.152 [m]
Span 5.10 [m]
Diameter 5.00 [m]
Fig. 2. Comparison between the eld data and BEM model for the Sandia 5 m turbine
for constant speed of 150 [rpm].
Fig. 3. 3D domain dimension.
Fig. 4. Sensitivity study results.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 9
Time discretization was dened according to a sensitivity study
presented in Fig. 4. A time step dt corresponding to a 2

[deg] of the
rotor revolution was employed for all the analyses. The analyses
were carried out for a total time corresponding to at least 8
complete revolutions of the rotor, to gain the independence of the
solution from the initial condition. Convergence to the periodic
behaviour of the system was monitored comparing the mean
power coefcient between two subsequent revolutions. For all the
considered cases, this parameter was kept below the 3% of the
mean Cp.
The air was modelled as constant propriety gas at 25 [

C]. A
specic attention was given to the turbulence model. The SAS
model was preferred to more classical turbulence models. This
model results in a compromise between LES and RANS formulation,
providing LES-like behaviour in unsteady regions of the ow eld
and standard RANS capabilities in stable ow regions [25]. In order
to further improve the near wall ow model accuracy, the
laminareturbulent transition were modelled employing the two
equation model developed by Langtry and Menter [26,27].
6. Results
6.1. BEM model
The torque characteristic of the turbine calculated using the
BEM model is plotted in Fig. 6.
Fig. 5. Mesh structure; a) 2D analyses computational domain; b) 2D rotor detail; c) 2D O-grid structure detail; d) 3D mesh on blade tip.
Table 3
elements dimension.
2D case 3D case
Stator
Maximum element 80 [mm] 80 [mm]
Elements on interface and wake 8 [mm] 12 [mm]
Rotor
Maximum element 8 [mm] 12 [mm]
Maximum element on blade 100 [mm] 150 [mm]
Maximum height on blade 40 [mm] 40 [mm]
Expansion ratio 1.1 1.15
Table 4
Mesh size per layer.
Stator
2D model 29,788 nodes
3D model 1,883,780 nodes
Rotor
2D model 434,508 nodes
3D model 6,986,796 nodes
Fig. 6. Tipespeed ratio curve predicted by the BEM model.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 10
The results suggest that the model drastically overestimated the
dead band of the turbine. The negative torque band extends from
l 0.2 to l 3. Similar results were obtained by Hill [10], which
reports two distinct negative torque bands between 1.2 and 3 for
a 0012 prole.
Fig. 7. Lift and trust forces calculated on the overall blade revolution; (a) l 1; (b) l 2.
Fig. 8. Lift characteristic of the NACA 0018 prole as described by different authors; a) lift coefcient versus incidence curves; b) characteristic points of the curves as function of the
Reynolds number.
Fig. 9. Modied aerodynamic database according to the ViternaeCorrigan correlation.
Fig. 10. Tipespeed ratio versus torque curve as predicted by the BEM model using
unmodied and modied aerodynamic data.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 11
The torque coefcient and the lift coefcient versus the blade
position over the overall revolution are shown in Fig. 7. The nega-
tive torque is caused by the sudden decrease of the lift in the after-
stall region [2], till the increase of the tipespeed ratio decreases the
angle of attack below the stall incidence (l > 3.5).
The drop of the lift in the after-stall region as reported by
Sheldahl and Klimas is, however, in contradiction with other
databases available in literature. Fig. 8 compares the lift coefcient
characteristic of the NACA0018 prole for Reynolds number in the
interval 80.000e150.000.
Sheldahl and Klimas data report both the lower maximum lift
and the higher lift loss in the after-stall region. The after-stall lift
drop is in accordance with recent data published by Timmer [28],
but the data appear to overestimate the maximum lift. The data
proposed by Jacobs and Sherman [29] present values of maximum
lift intermediate between the other two series of data. On the other
hand, the after-stall characteristic is strongly different and it shows
a smoother transition to the full stalled zone.
Until Timmer, and Jacobs and Shermans data could not be used
effectively for the present work due to the absence of low Reynolds
series [Re < 1e5]. To evaluate the inuence of the post-stall region
on the BEM results, the analyses were repeated using a database
derived from the Sheldahl and Klimas data, assuming the post-stall
Fig. 12. Start-up characteristics expected using tipespeed ratio versus torque curve
of Fig. 10.
Fig. 13. Tipespeed ratio versus torque curve. 2D CFD and BEM results. Fig. 14. Trust force comparison between CFD and BEM for l 1.
Fig. 11. Lift and Trust forces calculated over a whole blade revolution using modied aerodynamic data; (a) l 1; (b) l 2.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 12
Fig. 15. Blade loading charts for different blade positions.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 13
characteristic is in accordance with the model proposed by Viterna
and Corrigan [30]. The blend between the original data and the
post-stall model were arranged to obtain a post-stall characteristic
similar to the one observed in Jacobs and Sherman data. The
aerodynamic data obtained are compared to the original data in
Fig. 9.The torque curve obtained with the modied database is
reported in Fig. 10.
The local modication of the aerodynamic data had strong
impact on the torque curve. The torque characteristic obtained with
the modied data presents always positive torque except for l 1
Fig. 16. Adimensional vorticity development for step of 20 [

] of the rotor displacement.


Fig. 17. Flow streamlines in the 2D model.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 14
(M 3e4 Nm). The torque versus the angular blade position
(Fig. 11) still presents negative values, but the after-stall decrease of
the lift has reduced impact on the overall torque.
The expected turbine start-up characteristic was computed
by the torque curves obtained from the BEM model. The inte-
gration of the Newton law was obtained using explicit
RungeeKutta method of 2nd and 3rd order respectively. The
wind was assumed to be characterized by an average velocity of
6 [m/s], with the integration being carried out by rst consid-
ering a constant wind value and subsequently a wind perturbed
by a stochastic turbulence of 5 and 10%. The results are pre-
sented in the Fig. 12. The characteristics predicted using Shel-
dahl & Klimas data are clearly unable to predict the self-start of
the turbine for both constant and turbulent wind. On the
contrary, the torque curves obtained with the modied data
show different responses for constant and turbulent wind. The
small negative band (Dl y 0.1) prevents the acceleration of the
rotor to exceed speeds greater than l 0.97 when a constant
wind is considered. Otherwise, the superimposition of uctu-
ating terms to the reference wind velocity allows to overcome
the dead band and to accelerate the rotor. The start-up charac-
teristics show only some similarities with the experimental data
reported in Fig. 1, in the plateau and in the sudden acceleration
when lift regime is reached. Both the original and modied
data fail the comparison with experimental data in the linear
regime overestimating the torque generated in the drag driven
motion.
6.2. 2D CFD analyses
The torque prediction obtained from the 2D CFD analyses is
shown in Fig. 13. The torque characteristic appears to be in good
agreement with the original BEMmodel for l 1. For l >1, the CFD
results show a sudden change in the slope of the torque curve
towards positive values. In this case, the dead band is limited
approximately between 0.2 and 1.5. The BEM and CFD predictions
for trust and lift force for l 1 are compared in Fig. 14. Despite the
good agreement of the BEMand CFD torque mean values, the detail
comparison of the trust force highlights important differences.
The blade load charts at the different angular position of Fig. 15
can explain the differences between the two models and the
aerodynamic behaviour of the proles in CFD and the development
of the vortices on the rotor reported in Fig. 16.
Starting from the q 0, both the analyses describe the increase
in trust due to the increase of the blade lift (points a and a
0
in
Fig. 14) followed by the rapid decrease due to development of the
stall (points b and b
0
in Fig. 14). Even though the two curves
demonstrate qualitatively similar trends the CFD characteristic in
Fig. 14 is translated toward angular positions of about 10

higher.
This is explainable considering the velocity component orthogonal
to the undisturbed wind due to the vorticity induced by the wind
turbine and the expansion of the ow streamlines (Fig. 17) which
are computed in CFD model while they are not accounted accu-
rately in the BEM approach. The deection of the undisturbed
wind reduces the effective incidence on the blades for small q
values and causes the delay of the trust ramp in CFD model.
Consistently, the blade loading chart highlights the presence of
a small negative incidence when q 0

(Fig. 15). Differences


between the BEM and CFD for q 0 are also caused by the
curvature of the relative ow, which becomes manifestly visible
for small incidences. Consistently, the loading characteristic shows
the inversion of the load direction between the blade nose and the
trailing edge.
Moreover, the CFD data show a remarkable delay in the blade
stall that starts after about 40 degrees of rotation and 20

later than
the BEM model (Fig. 16). This increases the forces in a
0
and
decreases the trust in b
0
where stall occurs. The different blade stall
characteristic is directly related to the dynamic effects, which are
dominant at low tipespeed ratios [11,30] and are not adequately
considered in the BEM.
The after-stall region up to q 180

is characterized in the BEM


model by a continuous decrease in trust consistent with the
decrease of the relative velocity, while the CFD data show a more
complex trend. The partial ow reattachment in the suction side
near the leading edge (Fig. 15 q 80) causes a second increase of lift
force. This leads to a second peak of the trust c
0
at about 80 degrees
of rotation. Moreover, the stall vortices on the rear inner side of the
prole between 90 and 180

leads to the negative trust region d


0
.
On the second half of the rotation the blades encounter inci-
dences similar to the ones on the front side of the rotor, with
inversed order. The BEM description of the trust curve is then quite
symmetrical in respect of the position q 180. The negative peak e
corresponds to the peak b, and is associated to the full stall
condition for small incidence angels. The positive peak f corre-
sponds to the peak a, and to the rst ow reattachment to the
proles. Peaks e and f present smaller values than a and b,
consistently to the velocity reduction due to the use of the double
Fig. 18. Trust force on a blade revolution for 2D and 3D CFD analyses.
Fig. 19. Trust force on a blade revolution for 2D and the symmetric blade section of the
3D model.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 15
disk BEM model. The dynamic stall model also inuences the trust.
The decreasing values of incidence delays the reattachment of the
ow and consistently increases the negative peak value e and
postpones and decreases the peak value f.
The accordance between the BEM and CFD is acceptable for the
rst part of the downwind side. The main differences in the last part
could be related, as for small incidences in the upwind side, to the
dynamic effects. In the CFD analyses, the ow is not reattached to
the blade (Fig. 16), leading to a at zone of negative torque instead
of the peaks couple e and f.
6.3. 3D CFD analyses
A 3Danalysis of the considered rotor was carried out for l 1, to
highlight the three-dimensional effects on the start-up of the
Darrieus rotors. The mean torque obtained for the 3D analysis was
Fig. 20. Adimensional vorticity Uc=lV development for step of 20 [

] of the rotor displacement.


Fig. 21. 3D view of the dimensional vorticity (Uc=lV) for q 120 [

]; a) positive vorticity; b) negative vorticity.


A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 16
T
3D
0.0069 [Nm], which is signicantly higher than the one
obtained by the 2D analysis of T
2D
0.1633 [Nm]. Furthermore,
although negative, this value could be in agreement with the self-
start of the turbine as the result of the wind uctuations, as
previously discussed for BEM results. The comparison between 2D
and 3D trust curves are presented in Fig. 18. This comparison allows
to demonstrate the relevance of three-dimensional effects for low
tipespeed ratio and for the start-up characteristics.
To highlight the nite aspect-ratio impact and the 3D ow
effects on the start-up characteristic, the blades were divided into 6
regions of equal height H/12 0.05 [m] from the blade tip (S1) to
the symmetry plane (S6). The forces calculated on each blade
section f
i
were post-processed and scaled to the overall blade
length to compare on the same scale the torque of either the 3D or
2D cases:
F
i

H
H=12
f
i
(1)
The obtained trusts on the section near to the symmetry plane
are compared to the 2Dresults in Fig. 19. The comparison highlights
that the 3D effects do not inuence only the external parts of the
blade, but reaches the symmetry plane of the model. The four main
peaks of the curve from a to d are the more affected. In particular,
the rst peak a
00
is enforced, due to the increase of the dynamic
effects that delays the blade stall. Moreover, the trust in the after-
stall positions is signicantly reduced helping the increase of the
overall torque.
Both the reattachment peak c and the negative region d of the
2D model are effectively reduced in the 3D model. The cause can
be found comparing the vorticity map of the 2D model (Fig. 16)
with the corresponding map on the symmetry plane of the 3D
model showninFig. 20. Vortices appear reducedinintensity for all
the blade positions. Specically, the absence of a strong and stable
vortex near the blade trailing edge avoids the development of the
negative peak d
0
. The reduction of intensity of the vortex structure
appears to be related to the higher instability of the vortex in the
three-dimensional models, due to the presence of cross ow
components along the blade. The vorticites eld near the blade,
for q 120 is shown in Fig. 21. The vortex structures and the stalls
involve all the blade height and are highly three-dimensional.
Consequently the assumption of two-dimensional ow appears
to restrictive. The presence of a ow expansion in the axial
direction, neglected in 2D analyses, destabilizes the vortex
structure even in the symmetry section.
Fig. 22 shows the trust in the six sections on an entire blade
revolution, while histogram in Fig. 23 compares the mean trust of
the different sections. The intermediate sections between S3 and
the symmetry plane S6 (symmetry) present a similar tendency,
whereas the tip sections S1 and S2 show different trends with
a positive trust mean force. The positive mean torque is mainly due
to the absence of the post-stall negative peak b (see Fig. 22a). The
delay of the stall in the sections near the tip is caused by the
presence of a signicant cross ow from the symmetry plane
toward the blade tip, caused by both the streamline expansion and
the non-uniform blade loading between symmetry and tip. In
Fig. 24, the shears vectors on the external side of the blade are
plotted for different position of the blade in the range q 0

e
q 100

. The external side of the blade corresponds to the pres-


sure side for the upwind part of the rotor. Fig. 24 reports also the
colour map of the shear component along the turbine axis. Named
d, the angle between the wind velocity and the horizontal plane, the
presence of an angle d s 0 reduces the wind component on the
horizontal plane. Consistently, the local tipespeed ratio l* increases
according to:
l*
uR
Vsind
(2)
and the local incidence i* angle decreases:
i* atan

sin q
l* cos q

(3)
Fig. 22. Trust force on the six elements of the blade; a) sections S1 (tip) and S2; b) sections S3 and S4; c) section S5 and S6 (symmetry).
Fig. 23. Mean trust on a revolution for the six elements of the blade.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 17
The incidence reduction could be appreciated in Fig. 24, noticing
that for q 80

the stagnation point moves signicantly toward the


leading edges for sections S5 and S6. The incidence reduction due to
the secondary ow along the blade allows to avoid the stall in the
tip region, contributing to the increase of the overall torque.
The increase in the local tipespeed ratio l* near the tip section
appear to have positive effects on the start-up characteristic as long
as Cp(l*) Cp(l).
On the basis of these observations some hypotheses can be
outlined to explain the documented experimental start-up char-
acteristic [10]:
1) Linear regime: in this interval Cp(l) is greater than zero, due to
the drag driven ow. The high blade loading encourages
secondary ow, that increases the local l* in the blade tip
region. Cp(l*) is smaller than Cp(l), because the drag-driven
torque goes down quickly as l increases and assumes negative
values for l* y 1. Consistently, the tip effects slow down the
overall turbine acceleration, resulting in an almost constant
torque that is in the linear acceleration regime. During this
period the mean turbine tipespeed ratio is lower than the dead
band, but three-dimensional effects should be locally suf-
ciently strong to increase l* up to overcome the dead band;
2) Plateaux: in this region, the main part of the blade is charac-
terized by negative torques, because Cp(l) < 0. Nevertheless
the tip sections works with local l*, which is beyond the bead
band region, with a Cp(l*) > 0. As a result, the negative torque
caused by the central part of the blade is counterbalanced by
positive tip effects. The dead band is then reduced in ampli-
tude. If a negative band is still present, it could be overcome
thanks to the wind instability.
3) Lift and equilibrium: as the dead band goes over on the entire
blade, the Cp becomes positive and the turbine accelerates. As
the blade accelerates, the owincidence angle during the blade
revolution comes down as the blade loading, and the differ-
ences between l* and l. When the main part of the turbine
reaches the optimum working condition Cp(l) Cp
Max
, the tip
effects become dissipative and tip sections work with l*
slightly greater than l, and with Cp(l*) < Cp
Max
. In this region
the relative ow is characterized by high velocities and the tip
vortices, which were negligible for low l, than become the
reason of signicant losses, and reduce the expected Cp.
7. Conclusions
Different approaches to describe the self-start of an H-blade
Darrieus rotor are presented and compared in the present work.
The Blade Element Momentum approach was compared with two
and three-dimensional CFDsimulations. The tipespeed ratio versus
power coefcient curves and the evolution of the trust forces over
a blade revolution highlighted the limits and the strengths of each
approach.
Fig. 24. Vector plot of the shear stress on the pressure side for different blade position. Colour map refers to the horizontal component of the shear vector.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 18
The BEM model showed noteworthy limits to describe the self-
start behaviour of the tested geometry. This approach is strongly
relayed to the aerodynamic database given as input to the model
and suffers of the absence of well documented aerofoil character-
ization of aerodynamic proles for a very low Reynolds number.
The post-stall description zone, with incidences between 10 and 30
[

], was demonstrated to have relevant inuences on the dead band


zone of the rotor torque curve. Two different aerodynamic data-
bases gave opposite results, neglecting or demonstrating the
capability of the reference geometry to achieve self-start.
The 2D analysis highlighted the presence of a complex vortices
pattern, which interacted with the blades and strongly modied
the overall performance.
The CFD analyses proved the dynamic stall effected the ow
eld more so than the model implemented into the BEM-software
are able to predict. Furthermore 2D analyses allow to draw atten-
tion to the effects of the streamline divergence on the incidence
angle on the blades, and the effects of the relative ow curvature,
which are not accounted in the BEM models. 2D analyses described
a dead band for tipespeed ratio inside the range 0.2e1.5, neglecting
the capability of the reference geometry to achieve self-start.
3D analyses were conducted for tipespeed ratio equal to 1.
Results showed the great importance of 3D effects for small tipe-
speed ratio. When compared to the 2D results, three-dimensional
analyses showed a signicant reduction in the vortices intensity
inside the rotor. This is ascribed to the instability of vortices which
were subjects to secondary ows along the turbine axis. These ow
components were caused by the non-uniform blade loading, and
created a ow from the symmetry plane toward the blade tip.
Consistently the local incidence angle was reduced near the blade
tip and the stall was avoided. Both the reduction of the vortices
intensity and the tip effects increased the rotor expected mean
torque. The mean torque obtained was compatible with the self-
start of the considered geometry, assuming the presence of small
uctuation of the wind velocity.
The start-up capability of H-Darrieus appears to be the result of
many different causes, which include: the blade prole, the Rey-
nolds number, secondary ows and three-dimensional aero-
dynamic effects. Collected data suggest that turbine start-up could
take place when tip effects are sufcient to draw the turbine
outside the dead band. At the same time, they are dissipative for
tipespeed ratio less than 1, and should not be so pronounced to
avoid the acceleration of the turbine to tipespeed near the unity.
This delicate equilibrium could explain the presence of contradic-
tory experimental results.
References
[1] Ebert PR, Wood DH. Observations of the starting behaviour of a small
horizontal-axis wind turbine. Renewable Energy 1997;12(3):245e57.
[2] Baker JR. Features to aid or enable self starting of xed pitch low solidity
vertical axis wind turbines. Journal of Wind Engineering and Industrial
Aerodynamics 1983;15:369e80.
[3] Paraschivoiu I. Wind turbine design: with emphasis on Darrieus concept.
Montreal: Polythecnic International Press; 2002. 289e296.
[4] Watson G R. The self-starting capabilities of low-solidity xed pitch Darrieus
rotors. BWEA workshop; 1979 April 19e20, Craneld, UK, p. 32e39.
[5] Mays I, Musgrove PJ. Performance of the variable geometry vertical axis wind
turbine at high and low solidities. BWEA workshop; 1979 April 19e20,
Craneld, UK. p. 48e56.
[6] Ropatec. http://www.ropatec.com/index_n.php?pid01&lin3 [accessed
11.30.2011].
[7] Strickland JH. The Darrieus turbine: a performance prediction model using
multiple streamtubes. SAND75-0431. Albuquerque: Sandia National Labora-
tories; 1975.
[8] Blackwell BF, Sheldhl RE. Wind tunnel performance for the Darrieus turbine
whit NACA0012 blades. SAND76-0130. Albuquerque: Sandia National Labo-
ratories; 1976.
[9] Dominy R, Lunt P, Bickerdyke A, Dominy J. Selfstarting capability of a Darrieus
turbine. Proceedings of the institution of mechanical engineers, part A. Journal
of Power and Energy 2007;221:111e20.
[10] Hill N, Dominy J, Dominy R, Ingram G. Darrieus turbines: the physics of self-
starting. Proceedings of the institution of mechanical engineers; part A.
Journal of Power and Energy 2009;223:21e9.
[11] Paraschivoiu I, Allet A. Aerodynamic analyses of the darrieus wind turbines
including dynamic-stall effects. Journal of Propulsion and Power 1988;4:
472e7.
[12] Kirke BK. Evaluation of self-starting vertical axis wind turbines for stand-
alone applications. PhD thesis, School of Engineering, Grifth University,
Australia; 1998.
[13] Lunt PAV. An aerodynamic model for vertical axis wind turbine. MEmg
projects report. UK: School of Engineering, University of Durham; 2005.
[14] Kirke BK, Lazauskas L. Enhancing the performance of vertical axis wind
turbine using a simple variable pitch system. Wind Engineering 1992;15:
187e95.
[15] Templin RJ. Aerodynamic performance theory for the N RC vertical-axis wind
turbine. LTR-LA-160. National Research Council/National Aeronautical Estab-
lishment; 1974.
[16] Paraschivoiu I. Double-multiple streamtube model for Darrieus wind turbines.
Second DOE/NASA wind turbines dynamics workshop, NASA cp-2182, Clive-
land, Ohio; February 1981 19e25.
[17] Paraschivoiu I, Delclaux F. Double-multiple streamtube model with recent
improvements. AIAA Journal of Energy 1983;7:250e5.
[18] Paraschivoiu I, Delclaux F, Beguier C, Fraunie P. Aerodynamic analysis of the
Darrieus rotor including secondary effects. Journal of Energy 1983;7(5):
416e22.
[19] Paraschivoiu I, Fraunie P, Beguier C. Streamtube expansion effects on the
Darrieus wind turbine. AIAA Journal of Propulsion and Power 1985;1(2):
150e5.
[20] Glauert H. A general theory of the autogyro. ARCR R&M, No. 1111; 1926.
[21] Allet S, Halle I, Paraschivoiu I. Numerical simulation of dynamic stall around
an airfoil in Darrieus motion. Journal of Solar Energy Engineering 1999;121:
69e76.
[22] Sheldahl RE, Klimas PC. Aerodynamic characteristics of seven symmetrical
aerofoil sections through 180-degree angle of attack for use in aerodynamic
analysis of vertical axis wind turbines. SAND80-2114. Albuquerque: Sandia
National Laboratories; 1981.
[23] Sheldahl RE, Klimas PC, Feltz LV. Aerodynamic performance of a 5-metre-
diameter Darrieus turbine with extruded NACA-0015 blades. SAND80-0179.
Albuquerque: Sandia National Laboratories; 1980.
[24] Ansys CFX 12.0. Copyright Ansys Inc. http//www.ansys.com [accessed
11.30.2011].
[25] Egorov Y, Menter F. Development and application of SST-SAS turbulence
model in the DESIDER project. In. Second symposium on hybrid RANS-LES
methods, Corfu, Greece; 2007.
[26] Menter FR, Langtry RB, Likki SR, Suzen YB, Huang PG, Vlker S. A correlation
based transition model using local variables part 1emodel formulation.
ASME-GT2004-53452. Vienna, Austria: ASME TURBO EXPO; 2004.
[27] Menter FR, Langtry RB, Likki SR, Suzen YB, Huang PG, Vlker S. A correlation
based transition model using local variables part 2etest cases and industrial
applications. ASME-GT2004-53452. Vienna, Austria: ASME TURBO EXPO;
2004.
[28] Timmer WA. Two-dimensional low-Reynolds number wind tunnel results for
airfoil NACA 0018. Wind Engineering, 32(6), p. 525e537.
[29] Jacobs EN, Sherman A. Airfoil section characteristics as affected by changes of
the Reynolds number. NASA report, vol. 586; 1937.
[30] Viterna LA, Corrigan RD. Fixed pitch rotor performance of large horizontal axis
wind turbines. In. DOE/NASA workshop on large horizontal axis wind
turbines, Cleveland, Ohio; July 1981.
A. Rossetti, G. Pavesi / Renewable Energy 50 (2013) 7e19 19

Вам также может понравиться