Sie sind auf Seite 1von 5

Computational Fluid Dynamics JOURNAL vol.11 no.1 April 2002 (pp.

xxx-xxx)

DESIGN AND ANALYSIS OF WIND TURBINES USING


HELICOIDAL VORTEX MODEL

Jean-Jacques CHATTOT †

Abstract
The design and analysis of wind turbines is carried out using a helicoidal vortex model that
allows for accurate calculation of the induced velocities using the Biot-Savart law. The design
corresponds to the maximum power output for a given thrust and the distributions of circulation,
induced velocities, chord and twist of two- and three-bladed rotors are obtained. The analysis of
turbines at off-design conditions is based on the same helicoidal vortex model, however, the power
is not prescribed and an iteration is needed to insure that the vortex system is consistent with the
resulting power extracted from the air. 2-D data from experiments or viscous codes are used to
correct for viscous effects. Comparisons with published cases indicate that the method produces
useful results very efficiently.
Key Words: Please specify.

1 Introduction the optimization of propellers, where the details of the


helicoidal vortex structure, discretization and mini-
One of the key feature of the flow past wind turbines
mization equations can be found, Ref. [1]. Results of
and propellers is the helicoidal vortex structure shed
design cases are presented; the second part is devoted
by the blades, that induces velocities in the axial and
to the off-design analysis of the turbine. The equilib-
azimuthal directions which have a profound influence
rium equation and the solution procedure are given.
on the blade flow. For large Reynolds numbers the
It is shown how a consistency condition, derived from
phenomena depends primarily on the advance ratio
V 2P first principles, is enforced in order for the vortex sheet
adv = , and the power coefficient Pτ = . to be a free surface. Ye Zhiquan et Al., Ref. [2], use
ΩR ρΩ3 R5 the vortex theory to predict the aerodynamic perfor-
The vortex sheet is a free surface and its location mance of a turbine, however, no mention is made of
and strength depend on the conditions of the prob- the consistency condition. Their results and experi-
lem. The Euler and Navier-Stokes solvers cannot cap- mental data are compared with the present method.
ture the vortex sheets beyond 2 or 3 turns, due to The test cases demonstrate the ability of the method
the dissipation properties of the numerical schemes. to predict the performance of a turbine under fairly
Therefore, most of the interaction is lost hence the wide range of conditions, including situations where
flow cannot be accurately predicted. In the present substantial separation is present.
approach, the vortex structure is well represented and
the interaction is allowed to take place. The influence 2 Optimization of Turbines
coefficients from the multiple vortex sheets (from two-
and three-bladed rotor blades) are computed using the 2.1 Formulation
Biot-Savart law. This is done in a very efficient and The optimization of propellers presented in [1] is based
accurate way by treating the sheets as vortex lattices on the maximization of thrust (minimization of nega-
with variable discretization steps [1]. The paper is tive drag) for a given power transmitted to the fluid.
divided into two parts: the first part deals with the The corresponding situation for a turbine is to mini-
design aspects and is related to a previous paper on mize power (absorbed power is negative) for a given
thrust, advance ratio adv, and root location y0 . The
Received on December 9, 2001. vortex structure is well defined, since from first prin-
ciples, i.e. actuator disk theory, the convection speed
† University of California Davis, Davis, California
V + ub for the vortex sheet at the rotor is related to
95616

Copyright:
c Japan Society of CFD/ CFD Journal 2000
2 Jean-Jacques Chattot

thrust and power by and CD . The chord distribution is obtained in the


 process, by requiring the lift coefficient to be constant
 T = −2πρR2 (V + u )u along the blade at (Cl )opt corresponding to maximum
b b
, Cl
 P = 2πρR2 (V + u )2 u .
b b Cd
where ub is the induced axial velocity. The follow- 2.2 Application to the Design of Two
ing dimensionless coefficients are introduced: CD = Turbines
2T 2τ Pτ Pτ The result of the optimization is a distribution of cir-
− 2 2 , Cτ = 2 3
= 2
, η = Cp = − ,
ρV R ρV R adv πadv3 culation and induced velocities. In inviscid flow, there
where τ is the torque (negative) and η is the effi- is an infinite number of blade geometries that will pro-
ciency, as defined by Glauert [3], with a maximum duce the optimal distributions. They differ by chord,
ηmax = 16/27. The coordinates, velocities, circulation camber and twist. The viscous data corresponds to a
and chord are made dimensionless as x = R x, y = selected profile, at a representative Reynolds number,
Ry, z = Rz, u = V u , v = V v, w = V w,  c=
 Γ = V RΓ, based on the chord c(y) and incoming velocity q(y)
Rc. Unless otherwise indicated, the dimensionless vari- along the blade span. This approach requires some
ables are used in the rest of the paper and the tilde a priori knowledge of turbine blade profiles that can
is dropped for simplicity. The cost functional for tur- be used, if an efficient design is contemplated. The
bines is defined as Cost = Cτ + λCD , where λ is the optimization produces the chord and twist distribu-
Lagrange parameter. The analysis of the blade flow, tions that completely define the geometry. The test
using strip theory, yields the force and moment coef- cases presented below use the S809 airfoil starting at
ficients for one blade: y0 = 0.25, for which the geometry as well as a large
 set of experimental and theoretical data is available
1

 Cτ = 2 y0 Γ(y) (1 + u(y)) ydy from the NREL study [5]. The 2-D viscous data at

  y 

 1 Reynolds Re = 750, 000 are computed using Xfoil.
 + y0 q(y) + w(y) Cd (y)c(y)ydy They are shown in Figure 1. Note that, for turbines,
1 adv
y  , we have chosen the convention of negative incidence

 −2


C D = y Γ(y) + w(y) dy and negative lift coefficients. The optimum lift co-

 1
0 adv
efficient has been found to be (Cl )opt = −0.8979 at

+ y0 q(y) (1 + u(y)) Cd (y)c(y)dy
α = −6.35 deg. The design of a two-bladed turbine is
carried out for an advance ratio adv = 0.345 and a
where q(y) is the dimensionless incoming velocity in thrust coefficient CD = 0.643. The mesh system con-
the y = const. plane, i.e. sists of jx = 51 points. The distribution of circulation
 y 2 and induced velocities is presented in Figure 2. Note
q(y) = (1 + u(y))2 + + w(y) . that the induced velocities are fairly uniform over the
adv blade. The corresponding chord distribution is shown
in Figure 3. The twist, measured from the y-axis (neg-
2Γ ative) is presented in Figure 4. The efficiency is found
The 2-D viscous data, Cd (α), Cl (α) = are obtained
qc to be η = 0.1677. This is approximately 50% higher
from experimental data or Xfoil (Ref. [4]). than the efficiency of the NREL rotor at that design
0 point as seen in Figure 10.
0.1
−0.25

0.05
−0.5

0
gama(y),u(y) & w(y)

gama(y)
Cl

−0.75
u(y)
−0.05 w(y)
−1

−0.1
−1.25

−0.15
−1.5
0 0.05 0.1 0.15 0.2 0.25

Cd −0.2
0.2 0.4 0.6 0.8 1
y
Fig.1: S809 polar: Xfoil Re = 750, 000.
The discrete formulation, minimization equations Fig.2: Optimum circulation and induced velocities,
and Lagrange multiplier update are the same as for adv = 0.345, CD = 0.643
propellers, except for the exchange of the roles of Cτ
Design And Analysis of Wind Turbines Using Helicoidal Vortex Model 3

0.1
For the same working conditions, adv =
0.345, CD = 0.643, a three-bladed rotor is designed. 0.05

The results for the distributions of circulation and in- gama(y) 2 blades
duced velocities, chord and twist are presented in Fig- 0 u(y) 2 blades

gama(y), u(y) & w(y)


w(y) 2 blades
ures 5, 6 and 7, and compared with the results of the gama(y) 3 blades
u(y) 3 blades
−0.05
two-bladed turbine. As can be seen, the blade loading w(y) 3 blades

and the induced velocities for the three-bladed rotor −0.1


are smaller, as expected, and the blade projected area
is 65% of that of the two-bladed turbine, which is close −0.15

2
to the ratio. The efficiency, however, is increased to −0.2
3 0.2 0.4 0.6 0.8 1
η = 0.1749. y

3 Analysis of Turbines Fig.5: Optimum circulation and induced velocities,


In this part, the turbine geometry is given, from a de- adv = 0.345, CD = 0.643
sign calculation or from other manufacturing consid-
erations. The goal is to find the turbine power output line theory:
at off-design conditions, for a range of advance ratios.
Clm − Clm−1
The vortex structure need no longer be a regular screw Cl (α(y)) =Clm−1 + (α(y) − αm−1 )
surface that satisfy the minimum energy condition of αm − αm−1
Betz [6]. However, for reason of simplicity, and appro- = (Cl0 )m + (Cl1 )m α(y),
priately in cases where the blade loading is light, the
vortex sheet will be constructed as a regular screw. where the coefficients Clm correspond to the piecewise
linear approximation of the lift curve between points
m − 1 and m. They depend also on y. The lift co-
0.1 efficient is expressed in terms of the circulation, the
incoming velocity and the chord. The angle of attack
depends on the induced velocities. In discrete form,
xle(y) & xte(y)

0.05

at point yj , the equilibrium equation reads


0

Γj =
−0.05
   yj  
1 + wj
qj cj (Cl0 )m + (Cl1 )m arctan  adv  + tj   ,
−0.1
0.2 0.4 0.6 0.8 1
2 1 + uj
7
(1)
Fig.3: Chord distribution with quarter-chord along
the y-axis where
−40
 y 2
2 j
qj = (1 + uj ) + + wj
−50
adv
.
−60 The induced velocities depend, in turn, on Γ
through the following summations
t(y)

−70

 w = jx−1 (Γ
k j=1 j+1 − Γj ) aj,k
−80
 ,
u = jx−1
(Γ − Γ )b k j=1 j+1 j j,k
−90
0.2 0.4 0.6 0.8 1
y where the aj,k and bj,k are obtained from the Biot-
Savart formula [1]. Equation (1) is nonlinear, but the
Fig.4: Twist distribution nonlinearity is mild when the angle of attack is above
αcrit corresponding to the first local extremum. When
3.1 Formulation α(y) < αcrit along the blade, the equation of equilib-
The equilibrium equation is given by the condition rium becomes highly nonlinear and admits multiple
that the local lift coefficient is a function of the effec- solutions, as there are multiple values of α for a given
tive angle of attack, as is done in the Prandtl lifting value of Cl . It is necessary to add to equation (1) an
4 Jean-Jacques Chattot

xle(y) & xte(y) 2 blades −40


xle(y) & xte(y) 3 blades
t(y) 2 blades
t(y) 3 blades
0.1 −50
xle(y) & xte(y)

0.05 −60

t(y)
0 −70

−0.05 −80

−0.1 −90
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
y y

Fig.6: Chord distribution with quarter-chord along Fig.7: Twist distribution


the y-axis
twist distributions are well represented by
artificial viscosity or smoothing term. This is done by 

 c(y) = 0.194486 − 0.525251y + 0.919523y 2
∂2Γ 

adding µ∆y 2 to the right-hand side. The points 

∂y  −0.735398y 3 + 0.219878y 4
along the blade span are relaxed according to the fol-

 t(y) = −0.794454 − 4.01836y + 9.97529y 2
lowing iteration procedure: 



   −12.7413y 3 + 8.19633y 4 − 2.10289y 5
2µ δΓj
1+
yj+1 − yj−1 ω The blade is equipped with the NACA 4418. The root
  Γnj+1 − 2Γnj + Γn+1
j−1
location is at y0 = 0.2. The viscous data is shown
= F unj , wjn − Γnj + µ , in Figure 8. A range of values of the tip speed ra-
yj+1 − yj−1 1
tio (T SR = ) from 2 to 12 has been simulated.
where the upper index n refers to the iteration level adv
The results are compared with the theoretical and ex-
and δΓ = Γn+1 − Γn . F (u, w) represents the right-
perimental results of Ye Zhiquan et Al., in Figure 9.
hand side of equation (1), and ω is the relaxation
The agreement is good. For large advance ratios, i.e.
factor. Some under-relaxation is needed for stabil-
T SR < 5, the vortex methods evolve in a similar way
ity. With jx = 21 a value of µ = 0.2 is used. An
except for a shift. This may be due to the simplified
analysis calculation is carried out by prescribing the
equilibrium equation used in Ref. [2], where the term
geometry, the advance ratio and the power coefficient,
in the outer bracket in equation (1) is frozen when
Pτ . The latter is needed to construct the vortex struc-
α < αcrit . At low advance ratios, although the the-
ture, taking into account the clustering of the vortex
oretical results are close, the slopes are different. In
sheets, as the induced axial velocity decreases from ub
this range, the vortex interaction is more pronounced
at the rotor plane to 2ub in the Trefftz plane. The
and the consistency condition enforcement is crucial.
result of the calculation, however, yields a value of
It is not clear how this is handled in the vortex method
Pτ , obtained by integration along the blade, that is
of Ref. [2].
in general different from the input value. The consis-
0
tency condition requires, from first principles, that the
power coefficient corresponds to the helicoidal vortex −0.25

structure, a zeroth-order “equilibrium condition” for −0.5


the vortex sheets. This is done iteratively, using the
result for Pτ as input for the next calculation. It was −0.75
Cl

found that this is a stable procedure and convergence −1


of the consistency condition requires between 3 to 10
cycles. Note that, the first value is guessed, and must −1.25

16
be larger than Pτ min = − πadv3 . −1.5
27
−1.75
3.2 Applications 0 0.05 0.1 0.15 0.2 0.25

Cd
A curve fitting of the discrete geometric data of a
three-bladed wind turbine from Ye Zhiquan et Al. [2] Fig.8: NACA4418 polar: Xfoil Re = 500, 000.
is used with the present approach. The chord and The National Renewable Energy Laboratory
(NREL) has been testing its wind turbine in the
Design And Analysis of Wind Turbines Using Helicoidal Vortex Model 5

0.5
Ref.[2] theory
Ref.[2] experiments
from 1.8868 to 6.6667 has been investigated with
present method
the present approach. The theoretical results are
0.4
compared with the wind tunnel data in Figure 10.
The agreement is quite good, considering that, when
0.3
T SR < 4.26 a large portion of the blade is stalled.
Cp

This is reflected in the distribution of effective angle


0.2
of attack in Figure 11. There is a local extremum
for Cl at αcrit = −14.5 deg = −0.2531 rd, Cl =
0.1
−1.1435, Cd = 0.0589, corresponding to a large sepa-
rated region on the profile.
0
0 2 4 6 8 10 12

1/adv 4 Conclusion
Fig.9: Comparison of theoretical and experimental The design and analysis of turbines can benefit from
power coefficients the simpler and more efficient vortex approach. The
0.4
main feature is the integral representation of the vor-
NREL data
present analysis tex structure shed by the blades, which allows accu-
present design
rate evaluation of the induced velocities and correct
0.3 blade operation simulation. The results indicate that
this is the key issue. This approach can be consid-
ered a second-order method, where the actuator-disk
Cp

0.2
is the zeroth-order and the blade-element/momentum
theory is the first-order method. Nevertheless, the
0.1 approximations of the rigid treatment of the vortex
structure and the strip approach impose a limit on ac-
curacy at low Reynolds numbers and at the tip where
0
0 1 2 3 4 5 6 7 y-derivatives become large. A composite approach,
1/adv using a Navier-Stokes code in the near field and the
vortex method in the far field, seems at this time, the
Fig.10: Comparison of power coefficients with experi- only viable answer to this challenging problem.
mental data
REFERENCES
NASA Ames 80’ by 120’ wind tunnel. Some of the
data has been made available to assess the existing [1] Chattot, J.-J., “Optimization of Propellers Using
computational models. The geometry and flow config- Helicoidal Vortex Model,” CFD Journal, to appear
urations can be found on the web site [5]. the following 2002.
curve fitting formulae have been used: [2] Ye Zhiquan, Zhang Feng and Chen Yan,
 “Prediction of HAWT Aerodynamic Performance
 Using Vortex Theory,” Proceedings WINDPOWER

 c(y) = 0.171734 − 0.101038y
 ’98, American Wind Energy Association, 1998.
t(y) = −0.0416689 − 8.1586y + 18.3718y 2 [3] Glauert, H., The Elements of Aerofoil and

 Airscrew Theory, University Press, Cambridge,

 −21.4031y 3 + 12.6372y 4 − 3.0073y 5 1959.
[4] Drela, M., and Giles, M.B., “Viscous-Inviscid
0 Analysis of Transonic and Low Reynolds Number
−0.1
TSR=6.6667
TSR=5.4083
Airfoils,” AIAA Journal, Vol. 25, No. 10, 1987, pp.
TSR=4.6
TSR=3.797
1347-1355.
−0.2 TSR=3.3
TSR=2.8986 [5] Fingersh, L.J., Simms, D., Hand, M., Jager, D.,
TSR=2.7
−0.3
TSR=2.5145 Cotrell, J., Robinson, M., Schreck, S., Larwood, S.,
TSR=2.2
alpha(y)

TSR=1.8868 “Wind Tunnel Testing of NREL’s Unsteady


−0.4 Aerodynamics Experiment,” AIAA paper No.
−0.5
2001-0035. NREL data base,
http://wind2.nrel.gov/amestest/
−0.6 [6] Prandtl, L., and Betz, A., “Vier Abhandlungen
zur Hydrodynamik und Aerodynamik,” Gottingen
0 0.2 0.4 0.6 0.8 1
y
Nachr., Gottingen, Selbstverlag des Kaiser
Wilhelminstituts fur Stromungsforschung, 1927.
Fig.11: Evolution of α(y) with tip speed ratio
Using the polar of Figure 1, a range of T SR,

Das könnte Ihnen auch gefallen