You are on page 1of 8

Performance of an H-Darrieus in

the Skewed Flow on a Roof1


Sander Mertens
e-mail: s.mertens@citg.tudelft.nl

Gijs van Kuik


Gerard van Bussel
Delft University Wind Energy Research Institute,
Faculty of Civil Engineering and Geosciences,
Stevinweg 1, 2628 CN Delft

Application of wind turbines on roofs of higher buildings is a subject of increasing


interest. However, the wind conditions at the roof are complex and suitable wind turbines
for this application are not yet developed. This paper addresses both issues: the wind
conditions on the roof and the behavior of a roof-located wind turbine with respect to
optimized energy yield. Vertical Axis Wind Turbines (VAWTs) are to be preferred for
operation in a complex wind environment as is found on top of a roof. Since the wind
vector at a roof is not horizontal, wind turbines on a roof operate in skewed flow. Thus the
behavior of an H-Darrieus (VAWT) is studied in skewed flow condition. Measurements
showed that the H-Darrieus produces an increased power output in skewed flow. The
measurements are compared with a model based on Blade Element Momentum theory that
also shows this increased power output. This in contradiction to a HAWT in skewed flow
which suffers from a power decrease. The paper thus concludes that due to this property
an H-Darrieus is preferred above the HAWT for operation on a flat roof of higher
buildings. DOI: 10.1115/1.1629309

Introduction
Apart from the use in remote areas, the use of small wind
turbines for the built environment has gained in interest. This
originates from two sides:
the growing demand for renewable energy and
the increased interest among architects, project developers
and local governments in sustainable and zero- or low-energy
design buildings.
The interest in sustainable, zero- or low-energy design buildings is
caused by, among other reasons, more stringent rules on and the
awareness of the high-energy consumption of buildings. In The
Netherlands, the demand for renewable energy has been raised
because of two main reasons see Scheepers et al. 1 and Greenprices 2:
Europe signed the Kyoto protocol that demands the reduction
of the carbon dioxide emission. This resulted in national renewable electricity production goals
On the 1st of July 2001 a liberalized renewable electricity
market was introduced. This resulted in a large demand for
renewable energy.
According to Panofsky and Dutton 3, the roughness of the built
environment reduces the average wind speed at about 75% of the
average building height to zero. As a result, wind turbines in the
built environment require buildings that are reasonably higher
than the average building height. Wind turbines close to such
buildings can profit from the local flow deviation since they operate in wind accelerated by the building. Buildings may play an
active role in increasing the energy output and are not just used as
a foundation of the wind turbine. Such integration of building and
wind turbine will therefore be referred to as Building Augmented
Wind Turbines BAWTs.
There are three different categories of BAWTs: those situated
on a building; those situated between diffuser shaped buildings;
and those situated in a duct through a building see Mertens 4.
The first category can be explored in the near future; suitable
1
Copyright 2003 by Delft University of Technology. Published by the American
Institute of Aeronautics and Astronautics, Inc. and the American Society of Mechanical Engineers with permission.
Contributed by the Solar Energy Division of THE AMERICAN SOCIETY OF MECHANICAL ENGINEERS for publication in the ASME JOURNAL OF SOLAR ENERGY
ENGINEERING. Manuscript received by the ASME Solar Energy Division, December
19, 2002; final revision, July 17, 2003. Associate Editor: D. Berg.

Journal of Solar Energy Engineering

wind turbines are being designed but the buildings exist. The last
two categories are interesting for the far future because both
building and wind turbine should be designed almost from
scratch.
The use of an H-Darrieus wind turbine on buildings seems a
favorable proposition. The H-Darrieus does not suffer from frequent wind direction changes, architects like the design and the
integration of it in the building, and furthermore, as this paper will
show; the performance coefficient of an H-Darrieus can be higher
than the performance coefficient of a HAWT in similar conditions.
More specifically, this paper concerns an H-Darrieus in the
skewed flow on a flat roof of a mid- to high-rise building. Of
course the wind turbine could be tilted towards the average flow,
making normal operation possible, but this requires an expensive
tilt mechanism and knowledge of the expected skew angle. The
influence of the direction of the flow on the performance of an
H-Darrieus is described using a model based upon multiple stream
tube theory, a combination of momentum and blade element theories. The performance of an H-Darrieus in skewed flow is then
calculated with this model and compared with measurements of
the performance of the H-Darrieus in skewed flow and calculations of the performance of a HAWT in skewed flow.
When operating an H-Darrieus in rural areas, the skewness of
the flow is small compared to the skewness of the flow on roofs
and therefore the effect of skewness on the turbine performance
can be ignored. The effect of the skewed flow on the performance
of an H-Darrieus is thus strongly coupled with the skewness of the
flow on the roof, and the two topics should be discussed together.
We were interested in the effect of the skewed flow on the
performance of an H-Darrieus designed for siting on roofs of
higher buildings, a turbine for which we performed the aerodynamic design for Core International. We carried out some tests in
our open-jet wind tunnel and, to our surprise, the performance
coefficient increased in skewed flow. This raised our attention and
we decided to develop an analytical model in order to understand
the flow phenomena behind this.
The size of the skew angle is determined from a Computational
Fluid Dynamics CFD calculation, and the operation of the
H-Darrieus in skewed flow is discussed with help of an analytical
model based on the Blade Element Method BEM and measurements in the open-jet wind tunnel of the University of Technology
Delft.

Copyright 2003 by ASME

NOVEMBER 2003, Vol. 125 433

Downloaded 04 Jun 2009 to 131.180.23.87. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Skewed Flow Over Buildings


The boundary layer separates at the windward roof edge of the
building and the flow forms a separation bubble on the roof below the streamlines above the roof. As a consequence, the velocity vector of the flow above the separation bubble makes an angle
with the horizontal roof Fig. 1. This angle is subsequently referred to as the skew angle.
The skew angle is largest at the windward roof edge and decreases downwind. The local skew angle also depends on the
roughness of the upwind area. For high upwind roughness, say
upwind obstacles of the order of the building height, the building
is barely noticeable as a separate roughness obstacle and not much
mass flow is moved upwards. As a consequence, the skew angle
will be small. It can be concluded that buildings relatively low
with respect to the surrounding roughness objects can be associated with small skew angles and vice versa see the following
section on CFD calculations.

Computational Fluid Dynamics CFD Calculations


This section describes 3D Computational Fluid Dynamics
CFD calculations on the turbulent separation over a rectangular
model building, carried out with the commercial code Fluent 6.1.
The calculated streamlines are depicted in Fig. 1.
The reference building for the CFD calculation has a height of
20 m, a width of 30 m and a depth of 10 m. The dimensions of the
calculation domain, made nondimensional with the building
height are depicted in Fig. 2. The numbers given in parentheses
represent the nondimensional coordinates of the various corners of
the calculation domain.
For the turbulence model in the CFD calculation, the Reynolds Stress Model is used. Instead of assuming isotropic eddy
viscosity, the RSM calculates the individual Reynolds stresses,
using differential transport equations, together with an equation
for the dissipation rate. The Reynolds Stress Model with the
default parameters has superior performance for flows involving
separation according to Shih, et al. 5, and the Fluent Users
Guide 6.
Close to the wall, the nonequilibrium wall function approach is
used. The viscosity-affected region is not resolved, but instead
bridged by the nonequilibrium wall functions Launder and Spalding 7 and Kim and Choudhury 8, since these functions are
best suited for flows involving separation. In the nonequilibrium
wall functions the Launder and Spaldings loglaw for the mean
velocity is sensitive to pressure gradients. Furthermore, the twolayer-based concept is adopted to compute the budget of the turbulent kinetic energy in the wall-neighboring cells. By doing this,

Fig. 1 Visualization of the turbulent separation over a rectangular building in boundary layer flow

434 Vol. 125, NOVEMBER 2003

Table 1 Equivalent Surface Roughness Lengths


Roughness length m

Ground Cover
Long grass or rocky ground
Pasture land
Suburban housing
Forests, cities

0.05
0.20
0.6
1 6

the nonequilibrium wall functions partly account for nonequilibrium effects that are neglected in the standard wall functions Fluent 5 user guide 6.
The discretization is carried out with the first-order upwind
scheme and the velocity-pressure flow field is determined using
the SIMPLE model see Fluent 5 user guide 6.
The nonstructured mesh is built of 444663 hybrid tetra grid
cells, with the smallest cell size at the walls. The smallest cell
dimension is equal to or larger than double the roughness height.
Calculations with a second-order upwind scheme showed that numerical diffusion did not significantly influence the results of the
first order upwind scheme.
The input at the velocity inlet consists of the axial velocity
u(z), the turbulent kinetic energy K and the turbulent dissipation
rate (z) for a neutral boundary layer, all at a given height z. The
boundary conditions at the inlet are based on the surface roughness in the domain so the boundary layer is unchanged in the
entire domain. For a neutral boundary layer with ground surface
roughness z 0 , the logarithmic profile of the boundary layer reads

u
zd
u z * ln
.

zo

(1)

In this, u is the friction-velocity, is the Von Karman constant


*
taken as 0.4, and d is the displacement length, which Panofsky
and Dutton 3 claim is typically 75% of the average height of the
roughness elements l:
d0.75l.

(2)

The surface roughness z 0 is determined from measured wind profiles. Panofsky and Dutton 3 give the information in Table I for
associating a roughness length with the description of the surface.
According to Wilcox 9 the turbulence parameters at the inlet
are:
the turbulent kinetic energy K
K

u2
*
C

(3)

Fig. 2 Dimensions of the domain of the CFD calculation

Transactions of the ASME

Downloaded 04 Jun 2009 to 131.180.23.87. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

with k- model parameter C 0.09, and the turbulent dissipation (z)


z

u3
* .
zd

(4)

Negative or infinite values of (z) for zd are avoided by putting


them equal to the value of (z) at the center of the first cell close
the ground surface with (z)0.
The CFD calculations yield the separation streamlines on the
flat roof plotted in Fig. 3.
From the height of the separation streamlines and the distance
to the windward roof edge, the skew angle of the flow, as shown
in Fig. 4, is calculated. The wiggling of the curve is caused by
finite grid size. At the leading edge, the skew angle is equal to 90
deg, as a result of the deflection of the wind by the upwind wall.
From Fig. 4, it can be concluded that the skew angle of the flow
above the center of the flat roof of the building with a surrounding
roughness of z 0 0.03 m lies between 10 and 20 deg apart from
close to the windward roof edge. For higher buildings the skew
angle will increase, while for higher roughness the skew angle
will decrease.
The CFD calculation of the velocity above the center of the
roof shows that the velocity gradient in the flow above the separation bubble (z27 m) is comparable with the gradient in the
undisturbed flow see Fig. 5. It is common practice to treat this
gradient as a second order effect for the performance of VAWTs in
boundary layer flow since it can be showed see de Vries 10

that wind shear has a small effect on the power output of a VAWT.
The same is assumed for the H-Darrieus in the flow above the
roof.

The H-Darrieus in Skewed Flow


In the following, the single actuator disk multiple stream tube
model first developed by Wilson in 11 will be used. The basic
assumptions important for the model developed in this paper are
that:
the flow through the H-Darrieus can be divided into flow
through multiple stream-tubes that do not interact with each
other,
the deceleration of the air by the rotor can be modeled with
one actuator plate,
the chord length of the blades c is much smaller than the
radius R of the H-Darrieus.
Schematically, the H-Darrieus can be represented by two rotor
parts; one at the windward side and one at the leeward side of the
H-Darrieus. In single actuator disk multiple streamtube theory,
however, both parts are modelled with one actuator disk. For
skewed flow with skew angle , the flow at the rotor can be split
up as shown in Fig. 6 into two parts; one part of the flow will pass
only one rotor part subscript s and one part of the flow subscript
d will pass two rotor parts. In the following, the single rotor part
variables will have a subscript s and the double rotor part will
have a subscript d. The length of the parts depends on the skew
angle of the flow and the induction factor see the following discussion on induction factor. For a high skew angle and a wide
rotor, almost the whole H-Darrieus operates as two single rotor
parts.
According to the cross-flow principle mentioned in Jones and
Cohen 12, the lift force on a blade in skewed flow is determined
by the component perpendicular to the blade and is not dependent
on the component parallel to the blade. This cross-flow principle
is valid as long as skin friction is of no importance. The lift on a
blade of the double rotor part in skewed flow can then be calculated based on the velocity component perpendicular to the blade
V o,y,d , horizontal component which reads:
V o,y,d V o cos a d ,

(5)

where V o is the undisturbed velocity at the roof, is the skew


angle and a d is the induction factor of the double rotor part. The

Fig. 3 CFD calculations of the separation streamlines at a


model building with dimensions heightwidthdepth2030
10 m for different roughness z 0 of the upwind area

Fig. 4 CFD calculations of the skew angle deg at a model


building with dimensions heightwidthdepth203010 m
for different roughness z 0 of the upwind area

Journal of Solar Energy Engineering

Fig. 5 CFD calculation of the velocity profile above the center


of the roof compared to the undisturbed velocity profile

NOVEMBER 2003, Vol. 125 435

Downloaded 04 Jun 2009 to 131.180.23.87. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

induced velocity, which is by definition set equal to a d V o , is


assumed to be normal to the actuator disk see Fig. 7, since the
force caused by the actuator plate is a pressure force and thus
normal to the plate. According to Fig. 6 the flow through the
double rotor part has an angle of:

d arctan

sin
cos a d

(6)

with the normal vector at the actuator plate2.


The flow thus ceases to pass the rotor parts if the denominator
is equal to zero. We therefore demand that:

arccos a d .
On the other hand, the horizontal component of the velocity
downwind of the actuator disk (V o (cos 2ad)) equals zero or
becomes negative for too high an induction factor. This causes the
flow assumptions in the momentum theory to be violated. We
therefore demand that:
cos 2a d 0,
which gives
1
a d cos
2

1
F y,d c l,d R sin cV cd,d ,
2

(9)

with radius of the H-Darrieus R, rotational frequency , and rotational angle ( 0 corresponds to a blade moving parallel to
the wind direction, see Fig. 8.
The blade occupies a stream-tube width of:
dxRd sin .

(10)

From Equation 9, it can be seen that the absolute value of the


force parallel to the stream tube and with this, the local power
production repeats itself after 360 deg. Therefore the analysis
is limited to one revolution. For the double rotor part, the blade
passes the flow in the streamtube twice. The blade thus stays in
the streamtube for a fraction of 2d /2 of the total time of one
revolution. So, Equation 9 should be multiplied with d / to
find the average blade force in the streamtube. For the single rotor
part, this averaging factor is half this value since the blade passes
the flow just one time. From Equation 9, the average force
dF bl,d for B blades in the direction of the streamtube at the double
rotor part now follows as:
dF bl,d Bc l,d R sin cV cd,d

d
.
2

(11)

The lift L per unit blade length reads:


1
L V cd,d c l,d V 2cd,d c,
2

(7)

where is the air density, V cd,d is the resultant velocity on the


blade see Fig. 8, is the strength of the bound vortex of the
airfoil, c is the chord length of the blade, and c l,d is the lift
coefficient of the airfoil of the double rotor part. The lift coefficient of the airfoil is approximated with a function that takes
Reynolds effects and stall of the airfoil into account. The function
is fit to the measurements of the NACA 0018 profile that can be
found in Paraschivoiu 13.
The force on the blade per unit blade length can be expressed
as:

d V cd,d .
F

(8)

With equation 7 this results in a blade force in the y-direction at


the double rotor part of:
2
The first paper of the authors on the topic was published at the ASME Wind
Energy Symposium 2003 contribution nr. AIAA-2003-0523. In this paper, the
change of the undisturbed flow angle by the induction factor was not taken into
account.

Fig. 7 Schematic drawing of the skewed flow on an actuator


disk

Fig. 6 Schematic side view of an H-Darrieus in skewed flow.


Between the dotted lines: operating as a double rotor part d.
Above and below the dotted lines: operating as a single rotor
part s.

436 Vol. 125, NOVEMBER 2003

Fig. 8 Top view of the H-Darrieus in skewed flow

Transactions of the ASME

Downloaded 04 Jun 2009 to 131.180.23.87. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

The blade force can also be found with momentum theory. Glauert
momentum theory for yawed flow shows good agreement with
measurements on a HAWT in yawed flow as shown by Madsen,
Sorensen, and Schreck 14. It gives the thrust force perpendicular
to the actuator plate, which is assumed to lie on the y-axis as
shown in Burton et al. 15. The thrust force reads:
dF mo,d Rd sin W d 2a d V o

a d

1 Bc
V cd,d
.
c
4 R l,d W d
R

.
Vo

(15)

The same derivation is carried out for the single rotor part where
we have to keep in mind that the blade moves only one time
through the flow. The result is:
a s

1 Bc
V cd,s
c
.
8 R l,s W s

In the previous section, the flow angle of the double rotor part
is found to be different from the single rotor part different induction factor. The flow through the double rotor part pushes aside
the flow of the single rotor part see Fig. 9.
It is assumed that the flows do not mix but instead have an
average flow angle, which is the weighted average of the flow
angles calculated from the induction factors. This average flow
angle is denoted by .

(18)

h d Hh s H2R tan sin

(19)

for the double rotor part. Larger skew angles cause the blade
length of the double rotor part to disappear and the length of the
blade for the single rotor part to be limited to h s H for:
arcsin

H
H
arcsin
.
2R tan
2R tan

(20)

For outside these borders and max the blade length of the
different rotor parts can again be found with Equations 18 and
19.
The known blade lengths of the single and double rotor part
enable us to derive an implicit relation for the flow angle, which is
taken as a weighted average on the basis of those blade lengths.
The flow angle is thus defined as:

(16)

Rotor Performance

h s 2R tan sin
for the single rotor part and

(14)

where the Tip Speed Ratio is defined as:

(17)

H
2R

reads

(13)

Equating 11 and 12 gives an implicit relation for the induction


factor a d for B blades for a double rotor part in skewed flow:

maxarctan

(12)

for the double rotor plate, where W d is the total resulting velocity
towards the actuator plate see Fig. 7:
W d V o sin2 cos a d 2 .

According to Fig. 6 and Fig. 8 the blade length for flow angles
up to:

hs
Hh s

.
Hh s s Hh s d

(21)

The extracted power d P per unit blade length in a stream tube


with width dy can be found from the thrust force Equation 12
times the velocity through the rotor Equation 5:
d PdF mo,skew V o,y ,

(22)

since both the thrust force and the x-velocity point in the same
direction. From the power the performance coefficient calculated
on basis of the frontal area of the rotor reads:
dC P,d 2d sin W d a d

hd
cos a d
H

(23)

dC P,s 2d sin W s a s

hs
cos a s .
H

(24)

and

The viscous loss in performance for the double rotor part can be
found from the drag coefficient of the blades c d,d :
dC P,cd,d Bc d,d


V cd,d
Vo

ch d d

2RH 2

(25)

ch s d
.

2RH 2

(26)

and from c d,s for the single rotor part


dC P,cd,s Bc d,s


V cd,s
Vo

In order to calculate the total performance coefficient of a rotor


part dx within a stream tube we need to add the efficiencies and
losses of the different parts as:
dC P dC P,d 2dC P,s 2dC p,cd,d 2dC p,cd,s

Fig. 9 Side view of the H-Darrieus with illustration of the different flow angles

Journal of Solar Energy Engineering

(27)

where we counted the power production of the single rotor part


twice since there is one single rotor part at the windward side of
the rotor and one at the leeward side of the rotor. This fact is
crucial for the operation of the H-Darrieus in skewed flow. It
allows the power production to increase in skewed flow as a consequence of the increase in rotor area that experiences undisturbed
flow. The maximum performance coefficient of the total rotor is
the sum of the contributions of the single and double rotor part.
For maximum performance coefficient of the separate parts, the
single rotor part needs a higher TSR than the double rotor part.
NOVEMBER 2003, Vol. 125 437

Downloaded 04 Jun 2009 to 131.180.23.87. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

For maximum performance of the total rotor, the single rotor part
will thus operate at a bit too low TSR and the double rotor part
will operate at a bit too high TSR. However, the net result on the
performance of both rotor parts, each operating below maximum
performance, can be an increase of the turbine performance in
skewed flow!
The total performance coefficient, including drag losses, can be
found from integration of dC P :
C P

dC P d .

(28)

Wind Tunnel Tests


In order to validate the analytical model of the H-Darrieus in
skewed flow, tests were carried out in the open jet wind tunnel of
the TU Delft Wind Energy Section. The diameter of the wind
tunnel is 2.2 m and we tested a two bladed H-Darrieus of diameter
0.755 m, height 0.5 m and chord length 0.08 m see Fig. 10 at
wind tunnel velocity of 7 m/s. The airfoils of the Darrieus were
NACA 0018 profiles. At maximum power for zero skew angle, the
average induction factor at the double rotor area was calculated
equations 23 to 27 to be 0.35 see the Results section, below. The tests consisted of measurement of the number of revolutions per minute under different loads for different skew angles.
A Prony brake3 was used to load the H-Darrieus. For the estimation of the maximum performance of the H-Darrieus, the load was
increased with small steps until the rotor stopped turning. The
performance at the highest load was the maximum.

The net effect of the increased rotor area combined with the less
than optimally loaded single and double rotor parts, as discussed
after equation 27, thus causes the performance of the total rotor
to increase.
The calculations showed that the average induction factor at
maximum performance for zero skew angle is 0.35. Locally however, the induction factor is higher than 0.5 see Fig. 13. The
assumed flow pattern in momentum theory is therefore violated
zero velocity at the downwind part of the stream tube. For induction factors above approximately 0.4, one should use empirical
evaluations see for instance Burton et al. 15 of the thrust force
to overcome the breakdown of momentum theory. These relations,
however, are not known for skew angles and are thus not integrated in the current model for the operation of an H-Darrieus in
skewed flow. A first attempt to improve the model should therefore focus on the thrust force for heavily loaded H-Darrieus in
skewed flow.
Figure 12 shows that the predicted performance coefficient is
very high compared to the measurements. Furthermore, there is a
mismatch between data and prediction in the skew angle at which

Results
The measurements and calculations with the analytical model
Equations 23 to 27 of the performance coefficient and TipSpeed are presented in Figs. 11 and 12 as functions of the flow
skew angle, .
3
A Prony brake is a disk on the power axis with a pulley around the circumference
of the disk. One end of the pulley is fixed at a force sensor that measures the force in
the rope and at the other end of the rope a load causes friction around the circumference of the disk. This friction force and thus the torque are known by the fact that
the sum of the moments around the power axis is zero.

Fig. 11 Measured and calculated Tip Speed Ratio at maximum


aerodynamic efficiency in skewed flow with skew angle

Fig. 10 Photo of the test of the H-Darrieus in the open-jet wind


tunnel. The size of the Darrieus is overestimated because of the
camera point of view

438 Vol. 125, NOVEMBER 2003

Fig. 12 Measured and calculated performance coefficients in


skewed flow with skew angle

Transactions of the ASME

Downloaded 04 Jun 2009 to 131.180.23.87. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

the performance coefficient and the Tip Speed Ratio start to decrease measured to be about 25 deg. The model performance
coefficient prediction showed a better agreement with the measurements for a lift coefficient equal to 2 sin() maximum error
in performance coefficient below 15%. Any discussion about
possible causes of the mismatch between model and measurements, however, is speculative since the local induction factor
sometimes violates the flow pattern assumed in momentum
theory.
The presented model should provide a first basis for an assessment of the performance of an H-Darrieus in skewed flow. Many
modeling problems occur. One of the most important is undoubtedly the different flow angles through the single and double rotor
parts. To what extent does the flow through the double rotor part
move aside the flow through the single rotor part? The calculated
difference in flow angle can be up to 30 deg at the heavily loaded
center of the H-Darrieus. Besides focusing on the induction factor
of heavily loaded Darrieus in skewed flow, future efforts to refine
the model should thus focus on the problem of determining the
flow angle

The Horizontal Axis Wind Turbine in Skewed Flow

Fig. 13 Calculated induction factor vertical at zero skew


angle, plotted on the projected diameter of the H-Darrieus horizontal. Right part of the graph: blade moves against the wind

On the Research

According to the Wind Energy Handbook 15, Glauert gives


the thrust force of an actuator disk in yaw. This equation can be
used to calculate the performance coefficient of a HAWT in
skewed flow.
C P 4a 1a 2 cos a cos a .

(29)

According to Equation 29 the optimum performance coefficient


of a HAWT thus shows a decrease for increasing angle as shown
in Fig. 14.

Conclusions
The performance coefficient, C P , of an H-Darrieus in skewed
flow, based on the projected frontal rotor area at zero skew angle,
can increase above that for nonskewed flow. The effect has been
measured and modeled and can be explained with the threedimensional shape of the H-Darrieus, a geometry that causes an
effective increase of the energy extracting area in skewed flow.
The maximum performance coefficient of a slender H-Darrieus
(h/D1) in skewed flow increases less than a disk-like
H-Darrieus (h/D1) because the relative increase in area experiencing undisturbed flow is less.
The local induction factor in skewed flow can be well above the
permitted induction factor of momentum theory. Therefore, momentum theory is no longer valid and the local performance
predictions are questionable. The author has not found any useful
research on the thrust force for high induction factor in skewed
flow, so it appears that theory must be developed for this problem
area.
The angle of the flow through the disk is a function of the
induction factors through the single and double rotor part. Since
the induction factor of the double rotor part is the largest, the flow
through the double rotor part moves towards the flow through the
single rotor part, thus pushing aside the single-rotor flow. It is still
unknown to what extent this interaction takes place. Changes in
the model showed that this effect has a big influence on the performance.
For a given dimension of the H-Darrieus, the Tip Speed Ratio at
maximum performance increases in skewed flow. This increase
must be taken into account when designing an H-Darrieus for
skewed flow since the centrifugal force increases with the Tip
Speed Ratio squared.

The presented research is part of research into wind energy in


the built environment, which will be the subject of a report to be
published at the beginning of 2004. The work focuses on the
aerodynamic aspects of wind turbines close to buildings.
The authors thanks Simon Toet, Michiel Zaaijer and graduating
student Joto Andre Monteiro Sardo of the wind energy group of
Delft University of Technology for their indispensable assistance
in the presented research.

Nomenclature
a
C
B
c
cl
CP

cd
c P,cd
Fig. 14 Maximum aerodynamic efficiency of a HAWT in yawed
flow according to Glauert momentum theory for yawed flow
15

Journal of Solar Energy Engineering

d
D

induction factor of the H-Darrieus .


k- model turbulence model constant0.09 .
number of blades of the H-Darrieus .
chord length of the airfoil blade m
lift coefficient of the airfoil blade .
performance coefficient of the H-Darrieus defined on
the projected frontal area .
drag coefficient of the airfoil .
performance coefficient loss caused by the drag of
the airfoils .
displacement length m
diameter of the H-Darrieus m
NOVEMBER 2003, Vol. 125 439

Downloaded 04 Jun 2009 to 131.180.23.87. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

F bl The force parallel to the stream tube caused by the


airfoil in the stream tube found with blade element
theory per unit blade length N/m
F mo The force parallel to the stream tube caused by the
airfoil in the stream tube found with momentum
theory per unit blade length N/m
h height of the blade length of a rotor part double or
single rotor part N/m
H height of the H-Darrieus m
K turbulent kinetic energy m2 /s2
l average height of the roughness elements m
L lift force of the airfoil of the H-Darrieus per unit
blade length N/m
P power of the H-Darrieus per unit blade length Watt/
m
R H-Darrieus radius m
u undisturbed velocity in the boundary layer m/s
u 0 undisturbed velocity at a certain height in the boundary layer m/s
u friction velocity m/s
*
V o,y horizontal component of the total undisturbed velocity at the roof m/s
V o total component of the undisturbed velocity at the
roof m/s
V cd resulting velocity at the airfoil blade perpendicular to
the H-Darrieus axis m/s
W total resulting velocity at a rotor part m/s
x horizontal coordinate perpendicular to the direction of
the undisturbed wind vector .
y horizontal coordinate parallel to the direction of the
undisturbed wind vector .
z vertical coordinate perpendicular to the wind direction .
z 0 roughness height of the surface m
skew angle of the flow above the separation bubble
on the roof .
resulting flow angle through the rotor of the
H-Darrieus .

440 Vol. 125, NOVEMBER 2003

turbulent dissipation of kinetic energy .


rotational angle of the H-Darrieus .
Von Karman constant0.4 .
Tip Speed Ratio of the H-Darrieus .
Strength of the bound vortex of the airfoil

Subscripts
d used for the double rotor part .
s used for the single rotor part .

References
1 Scheepers, M.J.J. et al., 2001, Energie Markt Trends, ECN report number:
ECN-p-01-009.
2 www.greenprices.nl.
3 Panofsky, H.A., Dutton, J.A., 1984, Atmospheric Turbulence, Wiley, New
York.
4 Mertens, S., 2002, Wind Energy in Urban Areas, Refocus, the international
renewable energy magazine.
5 Shih, T.-H., Liou, W.W., Shabbir, A., and Zhu, J., 1995, A New k-e EddyViscosity Model for High Reynolds Number Turbulent FlowsModel Development and Validation, Comput. Fluids, 243, pp. 227238.
6 Fluent, 1998, Fluent 5 Users Guide 2, Fluent Inc., Lebanon, USA, 9-8.
7 Launder, B.E., and Spalding, D.B., 1974, The Numerical Computation of
Turbulent Flows, Comput. Methods Appl. Mech. Eng., 3, pp. 269289.
8 Kim, S.-E., and Choudhury, D., 1995, A Near-Wall Treatment Using Wall
Functions Sensitized to Pressure Gradient, ASME FED Vol. 217, Separated
and Complex Flows, ASME.
9 Wilcox, D.C., 1994, Turbulence Modeling for CFD, DCW Industries, Inc.,
pp. 127.
10 de Vries, O., 1979, Fluid Dynamic Aspects of Wind Energy Conversion,
AGARD AG 243, July.
11 Wilson, R.E., Lissaman, 1994, P.B.S., Applied Aerodynamics of Wind Power
Machines, Aerovironment, Inc, Pasadena, CA.
12 Jones, R.T., Cohen, D., 1957, Aerodynamics of Wings at High Speeds, Aerodynamic Components of Aircraft at High Speeds, High Speed Aerodynamics
and Jet Propulsion, Vol. VII, AF Donovan, HR Lawrence eds, Princeton
University Press, Princeton, NJ.
13 Paraschivoiu, I., 2002, Wind Turbine Design With Emphasis on Darrieus
Concept, Ecole polytechnique de Montreal, 2002.
14 Madsen, H.A., Sorensen, N.N., Schreck, S., 2003, Yaw Aerodynamics Analyzed With Three Codes in Comparison With Experiment, ASME Wind Energy
Symposium, contribution nr. AIAA-2003-0519.
15 Burton, T., Sharpe, D., Jenkins, N., Bossanyi, E., 2001, Wind Energy Handbook, J. Wiley & sons Ltd, England.

Transactions of the ASME

Downloaded 04 Jun 2009 to 131.180.23.87. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

You might also like