Sunteți pe pagina 1din 86

2D Investigation of the Lift & Drag

Characteristics of a NACA 632215


Aerofoil
September 2nd - December 18th 2008

Appendix Report
Group TEE1730
TEE1
Board of Studies of Energy
Aalborg University

Authors:
Lars Christian Riis Johansen

Rasmus Larsen Mosbk

Jakob Rabjerg Vang

Copies: 5
Number Pages:

Supervisor:
Matthias Mand

Abstract
This report concerns 2D investigation of the lift and drag characteristics of a NACA
632 215 aerofoil. An analytical solution for the lift coefficient is estimated as a function of the angle of incidence by assuming potential flow and thin aerofoil theory. The
numerical solution is obtained by using the Computational Fluid Dynamics software
Fluent 6.3 at following Reynolds number, Rec = 1.3315 105 (20 [ m
s ]). The lift and the
drag coefficients are calculated for different angles of incidence using the SST k
turbulence model. The experimental solution carried out by using strain gauge measurements for determine the lift and drag coefficient at varying angle of incidence. Lift
and drag coefficients are measured at Reynolds numbers, Rec = 6.8794 104 (10 [ m
s ]),
5 (20 [ m ]). An analysis of the measureRec = 1.0128 105 (15 [ m
])
and
Re
=
1.3315

10
c
s
s
ment is carried out by using uncertainty analysis of the results in order to verify the
data. Comparisons between the analytical, numerical and the experimental solution
are made in order to see if the results deviates from each other. The results showed
good agrement between the numerical solution and the experimental data, but there is
a large uncertainty on the experimental measurements at high angle of incidence. Flow
visualisation of the flow around the aerofoil is carried out by using string and smoke.

ii

Contents
Abstract

Nomenclature

ii

1 Problem Formulation

1.1

Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2 Determining CL from thin aerofoil theory

1
3

2.1

Potential Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2

Lifting Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.3

The Kutta Condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.4

Thin Aerofoil Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.5

Determining CL For the NACA 632 215 Aerofoil . . . . . . . . . . . . .

2.6

Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

12

3 Numerical Solution

13

3.1

Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

13

3.2

Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

14

3.3

Turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

18

3.4

Creation of the mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

25

3.5

Setup the case file . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

29

3.6

Grid Independence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

30

3.7

Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

37

4 Experiments
4.1
4.2
4.3

49

Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying


Reynolds Numbers and Angles of Incidence . . . . . . . . . . . . . . . .

49

Flow Visualisation of a NACA 632 215 Aerofoil at varying Angles of


Incidence Using String Attachments . . . . . . . . . . . . . . . . . . . .

70

Flow Visualisation of a NACA 632 215 Aerofoil at Varying Angles of


Incidence Using Smoke . . . . . . . . . . . . . . . . . . . . . . . . . . . .

72

iii

CONTENTS

iv

5 Comparison and Discussion of Results

75

6 Conclusion

79

Bibliography

80

Chapter 1

Problem Formulation
In this section the formal problem formulation for the project is given. The problem
formulation serves as the basis for the work presented in this report. The following
problem formulation has been chosen:
How do the coefficients of lift and drag of the NACA 632 215 aerofoil vary as a function
of the angle of attack.

1.1

Objective

The objective of this project is to develop curves for lift and drag characteristics of the
NACA 632 215 aerofoil. The curves are developed using three approaches. First an
analytical solution is obtained from potential flow theory. Then a numerical solution is
obtained using the FLUENT 6.3.26 CFD software. Finally wind tunnel measurements
are performed to validate the computational results. Discrepancies between the results
obtained are discussed.

1.1.1

Analytical Investigation

To obtain an analytical solution a Matlab script is developed. The script incorporates


lift calculations based on potential flow assumptions and thin aerofoil theory. All
calculations are performed assuming two dimensional flow.

1.1.2

Numerical Investigation

The numerical solution is to be obtained using FLUENT 6.3.26. A 2D geometry is


developed in AutoDesk Inventor. The geometry is imported into GAMBIT and a grid
is created. The solution obtained should be grid independent. This is checked by
solving the model with different grid refinements and comparing the results. When
grid independence is obtained the flowfield over the aerofoil is solved at different angles of incidence in order to obtain lift and drag characteristics. The development of
recirculation zones is also investigated.
1

1.1. Objective

1.1.3

Experimental Investigation

To validate the results obtained from the analytical and computational approaches the
coefficients of lift and drag are determined experimentally. A NACA 632 215 aerofoil is
mounted in a wind tunnel and the lift and drag characteristics are measured using strain
gauges. The uncertainties of the measurements are evaluated using uncertainty propagation. In addition to the quantitative investigation flow visualisation is performed
using smoke and strings taped to the wing.

Chapter 2

Determining CL from thin


aerofoil theory
This section deals with the determination of the lift generated by the the NACA 632
215 aerofoil through an analytical analysis. The calculation of the lift coefficient, CL ,
of the aerofoil is presented. The section contains a presentation the main assumptions,
derivation of the necessary equations and calculation of the lift coefficient. The source
for this section is [Anderson, 2001].

2.1

Potential Flow

The first assumption to be made in order to calculate CL is that of potential flow. This
implies that the flow is assumed to be inviscid, incompressible and irrotational. For twodimensional potential flow there exists a stream function [m2 /s] which describes the
flow. The stream function is constant along a streamline. The local velocity components
can be described from as:

uy =
x

ux =

(2.1)
(2.2)

Or in polar coordinates:

1
r

u =
r
ur =

(2.3)
(2.4)

satisfies the Laplace equation [Anderson, 2001]:


2 2
+
=0
x2
y 2
3

(2.5)

2.2. Lifting Flow

In potential flow complex flows can be modelled by superposition of elementary flows.


These are the source, the sink, the doublet, the vortex and the uniform flow. The
doublet is obtained by combining a source and a sink at the same point. Streamlines
for the different flows are shown in figure 2.1

Figure 2.1: Streamlines of elementary potential flows. The arrows denote the flow
direction. A: Vortex, B: Source, C: Sink, D: Doublet and E: Uniform flow.
Expressions for the stream function of the different flows are gives as in equations 2.62.9. The expressions are given in polar coordinates.

Vortex:
=

ln(r)
2

(2.6)

(2.7)

Source/sink:
=
Doublet:
sin()
2 r

(2.8)

= U r sin()

(2.9)

=
Uniform flow:

Here [m2 /s] is the vortex strength, [m2 /s] is the doublet strength, U [m/s] is the
the speed of the undisturbed stream and [m2 /s] is the source/sink strength. If > 0
Equation 2.7 represents the stream function of a source. If < 0 it represents the
stream function of a sink.

2.2

Lifting Flow

This section introduces the Kutta-Joukowski theorem which establishes the coupling
between circulation and lift generation.
As mentioned in section 2.1 the flow around objects can be simulated as a superposition
of elementary flows. An example of this is a cylinder, which can be simulated by a
doublet in a uniform flow. Adding the stream functions of the uniform flow and the
doublet we obtain:


sin()

= U r sin()
= U r sin() 1
2 r
2U r2

(2.10)

2. Determining CL from thin aerofoil theory

This is the stream function of a uniform flow over a cylinder of radius R =

2U .

Substituting R in equation 2.10 we get:




R2
= U r sin() 1 2
r

(2.11)

This flow situation is shown in figure 2.2. Notice the stagnation points at a and b. The
circular streamline corresponds to = 0

Figure 2.2: Streamlines of potential flows over a cylinder. The arrows denote the flow
direction.
The net force on a cylinder in a uniform potential flow is zero. This is due to the
assumption of zero viscosity. A lifting force can, however, be obtained by spinning the
cylinder. This is accomplished by superposition of a vortex onto the cylinder. Since
the value of the stream function is determined within an arbitrary constant [Anderson,
2001] the stream function for a vortex can be rewritten by adding a constant of the

value 2
ln(R). Thus equation 2.6 becomes:
=

r

ln(r)
ln(R) =
ln
2
2
2
R

(2.12)

The stream function for the spinning cylinder then becomes:


R2
= U r sin() 1 2
r



+

r

ln
2
R

(2.13)

The flow field around a spinning cylinder is show in figure 2.3. The stagnation points
a and b move towards the bottom of the cylinder as is increased.
The velocity components on the circumference can be obtained from Equation 2.13 by
applying Equations 2.3 and 2.4 and setting r = R. This yields:

ur = 0
u = 2U sin()

(2.14)

2R

(2.15)

The pressure distribution on the circumference of the circle is then determined from
Bernoullis Equation. If the elevation term is neglected we get:

2.3. The Kutta Condition

Figure 2.3: Streamlines of potential flows over a spinning cylinder.

p +

2
u2
U
= p() +
2
2

p = p() p

2 u2 )
(U

=
2

(2.16)

Where p [P a] is the deviation of the local static pressure from the upstream static
pressure. The lift per unit span is obtained by integrating p in the x-direction:

plower dx

L =
R

pupper dx
R

= R

plower sin()d + R

Z 2

= R

plower sin()d

(2.17)

p sin()d = U
0

Where subscripts upper and lower denote the upper and lower surface of the circle
respectively. L0 [N/m] is the lift per unit span.
Thus the lift generated by a spinning cylinder is directly proportional to the circulation.
Also the lift coefficient is given as
CL =

2
cU

(2.18)

where c = 2R is the chord length.


L0 = U is the Kutta-Joukowski theorem, which is used as the basis for determining
the lift generated by an aerofoil.

2.3

The Kutta Condition

An aerofoil generating lift can be modelled in a manner similar to that of the spinning
cylinder. The shape is different from that of the cylinder but equation 2.17 still holds.

2. Determining CL from thin aerofoil theory

The problem is how to choose the value of the circulation. The selected value must
result in physically meaningful flow around the aerofoil. Figure 2.4 shows the flow over
an aerofoil for two different values of . As with the case of the cylinder the stagnation
points a and b move as the value of is varied.

Figure 2.4: Streamlines of potential flow over an aerofoil at different values of circulation.
Both flow situations in figure 2.4 are valid potential flows. However only situation 2 is
physically meaningful. This is due to the fact that the flow velocity at the trailing edge
becomes infinite in situation 1. This means that the flow must always leave the trailing
edge smoothly as depicted in situation 2. This is known as the Kutta condition. This
condition ensures that there is only a single valid value of for a given flow over an
aerofoil.

2.4

Thin Aerofoil Theory

To obtain an analytical solution for the lift of a given aerofoil a number of simplification
must be made. Firstly it is assumed that the aerofoil is so thin that the thickness can
be neglected. Thus the aerofoil is approximated by a vortex sheet of strength (s) [m/s]
along the camber line. In order to satisfy the Kutta condition is chosen so that the
camber line becomes a streamline of the flow. For the camber line to be a streamline
the velocity across the camber line must be 0. Thus:
U,n + U,n (s) = 0

(2.19)

Where U,n [m/s] is the component of the free stream velocity normal to the camber
line and U,n (s) [m/s] is the induced velocity normal to the camber line. For small
angles and a camber line that only deviates slightly from the x-axis equation 2.19 can
be formulated as in Equation 2.20:

 Z c
dy
()d
U

=0
dx
0 x

(2.20)

dy
Where [rad] is the angle of attack, dx
is the slope of the camber line, c [m] is the
chord length and is a dummy variable of integration along the x-axis.

In order to determine an expression for the following transformation is made:

2.5. Determining CL For the NACA 632 215 Aerofoil

c(1 cos )
0
2
c(1 cos )
x=
0
2
=

(2.21)
(2.22)

Equation 2.20 then becomes:




Z
1
() sin()d
dy

=0
U
dx
2 0 cos() cos()

(2.23)

Solving for equation 2.24 is obtained:

1 + cos() X
A0
+
An sin(n)
sin()

() = 2U

!
(2.24)

n=1

A0 and An are found to be:


Z
1 dy
d
A0 =
0 dx
Z
2 dy
An =
cos(n)d
0 dx

(2.25)
(2.26)

The total circulation is given by:


Z
=
0

c
()d =
2

() sin()d

(2.27)

By inserting the expression from equation 2.24 and simplifying the following expression
is obtained:


= cU A0 + A1
2

(2.28)

By inserting equations 2.28, 2.25 and 2.26 in equation 2.18 the lift coefficient is then
determined to be:


Z
1 dy
(cos() 1)d
CL = 2 +
0 dx

(2.29)

The remaining task is to evaluate equation 2.29 for the aerofoil to be investigated.

2.5

Determining CL For the NACA 632 215 Aerofoil

Now the lift coefficient for the NACA 632 215 is determined from the equations developed in the previous sections. The geometry of the aerofoil is given by the coordinates

2. Determining CL from thin aerofoil theory

NACA 632 215 Coordinate Sets


Upper
n
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26

x/c
0.00000
0.00399
0.00637
0.01120
0.02348
0.04829
0.07323
0.09823
0.14834
0.19852
0.24875
0.29900
0.34926
0.39952
0.44977
0.50000
0.55019
0.60035
0.65047
0.70053
0.75055
0.80051
0.85043
0.90030
0.95014
1.00000

Lower

y/c
0.00000
0.01250
0.01528
0.01980
0.02792
0.03960
0.04847
0.05569
0.06682
0.07487
0.08049
0.08392
0.08530
0.08457
0.08194
0.07768
0.07203
0.06524
0.05751
0.04906
0.04014
0.03105
0.02213
0.01368
0.00616
0.00000

x/c
0.00000
0.00601
0.00863
0.01380
0.02652
0.05171
0.07677
0.10177
0.15166
0.20148
0.25125
0.30100
0.35074
0.40048
0.45023
0.50000
0.54981
0.59965
0.64953
0.69947
0.74945
0.79949
0.84957
0.89970
0.94986
1.00000

y/c
0.00000
-0.01150
-0.01388
-0.01766
-0.02420
-0.03328
-0.03999
-0.04535
-0.05336
-0.05895
-0.06259
-0.06448
-0.06470
-0.06315
-0.06004
-0.05562
-0.05013
-0.04382
-0.03691
-0.02962
-0.02224
-0.01513
-0.00867
-0.00334
0.00016
0.00000

Camber Line
x/c
0.00000
0.00500
0.00750
0.01250
0.02500
0.05000
0.07500
0.10000
0.15000
0.20000
0.25000
0.30000
0.35000
0.40000
0.45000
0.50000
0.55000
0.60000
0.65000
0.70000
0.75000
0.80000
0.85000
0.90000
0.95000
1.00000

y/c
0.00000
0.00050
0.00070
0.00107
0.00186
0.00316
0.00424
0.00517
0.00673
0.00796
0.00895
0.00972
0.01030
0.01071
0.01095
0.01103
0.01095
0.01071
0.01030
0.00972
0.00895
0.00796
0.00673
0.00517
0.00316
0.00000

Table 2.1: Coordinates determinig the geometry of the NACA 632 215 aerofoil [Abbott
and Von Doenhoff, 1960]
shown in Table 2.1. The coordinates for the camber line are obtained by taking average
of the x- and y-coordinates for the upper and the lower surfaces of the aerofoil.
The coordinates are plotted in Figure 2.5.
Since the camber line is only expressed in terms of discrete points an expression must
be obtained by interpolation. By using a cubic spline the expression becomes
y
x xn 3
x xn 2
x xn
= a1,n (
) + a2,n (
) + a3,n (
) + a4,n
cn
c
c
c
x
xn+1
xn
for

n = 1, 2, 3...25
c
c
c

(2.30)

where a1,n , a2,n , a3,n , a4,n are the coefficients presented in Table 2.2
Differentiating Equation 2.30 to obtain

d(y/c)
d(x/c) :




x xn 2
x xn
+ 2a2,n
+ a3,n
c
c
xn
x
xn+1
for

n = 1, 2, 3...25
c
c
c

d(y/c)
= 3a1,n
d(x/c) n

(2.31)

d(y/c)
In order to be able to insert d(x/c)
in Equation 2.29 we must make it a function of by
performing the transformation given in Equation 2.22. Equation 2.31 then becomes:



d(y/c)
1 cos() xn 2
1 cos() xn
= 3a1,n

+ 2a2,n (

) + a3,n
d(x/c) n
2
c
2
c
xn
1 cos()
xn+1
for

n = 1, 2, 3...25
c
2
c

(2.32)

2.5. Determining CL For the NACA 632 215 Aerofoil

10

Plot of NACA 632215 Coordinates


Upper Surface
Lower Surface
Camber Line

0.3
0.2

y/c

0.1
0
0.1
0.2
0.3
0

0.1

0.2

0.3

0.4

0.5
x/c

0.6

0.7

0.8

0.9

Figure 2.5: Plot of the NACA 632 215 coordinates


d(y/c)
is given by a function consisting of 25 segments each the integral of each
Since d(x/c)
section must be summed. With n obtained through Equation 2.33.

= cos

2x
1
c


(2.33)

a simple linear expression for CL is obtained by Equation 2.29:


!
25 Z
1 X n+1 dy
CL = 2 +
(cos() 1)d

dx
n=1 n


25 Z
1 X n+1
1 cos() xn 2
= 2 +
3a1,n

2
c
n=1 n




1 cos() xn
+2a2,n

+ a3,n (cos() 1)d


2
c


1
= 2 + 0.0288

(2.34)

This expression is plotted in figure 2.6 together with experimental data from [Abbott
and Von Doenhoff, 1960]. The data is for Rec = 3106 . As it can be seen the agreement
between the analytical solution and the experimental data is good for modest angles
of attack. However the thin aerofoil is not valid when the angle of attack becomes
too large. Around = 10 the effects of flow separation due to non-zero viscosity can
be seen in the experimental results. This represents itself as a reduction in the slope
of the CL curve. At larger angles of attack the discrepancy becomes larger and the
thin aerofoil theory eventually breaks down around = 15 deg which is the stall point
corresponding to Rec = 3 106 .

11

2. Determining CL from thin aerofoil theory

Spline Coefficients
n

a1

a2

a3

a4

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25

207.31284
207.31284
-10.11032
12.72396
1.34365
0.64529
0.55519
0.05583
0.17024
-0.01678
0.05690
0.02920
-0.01369
0.02555
-0.00851
0.00850
-0.02549
0.01348
-0.02841
-0.05985
0.02781
-0.21139
0.09773
-1.13955
-1.13955

-5.25808
-2.14838
-0.59354
-0.74519
-0.26804
-0.16727
-0.11887
-0.07723
-0.06886
-0.04332
-0.04584
-0.03731
-0.03293
-0.03498
-0.03115
-0.03243
-0.03115
-0.03497
-0.03295
-0.03721
-0.04619
-0.04202
-0.07373
-0.05907
-0.23000

0.12111
0.08408
0.07722
0.07053
0.05786
0.04698
0.03982
0.03492
0.02762
0.02201
0.01755
0.01339
0.00988
0.00649
0.00318
0.00000
-0.00318
-0.00648
-0.00988
-0.01339
-0.01756
-0.02197
-0.02776
-0.03440
-0.04885

0.00000
0.00050
0.00070
0.00107
0.00186
0.00316
0.00424
0.00517
0.00673
0.00796
0.00895
0.00972
0.01030
0.01071
0.01095
0.01103
0.01095
0.01071
0.01030
0.00972
0.00895
0.00796
0.00673
0.00517
0.00316

Table 2.2: Spline coefficients for interpolating the coordinates of the NACA 632 215
camberline

Figure 2.6: Plot of the Analytical Solution for CL as a function of compared to actual
experimental data at Rec = 3 106 [Abbott and Von Doenhoff, 1960]

2.6. Summary

2.6

12

Summary

In this section the thin aerofoil theory for estimating the lift characteristics for an
aerofoil has been presented. It is concluded that the analytical solution reflects the
data presented in the literature well for modest angles of attack. However it breaks
down around the stall point.

Chapter 3

Numerical Solution
In this chapter the numerical investigation of the lift and drag characteristics of the
NACA 632 215 aerofoil is presented. The chapter begins with a review of the governing
equations of incompressible, 2D fluid flow. Then the chosen schemes for discretization
of the governing equations are presented. This is followed by the selection of a proper
turbulence model. The meshing of the aerofoil as well as the grid independence analysis
are presented and evaluated. Eventually the results obtained are presented. The sources
for the theoretical part of the chapter are [Yin, 2008], [Fluent Inc., 2006] and [Versteeg
and Malalasekra, 2007].

3.1

Governing Equations

In this section the governing equations for the problem at hand. Since this project
concerns steady flow around an aerofoil at low Mach numbers the only equations of
interest are the Continuity Equation and the Momentum Equations. All these equations
can be divided into four terms and written in the form given in Equation 3.1.
()
+ (U) = (K) + S
{z
} |{z}
| t
{z } | {z } |

Transient

Convective

Diffusion

(3.1)

Source

T
where is a flow variable, = [ x
, y
, z
] is the divergence vector and K is the
diffusion coefficient. Equation 3.1 is called the general transport equation.

Since the flow is steady and incompressible the transient term is zero and is constant.
(U) = (K) + S
| {z } |
{z
} |{z}
Convective

Diffusion

(3.2)

Source

For the continuity equation = 1, K = 1 and S = 0. Equation 3.2 then becomes:

U=0

(3.3)

For the momentum equations in x- and y-directions we have the terms given in Table
3.1.
13

3.2. Discretization

14

x
y

ux
uy

p
+ SM,x
x
p
y + SM,y

Table 3.1: Variables for the Momentum Equations in x- and y-directions


This gives the Momentum Equations for 2D flow:
(ux U) = (ux )

p
+ SM,x
x

(3.4)

(uy U) = (uy )

p
+ SM,y
y

(3.5)

Equations 3.3, 3.4 and 3.5 must be solved for the flow field in orde to obtain the pressure
distribution on the aerofoil from which the values of CL and CD can be calculated. In
order for the computer to be able to do this the equations must be formulated as a set
of algebraic equations which the computer is able to solve iteratively.

3.2

Discretization

Since computers can only solve algebraic equations the governing equations are reduced
to algebraic form by the means of discretization. The entire domain is divided up into
a number of control volumes as shown in Figure 3.1. The node corresponding to a
control volume is denoted by C. The neighbouring nodes are denoted North (N), South
(S), East (E) and West (W). Lower case letters (n, s, e and w) denote centres of the
surfaces between the neighbouring cells.
The target is to reduce Equation 3.2 to an equation of the form in Equation 3.6 in
order to determine the value of at C from the values at the neighbouring nodes.
aC C =

anb nb

(3.6)

nb

This is done through three levels of approximations.


The procedure starts by integrating the general transport equation over the control
volume.
Z

Z
(U) dV =

CV

Z
(K) dV +

CV

S dV

(3.7)

CV

R
Here CV dV denotes integration over the control volume. For the convective and
diffusive terms the volume integrals are converted into surface integrals using the Gauss
Divergence Theorem.

XZ
k

Ak

(U) nk dA =

XZ
k

Ak

Z
(K) nk dA +

S dV
CV

(3.8)

15

3. Numerical Solution

NN

n
WW

EE

SS

Figure 3.1: Principle of a 2D grid for the finite volume method.


P R
Here k Ak dA denotes the summation of the integrals over each the control surfaces
bounding the control volume. nk denotes the normal vector of control surface k, where
k corresponds to n, s, e or w. Replacing all integrals by the product of average values
and the domain of integration yields:

X
k

(U) nk Ak =

(K) nk Ak + S VCV

(3.9)

Now no further simplification can be made without resorting to approximations. The


three levels of approximations necessary to discretize the general transport equation
are presented in the following sections.

3.2.1

First Level Approximation

The first approximation is that of determining the value of (U) nk . This is done by
mid-point approximation. This gives the result
(U) nk k uk

(3.10)

where k and uk are the value of and the normal velocity component at the centre
of control surface k. This approximation is second-order accurate.
Since Equation 3.6 does not involve the values at the control surface centres k must
be approximated from the values at the nodes. This can be done using a number of
different schemes. The relevant schemes available in Fluent 6.3 are summarised in the
following section.

3.2. Discretization

3.2.2

16

Second Level Approximation

This section summarises the schemes used for for approximating k . Figure 3.2 shows
the small section of the mesh to illustrate the notation used. k is the value at the
centre point of the face k. In this project two schemes are used. These are the Quick
scheme and the Second Order Upwind scheme.
QUICK
In order to get an accurate solution the QUICK scheme is applied for most of the
calculations. The QUICK scheme computes k by using the two upwind notes and the
downwind note. The version used in Fluent is given in Equation 3.11.

k,QU ICK


sup
sdown
=
up +
down
sup + sdown
sup + sdown


sup
sup,up + 2sup
up
up,up
+ (1 )
sup + sup,up
sup + sup,up

(3.11)

Here s [m] denotes the length of the cell. Subscripts correspond to the cells in Figure
3.2. The QUICK scheme is third order accurate [Yin, 2008]. However the solution may
become unstable at steep gradients [Versteeg and Malalasekra, 2007].
First Order Upwind
The Second Order Upwind (SOU) scheme approximates the value of k by Equation
3.12.
k,SOU = up + rup ()up

(3.12)

Here up and are the values at the upstream cell centre and rup is the vector from
the upstream cell centre to face k. The SOU scheme is less computationally expensive
than the QUICK scheme and more stable but may be less accurate.
Flow direction

k
up, up

up

down

Figure 3.2: Grid showing the notation for k schemes


Having approximated the cell face value in terms of the node values only one step
remains.

17

3. Numerical Solution

3.2.3

Third Level Approximation

The last value to be approximated is K nk from the diffusion term in Equation


3.9. The gradient at the control surface is obtained by midpoint approximation. This
gives:
(K) nk Kk ()k nk

(3.13)

Where Kk and ()k are taken at the centre of the control surface.
The value of ()k is obtained from the values at the neighbouring control volumes
by the means of central difference approximation.
()k =

()C + ()K
2

(3.14)

The problem then becomes that of approximating the control volume gradient. This is
done using the Green-Gauss theorem:
()C =

X
1
=
k Ak nk
VC

(3.15)

Here k denotes summation over all the control surfaces k bounding the control volume
C. VC denotes the volume of C and k denotes an average value of at control surface
k. This value is calculated either by the node based or the cell based approach. The
cell based approach is the most simple and computationally inexpensive. Here k is
calculated as the arithmetic mean of the values in the control volumes C and K sharing
face k.
K + C
k node =
2

(3.16)

The node based approach is more accurate and more computationally expensive. It
calculates k from the values at the nodes on the control surface.
k node =

nnodes

node

(3.17)

nodes

P
Where node is the value a node, nodes denotes summation over all the nodes on the
control surface. nnodes denotes the number of nodes on the surface, which is two for
two-dimensional problems. node is calculated by a weighted average of the values of
the control volumes that share the node. The expression is given in Equation 3.18.
node =

1 X
nCV

CV CV

(3.18)

CV

Here nCV denotes the number


Pof control volumes sharing the node. For two-dimensional
rectangular grids nCV = 4. CV denotes summation over all of the surrounding CVs.
CV is a weighting factor and CV is the value if for control volume CV.
Since the node based approach is the more accurate this is chosen for the calculations.

3.3. Turbulence

3.2.4

18

Summary

The discretization of the governing is performed through three levels of approximations.


The first level involves approximating the control surface value k . This is done by midpoint approximation.
The second level is approximation of at the control surface centre. This can be done
in several ways but for this work the QUICK and Second Order Upwind schemes are
used.
In the third level of approximation the choice to be made is which scheme to use when
approximating the control volume gradient ()C . This can be done using the node
based or the cell based approach. The node based approach is used because of the
higher accuracy.

3.3

Turbulence

In this section a turbulence model for use in the simulations is selected. The criteria
for selection are a trade-off between computational time and accuracy.
Generally there are three approaches to turbulence modelling in CFD [Versteeg and
Malalasekra, 2007]:
Reynold-Averaged Navier-Stokes Simulation (RANS)
Large eddie Simulation (LES)
Direct Numerical Simulation (DNS)
Reynolds-Averaged Navier-Stokes Simulation (RANS)
The RANS is focused of the mean flow and the effects of the turbulence on mean
flow properties. The Navier-Stokes equations are time average. Extra terms in the
transport equations appears as time-averaged or Reynolds-averaged. These extra terms
are modelled with turbulence models such as the k models and the k models.
[Versteeg and Malalasekra, 2007, p 65]
Large eddy Simulation (LES)
Large eddy Simulation (LES) is an intermediate form of turbulence model which tracks
the behaviour of the larger eddies. The model computes the turbulent flow by solving
the unsteady Navier-Stokes equations for the large eddies. The universal behaviour of
the small eddies is captured with a more compact model. The demand of computing
resources is fairly large compared to RANS. This CFD technique is usually used for
complex geometries. [Versteeg and Malalasekra, 2007, p 98]
Direct Numerical Simulation (DNS)
These simulations compute the mean flow and all turbulent velocity fluctuations. The
unsteady Navier-Stokes equations are solved with a sufficiently fine grid and suffi-

19

3. Numerical Solution

ciently small time steps to resolve the smallest turbulent eddies and the fastest fluctuations. These calculations are expensive in terms of computing resources. [Versteeg
and Malalasekra, 2007, p 110]
It is selected to use RANS Simulation since this approach is the least computationally
expensive.

3.3.1

Governing Equations for Reynolds Averaged Navier-Stokes

In the Reynolds Averaged Navier-Stokes equations the effects of turbulence are accounted for by modelling. For turbulent flow the velocity can be divided into components at shown in Equation 3.19.
ui = ui + u0i

(3.19)

where ui is the mean velocity and u0i is the fluctuating velocity and i = 1, 2 for the
directions. By substituting Equation 3.19 for the velocity components and taking the
time average we get the continuity equation given by Equation 3.20 and the momentum
equation given by Equation 3.21. Equation 3.20 and Equation 3.21 are called the
Reynolds Averaged Navier-Stokes equations [Fluent Inc., 2006].

+
(ui ) = 0
t
xi
 


uj

ui
2 ul

(ui ) =
+

+
ij
u0i u0j
+
t
xi xj
xj
xi
3 xl
xj

(3.20)

(3.21)

The Reynolds stresses, u0i u0j , in Equation 3.21 represent the effects of turbulence, and
has to be modelled separately in order to close Equation 3.21. The Reynolds stresses
can be modelled by the Boussinesq hypothesis given by Equation 3.22. The idea is to
introduce a turbulent viscosity t . The way t is determined depends on the chosen
turbulence model. [Fluent Inc., 2006].




uj
ui
2
uk
u0i u0j = t
+

k + t
ij
(3.22)
xj
xi
3
xk
Here k [m2 /s2 ] is the specific turbulent kinetic energy and ij = 1 for i = j ij = 0
for i 6= j. A list of different RANS turbulence models can be seen table 3.2. Some
of the turbulence models exist in different formulations. There exist several different
k and k models. In Fluent three different k models and two different k
models are available. The differences in the models is there formulation of the transport
equations.
Selection of the turbulence model
The turbulence model shall be selected to optimize the calculations of the Lift and
Drag coefficient. It is therefore important to select the right turbulence model. There
have been looked into different literature which has investigated turbulence models for
calculation of lift and drag coefficients for aerofoils. For external aerodynamics several
turbulence models including the Sparlart-Allmaras model, the standard k model

3.3. Turbulence

20

RANS Turbulence Models


Name of model
Mixing Length
Spalart-Allmaras
k
k
Algebraic Stress
Reynolds Stress

No. of extra transport equations


Zero
One
Two
Two
Two
Seven (3D)/Five (2D)

Table 3.2: Different RANS turbulence models with their number of transport equations
[Versteeg and Malalasekra, 2007, p 67]
and the SST k model are suitable. [Versteeg and Malalasekra, 2007]. The litterature
suggests that the SST-k model gives the best agreement with experimental results.
[Menter, 1992; Catalano and Amato, 2003]. With this in mind the SST k model is
selected.

3.3.2

The Shear Stress Transport k turbulence model

The Shear Stress Transport k model or SST k model is a hybrid model using
a transformation of the k model into a k model in the near-wall region and
the standard k model in the fully turbulent region far from the wall. The k in the
transport equation is the specific turbulent kinetic energy given by k = 21 (u02 + v02 + w02 )
where u0 , v 0 and w0 are the velocity fluctuations due to eddies for the x, y, and z
directions respectively. In the case of 2D flows the mean velocity in the z direction is
zero.
2

is the specific dissipation rate given by = k [ 1s ], where [ ms ] is the viscous dissipation


rate. The transport equation for k is given by equation 3.23


(k)
t
+ (kU ) = ( + )(k)
(3.23)
k
| t
{z } | {z }
|
{z
}
Convective
Transient

Diffusion

Gk
|{z}

Production

Yk
|{z}

Dissipation

The transport equation for is given by equation 3.24.




()
t
+ (U) = ( +
)()

| t
{z } | {z }
|
{z
}
Convective
Transient

Diffusion

G
|{z}

(3.24)

Production

Y
|{z}

Dissipation

The terms are explained below.

D
|{z}

Cross-diffusion

21

3. Numerical Solution

Transient terms
The first terms for the k and transport equations are the transient terms which is
the time rate of change of k or .

Convective terms
The second term is the convective term which governs the transport of k or by
convection.

Diffusion terms
The third term is the diffusion term which is transport of k or by turbulent diffusion.
Here is the viscosity, t is the eddy viscosity, given by Equation 3.25.


k 1
1 S F1
t =
,
max a1

(3.25)

Here S [ 1s ] is the strain rate magnitude, a1 is a model constant given by Table 3.3.

0 + Ret /Rk
1 + Ret /Rk


(3.26)

Ret is the turbulent Reynolds number given by Equation 3.27. is a coefficient that
damps the turbulent viscosity to obtain a low-Reynolds-number correction given by
, R , a constant given by Table 3.3. in Equation 3.28 is
Equation 3.26, here is
i
k
given by Equation 3.29 where i,1 and i,2 are constants given by Table 3.3.

i
0 =
3
i = F1 i,1 + (1 F1 )i,2

Ret =

(3.27)
(3.28)
(3.29)

For the diffusion term of the k transport equation given by Equation 3.23 k is given
by Equation 3.30. for the diffusion term of the transport equation is given by
Equation 3.31.
k =

1
F1 k,1 + (1 F1 )/k,2

(3.30)

1
F1 /,1 + (1 F1 )/,2

(3.31)

Here F 1 is given by Equation 3.32. Model constants k,1 , k,2 , ,1 and ,1 are given
by Table 3.3.

3.3. Turbulence

22

F1 = tanh(41 )
"
1 = min max
D+

(3.32)

k 500
,
0.09y y 2

!
,

4k
,2 D+ y 2



1 1 k
= max 2
, 1010
,2 xj xj

F2 = tanh(22 )
"
2 = max 2

#
(3.33)
(3.34)
(3.35)

k 500
,
0.09y y 2

#
(3.36)

Productive terms
The fourth term is the productive term which is the rate of production of k or . For
k is given by Equation 3.37 and the production for is given
k the production term, G
by Equation 3.38.


u
j

0
0
k = min u u
G
, 10 k
i j
xi
uj

G =
u0i u0j
t
xi


0 + Ret /Rk
=

1 + Ret /Rk

(3.37)
(3.38)
(3.39)

t is the kinematic turbulent viscosity. For the equation 3.39, Ret is given by Equation
3.27 and is given by Equation 3.26. R , 0 and are model constants given by
Table 3.3.
Dissipation terms
The fifth term is the dissipation term describes the rate of dissipation of k and omega.
For k it is given by Equation 3.40 and for by Equation 3.41.
Yk = k
2

Y = i


4/15 + (Ret /R )4

=
1 + (Ret /R )4

(3.40)
(3.41)
(3.42)
(3.43)

Here i is given by Equation 3.29.


The Cross-Diffusion Term
There is a sixth term for the transport equation. This is the Cross-Diffusion term in
Equation 3.24. Its value is given by Equation 3.44

23

3. Numerical Solution

D = 2(1 F1 ),2

1 k
xj xj

(3.44)

F1 is given by Equation 3.32 and ,2 is a model constant given by table 3.3.


The default model constants for the SST k model
=1

R = 2.95
= 0.09

= 0.52
k,1 = 1.176
i,1 = 0.075

= 1.5
,1 = 2
i,2 = 0.0828

R = 8
k,2 = 1.0
0 = 19

Rk = 6
,2 = 1.168
a1 = 0.31

Table 3.3: The default model constants for the SST k model in FLUENT 6.3.26
[Fluent Inc., 2006]

3.3.3

Boundary Layer Considerations

The value y + is the Reynolds based distance from the wall. This value is very important
factor for verifying calculations of flows. y + is given by Equation 3.45.
y+ = y

Here y [m] is the distance from the wall and u =

(3.45)
q

is the friction velocity [Yin,

2008]. is the wall shear stress.


We can also define a velocity at y dimensionless using the friction velocity as in Equation
3.46
u

u+ =
(3.46)
u
Viscous sub-layer
The viscous sub-layer is extremely thin, y + < 5, and we assume that the shear stress
= dd uy is constant and equal to the wall shear stress throughout the layer. In the
viscous sub-layer a linear relationship between the distance to the wall, y and the mean
velocity, u
given by Equation 3.47.
u
=

(3.47)

By making use of the definitions of u+ and y + this leads to equation 3.48 [Versteeg and
Malalasekra, 2007, p 58].
u+ = y +

(3.48)

Log-law layer
Outside the viscous layer a region exist where both the viscous and turbulent effects
are important. In this layer y + takes values between 30 and 500. In the Log-law layer

3.3. Turbulence

24

the relationship in Equation 3.49 is valid. Here is the shear stress and 0.4 is the
von Karmans constant. The additive constant is B ' 5.45. [Fluent Inc., 2006, p 58 to
59].
 y
1
u+ = ln(y + ) + B = 2.5ln u
+ 5.45
(3.49)

Buffer layer
Between the viscous layer and the log-law layer is a buffer layer. Here y + takes values
between 5 and 30. The mesh should be made either coarse or fine enough to prevent
the near wall cells from being placed in the buffer layer. [Fluent Inc., 2006, p 12-76]
Figure 3.3 shows the different layers for Near-Wall modelling.

Figure 3.3: The different y + layers. Modified from [Fluent Inc., 2006]

Demands for the SST k model


In order to resolve the boundary layer the y + values in the cells next to the wall
have to be in the order of the Viscous sub-layer which is y + < 5. The k models
use enhanced wall functions as the near-wall treatment. The enhanced wall functions
give a reasonable representation of the velocity profiles for y + values in the area of
3 < y + < 10, which corresponds to the border between the Viscous sub-layer and the
buffer layer. For the SST k model is preferable y + < 5, but y + < 10 is acceptable
[Fluent Inc., 2006, p12-84].

25

3.4

3. Numerical Solution

Creation of the mesh

The first step in generating the mesh is to draw the aerofoil in a CAD program and
the geometry is imported into Gambit. Four points called H, I, J and G are created to
generate four edges around the aerofoil. cf. 3.4. This is done to make a smooth mesh
around the aerofoil, since there must be more mesh points on highly curved surfaces
compared to straight lines.

+
+Gy

+Gz

Gx

Y
Z

Figure 3.4: The aerofoil imported in Gambit.


In order to get a good solution, the computational domain must be sufficiently large to
capture the relevant flow. The dimensions of the domain are inspired by [Franck Bertagnolio and Fuglsang, 2001, p7]. When the different points for the flow field are generated, edges between the points are made. The different edges create three different
faces which are named Rect1, Rect2 and Circ. All three faces are meshed. From the
outer edges farfields are created. This done so one can specify inlets and outlets for
the flow fields. After the mesh is created, boundary types are specified. Before we
can specify the boundary types we need to specify different farfields from the edges.
Three farfields are specified and can be seen in Figure 3.5. For farfield1 and farfield2
boundary type is specified as Velocity-Inlet since we want specify a velocity for the
wind tunnel. The boundary type for farfield3 is selected to be outflow. The inspiration
of the geometry and how to make the mesh is taken from [Bhaskaran, 2002]. Figure
+

rfi
e

ld

farfield2

fa

12.5c

Rect1

Circ

++
+ +

Gx

farfield3

Gy

21c

Rect2
Y
Z

farfield2
+

Figure 3.5: Faces, boundaries and dimensions for the flow field
3.6 shows the finished mesh with 150100 cells. The cell density is highest close to the

3.4. Creation of the mesh

26

aerofoil in order to resolve the boundary layer. In order to save on the computational
resources the mesh gets less dense away from the aerofoil. The more dense the cells, the
higher aspect ratio. The aspects ratio should typically not be larger than 20 to 100 in
important regions according to [Casey et al., 2000, p15]. This is violated in the dense
part of the mesh downstream of the aerofoil, but the effects on the flowfield around the
aerofoil are assumed to be negligible.

Figure 3.6: The mesh for the aerofoil with 150100 cells.
Grid

Dec 15, 2008

FLUENT
sstkw) the
The boundary layer needs to resolved in order for Fluent
to 6.3
be (2d,
ablepbns,
to predict
separation phenomena. Figure 3.7 shows a zoom of the aerofoil. Here the high mesh
density around the wing can be seen. Figure 3.8 shows a further zoom of the leading
edge of the aerofoil. A zoom of the trailing edge can be seen in figure 3.9.

27

Grid

Grid

3. Numerical Solution

Dec
15, 2008
Figure 3.7: The a zoom of mesh with 150100 cells around the
aerofoil.
FLUENT 6.3 (2d, pbns, sstkw)

Dec 15, 2008


Figure 3.8: The mesh with 150100 cells for the aerofoil at the leading
edge.
FLUENT 6.3 (2d, pbns, sstkw)

3.4. Creation of the mesh

Grid

28

15, 2008
Figure 3.9: The mesh for the aerofoil at the trailing edge with Dec
150100
cells.
FLUENT 6.3 (2d, pbns, sstkw)

29

3. Numerical Solution

3.5

Setup the case file

3.5.1

Boundary Conditions

As explained in section 3.4, the mesh is divided into three different farfields. Farfield1
and farfield2 are velocity inlet, and farfield3 is outflow. For farfield1 and farfield2 the
velocity specification method is selected to be Magnitude and Direction. The velocity
is set to U = 20[ m
s ] since it is the maximum velocity for the experiments in the wind
tunnel.
The turbulence specification methods is specified as Intensity and Length Scale.
The turbulence intensity defined as ratio of the root-mean-square (RMS) velocity fluc0
tuations, URM
,given by Equation 3.50. [Fluent Inc.,
S , and the mean flow velocity, u
2006]. It is selected to calculate the turbulence intensity from an estimated fully developed duct flow in the wind tunnel. This is done to give an estimation of the turbulence
intensity. The hydraulic diameter for the wind tunnel, DH , is calculated in Equation
3.51, where a is the height of the wind tunnel and b is the width of the wind tunnel.
I

0
1
URM
S
= 0.16(ReDH ) 8
u

(3.50)

2 454 103 m 456 103 m


2ab
=
= 455 103 m
(3.51)
a+b
454 103 m + 456 103 m
The Reynolds number for the hydraulic diameter is the calculated in Equation 3.52.
DH =

ReDH

kg
m
3 m
1.225 m
air U DH
3 20 s 455 10
=
=
= 623.11 103
s
air
1.79 105 N m2
1

I = 0.16(ReDH ) 8 = 0.16(ReDH ) 8 = 0.03 = 3%

(3.52)
(3.53)

The turbulent intensity is then calculated to be 3% in Equation 3.53. Turbulent intensities on 1% are usually considered low and turbulent intensities above 10% is usually
considered high.[Fluent Inc., 2006] Therefore a turbulent intensity of 3% seems reasonable. The length is achosen to be the hight of the total mesh which is 2.5 [m].

3.5.2

Reference Values

In Fluent the Reference Values must be specified in order to calculate the lift coefficient,
CL , and the drag coefficient, CD . The Reference Length is specified as the chord length
c = 0.1 m . The reference depth is inserted to be the span of the wing used in the
experiments which is s = 0.44 m. The reference area is specified as the aerofoil area
given by A = c s = 0.1 m 0.4 m = 0.044 m2 . The dimensions for the wing can be seen
in Figure 3.10

30

s=0.44m

3.6. Grid Independence

c=0.1m
Figure 3.10: Dimensions of the wing given by the span, s, and the chord length, c.

3.6

Grid Independence

This section details the procedure for assessing the independence of the solution on the
grid quality. This analysis is performed in order to minimize the error due to the quality
of the grid used. To this end five meshes of different quality have been developed. The
meshes are listed in Table 3.4
Mesh
Mesh
Mesh
Mesh
Mesh
Mesh

Number of Cells
1
2
3
4
5

4988
12150
45440
99846
150100

Table 3.4: List of the meshes tested for grid independence.

3.6.1

Convergence Histories

Converged solutions have been obtained for each of the meshes at = 0 and = 10 .
The convergence histories for CL and CD are plotted in Figures 3.11-3.15.

31

3. Numerical Solution

Figure 3.11: Convergence histories for CD and CL for Mesh 1

Figure 3.12: Convergence histories for CD and CL for Mesh 2

3.6. Grid Independence

Figure 3.13: Convergence histories for CD and CL for Mesh 3

Figure 3.14: Convergence histories for CD and CL for Mesh 4

32

33

3. Numerical Solution

Figure 3.15: Convergence histories for CD and CL for Mesh 5


As can be seen CL and CD converge asympthodically to a constant value after a number
of iterations. In some cases the CL and CD start to oscillate about some constant mean
value after a number of iterations. In these cases the problem is also assumed converged.
In order to even out the effects of any oscillations CL and CD are calculated as the
average value of the last 1000 iterations.

3.6.2

Comparison of Solutions

The values of CD and CL for the different solutions are plotted in Figures 3.16-3.17.

3.6. Grid Independence

Figure 3.16: Final values of CL as a function of number of cells in the mesh

Figure 3.17: Final values of CD as a function of number of cells in the mesh

34

35

3. Numerical Solution

As it can be seen the solutions do not change very much between meshes 3, 4 and 5.
This indicates that the quality of the solution will not improve substantially if the mesh
is refined further. The solution seems to be grid independent at grid sizes above 50000.

3.6.3

y + Values

However one more aspect of the mesh quality must be taken into account. The y +
values at the aerofoil surfaces for Meshes 3, 4 and 5 are plotted in Figures 3.18-3.20.

Figure 3.18: y + values at the aerofoil surfaces for Mesh 3

3.6. Grid Independence

Figure 3.19: y + values at the aerofoil surfaces for Mesh 4

Figure 3.20: y + values at the aerofoil surfaces for Mesh 5

36

37

3. Numerical Solution

As it can be seen the the y + values are below 5 for all the meshes at = 0 . However
at = 10 the y + values at the leading edge corresponding to x = 0 are in the region
of 9 for Mesh 3 and 6.5 for Mesh 4. This means that the SST k- turbulence model is
not valid at the leading edge in these situations. For Mesh 5 the maximum values of
y + are below 5 for both angles of attack.
Based on the observation that the solutions only change a few percent when the grid
becomes finer than 50000 cells and the above discussion of y + values Mesh 5 is selected
for computing the CL and CD characteristics for the aerofoil.

3.7

Numerical Results

In this section characteristics for CL and CD for the NACA 632 215 aerofoil is developed by solving the flowfield around the aerofoil at different angles of attack using the
mesh of 150100 cells. Converged solutions are obtained for 13 15 .

3.7.1

Convergence Histories

The convergence histories of the solutions are shown in this section. As it can be seen
some of the convergence plots show fluctuations in the values of CL and CD . These
fluctuations are divided into two groups according to their nature.
The first group are the fluctuations that present themselves as spikes in the plots which
otherwise seem to converge to a constant value. In order to solve this the simulations in
which the spikes are pronounced have been run again using the Second Order Upwind
scheme and double precision. These simulations exhibit smooth convergence curves.
After rerunning most of the simulations in this way the source of the spikes has been
identified as the use of single precision in the solver. Since time does not allow for
rerunning all the simulations the results for which the use of single precision and the
QUICK scheme has not produced pronounced spikes have been retained. This is deemed
justifiable since the values to which the solutions converge are almost independent of
the scheme used. Scheme and precision used for the final solutions are listed in Table
3.5.

Scheme

Prescision

Scheme

Prescision

15
14
13
12
11
10
8
6
4
2

SOU
SOU
SOU
SOU
SOU
QUICK
QUICK
QUICK
QUICK
QUICK

double
double
double
double
double
single
single
single
single
single

0
2
4
6
8
9
10
11
12
13

QUICK
QUICK
QUICK
SOU
SOU
SOU
SOU
SOU
SOU
SOU

single
single
double
double
double
double
double
double
double
double

Table 3.5: Schemes and precision used in in the simulations


The other group consists of periodic oscillations of otherwise converged solutions around

3.7. Numerical Results

38

a constant mean value. It is assumed that this results from vortex shedding effects
that the steady state solver used fails to capture. When this occurs, the mean value
about which the solution oscillates is assumed to be equal to the average value of CL
and CD respectively. It is recognised that the best approach would be to solve the
unstable cases using a transient solver to properly resolve the transient effects but this
is omitted due to time constraints. Since all solutions converge to some steady mean
value the above assumption is considered valid. The oscillations become visible around
11 > > 11 and become more pronounced as is increased. This corresponds
well with the expectation that separation and vortex shedding effects should be more
pronounced at higher angles of attack.

Figure 3.21: Convergence histories for CD and CL at = 14


Figures 3.21 show convergence histories for the aerofoil at = 14 . As can be seen the
convergence plot using single precision and the QUICK scheme exhibits large spikes.
In the plot using SOU the change in behaviour when changing from single precision to
to double precision can be seen. The solver is switched to double precision at around
1500 iterations. It also apparent that the mean values to which the solutions converge
are almost equal for both cases. To even out the effects of the fluctuations the mean
value over the last 1000 iterations is used.

3.7.2

Characteristics

The lift and drag characteristics extracted from the numerical solutions are presented
and discussed in this section.

39

3. Numerical Solution

Figure 3.22: Lift and drag characteristics for the NACA 632 215 extracted from the
numerical solutions.
As can be seen in Figure 3.22 the CL curve is almost linear in the region of 8 < <
8 . Outside this region the effects of flow separation are reflected in a reduction the
slope of the curve. The maximum lift is obtained at = 14 . Beyond this point the
effects of separation causes the aerofoil to stall. The effects of the aerofoil camber are
reflected in the nonzero value of CL at = 0.
The CD curve shows minimum drag at around = 0 . As with the CL curve two
almost linear regions 8 < < 0 and 0 < < 8 exist. Outside these regions the
slope increases rapidly due to flow separation.

3.7.3

Investigation of Flow Separation

In this section the flow separation on the surface of the aerofil is investigated in detail.
Plots of the pressure distribution, y + values and recirculation zones are given for different angles of attack in order to describe the development of flow separation. The
pressure distribution is illustrated through plots of the pressure coefficient on the upper
and lower surface of the aerofoil. The pressure coefficient used in the plots is calculated
from equation 3.54.
CP =

p pref
1
2
2 U

(3.54)

Here p is the local pressure on the surface and pref is the reference pressure. The value
of pref is specified as the value of p at the leading edge.
The recirculation zones are illustrated using a contour plot. The black regions correspond to flow in the negative x-direction.
At = 0 the flow around the aerofoil is smooth. This shows in the CP curves merging
smoothly at the trailing edge and the y + curves having no sharp minima. Note the
almost equal shape of the y + curves. Also there are no recirculation zones in Figure
3.25.
Increasing the angle of incidence to = 8 introduces some separation. In Figure 3.26
the plot of the y + on the upper surface shows that y + goes to zero around x = 0.7. At
the same point the slope of the CP curve on the upper surface is visibly reduced. Both

3.7. Numerical Results

40

Figure 3.23: y + values for the upper and lower surface of the aerofoil for = 0 .

Figure 3.24: Plot of pressure distribution Figure 3.25: Plot of the recirculation zones
for = 0 .
for = 0 .

Figure 3.26: y + values for the upper and lower surface of the aerofoil for = 8 .

41

3. Numerical Solution

Contours of X Velocity (m/s)

Dec 16, 2008


FLUENT 6.3 (2d, pbns, sstkw)

Figure 3.27: Plot of pressure distribution Figure 3.28: Plot of the recirculation zones
for = 8 .
for = 8 .
these figures indicate that the flow separates from the surface at this point. As Figure
3.27 shows, the recirculation zone is too thin to be visible. This corresponds well with
the observation that the slopes of the CL and CD curves in Figure 3.22 only change
slightly from = 6 to = 8 . The small black area just below the leading edge of
the aerofoil is caused by the position of the stagnation point on the lower surface. In
order to get to the upper side of the aerofoil some of the flow must go backward over
the leading edge.

Figure 3.29: y + values for the upper and lower surface of the aerofoil for = 10 .
At = 10 the separation bubble is visible as a black shadow above the trailing edge
in Figure 3.30. Figures 3.30 and 3.29 show that the point of separation has moved
forward on the aerofoil to around x = 0.06 [m]. The two minima around x = 0.005
[m] indicate the formation of a tiny separation bubble at this point. This bubble is,
however, too small to be visible on the contour plot.
= 14 is the angle of maximum lift. This is reflected in Figure 3.33 where the area
between the curves reaches its maximum value. The recirculation zone has become
more pronounced which correspond to the further decrease of the slope of the CL curve
in Figure 3.22. In Figure 3.32 the y + plot for the upper surface indicates that the
point of separation has moved further further upstream to around x = 0.0035 [m]. The

3.7. Numerical Results

42

Figure 3.30: Plot of pressure distribution Figure 3.31: Plot of the recirculation zones
for = 10 .
for = 10 .

Figure 3.32: y + values for the upper and lower surface of the aerofoil for = 14 .

Figure 3.33: Plot of pressure distribution Figure 3.34: Plot of the recirculation zones
for = 14 .
for = 14 .

43

3. Numerical Solution

presumed separation bubble on the front of the upper surface of the aerofoil has grown
in length but it is still not large enough to be visible in the contour plot or the pressure
plot.

Figure 3.35: y + values for the upper and lower surface of the aerofoil for = 15 .

Figure 3.36: Plot of pressure distribution Figure 3.37: Plot of the recirculation zones
for = 15 .
for = 15 .
At = 15 the recirculation zone has grown substantially. Also the point of separation
has moved further forward on the aerofoil to about x = 0.03 [m]. The area between the
curves in Figure 3.36 has decreased slightly compared to Figure 3.33. This is reflected
in the drop in CL in Figure 3.22.
Similar behaviour is observed when decreasing the angle below = 0 . To better
illustrate the variation of the pressure distribution and the corresponding lift generation
at the different angles of attack the plots for each angle are given in Figures 3.38-3.56.

3.7. Numerical Results

44

Figure 3.38: Plot of CP at = 13

Figure 3.39: Plot of CP at = 12

Figure 3.40: Plot of CP at = 11

Figure 3.41: Plot of CP at = 10

Figure 3.42: Plot of CP at = 9

Figure 3.43: Plot of CP at = 8

45

3. Numerical Solution

Figure 3.44: Plot of CP at = 6

Figure 3.45: Plot of CP at = 4

Figure 3.46: Plot of CP at = 2

Figure 3.47: Plot of CP at = 0

Figure 3.48: Plot of CP at = 2

Figure 3.49: Plot of CP at = 4

3.7. Numerical Results

46

Figure 3.50: Plot of CP at = 6

Figure 3.51: Plot of CP at = 8

Figure 3.52: Plot of CP at = 10

Figure 3.53: Plot of CP at = 11

Figure 3.54: Plot of CP at = 12

Figure 3.55: Plot of CP at = 13

47

3. Numerical Solution

Figure 3.56: Plot of CP at = 14

Figure 3.57: Plot of CP at = 15

3.7. Numerical Results

48

Chapter 4

Experiments
4.1

Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds Numbers and Angles of Incidence

4.1.1

Introduction

The purpose of the experiment is to determine the lift coefficient CL and drag coefficient
CD values of a NACA 632 215 aerofoil at varying wing chord Reynolds numbers Rec
and angles of incidence in a wind tunnel. A foam NACA 632 215 wing of uniform
cross section is used for the experiment cf. Figure 4.1. CL and CD values for many
standard aerofoils including the NACA 632 215 aerofoil are available in [Abbott and
Von Doenhoff, 1960]. A direct comparison is not possible since the data given is at
Reynolds numbers a factor 10 larger than those obtained in this experiment.

Figure 4.1: CAD image of the NACA 632 215 wing used in the experiment.

49

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
50

4.1.2

Theory

In the determination of the lift and drag characteristics of the NACA 632 215 aerofoil,
the dimensionless lift coefficient CL and drag coefficient CD are used. The use of dimensionless quantities simplifies the experimental investigation by combining variables
in dimensionless variables. The functional dependence of the parameters of principle
interest, here the lift, L [N], and drag, D [N], forces is thus reduced from several variables to a smaller number of dimensionless variables. This leads to fewer variables that
are to be varied in the experiment when determining the dependence of lift and drag
on other experiment parameters. As a result of dimensional analysis the lift and drag
forces are written in dimensionless form as defined in Equation 4.1 and 4.2.
CL =

CD =

(4.1)

1
2
2 U A

(4.2)

1
2
2 U A

h i
 
 
kg
where m
is the fluid density, U ms is the upstream fluid speed and A m2
3
is the wing planform area. Also obtained from dimensional analysis is the functional
dependence of CL and CD on the wing chord Reynolds number and angle of incidence
as expressed in Equation 4.3a, 4.3b, 4.4a and 4.4b. The function f itself cannot be
determined from dimensional analysis.

CL = f (, Rec )


L
U lc
= f ,
1
2

2 U A

(4.3a)
(4.3b)

CD = f (, Rec )


D
U lc
= f ,
1
2

2 U A

(4.4a)
(4.4b)

h i
kg
where lc [m] is the aerofoil chord length and ms
is the dynamic viscosity. A 2nd
order polynomial approximation of as a function of the flow temperature T [ C] and
flow static pressure p [Pa] used to evaluate is given in Equation 4.5. Equation 4.5 is
developed by linear regression of for T [10; 30] and p [101000; 101600] in the EES
software package. The statistics of the fit are given in Table 4.1.
(T, p) = 5.28051517 103 + 1.26974582 105 p
+ 1.80232428 1014 p2 5.60979713 104 T
5

+ 1.40244910 10

T 4.05720675 10

(4.5)

pT

A 2nd order polynomial approximation of for T [10; 30] is given in Equation 4.6.
The regression was performed in a similar manner as for Equation 4.5. The statistics
of the regression are given in Table 4.2.

51

4. Experiments

No. of points
RMS
Bias
R2

420
8.3048 106
1.6141 1015
100.0%

Table 4.1: The statistics for the linear regression of as a function of T [10; 30] and
p [101000; 101600].
No. of points
RMS
Bias
R2

420
1.0715 1011
1.0003 1020
100.0%

Table 4.2: The statistics for the linear regression of as a function of T [10; 30].

(T ) = 1.72936976 105 + 4.87039719 108 T 3.81943529 1011 T 2

(4.6)

The upstream velocity U is determined from the calibration data for the wind tunnel
shown in Figure 4.2. All pressures for the calibration data are in the unit inches water
coloumn denoted [in. H2 O].

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
52

Figure 4.2: Wind tunnel calibration data.

53

4. Experiments

where RP D [Pa] is the reference pressure difference. The RP D is measured between


the two pressure taps on the windtunnel. U and the static pressure p are calculated
from Equation 4.7 and 4.8 respectively cf. Figure 4.2.
s
U =

0.965 (RP D)
1
2

p = 0.1 (RP D)

(4.7)

(4.8)

Lift and Drag Force Measurement


The lift and drag force on the wing at a given Rec and are determined from free
body force considerations and the theory of mechanics of materials. The forces are
determined by measuring the strain in a thin plate to which the wing is connected.
The measurement of strain is performed with metal foil strain gauges. The wing is
screwed on a threaded bolt which is welded to a thin plate. Two sets of strain gauges
are mounted on the centerline of the plate at different distances along the centerline.
Each pair of strain gauges in a half bridge is mounted at the same distance along the
centerline on either side of the plate. The setup is illustrated in Figure 4.3.

Figure 4.3: Orthogonal projection diagrams of the wing mounted on the test stand.
Only part of the wing is shown.
Due to the flow over the wing a nonzero net pressure distribution will act upon the
wing. Summing the pressure over the entire wing area leads to a resultant force F . This
force can be divided into a drag component and lift component as shown in Figure 4.4.
The drag force, D, is defined as the component acting parallel to and in the same
direction as the upstream flow. The lift force, L, is defined as the force component
acting in the plane normal to the upstream flow.

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
54

Figure 4.4: The lift, L, and drag, D, components of the force F acting on the wing
shown.
The determination of the lift and drag components is complicated by the limitation in
the experiment, to determine only the component of F acting normal to the faces of
the plate on which the strain gauges are mounted in the case of 6= 0. The remaining
component parallel to the previously mentioned sides cannot be determined from the
one measurement alone. Figure 4.5 illustrates the drag and lift components normal and
parallel to the strain gauge plate faces when > 0 when the wing chord is parallel to
the strain gauge plate faces.

Figure 4.5: The left figure illustrates the wing chord aligned parallel the strain gauge
plate faces. The right figure illustrates the lift, L1 , and drag, D1 , force components
acting normal (Ln1 and Dn1 respectively) and parallel (Lc1 and Dc1 respectively) to
the strain gauge plate faces when 1 > 0.
The force components Ln1 and Dn1 are determined from Figure 4.5 and are given in
Equation 4.9 and 4.10.
Ln1 = L1 cos 1

(4.9)

Dn1 = D1 sin 1

(4.10)

55

4. Experiments

The measured component of F1 normal to the strain gauge plate faces given in terms
of Ln1 and Dn1 is given in Equation 4.11.
F1 = Ln1 + Dn1 = L1 cos 1 + D1 sin 1

(4.11)

In order to determine L and D a second independent measurement is needed. The


plate is therefore turned 90 as shown in Figure 4.6.

Figure 4.6: The left figure illustrates the wing chord aligned normal to the strain gauge
plate faces. The right figure illustrates the lift, L2 , and drag, D2 , force components
acting normal (Ln2 and Dn2 respectively) and parallel (Lc2 and Dc2 respectively) to
the strain gauge plate faces when 2 > 0.
The force components Ln2 and Dn2 are determined from Figure 4.6 and are given in
Equation 4.12 and 4.13.
Ln2 = L2 sin 2

(4.12)

Dn2 = D2 cos 2

(4.13)

The measured component of F2 normal to the strain gauge plate faces given in terms
of Ln2 and Dn2 is given in Equation 4.14.
F2 = Ln2 + Dn2 = L2 sin 2 + D2 cos 2

(4.14)

For the case of 1 = 2 = it follows that L1 = L2 = L and D1 = D2 = D.


Substituting the aforementioned identitites in Equation 4.11 and 4.14, Equation 4.15
and 4.16 are obtained.
L cos + D sin = F1

(4.15)

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
56

L sin + D cos = F2

(4.16)

Solving Equation 4.15 and 4.16 for L and D, Equation 4.17 and 4.18 are obtained.
L = F1 cos F2 sin

(4.17)

D = F1 sin + F2 cos

(4.18)

Solid Mechanics
The force F acting on the wing results in a strain in the plate in the xaxis direction
denoted x . Since there is a correlation between the strain in the plate and the net force
acting on the wing due to the net pressure distribution on the wing, F can be determined
from x [Gere, 2004]. The strain measured by the strain gauges is proportional to the
normal stress in the xaxis direction in the plate at the location of the strain gauges.
Since the stresses in the plate are within the elastic stress regime, the normal stress in
the plate at the strain gauges is determined from Hookes law given in Equation 4.19.
x = Ex

(4.19)

where x [MPa] is the xaxis component of the normal stress and E [MPa] is Youngs
modulus. The normal stress x at the location of the strain gauges is proportional
to the bending moment at a given position in the plate and can be determined from
Equation 4.20.
x =

My
I

(4.20)

where M [N m] is the bending moment, y [m] is the distance from the neutral axis
to the point where x is to be determined in the yaxis direction when the beam is
undeformed and I m4 is the cross sectional moment of inertia with respect to the
zaxis. I is given by Equation 4.21 for a rectangular cross section such as the plate
cross section shown in Figure 4.7.

Figure 4.7: The plate cross section with the neutral axis and dimensions used in Equation 4.21.

I=

bt3
12

(4.21)

To obtain the bending moment acting in the plate at the strain gauges, the forces acting
on the free body consisting of the plate, bolt and wing are to be determined. The free
body is shown in Figure 4.8.

57

4. Experiments

Figure 4.8: Free body diagram of the wing, bolt and plate body.
The pressure distribution is assumed to be symmetrical about the lengthwise midpoint
so that the resulting pressure force F acts in the wing lengthwise midpoint, a distance
xF [m] from the bottom of the plate. Applying the condition of force equilibrium in the
yaxis and moment equilibrium about the point s Equation 4.22 and 4.23 are obtained
respectively.
X

Fy = 0 Rs F = 0 Rs = F

(4.22)

Ms = 0 Ms + xF (F ) = 0 Ms = xF F

(4.23)

The bending moment acting at the strain gauges is determined by regarding a the plate
cut at a distance xA from the start of the plate. The resulting free body is shown in
Figure 4.9.

Figure 4.9: Free body diagram of the plate cut a distance xA from the start of the
plate.
Applying the condition of moment equilibrium about the point A, Equation 4.24 is
obtained.
X

MA = 0 MA + M s xA Rs = 0
MA = Ms + xA Rs = xF F + xA F
= F (xA xF )

(4.24)

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
58
Chord length, c
Wing length, lw
Plate width, b
Plate thickness, t
Distance strain gauge pair 1, x1
Distance strain gauge pair 2, x2
Distance to force, xF
Youngs modulus, E
Correction factor for strain gauge set 1, kc1
Correction factor for strain gauge set 2, kc2

0.1 [m]
0.44 [m]
0.0205 [m]
0.0022 [m]
0.019 [m]
0.0345 [m]
0.294 [m]
200 [GPa]
0.7273
0.7533

Table 4.3: Constants used to calculate F .


The bending moment at either of the strain gauge sets is obtained by substituting the
distance of the strain gauge, respectively x1 [m] and x2 [m], shown in Figure 4.8, for xA
in Equation 4.24. Substituting Equation 4.20 in Equation 4.19, substituting Equation
4.21 and Equation 4.24 in the result and isolating F , Equation 4.25 is obtained.
F =

bt3 Ex
12y (xA xF )

(4.25)

Due to a systematic error in the force measurements a calibration is performed by


hanging weights on the stand and calculating a correction factor for Equation 4.25
in order to obtain the correct force. A correction factor is obtained for the force
measurements based on the strain gauge pairs 1 and 2, kc1 and kc2 respectively. The
correction factors are given in Table 4.3. The corrected force is then calculated using
Equation 4.26 and 4.27 for strain gauge pairs 1 and 2 respectively.
F =

kc1 bt3 Ex1


12y (xA xF )

(4.26)

F =

kc2 bt3 Ex2


12y (xA xF )

(4.27)

where x1 and x2 are the strain measurements from the strain gauge pairs 1 and 2
respectively.
Strain Measurement
The strain measured using a strain gauge is determined from the change in electrical
resistance R [] in the strain gauge. This is expressed in Equation 4.28 [Hoffmann,
1989].
R
q
=  (1 + 2v) +
R0
q

(4.28)

where v is Poissons ratio and q is he resistivity. The change in resistivity is due to


microstructural changes in the strain gauge. An important property of strain gauges

59

4. Experiments

is the gauge factor k. The gauge factor is the proportionality constant of the strain to
the relative resistance change as expressed in Equation 4.29.
k=

R/R0


(4.29)

For the strain gauges used k = 2. In order to measure the change in resistance the
Wheatstone bridge is commonly used. The circuit is shown in Figure 4.10.

Figure 4.10: The Wheatstone bridge circuit [Hoffmann, 1989].


Here Vs [V] is the bridge excitation voltage, Vo [V] is the bridge output voltage and
R1..4 [] are the bridge arm resistors. Any combination of the bridge arms can be
strain gauges. For a change in the resistance in all the bridge arms, R1..4 , the relative
bridge output voltage, Vo /Vs is determined from Equation 4.30 [Hoffmann, 1989].
Vo
R1 + R1
R4 + R4
=

Vs
R1 + R1 + R2 + R2 R3 + R3 + R4 + R4

(4.30)

For small changes in resistance on the order of 103 the approximation of Equation
4.30 given in Equation 4.31 is sufficiently accurate [Hoffmann, 1989].
Vo
1
=
Vs
4

R1 R2 R3 R4

R1
R2
R3
R4


(4.31)

Substituting Equation 4.29 in Equation 4.31 results in Equation 4.32.


Vo
k
= (1 2 + 3 4 )
Vs
4

(4.32)

where 1..4 are the strains measured by strain gauges 1 to 4 in the Wheatstone bridge.
The measurement system used is a half bridge circuit cf. Figure 4.11. Here R1 and R2
are strain gauges and R3 and R4 are normal resistors. Since the two strain gauges are

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
60
mounted on opposite sides of the plate, they will measure an equal magnitude strain
with opposite signs. Therefore Equation 4.32 can be reformulated to Equation 4.33.
k
Vo
= ()
Vs
2

(4.33)

Figure 4.11: The half bridge Wheatstone circuit. R1 and R2 are strain gauges and R3
and R4 are normal resistors.
In general the bridge outout voltage needs to be amplified. A general strain measurement system is shown in Figure 4.12.

Figure 4.12: The strain gauge measuring system.

4.1.3

Equipment

NACA 632 215 foam wing

61

4. Experiments

m
s

10
15
20
Table 4.4: The U values the wing is tested at.
Plint & Partners wind tunnel, SN: WT1474
Strain gauge stand
4 strain gauges arranged in two half bridges
Resistance: 120 []
Gauge factor: 2:1
2 laptops
CatmanEasy software
LabView software
Differential pressure transducer
HBM Spider8 DAQ system
National Instruments NI cDAQ9172 system
National Instruments NI 9201 analog input module
National Instruments NI 9211 thermocouple module
Ktype thermocouple

4.1.4

Procedure

The wing is placed in the wind tunnel on the strain gauge stand as shown in Figure
4.13. A sampling frequency of 25 [Hz] is specified for the strain gauge data acquisition
(DAQ) system and 25 [Hz] for the thermometer and differential pressure transducer
DAQ system. The wind tunnel is turned on. A measurement of the pressure difference
between
  pressure taps 1 and 2 in Figure 4.13 is performed. The flow velocity magnitude
U ms in the wind tunnel at pressure tap 2 is then adjusted as closely as possible to
the first value given in Table 4.4. The wing is aligned so that = 0 . This is done using
the angle arm shown in Figure 4.13. The pressure difference between pressure taps 1
and 2, the temperature and strain are recorded for 10 [s]. This procedure is repeated
for increments of of 1 2 in the interval [15 ; 15 ]. The strain gauge stand is then
turned 90 counter clockwise as shown in Figure 4.14. The above testing procedure
is repeated. The wing is tested with the preceding procedures for the remaining U
values specified in Table 4.4.

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
62

Figure 4.13: Diagram of the test configuration for determination of Cl and Cd using a
strain gauge stand.

Figure 4.14: The strain gauge stand as seen from above. The diagram to the left
illustrates how the stand is arranged when turned 90 after the first set of measurements
for = [15 ; 15 ].

63

4.1.5

4. Experiments

Uncertainty Analysis and Results

The measurements obtained include both a systematic error and a random error. An
estimate of the probable error, the uncertainty, is therefore necessary in order to determine the relative quality of the measurements. The systematic and random uncertainties, eB and eP respectively, are determined in the following. The total uncertainty, eT ,
associated with the data is a combination of both the systematic and random uncertainty. One method of combining n number of uncertainties is the rootsumsquares
(RSS) method given in Equation 4.34 [Figliola and Beasley, 2006].

etotal

v
u n
uX
=t
e2

(4.34)

i=1

Using the RSS method to determine the combined uncertainty due to eS and eR leads
to Equation 4.35 for eT .
q
eT = e2S + e2R

(4.35)

Since CL and CD are functions of variables each with an associated systematic and
random error an estimate of the error propagation from these variables to the final
values of CL and CD is needed. Given a function such as that in Equation 4.36 the
uncertainty ey due to the propagation of error is determined from Equation 4.37.
y = f (x1 , x2 , ..., xn )
"
ey =

n
X

(4.36)

#( 1 )
2

(i exi )2

(4.37)

i=1

where i is a sensitivity index determined from Equation 4.38.



y
i =
xi x1..n =x1..n

(4.38)

i is evaluated with the mean values of the variables in the expression as indicated
in Equation 4.38. The sensitivity index gives an indication of the sensitivity of ey to
the uncertainties of the variables of which y is a function. The larger a given i , the
larger the effect of the associated exi is upon ey . Equation 4.39 and 4.40 are the full
expressions for CL and CD respectively used to calculate the sensitivity indexes.
CL =

kc Ebt3 (p cos () + n sin ())


1
106
6
U 2 Ay (xA xF )

(4.39)

CD =

kc Ebt3 (p cos () n sin ())


1
106
6
U 2 Ay (xA xF )

(4.40)

where p and n are the strain measurements with the strain gauge plate parallel and
normal to the flow respectively.

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
64
Random Uncertainty
During repeated measurements random errors occur in the data measured. To determine the random uncertainty in the results the sample standard deviation, s, along with
the tdistribution is used with a 95% confidence interval. The random uncertainty is
then calculated from Equation 4.41.
s
eR = t,0.975
n

(4.41)

where t,0.975 is the 97.5 percentile of the tdistribution with degrees of freedom and
n is the sample size. The number of degrees of freedom is determined by the relation
in Equation 4.42.
=n1

(4.42)

The random uncertainty is calculated for the following variables:


n
p
U

In order to determine the propagated uncertainty due to the variables mentioned above,
the sensitivity indexes given in Equation 4.43, 4.44, 4.46 and 4.45 are needed for CL .
CL
1
kc Ebt3 sin ()
= 106
n
6
U 2 Ay (xA xF )

(4.43)

CL
1
kc Ebt3 cos ()
= 106
p
6
U 2 Ay (xA xF )

(4.44)

kc Ebt3 (p cos () + n sin ())


CL
1
= 106
U
3
U 3 Ay (xA xF )

(4.45)

kc Ebt3 (p cos () + n sin ())


1
CL
= 106

6
2 U 2 Ay (xA xF )

(4.46)

For CD the sensitivity indexes given in Equation 4.47, 4.48, 4.50 and 4.49 are needed.
CD
1
kc Ebt3 sin ()
= 106
n
6
U 2 Ay (xA xF )

(4.47)

1
kc Ebt3 cos ()
CD
= 106
p
6
U 2 Ay (xA xF )

(4.48)

kc Ebt3 (p cos () n sin ())


CD
1
= 106
U
3
U 3 Ay (xA xF )

(4.49)

65

4. Experiments

kc Ebt3 (p cos () n sin ())


1
CD
= 106

6
2 U 2 Ay (xA xF )

(4.50)

These sensitivity indexes are evaluated for both sets of strain measurements. The random uncertainties eR,CL1 , eR,CL2 , eR,CD1 and eR,CD2 are then evaluated using Equation
4.51, 4.52, 4.53 and 4.54, which are derived using Equation 4.37.

"
eR,CL1 =

+

CL1
eR,

"
eR,CL2 =

+

eR,CD1 =

eR,CD2 =

+

CD2
eR,


+


+

2


+


+

(4.52)

(4.53)

2
CD2
eR,2,p
2,p
 #( 1 )

CD2
eR,U
U

(4.51)

2
CD1
eR,1,p
1,p
 #( 1 )

CD1
eR,U
U

2

2
CL2
eR,2,p
2,p
 #( 1 )

CL2
eR,U
U

2

2

CD2
eR,2,n
2,n


+

2
CL1
eR,1,p
1,p
 #( 1 )

CL1
eR,U
U

2

2

CD1
eR,1,n
1,n


+

CD1
eR,

"

2

2

CL2
eR,2,n
2,n

CL2
eR,

"

CL1
eR,1,n
1,n

(4.54)

where eR,1,n , eR,1,p , eR,2,n , eR,2,p , eR,U , and eR, are evaluated by applying Equation
4.41 to the variables listed above as given in Equation 4.55, 4.56, 4.57, 4.58, 4.59, 4.60.
s
eR,1,n = t1,n ,0.975 1,n
n1,n

(4.55)

s
eR,1,p = t1,p ,0.975 1,p
n1,p

(4.56)

s
eR,2,n = t2,n ,0.975 2,n
n2,n

(4.57)

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
66
s
eR,2,p = t2,p ,0.975 2,p
n2,p

(4.58)

sU
eR,U = tU ,0.975
n U

(4.59)

s
eR, = t ,0.975
n

(4.60)

Systematic Uncertainty
The systematic error is caused by constant deviation of variables and constants from
their true value as opposed to the random error estimated in the preceding section.
The determination of the systematic uncertainties eS,CL1 , eS,CL2 , eS,CD1 and eS,CD2 is
performed using Equation 4.37. In the case of systematic uncertainties the variable
uncertainties used in Equation 4.37 are estimates. The variables and estimates used
are listed in the following:
eS,kc1 = 0.05
eS,kc2 = 0.05
eS, = 2 [ ]
 
eS,U = 0.2 ms
h i
kg
eS, = 0.1 m
3
 
eS,A = 0.0001 m2
The systematic error in the determination of the force F is calibrated as described
previously. Therefore the systematic error in the variables and constants apart from
kc1 and kc2 in Equation 4.26 and 4.27 are disregarded since they are eliminated by the
correction factors. The remaining systematic error in the calculations of F are due to
the systematic error in the correction factors. The sensitivity indexes needed for CL
are given in Equation 4.61, 4.62 and 4.63.
Ebt3 (p cos () + n sin ())
CL
1
= 106
kc
6
U 2 Ay (xA xF )

(4.61)

kc Ebt3 (p cos () + n sin ())


CL
1
= 106
A
6
U 2 A2 y (xA xF )

(4.62)

kc Ebt3 (p sin () + n cos ())


CL
1
= 106

6
U 2 Ay (xA xF )

(4.63)

The sensitivity indexes for CD are given in 4.64, 4.65 and 4.66.
Ebt3 (p cos () n sin ())
1
CD
= 106
kc
6
U 2 Ay (xA xF )

(4.64)

67

4. Experiments

kc Ebt3 (p cos () n sin ())


CD
1
= 106
A
6
U 2 A2 y (xA xF )

(4.65)

kc Ebt3 (p sin () n cos ())


CD
1
= 106

6
U 2 Ay (xA xF )

(4.66)

The systematic uncertainties eS,CL1 , eS,CL2 , eS,CD1 and eS,CD2 are given in Equation
4.67, 4.68, 4.69 and 4.70.

"
eS,CL1 =

CL1
eS,kc1
kc1

2
+


"
eS,CL2 =

CL2
eS,kc2
kc2

eS,CD1 =

CD1
eS,kc1
kc1

eS,CD2 =

CD2
eS,kc2
kc2


+

2

2

CD1
eS,


+

CD2
eS,U
U

(4.68)

(4.69)

2
CD2
eS,A
A
 #( 1 )
2

(4.67)

2
CD1
eS,A
A
 #( 1 )

CD1
eS,U
U

2


+


+

CD2
eS,

CD2
eS,

2

2
CL2
eS,A
A
 #( 1 )

CL2
eS,U
U

2


+

2
CL1
eS,A
A
 #( 1 )

CL1
eS,U
U

2

2


+

CD1
eS,

2


+

2

CL2
eS,

CL2
eS,

2

"


+

CL1
eS,

CL1
eS,

2

"

(4.70)

Total Uncertainty
The total uncertainty eT,CL1 , eT,CL2 , eT,CD1 and eT,CD2 are determined using Equation
4.35 and are given in Equation 4.71, 4.72, 4.73 and 4.74.
eT,CL1 =

q
e2S,CL1 + e2R,CL1

(4.71)

eT,CL2 =

q
e2S,CL2 + e2R,CL2

(4.72)

4.1. Lift and Drag Characteristics of a NACA 632 215 Aerofoil at Varying Reynolds
Numbers and Angles of Incidence
68

eT,CD1 =

q
e2S,CD1 + e2R,CD1

(4.73)

eT,CD2 =

q
e2S,CD2 + e2R,CD2

(4.74)

The final CL and CD values are calculated as the arithmetic means of the two sets of
values as shown in Equation 4.75 and 4.76
CL =

CL1 + CL2
2

(4.75)

CD =

CD1 + CD2
2

(4.76)

The uncertainty of CL and CD are calculated using Equation 4.37. Equation 4.77 and
4.78 give the total uncertainty for CL and CD .
s
eT,CL =
s
eT,CD =

4.1.6

1
eT,CL1
2

2

1
eT,CD1
2

2


+


+

1
eT,CL2
2

2

1
eT,CD2
2

(4.77)

2
(4.78)

Results

 
Figure 4.15 and 4.16 show CL and CD respectively at U = 10 ms  corresponding
to

m
4
Rec = 6.879410 . Figure 4.17 and 4.18 show CL and CD at U = 15 s corresponding
 
to Rec = 1.0128 105 . Figure 4.17 and 4.18 show CL and CD at U = 15 ms
corresponding to Rec = 1.3315 105 . CL and CD exhibit similar tendencies to data at
Reynolds numbers
1960]. For 0
 m  a factor 10 larger
 m  [Abbott and Von Doenhoff,

at U = 10 s and U = 15 s CL > 0 and for < 0 CL < 0. Both data


m
sets appear to stall at = 11 12 . There is not sufficient
data
for
U
=
15

s
 
to determine the stall point for < 0. For U = 20 ms , CL 0 for 0 and
appears to stall at = 9 and = 14 . This does not agree with the fact that
for = 0 cambered aerofoils should have CL > 0 [Abbott and Von Doenhoff, 1960;
Houghton and Carpenter, 2003]. The errors can be seen to increase for || > 0 . The
systematic uncertainty is dominant with respect to the total uncertainty. Furthermore
the systematic error is proportional to the strain terms. The strain terms increase in
absolute value for an increasing absolute value in the angle of attack therefore leading
to an increasing total uncertainty.
The CD data for all U values shows a parabolic tendency which is in agreement with
the literature for higher Reynolds numbers [Abbott and Von Doenhoff, 1960]. For all CD
data sets there are negative values which are not physically possible. The uncertainties
can however account for these irregularities. As for CL , the CD uncertainties increase
with increasing absolute angle of attack. The explanation for this phenomenon is the
same as for the increasing CL uncertainties.

69

4. Experiments

Figure
4.15: CL at Rec = 6.8794 104 Figure
4.16: CD at Rec = 6.8794 104
 m 
 m 
10 s with errors shown.
10 s with errors shown.

Figure
4.17: CL at Rec = 1.0128 105 Figure
4.18: CD at Rec = 1.0128 105
 m 
 m 
15 s with errors shown.
15 s with errors shown.

Figure
4.19: CL at Rec = 1.3315 105 Figure
4.20: CD at Rec = 1.3315 105
 m 
 m 
20 s with errors shown.
20 s with errors shown.

4.2. Flow Visualisation of a NACA 632 215 Aerofoil at varying Angles of Incidence
Using String Attachments
70

4.2
4.2.1

Flow Visualisation of a NACA 632 215 Aerofoil at varying Angles of Incidence Using String Attachments
Introduction

The purpose of the experiment is to qualitatively analyse the flow around a NACA
632 215 aerofoil. This is done using a NACA 632 215 foam wing with string taped at
one end to the wing. The end of the string not taped is thus able to follow the air
flow. The wing is placed in a wind tunnel and the air flow around the wing at angles of
incidence up to the stall point is investigated by inspection of strings attached to the
midsection and tip of the wing.

4.2.2

Equipment

NACA 632 215 foam wing with string attachments


Plint & Partners wind tunnel, SN: WT1474
Strain gauge stand
Digital camera
Laptop
LabView software
Differential pressure transducer
National Instruments NI cDAQ9172 system
National Instruments NI 9211 thermocouple module
Ktype thermocouple

4.2.3

Procedure

The NACA 632 215 foam wing is mounted on a strain gauge stand and placed in a wind
tunnel as shown in Figure 4.21. The wind tunnel is turned on. The pressure difference
between
  pressure tap 1 and 2 is measured. U is adjusted so as to achieve a value of
15 ms . The angle of incidence is set to 0 . The digital camera is placed as shown
in Figure 4.22. The string movement is observed and captured by the digital camera.
The angle of incidence is varied through the interval [15 ; 15 ] and the movement of
the strings is observed and captured using the camera.

4.2.4

Results

At the wing tips tip vortices could be observed. The strength of the vortices were
observed to increase with increasing angle of incidence. At angles of incidence below
= 10 the strings showed smooth flow. The string showed eddies from an angle of
incidence, = 12 . From = 14 there was a very high recirculation and you could
see on the strings that the eddies was a lot bigger.

71

4. Experiments

Figure 4.21: The experiment arrangement.

Figure 4.22: The wind tunnel as seen from above. Placement of the camera is shown.
Optical access for the camera is provided by the plexiglas wall wind tunnel.

4.3. Flow Visualisation of a NACA 632 215 Aerofoil at Varying Angles of Incidence
Using Smoke
72

4.3
4.3.1

Flow Visualisation of a NACA 632 215 Aerofoil at Varying Angles of Incidence Using Smoke
Introduction

The purpose of the experiment is to qualitatively analyse the flow around a NACA
632 215 aerofoil. This is done using a NACA 632 215 foam wing with smoke injected
into the air flow upstream of the wing. The wing is placed in a wind tunnel and the
air flow around the wing at angles of incidence up to the stall point is investigated by
inspection of the smoke flow around the midsection of the wing. Specifically, the flow
close to and at the stall point of the wing is investigated visually by inspection of the
smoke flow.

4.3.2

Equipment

NACA 632 215 foam wing


Plint & Partners wind tunnel, SN: WT1474
Strain gauge stand
Digital camera
Smoke generator
Laptop
LabView software
Differential pressure transducer
National Instruments NI cDAQ9172 system
National Instruments NI 9211 thermocouple module
Ktype thermocouple

4.3.3

Procedure

The NACA 632 215 foam wing is mounted on a strain gauge stand and placed in a wind
tunnel as shown in Figure 4.23. The wind tunnel is turned on. The pressure difference
between
 m  pressure tap 1 and 2 is measured. U is adjusted so as to achieve a value of
15 s . The smoke generator is turned on. The position of the smoke nozzle is adjusted
so the smoke jet flows around the wing at the vertical midpoint. The position of the
smoke nozzle in the horizontal plane is adjusted so that the smoke jet hits the leading
edge of the wing. The angle of incidence is set to 0 . The digital camera is placed as
shown in Figure 4.23. The smoke flow is observed and captured by the digital camera.
The angle of incidence is varied through the interval [15 ; 15 ] and the flow of the
smoke around the wing is observed and captured using the camera. The experiment
setup is shown in Figure 4.23 and 4.24.

73

4. Experiments

Figure 4.23: The experiment arrangement.

Figure 4.24: The wind tunnel as seen from above. The plexiglass window used to
supply optical access to camera is shown. Placement of the laser is shown.

4.3. Flow Visualisation of a NACA 632 215 Aerofoil at Varying Angles of Incidence
Using Smoke
74

4.3.4

Results

During the flow visualisation weak recirculation was observed at the tip of the trailing
edge at an angle of incidence = 10 . The at higher angles of incidence the degree
of recirculation was observed to increase slightly. At = 15 there was a sudden
transition to full separation on the upper surface of the wing. At this angle the wing
started to oscillate heavily causing the flow to vary between partially attached and fully
separated.

Chapter 5

Comparison and Discussion of


Results
In this chapter the results obtained through the analytical, numerical and experimental investigations are compared. Any discrepancies between the different results are
accounted for and analysed.

5.0.5

Lift Characteristics

The lift characteristics obtained from the three different approaches are plotted in
Figures 5.1, 5.1 and 5.1.

Figure 5.1: Analytical solu- Figure 5.2: Numerical solu- Figure 5.3: Analytical solution for CL vs.
tion for CD vs.
tion for CL vs.
The numerical and experimental results show similar trends whereas the analytical
solution breaks down at high angles of incidence due to the simplifications used in the
analytical analysis cf. chapter 2.
Figure 5.4 shows the graphs for the experimental data at 20 [m/s] and the analytical
solution. As it can be seen the agreement between the two are not very good. The analytical approach fails to predict the correct slope of the curve. Some of this discrepancy
may be a result of the uncertainty of the measurements since the graph of the analytical
solution lies within the error bars except for the area around = 0[ ]. It however, the
cause of the discrepancy may also be the assumption of inviscid flow employed in the
analytical solution. Since viscous effects are more important at lower Reynolds numbers this seems likely. This hypothesis is further justified by the much better agreement
75

76

Figure 5.4: Comparison of analytical and experimental results for CL vs.

between the analytical solution and CL data obtained at higher Reynolds Numbers cf.
Section 2.5.
It is clear from Figure 5.5 that the agreement between the numerical results and the
experiment is much better than the agreement between the experiment and the analytical solution. The slopes of the curves agree very well in the linear region. Also the
curves almost coincide around the stall point for > 0 and both curves have maximum
lift at = 14 [ ]. At negative angle of attack the numerical solution however fails to
capture the stalling of the aerofoil at = 13 [ ]. Given the large uncertainty of the
measurements at this angle the disagreement is not surprising.
One common discrepancy between the experimental and the theoretical results is the
difference in lift at = 0. Since cambered aerofoils produce positive nonzero lift at
= 0 [ ] [Abbott and Von Doenhoff, 1960; Houghton and Carpenter, 2003] the main
source of this discrepancy is assumed to be some overlooked experimental error.

5.0.6

Drag Characteristics

The CD curves for the experiment and the numerical analysis are compared in Figure
5.6. The plot shows that the values of CL at = 0 [ ] agree perfectly. For 6= 0 [ ] there
is a discrepancy between the experimental and the numerical results which grows with
increasing absolute value of . However given the similar shapes of the CD curves as
well as the non-physical negative values of CD obtained in the interval 6 [ ] < < 10 [ ]
and large relative errors of the experimental values it is assumed that the numerical
solution accurately represents the actual variation of CD as a function of

77

5. Comparison and Discussion of Results

Figure 5.5: Comparison of numerical and experimental results for CL vs.

Figure 5.6: Comparison of numerical and experimental results for CD vs.

78

5.0.7

Summary

Generally the numerical and experimental data are found to agree very well. The
analytical solution does not agree well with the experiments. This is mainly contributed
to the limited validity of potential flow assumptions at low Reynolds Numbers.

Chapter 6

Conclusion
In this report the aerodynamic characteristics of a NACA 632 215 aerofoil have been
investigated. Lift and drag characteristics have been developed using experimental,
analytical and numerical approaches. For the experiments a foam wing has been produced and tested in a wind tunnel at three different wind velocities corresponding to
Rec = 6.88 104 , Rec = 1.01 105 and Rec = 1.33 105 . Lift and drag has been measure
using a strain gauge stand. The resulting curves for the lift and drag coefficients show
trends similar to curves found in the literature but are not directly comparable since
the data in the literature has been obtained at Rec > 106 . Uncertainty analysis of the
experimental data has shown bias errors increasing dramatically at angles of attack
different from 0. The relative errors are largest for the drag component for which some
unphysical negative values have been obtained for all sets of data. The lift curves are
in better agreement with the theory even though unexpected negative lift at zero angle
of incidence has been observed.
For the analytical approach thin aerofoil theory has been employed in order to develop
a linear expression to approximate the linear region of the lift curve. The solution has
been shown to be in good agreement with data for Rec = 3.0106 given in the literature.
The agreement with the experimental data is not good however. The slope of the lift
curve predicted by thin aerofoil theory is somewhat steeper than the one observed at
Rec = 1.33 105 .
The numerical investigation has been carried out for two-dimensional flow over the
aerofoil at Rec = 1.33 105 using the SST-k turbulence model. The lift and and
drag characteristics obtained show good agreement with the experimental data. The
best agreement is obtained for the lift curve around the stall point for > 0.
In order to improve the results further investigation of the following areas could be
interresting:
In order to reduce the errors in the measurements the experiment should be redesigned.
It is expected that the drag measurements could be improved notably by using another
measuring technique such as CTA. Also the accuracy of the strain gauge measurements
could be improved by reducing the need for manual alignment of the stand and the
wing.
For the numerical part the sensitivity of the solution to variation of the model parameters could be of interest. Investigating the effects of different turbulence models on the
solutions could be desirable.
79

Bibliography
Abbott, I. H. and A. E. Von Doenhoff (1960). Theory of Wing Sections. Dover Publications Inc.
Anderson, J. D. (2001). Fundamentals of Aerodynamics (3rd ed.). McGraw-Hill.
Bhaskaran, R. (2002). Flow over an airfoil, fluent tutorial. Website visited 8th Dec.
2008. http://courses.cit.cornell.edu/fluent/airfoil/step2.htm.
Casey, M., T. Wintergerste, and S. Innotec (2000). Best Practice Guidelines. ERCOFTAC - European Research Community On Flow, Turbulence And Combustion.
Catalano, P. and M. Amato (2003). An evaluation of rans turbulence modelling for
aerodynamic applications. Aerospace Science and Technology 7, 493509.
Figliola, R. S. and D. E. Beasley (2006). Theory and Design for Mechanical Measurements (4th ed.). John Wiley & Sons, Inc.
Fluent Inc. (2006). FLUENT 6.3 Users Guide. Centerra Resource Park , 10 Cavendish
Court, Lebanon, NH 03766: Fluent Inc.
Franck Bertagnolio, Niels Srensen, J. J. and P. Fuglsang (2001). Wind Turbine Airfoil
Catalogue. Ris National Laboratory.
Gere, J. M. (2004). Mechanics Of Materials (6th ed.). Nelson Thomson Learning.
Hoffmann, K. (1989). An Introduction to Measurements using Strain Gages. Hottinger
Baldwin Messtechnik GmbH.
Houghton, E. L. and P. W. Carpenter (2003). Aerodynamics for Engineering Students
(5th ed.). Butterworth-Heinemann.
Menter, F. R. (1992). Improved Two-Equation k-omega Turbulence Models for Aerodynamic Flows. NASA, Armes Research Center, Mott Field, California.
Versteeg, H. and W. Malalasekra (2007). An Introduction to Computational Fluid
Dynamics: The Finite Volume Method (2nd ed.). Prentice Hall.
Yin, C. (2008). Numerical fluid mechanics. Slides for the course in Numerical Fluid
Mechanics for TEE1 and INTRO.

80

S-ar putea să vă placă și