You are on page 1of 24

Appendix A-Blade Element Momentum Theory 176

APPENDIX A
Appendix A-Blade Element
Momentum Theory
Appendix A-Blade Element Momentum Theory 177
A.1. Introduction
With the availability of large computing power and developments in grid generation and
numerical algorithms, it is very tempting to compute the turbulent flow around a rotating
blade using the equations of motion in their differential forms, (Reynolds-averaged Navier-
Stokes equations). These equations need a model of turbulence to make them a closed set.
A large number of models of turbulence have been proposed, for example, mixing length
models, one-equation models, two-equation models, Reynolds stress models, algebraic
stress models, etc. Tulapurkara (1996) gives a review of these models in computation of
flow past aerofoils and wings.
A rotating blade is experiencing a wide range of angle of attacks along its span, from
negative values to post-stall values. A high angle of attack causes a strong adverse-
pressure-gradient separating flow in which no turbulent model can be applied accurately.
Applying the equations of motion in their integral forms (linear and angular momentum)
around the wind turbine (rather than the blade) together with the experimental data of blade
aerodynamic characteristics is an alternative to solve the flow field and find the blade
loading. This method is known as Blade Element Momentum Theory, BEMT.
A.2. Induced, total and relative velocity fields
Extracting energy from wind slow downs the wind speed and rotation of blades induces
some circumferential and radial velocities to the wind. Wind speed retardation depends on
the amount of the extracted energy (wind turbine loading), while the induced
circumferential velocity depends on the wind turbine angular velocity. Induced radial
velocity due to centrifugal forces is much smaller than the two other components and is
usually neglected.
Flow in the plane of a wind turbine rotor, from now on referred as disk, can be considered
as a combination of upstream mean velocity field, w V and induced velocity field, i V .

i w V V V + = (A.1)
Figure (A.1) shows the coordinate systems used to represent velocity fields. In this figure:
Appendix A-Blade Element Momentum Theory 178
r x : Disk (rotor) system of coordinates
x : Rotor axis
r : Rotor plane
r : Disk radial coordinate, 0<r R; R: rotor radius
s t n : Blade system of coordinates
n: Normal to the blade
s t : Blade plane
t n : Aerofoil plane
s : Blade span-wise coordinates
: Yaw angle
: Conning angle
: Azimuth angle
: Rotor angular velocity
Here it is assumed that wind velocity has no vertical components ( 0 =
Z
w
V ).
Using the s t n system of coordinates, wind velocity w V at a general point P can be
expressed as
( ) ( ) ( )
s w t w n w P
w e V e V e V V sin sin cos sin cos | + + + + + = (A.2)
Conning angle is usually very small and therefore cos sin << , but yaw angle can be
large enough such that sin and cos have the same order of magnitude. Neglecting terms
including sin in comparison with the terms including cos , Equation (A.2) can be
rewritten in a simpler form as

Appendix A-Blade Element Momentum Theory 179
s w t w n w P
w e V e V e V V sin cos sin cos cos sin cos cos | + + = (A.3)


Figure A.1- Upstream wind velocity components in different coordinate systems
Neglecting radial induction component, induced velocity i V has only two components of
x
i
V and

i
V in the disk plane.
x
i
V is directed opposite to the axial component of the wind
velocity and

i
V is due to induced angular velocity in the opposite direction of the rotor
angular velocity.
Induction factors are defined as
x
w
x
i
V V a = (A.4)
= r V a
i

(A.5)
where a and a are axial and rotational or swirl induction factors respectively. One can
write the induced velocity vector at point P in r x system of coordinates in terms of
axial and rotational induction factors and then transform it to s t n system as
t
x
w P
i e a r i aV V

| = (A.6.a)
s-t


x

w
V
x
i
V
( ) + =

sin
w
t s
w
V V
r
( ) + = cos
w
n
w
V V

n

i
V
( ) sin sin +
w
V
z
P
s-t
s

) ( t
s
i
V
( ) cos sin +
w
V

x
Appendix A-Blade Element Momentum Theory 180
t n w P
i e a r e aV V cos cos | = (A.6.b)
Finally substituting w V from Equation (A.3) and i V from Equation (A.6.b) back into
Equation (A.1), velocity field in the plane of blade can be written as
{ } { } { }
s w t w n w
e V e a r V e a V V sin cos sin cos cos sin ) 1 ( cos cos + + =
(A.7)
For a moving blade the relative velocity of the flow at a point P located on the blade is
P
blade
P
flow
P
rel V V V | | | = (A.8)
where
t P
blade e r V | = (A.9)
Combining Equations (A.7),(A.8) and (A.9) leads to
{ }
n w
rel e a V V ) 1 ( cos cos =
{ }
t w
e a r V ) 1 ( cos cos sin + +
{ }
s w
e V sin cos sin + (A.10)
Figure (A.2) shows the relative velocity in the plane of blade aerofoil. Inflow angle and
normalised in-plane relative velocity can be derived from Figure (A.2) as

cos cos sin ) 1 (


) 1 ( cos cos
tan
+

=
a
a
r
(A.11)
and


sin
) 1 ( cos cos | a
V
V
w
t n rel

=

(A.12)
where local velocity ratio,
r
is defined as
Appendix A-Blade Element Momentum Theory 181
w
r
V
r
= (A.13.a)
or in terms of blade span coordinate s (
cos
r
s = ):
w
r
V
s
=

cos
(A.13.b)

Figure A.2- Relative velocity in aerofoil plane
A.2.1. Special case Zero yaw
In the case of zero yaw angle, 0 = , Equations (A.4), (A.10), (A.11) and (A.12) can be re-
written as
w
x
i
V V a = (A.14)
t n w rel
e a r e a V V ) 1 ( ) 1 ( cos + = (A.15)
) 1 (
) 1 ( cos
tan
a
a
r
+


(A.16)

sin
) 1 ( cos | a
V
V
w
t n rel

=

(A.17)
and the velocity diagram changes to Figure (A.3).

n
t
t n rel
V

|

) 1 ( cos cos a V
w

cos cos sin ) 1 (
w
V a r +
Appendix A-Blade Element Momentum Theory 182
Now, flow through the rotor disk can easily be found by superposing the x-components of
the upstream wind and the induced velocities.
) 1 ( a V V V V
w
x
i
x
w d
= = (A.18)


Figure A.3- Relative velocity in aerofoil plane; zero yaw
A.3. Momentum Theory
Momentum theory applied to the wind turbine aerodynamic is based on three basic
assumptions:
1. Axi-symmetric flow
2. Steady flow
3. Frictionless flow
Figure (A.4) shows an axi-symmetric flow through a wind turbine and typical qualitative
variations of pressure, velocity and rotation between upstream of a wind turbine and far
wake behind it.
n
t

) 1 ( cos a V
w

) 1 ( a r +
t n rel rel
V V

= |
Appendix A-Blade Element Momentum Theory 183

Figure A.4- Pressure, velocity and rotation distributions
A.3.1. Thrust and torque coefficients
Applying the x-component of linear momentum equation to the annulus control volume
shown in Figure (A.5), gives thrust force as ) (
FW rotor
V V dQ dT =

,where the volume
flow rate
disk d
dA V dQ = ,
w
V V =

and ) 1 ( a V V
w d
= and therefore
disk FW w w rotor
dA V V a V dT ) )( 1 ( = (A.19)
Applying the energy equation for the same control volume gives the turbine power.
) ( 5 . 0
2 2
FW w disk d
V V dA V dP = (A.20)
Disk Far Wake
Pressure, P
Velocity, V
Rotation,
Appendix A-Blade Element Momentum Theory 184

Figure A.5- Annulus control volume; Linear momentum balance
Turbine power can also be obtained by multiplying the thrust force and flow velocity at the
disk.
d
dTV dP = (A.21)
Combining Equations (A.19), (A.20) and (A.21) concludes
w FW
V a V ) 2 1 ( = (A.22)
By substituting
FW
V back in Equation (A.19),
rotor
dT will be determined in terms of the
wind velocity at the upstream and the axial induction factor.
disk w rotor
dA V a a dT
2
) 1 ( 2 = (A.23)
Thrust coefficient by definition is
disk w
T
dA V
dT
C
2
2
1

= (A.24)
And therefore as a result of Momentum Theory it becomes:
) 1 ( 4 a a C
T
= (A.25)
CV
FW
dQV dT

dQV
Appendix A-Blade Element Momentum Theory 185
To determine torque coefficient by the Momentum Theory one can start from applying the
angular momentum equation about the x-axis for the control volume shown in Figure (A.6)
to find a relation between the rotation in far wake and circumferential velocity at disk as
( ) ( ) ( )


disk disk disk FW
rV rr r r 2 2 / 2
2 2
= = = .
Since the circumferential velocity

disk
V is only due to induction, one can substitute

i
V V =
from Equation (A.5) in the above equation to find( )
FW
r
2
.
( ) = a r r
FW
2 2
2 (A.26)


Figure A.6- Annulus control volume; Angular momentum balance between disk and Far
Wake
Applying angular momentum equation about xaxis for the control volume shown in
Figure (A.7), the applied torque on the rotor will be determined.
( ) ( ) { }

=
2 2
r r dQ dM
FW x
(A.27)
Combining Equations (A.26) and (A.27) gives the rotor torque as
disk w x
dA r a a V dM
2
) 1 ( 2 = (A.28)
Torque coefficient is defined as
CV
FW
r dQ ) (
2

disk
r dQ ) (
2

Appendix A-Blade Element Momentum Theory 186
disk w
x
M
rdA V
dM
C
2
2
1

= , (A.29)
and finally as a result of the Momentum Theory it becomes
) 1 ( 4 a a C
r M
= (A.30)


Figure A.7- Annulus control volume; Angular momentum balance
A.3.2. Tip and Hub Losses
In momentum theory, the axi-symmetric flow is the basic assumption, which holds if the
turbine rotor has an infinite number of blades with zero chord length. In the case of a real
turbine with a finite number of blades, the induced velocity on the blades is different from
the mean induced velocity in the flow annulus and therefore circumferential symmetry
does not hold. The non-uniformity of the induced flow field makes the actual local
T
C and
M
C to be smaller than the expected values by the optimum actuator disk theory. The
departure of the induced velocity,
T
C and
M
C from their momentum theory values is more
significant near the tip and root of the blade. These deviations from the uniform induced
velocity flow field are called tip and hub losses. The overall loss factor, F is defined as
hub tip
F F F = (A.31)
CV
FW
r dQ ) (
2

dM

) (
2
r dQ
Appendix A-Blade Element Momentum Theory 187
In which
tip
F is unity at inboard parts of the blade and takes smaller values near the tip of
the blade and
hub
F is unity at outboard parts of the blade and takes smaller values near the
hub of the blade. Overall loss factor F can be applied on the induced velocity or disk
velocity.
A.3.3. Loss Models
Depending on which parameter or parameters are affected by F , different models will be
introduced. Most of commercial codes use two models that are known as
1. Classical or Wilson
2. Standard or Advanced or Wilson-Walker.
a. Classical Model
In Classical model, loss factor F is applied to the axial induced velocity:
w
x
i
aFV V = (A.32)
and therefore
) 1 ( aF V V
w d
= (A.33)
) 2 1 ( aF V V
w FW
= (A.34)
) 1 ( 4 aF Fa C
T
= (A.35)
) 1 ( 4 aF a C
r M
= (A.36)
According to Figure (A.8), which shows the relative velocity diagram for this model, the
normalised relative velocity is given by Equation (A.37).


sin
cos ) 1 ( | aF
V
V
w
t n rel

=

(A.37)
b. Wilson-Walker Model
Appendix A-Blade Element Momentum Theory 188
In Wilson-Walker model, loss factor F is directly applied to the disk
velocity, ) 1 ( a FV V
w d
= and the difference between the free stream velocity and far wake
velocity is defined as
w FW w
aV V V 2 = . With the above assumptions thrust and torque
coefficients can be calculated as:
) 1 ( 4 a Fa C
T
= (A.38)
) 1 ( 4 a F a C
r M
= (A.39)
For this model the relative velocity diagram becomes as shown in Figure (A.9) and the
normalised relative velocity becomes as given by Equation (A.40).


sin
cos ) 1 ( | F a
V
V
w
t n rel

=

(A.40)

Figure A.8- Relative velocity in aerofoil plane; zero yaw; Classical model

Figure A.9- Relative velocity in aerofoil plane; zero yaw; Wilson-Walker model
n
t

) 1 ( a r +
t n rel rel
V V

= |
) 1 ( cos a F V
w

n
t

) 1 ( a r +
t n rel rel
V V

= |
) 1 ( cos aF V
w

Appendix A-Blade Element Momentum Theory 189
A.3.4. Prandtl Tip and Hub loss factors
Among some theories for estimating the tip and hub loss factors, Prandtl theory is simple
and efficient and also gives acceptable results. In Prandtl theory tip and hub loss factors are
defined as
{ } ) exp( cos
2
1
tip tip
f F =

if 7
tip
f (A.41.a)
1 =
tip
F if ) 5 . 0 ( 5 . 0
85 . 0
+ =
tip
new
tip
F F (A.41.b)
Where
sin 2
) (
r
r R B
f
tip

= (A.41.c)
and
{ } ) exp( cos
2
1
hub hub
f F =

if 7
hub
f (A.42.a)
1 =
hub
F if 7 >
hub
f (A.42.b)
Where
sin 2
) (
hub
hub
hub
R
R r B
f

= (A.42.c)
In the above equations, B is the number of blades, R and
hub
R are rotor and hub radii and
is the inflow angle from Equation (A.16).
A.3.4.1. Xu modification
An improvement to Prandtls tip loss factor has presented by Xu (2001) as follows
a. Pre-stalled condition (
s
< ):
) 5 . 0 ( 5 . 0
85 . 0
+ =
tip
new
tip
F F if 1 / 7 . 0 R r (A.43.a)
Appendix A-Blade Element Momentum Theory 190
R
F r
F
R r tip new
tip
7 . 0
) | 1 (
1
7 . 0 / =

= if 7 . 0 / < R r (A.43.b)
b. Post-stalled condition (
s
):
8 . 0 =
new
tip
F if 1 / 8 . 0 R r (A.43.c)
1 =
new
tip
F if 8 . 0 / < R r (A.43.d)
A.3.5. Heavy loading (High axial induction factor)
Momentum theory predicts a parabolic variation for thrust coefficient with a maximum
value of 1 at 5 . 0 = a , while the experimental data show that
T
C keeps increasing for
5 . 0 > a . For small axial induction factors, 4 . 0 0 < <
c
a a , known as light loading state,
predicted thrust coefficient by the momentum theory is in a good agreement with the
experimental data. In the case of heavy loading state, where
c
a a > , predicted
T
C departs
dramatically from its actual value. In the extreme loading situation, 1 = a , wind turbine acts
as a drag driven device with a thrust coefficient of 2 ) (
max ,
= =
Drag T T
C C rather than
0 =
T
C as predicted by Equation (A.25). Extrapolating Equation (A.25), with a maximum
value of 2 =
T
C at 1 = a , predicts reasonable values for
T
C . Separating light and heavy
loading state, Equations (A.35) and (A.38) can be re-written as follows
a. Classical model
) 1 ( 4 aF Fa C
T
= if
c
a a (A.44.a)
2 1
2
0
A a A a A C
T
+ + = if
c
a a > (A.44.b)
where
2
2
0
) 1 (
) 2 ( 4 4 2
c
c c
a
a a F F
A

+
= (A.44.c)
2
2 2
1
) 1 (
8 ) 1 ( 4 4
c
c c c
a
a F a F a
A

+ +
= (A.44.d)
Appendix A-Blade Element Momentum Theory 191
2
2 2 2 2
2
) 1 (
4 4 2
c
c c c
a
a F Fa a
A

+
= (A.44.e)
b. Wilson-Walker model
) 1 ( 4 a Fa C
T
= if
c
a a (A.45.a)
2 1
2
0
B a B a B C
T
+ + = if
c
a a > (A.45.b)
where
F
a
B
c
4
) 1 (
2
2
0

= (A.45.c)
F
a
a
B
c
c
4
) 1 (
4
2
1
+

= (A.45.d)
2
2
) 1 (
2 4
2
c
c
a
a
B


+ = (A.45.e)
A.4. Blade Element Force Analysis
Figure (A.10) shows a blade segment (element) subjected to the aerodynamic forces in the
same system of coordinates as introduced in Figure (A.1). Assuming 2-dimentional flow
on the aerofoil and neglecting radial forces on the blade ( 0 =
s
dF ), thrust force on the
element can be obtained as cos
n
dF dT = or
( ) cos sin cos dD dL dT + = (A.46)
Lift and drag coefficients are defined as
( )
e
t n
rel
L
dA V
dL
C

=
2
2
1

(A.47)
( )
e
t n
rel
D
dA V
dD
C

=
2
2
1

(A.48)

Appendix A-Blade Element Momentum Theory 192

Figure A.10- Blade element force analysis
where ( )
t n rel
V

is the relative velocity in the t n plane (see Figures (A.2) and (A.3)) and
cos cdr cds dA
e
= = is the element area. Combining Equations (A.46), (A.47) and (A.48)
gives thrust force on a blade element as
( ) ( )dr C C V c dT
D L t n rel
sin cos
2
1
2
+ =

(A.49)
and for a turbine with B blades it becomes
( ) ( )dr C C V Bc dT
D L t n rel
sin cos
2
1
2
+ =

(A.50.a)
or in terms of span coordinate
( ) ( ) cos sin cos
2
1
2
ds C C V Bc dT
D L t n rel
+ =

(A.50.b)
Using Equations (A.24) and (A.17), thrust coefficient can be written as
( )


2
2 2
sin
sin cos ) 1 ( cos
D L r
T
C C a
C
+
= (A.51)
where
r
, local solidity ratio, is defined as

cos 2 2 s
Bc
r
Bc
r
= = (A.52)
Rotor Plane
Element
Blade Span

dD
t
dF
dL
n
dF
n
s
dF

dT
s
t
n
dF

x
r
Appendix A-Blade Element Momentum Theory 193
Aerodynamic forces on the blade element also produce a torque about the rotor axis equal
to
t x
rdF dM = (Figure (A.10)). Recalling Equations (A.47) and (A.48), for a turbine
with B blades the generated torque about the rotor axis can be expressed as
( ) ( ) cos cos sin
2
1
2
rdr C C V Bc dM
D L t n rel x
=

(A.53.a)
or in terms of span coordinate
( ) ( ) cos cos sin
2
1
2
sds C C V Bc dM
D L t n rel x
=

(A.53.b)
Inserting the above result into the definition of the torque coefficient
M
C , Equation (A.29),
yields to
( )


2
2 3
sin
cos sin ) 1 ( cos
D L r
M
C C a
C

= (A.54)
A.4.1. Blade Aerodynamic Characteristics
Lift and drag coefficients are functions of the angle of attack and Reynolds number. Angle
of attack is in turn a function of the velocity field and the blade geometry and can be
expressed as
+ + = pitch
e 0
(A.55)
In the above equation is the inflow angle, (Equation (A.16)),
e
and
0
stand for the
blade elastic twist and pre-twist, pitch is the blade pitch angle and accounts for the
angle of attack corrections as given by the following equation.
m c
+ = (A.56)
In Equation (A.56)
c
refers to cascade correction and
m
refers to the other corrections.
A.4.1.1. Cascade Correction,
c

Cascade correction to the angle of attack has two components
2 1
+ =
c
(A.57)
Appendix A-Blade Element Momentum Theory 194
where
1
accounts for the effect of finite aerofoil thickness and
2
accounts for the
effect of finite aerofoil width
rc
A B
a

2
cos
0
1
= (A.58)
)
`

=

R
r a
R a
r a ) 1 (
tan
) 2 1 (
) 1 (
tan
4
1
1 1
2
(A.59)
0
is the inflow angle prior to rotational induction, ( 0 = a in Figure (A.3)) and
a
A is the
aerofoil cross section area, normally taken as
max
68 . 0 ct A
a
, where
max
t is the maximum
thickness of the aerofoil.
A.4.1.2. Effect of Finite Aspect Ratio, Pre-Stall Condition
The effect of finite aspect ratio at small angle of attacks can be considered as either
Estimating the finite aerofoil data by applying Lanchester-Pradtl theory to 2-D data
L L
C C = (A.60)
AR
C
C C
L
D D

2
+ = (A.61)
AR
C
L

+ = (A.62)
where
L
C ,
D
C and are infinite aerofoil data and AR stands for the aspect ratio, or using
2-D data and applying a tip loss factor to the HAWT aerodynamic model.
A.4.1.3. Effect of Finite Aspect Ratio, Post-Stall Condition
For larger angle of attacks, two models can be used to estimate lift and drag coefficients of
an aerofoil.
a. Viterna-Corrigan model
Viterna-Corrigan model (Viterna and Corrigan, 1981) uses only three values from the 2-
Dim data to estimate lift and drag coefficients in post-stall condition, 90 < <
s
.
Appendix A-Blade Element Momentum Theory 195


sin
cos
cos sin
2
max , L D L
K C C + = (A.63)
cos sin
2
max , D D D
K C C + = (A.64)
where
( )
s
s
s s D stall L L
C C K


2
max , ,
cos
sin
cos sin = (A.65)
s
s D stall D
D
C C
K


cos
sin
2
max , ,

= (A.66)
AR C
D
018 . 11 . 1
max ,
+ = if 50 AR (A.67.a)
01 . 2
max ,
=
D
C if 50 > AR (A.67.b)
For other angle of attacks in the range of [-180, 180] reflected values for
D
C and reflected-
reduced values for
L
C can be used, (see Figure (A.11)).
b. Flat-Plate Theory
Flat-plate theory estimates post-stall aerodynamic characteristics of an aerofoil as
cos
N L
C C = (A.68)
sin
N D
C C = (A.69)
where

+
= 98 . 1 ,
sin
238 . 0
222 . 0
1
min

N
C if 1 F (A.70.a)
|

\
|
=
F
F
C
N
2
) 1 ( 22 . 12
tanh 81 . 0 98 . 1 if 1 5 . 0 < < F (A.70.b)
Appendix A-Blade Element Momentum Theory 196
17 . 1 =
N
C if 5 . 0 F (A.70.c)

-1
-0.5
0
0.5
1
1.5
-180 -150 -120 -90 -60 -30 0 30 60 90 120 150 180
Angle of attack
A D C B C D
Drag
Lift
A: Available data
B: Viterna model
C: Viterna model with reflection and reduction
D: Interpolation between limits

Figure A.11- Viterna post-stall model
A.5. Blade Element Momentum Theory, BEMT
Equating thrust and torque coefficients obtained from the blade element force analysis
(with the assumption of zero drag force) and those obtained from momentum theory is the
base of the BEMT. Neglecting drag force in Equations (A.51) and (A.54), thrust and torque
coefficients will become


2
2 2
sin
cos ) 1 ( cos
0
L r
T
C a
C

= (A.71)


2
2 2
sin
sin ) 1 ( cos
0
L r
M
C a
C

= (A.72)
Depending on which brake state model in the momentum theory is used, formulation for
BEMT will be different.
a. Classical Model
Combining Equations (A.71) and (A.72) with equations (A.44) and (A.36) gives
Appendix A-Blade Element Momentum Theory 197
2
1 1
0
T
C
a

= if 96 . 0 4 . 0
0

T
C a (A.73.a)
0
2 0
2
1 1
2
) ( 4
0
A
C A A A A
a
T
+
= if 96 . 0 4 . 0
0
> >
T
C a (A.73.b)


cos
tan
r
aF
a = (A.74)
b. Wilson-Walker Model
Combining Equations (A.71) and (A.72) with Equations (A.45) and (A.39) gives:
2
1 1
0
F
C
a
T

= if F C a
T
96 . 0 4 . 0
0
(A.75.a)
0
2 0
2
1 1
2
) ( 4
0
B
C B B B B
a
T
+
= if F C a
T
96 . 0 4 . 0
0
> > (A.75.b)


cos
tan
r
a
a = (A.76)
6 Equations (A.16), (A.31), (A.55), (A.71), (A.73)/(A.75), (A.74)/(A.76) and two set of
tabulated data for
L
C and
D
C can be solved to find a , a , F , , ,
L
C ,
D
C and
0
T
C .
Knowing a , ,
L
C and
D
C one can use Equations (A.50) and (A.51) to calculate T and
T
C and Equations (A.53) and (A.54) to find M and
M
C . Having rotor torque M , turbine
mechanical power, P can be easily calculated by
M dM P = =

(A.77)
and the power coefficient
P
C can be determined from the following equation.
2 3 3
2
1
2
1
R V
P
A V
P
C
w rotor w
P

= = (A.78)
Appendix A-Blade Element Momentum Theory 198
A.6. Flap bending
Referring to Figure (A.12), tangential and normal force increments on the blade element at
location ,
t
dF and
n
dF , are given by Equation (A.79).
( )






.
2
. .
sin cos
cos sin
2
1
sin cos
cos sin
at
D
L
t n rel
at at
n
t
C
C
d V c
dD
dL
dF
dF
|
|

\
|
)
`


=
|
|

\
|
)
`


=
)
`


(A.79)

Figure A.12- Blade element force analysis for flap bending calculation
Flap-bending moment produced by these forces at location s can be calculated as:
( ) ( ) ( ) ( )

=
=
|

\
|
+ =
s
R
n t s at FB
s
R
dF dF M


cos
0
.
cos
sin cos (A.80)
In the above equation, |

\
|
s
R

cos
is the moment arm and is the angle between the
normal axis and the chord line at s - location which can be expresses in terms of the inflow
and attack angles as:

s at.
) ( = (A.81)
r
Blade
element

s

x
R

dD
dL
n
t
Chord line at
s -location

n
dF
t
dF
Appendix A-Blade Element Momentum Theory 199
A.7. References
Tulapurkara, E.G., 1996. Turbulence Models for the Computation of Flow past Airplanes.
Progress in Aerospace Sciences, 33, pp 71-165.
Viterna, L.A., Corrigan, R.D., 1981. Fixed Pitch Rotor Performance of Large Horizontal
Axis Wind Turbines. In: DOE/NASA workshop on Large HAWT's, Cleveland, Ohio, 1981.
Xu, G., 2001. Computational Studies of Horizontal Axis Wind Turbines. Doctoral Thesis,
Georgia Institute of Technology.

You might also like